Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Waste Management 158 (2023) 76–83

Contents lists available at ScienceDirect

Waste Management
journal homepage: www.elsevier.com/locate/wasman

Nitrogen availability in biochar-based fertilizers depending on activation


treatment and nitrogen source
Raúl Castejón-del Pino *, María L. Cayuela, María Sánchez-García, Miguel A. Sánchez-Monedero
Department of Soil and Water Conservation and Organic Waste Management, CEBAS-CSIC, Campus Universitario de Espinardo, 30100 Murcia, Spain

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

Keywords: Different activation and N-doping treatments were used to produce biochar-based fertilizers (BBFs) with
Agricultural waste increased N concentration and slow N release. Pristine biochars were produced by pyrolysis of olive tree pruning
Olive tree pruning feedstock at low and high temperatures (400 and 800 ◦ C). These biochars were activated either by ultra­
N-doped biochar
sonication, or oxidation with hydrogen peroxide (H2O2) or nitric acid (HNO3) to increase their N retention
Available N
Nitrogen use efficiency
potential. Subsequently biochars were enriched with N with either urea or ammonium sulfate. The activation of
Oxidizing treatment low-temperature biochars with HNO3 was the most effective treatment leading to new surface carboxylic groups
that facilitated the later enrichment with N. When treated with urea, BBFs reached 7.0 N%, whereas the H2O2
activation only allowed an increase up to 2.0 N%. The use of urea as the external N source was the most efficient
for incorporating N. Urea treated biochars had a water-soluble fraction that represented up to 14.5 % of the total
N. The hydrolyzable N fraction, composed by amides and simple N heterocycles originated by the N-doping
treatments, and nitro groups generated from HNO3 activation represented up to 60 % of the total N. This study
relates the N chemical forms in the new BBFs to potential N availability in soil. The presence of water-soluble,
hydrolyzable and non-hydrolyzable N implied that these BBFs may supply N that would be progressively
available for plants, acting as slow-release fertilizers.

1. Introduction on the biochar surface through cation exchange capacity sites, which
regulates the release of N and prevent N losses (Cai et al., 2016).
Biochar-based fertilizers (BBFs) are a new type of engineered fertil­ BBFs represent a new approach to optimize biochar-fertilizer in­
izers, where biochar is formulated with mineral fertilizers or beneficial teractions. Different methods have been developed for their production.
microorganisms, for precise plant nutrition and the enhancement of soil The most common strategy is a physical mixture of biochar, synthetic
properties derived from biochar application (Calabi-Floody et al., 2018; fertilizer and an adhesive or a coating material (Sim et al., 2021).
Rombel et al., 2022). These new BBFs have been proposed as a strategy Recently, a new promising approach is the production of engineered
to increase nitrogen use efficiency (NUE), to improve soil properties, and biochars, where N is incorporated into the structure of the BBFs, either
to sequester C in soil simultaneously, in comparison to conventional by pre- or post-pyrolysis processes (Ndoung et al., 2021).
mineral fertilizers (Joseph et al., 2021; Melo et al., 2022; Zhou et al., A variety of methods can be used as pre-pyrolysis treatments: direct
2021). pyrolysis of N-rich feedstock (Piash et al., 2021), using NH3 as the py­
The co-application of biochar and synthetic fertilizers in soil has been rolysis carrier gas (Yu et al., 2018), or preloading a feedstock with an
the traditional approach utilized to exploit the benefits of their inter­ exogenous source of N such as ammonium nitrate (Zhu et al., 2018) or
action. Biochar applied in combination with mineral N fertilizers has the urea (Wang et al., 2019a). These N-doped biochars are typically used for
potential to increase NUE by reducing N2O emissions and NO–3 leaching, the removal of contaminants in water or soil (Shakoor et al., 2021; Yu
improving N uptake by plants (Marcińczyk and Oleszczuk, 2022), and et al., 2022), but with limited agricultural applications. N-enriched
increasing crop yield, as compared to the application of synthetic fer­ biochars produced by this strategy usually contain most of the N in a
tilizer alone (Bai et al., 2022; Melo et al., 2022). One of the most stable and recalcitrant form that is hardly available for plants. Rasse
recognized mechanisms behind this interaction is the adsorption of NH+ 4 et al. (2022) also identified the insufficient sorption capacity of biochar

* Corresponding author.
E-mail address: rcastejon@cebas.csic.es (R. Castejón-del Pino).

https://doi.org/10.1016/j.wasman.2023.01.007
Received 13 October 2022; Received in revised form 15 December 2022; Accepted 6 January 2023
Available online 13 January 2023
0956-053X/© 2023 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/).
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

