You are on page 1of 10

Journal of Environmental Management 317 (2022) 115385

Contents lists available at ScienceDirect

Journal of Environmental Management


journal homepage: www.elsevier.com/locate/jenvman

Impacts of pyrolysis temperatures on physicochemical and structural


properties of green waste derived biochars for adsorption of potentially
toxic elements
Nisar Muhammad a, b, c, Liya Ge c, Wei Ping Chan c, Afsar Khan d, Mohammad Nafees b, *,
Grzegorz Lisak c, e, **
a
Department of Environmental Science, Gomal University, Dera Ismail Khan 29050, Pakistan
b
Department of Environmental Sciences, University of Peshawar, Peshawar, 25120, Pakistan
c
Residues and Resource Reclamation Centre, Nanyang Environment and Water Research Institute, Nanyang Technological University, 1 Cleantech Loop, Singapore,
637141, Singapore
d
Department of Chemistry, University of Baltistan, Skardu, 16250, Pakistan
e
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore, 637141, Singapore

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

Keywords: This study comparatively investigated the influence of changes in pyrolysis temperature on the physicochemical,
Pyrolysis temperature structural, and adsorptive properties of biochars derived from a green waste (Cynodon dactylon L.). For this
Biochars purpose, the biophysically dried green wastes were pyrolyzed at 400 ◦ C, 600 ◦ C, and 800 ◦ C under the same
Potentially toxic elements
pyrolysis conditions. The results revealed that the physicochemical and structural properties were varied,
Adsorptive performance
Green waste
depending upon the pyrolysis temperatures. With the increase of pyrolysis temperature, the surface functional
groups were escaped, the structure became more porous (pore volume of 0.089 ± 0.001), the metal oxides were
remained consistent, and the biochars turned into more alkaline nature (pH of 11.9 ± 0.2). Furthermore, as
referring to the adsorptive performance for potentially toxic elements, with experimental adsorption capacity of
up to 33.7 mg g− 1 and removal rate up to 96% for a multi-metals containing solution, the biochars pyrolyzed at
high temperature (800 ◦ C) was significantly (p < 0.05) higher than those pyrolyzed at low temperature (400 ◦ C).
According to the physicochemical and structural properties, and the adsorptive performances of the biochars, the
optimal pyrolysis temperature was herein recommended to be 800 ◦ C.

1. Introduction productivity (Abujabhah et al., 2016; Partey et al., 2016; Venegas et al.,
2016). As the environmental applications of BCs largely depend on their
Biochar (BC) is a carbon (C) enriched solid product generated from physicochemical and structural characteristics, these properties should
the pyrolysis of different biomass (Muhammad et al., 2021) at a rela­ be carefully investigated and evaluated prior to the applications (Angin
tively low temperature (Lehmann and Joseph, 2009). It is different from and Şensöz, 2014). BCs can be produced by the pyrolysis of different raw
other biomass pyrolysis products, in which the main usage is to store C in materials (mainly including woods and the biomass of crops), high-yield
the soil for a long time, rather than to create raw materials for processing green wastes, agricultural by-products (such as straws of cereal and
industries or fuels (Cross et al., 2010). In the Amazon basin, BC has been peanut), as well as other organic wastes (such as sewage sludges and
used as a soil amendment for more than 2000 years (Kim et al., 2007). It food wastes) (Méndez et al., 2013; Yang et al., 2017). Generally, woody
has a wide range of environmental applications, which are not limited to feedstocks are preferentially used for BC productions, because they have
the fixation of carbon dioxide, the remediation of degraded soils, the high carbon contents, and contain a large amount of cellulose, hemi­
adsorption of contaminants, and the improvement of crop yield and cellulose and lignin, as well as a small amount of other organic and

* Corresponding author.
** Corresponding author. Residues and Resource Reclamation Centre, Nanyang Environment and Water Research Institute, Nanyang Technological University, 1
Cleantech Loop, Singapore, 637141, Singapore.
E-mail addresses: nafees@uop.edu.pk (M. Nafees), g.lisak@ntu.edu.sg (G. Lisak).

https://doi.org/10.1016/j.jenvman.2022.115385
Received 26 October 2021; Received in revised form 20 May 2022; Accepted 20 May 2022
Available online 28 May 2022
0301-4797/© 2022 Elsevier Ltd. All rights reserved.
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

