Silver Adsorption On Biochar Produced From Spent Coffee Grounds: Validation by Kinetic and Isothermal Modelling
Silver Adsorption On Biochar Produced From Spent Coffee Grounds: Validation by Kinetic and Isothermal Modelling
Silver Adsorption On Biochar Produced From Spent Coffee Grounds: Validation by Kinetic and Isothermal Modelling
https://doi.org/10.1007/s13399-022-03491-0
ORIGINAL ARTICLE
Abstract
This study investigates silver adsorption on biochar produced from pyrolysis of spent coffee grounds (SCGs). Biochars were
produced from SCGs at temperatures between 500 and 1000 °C. SCG-derived biochars were then characterised by different
analytical methods, such as Brunauer-Emmet-Teller (BET), Fourier transform infrared (FTIR), X-ray diffraction (XRD), and
investigated for silver removal. The results revealed that the biochar produced at 500 °C offered a maximum surface area of
40.1 m2/g with a yield of 23.48% biochar and the highest silver adsorption capacity of 49.0 mg/g with 99.9% silver removal
efficiency. The morphology of adsorbed silver on biochar was determined using scanning electron microscopy–energy-
dispersive spectrometry (SEM–EDS), and XRD analyses, which showed an even distribution of silver on the biochar surface.
Furthermore, X-ray photoelectron spectroscopy (XPS) confirmed that part of the silver ions was reduced to form metallic
silver (Ag0)/silver nanoparticles (Ag NPs) during adsorption. The kinetics and isothermal evaluation suggested that silver
adsorption was dominated by the pseudo-second-order model and Langmuir isotherm, which means that silver adsorption
was mainly dominated by chemisorption and monolayer on biochar surface. Overall, this study suggests that 500 °C was
the most feasible pyrolysis temperature to produce SCG-derived biochar with suitable physicochemical properties that can
efficiently adsorb silver species from wastewater.
Keywords Pyrolysis kinetics · Biochar application · Adsorption kinetics · Spent coffee grounds · Silver pollution · Heavy
metal removal
13
Vol.:(0123456789)
Biomass Conversion and Biorefinery
wastewater to eliminate heavy metals like silver and other a key concern for climate change through global warming
contaminants in wastewater prior to its further applications [29]. A circular economy approach for SCGs will not only
or final discharge [10, 11]. reduce the environmental impact by eliminating pollution
Several advanced technologies, such as the activated from landfilling but also transform SCGs into value-added
sludge process, sequencing batch reactor (SBR), ion products.
exchange, and reverse osmosis, are available to eliminate Pyrolysis of SCGs is an environmentally sustainable
silver from wastewater, but most of these technologies are process to convert waste into value-added resources [30],
considered highly expensive and inefficient to remove sil- which could also be a potential route for safe disposal of
ver particles completely from the wastewater. In addition, about eight million tonnes of SCGs per year [31]. Optimisa-
current wastewater treatment facilities cannot completely tion of pyrolysis conditions, such as pyrolysis temperature,
remove silver due to its low concentration (µg/L level) and can improve the physicochemical properties of biochar to
its toxic effect on the bacterium-dominant treatment process enhance adsorption capacity. Moreover, the bio-oil fraction
[12]. Adsorption has been gaining substantial interest for can alleviate the pyrolysis energy cost and make the process
the efficient removal of silver from aqueous media, mainly economically more viable [32, 33].