as a limitation for the production of BBFs with enough N concentration et al. (2017). This treatment was performed at 80 ◦ C for 1 h and shaken
and the capacity to perform as slow release fertilizer. at 50 rpm (1:25 w/v) in a water bath (WNB 14, Memmbert, Germany).
Post-pyrolysis treatments to produce N-rich biochars are heteroge­ After the activation treatments, samples were washed with deionized
neous, and produce different end materials with slow-release perfor­ water three times and dried at 105 ◦ C until a constant weight was
mance (Marcińczyk and Oleszczuk, 2022). In order to optimize N obtained.
incorporation, a previous activation step is recommended (Rasse et al., Eventually, the activated biochars were supplemented with N either
2022). The activation treatments can be separated between physical and as urea (U) or ammonium sulfate (AS). Biochars were submerged (1:10
chemical (Jellali et al., 2022): i) physical modification is related to the w/v) in a 2 M urea or ammonium sulfate solution at 80 ◦ C for 1.5 h in a
increase in the specific surface area and the total pore volume by ball- thermoreactor (ECO 16, Velp Scientifica, Italy). After the N-doping
milling, ultrasonication, and CO2 or steam activation, ii) chemical treatments, samples were washed with deionized water three times and
modification by acidic, salt and alkaline solutions, or a mixture of dried at 80 ◦ C until a constant weight was obtained. The complete
chemicals, to increase the number of functional groups. Oxygenated description of the different biochar-based fertilizers and the identifica­
functional groups facilitate surface chemical interactions, notably tion codes are summarized in the Supplementary Material in Table S1.
increasing the active binding sites (Yang et al., 2019). Either acid,
alkaline or oxidizing modifications enhance the formation of O-func­
tional groups such as carboxylic, lactonic, phenolic, and carbonyl 2.3. Characterization of BBFs
groups, which can interact with N (Wang and Wang, 2019). The complex
mechanisms behind the retention of N into biochar represent an op­ An automatic elemental analyzer (CHNS-932, LECO, USA) was used
portunity to produce controlled N-release fertilizers with different N to measure the elemental content of carbon (C), hydrogen (H) and ni­
bioavailability (Bakshi et al., 2021). trogen (N) in the samples. The sulfur concentration was below the limit
In this work, we propose the production of a new type of engineered of detection (0.2 %). Before analysis, the samples were ball-milled to
fertilizers by the chemical incorporation of mineral N fertilizers on the <0.5 mm. Ash content was analyzed according to ASTM D1762-84
activated surface of biochar and we relate the chemical characterization (2021) following the next equation: Ash (%) = (calcined material (g) /
of the incorporated N forms in biochar with their bioavailability. To dried material (g)) × 100. Oxygen (O) concentration was determined by
achieve this aim, we produced BBFs through the post-pyrolysis treat­ subtraction (O (%) = 100 – C (%) – H (%) – N (%) – ash (%)). Electrical
ments of biochars, which were pyrolyzed at two highest treatment conductivity (EC) and pH were determined in a 1:20 (w/v) water-
temperatures (HTT) of 400 and 800 ◦ C. The surface of these biochars soluble extract after the sample was shaken for 2 h, centrifuged for
was activated to maximize their N retention capacity either by soni­ 20 min at 4000 rpm, and paper-filtered.
cation or by oxidation with hydrogen peroxide or nitric acid. Urea and An X-ray Photoelectron Spectrometer (XPS) analysis was performed
ammonium sulfate were used as N sources to enrich the biochars. We to obtain the surface functional group distribution and atomic elemental
postulated that (i) the physicochemical activation influences the N composition of the BBFs, using an XPS system (K-Alpha, Thermo Sci­
enrichment of the BBFs and (ii) the N chemical forms incorporated with entific, UK) with a monochromatic Al Kα radiation source. The quanti­
the added N sources may define the N bioavailability in BBFs. tative analysis of the XPS results was performed by calculating the
integral of each peak, where baselines were adjusted following the
2. Materials and methods Shirley method. The peaks were detected in Lorentzian-Gaussian type
curves. The XPS data were processed with the available software
2.1. Biochar production XPSPEAK 4.1. Before XPS analysis, samples were ball-milled < 0.5 mm.
Fourier Transform Infrared (FTIR) spectroscopy was performed with
Olive tree pruning was selected as feedstock for the production of a Perkin-Elmer 1430 spectrophotometer to evaluate the surface func­
biochar. Pruning consisted of tree branches with a diameter of 5–15 mm, tional groups. The KBr pellet method was used for this. Briefly, 1–2 mg
which were separated from leaves. Feedstock was collected in an olive of the sample were mixed with 200 mg of dried KBr and compressed to
orchard located in the South-East of Spain. Prior to biomass pyrolysis, form pellets. 10 scans were recorded per sample. Before the IR analysis,
the feedstock was air-dried to a moisture content below 10 % and milled the samples were ball-milled < 0.5 mm.
to a particle size of <6 mm. Dried and milled feedstock was slow- Three N fractions were defined depending on its bioavailability, as
pyrolyzed (a temperature ramp of 5 ◦ C min− 1 and 2 h of residence referred to the grade of N availability to plants: water-soluble N (shortly
time) in a rotatory tube furnace (RSR-B 80/500/11, Nabertherm, Ger­ available N for plants), hydrolyzable N (gradually available N) and non-
many) at two HTT of 400 and 800 ◦ C (B-400 and B-800, respectively) in hydrolyzable N (hardly available N). Water-soluble N was determined in
an oxygen-free environment, maintained with an Ar flux. Further in­ a deionized water extraction, which was shaken for 2 h (1:20 w/v),
formation on the pyrolysis conditions and the physicochemical proper­ centrifuged for 15 min at 4000 rpm, filtered (<0.45 μm), and measured
ties of feedstock and biochars was provided in a previous study by by an aqueous elemental analyzer (multi N/C 3100, Analytik Jena,
Sánchez-García et al. (2019). Pristine biochars were washed twice by Germany). Hydrolyzable and non-hydrolyzable N fractions were
stirring for 10 min with deionized water (1:10 w/w), dried at 105 ◦ C, analyzed following an acid digestion according to Wang et al. (2012).
milled and sieved (<1 mm). This procedure allows us to separate the bio-reactive N pool from the
chemically stable N of the BBFs. Briefly, a 0.5 g sample of biochar was
2.2. Synthesis of the BBFs treated with 25 mL of 6 M HCl and 0.1 % phenol plus 2 drops of octyl
alcohol in a 100 mL Pyrex cation digestion tube. The 1:50 mixture was
Three activation treatments were applied to biochars prepared at blended with a vortex mixer, covered with a reflux funnel, and placed on
both temperatures: ultrasonication (treatment 1) and oxidation either an Al digestion block (Bloc Digest 12, J.P. Selecta, Spain) for 24 h at
with hydrogen peroxide (treatment 2) or nitric acid (treatment 3). 105 ◦ C. Non-hydrolyzable N was determined by elemental analysis in
Treatment 1: Ultrasonic treatment at room temperature for 0.5 h using a the biochar after the acid-digestion, whereas hydrolyzable N was esti­
1:10 mass ratio of biochar:deionized water in an ultrasonic device set at mated as follows:
80 W and 50 Hz (LT − 80 PRO, TQTECH., Spain). Treatment 2: 5 % H2O2 Hydrolyzable N = Total N – non-hydrolyzable N – water-soluble N.
treatment following a methodology modified from Mia et al. (2017).
This treatment was performed at 80 ◦ C for 1.5 h, and shaken at 80 rpm Additionally, mineral N-fractions were determined as follows: NO–3
(1:25 w/v) in a water bath (WNB 14, Memmbert, Germany). Treatment and NO–2 were analyzed by ion chromatography (ICS 2100, Dionex, USA)
3: 40 % HNO3 treatment following a methodology modified from Güzel in the same water extract used for water-soluble N. NH+
4 was extracted

77
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

by shaking the samples for 2 h (1:20 w/v) with 2 M KCl, centrifuging for were also grouped according to pyrolysis temperature in the 0.05–0.20
15 min at 4000 rpm, and filtering (<0.45 μm). NH+ 4 was measured by a range in biochars made from B-800, and 0.20–0.45 in biochars derived
colorimetric method based on Bethelot’s reaction (Sommer et al., 1992). from B-400. Surface activating treatments also had a strong impact on
Two replicates were performed in each analysis to calculate the the O/C molar ratios of the BBFs, which was more evident in biochars
means and the standard deviations. made at 400 ◦ C. Furthermore, the biochars activated with HNO3 showed
the highest molar O/C ratio at both HTT, caused by the increase in O and
3. Results the decrease in C concentrations. The source of N did not affect the
molar ratios.
3.1. Elemental and physicochemical characterization of BBFs Pristine biochars produced at 400 ◦ C and 800 ◦ C had a basic pH of
8.32 and 10.69, respectively (Table 1). Both oxidizing treatments, HNO3
Biochar-based fertilizers were prepared from olive tree pruning and H2O2, decreased the pH values, especially in biochars modified with
biochars (B-400 and B-800), which were characterized by a rich organic HNO3, reaching 3.40 for B3-400 and 4.53 for B3-800. Generally, urea-
composition, and low N and ash concentrations (Table 1). Nitric acid treated biochars showed higher pH values (between 5.46 and 8.99)
oxidation was the most effective treatment for incorporating N into the than ammonium sulfate-enriched biochars (between 4.65 and 7.0).
biochar structure. The oxidation with HNO3, without any additional Although most BBFs had a circumneutral pH, B3U-400 and B3AS-400
source of N, increased N concentration from 0.68 to 4.31 % in B3-400 were moderately acidic, with a pH of 5.46 and 4.65, respectively. The
and from 0.81 to 1.97 % in B3-800. The addition of either urea or electrical conductivity (EC) of pristine biochars was low for both B-400
ammonium sulfate further increased N concentrations up to 7.02 % and and B-800, with a similar value of 0.73 and 0.59 dS/cm, respectively
5.39 % in B3U-400 and B3AS-400, respectively. The addition of extra N (Table S2). The biochar’s EC increased with the N content, especially
was less effective in biochars prepared at 800 ◦ C, where the N concen­ when ammonium sulfate was used.
tration slightly increased up to 2.79 % and 1.62 % in B3U-800 and B3AS-
800, respectively. At both HTT, the use of urea as an added N source was 3.2. Surface functionalization of BBFs
more effective than ammonium sulfate in incorporating N into the BBF
structure. The surface functionalization of the BBFs was assessed by both XPS
The activation with H2O2 slightly facilitated the incorporation of the and FTIR spectroscopy. The FTIR spectra of pristine biochar (B-400)
added N in biochars produced at 400 ◦ C (B2U-400 and B2AS-400), in were dominated by the presence of two main bands at 3430 and 1600
comparison to the non-activated biochar (BU-400). The incorporation of cm− 1 (Fig. 1), corresponding to the stretching vibrations of O–H and
N in H2O2-treated biochar was even less effective in biochars prepared at aromatic C–– C/C–C bonds, respectively (Johnston, 2017). In the case of
800 ◦ C. The sonication treatment did not show any significant effect B-800, these bands were less intense, especially the one at 1600 cm− 1
neither in 400 nor in 800 ◦ C biochars. (Figure S2). The oxidation treatments enhanced the O–H band, espe­
The Van Krevelen diagram can group biochars according to their cially HNO3 treatments. Furthermore, the FTIR spectra of the HNO3-
degree of aromaticity and the content of O-functional groups treated biochars (B3-400, B3U-400 and B3AS-400) were characterized
(Figure S1). For all modified biochars, molar H/C ratios were under the by three peaks at 1720, 1532 and 1385 cm− 1, corresponding to C– –O
0.7 limit established by the IBI Biochar Standard (2015). BBFs were stretching vibration from carboxylic acid groups (Li et al., 2014),
grouped according to pyrolysis HTT. The aromaticity of the biochars was asymmetric –NO2 and symmetric –NO2, respectively (Fan et al., 2018;
mostly affected by the pyrolysis temperature, regardless of the activa­ Qian and Chen, 2014) (Fig. 1). In contrast, the sonication treatments did
tion treatment, except for the HNO3 treatment, which caused a signifi­ not change the spectra as compared to the non-activated biochars.
cant decrease in the H concentrations in the biochars. BBFs made at The differences observed in the FTIR spectra of the BBFs were driven
800 ◦ C were grouped between 0.1 and 0.2H/C ratios, whereas those at by the oxidation treatments rather than the added N source, since
400 ◦ C were found between 0.5 and 0.6. Similarly, molar O/C ratios ammonium sulfate and urea-enriched biochars had similar FTIR spectra.

