You are on page 1of 12

Chemosphere 306 (2022) 135310

Contents lists available at ScienceDirect

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

From waste to fertilizer: Nutrient recovery from wastewater by pristine and


engineered biochars
Marta Marcińczyk a, Yong Sik Ok b, Patryk Oleszczuk a, *
a
Department of Radiochemistry and Environmental Chemistry, Faculty of Chemistry, Maria Curie-Skłodowska University, 3 Maria Curie-Skłodowska Square, 20-031
Lublin, Poland
b
Korea Biochar Research Center, APRU Sustainable Waste Management Program & Division of Environmental Science and Ecological Engineering, Korea University,
Seoul 02841, Republic of Korea

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Wastewater is the valuable sources of


nutrients.
• New knowledge on the biochar use to
recover nutrients from wastewater was
discussed.
• Nutrient adsorption is determine by:
biochar’s properties and solution’s pH.
• Potential of biochar as a slow-release
fertilizer was presented.
• The use of biochar-based fertilizer can
be economically and agriculturally
viable.

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

Handling Editor: Y Yeomin Yoon Biochar application for the recovery of nutrients from wastewater is a sustainable method based on a circular
economy. Wastewater, food wastewater, and stormwater are valuable sources of nutrients (i.e., PO3− 4 , NO3 , and

Keywords: NH+ 4 ). The unique properties of biochar, such as its large specific surface area, pH buffering capacity, and ion-
Biochar exchange ability, make it a cost-effective and environmentally friendly adsorbent. Biochar engineering im­
Wastewater
proves biochar properties and provide targeted adsorbents. The biochar-based fertilizers can be a sustainable
Fertilizer
alternative to traditional fertilization. The aim of the study was to compare the potential of pristine and engi­
Circular economy
Nutrients recovery neered biochars to recover nutrients from wastewater and to determine the factors which may affect this process.
Engineered biochar can be used as a selective adsorbent from multicomponent solutions. Adsorption on engi­
neered biochar can be also regulated by additional mechanisms: surface precipitation and ligand/ion exchange.
Metal modification (e.g. Mg, Fe) enhances PO3− 4 and NO3 adsorption capacity, and thus may provide the extra

plant macro-/micronutrients. The desorption mechanism, which is the basis for nutrient release are strongly pH
depended. The use of biochar-based fertilizer can have economic and agricultural benefits when using waste
materials and reducing pyrolysis energy costs.

* Corresponding author.
E-mail address: patryk.oleszczuk@mail.umcs.pl (P. Oleszczuk).

https://doi.org/10.1016/j.chemosphere.2022.135310
Received 25 April 2022; Received in revised form 8 June 2022; Accepted 9 June 2022
Available online 14 June 2022
0045-6535/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

1. Introduction discussed in the paper may contribute to the development of application


methods for the recovery of nutrients.
Natural resources, such as soil and mineral deposits, are depleted to
meet the needs of the growing population. It is estimated that phos­ 2. Biochar
phorus (P) reserves will be depleted during the next 50–100 years, with
the P demand gradually increasing by 2.5% per year (Xu et al., 2018). Biochar is a solid material produced by the carbonization of biomass
Inappropriate and excessive use of fertilizers causes their accumulation (Mohan et al., 2014). Biomass resources, such as agricultural residue,
in the soil (Jie et al., 2002), which leads to their degradation. Accidental food processing waste, solid waste, animal waste, and municipal sludge,
loss of nutrients from nitrogen fertilizers results in elevated concentra­ are thermally decomposed under oxygen-limited conditions through
tions of greenhouse gases as a result of N2O emissions, groundwater pyrolysis (Lee et al., 2019). Biochar has unique properties such as a large
pollution, and eutrophication of surface waters (Coskun et al., 2017). specific surface area, pH buffering capacity, and ion-exchange ability,
Therefore, it is important to identify, evaluate, and implement a sus­ and is perceived as a cost-effective, environmentally friendly, and effi­
tainable strategy to ensure safety in the fertilizer market. An important cient multifunctional adsorbent (Huang et al., 2020a). Furthermore,
part of the newest growth strategy (e.g., the European Green Deal biochar has been shown to have a positive effect on many physico­
(EGD)) is circular economy, which is an economic model focused on chemical and biological soil properties by improving ion exchange
resource efficiency (Klein et al., 2022) that includes the perspectives of properties, increasing soil pH, and enhancing water-holding capacity
both raw material management and waste management (UE COM no. (Ye et al., 2020b). As a soil amendment, biochar can be used as a fer­
398, 2014). One of the possibilities consistent with the circular economy tilizer to bind toxic contaminants, such as heavy metals (Song et al.,
is the use of biochar as both a fertilizer and an adsorbent for nutrient 2020), and stimulate soil microbial activity (El-Naggar et al., 2018). The
recovery. Biochar has a long agricultural history (Kookana et al., 2011) biomass type, pretreatment method, and pyrolysis conditions (temper­
as the main factor leading to high fertility in the Terra Preta de Índio ature, duration, and carrier gas) affect the physicochemical properties of
(Amazonian dark earth). This unique soil has a significantly high biochar (Cao et al., 2013). Biochar prepared from manure is usually
nutrient content, especially P and N. Biochar contributes to the richer in nutrients, has a higher pH, and is characterized by a higher
improvement of soil properties (Agbede and Oyewumi, 2022; L. Ye et al., specific surface area than biochar obtained from highly cellulosic feed­
2020a), carbon sequestration, reduction of greenhouse gas emissions, stock such as wood (Ghodake et al., 2021). Biochars produced at high
and removal of organic and heavy metal pollutants (Krasucka et al., temperatures (>550 ◦ C) are generally characterized by a larger specific
2021; Song et al., 2020; Su et al., 2021). The use of nutrients contained surface area and higher pH but have lower nutrient content than bio­
in different waste streams to produce fertilizers is strongly recom­ chars produced at lower temperatures (<550 ◦ C) (Ippolito et al., 2020).
mended by the circular economy (Smol, 2021). Agriculture, forestry, Pristine biochar has some limitations as an efficient fertilizer owing to its
and fish waste, household waste, and wastewater are only a few exam­ physicochemical properties, such as a negatively charged surface, low
ples of nutrient-rich wastes that can be used for fertilizer purposes. specific surface area, and lack of acidic functional groups (Kavitha et al.,
These materials provide valuable nutrients for plant growth, including 2018; Takaya et al., 2016), which limits its utilization in reducing the
N, P, K, Ca, and Mg. However, the use of waste is not only affected by the leaching of anionic nutrients (NO3- and PO43-) from soils. However,
presence of elements that are valuable for plants but also by the presence engineered biochars or biochar composites can be a good replacement
of pollutants such as Pb, Hg, and Zn (Qasem et al., 2021). Therefore, for pristine biochar because these biochars have the potential for various
there is a need to search for sorbents that effectively recover nutrients applications in the soil (as pristine biochar), but more importantly, may
from waste as well as reduce the penetration of harmful substances into contain nutrients and slowly release them from the matrix, thereby
the environment. Moreover, biochar seems to be an excellent tool to improving their functionality Wang et al., 2022). Engineered biochar or
fulfill the 17 Sustainable Development Goals (SDGs) of the United Na­ biochar composites are the form of pristine biochar modified with ma­
tions, especially SDG15 (protecting life on land) and SDG2 (promoting terials such as metals, nanomaterials, microorganisms, and hydroxides
sustainable agriculture). Several conventional methods are used to (Wang et al., 2021) (Fig. 1). The development of EngBCs has gained
recover nutrients from wastewater, including biochemical degradation, significant attention recently, and numerous reports have shown the
chemical precipitation, flocculation/coagulation, and artificial wetlands efficient removal of heavy metals, toxic dyes, and contaminants from
(Chrispim et al., 2019; Santos and Pires, 2018). These methods usually municipal wastewater and industrial effluent (Kun Luo et al., 2021;
remove valuable nutrients from wastewater, with high operational and Yaashikaa et al., 2019).
maintenance costs and massive waste production (Wang and Wang,
2019). The efficient techniques for P recovery from wastewater are the 3. Nutrient sources
Integrated Fixed-Film Activated Sludge Systems with Enhanced Bio­
logical Phosphorus removal (IFAS-EBPR) and struvite precipitation Phosphorus is a non-renewable resource; the conversion of N to
(Bashar et al., 2018). Removal of nitrogen (N) from sewage is based on fertilizer using the Haber–Bosch process is energy intensive and nutrient
physical (membrane processes, filtration, adsorption, stripping) or bio­ losses cause environmental eutrophication (Lim and Welty, 2018).
logical methods (Bio-electrochemical systems) (Beckinghausen et al., Therefore, the recovery of nutrients from wastewater has attracted
2020). The costs of wastewater treatment are largely determined by considerable attention. The potential sources of macro- (N, P, K, Ca, Mg,
energy prices, e.g. IFAS-EBPR: 60 € P kg− 1 (Stávková and Maroušek, and Na) and micronutrients (Cu, Fe, Mn, and Zn) that can be used for
2021) and vacuum stripping 470 US $ t− 1 (Beckinghausen et al., 2020). fertilization are municipal, industrial, and agricultural waste and
Adsorption using biochar is seems to be cost-effective, user-friendly, and wastewater. Wastewater is considered a source of nutrients because of
efficient method for removing nutrients from wastewater. Therefore, the its high concentration and high production. The global wastewater
aim of this study was to collect existing data on the possible applications production is estimated to be 359.4 × 109 m3 yr− 1, of which 63% is
of biochar to recover nutrients from wastewater and present its potential collected and 52% is treated using various methods (Jones et al., 2021).
for use as a slow/control-release fertilizer. Special attention was directed Moreover, 48% of global wastewater production is released into the
to the sources of nutrient-rich wastewater and their use for fertilizer environment untreated (Jones et al., 2021). The total recovery of P from
purposes. This paper broadly discusses the adsorption and desorption wastewater would fulfill 15–20% of the global demand for this element
processes of macro- and micronutrients using biochar and engineered (Jing et al., 2019). The main types of wastewater are domes­
biochars (EngBCs). The green entrepreneurship, i.e., cooperation be­ tic/municipal (gray or black), industrial, and stormwater runoff (Dia­
tween scientists and practitioners, should be the basis for developing z-Elsayed et al., 2019). Municipal wastewater (particularly black water)
new techniques and practice (Muo and Azeez, 2019). The research is an excellent source of N and P (Yu et al., 2020). Human urine can also

2
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

Fig. 1. Biochar modification methods.

serve as a source of nutrients (Jena et al., 2021). Urine constitutes less digestion (Santos and Pires, 2018) and membrane hybrids are effective
than 1% of the total wastewater (Pathy et al., 2021) but approximately processes with high operational costs and difficult technical feasibility
80% of the N, 50% of the P, and 70% of the K in municipal wastewater is (Ye et al., 2020). The present review focuses on the implementation of
from urine (Wilsenach et al., 2003). Food wastewater, produced during wastewater nutrient recovery using adsorption on biochar and engi­
the processing of food as a result of dehydration or anaerobic degrada­ neered biochar and biochar composites. Biochar has the potential to
tion, is also rich in nutrients (Thant Zin and Kim, 2021). The use of recover nutrients from wastewater but can also be an additional source
biowaste composting provides essential nutrients at high concentrations of macro- and micronutrients, which can significantly increase its
(Antonangelo et al., 2020; Oldfield et al., 2018). Stormwater also con­ fertilization properties (Marcińczyk and Oleszczuk, 2022). The appli­
tains nutrients in various forms including soluble ions (NH+
4 , NO3 , NO2 ,
− −
cation of nutrient-loaded biochar in agricultural fields can also act as a
3−
and PO4 ), precipitates (with metals or carbonates), colloids, and par­ slow/control-release fertilizer.
ticulate matter (particles larger than 0.45 μm) (Tondera et al., 2018).
However, it should be noted that only nutrients recovered from stable 4. Nutrients adsorption using pristine biochar and modified
sources (e.g. municipal, fermentation residues) can be an alternative to biochar
traditional sources (Maroušek and Gavurová, 2022; H. Yang et al.,
2020). The use of waste as a source of nutrients may face one technical Adsorption techniques are considered promising for recovering nu­
(legislative constraints), economic (high transportation and storage trients from wastewater. The simplicity of the procedure, cost-
costs) and practical (large volumes) barriers (Vaneeckhaute et al., effectiveness of sorbents, and economics (Loganathan et al., 2014)
2017). indicate the advantages of this technique (Fig. 2). The sorption method
In the past decade, many review studies have included detailed in­ is relatively more effective at low concentrations; post-treatment pro­
formation on nutrient recovery from wastewater (Fig. 2). Microalgae- cedures are often not required, and no sludge is produced during or after
based wastewater treatment processes (Li et al., 2019), anaerobic the process (Bhatnagar and Sillanpää, 2011). The high cost of sorbent

Fig. 2. Nutrient recovery from wastewater.