inorganic compounds (Suliman et al., 2016). Nevertheless, previous groundwater treatment before performing the irrigation on the farmland
studies had also recommended feedstocks with higher contents of cel­ or alternatively, for soil conditioning using this locally self-grown waste
lulose, hemicellulose and lignin compounds, which resulted compara­ grass derived biochar, produced from high or low pyrolysis temperature,
tively higher yield and efficacy upon the applications (Xu and Chen, respectively. This study is also part of a larger research effort to explore
2013; Zhao et al., 2017). the use of indigenous materials for contaminated river (or groundwater)
The physicochemical and structural properties of BCs depend on treatment, wastewater treatment, bioremediation and soil amendment,
both types of feedstocks and pyrolysis conditions (Bouraoui et al., 2015; especially for the rural areas. Given the importance mentioned above,
Zhang et al., 2015). Among different operating parameters, the pyrolysis the green waste was subjected to a wide pyrolysis temperature range
temperature plays a crucial role, because it affects the yield, properties, from 400 to 800 ◦ C. The produced BCs were comprehensively investi­
and nature of BCs (Yuan et al., 2013). The pyrolysis temperature has gated and compared for (i) physicochemical properties, (ii) structural
significant effects on the surface functional groups, pore size, pore vol­ analyses, and (iii) potentially adsorptive performances of the toxic ele­
ume, surface area, yield, elemental composition, thermal stability, and ments, with pyrolysis temperature as a critical influencer in the BC py­
surface charge of BCs (Guizani et al., 2017b; Shaaban et al., 2014). The rolysis process. Furthermore, a wide concentration range of PTEs
high pyrolysis temperature is positively correlated with carbon (25–250 mg L− 1) with multiple metals (7 elements) and at different
(elemental carbon, fixed carbon and dissolved organic carbon), ash and concentrations (6 points) were tested to comprehensively explore the
mineral contents, as well as the surface area, pore size, pore volume, and application potentials of this less studied green material.
cation exchange capacity (CEC) of BCs, but negatively correlated with
hydrogen, nitrogen, sulfur and oxygen contents, as well as yield, mois­ 2. Methods and materials
ture and volatile substances of BCs (Liang et al., 2016; Muhammad et al.,
2020; Tag et al., 2016). In brief, BCs pyrolyzed at high temperatures are 2.1. Preparation of biochar
highly aromatic and difficult to decompose, which are potentially suit­
able for removing and treating contaminants (Muhammad et al., 2020). Cynodon dactylon L. was pyrolyzed in a fixed-bed reactor (horizontal)
Previous studies also demonstrated that the application of high tem­ with a length of 100 cm and an inner diameter of 5.5 cm at different
perature (700–900 ◦ C) during pyrolysis might produce biochar with temperatures (400 ◦ C, 600 ◦ C, and 800 ◦ C). A heating rate of 10 ◦ C min− 1
superior performance in heavy metals adsorption (Chatterjee et al., was applied to reach desired temperatures, then held to perform slowly
2020; Li et al., 2019; Tan et al., 2014). In contrast, BCs pyrolyzed at low pyrolyzed with 100 L min− 1 N2 with a residence time of 20 min, and
temperatures have relatively less carbonized and condensed structures finally cooled down to 50 ◦ C under continuous nitrogen gas flow. The
of carbon, which are suitable for improving soil fertility (Zambon et al., solid products (BCs) were collected, ground and sieved through 40 mesh
2016). Recent studies also point toward the need to systematically assess (0.45 mm). The BCs obtained under the pyrolysis temperatures of
the source and availability of feedstock, optimization of the biochar 400 ◦ C, 600 ◦ C, and 800 ◦ C were subsequently denoted as 400BC,
production process, development of low-cost adsorbent, demonstration 600BC, and 800BC, respectively.
at the field, possible regeneration of adsorbent, the complexity of the
interaction of adsorption conditions, and the applicability of such 2.2. Physicochemical analysis of biochar
technology at rural areas and developing countries (Baltrėnaitė-Gedienė
et al., 2020; El-Naggar et al., 2021; Hu et al., 2021). The pH value of BCs was determined by mixing 1:5 (w/v) of the BCs
Bermuda grass (Cynodon dactylon L.) is a perennial grass that grows and deionized (DI) water, and shaking on a mechanical shaker for 1 h
naturally. As one of the most common grasses throughout the world, (Enders and Lehmann, 2017) with a pH meter (Seven2Go™). The water
Bermuda grass is a salt-tolerant, drought-resistant, potentially invasive holding capacity (WHC) (Eq. (1)) was determined through the gravi­
and facultative halophyte species, which even grows in wastelands, metric method (Nguyen and Lehmann, 2009). The CEC was calculated as
barren lands and other native species dominated grasslands (Kovács the sum of all cations in meq/100 g dry BC as measured through a flame
et al., 2002). It has an overgrowth property and can cover an area up to atomic absorption spectrophotometer (FAAS, PerkinElmer 700) (Gaskin
25 m2 within 1.5 years (Horowitz, 1996). Compared with other grasses, et al., 2008). For dissolved organic carbon (DOC), 1.0 g of BCs was
it has the highest concentrations of total elemental contents for C, cal­ mixed with 20 mL of CaCl2 solution (0.01 M) and shaken overnight. The
cium (Ca), potassium (K), magnesium (Mg), sodium (Na), and phos­ homogenized mixture was centrifuged at 10,000 rpm for 10 min, and the
phorous (P), which makes it a more competitive species (Kovács et al., supernatant was filtered for measurement (Cabrera et al., 2011) through
2002). Due to the overgrowth characteristics, Bermuda grasses are cut the TOC-V CSH analyzer (Shimadzu). The dried BC samples were ground
routinely and discarded as green wastes commonly burned in the open for elemental analysis, in which C%, H%, N% and S% were determined
environment, which becomes an important source of carbon dioxide in through the CHNS analyzer (PerkinElmer 2400 Series II CHNS/O). The
the atmosphere. Therefore, the use of green waste (Cynodon dactylon L.) O% was further calculated by Eq. (2) (Aslam et al., 2017). The contents
for BC production is a reliable environmental-friendly solution to this of PTEs (As (III), Cd, Co, Cr (III), Ni, Pb and Zn) and other main elements
problem, which can be used to remove and immobilize contaminants in (Ca, Mg, Na, K and P) for the BC samples were determined through the
the contaminated soil and water. application of microwave-assisted digestion (MAD) with method 3052
To the best of our knowledge, no previous study had been performed for siliceous and organically based matrices (ASTM, 1996) and mea­
to compare the BCs produced from the green waste, Bermuda grass surement through an inductively coupled plasma-atomic emission
(Cynodon dactylon L.) at different pyrolysis temperatures and use them spectrometer (ICP-AES, PerkinElmer, Optima 8300).
to remove PTEs from wastewater or contaminated groundwater. This ms − md − mw
study aims to comprehensively and scientifically establish the properties WHC% = × 100 (1)
md
of biochar generated from Bermuda grass with close relation to the
characteristics of the feedstock itself. Pyrolysis at a wider temperature O% = 100% − (C% + H% + N% + S% + Ash%) (2)
range is attempted to explore the different potential applications of the
resulted biochar. Systematic measurement and reporting of the physi­ where, ms is the mass of filter paper saturated with DI water and sand-BC
cochemical and structural properties will help to promote the actual mixture, md is mass of dried filter paper and sand-BC mixture, and mw is
usage of the BC on site in the future. Comparative assessment with other the mass of wet filter paper used in Eq. (1).
types of biochar will allow the potential users to reach informed de­
cisions based on the local setting and needs. For instance, the use of BC
from Bermuda grass as a low-cost adsorbent for contaminated