Table 1
Elemental content (C, H, O and N in %), ash (%) and pH of the pristine biochars produced at 400 and 800 ◦ C, the HNO3-activated biochars without subsequent N
addition (treatment 3) and the sixteen final materials doped with urea (U) or ammonium sulfate (AS) after ultrasonic (1), H2O2 (2) or HNO3 (3) treatments.
%C %H %O %N %Ash pH

Pristine biochars
B-400 73.2 ± 0.7 3.86 ± 0.63 17.6 ± 1.4 0.68 ± 0.07 4.6 ± 0.0 8.32 ± 0.06
B-800 82.0 ± 0.4 1.04 ± 0.11 12.0 ± 0.1 0.81 ± 0.03 4.1 ± 0.2 10.7 ± 0.12
Only treated with HNO3
B3-400 57.9 ± 0.5 2.51 ± 0.27 31.3 ± 0.94 4.31 ± 0.03 3.9 ± 0.1 3.40 ± 0.07
B3-800 76.2 ± 0.7 0.88 ± 0.14 18.0 ± 1.16 1.97 ± 0.07 3.0 ± 0.5 4.53 ± 0.17
Urea
BU-400 69.0 ± 0.7 3.16 ± 0.07 20.3 ± 0.6 1.51 ± 0.00 6.0 ± 0.0 8.05 ± 0.01
B1U-400 69.5 ± 0.0 3.54 ± 0.07 18.3 ± 0.6 1.61 ± 0.04 7.0 ± 0.7 7.82 ± 0.08
B2U-400 68.2 ± 2.0 3.51 ± 0.07 23.9 ± 2.3 2.04 ± 0.02 2.3 ± 0.2 6.98 ± 0.13
B3U-400 56.8 ± 1.0 2.68 ± 0.19 32.4 ± 0.5 7.02 ± 0.02 1.1 ± 0.3 5.46 ± 0.06
BU-800 84.7 ± 0.7 0.93 ± 0.08 9.3 ± 0.4 1.53 ± 0.04 3.6 ± 0.1 8.62 ± 0.18
B1U-800 83.1 ± 0.1 1.28 ± 0.30 10.5 ± 1.1 1.73 ± 0.67 3.3 ± 0.2 8.99 ± 0.11
B2U-800 80.4 ± 0.8 0.88 ± 0.01 14.1 ± 2.1 1.33 ± 0.01 3.3 ± 1.2 8.29 ± 0.05
B3U-800 76.9 ± 0.8 1.09 ± 0.07 17.3 ± 0.2 2.79 ± 0.02 1.9 ± 0.7 7.18 ± 0.01
Ammonium sulfate
BAS-400 68.4 ± 1.6 3.07 ± 0.07 25.7 ± 1.4 0.88 ± 0.07 2.0 ± 0.3 6.90 ± 0.15
B1AS-400 66.5 ± 1.1 3.24 ± 0.07 28.1 ± 1.0 0.98 ± 0.02 1.2 ± 0.0 6.64 ± 0.02
B2AS-400 70.5 ± 0.6 2.97 ± 0.03 23.9 ± 0.9 1.03 ± 0.02 1.5 ± 0.3 6.04 ± 0.18
B3AS-400 57.5 ± 0.5 2.67 ± 0.21 34.0 ± 0.5 5.39 ± 0.02 0.2 ± 0.3 4.56 ± 0.01
BAS-800 86.3 ± 0.2 0.92 ± 0.09 7.7 ± 0.0 0.90 ± 0.06 4.2 ± 0.3 7.00 ± 0.07
B1AS-800 84.1 ± 1.4 0.89 ± 0.09 9.3 ± 0.6 0.97 ± 0.03 4.6 ± 0.7 7.01 ± 0.04
B2AS-800 82.6 ± 0.2 1.02 ± 0.07 12.4 ± 0.1 0.73 ± 0.04 3.3 ± 0.2 6.77 ± 0.02
B3AS-800 75.4 ± 0.0 1.25 ± 0.03 19.4 ± 0.1 1.62 ± 0.02 2.4 ± 0.2 6.16 ± 0.08

78
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

Fig. 1. a) Region between 4000 and 750 cm− 1 of infrared spectra of pristine biochar made at 400 ◦ C (B-400), biochar only treated with HNO3 (B3-400) and non-
treated biochar or activated with treatments 1, 2 and 3 (ultrasonic, H2O2 and HNO3, respectively) with final N-enrichment with urea (BU-400, B1U-400, B2U-400 and
B3U-400). b) FTIR spectra of B-400, B3-400 and BBFs made at 400 ◦ C treated with ammonium sulfate.

The only difference was the peak at 1590–1600 cm− 1, which showed a treated biochars, which exhibited a peak at 1385 cm− 1, corresponding to
rise after the urea treatments, due to the formation of amide or amine N–O stretching vibration from –NO2 groups.
groups (Bamdad et al., 2021; Shi et al., 2020). Only minor differences XPS was used to describe the surface chemical composition of C, O
were found in BBFs produced with B-800 due to their high recalcitrant and N content by studying the binding energies of C1s, O1s and N1s
structure (Figure S2). The only difference was observed in the HNO3- photoelectrons of the carbonaceous materials (Table S3). C1s were

Fig. 2. N1s XPS over the binding energy range 410–395 eV of pristine biochar made at 400 ◦ C (B-400), B-400 with urea treatment (BU-400), B-400 with HNO3
treatment (B3-400) and B-400 with HNO3 and urea treatments (B3U-400).