3
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

regeneration may affect the practicality of adsorption methods. parameters such as atomic ratios H/C and O/C. High temperature
Furthermore, sorbents are usually characterized by a low selectivity for removes H- and O-containing functional groups, increases the aroma­
coexisting ions. Adsorption techniques are sensitive to pH and temper­ ticity and decreases polarity of biochars compered to low temperature
ature (Fidel et al., 2018). The use of waste-produced biochar reduces the pyrolysis conditions (<500 ◦ C).
costs of adsorption techniques and also provides selective, engineered
sorbents (Panwar and Pawar, 2022). 4.2. Ammonium adsorption using engineered biochar

4.1. Ammonium adsorption using pristine biochar Pristine biochar has a limited ability to bind nutrients and hence the
recently growing interest in EngBCs, which after appropriate modifica­
Ammonium ions are the main form of N (NH−4 N) in water. Yang et al. tion, achieve an even greater ability to adsorb NH−4 N ions (Thant Zin and
(2018) determined the ability of biochars produced at 300 ◦ C (PS300) Kim, 2021). EngBCs can be used to remove NH−4 N from water/waste­
and 500 ◦ C (PS500) from pine sawdust and at 550 ◦ C from wheat straw water; however, after adsorption they can be used as fertilizer as a
(WS550) to adsorb NH−4 N ions from wastewater. PS300 was character­ source of NH−4 N in soil. Modifications of biochar increase its cation ex­
ized by the highest affinity for NH−4 N ions, adsorbing NH−4 N ions at a change capacity (CEC) and introduce additional functional groups
level of 5.38 mg g− 1. Significantly lower adsorption values were (Zhang et al., 2020), which have a notable impact on the mechanisms
observed for PS550 (3.37 mg g− 1) and WS550 (2.08 mg g− 1). The au­ responsible for NH−4 N adsorption. Natural clay minerals, such as mont­
thors demonstrated that the adsorption of NH−4 N primarily occurs based morillonite (Mt) and bentonite, are characterized by a higher specific
on chemical bonds and electrostatic interactions between the negatively surface area, significant CEC, and adsorption capacity relative to bio­
charged biochar surface and positively charged NH−4 N (Yang et al., chars. Because of this capacity, these minerals have been used to
2018). The higher H/C ratio indicates relatively lower aromatic struc­ fabricate biochar composites (Chen et al., 2017). Clay minerals are
ture (higher organic carbon content) and may suggest higher adsorption formed by tetrahedral layers of silica, in which water molecules and
sites potentially available for adsorption of nutrients. Additionally the metal cations occur (Ismadji et al., 2016). Chen et al. (2017)
high O/C indicates the oxygen-containing functional groups and co-pyrolyzed bamboo powder and Mt. The pyrolysis process was carried
hydrophilic/polar character of biochars (Gai et al., 2014; Lou et al., out in the temperature range of 300–500 ◦ C, in a N atmosphere. The
2016). The higher values of H/C and O/C, which characterize PS300, montmorillonite-biochar (MtBC) composites obtained were tested for
indicate the presence of numerous functional groups on the surface of their efficiency in removing NH−4 N. The adsorption of NH−4 N on the
the biochar and its low carbonization (Yang et al., 2018), which resulted MtBC was 12.5 mg g− 1, which was more than five times higher than that
in the highest Qmax (maximum adsorption capacity) values for NH−4 N on unmodified biochar. This positive effect of NH−4 N adsorption was
(Yang et al., 2018). Similar conclusions were drawn by Hu et al. (2020), achieved owing to the interaction of NH−4 N with the functional groups
who investigated the possibility of removing NH−4 N from aqueous so­ present on the surface of Mt and the binding (intercalation) of ions in the
lutions using biochars produced from fruit skin (orange, pineapple, and interlayer spaces of Mt. In another study conducted using bentonite as a
pitaya) at 300, 400, 500, and 600 ◦ C. The study found higher adsorption component of a biochar composite, the adsorption of NH−4 N was found
values for the biochars produced at lower temperatures compared to to be at a level of 23.7 mg g− 1. For comparison, the adsorption of the
those produced at higher temperatures, regardless of the type of feed­ same ions on the biochar or bentonite was 9.49 and 12.4 mg g− 1,
stock used. The highest adsorption capacity (5.60 mg g− 1) was deter­ respectively (Ismadji et al., 2016). A higher specific surface area and
mined for the biochar prepared from pineapple peel (300 ◦ C; H/C = more developed porous structure of the bentonite-biochar composite
0.06; O/C = 0.29). Xu et al. (2019) suggested that biochar characterized were the factors determining its higher adsorption capacity compared to
by high H/C content is more effective for the adsorption of inorganic that of bentonite or biochar alone. The mechanism of binding NH−4 N is
pollutants and high O/C ratio (represents high hydrophilicity and po­ based on van der Waals interactions and ion exchange with the metal
larity) is recommended to the adsorption of polar soluble pollutants such cations present in the interlayer spaces of bentonite (Ismadji et al.,
as ammonium. Furthermore, the study (Hu et al., 2020) showed that the 2016).
pH of the solution had a significant effect on NH−4 N adsorption. Kizito Another biochar modification method involves the use of oxidants
et al. (2015) demonstrated that the adsorption efficiency of NH−4 N had (acids and H2O2), which are designed to increase the percentage of ox­
the lowest values in low pH ranges (3–4), increasing in the pH range ygen functional groups (Han et al., 2021). Specific oxidants provide
(4–8). When the pH exceeded 8, the adsorption decreased again. The advantages and disadvantages. For example, H2O2 minimizes the pos­
authors explained the behavior of NH−4 N ions through the fact that at sibility of accompanying reactions such as precipitation or complexa­
low pH values, some functional groups (COO- and OH-) adopt a positive tion, which may occur when acids are applied (H2SO4, HNO3) (Wang
charge as a result of protonation, which leads to the repulsion of posi­ et al., 2015a, 2015b). In a study, Wang et al. (2020) used H2O2 to modify
tively charged NH−4 N ions from the biochar matrix (Gong et al., 2019). biochar produced from peanut shells at a temperature of 500 ◦ C. In this
Moreover, in a low pH solution, there is a high concentration of H+ ions way, an increase in sorption affinity was achieved toward NH−4 N from
that compete with NH−4 N ions for active sites on the biochar surface 24.6 mg g− 1 for the pristine biochar to 123.3 mg g− 1 for the H2O2-mo­
(Tang et al., 2019). At a high pH (>8), NH−4 N ions are converted to NH3, dified biochar. This enhanced adsorption was linked to a change in the
which substantially reduces the adsorption of these ions because the surface properties of the biochar because of H2O2 modification. Nitro­
electrostatic interaction mechanism is no longer efficient (Tang et al., gen, located in functional groups on the biochar surface, initially having
2019). It was also noted that with increasing biochar mass, in the range pyrrolic character, changing to pyridine after adsorption of NH−4 N,
of 0.1–1 g, the amount of adsorbed NH−4 N increased, which is a result of which indicated that chemical adsorption predominantly occurred (π-π
a higher number of available adsorption sites (Tang et al., 2019). bonds and hydrogen bonds). Another method for increasing the biochar
Nonetheless, the application of biochar rates exceeding 1 g no longer adsorption capacity for NH−4 N is neutralization of the acidic surface.
increased the adsorption of NH−4 N and could even decrease it (Tang Because the affinity of alkali metal ions for the functional groups of
et al., 2019). This was because of the aggregation of biochar particles biochar is generally lower than that for H+ ions, they can be more easily
and the decrease in available active sites on the adsorbent. The replaced with NH−4 N (Han et al., 2021). In a study by Wang et al. (2015a,
adsorption of NH−4 N ions also increased with a decrease in biochar 2015b), maple wood-derived biochar was oxidized using H2O2. The
particle size (Tang et al., 2019). The Qmax determined for NH−4 N for determined adsorption of NH−4 N ions on the modified biochar increased
different biochars can adopt a wide range of values (Table S1), which from 1.77 mg g− 1 in pH 3.69–5.44 mg g− 1 after pH neutralization to 7.
predominantly depends on the type of biochar used. The pyrolysis Modification of biochar with HCl or H2SO4 (Wang et al., 2020) also
temperature plays the important role in biochar’s physicochemical increases its adsorption capacity for NH−4 N, which in such cases can