because it is a simple, economical, and eco-friendly pro- Biochar produced from SCGs could be highly advanta-
cess [13]. Recently, many natural and synthetic adsorbents geous to support the circular economy of wastes and miti-
have been investigated for silver adsorption, for instance, gate heavy metal pollution at the same time. Based on the
activated carbons [14], clays [15], biowaste materials [16, literature review, only one investigation was found where
17], cellulosic materials [18, 19], zeolites [20], graphene SCG-derived biochar was used for silver removal in 2016
[21, 22], and biochar [13, 23]. However, compared to other [34]. More research could open better possibilities for SCG
adsorbents, biochar has been demonstrated as a promising waste disposal and mitigation of silver pollution. Therefore,
adsorbent for heavy metals because of its low production we aimed to prepare biochar using SCGs as feedstock. SCGs
cost, negligible environmental footprint, and significantly were pyrolyzed at a wide range of temperatures between
high adsorption capacity [24]. 500° and 1000 °C, biochar physicochemical properties were
Biochar can easily adsorb heavy metals [23]. The primary analysed, and the silver adsorption performance was investi-
interactions between biochar and heavy metal adsorption gated. A detailed adsorption kinetics evaluation and isother-
constitute electrostatic interactions, ion exchange, precipita- mal study of silver adsorption were also conducted to better
tion, and chemisorption [13]. Biochar is known to contain understand the adsorption mechanisms.
a net negative charge and can bind to positively charged
silver ions, thus could ultimately play a key role in silver
adsorption [25]. Functional groups present on the biochar
2 Materials and methods
surface help the reduction of heavy metals. For example, a
study confirmed that carboxyl groups in the biochar reduced
2.1 Materials
CrIV into C rIII [26]. In contrast, physicochemical properties
of biochar such as pore size, particle size, surface area, and
Wet SCGs were collected from Juliette’s Espresso, a cof-
the number and type of functional groups affect the adsorp-
fee shop at the JCU Townsville Bebegu Yumba campus,
tion capacity of heavy metals. For instance, high surface
Queensland, Australia. SCG samples were dried in an oven
area and porosity may provide additional sites and enhance
at 105 °C overnight to remove residual moisture. Silver
the adsorption of silver species [25]. Numerous biomass-
nitrate (AgNO3), 99.99% wt/wt solid, was outsourced from
derived biochar applied to remove various heavy metals,
Sigma-Aldrich (S6506-25G, lot#MKCJ2718).
such as arsenic (As), cadmium (Cd), lead (Pb), nickel (Ni),
zinc (Zn), copper (Cu), chromium (Cr), antimony (Sb), and
mercury (Hg) from wastewater [27]. 2.2 Methods
Spent coffee grounds (SCGs) are the waste by-product of
the coffee industry. Coffee grounds (CGs) are coffee beans 2.2.1 Characterisation of SCG biomass wastes
that have been crushed to a powder. Coffee is prepared by
forcing boiling water through the CGs. After the water has The composition of biomass wastes plays a vital role in the
been forced through, what remains is SCGs. An average cafe final properties of biochar, particularly the textural prop-
in a large Australian metropolitan area in 2016 produced erties like surface area and pore volume that enhance the
over three tonnes of SCGs (wet weight) per year, over 90% adsorption of a contaminant. Therefore, proximate and ulti-
of which directly went to landfilling as SCGs do not have mate analyses of SCG biomass were carried out to deter-
significant market value [28]. Landfilling of SCGs or any mine the percentage of moisture, volatiles, fixed carbon, ash
organic/food waste causes greenhouse gas emission that is content, elemental carbon (C), hydrogen (H), and nitrogen
13
Biomass Conversion and Biorefinery
(N) content. Oxygen (O) content was calculated from mass production to be posterior used for silver adsorption.
difference and considering the total mass of 100. This wide temperature range was selected based on our
Ultimate or elemental analysis (EA) was done in a previous studies and literature review. At least three
Costech Analytical Elemental Analyser 4010 (Valencia, CA, pyrolysis experiments were performed for each pyrolysis
USA) fitted with a zero-blank auto-sampler. The analyser condition. After each pyrolysis experiment, biochar was
uses dumas combustion to form CO2 and N2 that are detected weighed, grounded, and collected for characterisation
by a thermal conductivity detector (TCD). Standards were and adsorption testing.