79
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

decomposed into three peaks at 284.5 eV (aromatic and aliphatic C–C/ of hydrolyzable N in B2U-400, as compared to BU-400 and B1U-400.
C–H bonds), 286.1–286.3 eV (C–O) and 288.2–288.5 eV (C– – O) BBFs with the HNO3 treatment showed the highest content of poten­
(Zhang et al., 2021). The total C1s concentration did not change tially bioavailable N (water-soluble and hydrolyzable N), especially
significantly between treatments. In all cases, the XPS results were those produced from biochar made at 400 ◦ C, which reached 39.3 and
dominated by the C– – C/C–C signal (aromatic and aliphatic C), which 24.4 N g kg− 1 of potentially bioavailable N in B3U-400 and B3AS-400,
represented up to 80 % of the total C concentration. However, after the respectively.
HNO3 and H2O2 treatments, the C– – C/C–C relative percentage The added N source highly influenced hydrolyzable N, as urea was
decreased due to the oxidizing effect, while the peaks corresponding to more efficient for incorporating N than ammonium sulfate, and espe­
functionalized C, such as C–O and C– – O, increased. This effect was cially for incorporating hydrolyzable N forms. N in ammonium sulfate
clearly observed in HNO3-treated biochars, B3-400 and B3-800. The samples was basically non-hydrolyzable, except with nitric acid acti­
functionalization of surface was also observed in the O1s spectra, which vation and subsequent ammonium sulfate treatment. In urea-treated
presented three peaks at 531.1–531.3 eV (C– – O bonds), 532.5–532.7 eV biochars, the concentration of water-soluble N was higher than the
(O–C– – O) and 533.7 eV (C–O) (Zhang et al., 2020) (Table S3). In sum of the inorganic soluble N. For instance, the water-soluble N of B3U-
general, the relative percentage of O functional groups only showed 400 was 10.19 N g kg− 1, whereas the sum of NO–3-N and NH+ 4 -N only
minor differences between the samples. Nitric acid treatments (B3-400 reached 1.49 N g kg− 1 (Table S2), meaning that other soluble N sources
and B3-800) produced an increase in O–C– – O and C–O stretching vi­ apart from the mineral forms, dominated the water-soluble fraction.
bration in comparison to pristine biochars. BBFs prepared with urea, Pyrolysis HTT affected N availability of the BBFs, as those made with
presented a higher percentage of C– – O, which increased gradually with B-800 showed not only lower total N concentrations, but also lower
the abundance of N in the samples, as compared with those prepared water-soluble and hydrolyzable N than those produced with B-400. Only
with ammonium sulfate. BBFs treated with nitric acid showed a relevant bioavailable N fraction
In the N1s spectra (Fig. 2), three peaks were detected in all the in the BBFs produced with B-800, with the ammonium sulfate treatment
samples at 398.7 (pyridinic N bonds), 399.9–400.1 (amides, amines and being unsuccessful in increasing N concentration in biochar produced at
pyrrolic N) and 401.2–401.5 eV (graphitic N) (Leng et al., 2020; Wang high pyrolysis temperatures.
et al., 2019b). In the case of HNO3-treated biochars, an additional peak
associated to the bonding stretching of –NO2 was observed at 405.5 eV 4. Discussion
in BBFs made with biochar at 400 ◦ C (Fig. 2, Table S3), and at 406.5 eV
in samples made with biochar at 800 ◦ C. Fig. 2 also shows that amide 4.1. Influence of activation treatments on the N enrichment of BBFs
and pyrrolic N increased after both HNO3 and urea treatments. The
pyridinic N peak slightly increased after HNO3 treatment, but increased In agreement with our first hypothesis, the activation of biochar
more strongly after the urea treatment. The graphitic N peak hardly before the addition of an external source of N was a suitable strategy for
changed by the treatments, and lastly, the presence of ammonium was enhancing N retention. Whereas the addition of N source to non-
not proven (Fig. 2). activated biochars achieved N concentrations of up to 1.5 %, the com­
bination of the oxidation with HNO3 and the addition of N sources
3.3. N availability in BBFs produced BBFs with up to 7.0 % N. Activation with H2O2, and especially
HNO3, are known to increase biochar surface functional acid groups. Li
N in the BBFs was separated into three different fractions according and Li (2019) described that HNO3 treatment produced carboxyl,
to its availability: water-soluble N, hydrolyzable N, and non- carbonyl and hydroxyl groups, as a consequence of the oxidation of
hydrolyzable N (Fig. 3). N in pristine biochars was mostly non- aromatic and aliphatic C on the biochar surface. In our case, an
hydrolyzable. Modified biochars not only presented a high percentage enhancement of –COOH and –OH groups were observed in the FTIR
of N, but also exhibited an enhanced proportion of bioavailable N, spectra of biochar oxidized with either HNO3 (B3-400) or H2O2 (B2-400)
especially those prepared with HNO3 and/or urea. Ultrasonic-treated (Figure S3). In addition, the XPS analysis confirmed an increase in these
biochars showed a high proportion of non-hydrolyzable N, similar to functional groups in B3-400. The degree of oxidation was unequal for
pristine biochars. The oxidation with H2O2 increased the concentration both oxidizing treatments, as HNO3 oxidation was stronger than H2O2,

Fig. 3. Water-soluble N, hydrolyzable N and non-hydrolyzable N of pristine biochars, biochars treated with HNO3, and the sixteen BBFs, expressed in N g kg− 1. U and
AS represent urea and ammonium sulfate treatments, respectively. 1, 2, and 3 correspond to ultrasonic, H2O2 and HNO3 activations, respectively. 400 ◦ C and 800 ◦ C
are the pyrolysis HTT of the biochars.