4
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

reach Qmax values of 100 mg g− 1 and 120 mg g − 1, respectively. It was adsorption of NO−3 N on biochars produced at different temperatures
suggested (Wang et al., 2020) that biochar with high porous structure (200, 350, 500, and 700 ◦ C) from feedstocks with a high content of
(high SBET) effectively adsorb nitrogen ions and exhibit the fast lignin and cellulose. The composition of the biopolymers in the feed­
adsorption rates. Alkali treatment (NaOH) of biochar also increases the stock may determine biochar properties such as their surface area and
negative charges on it surfaces and enhances the ammonium adsorption. pores size distribution, and thus their sorption characteristics. Higher
Liu et al. (2016) reported that NH4–N adsorption by modified (1 M adsorption was observed for the lignin biochars produced at higher
NaOH) biochars produced from peanut shells, corncobs, or cotton stalks temperatures. However, the adsorption levels achieved in these exper­
was respectively 1.3–1.7-2.5-times higher than on pristine biochars. In iments were very low, ranging from 0.029 to 0.580 mg g− 1.
addition modified biochars demonstrated extremely good NH4–N
adsorption ability at high temperature (50 ◦ C), indicating that it can be 4.4. Nitrate adsorption using engineered biochar
used in high temperature environment for ammonium removal. Biochar
modifications for NH−4 N adsorption applications can also be based on the Biochar modifications aimed at improving NO−3 N adsorption are
impregnation of biochar with MgO. Gong et al. (2017) impregnated designed to reduce the electrostatic repulsion between NO−3 ions and the
biochar prepared from Phragmites australis using MgCl2, which increased negatively charged biochar surface (Zhang et al., 2020). One method for
its adsorption capacity from approximately 1 to 30 mg g− 1 NH−4 N. The engineering biochar to increase its NO−3 N adsorption capacity is Al
main mechanism responsible for NH−4 N adsorption using Mg-coated modification (Yin et al., 2018a,b). As a result of this type of modifica­
biochar was ion exchange between Mg2+ and NH+ 4 . The literature data tion, Yin et al. (2018a,b) obtained NO−3 N adsorption capacity that was
included in the table clearly indicate that biochar modification affects five times higher (89.6 mg g− 1) than that of pristine biochar. The authors
the NH−4 N adsorption capacity. used pre-pyrolysis modification, during which the initial feedstock,
which consisted of poplar pieces, was impregnated with an AlCl3 solu­
4.3. Nitrate adsorption using pristine biochar tion, dried, and pyrolyzed at a temperature of 550 ◦ C in a N atmosphere.
The main reason for the substantial increase in adsorption was the
Apart from NH−4 N, NO−3 anions are nitrogen-containing ions (NO−3 N) development of the specific surface area and a change in the charge of
that occur most frequently in aqueous solutions. The electrostatic the Al-coated biochar surface to become positive, which attracted
repulsion between the negatively charged biochar surface and NO−3 N negatively charged NO−3 . Yin et al. (2018a,b) investigated NO−3 N
can be an obstacle to the adsorption of these ions by biochar (Zhang adsorption on Al- and Mg-modified biochar composites. The Al-coated
et al., 2020). Nonetheless, research has shown that adsorption of NO−3 N biochar, for which the maximum adsorption capacity (40.6 mg g− 1)
using biochars is possible. Alsewaileh et al. (2019) compared NO−3 N was 8 and 2.5 times higher than for the pristine biochar and for the
adsorption on biochars produced from date palms at temperatures of Mg-coated biochar, respectively, proved to be the most efficient. When
300 and 700 ◦ C. The biochar produced at 700 ◦ C exhibited twice the modifying biochar produced from peanut shells at a temperature of
adsorption (7.73 mg g− 1) of that generated at 300 ◦ C. The better 600 ◦ C with MgCl2, Zhang et al. (2012) found the maximum NO−3 N
adsorption capacities of biochars produced at higher temperatures are adsorption capacity to be 95 mg g− 1. In an aqueous environment, MgO
usually due to their higher specific surface area (Alsewaileh et al., 2019) on the biochar surface adopts a positive charge and can therefore effi­
and thus, a higher number of active sites are capable of adsorbing NO−3 N. ciently combine with negatively charged NO−3 N (Yao et al., 2011). When
Moreover, with increasing pyrolysis temperature, the number of modifying biochar derived from Conocarpus green waste (600 ◦ C) with
oxygen-containing functional groups and thus the electrostatic repulsion MgCl2, Usman et al. (2016) did not achieve high adsorption values as in
between the biochar and NO−3 N ion can become weaker (Alsewaileh the studies cited previously. Although the techniques used managed to
et al., 2019). In such cases, the adsorption of negatively charged NO−3 N is increase the biochar adsorption capacity by nearly three times (Usman
possible based on covalent and electrostatic interactions. Solution pH et al., 2016), it was not at a satisfactory level at 0.8 mg g− 1.
also plays an important role in the adsorption of NO−3 N by biochars Lanthanum is another metal that has been used to modify biochar to
(Fidel et al., 2018). At low pH (<7), H+ ions can accumulate on the increase its adsorption capacity for NO−3 N. Owing to the La modification
biochar surface; thus, the number of positively charged active sites in­ of biochar derived from oak sawdust (600 ◦ C), Wang et al. (2015a,
creases, which also results in increased adsorption of negatively charged 2015b) obtained an almost 12-fold increase in NO−3 N adsorption of 100
NO−3 N ions. At pH (>7), there is competition between OH− anions and mg g− 1. The significant increase in adsorption was attributed to an in­
NO−3 N, which reduces the adsorption. Research has revealed that the crease in basic functional groups on the modified biochar, which facil­
feedstock used to produce biochar plays an essential role in the itated chemisorption. The O/C (0.23) and (O + N)/C (0.24) ratios in
adsorption of NO−3 N. Zhao et al. (2017) compared the NO−3 N adsorption La-biochars (600 ◦ C) were both higher than those of pristine biochar
capacity of biochar produced from corncobs, peanut shells, and corn (600 ◦ C) (O/C = 0.15 and (O + N)/C = 0.16) and confirmed the higher
waste at a temperature of 600 ◦ C. These authors found that corncob content of basic functional groups. Research has shown that modifica­
biochar was the most efficient for NO−3 N adsorption. The maximum tion using concentrated HCl also causes an increase in the biochar
sorption capacity determined for this biochar was 14.5 mg g− 1 NO−3 N, adsorption capacity for NO−3 ions (Chintala et al., 2013). Using
which was approximately nine times higher than those determined for HCl-modified biochar from Ponderosa pine wood (650 ◦ C), Chintala et al.
the other biochars. Hafshejani et al. (2016) also found a similar (2013) found the adsorption capacity to increase from 2.58 mg g− 1 to
maximum sorption capacity value of 11.6 mg g− 1 for sugarcane bagasse 9.74 mg N g− 1. Chemical modification may be provided also with using
biochar produced at 300 ◦ C. Nonetheless, the majority of studies have N, N-dimethylformamide (Hafshejani et al., 2016). During this modifi­
reported either marginal or no adsorption of NO−3 N on pristine biochar. cation the additional positively charged functional groups are intro­
When studying NO−3 N adsorption on mustard straw (300 ◦ C) and wheat duced to biochars which enhance NO3–N adsorption (over 2 times)
straw (300 ◦ C) biochar as well as on commercial activated carbon, compared to pristine biochar (Table S2). However, the presence of
Mishra and Patel (2009)obtained very low adsorption values, which co-existing ions (e.g., phosphate, carbonate, sulfate or chloride) may
were as follows: 1.3 mg g− 1, 1.1 mg g− 1, and 1.22 mg g− 1, respectively reduce the NH4–N adoption capacity. Multivalent anion with higher
(Mishra and Patel, 2009). Investigating adsorption of NO−3 N using charge density was adsorbed more favorable than monovalent anion.
wheat-straw, corn-straw, and peanut-shell biochars produced at Another modification method contributing to the enhanced NO−3 N
different temperatures (400, 500, 600, and 700 ◦ C), Gai et al. (2014) adsorption by biochar is the use of clay minerals. Viglašová et al. (2018)
observed no adsorption. A study conducted by Hale et al. (2013) also proposed the synthesis of bamboo-based MtBC composites. The feed­
demonstrated marginal adsorption of NO−3 N on biochars derived from stock was combined with Mt suspension, dried, and pyrolyzed at 460 ◦ C.
cacao shells and corncob (300–350 ◦ C). Yang et al. (2017) studied Owing to the addition of Mt to biochar, a two-fold increase in NO−3 N

5
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

adsorption was obtained at a level of 9 mg g− 1 (Yu et al., 2020). The introduction of additional metals e.g., Mg or Fe, which increase the
adsorption of NO−3 N on clay mineral composites is based on complex content of the mineral fraction; largely participates in the binding of P.
mechanisms that combine electrostatic interactions, ionic bonding, and Zhang et al. (2021) modified sludge biochars with Mg and Mg–Fe. The
ion exchange (Viglašová et al., 2018). When comparing the biochar adsorption of PO−4 P on the Mg-biochar composite was 142.3 mg g− 1, and
modification methods in terms of biochar applications for NO−3 N lower values were determined for the Mg–Fe biochar (35.4 mg g− 1);
adsorption, it should be noted that the methods based on Al or Mg however, both results were much higher than those for the pristine
addition are more beneficial because they do not cause a significant biochar (2.3 mg g− 1). The authors suggested that the process of PO−4 P
increase in adsorption relative to the initial biochar; however, they also adsorption on the composites is complex and involves a range of
sorb NO−3 N several times better than the biochars modified using other mechanisms, such as protonation, electrostatic attraction, surface
methods (Table S2). Generally, NO3–N adsorption is determined by inner-sphere complexation, and precipitation. The impregnation of
presence the basic functional groups (high N content and (O + N)/C Mg-biochar seems to be an effective method to increase PO−4 P adsorp­
ratios), electrostatic interaction (outer-sphere complexation mecha­ tion, which was also confirmed in a study by Shin et al. (2020). The
nism) or high surface area (Yang et al., 2017). modification of biochar from ground coffee bean waste with Mg resulted
in an adsorption of 56 mg g− 1. Xiao et al. (2020) also indicated that
4.5. Phosphate adsorption using pristine biochar Mg-coated biochar composites were efficient for PO−4 P removal.
Nevertheless, the values obtained were significantly lower (21.8 mg g− 1)
Phosphate ions, which are another macronutrient essential for plant than those presented in previous studies but almost 14 times higher than
growth (Yang et al., 2020), are best assimilated by plants as H2PO−4 and that of the pristine biochar. Modification proposed by Peng et al. (2021)
HPO2− 4 . Depending on the pH of the environment, inorganic forms of P involved the formation of Mg/Al layered double hydroxides (LDHs),
occur as H2PO−4 , HPO2− 3−
4 , and PO4 (PO4 P) ions (Zhang et al., 2020).