2
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

2.3. Structural analysis of biochar where Ads% is the adsorption efficiency of the BCs, Ci is the initial
concentrations of PTEs in the diluted solutions and Cf is the final con­
Approximately 10 mg of each BC was dispersed into DI water (20 mL) centrations of PTEs in the filtered solutions.
for 30 min with ultrasound, and then taken onto a carbon-coated copper To examine the statistical significance of two different sets of data, a
grid. After drying, the micromorphology was observed by a scanning two-sample t-test was performed for the calculation of p-value by
electronic microscope (SEM, Hitachi). The pore volume, pore size and GraphPad Software, and a significance level of 0.05 was set in the cur­
surface area were measured by a surface area analyzer (NOVA2200e, rent study.
Quantachrome) through nitrogen adsorption/desorption at 77.4 K, ac­
cording to Barrett-Joyner-Halenda (BJH) and Brunauer-Emmett-Teller 3. Results and discussion
(BET) methods, respectively. Before analysis, the BCs samples were
degassed under vacuum conditions at 180 ◦ C for 24 h (Tan et al., 2014). 3.1. Physicochemical properties of biochars
Energy dispersive X-ray (EDX) was used to detect elemental mapping
and chemical composition of particles, which was determined by a super The properties of prepared BCs (mean ± SD) are shown in Table 1.
ultra-thin window X-ray analyzer (Silicon drift detector, Quantax, The BCs were all alkaline and with increasing pH values for BCs
Bruker) under 20 kV voltage at the highest possible resolution for generated at the higher temperatures. Among the generated BCs, 800BC
maximum percentage measurement of the elements (Purakayastha was of the highest pH value at 11.88. This could be resulted from the
et al., 2015). The mineralogical composition of BCs was determined by transformation of minerals and metal oxides into alkaline media during
X-ray diffraction (XRD) analysis. A Bruker 2D phaser system (D5005, the pyrolysis (Tan et al., 2014). The correlation of higher pH value to the
Bruker Germany) equipped with a Lynxeye detector was used for increased of pyrolysis temperature was observed in a previous study on
obtaining the XRD diffractogram from 5 to 50◦ 2θ at a scan rate of 40s sewage sludge (Méndez et al., 2013), which is caused by the release of
per 0.05◦ 2θ step size. The diffractograms of BCs were identified by alkali metal salts from the biomass materials (Chen et al., 2011) and the
comparing the 2θ values to ICDD-PDF software Diffrac Eva. The iden­ reduction of acidic functional groups on the surface during the loss of
tification of the minerals was also achieved by comparing to the stan­ oxygen (Filippis et al., 2013). Generally, the alkaline pH of BCs
dards files of joint committee on powder diffraction (Kim et al., 2013; contributed to the immobilization of PTEs, which is attributed to the
Srinivasan et al., 2015). For fourier transformed infra-red spectroscopy conversion of PTE into sediments with low solubility (Kistler et al.,
(FTIR), the BC samples were prepared by mixing 1.0 mg of each BC with 1987). Furthermore, the alkaline pH of BCs contributed to avoiding their
200 mg of KBr (FTIR reagent) and making pellets (7 × 13 mm) with own PTEs from being leached into the soil, and thus guaranteed the
hydraulic press at a pressure of 10 tons/cm2 for 2 min. The pellets were safety of soil.
then scanned the infra-red absorbance in the range of 400–4000 cm− 1 SEM images (Fig. 2) illustrated and supported that the surface area,
using GX spectrometer (PerkinElmer, USA) (Tan et al., 2014). The BC pore volume and pore size increased considerably with the rising in
samples were analyzed by a Renishaw inVia Raman microscope (514 nm pyrolysis temperature. Other research groups had reported similar
laser diode) using 4 cm− 1 spectral resolution with 10% laser power and findings, such as BCs derived from sewage sludge (Méndez et al., 2013;
15 min acquisition time. The Raman spectra were fitted in the range of Tan et al., 2014) or derived using different biomass materials such as
400–3500 cm− 1 by WiRE Raman software (version 3.6) (Mohanty et al., peanut, canola, corn and soybean (Yuan et al., 2011). The CEC of BCs
2013). The X-ray photoelectron spectroscopy (XPS) analysis of the BCs produced from Cynodon dactylon L. correlated well with the rising in
was conducted through achromatic X-ray radiation (1253.6 eV) with an pyrolysis temperature (p < 0.05). This is different from the previous
axis-165 (Kratos analytical Inc.). The curves of C1s and O1s were fitted studies where the CEC could either increase or decrease with the py­
by CasaXPS software (version 2.3.19, Casa Software Ltd.). For C1s and rolysis temperature depending on the feedstock and the temperature
O1s, the binding energy of 284.5 eV and 531.0 eV were used as the range applied (Méndez et al., 2013; Tan et al., 2014; Tomczyk et al.,
reference charges, respectively (Suliman et al., 2016). The thermogra­ 2020). Nevertheless, the CEC values (from 51.2 to 94.8 cmol kg− 1) of
vimetric analysis (TGA) and derivative thermogravimetric (DTG) ana­ BCs are generally comparable to the other biomass waste materials
lyses were achieved by measuring from 25 to 1000 ◦ C at a heating rate of (from 4.6 to 122.0 cmol kg− 1). The concentrations of DOC were 19.58,
10 ◦ C min− 1 and continuous flow of nitrogen gas (20 mL min− 1) using 22.28 and 25.94 mg kg− 1 after pyrolysis at temperatures of 400 ◦ C,
TGA/DTG 1 STAR® system (Mettler-Toledo, Switzerland) (Tan et al., 600 ◦ C and 800 ◦ C, respectively, which increased with the rising of
2014). pyrolysis temperature, comparable to other biomass waste materials
(Cheng et al., 2016). The concentrations of C (from 56.6 to 72.8% with
2.4. Adsorption performance of BCs for potentially toxic elements (PTEs) the increase in temperature) and P (from 2030 to 2920 mg kg− 1) in the
BC pyrolyzed at 800 ◦ C were relatively higher than those at lower
Stock solutions of As(III), Cd(II), Co(II), Cr(III), Ni(II), Pb(II) and Zn temperatures, which is consistent with the findings for BCs derived from
(II) (1000 mg L− 1 each) were prepared from their parental salts in DI branches on C, from 62.2 to 80.0% and for P, from 0.21 to 0.34%,
water. From the stock solutions, 250, 200, 150, 100, 50 and 25 mg L− 1 respectively (Zhao et al., 2017).
solutions were prepared through further dilutions. Each BC (100 mg) With the increase in pyrolysis temperature, the surface functional
was put into a glass bottle and added 10 mL of the diluted solutions of groups containing N, H, O and S were reduced with the release of vol­
each PTE to do multi-element mixed adsorption. The BC and diluted atiles. This is the reason why the concentrations of these elements were
solution were intensively mixed overnight by vortex and continuously rapidly dropped as temperatures increased, while the C was remained
stirred for 4 h at 25 ± 2 ◦ C. The suspensions were then passed through and converted into the fixed carbon (FC). Therefore, microstructures
filters (0.45 μm) and analyzed for concentrations of PTEs through the developed greatly, the atomic ratios (O/C and H/C) decreased and the
ICP-AES (ASTM, 1996). The ICP-AES was re-calibrated for standardi­ amorphous carbon increased, similar to the findings for BCs derived
zation (R2 > 0.99) after analysis of each 10 samples. All the adsorption from different sewage sludges (Tan et al., 2014). The decrease of O/C
experiments were performed in triplicates, and blanks were measured and H/C ratios also indicated that, during the pyrolysis of BCs, dehy­
for comparative analysis. The adsorption efficiency of BCs was deter­ drative polycondensation and dehydrogenative polymerization led to
mined as: the massive loss of oxygen and aliphatic hydrogen (Filippis et al., 2013).
The low O/C molar ratio of BC showed that less polar groups with higher
Ci − Cf
Ads% = × 100% (3) hydrophilicity were available (Chun et al., 2004). Although the H/C
Ci
molar ratio of the BCs as the degree of carbonization is always lower
than 0.5, it was shown that high aromaticity and carbonization could

3
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Table 1
Physicochemical properties of biochars pyrolyzed at 400 ◦ C, 600 ◦ C and 800 ◦ C.
Parameter 400BC 600BC 800BC Parameter 400BC 600BC 800BC

pH 9.1 ± 0.2 10.3 ± 0.2 11.9 ± 0.2 Ash (%) 13.2 ± 0.13 17.8 ± 0.18 22.5 ± 0.26
WHC (%) 42.8 ± 0.7 48.7 ± 0.8 56.3 ± 0.9 VM (%) 28.7 ± 0.1 20.9 ± 0.1 13.7 ± 0.1
Surface area (m2 g− 1) 17.3 ± 0.3 36.5 ± 0.6 48.8 ± 0.8 FC (%) 53.0 ± 0.1 61.3 ± 0.1 67.8 ± 0.1
Pore volume (cm3 g− 1) 0.041 ± 0.001 0.052 ± 0.001 0.089 ± 0.001 P (mg kg− 1) 2030 ± 46 2440 ± 53 2920 ± 68
Pore diameter (nm) 3.735 ± 0.05 3.759 ± 0.06 3.778 ± 0.08 Ca (mg kg− 1) 5640 ± 120 8640 ± 190 11680 ± 221
CEC (cmol kg− 1) 51.2 ± 1.7 71.4 ± 1.8 94.8 ± 1.8 Mg (mg kg− 1) 2430 ± 53 3630 ± 72 4420 ± 88
DOC (mg kg− 1) 19.6 ± 0.3 22.3 ± 0.4 25.9 ± 0.4 Na (mg kg− 1) 228 ± 5.7 363 ± 8.5 472 ± 11.6
C (%) 56.6 ± 0.9 65.3 ± 1.1 72.8 ± 1.2 K (mg kg− 1) 8130 ± 190 10270 ± 220 13820 ± 260
H (%) 3.56 ± 0.06 2.78 ± 0.05 1.98 ± 0.03 As (mg kg− 1) 0.02 ± 0.001 0.03 ± 0.001 0.03 ± 0.001
N (%) 1.88 ± 0.03 1.62 ± 0.03 1.22 ± 0.02 Cd (mg kg− 1) 0.03 ± 0.001 0.04 ± 0.002 0.04 ± 0.002
S (%) 1.93 ± 0.03 1.67 ± 0.03 1.19 ± 0.02 Co (mg kg− 1) 0.02 ± 0.001 0.02 ± 0.001 0.03 ± 0.002
O (%) 14.1 ± 0.23 10.6 ± 0.17 7.23 ± 0.12 Cr (mg kg− 1) 0.05 ± 0.004 0.07 ± 0.005 0.08 ± 0.005
O/C 0.25 ± 0.004 0.16 ± 0.003 0.10 ± 0.002 Ni (mg kg− 1) 0.04 ± 0.001 0.05 ± 0.001 0.06 ± 0.001
H/C 0.06 ± 0.001 0.04 ± 0.001 0.03 ± 0.001 Pb (mg kg− 1) 0.08 ± 0.01 0.13 ± 0.03 0.18 ± 0.04
Yield (%) 38.3 ± 1.06 34.3 ± 1.00 32.6 ± 0.93 Zn (mg kg− 1) 4.27 ± 0.70 6.16 ± 0.96 7.94 ± 1.30
Moisture (%) 1.36 ± 0.06 1.33 ± 0.04 1.29 ± 0.03