80
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

as reflected by the increase in the intensity of the –COOH peak observed but, in our case, even using higher HNO3 concentration, we did not find
in the FTIR spectra of B3-400 compared with B2-400 (Figure S3). The any significant increase.
increase in surface acidic functional groups in the oxidized biochars BBFs treated with urea showed a remarkable amount of water-
improved its reactivity, favoring chemical and electrostatic interactions soluble N, with inorganic N being a minor fraction of the total water-
with the added N sources to produce the enriched BBFs, in accordance soluble N (Fig. 3, Table S2). This result suggested the availability of
with Wang et al. (2020). Figure S4 was included in Supplementary organic N in urea form. Sajjadi et al. (2019) noted H-bonding between
Material summarizing the main chemical reactions produced by the carboxyl and hydroxyl groups and urea. Aside from the chemical
activation treatments and the added N source. bonding of N, activated biochars may retain N by other physicochemical
The addition of urea led to the formation of amides, which represents mechanisms. Otieno et al. (2021) found that the surface area and porous
one of the largest N fractions incorporated into the BBFs’ surface, as structure of biochar can contribute in retaining ammonium in solution,
reflected in the enhancement in the amide/pyrrolic peak observed in the but in our case, the XPS and FTIR spectra rejected significant NH+ 4
XPS analysis (Fig. 2), and in the FTIR peak at 1590–1600 cm− 1 (Fig. 1). concentrations. In fact, neither NO–3 nor NH+ 4 were the main water-
The interaction between the carboxylic acid groups of the biochar with soluble N forms in urea-treated biochars, suggesting that a consider­
amine-N compounds to form amides has already been described for able amount of urea was trapped in the biochar pores. Liu et al. (2019)
ethylenediamine (Zahedifar et al., 2021) and urea (Chhatwal et al., described how the specific surface area diminished after the urea
2020). Amine-N may interact either by a nucleophilic reaction with a application, suggesting that urea particles filled the micro and meso­
carbonyl group, or by a condensation reaction with carboxyls, forming pores of the biochar. Nevertheless, HNO3-BBFs, which showed the
amides in both cases (Li and Li, 2019). Another fact that demonstrates highest water-soluble N contents, are characterized by decreasing both
the formation of amides after urea treatment was the perceptible surface area and total pore volume (Chacón et al., 2020; Yang et al.,
carboxyl peak in B2-400 and B2AS-400, and its inexistence in B2U-400 2019), and the transformation of the micro- and mesopores to macro­
(Fig. 1). This information suggested the conversion of carboxyl to an pores (Güzel et al., 2017; Vu et al., 2017). These macropores may have
amide peak in the urea samples and not in the ammonium sulfate ones. been responsible for the physical retention of urea, as previously re­
The urea treatment also increased pyridinic N concentration as ported by Liu et al. (2019).
observed in the XPS analysis (Fig. 2). O-containing functional groups can The ultrasonic treatment was less effective for incorporating N than
react with amines through Maillard reactions, producing pyridinic and the oxidation treatments. Sonication has been previously used to clean
other heterocyclic N groups (Chen et al., 2018). The formation of pyr­ the biochar surface, increasing its porosity (Sajjadi et al., 2019). These
idinic N after the urea treatment was previously reported to occur at authors used sonication as a pretreatment for the activation of biochar
high temperatures during pyrolysis (Hulicova-Jurcakova et al., 2009), with urea as a strategy to increase its adsorption capacity. The final N
with the highest pyridinic N formation observed at 600 ◦ C (Wang et al., concentration of biochars were in the 1.80 and 2.23 % range, which is in
2019b). In our study, we obtained a large proportion of pyridinic N, in line with the N concentrations of the sonicated BBFs in our study (be­
the range between 21.4 and 47.9 %, in BBFs treated with urea, even at tween 0.97 and 1.73 %).
relatively low temperatures (80 ◦ C). Pyrolysis HTT also influenced N enrichment, as lower temperatures
BBFs obtained with ammonium sulfate as the external N source of biochar production led to higher N retention. Condensed aromatic C is
showed similar FTIR and XPS peaks than those treated with urea, but the the most common structure found in biochars produced at a high py­
amount of the N forms differed as compared with urea-treated BBFs. The rolysis temperature, in which aliphatic structures and functional groups
XPS analysis of biochars treated with ammonium sulfate showed lower (such as carbonyls, carboxyls, ketones and esters) are almost absent
amide/pyrrolic peaks than urea-treated biochars, suggesting a minor (Hassan et al., 2020). Due to the high chemical stability of high tem­
formation of amides in ammonium sulfate BBFs. In the case of B3U-400, perature biochars, oxidizing treatments were more effective in low
which had the highest concentration of carboxylic acid groups, it pyrolysis-temperature biochars. Our results are in agreement with Singh
showed a higher amide/pyrrolic peak than B3AS-400, confirming that it et al. (2020), who found, in a study comparing urea sorption in biochars
did not form a significant amount of amide groups (Table S3). made at different temperatures, the highest values for biochar made at
Biochar has traditionally been considered to adsorb NH+ 4 mainly 450 ◦ C, associated to the optimum equilibrium between functionality,
through H-bonding and electrostatic interactions (Cai et al., 2016), porosity and cation exchange capacity. Nevertheless, the application of
however, the XPS analysis demonstrated the absence of NH+ 4 in both activation treatments is essential for biochar produced at a high tem­
urea and ammonium sulfate treatments, evidencing that biochar cation perature, where the combination of HNO3 and urea treatments showed a
exchange capacity was not an important mechanism for N retention in remarkable hydrolyzable N concentration despite the highly stable C
our case. KCl extracts further demonstrated the absence of significant structure.
concentrations of NH+ 4 in cation exchange sites (Table S2), which con­
tradicts the traditional view. Other less-studied mechanisms, involving 4.2. N Bioavailability in BBFs
the retention of NH+ 4 with O-functional groups, may be relevant as
suggested by Takaya et al. (2016) for hydrochar and low temperature N-doping of biochar through surface functionalization with N com­
biochars. pounds has been mostly used as a strategy to remove pollutants from the
Apart from the interaction of the added N with carboxylic groups, the environment (Yang et al., 2019). The adsorption properties of N-doped
incorporation of nitro groups in the modified biochars represented up to biochars rely on the presence of amines, pyrrolic and pyridinic N, that
50 % of the N functional groups in B3-400 and B3-800, as measured in contain free electrons that can interact with organic and inorganic
the XPS analysis (Table S3). The nitration of the aromatic rings to form pollutants (Wang and Wang, 2019).
nitro groups represents another N retention mechanism derived from the In the case of BBFs, the added N needs to be plant-available to behave
oxidation with nitric acid, as reported by Sajjadi et al. (2018). In as a slow or controlled N release fertilizer. For this reason, a different
concentrated nitric acid solution, nitro-aromatic compounds are formed approach is needed to reduce the recalcitrance of the added N and in­
between NO–2 and aromatic C rings by electrophilic aromatic substitution crease its bioavailability. In this study, we hypothesized that the added
(Yang et al., 2019). The oxidation with HNO3 also originated an increase N source may influence the availability of the N in BBFs. The urea
in pyrrolic and pyridinic N, as observed in the N1s spectra (Fig. 2). The treatments showed a high amount of hydrolyzable N, while ammonium
increase in heterocyclic N alongside the formation of nitro groups has sulfate BBFs only reached a relevant N concentration when HNO3 was
been previously reported (Lu et al., 2021; Zheng et al., 2020). Li and Li used as the activation treatment. Approximately two thirds of the N in
(2019), in a biochar pre-treated with H3PO4, found graphitic N enrich­ urea BBFs were potentially bioavailable in the soil. In the case of the
ment with increasing the concentration of HNO3 treatment up to 25 %, most efficient BBFs (B3U-400), about 15 % of the total N was soluble in