composed of both positively and negatively charged layers, and water
Similar to the case of NO−3 N ions, the adsorption of PO−4 P on pristine filling the interlayer spacing on the surface of the biochar (corn straw,
biochar is usually inhibited owing to the electrostatic repulsion of anions 950 ◦ C). The adsorption of PO−4 P on the obtained composite was found
by the negatively charged biochar surface. Phosphorus adsorption on to be 286.2 mg g− 1. Yang et al. (2019) also confirmed the high efficiency
biochar is dependent on the concentration and availability of cations in of Mg/Al LDH biochar composites for PO−4 P adsorption. The adsorption
its inorganic fraction. Thus, biochars with low ash content exhibit a low of PO−4 P using the produced biochar composite (corn stalk, 600 ◦ C) was
ability to bind P because they contain small amounts of metals (Shep­ 152 mg g− 1. The high results for PO−4 P adsorption on the BC-LDHs
herd et al., 2017). Furthermore, the mechanisms of P binding by bio­ composites can be explained by the good dispersion of colloidal LDH
chars are related to an ash fraction rather than a functional carbon flakes in the aqueous solution, which increased the number of active
group. As reported by Hale et al. (2013), cacao shell and corncob bio­ sites and enabled efficient adsorption on the composites. Huang et al.
chars with an ash content of 8–17% adsorbed PO−4 P in low amounts of (2020b) modified biochar with lanthanum and chitosan, achieving
0.022–0.748 mg g− 1. Jung et al. (2015) suggested that Mg and Ca PO−4 P adsorption at 109 mg g− 1. The adsorption mechanism was based
contents in biochars are of key importance for PO−4 P sorption by bio­ on electrostatic interactions, but Lewis acid-base interactions and ligand
chars, particularly the Mg/P and Ca/P ratios. The authors observed that exchange also played a significant role. Particularly, very high PO−4 P
the biochars tested (prepared from soybean stover, bamboo wood, and adsorption results were obtained by Novais et al. (2018) for biochars
maize residue), with low Mg/P (<1.5) and Ca/P (<5.0) ratios, poorly derived from poultry manure (350 ◦ C) and sugarcane straw (350 ◦ C) and
adsorbed PO−4 P, and caused the release of P from the biochar. A con­ doped with Al. Doping with Al was performed after pyrolysis as a result
trasting relationship was observed in peanut shell biochar, which was of the saturation of the biochars with an AlCl3 solution (biochar mass:
characterized using higher Mg/P (3.46) and Ca/P (47.60) ratios and solution of 1:1000). The Al-doped biochars prepared from poultry
adsorbed PO−4 P at a level of 6.79 mg g− 1. Kizito et al. (2017) compared manure and sugarcane straw were characterized by very high PO−4 P
the adsorption of PO−4 P from digested liquid swine manure onto biochar adsorptions of 701.65 mg g− 1 and 758.96 mg g− 1, respectively. The
obtained from wood, corncobs, rice husks, and sawdust. Adsorption of authors suggested that the use of Al allows the negatively charged bio­
PO−4 P did not differ significantly between the feedstocks and ranged char surface to be coated with Al3+ cations and hydrolysis products of its
from 8.1 to 12.0 mg g− 1. The authors suggested that the main adsorption salts, thus generating a positive surface charge that facilitates the
mechanism was based on the Ca–P precipitation reaction. They also binding of the negatively charged ions. The negative charge of biochar’s
suggested that the presence of other ions, such as Na+ or K+, in the surface can be also eliminated by biochar treatment using ZnCl2 (Park
digested swine manure (from which PO4–P was adsorbed) could have et al., 2015). The maximum PO−4 P adsorption capacity was enhanced
contributed to the protonation of the negatively charged biochar surface after Zn activation to 15.46 mg kg− 1, whereas pristine biochar release P
and the formation of a local positive charge, which facilitated the (− 12.38 mg kg− 1). These examples (Table S3) show how large differ­
adsorption of negative PO−4 P on the biochar surface. PO−4 P adsorption ences in PO−4 P adsorption can be obtained using various biochar modi­
was also found to be dependent on the pyrolysis temperature at which fications. Compared with adsorption on pristine biochar, adsorption on
the biochar was produced (Fang et al., 2014; Peng et al., 2012; Trazzi engineered biochar can be more than 1000 times higher (Novais et al.,
et al., 2016) (Table S3). The higher the pyrolysis temperature is, the 2018). The utilization of metals to fabricate biochar composites forms a
higher the PO−4 P adsorption values. bridge that enables the adsorption of anions on the negatively charged
biochar surface. However, when selecting a modification method, effi­
4.6. Phosphate adsorption using engineered biochar ciency as well as the possibility of safe environmental application must
be considered. This particularly applies to modification with Al, which
As indicated for the previously discussed ions, biochar modification exhibits toxicity that can disturb plant growth and soil microbiological
using metals may be a solution for enhanced efficiency in removing processes (Shetty et al., 2021), when this type of composite is considered
PO−4 P (Zhang et al., 2021). Wang et al. (2021) proposed a solution for for use as a secondary fertilizer with controlled release of PO4. Wang
the problem of low adsorption on pristine biochar. Biochars produced et al. (2021) proposed a different solution for the problem of low
from oilseed rape straw (700 ◦ C) were modified with coal gangue. adsorption on pristine biochar. Oilseed rape straw biochar (700 ◦ C) was
Owing to this modification, adsorption of 7.9 mg g− 1 PO−4 P was ach­ modified with coal gangue. Owing to the modification used, an
ieved, which was nearly five times higher than that on the pristine adsorption of 7.9 mg g− 1 PO−4 P was achieved, which was nearly five
biochar. The authors explained the process of adsorption on this com­ times higher than that of the pristine biochar. The authors explained the
posite based on the mechanisms of ion exchange, surface precipitation, process of adsorption on this composite based on the mechanisms of ion
and electrostatic attraction. Modifications may also involve the exchange, surface precipitation, and electrostatic attraction (Table 1).

6
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

Table 1
Mechanism of NH4–N, NO3–N and PO4–P adsorption/desorption on/from biochar.

4.7. Combined adsorption 5. Nutrient desorption from pristine biochar and modified
biochar
In addition to these discussed possibilities of recovering single ions as
nutrients, it is also possible to simultaneously adsorb N and P ions on In the fabrication of biochar-based slow-release fertilizers, the pre­
biochar composites. An example is the use of Mg-modified biochar to viously described methods for enriching biochars with nutrients are
recover PO−4 P and NH−4 N from aqueous solutions based on the precipi­ important as are the processes of desorption of these nutrients, which
tation reaction (Equation (1)): enable essential elements to be transferred to the soil and subsequently
assimilated by the plants. The following section presents information
Mg2+ + NH4+ + HPO2−4 + 6 H2 O→MgNH4 PO4 *6H2 O + H + (1)
regarding the efficiency of nutrient desorption from biochar, desorption
Struvite (MgNH4PO4∙6H2O) obtained from this reaction covers the mechanisms, and factors affecting desorption.
biochar surface (Thant Zin and Kim, 2021) and can also be a source of N,
P, and Mg. Therefore, such materials can act as slow-release compound 5.1. Ammonium desorption
fertilizers. Xu et al. (2018) proposed the use of MgO-coated wood waste
biochar for the simultaneous recovery of N and P from urea. The Yin et al. (2020) used biochars produced from sewage sludge and
adsorption of N and P was 47.5 mg g− 1 and 116 mg g− 1, respectively. walnut shells at a temperature of 600 ◦ C to evaluate NH−4 N adsorp­
Gong et al. (2017) used Mg-coated biochar obtained from P. australis for tion/desorption processes. In their study, the authors investigated the
the simultaneous removal of PO−4 P and NO−3 N from water. The authors effect of pH on NH−4 N desorption from biochar. The desorption of NH−4 N
found the sorption to be 100 mg g− 1 for PO−4 P and 30 mg g− 1 for NH−4 N. gradually decreased with increasing solution pH from 2.0 to 11.5. The
Chandra et al. (2020) and Yin et al. (2018a,b) compared the adsorption maximum NH−4 N desorption of 2.2% was achieved at a pH of 2. The
of NH−4 N, PO−4 P, and NO−3 N from eutrophic water onto biochar using authors suggested (Yin et al., 2020) that NH−4 N desorption primarily
single- and three-component systems. The authors found no differences occurred based on ion exchange, which was confirmed by its greatest
in the levels of NH−4 N and PO−4 P adsorption between these systems. intensity at low pH values (pH = 2). Desorption was also observed to
Nonetheless, they observed a 50% decrease in NO−3 N adsorption in the increase slightly in a strongly alkaline environment because the OH−
three-component system. The absence of differences in NH−4 N and PO−4 P ions present in this environment electrostatically attracted NH−4 N.
desorption indicates a lack of competition between these ions, which is Desorption of NH−4 N occurred in two stages, with the first lasting 4 h and
due to their binding/interaction with different active sites. The reduced being the most intensive. Aghoghovwia et al. (2020) determined the
adsorption of NO−3 N may result from the greater affinity of PO−4 P for the NH−4 N desorption from biochars obtained from various feedstocks,
proposed functional groups, because of which the active sites are including grape pip biochar, grape skin biochar, pine wood biochar, and
occupied by competitive ions. The generation of adsorbents with the rubber tire biochar, using 2 M KCl. Desorption solutions are commonly
ability to simultaneously bind nutrients is of particular interest for the used to determine mineral nitrogen in soils (Aghoghovwia et al., 2020).
treatment of water contaminated with many nutrients, such as eutrophic Within the first 24 h, depending on the feedstock used to produce the
water or wastewater. biochars, desorption ranged as 53–60% of the initially adsorbed NH−4 N
(0.6–9.4 mg g− 1). The authors suggest that because adsorption was
based on electrostatic and van der Waals interactions, desorption was
possible through ion exchange with K+ ions (Aghoghovwia et al., 2020).
The weakest binding and hence the highest desorption, were obtained

7
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

for the grape pip biochar (60.3%) and rubber tire biochar (58.4%). The et al., 2015a, 2015b), which was associated with N losses as a result of
adsorption of NH−4 N on these biochars mainly resulted from electrostatic the release of gaseous NH3. The higher the level of biochar oxidation
interactions because a net negative charge and the presence of carboxyl achieved by increasing the H2O2 concentration and biochar contact
groups were observed; therefore, NH−4 N could be easily desorbed by K+ time, the lower the NH−4 N desorption from the biochar. As shown in the
ions (Table 1). Gong et al. (2019) investigated NH−4 N desorption from above examples, desorption of NH−4 N from biochar occurs gradually, but
biochar prepared from rice straw at temperatures of 300, 400, 600, and in most cases, maximum NH−4 N is released from the biochar matrix
800 ◦ C. The range of NH−4 N desorption was dependent on the biochar relatively quickly (from 24 to 72 h), which limits the use of biochar as a
production temperature to the greatest extent, which determined its slow/control-release fertilizer. The NH−4 N desorption process can be
physicochemical properties. Desorption of NH−4 N was highest for the controlled to some degree by the selection of substrates, pH, and
biochar produced at 800 ◦ C (60.2%), while desorption from BC600, appropriate modifications. Under environmental conditions, controlling
which had most efficiently adsorbed NH−4 N (15.61 mg g− 1), occurred at pH or moisture content can be difficult. However, the presented studies
46.2%. Nutrients physically adsorbed on the biochar surface are des­ demonstrate the potential of reusing biochars that were initially used to
orbed easily, whereas the opposite phenomenon is observed in the case remove NH−4 N from water and subsequently as fertilizer. Nevertheless, a
of chemical adsorption. Adsorption on biochars produced at high tem­ range of studies on the desorption kinetics and environmental processes
peratures usually occurs through physical sorption, which explains the that may impact the release of NH−4 N are required. NH−4 N is bound by
higher desorption from these biochars (Gong et al., 2019). Wang et al. biochars with little strength, which creates the risk that NH−4 N will be
(2015a, 2015b) studied NH−4 N desorption from maple wood biochar quickly released after the application of this type of fertilizer to soils
(500 ◦ C) in aqueous and KCl solutions. The authors reported that less (Fig. 3). Therefore, the direct application of secondary biochar is in
than 27% of the initially adsorbed NH−4 N was desorbed in water within question; currently, biochar will probably require additional treatments
24 h, whereas the use of KCl solution for extraction allowed 99% that will increase the strength of interaction with the matrix, due to
desorption to be achieved. Desorption to the KCl solution was dependent which NH−4 N will be leached more slowly and in a more controlled
on pH, and as suggested by the authors, at pH above 7, the NH−4 N ions manner. One should not exclude the direct application of secondary
were converted to volatile NH3, resulting in incomplete recovery of biochars to soils; however, this requires additional studies using
NH−4 N. A recent study by Vendra Singh et al. (2020) also reported high different feedstocks and more complex biochars, wherein bonding will
levels of urea desorption from biochar. The authors indicated that within be stronger, which will provide gradual release of NH−4 N.
72 h rice straw biochar prepared at 450 ◦ C desorbed almost 98% of urea
from the water. The desorption process was dependent on urea
adsorption and diffusion, as well as on surface parameters, such as the 5.2. Nitrate desorption
porous structure, because a microporous structure favors the storage and
retention of nutrients in biochars. Chen et al. (2017) studied NH−4 N Most studies have shown that desorption of NO−3 N from pristine
desorption from a composite of biochar (bamboo powder, 300–500 ◦ C) biochar is marginal or reaches very low values (Aghoghovwia et al.,
with Mt. Weak interactions of NH−4 N with the biochar, based on van der 2020; Hale et al., 2013). Using KCl as an extractant, Aghoghovwia et al.
Waals interactions and surface adsorption, allowed the adsorbate to be (2020) determined the desorption of NO−3 N from biochar prepared from
leached evenly and easily, with the obtained desorption results ranging plant waste (grape pip, grape skin, and pine wood) and rubber tires to
as 0.30–4.92% of the initial adsorbed amount. Desorption of NH−4 N from range 0.01–0.23%, depending on the type of biochar. According to these
biochar (maple wood, 550 ◦ C) modified with 30% H2O2 was determined authors (Aghoghovwia et al., 2020), biochar catalyzes the reactions of
by Wang et al. (2015a, 2015b) in two stages. A sample was first nitrogen compounds owing to the transfer of electrons in the aromatic
extracted with water and then with KCl solution. NH−4 N desorption in structure of the biochar, which causes the formation of nitrogen forms
the water was very low, accounting for less than 27% of the adsorbed other than NO−3 N. The biochar reduced NO−3 N, probably to nitrites or
NH−4 N for the biochar with an unadjusted pH and less than 16% for the gaseous forms, which resulted in reduced desorption. Hale et al. (2013)
biochar with an adjusted pH (=7). At a low pH, KCl application allowed observed no desorption of NO−3 N from cacao shell biochar produced at a
the adsorption of NH−4 N to recover almost completely. However, this temperature of 350 ◦ C. Chintala et al. (2013) observed that the
was not possible at pH greater than 7, as explained by the authors (Wang desorption of NO−3 N was significantly dependent on the solution pH and
amount of previously adsorbed NO−3 N. In addition, biochar activation

Fig. 3. Biochar-based fertilizer obtained through adsorption from contaminated wastes.