used to formulate a calibration curve from which the C and Biochar yield (%) was calculated after each pyrolysis by
N percentages of the biomass sample were determined. the following Eq. (3):
Thermogravimetric analysis (TGA) is an important tech-
BC
nique to understand the thermal decomposition behaviour of Biochar yield = × 100 (3)
DBW
a substance. The ‘Discovery TA/SDT650’ analyser was used
for thermogravimetric analyses of SCG biomass samples. where BC is the weight of dry biochar and BDW is the
All thermal analyses were carried out in a nitrogen environ- weight of dry SCG feedstock.
ment (N2 flow rate, 50 ml/min) from 25 to 1000 °C with a The pH of biochar often influences the pH of the adsorp-
heating rate of 10 °C/min. The TGA curves were obtained tion media, which might impact the performance of the
from the plot of a residual mass percent against tempera- adsorbent. To determine the pH of each biochar sample,
ture. The moisture content (%) and volatile matters (%) were 0.5 g of biochar was mixed in 10 ml of double-distilled water
calculated from the TGA curve. Moisture content (%) was and shaken in an ultrasonic bath for 90 min to ensure good
calculated from the mass loss to 200 °C, and volatile matters contact between the biochar and water [35]. The pH was then
(%) were calculated from the mass loss from 200 to 500 °C. measured and recorded by the Thermo Scientific (Model:
To study pyrolysis kinetics, TGA and DTGA (derivative Orion Star A215) pH metre with an Orion 8157BNUMD
TGA) curves were further recorded at three different heating probe.
rates, 5, 10, and 15 °C/min. For each TGA analysis, about Specific surface area (Brunauer-Emmet-Teller, BET sur-
20 mg of the SCG biomass sample was analysed. face area) is a key property of any adsorbent material [36].
Ash content was determined by heating around 10 g of BET surface area, total volume, and pore size of pores were
SCG biomass sample (dry sample) at 600 °C for 6 h in a con- measured by an N 2 adsorption and desorption phenomena
ventional furnace (air atmosphere). The sample was removed using the Quantachrome Autosorb iQ3 gas adsorption ana-
from the furnace (at room temperature) and cooled in a des- lyser. Approximately 0.2 g of biochar was used for each run,
iccator for an additional hour, then weighed. Ash content and all biochar samples were analysed at least twice. Prior to
was calculated from Eq. 1: each analysis, the sample was firstly degassed under vacuum
conditions overnight at room temperature and then at 250 °C
B−C
Ash content(%) = × 100% (1) for an hour. The BET model was used to calculate specific
A−C
surface area and the BJH (Barrett-Joyner-Halenda) adsorp-
where B is the mass of the crucible plus the mass of biomass tion model for pore size and total pore volume.
after heat treatment, A is the mass of the crucible plus the A Fourier transform infrared (FTIR) is the equipment
initial mass of the biomass, and C is the mass of the crucible. for identifying surface functional groups, which attribute
Fixed carbon content (%) was calculated as below: the surface polarity or the surface charge for chemisorp-
(2)
Fixed carbon content (%) = 100 % − (Moisture % + Volatiles % + Ash %)
2.2.2 Pyrolysis experiments and characterisation of biochar tion [37]. A Thermo Science i7ATR- FTIR analyser was
used to estimate chemical functional groups presented
Pyrolysis of SCG biomass was done in a nitrogen envi- in different biochar samples. Prior to the experiment,
ronment (N2 flow rate, 3 L/min) using a fixed bed reactor biochar samples were manually grounded to improve the
and a conventional electrical tube (quartz) furnace–Ter- homogeneity and obtain a fine powder (less than 75 µm)
molab (Portugal). For each experiment, about 10 g of for the chemical analysis. A small amount of powdered
biomass waste was weighed in a quartz crucible and sample was placed on a diamond ATR sample holder, and
placed in the tube furnace. SCG biomass sample was a total of 64 scans were collected per sample between
pyrolysed at 500, 600, 700, 800, 900, and 1000 °C, with 500 and 4000 c m −1.
a heating rate of 10 °C/min and a residence time of 2 h X-ray diffraction (XRD) analysis was used to understand
to find the best pyrolysis temperature for SCG biochar the structural arrangement, or the crystallinity of biochar
13
Biomass Conversion and Biorefinery
obtained after pyrolysis. The Bruker Phaser D2 X-ray Powder 3 Results and discussion
Diffractometer, Cu radiation, was used. XRD spectra were
obtained from 5° to 65° with a scan step size of 0.02° and a 3.1 Characterisation of SCG biomass wastes
time step of 1 s.