81
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

water, mostly in organic forms, as the inorganic N available for plants Declaration of Competing Interest
represented 2.1 % of the N. This indicates that urea was either trapped in
pores as found by Liu et al. (2019); or retained by hydrogen bonding The authors declare that they have no known competing financial
interaction in accordance with Sajjadi et al. (2019) and Chen et al. interests or personal relationships that could have appeared to influence
(2020). Our results suggest that a small amount of NO–3-N and NH+ 4 -N the work reported in this paper.
would be plant-available at first, and the hydrolysis of the trapped urea
could start some days later in the short term (Sigurdarson et al., 2018). Data availability
A large proportion of the added N could be slowly available for
plants, since about 50 % of N in B3U-400 was hydrolyzable in acid, Data will be made available on request.
which is considered a fraction that could be mineralized over time in soil
(Otto et al., 2013). Amides and nitro groups found on the biochar’s Acknowledgments
surface may be the most relevant functional groups included in the N
hydrolyzable fractions of BBFs. Piash et al. (2021) described that in a soil We gratefully acknowledge the financial support of the projects
incubation experiment with an amide-rich biochar, most of the amide RTI2018-099417-BI00 and PID2021-128896OB-I00 from the Spanish
content was released after the first 120 days, being available for plant Ministry of Science, Innovation and Universities, co-funded with EU
use in a reasonable crop growing period. We found that acid digestion FEDER funds. The authors are very grateful to Mr. Mario Fon for editing
degraded most nitro groups, as confirmed by FTIR, where the peaks the final version of the manuscript.
associated to nitro groups (1385 and 1532 cm− 1) reduced their intensity
after hydrolysis (Figure S5). Accordingly, Marvin-Sikkema and de Bont
Appendix A. Supplementary material
(1994) previously reported that microorganism degrade numerous
nitro-aromatic compounds in soil. The aromatic N contained in biochar
Supplementary data to this article can be found online at https://doi.
is generally considered unavailable (Jassal et al., 2015). However, our
org/10.1016/j.wasman.2023.01.007.
results suggest that a small fraction of the heterocyclic N of BBFs,
especially low molecular weight pyrrols and pyridines, may also be
included in the hydrolyzable N fraction. Even though aromatic N in References
biochar is generally considered to be mostly recalcitrant (Jassal et al.,
Bai, S.H., Omidvar, N., Gallart, M., Kämper, W., Tahmasbian, I., Farrar, M.B., Singh, K.,
2015), Wang et al. (2012) reported that low molecular weight hetero­ Zhou, G., Muqadass, B., Xu, C.-Y., Koech, R., Li, Y., Nguyen, T.T.N., van Zwieten, L.,
cyclic N can decompose in soil. 2022. Combined effects of biochar and fertilizer applications on yield: A review and
The non-hydrolyzable N fraction represented about one third of the meta-analysis. Sci. Total Environ. 808, 152073 https://doi.org/10.1016/j.
scitotenv.2021.152073.
total N in urea BBFs. This remaining N, composed of graphitic and the Bakshi, S., Banik, C., Laird, D.A., Smith, R., Brown, R.C., 2021. Enhancing biochar as
non-hydrolyzable fraction of pyridinic and pyrrolic N, can be considered scaffolding for slow release of nitrogen fertilizer. ACS Sustain. Chem. Eng. 9,
to be recalcitrant. Aromatic N is mostly classified as a stable form, 8222–8231. https://doi.org/10.1021/acssuschemeng.1c02267.
Bamdad, H., Papari, S., MacQuarrie, S., Hawboldt, K., 2021. Study of surface
although some authors revealed that aging, weathering and microbial heterogeneity and nitrogen functionalizing of biochars: Molecular modeling
activity could mineralize it, being available for plant nutrition (Williams approach. Carbon 171, 161–170. https://doi.org/10.1016/j.carbon.2020.08.062.
et al., 2019). The biochar structure protects aromatic heterocyclic N, and Cai, Y., Qi, H., Liu, Y., He, X., 2016. Sorption/desorption behavior and mechanism of NH4
+
by biochar as a nitrogen fertilizer sustained-release material. J. Agric. Food Chem.
it only becomes available when the biochar structure starts to degrade in 64, 4958–4964. https://doi.org/10.1021/acs.jafc.6b00109.
soil. Mineralization of heterocyclic N may be possible on longer periods; Calabi-Floody, M., Medina, J., Rumpel, C., Condron, L.M., Hernandez, M., Dumont, M.,
however, the C ring could be opened by hydroxylation, producing Mora, M. de la L., 2018. Smart fertilizers as a strategy for sustainable agriculture, in:
Advances in Agronomy. Elsevier, pp. 119–157. https://doi.org/10.1016/bs.agron.20
aliphatic N compounds available for microbial use (Torres-Rojas et al.,
17.10.003.
2020). Chacón, F.J., Sánchez-Monedero, M.A., Lezama, L., Cayuela, M.L., 2020. Enhancing
biochar redox properties through feedstock selection, metal preloading and post-
pyrolysis treatments. Chem. Eng. J. 395, 125100 https://doi.org/10.1016/j.
5. Conclusions
cej.2020.125100.
Chen, W., Chen, Y., Yang, H., Li, K., Chen, X., Chen, H., 2018. Investigation on biomass
This study provided several modified biochars obtained from olive nitrogen-enriched pyrolysis: Influence of temperature. Bioresour. Technol. 249,
tree pruning with a potential use as slow-release N fertilizers. The 247–253. https://doi.org/10.1016/j.biortech.2017.10.022.
Chen, X., Li, H., Liu, W., Meng, Z., Wu, Z., Wang, G., Liang, Y., Bi, S., 2020. Low-
biochar-based fertilizers contain heterogeneous chemical N forms, temperature constructing N-doped graphite-like mesoporous structure biochar from
which may regulate the N release. In general, oxidation treatments furfural residue with urea for removal of chlortetracycline from wastewater and
enhanced the incorporation of N from the added N source, as compared hydrothermal catalytic degradation mechanism. Colloids Surf. Physicochem. Eng.
Asp. 600, 124873 https://doi.org/10.1016/j.colsurfa.2020.124873.
to sonicated and non-activated biochar. Biochars treated with nitric acid Chhatwal, A.R., Lomax, H.V., Blacker, A.J., Williams, J.M.J., Marcé, P., 2020. Direct
showed the highest concentration of O-functional groups, which were synthesis of amides from nonactivated carboxylic acids using urea as nitrogen source
responsible for an important part of N retention. Aside from the incor­ and Mg(NO 3) 2 or imidazole as catalysts. Chem. Sci. 11, 5808–5818. https://doi.
org/10.1039/D0SC01317J.
poration of external sources of N, the oxidation with HNO3 increased N Fan, Q., Sun, J., Chu, L., Cui, L., Quan, G., Yan, J., Hussain, Q., Iqbal, M., 2018. Effects of
content by itself, forming nitro groups and pyrrolic N. The combination chemical oxidation on surface oxygen-containing functional groups and adsorption
of HNO3 and urea led to the highest N enrichment in the biochar-based behavior of biochar. Chemosphere 207, 33–40. https://doi.org/10.1016/j.
chemosphere.2018.05.044.
fertilizers (BBFs). About two thirds of the N of the BBFs treated with urea
Güzel, F., Sayğılı, H., Akkaya Sayğılı, G., Koyuncu, F., Yılmaz, C., 2017. Optimal
would be potentially available for plants with different release rates. The oxidation with nitric acid of biochar derived from pyrolysis of weeds and its
water-soluble N, mostly in the form of trapped urea that would be firstly application in removal of hazardous dye methylene blue from aqueous solution.
J. Clean. Prod. 144, 260–265. https://doi.org/10.1016/j.jclepro.2017.01.029.
available, represented a fraction of 14.5 %. Hydrolyzable N groups, as
Hassan, M., Liu, Y., Naidu, R., Parikh, S.J., Du, J., Qi, F., Willett, I.R., 2020. Influences of
amides, nitro groups and low weight molecular N that would be slowly feedstock sources and pyrolysis temperature on the properties of biochar and
released in soils for plants nutrition, represented approximately 50 % of functionality as adsorbents: A meta-analysis. Sci. Total Environ. 744, 140714
the total N. These findings provide new insights to design future field https://doi.org/10.1016/j.scitotenv.2020.140714.
Hulicova-Jurcakova, D., Seredych, M., Lu, G.Q., Bandosz, T.J., 2009. Combined effect of
experiments to evaluate the slow-release performance, the efficient use nitrogen- and oxygen-containing functional groups of microporous activated carbon
of the N of these engineered BBFs and their economic feasibility. on its electrochemical performance in supercapacitors. Adv. Funct. Mater. 19,
438–447. https://doi.org/10.1002/adfm.200801236.
IBI International Biochar Initiative, 2022. Product Definition and Specification
Standards, version 2.1 (23 November 2015) [WWW Document]. URL http://www.
biochar-international.org/characterizationstandard.