resist the further decomposition of BCs (Yuan et al., 2013). As the py­
rolysis temperature increased, the concentrations of major elements
were increased. Compared with 400 ◦ C, the concentrations at 800 ◦ C
were increased 51.8% for calcium, 45.0% for magnesium, 51.7% for
sodium and 41.1% for potassium. These observations were comparable
with the BCs derived from apple trees (Zhao et al., 2017).
Moreover, the yield, moisture content and VM content of BCs
decreased, while the ash content increased with the rising in pyrolysis
temperature. The devolatilization of volatiles at high pyrolysis temper­
ature increased the FC contents in BCs (Yang and Sheng, 2012), whereas
FC was considered a measurement to assess the recalcitrance and sta­
bility of BCs (Mašek et al.; Zhao et al., 2013). In addition, BCs with low
VM contents were found to be suitable for agronomic applications, due
to their role in nitrification and microbial activity (Deenik et al., 2009).
A number of studies had reported similar findings for BCs derived from
sewage sludge and crops, respectively (Méndez et al., 2013; Yuan et al.,
2013). The PTE concentrations of BCs were also investigated to identify
the risk of leaching into the soil, but these concentrations are within a
safe range (He et al., 2010; Tan et al., 2014).

Fig. 1. Nitrogen adsorption-desorption isotherms and pore size distribution


3.2. Structural analysis of biochar curves (inside) of biochars.

3.2.1. Micromorphology and surface characteristics


Ca from 2.97% to 3.45%, while a decrease in the intensity of O from
As shown in the SEM images (Fig. 2-up), the pore volume and pore
22.91% to 11.99% and S from 2.44% to 1.47%, with the increase of
size were increased with the rising of pyrolysis temperature (from
pyrolysis temperature. These findings were further supported by the
400 ◦ C to 800 ◦ C), which supported the physicochemical results of pore
results from CHNS and ICP-AES (Table 1). Similar findings for C, O and S
volume and pore diameter (Table 1). Based on physicochemical results,
intensities were found in the BCs prepared from green (leaf) waste with
the pore volume was significantly increased with the rising of pyrolysis
the increase in pyrolysis temperature (Sahota et al., 2018).
temperature (p < 0.05), which was comparable to previous studies (Tan
et al., 2014; Yuan et al., 2011). With the increase in pyrolysis temper­
3.2.3. XRD analysis
ature, the N2 adsorption of the BCs also significantly (p < 0.05) increased
The XRD patterns showed the intensities of the diffracted beams
(Fig. 1). According to the IUPAC classification, the N2
(Bragg’s law) or crystalline phases on the surface of BCs. All the spectra
adsorption-desorption isotherms corresponded to the type IV isotherm
(except 40◦ ) of the BCs intensities (Fig. 3) are sharp with a compara­
(Pavan and Andrew, 2019; Sing et al., 1985). These observations indi­
tively broad peak at 40◦ , indicating crystalline structure and carbon-rich
cated that (i) mesoporous structure (2–50 nm) was found on the BCs, (ii)
phase in the BCs (Srinivasan et al., 2015). Moreover, all the BCs showed
narrow necks and wide bodies (ink bottle pores) were observed, and (iii)
no peak in 15–20 region of 2θ, which confirmed the complete decom­
in the mesopores, capillary condensation occurred (Tan et al., 2014).
position of celluloses (Regmi et al., 2012). The identified peaks belong to
The distribution of pores was at the peak position of 3.82 nm, as shown
sylvite (KCl) or apatite (Ca5(PO4)3 (Cl,F,OH), sulphate (SO2− 4 ), whe­
by the pore size curves. In addition, the pore structures were observed as
wellite (CaC2O4.H2O), calcite (CaCO3), pericalse (MgO) and quartz
uniform and channelized. The formation of mesopores was consistent
(SiO2) at the 2θ region of 28.00, 32.01, 37.02, 40.00, 43.04 and 48.02,
with the BCs derived from sewage sludge (Rio et al., 2005).
respectively (Fig. 3). These peaks (compounds) were found in the BCs
prepared from timothy grasses (Mohanty et al., 2013).
3.2.2. EDX analysis
The EDX spectra could provide quantitative and qualitative analysis
3.2.4. FTIR analysis
of the pyrolyzed BCs (Purakayastha et al., 2015). The EDX spectra
The FTIR spectra (Fig. 4) showed a number of functional groups on
(Fig. 2-down) of the BCs showed an increase in the intensity of C from
the surface of the BCs, which played an essential role in the
62.77% to 74.14%, N from 1.25% to 2.19%, K from 4.63% to 5.21% and

4
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Fig. 2. SEM images (up) and EDX spectra (down) of 400BC, 600BC and 800BC.

Fig. 3. XRD spectra of 400BC, 600BC and 800BC.

Fig. 4. FTIR spectra of 400BC, 600BC and 800BC.


immobilization of PTEs by forming complexes (Ahmad et al., 2014; Lee
et al., 2010). The peaks in the 400-1500 cm− 1 (fingerprint region) exist
at 466.73, 619.09, 783.03, 1105.11, 1193.12 and 1434.91, while the pyrolysis temperature from 400 ◦ C to 800 ◦ C, which could justify the
peaks from 2500 to 4000 cm− 1 exist at 2558.71, 2852.46, 2919.96, results of CHNS for N, H and O (functional groups elements) and O/C
2962.39 and 3425.27 cm− 1 showed the availability of single bond and H/C ratios as shown in Table 1. The decreasing of functional groups
compounds, while the peaks from 1500 to 2000 cm− 1 (1647.06 and with the increase of pyrolysis temperature is comparable to findings
1872.71 cm− 1) indicated the presence of C– – O double bond stretching described in a previous study (Tan et al., 2014), where dehydrogenation
(carbonyl group) (Darban et al., 2016; Mohanty et al., 2013; Zhao and and reduction of carboxyl groups were observed.
Nartey). The peaks in various regions were attributed to different
components in the feedstock, such as 400-800 cm− 1 attributed to aro­ 3.2.5. Raman analysis
matic hydrogen, the C–H bond from lignin compound, 1100-1200 cm− 1 The crystalline and amorphous properties (structure evolution) of
attributed to C–O stretching from cellulose and C–C stretching from BCs were analyzed by Raman spectroscopy, due to its sensitivity to both
alkenes function, 1300–1500 attributed to O–H, C–H, and C–O stretch the structures (Mohanty et al., 2013). The spectra from 500 to 3000
from cellulose, hemicellulose, and alcohol groups, respectively. Peaks at cm− 1 indicated various types of bands during the curve fitting (Fig. 5).
1500-2000 cm− 1 indicated the presence of C– – O stretching from the The spectra of the BCs with three peaks at 1344.99 cm− 1, 1579.36 cm− 1
aromatic structure of hemicellulose and 2500–4000 attributed to and 2699.16 cm− 1 were commonly referred to “D”, “G” and “2D” bands.
alcohol compound or hydrogen bonded OH stretching from lignocellu­ The “D” band at 1344.99 cm− 1 was indicated as the disordered (defect)
lose compounds (Regmi et al., 2012; Suryanto et al., 2014). Further­ region of carbon for methyl groups (C–C vibration), the “G” band at
more, the surface functional groups were decreased with the increase of 1579.36 cm− 1 was the graphitic region for the in-pane vibrations of