8
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

significantly affects desorption (Chintala et al., 2013). Biochars pre­ composites produced using biochar from ground coffee bean waste py­
pared from Pinus ponderosa L. wood chips and Panicum virgatum L. at rolyzed at a temperature of 500 ◦ C. The authors used pre-pyrolysis
650 ◦ C were activated using HCl. A study (Chintala et al., 2013) revealed modification and impregnated ground coffee bean waste with MgCl2
that the desorption of NO−3 N from activated biochars was clearly higher solution. Desorption of PO4–P occurred at the same rate in both water
than that from nonactivated biochars. Desorption from activated Pinus and the 2% citric acid solution. The Mg-biochar composite released
ponderosa-derived biochar was nearly 75%, whereas that from approximately 70% of the initially adsorbed PO4–P within 12 h. At the
non-activated biochar was 65%. The authors observed that the desorp­ same time, desorption in 2% citric acid resulted in the release of 90% of
tion of NO−3 N was higher at pH 9 than at pH 4. Increased desorption at a adsorbed PO4–P. The obtained results show that an acidic environment
high pH can be associated with the formation of a negative charge on the (2% citric acid, pH 2.5) was more beneficial for the desorption of PO4–P,
biochar surface, which acts repulsively on negatively charged NO−3 N. which was associated with the higher solubility of Mg–P (Mg3(PO4)2),
Moreover, a higher concentration of OH− ions at a high pH can cause the MgHPO4) at acidic pH. The desorption of PO4–P is strongly dependent
displacement of NO−3 N from the sorption sites. Gong et al. (2019) found on the pH of the desorbing solution, which limits soil environmental
that biochar production temperature plays an important role in the conditions, but also on pyrolysis conditions or the initial concentration
desorption of NO−3 N from biochars. The authors compared the desorp­ used in the adsorption, which makes it possible to design biochar ad­
tion of NO−3 N from rice straw biochars produced at 300, 400, 600, and sorbents and adjust their properties to the soil requirements.
800 ◦ C. As previously mentioned, physical adsorption predominantly
occurs on biochars produced at higher temperatures (>600 ◦ C), which is 6. Economic evaluation
a process that weakly binds NO−3 N ions, and thus, such adsorbed nu­
trients are desorbed more easily. Comparing NO−3 N desorption from the The production costs of biochar depend on many factors i.e., feed­
biochars prepared at two extreme temperatures, 300 ◦ C and 800 ◦ C, stock used, production automation, energy costs and logistics, and range
Gong et al. (2019) found desorption to be nearly twice that of the bio­ from 30 to 800 € t− 1 (Maroušek and Trakal, 2022). The energy balance
char produced at a temperature of 800 ◦ C (64% of the initially adsorbed of biochar can be regulated by the choice of pyrolysis conditions, e.g., at
NO−3 N). The present study demonstrates that desorption of NO−3 N from temperature below 500 ◦ C hemicellulose and lignin decomposition is an
biochars can be substantially lower than that of NH+ 4 , and that it is exothermic reaction and cellulose is an endothermic reaction, while at
strongly dependent on the pH of the desorbing solution. This may cause higher temperatures the opposit trend occurs (Kwapinski, 2019). The
problems when trying to obtain secondary controlled-release fertilizers, biochar production costs seem to be lower compared to synthetic fer­
especially compound fertilizers. Adjustment of the relevant parameters tilizer prices. According to the World Bank (April 2022) the price of
that consider the desorption of different ions may require additional nitrogen fertilizers (urea) is as high as 925 US $ t-1 and phosphorus
research that would include biochars produced at different temperatures fertilizers 954 US $ t-1. It is estimated that such high prices of fertilizers
and their modifications. will remain or even increase due to high canning of energy resources
(coal, natural gas) and disturbances in the supply chain of raw materials
5.3. Phosphorus desorption (phosphorus, ammonia, sulfur). The nutrient losses from synthetic fer­
tilizers can be as high as 50% (Dimkpa et al., 2020), which results in low
Studies have also been conducted on PO−4 P ions. Yin et al. (2020) utilization of fertilizers and an increase in the costs of agricultural pro­
investigated the desorption of PO4–P ions from biochar produced from duction. On the other hand significant changes are achieved in crop
municipal sewage sludge and walnut shells at a temperature of 600 ◦ C. productivity when high biochar (pristine) rates, ranging 2.5–20 ton
Desorption of PO4–P was clearly dependent on the initial PO−4 P con­ ha− 1, are applied to soil (Joseph et al., 2013). Enriching biochar with
centration used in the adsorption process, which is understandable nutrients can lower this dosage level. Moreover, a one-time application
because it determines the amount of desorbed nutrients. This also in­ of biochar can provide fertilization for several seasons, reducing agri­
dicates that desorption depends on the amount of adsorbed nutrients, cultural and energy costs (Shaviv, 2001). However, attention should also
which can form the basis for preparing fertilizers with a specific amount be paid to the fact that some of the physicochemical properties of bio­
of released nutrients. The highest desorption level was achieved at pH 2, chars (water-binding capacity, porosity, specific area) have changed due
where 7.58% of the initially adsorbed PO−4 P was desorbed from the to aging in soil (Maroušek and Trakal, 2022). Furthermore, the final cost
biochar within 24 h. However, the authors (Yin et al., 2020) observed of secondary biochar is related to the cost of modifying factor e.g., metal
that with the adsorption of a lower amount of PO−4 P (14.2 mg g− 1), salts (Maroušek and Gavurová, 2022). The growing production of
desorption was not proportionate, at only 2.1%. These differences in wastewater (Jones et al., 2021) not only has an adverse impact on the
desorption were linked to the easier desorption of nutrients from the environment (e.g., eutrophication), but also requires the large financial
successive adsorption layers and the large difference between the con­ outlays (Muga and Mihelcic, 2008). Additionally, the average fee for
centration of the adsorbed ions and the aqueous solution used as a biowaste landfilling in the EU is currently about 30 € t− 1 (Favoino and
desorbate. Moreover, PO−4 P formed on the biochar surface were spar­ Giavini, n.d.). Biochar-based fertilizer production can be profitable
ingly soluble in neutral solution; hence, after the application of such segment if: 1) waste are used for biochar production; 2) pyrolysis pro­
biochar in soil, their biological availability can also be limited. With cess involve waste heat and; 3) the obtained biochar is rich in nutrients
increasing pH, the amount of PO−4 P desorbed from the biochar decreased (from waste); 4) biochar-based fertilizers are used for targeted fertil­
because of the poor solubility of PO−4 P in a basic environment and ization (Stávková and Maroušek, 2021). Nevertheless, further studies
electrostatic repulsion between OH− and PO−4 P, which did not promote are required to evaluate the economic and agricultural profitability of
the leaching of these ions into the solution. Trazzi et al. (2016) also using biochar-based fertilizers.
investigated the effect of the initial PO−4 P concentration and biochar
production conditions on PO4–P desorption. Biochar was prepared from 7. Conclusions and recommendations for further research
sugarcane and Miscanthus at 300, 500, and 700 ◦ C. The authors found
that PO−4 P desorption was dependent on the initial concentration of the Utilization of the characteristic properties of biochar, that is, high
solution used for adsorption and the biochar pyrolysis temperature. specific surface area, porosity, and the presence of various functional
Desorption from sugarcane biochar (700 ◦ C) ranged as 22.4–45.6% for groups, allows us to obtain an excellent matrix for recovering nutrients
initial solution concentrations of 25 and 400 ppm. With increasing py­ from wastewater. Biochar adsorption processes have an advantage over
rolysis temperature ranging as 300–700 ◦ C, a 3–12% increase in other methods, especially because of the simplicity of the procedure and
desorption was observed. Shin et al. (2020) studied the effect of different cost-effective and eco-friendly adsorbent. The main mechanisms that
extractants on the range of PO4–P desorption from Mg-biochar determine adsorption on pristine biochar are electrostatic adsorption,