82
R. Castejón-del Pino et al. Waste Management 158 (2023) 76–83

Jassal, R.S., Johnson, M.S., Molodovskaya, M., Black, T.A., Jollymore, A., Sveinson, K., Shi, W., Ju, Y., Bian, R., Li, L., Joseph, S., Mitchell, D.R.G., Munroe, P.,
2015. Nitrogen enrichment potential of biochar in relation to pyrolysis temperature Taherymoosavi, S., Pan, G., 2020. Biochar bound urea boosts plant growth and
and feedstock quality. J. Environ. Manage. 152, 140–144. https://doi.org/10.1016/ reduces nitrogen leaching. Sci. Total Environ. 701, 134424 https://doi.org/
j.jenvman.2015.01.021. 10.1016/j.scitotenv.2019.134424.
Jellali, S., El-Bassi, L., Charabi, Y., Usman, M., Khiari, B., Al-Wardy, M., Jeguirim, M., Sigurdarson, J.J., Svane, S., Karring, H., 2018. The molecular processes of urea
2022. Recent advancements on biochars enrichment with ammonium and nitrates hydrolysis in relation to ammonia emissions from agriculture. Rev. Environ. Sci.
from wastewaters: A critical review on benefits for environment and agriculture. Biotechnol. 17, 241–258. https://doi.org/10.1007/s11157-018-9466-1.
J. Environ. Manage. 305, 114368 https://doi.org/10.1016/j.jenvman.2021.114368. Sim, D.H.H., Tan, I.A.W., Lim, L.L.P., Hameed, B.H., 2021. Encapsulated biochar-based
Johnston, C.T., 2017. Biochar analysis by Fourier-transform infra-red spectroscopy, in: sustained release fertilizer for precision agriculture: A review. J. Clean. Prod. 303,
Biochar: A Guide to Analytical Methods. Csiro Publishing, pp. 199–213. 127018 https://doi.org/10.1016/j.jclepro.2021.127018.
Joseph, S., Cowie, A.L., Van Zwieten, L., Bolan, N., Budai, A., Buss, W., Cayuela, M.L., Singh, S.V., Chaturvedi, S., Dhyani, V.C., Kasivelu, G., 2020. Pyrolysis temperature
Graber, E.R., Ippolito, J.A., Kuzyakov, Y., Luo, Y., Ok, Y.S., Palansooriya, K.N., influences the characteristics of rice straw and husk biochar and sorption/desorption
Shepherd, J., Stephens, S., Weng, Z., (Han), Lehmann, J., 2021. How biochar works, behaviour of their biourea composite. Bioresour. Technol. 314, 123674 https://doi.
and when it doesn’t: A review of mechanisms controlling soil and plant responses to org/10.1016/j.biortech.2020.123674.
biochar. GCB Bioenergy 13, 1731–1764. https://doi.org/10.1111/gcbb.12885. Sommer, S.G., Kjellerup, V., Kristjansen, O., 1992. Determination of total ammonium
Leng, L., Xu, S., Liu, R., Yu, T., Zhuo, X., Leng, S., Xiong, Q., Huang, H., 2020. Nitrogen nitrogen in pig and cattle slurry: Sample preparation and analysis. Acta Agric. Scand.
containing functional groups of biochar: An overview. Bioresour. Technol. 298, Sect. B - Soil Plant Sci. 42, 146–151. https://doi.org/10.1080/09064719209417969.
122286 https://doi.org/10.1016/j.biortech.2019.122286. Takaya, C.A., Fletcher, L.A., Singh, S., Anyikude, K.U., Ross, A.B., 2016. Phosphate and
Li, B., Li, K., 2019. Effect of nitric acid pre-oxidation concentration on pore structure and ammonium sorption capacity of biochar and hydrochar from different wastes.
nitrogen/oxygen active decoration sites of ethylenediamine -modified biochar for Chemosphere 145, 518–527. https://doi.org/10.1016/j.chemosphere.2015.11.052.
mercury(II) adsorption and the possible mechanism. Chemosphere 220, 28–39. Torres-Rojas, D., Hestrin, R., Solomon, D., Gillespie, A.W., Dynes, J.J., Regier, T.Z.,
https://doi.org/10.1016/j.chemosphere.2018.12.099. Lehmann, J., 2020. Nitrogen speciation and transformations in fire-derived organic
Li, Y., Shao, J., Wang, X., Deng, Y., Yang, H., Chen, H., 2014. Characterization of matter. Geochim. Cosmochim. Acta 276, 170–185. https://doi.org/10.1016/j.
modified biochars derived from bamboo pyrolysis and their utilization for target gca.2020.02.034.
component (furfural) adsorption. Energy Fuels 28, 5119–5127. https://doi.org/ Vu, T.M., Trinh, V.T., Doan, D.P., Van, H.T., Nguyen, T.V., Vigneswaran, S., Ngo, H.H.,
10.1021/ef500725c. 2017. Removing ammonium from water using modified corncob-biochar. Sci. Total
Liu, X., Liao, J., Song, H., Yang, Y., Guan, C., Zhang, Z., 2019. A biochar-based route for Environ. 579, 612–619. https://doi.org/10.1016/j.scitotenv.2016.11.050.
environmentally friendly controlled release of nitrogen: Urea-loaded biochar and Wang, H., Guo, W., Liu, B., Wu, Q., Luo, H., Zhao, Q., Si, Q., Sseguya, F., Ren, N., 2019a.
bentonite composite. Sci. Rep. 9, 9548. https://doi.org/10.1038/s41598-019- Edge-nitrogenated biochar for efficient peroxydisulfate activation: An electron
46065-3. transfer mechanism. Water Res. 160, 405–414. https://doi.org/10.1016/j.
Lu, Y., Hu, Y., Tang, L., Xie, Q., Liu, Q., Zhong, L., Fu, L., Fan, C., 2021. Effects and watres.2019.05.059.
mechanisms of modified biochars on microbial iron reduction of Geobacter Wang, J., Wang, S., 2019. Preparation, modification and environmental application of
sulfurreducens. Chemosphere 283, 130983. https://doi.org/10.1016/j. biochar: A review. J. Clean. Prod. 227, 1002–1022. https://doi.org/10.1016/j.
chemosphere.2021.130983. jclepro.2019.04.282.
Marcińczyk, M., Oleszczuk, P., 2022. Biochar and engineered biochar as slow- and Wang, T., Camps Arbestain, M., Hedley, M., Bishop, P., 2012. Chemical and bioassay
controlled-release fertilizers. J. Clean. Prod. 339, 130685 https://doi.org/10.1016/j. characterisation of nitrogen availability in biochar produced from dairy manure and
jclepro.2022.130685. biosolids. Org. Geochem. 51, 45–54. https://doi.org/10.1016/j.
Marvin-Sikkema, F.D., de Bont, J.A.M., 1994. Degradation of nitroaromatic compounds orggeochem.2012.07.009.
by microorganisms. Appl Microbiol Biotechnol. 42, 499–507. https://doi.org/ Wang, X., Liu, Y., Zhu, L., Li, Y., Wang, K., Qiu, K., Tippayawong, N., Aggarangsi, P.,
10.1007/BF00173912. Reubroycharoen, P., Wang, S., 2019b. Biomass derived N-doped biochar as efficient
Melo, L.C.A., Lehmann, J., Carneiro, J.S., da, S., Camps-Arbestain, M., 2022. Biochar- catalyst supports for CO2 methanation. J. CO2 Util. 34, 733–741. https://doi.org/
based fertilizer effects on crop productivity: a meta-analysis. Plant Soil 472, 45–58. 10.1016/j.jcou.2019.09.003.