5
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

presented in Table S1 (Supplementary Material). The C1s peaks were


measured at 284.90 (eV) and 288.00 (eV), while the O1s peak was at
533.00 (eV). The C1s peak of the BCs was deconvoluted into main peaks
and attributed to aliphatic carbon as C–C (alkanes) and C– – C (alkenes) at
284.90 (eV), and aromatic carbon as C– – C (graphene) at 288.00 (eV). At
the same time, the O1s peak was convoluted into a single peak at 533.00
(eV), which was attributed to C– – O (O-containing functional groups)
(Moulder et al., 1992; Srinivasan et al., 2015). It could be observed from
the spectra of C1s that all the BCs were rich in C and the intensities
remained high, while the intensity of O1s decreased lower, which could
be attributed to the increase in carbonization pyrolysis temperature
from 400 ◦ C to 800 ◦ C. These results were consistent with the previous
findings (Srinivasan et al., 2015). Furthermore, the XPS results could be
positively correlated to the EDX (Fig. 2-down), Raman (Fig. 5) and CHNS
(Table 1) results of all the BCs. All these findings showed that C was a
dominant element and O was in decreasing order with the increase in
pyrolysis temperature from 400 ◦ C to 800 ◦ C.

3.2.7. Thermal stability


The TGA and DTG curves of the BCs (Fig. 6) showed that distinctive
Fig. 5. Raman spectra of 400BC, 600BC and 800BC. weight loss started from 300 ◦ C and reached a maximum at 900 ◦ C. The
TGA curves (Fig. 6-A) were in the thermal stability order of 400BC <
sp2-bonded graphitic carbon structure of aromatic carbon (C– – C) and 600BC < 800BC from 200 to 900 ◦ C, which confirmed that 800BC was
the “2D” band at and 2699.16 cm− 1 was actually the second order of the comparatively more stable than 600BC and 400BC. The DTG curves
“D” band (one half of the 2D band), which also represented the methyl (Fig. 6-B) indicated the loss of weight rate from 100 to 1000 ◦ C within a
groups (C–C vibration) (Smith et al., 2016; Tuinstra and Koenig, 1970). time interval of 10–100 min. The maximum weight loss was recorded at
The band area of “G” peaks was less than “D” peaks in all the BCs, which 450 ◦ C, 750 ◦ C and 950 ◦ C for 400BC, 600BC and 800BC, respectively.
showed semicircle breathing of aromatic rings with the loss of O-con­ The absolute maximum weight loss of the BCs was in the order of 400BC
taining functional groups in BCs (Mohanty et al., 2013). Furthermore, > 600BC > 800BC, indicating the more availability of functional groups
the “D”, “G” and “2D” peaks in the Raman spectra of all the BCs reflected on the surface of 400BC than 600BC and 800BC, which supported the
the impacts of pyrolysis temperature. For instance, the intensities at results of FTIR spectra as shown in Fig. 5. Similar findings of TGA and
1344.99 cm− 1, 1579.36 cm− 1 and 2699.16 cm− 1 peaks increased in the DTG with the increase of pyrolysis temperature for BCs derived from
order of 800BC > 600BC > 400BC with 400 ◦ C, 600 ◦ C and 800 ◦ C sewage sludge (Tan et al., 2014).
(Fig. 5), which could justify the CHNS results of C contents (Table 1).
Moreover, the “2D” peak was lower than “D” peak in all the BCs, due to
the transition of non-crystalline graphite to sp2-amorphous carbon 3.3. Adsorptive performance of biochar on PTEs
structure (Jerng et al., 2011). Similar findings with the increase of py­
rolysis temperature had been reported by Guizani et al. (2017a) for BCs The adsorptive performances of all BCs were at significant (p < 0.05)
derived from beech wood. levels for all the PTEs with different concentration levels (i.e., 25, 50,
100, 150, 200 and 250 mg L− 1) as shown in Fig. 7 and the sorption
3.2.6. XPS analysis isotherms of biochar (Fig. S1 and S2) can be found in the Supplementary
The XPS surface quantitative spectroscopic results of the BCs were Material. Among the BCs, the highest removal rate (up to 96%) was
offered by the 800BC, followed by 600BC and 400BC. As shown in Fig. 7

Fig. 6. TGA (A) and DTG (B) curves of 400BC, 600BC and 800BC.

6
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Fig. 7. Adsorption performance of 400BC, 600BC and 800BC from 25 mg L− 1 solution (a), 50 mg L− 1
solution (b), 100 mg L− 1
solution (c), 150 mg L− 1
solution (d),
200 mg L− 1 solution (e), and 250 mg L− 1 solution (f) of As, Cd, Co, Cr, Ni, Pb and Zn.

(f), a fair adsorption was recorded in the case of 250 mg L− 1 solution for higher adsorption performance of BCs could be achieved. The findings of
all the BCs. The overall results revealed that the increase in pyrolysis this study revealed that the green waste (Cynodon dactylon L.) pyrolyzed
temperature was positively correlated with the adsorption of PTEs. BCs (especially for 800BC) demonstrated competitive adsorption per­
Similarly, the lower concentration of PTE(s) in the target solution, the formance (for PTEs) as compared to other studies reported as shown in

7
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Table 2
Comparison of the adsorption performance of biochar.
Feedstock Temperature Initial concentration of PTE (mg L− 1) Adsorption capacity (mg g− 1) Removal rate (%) Reference
(◦ C)

Perilla leaf 300–700 ◦ C As at 0.05–7.0 mg L− 1. Up to 3.8 mg g− 1


Up to 90% Niazi et al.,
(2018)
− 1 − 1
Potato peel 450 ◦ C As, Cd, Cu, Pb at 5 mg L , respectively. Up to 2.1 mg g for single metal and Up to 41% for single metal Hussain et al.,
up to 5.4 mg g− 1 for multi-metals and up to 27% for multi- (2017)
metals
1 1
Sewage 500–900 ◦ C Cd at 0–200 mg L− Up to 24.0 mg g− Up to 100% Tan et al.,
sludge (2014)
− 1 − 1
Dairy 350 C◦
Cd, Cu, Pb, Zn at 1–5 mM (or 64–1036 mg L , for each Up to 165.8 mg g for single metal Up to 100% for single metal Xu et al.,
manure metal as calculated based on the respective molecular and up to 158.0 mg g− 1 for multi- and up to 79% for multi- (2013)
weights). metals. metals
Rice husk Up to 29.0 mg g− 1 for single metal Up to 18% for single metal
and up to 18.6 mg g− 1 for multi- and up to 11% for multi-
metals. metals
Bermuda 400–800 ◦ C As, Cd, Co, Cr, Ni, Pb and Zn at 25–250 mg L− 1, for Up to 33.7 mg g− 1 for multi-metals. Up to 96% for multi-metals. Current study
grass each metal, respectively.