9
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

physical adsorption, and surface adsorption on macro- and micropores. Aghoghovwia, M.P., Hardie, A.G., Rozanov, A.B., 2020. Characterisation, adsorption and
desorption of ammonium and nitrate of biochar derived from different feedstocks.
The maximum adsorption capacity strongly depends on the biochar type
Environ. Technol. 43 (5), 774–787. https://doi.org/10.1080/
(i.e., biochar feedstock and pyrolysis temperature). Pristine biochar has 09593330.2020.1804466, 0(0).
some limitations (for example, negatively charged surface, low specific Alsewaileh, A.S., Usman, A.R., Al-Wabel, M.I., 2019. Effects of pyrolysis temperature on
surface area, and lack of acidic functional groups); therefore, it is rec­ nitrate-nitrogen (NO3− -N) and bromate (BrO3− ) adsorption onto date palm
biochar. J. Environ. Manag. 237, 289–296. https://doi.org/10.1016/j.
ommended to modify biochar. Metal modification (e.g., Mg, Al, Fe, Zn), jenvman.2019.02.045.
acid-base treatment, H2O2 modification or combining biochars with Antonangelo, J.A., Sun, X., Zhang, H., 2020. The roles of co-composted biochar (COMBI)
minerals (e.g., bentonite, montmorillonite) effectively enhanced the in improving soil quality, crop productivity, and toxic metal amelioration.
J. Environ. Manag. 277, 111443. https://doi.org/10.1016/j.jenvman.2020.111443.
adsorptive properties of biochar. Engineered biochar exhibits selective Bashar, R., Gungor, K., Karthikeyan, K.G., Barak, P., 2018. Cost effectiveness of
adsorption, which is important for adsorption from multicomponent phosphorus removal processes in municipal wastewater treatment. Chemosphere
solutions. Adsorption on EngBCs depends on additional mechanisms, 197, 280–290. https://doi.org/10.1016/j.chemosphere.2017.12.169.
Beckinghausen, A., Odlare, M., Thorin, E., Schwede, S., 2020. From removal to recovery:
particularly ligand and ion exchange, hydrogen bonding, and surface an evaluation of nitrogen recovery techniques from wastewater. Appl. Energy 263,
precipitation. Desorption processes are strongly influenced by biochar 114616. https://doi.org/10.1016/j.apenergy.2020.114616.
type, pH, and type of desorbing solution. Electrostatic interaction and Bhatnagar, A., Sillanpää, M., 2011. A review of emerging adsorbents for nitrate removal
from water. Chem. Eng. J. 168 (2), 493–504. https://doi.org/10.1016/j.
ion exchange are the main mechanism effected desorption processes. cej.2011.01.103.
Direct application is difficult on an agriculture scale and may impose Cao, X., Ro, K.S., Libra, J.A., Kammann, C.I., Lima, I., Berge, N., Li, L., Li, Y., Chen, N.,
restrictions in the secondary biochar application. Future research should Yang, J., Deng, B., Mao, J., 2013. Effects of biomass types and carbonization
conditions on the chemical characteristics of hydrochars. J. Agric. Food Chem. 61
focus on a better understanding of the desorption kinetics and envi­
(39), 9401–9411. https://doi.org/10.1021/jf402345k.
ronmental processes that may impact the release of nutrients in the soil. Chandra, S., Medha, I., Bhattacharya, J., 2020. Potassium-iron rice straw biochar
Direct application of secondary biochar to soil can cause unwanted composite for sorption of nitrate, phosphate, and ammonium ions in soil for timely
phenomena, such as the uncontrolled loss of nutrients, providing a lower and controlled release. Sci. Total Environ. 712, 136337. https://doi.org/10.1016/j.
scitotenv.2019.136337.
utilization rate and higher application costs. Furthermore, nutrient high Chen, L., Chen, X.L., Zhou, C.H., Yang, H.M., Ji, S.F., Tong, D.S., Zhong, Z.K., Yu, W.H.,
concentration in soil manifested as specific toxicity may cause plant Chu, M.Q., 2017. Environmental-friendly montmorillonite-biochar composites:
stress. Nutrients released to the soil too quickly from the secondary facile production and tunable adsorption-release of ammonium and phosphate.
J. Clean. Prod. 156, 648–659. https://doi.org/10.1016/j.jclepro.2017.04.050.
biochar will not be fully assimilated by the plants, and an excess of Chintala, R., Mollinedo, J., Schumacher, T.E., Papiernik, S.K., Malo, D.D., Clay, D.E.,
nutrients can lead to groundwater contamination, and eutrophication of Kumar, S., Gulbrandson, D.W., 2013. Nitrate sorption and desorption in biochars
surface waters. The direct application of secondary biochars to soils from fast pyrolysis. Microporous Mesoporous Mater. 179, 250–257. https://doi.org/
10.1016/j.micromeso.2013.05.023.
should not be ruled out; however, additional studies are required using Chrispim, M.C., Scholz, M., Nolasco, M.A., 2019. Phosphorus recovery from municipal
different feedstocks and more complex biochars in which bonding will wastewater treatment: critical review of challenges and opportunities for developing
be stronger, which will provide gradual nutrient release. Evaluation of countries. J. Environ. Manag. 248, 109268. https://doi.org/10.1016/j.
jenvman.2019.109268.
the economic benefits of secondary biochar application requires further Coskun, D., Britto, D.T., Shi, W., Kronzucker, H.J., 2017. Nitrogen transformations in
analyses. modern agriculture and the role of biological nitrification inhibition. Nature Plants
3, 17074. https://doi.org/10.1038/nplants.2017.74.
Diaz-Elsayed, N., Rezaei, N., Guo, T., Mohebbi, S., Zhang, Q., 2019. Wastewater-based
Credir author statement resource recovery technologies across scale: a review. Resour. Conserv. Recycl. 145,
94–112. https://doi.org/10.1016/j.resconrec.2018.12.035.
Marta Marcińczyk: Writing – original draft, Yong Sik Ok: Review and Dimkpa, C.O., Fugice, J., Singh, U., Lewis, T.D., 2020. Development of fertilizers for
enhanced nitrogen use efficiency – trends and perspectives. Sci. Total Environ. 731,
editing, Patryk Oleszczuk: Conceptualization, Writing – original draft, 139113. https://doi.org/10.1016/j.scitotenv.2020.139113.
Review and editing, Supervision. El-Naggar, A., Lee, S.S., Awad, Y.M., Yang, X., Ryu, C., Rizwan, M., Rinklebe, J.,
Tsang, D.C.W., Ok, Y.S., 2018. Influence of soil properties and feedstocks on biochar
potential for carbon mineralization and improvement of infertile soils. Geoderma
Declaration of competing interest 332, 100–108. https://doi.org/10.1016/j.geoderma.2018.06.017.
Fang, C., Zhang, T., Li, P., Jiang, R., Wang, Y., 2014. Application of magnesium modified
corn biochar for phosphorus removal and recovery from swine wastewater. Int. J.
The authors declare that they have no known competing financial Environ. Res. Publ. Health 11 (9), 9217–9237. https://doi.org/10.3390/
interests or personal relationships that could have appeared to influence ijerph110909217.
Favoino, E., & Giavini, M. (n.d.). This Report Was Commissioned by the Bio-Based
the work reported in this paper.
Industries Consortium (BIC). 50.
Fidel, R.B., Laird, D.A., Spokas, K.A., 2018. Sorption of ammonium and nitrate to
Data availability biochars is electrostatic and pH-dependent. Sci. Rep. 8 (1), 17627. https://doi.org/
10.1038/s41598-018-35534-w.
Gai, X., Wang, H., Liu, J., Zhai, L., Liu, S., Ren, T., Liu, H., 2014. Effects of feedstock and
No data was used for the research described in the article. pyrolysis temperature on biochar adsorption of ammonium and nitrate. PLoS One 9
(12), e113888. https://doi.org/10.1371/journal.pone.0113888.
Ghodake, G.S., Shinde, S.K., Kadam, A.A., Saratale, R.G., Saratale, G.D., Kumar, M.,
Acknowledgments
Palem, R.R., Al-Shwaiman, H.A., Elgorban, A.M., Syed, A., Kim, D.-Y., 2021. Review
on biomass feedstocks, pyrolysis mechanism and physicochemical properties of
This work was supported by the National Science Center (Poland) biochar: state-of-the-art framework to speed up vision of circular bioeconomy.
under the SHENG 1 grant [grant number UMO-2018/30/Q/ST10/ J. Clean. Prod. 297, 126645. https://doi.org/10.1016/j.jclepro.2021.126645.
Gong, Y.-P., Ni, Z.-Y., Xiong, Z.-Z., Cheng, L.-H., Xu, X.-H., 2017. Phosphate and
00060]. ammonium adsorption of the modified biochar based on Phragmites australis after
phytoremediation. Environ. Sci. Pollut. Control Ser. 24 (9), 8326–8335. https://doi.
org/10.1007/s11356-017-8499-2.
Appendix A. Supplementary data Gong, H., Tan, Z., Zhang, L., Huang, Q., 2019. Preparation of biochar with high
absorbability and its nutrient adsorption–desorption behaviour. Sci. Total Environ.
Supplementary data to this article can be found online at https://doi. 694, 133728. https://doi.org/10.1016/j.scitotenv.2019.133728.
Hafshejani, L.D., Hooshmand, A., Naseri, A.A., Mohammadi, A.S., Abbasi, F.,
org/10.1016/j.chemosphere.2022.135310.
Bhatnagar, A., 2016. Removal of nitrate from aqueous solution by modified
sugarcane bagasse biochar. Ecol. Eng. 95, 101–111. https://doi.org/10.1016/j.
References ecoleng.2016.06.035.
Hale, S.E., Alling, V., Martinsen, V., Mulder, J., Breedveld, G.D., Cornelissen, G., 2013.
The sorption and desorption of phosphate-P, ammonium-N and nitrate-N in cacao
Agbede, T.M., Oyewumi, A., 2022. Benefits of biochar, poultry manure and
shell and corn cob biochars. Chemosphere 91 (11), 1612–1619. https://doi.org/
biochar–poultry manure for improvement of soil properties and sweet potato
10.1016/j.chemosphere.2012.12.057.
productivity in degraded tropical agricultural soils. Resour. Environ. Sustain. 7,
100051. https://doi.org/10.1016/j.resenv.2022.100051.