https://doi.org/10.1007/s11104-021-05276-2. Wang, Z., Li, J., Zhang, G., Zhi, Y., Yang, D., Lai, X., Ren, T., 2020. Characterization of
Mia, S., Dijkstra, F.A., Singh, B., 2017. Aging induced changes in biochar’s functionality acid-aged biochar and its ammonium adsorption in an aqueous solution. Materials
and adsorption behavior for phosphate and ammonium. Environ. Sci. Technol. 51, 13, 2270. https://doi.org/10.3390/ma13102270.
8359–8367. https://doi.org/10.1021/acs.est.7b00647. Williams, E.K., Jones, D.L., Sanders, H.R., Benitez, G.V., Plante, A.F., 2019. Effects of 7
Ndoung, O.C.N., de Figueiredo, C.C., Ramos, M.L.G., 2021. A scoping review on biochar- years of field weathering on biochar recalcitrance and solubility. Biochar 1,
based fertilizers: Enrichment techniques and agro-environmental application. 237–248. https://doi.org/10.1007/s42773-019-00026-1.
Heliyon 7, e08473. https://doi.org/10.1016/j.heliyon.2021.e08473. Yang, X., Wan, Y., Zheng, Y., He, F., Yu, Z., Huang, J., Wang, H., Ok, Y.S., Jiang, Y.,
Otieno, A.O., Home, P.G., Raude, J.M., Murunga, S.I., Ngumba, E., Ojwang, D.O., Gao, B., 2019. Surface functional groups of carbon-based adsorbents and their roles
Tuhkanen, T., 2021. Pineapple peel biochar and lateritic soil as adsorbents for in the removal of heavy metals from aqueous solutions: A critical review. Chem. Eng.
recovery of ammonium nitrogen from human urine. J. Environ. Manage. 293, J. 366, 608–621. https://doi.org/10.1016/j.cej.2019.02.119.
112794 https://doi.org/10.1016/j.jenvman.2021.112794. Yu, H., Zhang, Y., Zhan, J., Tang, C., Zhang, X., Huang, H., Ye, D., Wang, Y., Li, T., 2022.
Otto, R., Mulvaney, R.L., Khan, S.A., Trivelin, P.C.O., 2013. Quantifying soil nitrogen A composite amendment benefits rice (Oryza sativa L.) safety and production in
mineralization to improve fertilizer nitrogen management of sugarcane. Biol. Fertil. cadmium-contaminated soils by unique characteristics after oxidation modification.
Soils 49, 893–904. https://doi.org/10.1007/s00374-013-0787-5. Sci. Total Environ. 806, 150484 https://doi.org/10.1016/j.scitotenv.2021.150484.
Piash, M.I., Iwabuchi, K., Itoh, T., Uemura, K., 2021. Release of essential plant nutrients Yu, W., Lian, F., Cui, G., Liu, Z., 2018. N-doping effectively enhances the adsorption
from manure- and wood-based biochars. Geoderma 397, 115100. https://doi.org/ capacity of biochar for heavy metal ions from aqueous solution. Chemosphere 193,
10.1016/j.geoderma.2021.115100. 8–16. https://doi.org/10.1016/j.chemosphere.2017.10.134.
Qian, L., Chen, B., 2014. Interactions of aluminum with biochars and oxidized biochars: Zahedifar, M., Seyedi, N., Shafiei, S., Basij, M., 2021. Surface-modified magnetic biochar:
Implications for the biochar aging process. J. Agric. Food Chem. 62, 373–380. Highly efficient adsorbents for removal of Pb(ІІ) and Cd(ІІ). Mater. Chem. Phys. 271,
https://doi.org/10.1021/jf404624h. 124860 https://doi.org/10.1016/j.matchemphys.2021.124860.
Rasse, D.P., Weldon, S., Joner, E.J., Joseph, S., Kammann, C.I., Liu, X., O’Toole, A., Zhang, Y., Liu, N., Yang, Y., Li, J., Wang, S., Lv, J., Tang, R., 2020. Novel carbothermal
Pan, G., Kocatürk-Schumacher, N.P., 2022. Enhancing plant N uptake with biochar- synthesis of Fe, N co-doped oak wood biochar (Fe/N-OB) for fast and effective Cr(VI)
based fertilizers: limitation of sorption and prospects. Plant Soil. https://doi.org/ removal. Colloids Surf. Physicochem. Eng. Asp. 600, 124926 https://doi.org/
10.1007/s11104-022-05365-w. 10.1016/j.colsurfa.2020.124926.
Rombel, A., Krasucka, P., Oleszczuk, P., 2022. Sustainable biochar-based soil fertilizers Zhang, Y., Zheng, Y., Yang, Y., Huang, J., Zimmerman, A.R., Chen, H., Hu, X., Gao, B.,
and amendments as a new trend in biochar research. Sci. Total Environ. 816, 151588 2021. Mechanisms and adsorption capacities of hydrogen peroxide modified ball
https://doi.org/10.1016/j.scitotenv.2021.151588. milled biochar for the removal of methylene blue from aqueous solutions. Bioresour.
Sajjadi, B., Broome, J.W., Chen, W.Y., Mattern, D.L., Egiebor, N.O., Hammer, N., Technol. 337, 125432 https://doi.org/10.1016/j.biortech.2021.125432.
Smith, C.L., 2019. Urea functionalization of ultrasound-treated biochar: A feasible Zheng, Y., Wang, Z., Liu, C., Tao, L., Huang, Y., Zheng, Z., 2020. Integrated production of
strategy for enhancing heavy metal adsorption capacity. Ultrason. Sonochem. 51, aromatic amines, aromatic hydrocarbon and N-heterocyclic bio-char from catalytic
20–30. https://doi.org/10.1016/j.ultsonch.2018.09.015. pyrolysis of biomass impregnated with ammonia sources over Zn/HZSM-5 catalyst.
Sajjadi, B., Zubatiuk, T., Leszczynska, D., Leszczynski, J., Chen, W.Y., 2018. Chemical J. Energy Inst. 93, 210–223. https://doi.org/10.1016/j.joei.2019.03.007.
activation of biochar for energy and environmental applications: a comprehensive Zhou, J., Qu, T., Li, Y., Van Zwieten, L., Wang, H., Chen, J., Song, X., Lin, Z., Zhang, X.,
review. Rev. Chem. Eng. 35, 777–815. https://doi.org/10.1515/revce-2018-0003. Luo, Y., Cai, Y., Zhong, Z., 2021. Biochar-based fertilizer decreased while chemical
Sánchez-García, M., Cayuela, M.L., Rasse, D.P., Sánchez-Monedero, M.A., 2019. Biochars fertilizer increased soil N2O emissions in a subtropical Moso bamboo plantation.
from Mediterranean agroindustry residues: Physicochemical properties relevant for CATENA 202, 105257. https://doi.org/10.1016/j.catena.2021.105257.
C sequestration and soil water retention. ACS Sustain. Chem. Eng. 7, 4724–4733. Zhu, S., Huang, X., Ma, F., Wang, L., Duan, X., Wang, S., 2018. Catalytic removal of
https://doi.org/10.1021/acssuschemeng.8b04589. aqueous contaminants on N-doped graphitic biochars: Inherent roles of adsorption
Shakoor, M.B., Ye, Z.-L., Chen, S., 2021. Engineered biochars for recovering phosphate and nonradical mechanisms. Environ. Sci. Technol. 52, 8649–8658. https://doi.org/
and ammonium from wastewater: A review. Sci. Total Environ. 779, 146240 https:// 10.1021/acs.est.8b01817.
doi.org/10.1016/j.scitotenv.2021.146240.

83

You might also like