Table 2, such as up to 90% removal of As with biochar pyrolyzed at Na and K) were significantly increased; and (iii) The O/C and H/C ratios
700 ◦ C and 80% with biochar pyrolyzed at 300 ◦ C using Perilla leaf as were reduced. All these characteristics made the BCs optimum, which is
feedstock (Niazi et al., 2018), more than 99% Pb, and up to 51% Cd, Zn why the 800 ◦ C pyrolyzed BC showed comparatively better performance
49% and Cu 62% removal with dairy manured BCs (Xu et al., 2013), in terms of PTEs adsorption than the 600 ◦ C and 400 ◦ C pyrolyzed BCs.
removal of up to 40% Pb, 42% As, 20% Cu and 5% Cd with potato peels
derived BC (Hussain et al., 2017), and removal of up to 55%, 53% and Credit author statement
25% As by BCs derived from organic solid waste, sewage sludge and rice
husk, respectively (Agrafioti et al., 2014). Nisar Muhammad: Conceptualization; Investigation; Methodology;
There are multiple possible reasons for the observations on the su­ Data curation; Formal analysis; Validation; Visualization; Writing –
perior adsorption performance of the 800BC. These may include: (i) the original draft; Writing – review & editing; Liya Ge: Investigation;
presence of main elements in the biochars and their cation exchange Methodology; Data curation; Formal analysis; Writing – review & edit­
capacity (Tan et al., 2014; Xu et al., 2013), (ii) comparatively alkaline ing; Wei Ping Chan: Investigation; Methodology; Visualization; Writing
pH of the biochars (Tan et al., 2014), (iii) higher surface area of the – review & editing; Afsar Khan: Conceptualization; Methodology; Vali­
biochars, and (iv) larger pore size and pore volume of biochars (Agrafioti dation; Writing – review & editing; Mohammad Nafees: Conceptuali­
et al., 2014). The adsorptive performance of BCs was positively corre­ zation; Funding acquisition; Project administration; Resources;
lated with the presence of main elements (Ca, Mg, Na and K) and their Supervision; Validation; Writing – review & editing; Grzegorz Lisak:
CEC value. Therefore, the maximum release of these elements (cations) Conceptualization; Funding acquisition; Project administration; Re­
that occurred from BCs can provide more space for PTEs adsorption (Tan sources; Supervision; Validation; Writing – review & editing.
et al., 2014; Xu et al., 2013) and contribute to the increase of ion ex­
change. Similarly, for the pH values, the higher the temperature applied
during pyrolysis, the higher the pH of BCs generated, and therefore the Declaration of competing interest
higher PTEs adsorptive capacity could be achieved (Tan et al., 2014).
The 800BC has comparatively higher alkalinity (pH of 11.9 ± 0.2) than The authors declare that they have no known competing financial
the 600BC (pH of 10.3 ± 0.2) and 400BC (pH = 9.1 ± 0.2), which could interests or personal relationships that could have appeared to influence
be the possible reason to show relatively better adsorptive performance. the work reported in this paper.
Moreover, the higher surface area and pore volume of BCs also allowed
the 800BC to absorb more PTEs through site binding and pore trapping Acknowledgments
mechanisms, respectively (Agrafioti et al., 2014), as shown in Table 1.
This research was financially supported by the Higher Education
4. Conclusion Commission Pakistan under the International Research Support Initia­
tive Program (Grant No: 1-8/HEC/HRD/2018/8939, PIN: IRSIP 41
In this study, Bermuda grass (Cynodon dactylon L.) was applied as a BMS83). We acknowledge the collaborative contributions from
potentially overgrown grass waste to be converted into an useful biochar Department of Environmental Sciences, University of Peshawar Pakistan
material. It can be used in water/wastewater treatment and purification and Nanyang Technological University of Singapore for carrying out this
to remove PTEs (especially high temperature pyrolyzed BC, such as study.
800BC) for water reuse, or soil conditioning for restoration of degraded
(applicable for the low temperature pyrolyzed BC, such as 400BC). The Appendix A. Supplementary data
impacts of pyrolysis temperatures on the properties of BCs were also
systematically investigated. Three types of BCs were pyrolyzed from the Supplementary data to this article can be found online at https://doi.
green waste in a horizontal reactor through slow pyrolysis processes at org/10.1016/j.jenvman.2022.115385.
400 ◦ C, 600 ◦ C and 800 ◦ C, with a heating rate of 10 ◦ C min− 1 and 20
min residence time. The results showed that improvements in the References
physicochemical, structural and PTEs adsorptive properties of the BCs
were observed, with the increase of the pyrolysis temperature from Abujabhah, I.S., Bound, S.A., Doyle, R., Bowman, J.P., 2016. Effects of biochar and
compost amendments on soil physico-chemical properties and the total community
400 ◦ C to 800 ◦ C. The possible reasons include: (i) the surface functional
within a temperate agricultural soil. Appl. Soil Ecol. 98, 243–253.
groups were lost, which increased the pore size, pore volume and surface Agrafioti, E., Kalderis, D., Diamadopoulos, E., 2014. Arsenic and chromium removal
area; (ii) The concentrations of DOC, CEC and main elements (Ca, Mg, from water using biochars derived from rice husk, organic solid wastes and sewage
sludge. J. Environ. Manag. 133, 309–314.