10
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

Han, B., Butterly, C., Zhang, W., He, J., Chen, D., 2021. Adsorbent materials for Lou, K., Rajapaksha, A.U., Ok, Y.S., Chang, S.X., 2016. Sorption of copper(II) from
ammonium and ammonia removal: a review. J. Clean. Prod. 283, 124611. https:// synthetic oil sands process-affected water (OSPW) by pine sawdust biochars: effects
doi.org/10.1016/j.jclepro.2020.124611. of pyrolysis temperature and steam activation. J. Soils Sediments 16 (8), 2081–2089.
Hu, X., Zhang, X., Ngo, H.H., Guo, W., Wen, H., Li, C., Zhang, Y., Ma, C., 2020. https://doi.org/10.1007/s11368-016-1382-9.
Comparison study on the ammonium adsorption of the biochars derived from Luo, Kun, Pang, Ya, Wang, Dongbo, Li, Xue, Wang, Liping, Lei, Min, Huang, Qi, Yang, Qi,
different kinds of fruit peel. Sci. Total Environ. 707, 135544. https://doi.org/ 2021. A critical review on the application of biochar in environmental pollution
10.1016/j.scitotenv.2019.135544. remediation: role of persistent free radicals (PFRs). J. Environ. Sci. (China) 108,
Huang, Y., Lee, X., Grattieri, M., Yuan, M., Cai, R., Macazo, F.C., Minteer, S.D., 2020a. 201–216. https://doi.org/10.1016/j.jes.2021.02.021.
Modified biochar for phosphate adsorption in environmentally relevant conditions. Marcińczyk, M., Oleszczuk, P., 2022. Biochar and engineered biochar as slow- and
Chem. Eng. J. 380, 122375. https://doi.org/10.1016/j.cej.2019.122375. controlled-release fertilizers. J. Clean. Prod. 339, 130685. https://doi.org/10.1016/
Huang, Y., Lee, X., Grattieri, M., Yuan, M., Cai, R., Macazo, F.C., Minteer, S.D., 2020b. j.jclepro.2022.130685.
Modified biochar for phosphate adsorption in environmentally relevant conditions. Maroušek, J., Gavurová, B., 2022. Recovering phosphorous from biogas fermentation
Chem. Eng. J. 380, 122375. https://doi.org/10.1016/j.cej.2019.122375. residues indicates promising economic results. Chemosphere 291 (Pt 1), 133008.
Ippolito, J.A., Cui, L., Kammann, C., Wrage-Mönnig, N., Estavillo, J.M., Fuertes- https://doi.org/10.1016/j.chemosphere.2021.133008.
Mendizabal, T., Cayuela, M.L., Sigua, G., Novak, J., Spokas, K., Borchard, N., 2020. Maroušek, J., Trakal, L., 2022. Techno-economic analysis reveals the untapped potential
Feedstock choice, pyrolysis temperature and type influence biochar characteristics: a of wood biochar. Chemosphere 291, 133000. https://doi.org/10.1016/j.
comprehensive meta-data analysis review. Biochar 2 (4), 421–438. https://doi.org/ chemosphere.2021.133000.
10.1007/s42773-020-00067-x. Mishra, P.C., Patel, R.K., 2009. Use of agricultural waste for the removal of nitrate-
Ismadji, S., Tong, D.S., Soetaredjo, F.E., Ayucitra, A., Yu, W.H., Zhou, C.H., 2016. nitrogen from aqueous medium. J. Environ. Manag. 90 (1), 519–522. https://doi.
Bentonite hydrochar composite for removal of ammonium from Koi fish tank. Appl. org/10.1016/j.jenvman.2007.12.003.
Clay Sci. 119, 146–154. https://doi.org/10.1016/j.clay.2015.08.022. Mohan, D., Sarswat, A., Ok, Y.S., Pittman, C.U., 2014. Organic and inorganic
Jena, J., Das, T., Sarkar, U., 2021. Explicating proficiency of waste biomass-derived contaminants removal from water with biochar, a renewable, low cost and
biochar for reclaiming phosphate from source-separated urine and its application as sustainable adsorbent – a critical review. Bioresour. Technol. 160, 191–202. https://
a phosphate biofertilizer. J. Environ. Chem. Eng. 9 (1), 104648. https://doi.org/ doi.org/10.1016/j.biortech.2014.01.120.
10.1016/j.jece.2020.104648. Muga, H.E., Mihelcic, J.R., 2008. Sustainability of wastewater treatment technologies.
Jie, C., Jing-zhang, C., Man-zhi, T., Zi-tong, G., 2002. Soil degradation: a global problem J. Environ. Manag. 88 (3), 437–447. https://doi.org/10.1016/j.
endangering sustainable development. J. Geogr. Sci. 12 (2), 243–252. https://doi. jenvman.2007.03.008.
org/10.1007/BF02837480. Muo, I., Azeez, A., 2019. Green entrepreneurship: literature review and agenda for future
Jing, H.-P., Li, Y., Wang, X., Zhao, J., Xia, S., 2019. Simultaneous recovery of phosphate, research. https://doi.org/10.2478/ijek-2019-0007.
ammonium and humic acid from wastewater using a biochar supported Mg(OH)2/ Novais, S.V., Zenero, M.D.O., Barreto, M.S.C., Montes, C.R., Cerri, C.E.P., 2018.
bentonite composite. Environ. Sci.: Water Res. Technol. 5 (5), 931–943. https://doi. Phosphorus removal from eutrophic water using modified biochar. Sci. Total
org/10.1039/C8EW00952J. Environ. 633, 825–835. https://doi.org/10.1016/j.scitotenv.2018.03.246.
Jones, E.R., van Vliet, M.T.H., Qadir, M., Bierkens, M.F.P., 2021. Country-level and Oldfield, T.L., Sikirica, N., Mondini, C., López, G., Kuikman, P.J., Holden, N.M., 2018.
gridded estimates of wastewater production, collection, treatment and reuse. Earth Biochar, compost and biochar-compost blend as options to recover nutrients and
Syst. Sci. Data 13 (2), 237–254. https://doi.org/10.5194/essd-13-237-2021. sequester carbon. J. Environ. Manag. 218, 465–476. https://doi.org/10.1016/j.
Joseph, S., Graber, E.R., Chia, C., Munroe, P., Donne, S., Thomas, T., Nielsen, S., jenvman.2018.04.061.
Marjo, C., Rutlidge, H., Pan, G.X., Li, L., Taylor, P., Rawal, A., Hook, J., 2013. Panwar, N.L., Pawar, A., 2022. Influence of activation conditions on the physicochemical
Shifting paradigms: development of high-efficiency biochar fertilizers based on properties of activated biochar: a review. Biomass Convers. Biorefin. 12 (3),
nano-structures and soluble components. Carbon Manag. 4 (3), 323–343. https:// 925–947. https://doi.org/10.1007/s13399-020-00870-3.
doi.org/10.4155/cmt.13.23. Park, J.H., Ok, Y.S., Kim, S.H., Cho, J.S., Heo, J.S., Delaune, R.D., Seo, D.C., 2015.
Jung, K.-W., Hwang, M.-J., Ahn, K.-H., Ok, Y.-S., 2015. Kinetic study on phosphate Evaluation of phosphorus adsorption capacity of sesame straw biochar on aqueous
removal from aqueous solution by biochar derived from peanut shell as renewable solution: influence of activation methods and pyrolysis temperatures. Environ.
adsorptive media. Int. J. Environ. Sci. Technol. 12 (10), 3363–3372. https://doi.org/ Geochem. Health 37 (6), 969–983. https://doi.org/10.1007/s10653-015-9709-9.
10.1007/s13762-015-0766-5. Pathy, A., Ray, J., Paramasivan, B., 2021. Challenges and opportunities of nutrient
Kavitha, B., Reddy, P.V.L., Kim, B., Lee, S.S., Pandey, S.K., Kim, K.-H., 2018. Benefits and recovery from human urine using biochar for fertilizer applications. J. Clean. Prod.
limitations of biochar amendment in agricultural soils: a review. J. Environ. Manag. 304, 127019 https://doi.org/10.1016/j.jclepro.2021.127019.
227, 146–154. https://doi.org/10.1016/j.jenvman.2018.08.082. Peng, F., He, P.-W., Luo, Y., Lu, X., Liang, Y., Fu, J., 2012. Adsorption of phosphate by
Kizito, S., Wu, S., Kipkemoi Kirui, W., Lei, M., Lu, Q., Bah, H., Dong, R., 2015. Evaluation biomass char deriving from fast pyrolysis of biomass waste. Clean: Soil, Air, Water
of slow pyrolyzed wood and rice husks biochar for adsorption of ammonium 40 (5), 493–498. https://doi.org/10.1002/clen.201100469.
nitrogen from piggery manure anaerobic digestate slurry. Sci. Total Environ. 505, Peng, Y., Sun, Y., Hanif, A., Shang, J., Shen, Z., Hou, D., Zhou, Y., Chen, Q., Ok, Y.S.,
102–112. https://doi.org/10.1016/j.scitotenv.2014.09.096. Tsang, D.C.W., 2021. Design and fabrication of exfoliated Mg/Al layered double
Kizito, S., Luo, H., Wu, S., Ajmal, Z., Lv, T., Dong, R., 2017. Phosphate recovery from hydroxides on biochar support. J. Clean. Prod. 289, 125142. https://doi.org/
liquid fraction of anaerobic digestate using four slow pyrolyzed biochars: dynamics 10.1016/j.jclepro.2020.125142.
of adsorption, desorption and regeneration. J. Environ. Manag. 201, 260–267. Qasem, N.A.A., Mohammed, R.H., Lawal, D.U., 2021. Removal of heavy metal ions from
https://doi.org/10.1016/j.jenvman.2017.06.057. wastewater: a comprehensive and critical review. NPJ Clean Water 4 (1), 1–15.
Klein, N., Ramos, T.B., Deutz, P., 2022. Advancing the circular economy in public sector https://doi.org/10.1038/s41545-021-00127-0.
organisations: employees’ perspectives on practices. Circ. Econ. Sustain. 2, 759–781. Santos, F.M., Pires, J.C.M., 2018. Nutrient recovery from wastewaters by microalgae and
https://doi.org/10.1007/s43615-021-00044-x. its potential application as bio-char. Bioresour. Technol. 267, 725–731. https://doi.
Kookana, R.S., Sarmah, A.K., Van Zwieten, L., Krull, E., Singh, B., 2011. Chapter three - org/10.1016/j.biortech.2018.07.119.
biochar application to soil: agronomic and environmental benefits and unintended Shaviv, A., 2001. Advances in controlled-release fertilizers. In: Advances in Agronomy,
consequences. In: Sparks, D.L. (Ed.), Advances in Agronomy, vol. 112. Academic vol. 71. Academic Press, pp. 1–49. https://doi.org/10.1016/S0065-2113(01)71011-
Press, pp. 103–143. https://doi.org/10.1016/B978-0-12-385538-1.00003-2. 5.
Krasucka, P., Pan, B., Sik Ok, Y., Mohan, D., Sarkar, B., Oleszczuk, P., 2021. Engineered Shepherd, J.G., Joseph, S., Sohi, S.P., Heal, K.V., 2017. Biochar and enhanced phosphate
biochar – a sustainable solution for the removal of antibiotics from water. Chem. capture: mapping mechanisms to functional properties. Chemosphere 179, 57–74.
Eng. J. 405, 126926. https://doi.org/10.1016/j.cej.2020.126926. https://doi.org/10.1016/j.chemosphere.2017.02.123.
Kwapinski, W., 2019. 2—char production technology. In: Jeguirim, M., Limousy, L. Shetty, R., Vidya, C.S.-N., Prakash, N.B., Lux, A., Vaculík, M., 2021. Aluminum toxicity in
(Eds.), Char and Carbon Materials Derived from Biomass. Elsevier, pp. 39–68. plants and its possible mitigation in acid soils by biochar: a review. Sci. Total
https://doi.org/10.1016/B978-0-12-814893-8.00002-X. Environ. 765, 142744. https://doi.org/10.1016/j.scitotenv.2020.142744.
Lee, J., Sarmah, A.K., Kwon, E.E., 2019. Chapter 1—production and formation of Shin, H., Tiwari, D., Kim, D.-J., 2020. Phosphate adsorption/desorption kinetics and P
biochar. In: Ok, Y.S., Tsang, D.C.W., Bolan, N., Novak, J.M. (Eds.), Biochar from bioavailability of Mg-biochar from ground coffee waste. J. Water Proc. Eng. 37,
Biomass and Waste. Elsevier, pp. 3–18. https://doi.org/10.1016/B978-0-12-811729- 101484. https://doi.org/10.1016/j.jwpe.2020.101484.
3.00001-7. Smol, M., 2021. Transition to circular economy in the fertilizer sector—analysis of
Li, K., Liu, Q., Fang, F., Luo, R., Lu, Q., Zhou, W., Huo, S., Cheng, P., Liu, J., Addy, M., recommended directions and end-users’ perception of waste-based products in
Chen, P., Chen, D., Ruan, R., 2019. Microalgae-based wastewater treatment for Poland. Energies 14 (14), 4312. https://doi.org/10.3390/en14144312.
nutrients recovery: a review. Bioresour. Technol. 291, 121934. https://doi.org/ Song, S., Arora, S., Laserna, A.K.C., Shen, Y., Thian, B.W.Y., Cheong, J.C., Tan, J.K.N.,
10.1016/j.biortech.2019.121934. Chiam, Z., Fong, S.L., Ghosh, S., Ok, Y.S., Li, S.F.Y., Tan, H.T.W., Dai, Y., Wang, C.-
Lim, T.C., Welty, C., 2018. Assessing variability and uncertainty in green infrastructure H., 2020. Biochar for urban agriculture: impacts on soil chemical characteristics and
planning using a high-resolution surface-subsurface hydrological model and site- on Brassica rapa growth, nutrient content and metabolism over multiple growth
monitored flow data. Front. Built Environ. 4. https://www.frontiersin.org/art cycles. Sci. Total Environ. 727, 138742. https://doi.org/10.1016/j.
icle/10.3389/fbuil.2018.00071. scitotenv.2020.138742.
Liu, Z., Xue, Y., Gao, F., Cheng, X., Yang, K., 2016. Removal of ammonium from aqueous Stávková, J., Maroušek, J., 2021. Novel sorbent shows promising financial results on P
solutions using alkali-modified biochars. Chem. Speciat. Bioavailab. 28 (1–4), recovery from sludge water. Chemosphere 276, 130097. https://doi.org/10.1016/j.
26–32. https://doi.org/10.1080/09542299.2016.1142833. chemosphere.2021.130097.
Loganathan, P., Vigneswaran, S., Kandasamy, J., Bolan, N.S., 2014. Removal and Su, J., Weng, X., Luo, Z., Huang, H., Wang, W., 2021. Impact of biochar on soil
recovery of phosphate from water using sorption. Crit. Rev. Environ. Sci. Technol. 44 properties, pore water properties, and available cadmium. Bull. Environ. Contam.
(8), 847–907. https://doi.org/10.1080/10643389.2012.741311. Toxicol. 107 (3), 544–552. https://doi.org/10.1007/s00128-021-03259-8.