8
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Ahmad, W., Najeeb, U., Zia, M.H., 2014. Soil Contamination with Metals. Elsevier Inc., Lehmann, J., Joseph, S., 2009. Biochar for Environmental Management: Science and
pp. 37–61 Technology, second ed. Earth Scan.
Angin, D., Şensöz, S., 2014. Effect of pyrolysis temperature on chemical and surface Li, C., Zhu, X., He, H., Fang, Y., Dong, H., Lü, J., Li, J., Li, Y., 2019. Adsorption of two
properties of biochar of Rapeseed (Brassica napus L.). Int. J. Phytoremediation 16, antibiotics on biochar prepared in air-containing atmosphere: influence of biochar
684–693. porosity and molecular size of antibiotics. J. Mol. Liq. 274, 353–361.
Aslam, Z., Khalid, M., Naveed, M., Shahid, M., 2017. Evaluation of green waste and Liang, C., Gascó, G., Fu, S., Méndez, A., Paz-Ferreiro, J., 2016. Biochar from pruning
popular twigs biochar produced at different pyrolysis temperatures for remediation residues as a soil amendment: effects of pyrolysis temperature and particle size. Soil
of heavy metals contaminated soil. Int. J. Agric. Biol. 19, 1427–1436. Tillage Res. 164, 3–10.
ASTM, 1996. Microwave Assisted Acid Digestion of Siliceous and Organically Based Mašek, O., Brownsort, P., Cross, A., Sohi, S., Influence of Production Conditions on the
Matrices, 3052-3051-3052-3020. Yield and Environmental Stability of Biochar. Elsevier, pp. 151-155.
Baltrėnaitė-Gedienė, E., Marčiulaitienė, E., Pranskevičius, M., Titova, J., Bhatnagar, A., Méndez, A., Terradillos, M., Gascó, G., 2013. Physicochemical and agronomic properties
Abu-Danso, E., 2020. Physicochemical properties of pyrogenic carbonaceous of biochar from sewage sludge pyrolysed at different temperatures. J. Anal. Appl.
product, biochar, syngenetically modified for its use in adsorption systems. Pyrol. 102, 124–130.
J. Environ. Eng. 146, 04020078. Mohanty, P., Nanda, S., Pant, K.K., Naik, S., Kozinski, J.A., Dalai, A.K., 2013. Evaluation
Bouraoui, Z., Jeguirim, M., Guizani, C., Limousy, L., Dupont, C., Gadiou, R., 2015. of the physiochemical development of biochars obtained from pyrolysis of wheat
Thermogravimetric study on the influence of structural, textural and chemical straw, timothy grass and pinewood: effects of heating rate. J. Anal. Appl. Pyrol. 104,
properties of biomass chars on CO2 gasification reactivity. Energy 88, 703–710. 485–493.
Cabrera, A., Cox, L., Spokas, K.A., Celis, R., Hermosín, M.C., Cornejo, J., Koskinen, W.C., Moulder, J.F., Stickle, W.F., Sobol, P.E., Bomben, K.D., 1992. Handbook of Photoelectron
2011. Comparative sorption and leaching study of the Herbicides Fluometuron and Spectroscopy.
4-Chloro-2-methylphenoxyacetic acid (MCPA) in a soil amended with biochars and Muhammad, Nisar, Ge, Liya, Haya Khan, Muhammad, Chan, Wei Ping,
other sorbents. J. Agric. Food Chem. 59, 12550–12560. Bilal, Muhammad, Lisak, Grzegorz, Nafees, Muhammad, 2021. Effects of different
Chatterjee, R., Sajjadi, B., Chen, W.-Y., Mattern, D.L., Hammer, N., Raman, V., Dorris, A., biochars on physicochemical properties and immobilization of potentially toxic
2020. Effect of pyrolysis temperature on physicochemical properties and acoustic- elements in soil - A geostatistical approach. Chemosphere 277, 130350.
based amination of biochar for efficient CO2 adsorption. Front. Energy Res. 8, 85. Muhammad, N., Nafees, M., Khan, M.H., Ge, L., Lisak, G., 2020. Effect of biochars on
Chen, X., Chen, G., Chen, L., Chen, Y., Lehmann, J., McBride, M.B., Hay, A.G., 2011. bioaccumulation and human health risks of potentially toxic elements in wheat
Adsorption of copper and zinc by biochars produced from pyrolysis of hardwood and (Triticum aestivum L.) cultivated on industrially contaminated soil. Environ. Pollut.
corn straw in aqueous solution. Bioresour. Technol. 102, 8877–8884. 260, 1–12.
Cheng, Q., Huang, Q., Khan, S., Liu, Y., Liao, Z., Li, G., Ok, Y.S., 2016. Adsorption of Cd Nguyen, B.T., Lehmann, J., 2009. Black carbon decomposition under varying water
by peanut husks and peanut husk biochar from aqueous solutions. Ecol. Eng. 87, regimes. Org. Geochem. 40, 846–853.
240–245. Niazi, N.K., Bibi, I., Shahid, M., Ok, Y.S., Burton, E.D., Wang, H., Shaheen, S.M.,
Chun, Y., Sheng, G., Chiou, G.T., Xing, B., 2004. Compositions and sorptive properties of Rinklebe, J., Lüttge, A., 2018. Arsenic removal by perilla leaf biochar in aqueous
crop residue-derived chars. Environ. Sci. Technol. 38, 4649–4655. solutions and groundwater: an integrated spectroscopic and microscopic
Cross, A., Sohi, S., Borlinghaus, M., 2010. Influence of feedstock and production examination. Environ. Pollut. 232, 31–41.
conditions on the stability of biochar in soils. Fuel 12, 13895-13895. Partey, S.T., Saito, K., Preziosi, R.F., Robson, G.D., 2016. Biochar use in a legume–rice
Darban, A.K., Arabyarmohammadi, H., Abdollahy, M., Ayati, B., 2016. The role of rotation system: effects on soil fertility and crop performance. Arch. Agron Soil Sci.
nanoporous biochars functional groups for immobilization of heavy metals in 62, 199–215.
aqueous solution. Mater. Sci. Forum 860, 43–46. Pavan, M.V.R., Andrew, R.B., 2019. BET Surface Area Analysis of Nanoparticles, 2.3.2-
Deenik, J.L., McClellan, A.T., Uehara, G., 2009. Biochar volatile matter content effects on 2.3.7.
plant growth and nitrogen transformations in A tropical soil. Western Nutrient Purakayastha, T.J., Kumari, S., Pathak, H., 2015. Characterisation, stability, and
Management Conference 8, 26–31. microbial effects of four biochars produced from crop residues. Geoderma 239–240,
El-Naggar, A., Ahmed, N., Mosa, A., Niazi, N.K., Yousaf, B., Sharma, A., Sarkar, B., 293–303.
Cai, Y., Chang, S.X., 2021. Nickel in soil and water: sources, biogeochemistry, and Regmi, P., Garcia Moscoso, J.L., Kumar, S., Cao, X., Mao, J., Schafran, G., 2012. Removal
remediation using biochar. J. Hazard Mater. 419, 126421. of copper and cadmium from aqueous solution using switchgrass biochar produced
Enders, A., Lehmann, J., 2017. Biochar: A Guide to Analytical Methods. Google Books. via hydrothermal carbonization process. J. Environ. Manag. 109, 61–69.
Filippis, P.D., Palma, L.D., Petrucci, E., Scarsella, M., Verdone, N., 2013. Production and Rio, S., Faur-Brasquet, C., Le Coq, L., Le Cloirec, P., 2005. Structure characterization and
characterization of adsorbent materials from sewage sludge by pyrolysis. Chem. Eng. adsorption properties of pyrolyzed sewage sludge. Environ. Sci. Technol. 39,
Trans. 32, 205–210. 4249–4257.
Gaskin, J.W., Steiner, C., Harris, K., Das, K.C., Bibens, B., 2008. Effect of low-temperature Sahota, S., Vijay, V.K., Subbarao, P.M.V., Chandra, R., Ghosh, P., Shah, G., Kapoor, R.,
pyrolysis conditions on biochar for agricultural use. Am. Soc. Agric. Biol. Eng. 51, Vijay, V., Koutu, V., Thakur, I.S., 2018. Characterization of leaf waste based biochar
2061–2069. for cost effective hydrogen sulphide removal from biogas. Bioresour. Technol. 250,
Guizani, C., Haddad, K., Limousy, L., Jeguirim, M., 2017a. New insights on the structural 635–641.
evolution of biomass char upon pyrolysis as revealed by the Raman spectroscopy and Shaaban, A., Se, S.M., Dimin, M.F., Juoi, J.M., Mohd Husin, M.H., Mitan, N.M.M., 2014.
elemental analysis. Carbon 119, 519–521. Influence of heating temperature and holding time on biochars derived from rubber
Guizani, C., Jeguirim, M., Valin, S., Limousy, L., Salvador, S., 2017b. Biomass chars: the wood sawdust via slow pyrolysis. J. Anal. Appl. Pyrol. 107, 31–39.
effects of pyrolysis conditions on their morphology, structure, chemical properties Sing, K.S.W., Everett, D.H., Moscou, L., Pierotti, R.A., Rouquerol, J., Siemieniewska, T.,
and reactivity. Energies 10, 796-796. 1985. International union of pure and applied chemistry physical chemistry division
He, Y.D., Zhai, Y.B., Li, C.T., Yang, F., Chen, L., Fan, X.P., Peng, W.F., Fu, Z.M., 2010. The commission on colloid and surface chemistry including catalysis* reporting
fate of Cu, Zn, Pb and Cd during the pyrolysis of sewage sludge at different physisorption data for gas/solid systems with special reference to the determination
temperatures. Environ. Technol. 31, 567–574. of surface area a. Pure Appl. Chem. 57, 603–619.
Horowitz, M., 1996. Bermudagrass (Cynodon dactylon): a history of the weed and its Smith, M.W., Dallmeyer, I., Johnson, T.J., Brauer, C.S., McEwen, J.S., Espinal, J.F.,
control in Israel. Phytoparasitica 24, 305–320. Garcia-Perez, M., 2016. Structural analysis of char by Raman spectroscopy:
Hu, S., Chen, C., Chen, X., Jing, F., Wen, X., 2021. A Field Study of Biochar Application improving band assignments through computational calculations from first
Impact on Adsorption and Accumulation of Cd in Paddy Soil and Rice. Archives of principles. Carbon 100, 678–692.
Agronomy and Soil Science, pp. 1–12. Srinivasan, P., Sarmah, A.K., Smernik, R., Das, O., Farid, M., Gao, W., 2015. A feasibility
Hussain, A., Maitra, J., Khan, K.A., 2017. Development of biochar and chitosan blend for study of agricultural and sewage biomass as biochar, bioenergy and biocomposite
heavy metals uptake from synthetic and industrial wastewater. Appl. Water Sci. 7, feedstock: production, characterization and potential applications. Sci. Total
4525–4537. Environ. 512–513, 495–505.
Jerng, S.K., Yu, D.S., Lee, J.H., Kim, C., Yoon, S., Chun, S.H., 2011. Graphitic carbon Suliman, W., Harsh, J.B., Abu-Lail, N.I., Fortuna, A.M., Dallmeyer, I., Garcia-Perez, M.,
growth on crystalline and amorphous oxide substrates using molecular beam 2016. Influence of feedstock source and pyrolysis temperature on biochar bulk and
epitaxy. Nanoscale Res. Lett. 6, 1–6. surface properties. Biomass Bioenergy 84, 37–48.
Kim, J.S., Sparovek, G., Longo, R.M., De Melo, W.J., Crowley, D., 2007. Bacterial Suryanto, H., Marsyahyo, E., Irawan, Y.S., Soenoko, R., 2014. Morphology, structure, and
diversity of terra preta and pristine forest soil from the Western Amazon. Soil Biol. mechanical properties of natural cellulose fiber from mendong grass (Fimbristylis
Biochem. 39, 684–690. globulosa). J. Nat. Fibers 11, 333–351.
Kim, P., Johnson, A.M., Essington, M.E., Radosevich, M., Kwon, W.T., Lee, S.H., Rials, T. Tag, A.T., Duman, G., Ucar, S., Yanik, J., 2016. Effects of feedstock type and pyrolysis
G., Labbé, N., 2013. Effect of pH on surface characteristics of switchgrass-derived temperature on potential applications of biochar. J. Anal. Appl. Pyrol. 120, 200–206.
biochars produced by fast pyrolysis. Chemosphere 90, 2623–2630. Tan, C., Yaxin, Z., Hongtao, W., Wenjing, L., Zeyu, Z., Yuancheng, Z., Lulu, R., Chen, T.,
Kistler, R.C., Brunner, P.H., Widmer, F., 1987. Behavior of chromium, nickel, copper, Zhang, Y.Y., Wang, H., Lu, W., Zhou, Z., Zhang, Y.Y., Ren, L., 2014. Influence of
zinc, cadmium, mercury, and lead during the pyrolysis of sewage sludge. Environ. pyrolysis temperature on characteristics and heavy metal adsorptive performance of
Sci. Technol. 21, 704–708. biochar derived from municipal sewage sludge. Bioresour. Technol. 164, 47–54.
Kovács, M., Engloner, A., Neámeth, N., Szirmai, O., Turcsányi, G., 2002. Chemical Tomczyk, A., Sokołowska, Z., Boguta, P., 2020. Biochar physicochemical properties:
composition of Bermuda grass (Cynodondactylon) in Hungary. Acta Agron. Hung. pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Biotechnol. 19,
50, 151–156. 191–215.
Lee, J.W., Kidder, M., Evans, B.R., Paik, S., Buchanan, A.C., Garten, C.T., Brown, R.C., Tuinstra, F., Koenig, J.L., 1970. Characterization of graphite fiber surfaces with Raman
2010. Characterization of biochars produced from cornstovers for soil amendment. spectroscopy. J. Compos. Mater. 4, 492–497.
Environ. Sci. Technol. 44, 7970–7974.