11
M. Marcińczyk et al. Chemosphere 306 (2022) 135310

Takaya, C.A., Fletcher, L.A., Singh, S., Anyikude, K.U., Ross, A.B., 2016. Phosphate and Yaashikaa, P.R., Senthil Kumar, P., Varjani, S.J., Saravanan, A., 2019. Advances in
ammonium sorption capacity of biochar and hydrochar from different wastes. production and application of biochar from lignocellulosic feedstocks for
Chemosphere 145, 518–527. https://doi.org/10.1016/j.chemosphere.2015.11.052. remediation of environmental pollutants. Bioresour. Technol. 292, 122030. https://
Tang, Y., Alam, M.S., Konhauser, K.O., Alessi, D.S., Xu, S., Tian, W., Liu, Y., 2019. doi.org/10.1016/j.biortech.2019.122030.
Influence of pyrolysis temperature on production of digested sludge biochar and its Yang, J., Li, H., Zhang, D., Wu, M., Pan, B., 2017. Limited role of biochars in nitrogen
application for ammonium removal from municipal wastewater. J. Clean. Prod. 209, fixation through nitrate adsorption. Sci. Total Environ. 592, 758–765. https://doi.
927–936. https://doi.org/10.1016/j.jclepro.2018.10.268. org/10.1016/j.scitotenv.2016.10.182.
Tondera, K., Blecken, G., Tournebize, J., Mander, Ü., Tanner, C., 2018. Nutrient removal Yang, H.I., Lou, K., Rajapaksha, A.U., Ok, Y.S., Anyia, A.O., Chang, S.X., 2018.
from variable stormwater flows, pp. 31–55. https://doi.org/10.1007/978-3-319- Adsorption of ammonium in aqueous solutions by pine sawdust and wheat straw
70013-7_3. biochars. Environ. Sci. Pollut. Control Ser. 25 (26), 25638–25647. https://doi.org/
Trazzi, P.A., Leahy, J.J., Hayes, M.H.B., Kwapinski, W., 2016. Adsorption and desorption 10.1007/s11356-017-8551-2.
of phosphate on biochars. J. Environ. Chem. Eng. 4 (1), 37–46. https://doi.org/ Yang, F., Zhang, S., Sun, Y., Tsang, D.C.W., Cheng, K., Ok, Y.S., 2019. Assembling
10.1016/j.jece.2015.11.005. biochar with various layered double hydroxides for enhancement of phosphorus
Usman, A.R.A., Ahmad, M., El-Mahrouky, M., Al-Omran, A., Ok, Y.S., Sallam, A. Sh, El- recovery. J. Hazard Mater. 365, 665–673. https://doi.org/10.1016/j.
Naggar, A.H., Al-Wabel, M.I., 2016. Chemically modified biochar produced from jhazmat.2018.11.047.
conocarpus waste increases NO3 removal from aqueous solutions. Environ. Yang, H., Ye, S., Zeng, Z., Zeng, G., Tan, X., Xiao, R., Wang, J., Song, B., Du, L., Qin, M.,
Geochem. Health 38 (2), 511–521. https://doi.org/10.1007/s10653-015-9736-6. Yang, Y., Xu, F., 2020. Utilization of biochar for resource recovery from water: a
Vaneeckhaute, C., Lebuf, V., Michels, E., Belia, E., Vanrolleghem, P.A., Tack, F.M.G., review. Chem. Eng. J. 397, 125502. https://doi.org/10.1016/j.cej.2020.125502.
Meers, E., 2017. Nutrient recovery from digestate: systematic technology review and Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X., Pullammanappallil, P., Yang, L.,
product classification. Waste Biomass Valorizat. 8 (1), 21–40. 2011. Removal of phosphate from aqueous solution by biochar derived from
Vendra Singh, S., Chaturvedi, S., Dhyani, V.C., Kasivelu, G., 2020. Pyrolysis temperature anaerobically digested sugar beet tailings. J. Hazard Mater. 190 (1), 501–507.
influences the characteristics of rice straw and husk biochar and sorption/desorption https://doi.org/10.1016/j.jhazmat.2011.03.083.
behaviour of their biourea composite. Bioresour. Technol. 314, 123674. https://doi. Ye, Y., Ngo, H.H., Guo, W., Chang, S.W., Nguyen, D.D., Zhang, X., Zhang, J., Liang, S.,
org/10.1016/j.biortech.2020.123674. 2020. Nutrient recovery from wastewater: from technology to economy. Bioresour.
Viglašová, E., Galamboš, M., Danková, Z., Krivosudský, L., Lengauer, C.L., Hood- Technol. Rep. 11, 100425. https://doi.org/10.1016/j.biteb.2020.100425.
Nowotny, R., Soja, G., Rompel, A., Matík, M., Briančin, J., 2018. Production, Ye, L., Camps-Arbestain, M., Shen, Q., Lehmann, J., Singh, B., Sabir, M., 2020a. Biochar
characterization and adsorption studies of bamboo-based biochar/montmorillonite effects on crop yields with and without fertilizer: a meta-analysis of field studies
composite for nitrate removal. Waste Manag. 79, 385–394. https://doi.org/ using separate controls. Soil Use Manag. 36 (1), 2–18. https://doi.org/10.1111/
10.1016/j.wasman.2018.08.005. sum.12546.
Wang, L., Ok, Y.S., Tsang, D.C.W., Alessi, D.S., Rinklebe, J., Mašek, O., Bolan, N.S., Ye, L., Camps-Arbestain, M., Shen, Q., Lehmann, J., Singh, B., Sabir, M., 2020b. Biochar
Hou, D., 2022. Biochar composites: emerging trends, field successes and effects on crop yields with and without fertilizer: a meta-analysis of field studies
sustainability implications. Soil Use Manag. 38 (1), 14–38. https://doi.org/10.1111/ using separate controls. Soil Use Manag. 36 (1), 2–18. https://doi.org/10.1111/
sum.12731. sum.12546.
Wang, J., Wang, S., 2019. Preparation, modification and environmental application of Yin, Q., Ren, H., Wang, R., Zhao, Z., 2018a. Evaluation of nitrate and phosphate
biochar: a review. J. Clean. Prod. 227, 1002–1022. https://doi.org/10.1016/j. adsorption on Al-modified biochar: influence of Al content. Sci. Total Environ.
jclepro.2019.04.282. 631–632, 895–903. https://doi.org/10.1016/j.scitotenv.2018.03.091.
Wang, B., Lehmann, J., Hanley, K., Hestrin, R., Enders, A., 2015a. Adsorption and Yin, Q., Wang, R., Zhao, Z., 2018b. Application of Mg–Al-modified biochar for
desorption of ammonium by maple wood biochar as a function of oxidation and pH. simultaneous removal of ammonium, nitrate, and phosphate from eutrophic water.
Chemosphere 138, 120–126. https://doi.org/10.1016/j.chemosphere.2015.05.062. J. Clean. Prod. 176, 230–240. https://doi.org/10.1016/j.jclepro.2017.12.117.
Wang, Z., Guo, H., Shen, F., Yang, G., Zhang, Y., Zeng, Y., Wang, L., Xiao, H., Deng, S., Yin, Q., Liu, M., Li, Y., 2020. Desorption characteristics of phosphate and ammonium
2015b. Biochar produced from oak sawdust by Lanthanum (La)-involved pyrolysis from sludge-based biochar. Environ. Technol. 1–11. https://doi.org/10.1080/
for adsorption of ammonium (NH4+), nitrate (NO3− ), and phosphate (PO43− ). 09593330.2020.1858179, 0(0).
Chemosphere 119, 646–653. https://doi.org/10.1016/j.chemosphere.2014.07.084. Yu, J., Hu, H., Wu, X., Zhou, T., Liu, Y., Ruan, R., Zheng, H., 2020. Coupling of biochar-
Wang, Z., Li, J., Zhang, G., Zhi, Y., Yang, D., Lai, X., Ren, T., 2020. Characterization of mediated absorption and algal-bacterial system to enhance nutrients recovery from
acid-aged biochar and its ammonium adsorption in an aqueous solution. Materials swine wastewater. Sci. Total Environ. 701, 134935. https://doi.org/10.1016/j.
13 (10), 2270. https://doi.org/10.3390/ma13102270. scitotenv.2019.134935.
Wang, B., Ma, Y., Lee, X., Wu, P., Liu, F., Zhang, X., Li, L., Chen, M., 2021. Zhang, M., Gao, B., Yao, Y., Xue, Y., Inyang, M., 2012. Synthesis of porous MgO-biochar
Environmental-friendly coal gangue-biochar composites reclaiming phosphate from nanocomposites for removal of phosphate and nitrate from aqueous solutions. Chem.
water as a slow-release fertilizer. Sci. Total Environ. 758, 143664. https://doi.org/ Eng. J. 210, 26–32. https://doi.org/10.1016/j.cej.2012.08.052.
10.1016/j.scitotenv.2020.143664. Zhang, M., Song, G., Gelardi, D.L., Huang, L., Khan, E., Mašek, O., Parikh, S.J., Ok, Y.S.,
Wilsenach, J.A., Maurer, M., Larsen, T.A., van Loosdrecht, M.C.M., 2003. From waste 2020. Evaluating biochar and its modifications for the removal of ammonium,
treatment to integrated resource management. Water Sci. Technol. 48 (1), 1–9. nitrate, and phosphate in water. Water Res. 186, 116303. https://doi.org/10.1016/j.
https://doi.org/10.2166/wst.2003.0002. watres.2020.116303.
Xiao, R., Zhang, H., Tu, Z., Li, R., Li, S., Xu, Z., Zhang, Z., 2020. Enhanced removal of Zhang, M., Yang, J., Wang, H., Lv, Q., Xue, J., 2021. Enhanced removal of phosphate
phosphate and ammonium by MgO-biochar composites with NH3⋅H2O hydrolysis from aqueous solution using Mg/Fe modified biochar derived from excess activated
pretreatment. Environ. Sci. Pollut. Control Ser. 27 (7), 7493–7503. https://doi.org/ sludge: removal mechanism and environmental risk. Environ. Sci. Pollut. Control
10.1007/s11356-019-07355-5. Ser. 28 (13), 16282–16297. https://doi.org/10.1007/s11356-020-12180-2.
Xu, K., Lin, F., Dou, X., Zheng, M., Tan, W., Wang, C., 2018. Recovery of ammonium and Zhao, H., Xue, Y., Long, L., Hu, X., 2017. Adsorption of nitrate onto biochar derived from
phosphate from urine as value-added fertilizer using wood waste biochar loaded agricultural residuals. Water Sci. Technol. 77 (2), 548–554. https://doi.org/
with magnesium oxides. J. Clean. Prod. 187, 205–214. https://doi.org/10.1016/j. 10.2166/wst.2017.568.
jclepro.2018.03.206. Zin, M.M.Thant, Kim, D.-J., 2021. Simultaneous recovery of phosphorus and nitrogen
Xu, D., Cao, J., Li, Y., Howard, A., Yu, K., 2019. Effect of pyrolysis temperature on from sewage sludge ash and food wastewater as struvite by Mg-biochar. J. Hazard
characteristics of biochars derived from different feedstocks: a case study on Mater. 403, 123704. https://doi.org/10.1016/j.jhazmat.2020.123704.
ammonium adsorption capacity. Waste Manag. 87, 652–660. https://doi.org/
10.1016/j.wasman.2019.02.049.

12

You might also like