9
N. Muhammad et al. Journal of Environmental Management 317 (2022) 115385

Venegas, A., Rigol, A., Vidal, M., 2016. Changes in heavy metal extractability from Yuan, J.H., Xu, R.K., Zhang, H., 2011. The forms of alkalis in the biochar produced from
contaminated soils remediated with organic waste or biochar. Geoderma 279, crop residues at different temperatures. Bioresour. Technol. 102, 3488–3497.
132–140. Zambon, I., Colosimo, F., Monarca, D., Cecchini, M., Gallucci, F., Proto, A.R., Lord, R.,
Xu, X., Cao, X., Zhao, L., 2013. Comparison of rice husk- and dairy manure-derived Colantoni, A., 2016. An innovative agro-forestry supply chain for residual biomass:
biochars for simultaneously removing heavy metals from aqueous solutions: role of physicochemical characterisation of biochar from olive and hazelnut pellets.
mineral components in biochars. Chemosphere 92, 955–961. Energies 9.
Xu, Y., Chen, B., 2013. Investigation of thermodynamic parameters in the pyrolysis Zhang, J., Liu, J., Liu, R., 2015. Effects of pyrolysis temperature and heating time on
conversion of biomass and manure to biochars using thermogravimetric analysis. biochar obtained from the pyrolysis of straw and lignosulfonate. Bioresour. Technol.
Bioresour. Technol. 146, 485–493. 176, 288–291.
Yang, H., Sheng, K., 2012. Characterization of biochar properties affected by different Zhao, B., Nartey, O.D., Characterization and Evaluation of Biochars Derived from
pyrolysis temperatures using visible-near-infrared spectroscopy. ISRN Spectrosc. Agricultural Waste Biomass from Gansu, China, pp. 1-17.
2012, 1–7. Zhao, L., Cao, X., Mašek, O., Zimmerman, A., 2013. Heterogeneity of biochar properties
Yang, X., Wang, H., Strong, P.J., Xu, S., Liu, S., Lu, K., Sheng, K., Guo, J., Che, L., He, L., as a function of feedstock sources and production temperatures. J. Hazard Mater.
Ok, Y.S., Yuan, G., Shen, Y., Chen, X., 2017. Thermal properties of biochars derived 1–9, 256-257.
from Waste biomass generated by agricultural and forestry sectors. Energies 10. Zhao, S.X., Ta, N., Wang, X.D., 2017. Effect of temperature on the structural and
Yuan, H., Lu, T., Zhao, D., Huang, H., Noriyuki, K., Chen, Y., 2013. Influence of physicochemical properties of biochar with apple tree branches as feedstock
temperature on product distribution and biochar properties by municipal sludge material. Energies 10, 1–15.
pyrolysis. J. Mater. Cycles Waste Manag. 15, 357–361.

10

You might also like