The results of ultimate and proximate analyses of SCG bio-
2.2.3 Silver adsorption capacity and silver removal mass samples are presented in Table S1. The ultimate analysis
efficiency of SCG biochar data indicate that SCGs contained a high percentage of ele-
mental carbon (C) of 47.5% and fixed carbon of 18.8%. Both
To evaluate the silver adsorption capacity, 0.1 g of each carbon content values obtained in this study were similar to
biochar was added into 50 mL of 50 mg/L silver solution. the previous studies; for instance, elemental carbon and fixed
Silver concentration in the aqueous solution was meas- carbon were reported at 18.58% and 43.68%, respectively, for
ured before and after adsorption by inductively coupled rice straw [38]. In addition, SCGs had 71.5% of volatiles and a
plasma–atomic emission spectroscopy (ICP-AES). An low oxygen (O) content, approximately 43.7%, which might be
Agilent 5100 Varian Liberty Series II (Singapore) was advantageous to produce the bio-oil with a high heating value
used, and a series of silver standards were measured at a [39]. On the other hand, the proximate analysis revealed that
wavelength of 338.289 nm to calibrate the instrument. An SCGs contain only 2.5% ash, indicating their suitability for
independent silver standard solution was measured along biochar production via pyrolysis. Based on the ultimate and
with every batch of sample solution for quality control and proximate analysis data, it can be suggested that SCGs might
quantification purposes. The silver adsorption capacity (Qt be a suitable biomass to produce biochar.
in mg/g of biochar) at a given time t was calculated by TGA and DTGA analyses in Figure S1 show the ther-
Eq. (4): mal decomposition behaviour of SCG biomass. TGA and
( ) DTGA curves can be separated into three stages, as shown
C0 − Ct × V
Qt = (4) in Table S2. Stage 1 (up to 200 °C) is mainly attributed to
m the loss of moisture and other volatile substances extractable
where C0 (mg/L) is the initial concentration of the silver in from water. For instance, SCG biomass released a moisture
the aqueous solution, Ct (mg/L) is the concentration after content of 4.73%. Stage 2 is considered the main stage where
adsorption at time t, V (L) is the volume of the solution, the majority of the weight loss occurs, contributing to the
and m (g) is the amount of biochar used. The silver removal depolymerisation of biomass components, such as hemicel-
percentage (%) was obtained by Eq. (5). lulose, cellulose, and lignin. Hemicellulose is known to be
completely depolymerised at 350 °C, while the degradation
of cellulose starts above 180 °C and is completely degraded at
( )
C0 − Ct
% of silver removal = × 100 (5) 400 °C [29]. In contrast, depolymerisation of lignin polymers
C0
starts above 180 °C and is completely decomposed by 800 °C
After silver adsorption, biochar was separated from [40]. In stage 2, SCGs had a mass loss of 72.40%, correspond-
the liquid phase using filter paper that was then dried at ing to the temperature range of 200–550 °C. Similar results
105 °C overnight and analysed by XRD at room tempera- were reported in a previous study; a mass loss of 75% for
ture. Scanning electron microscopy combined with energy- SCGs was reported at this stage by Li et al. [29]. In addition,
dispersive spectroscopy (SEM–EDS) was used to examine SCG mass loss of up to 330 °C was attributed to hemicel-
the microstructure and silver distribution on biochar before lulose decomposition, and the last DTGA peak was found
and after adsorption tests. A Jeol JSM5410LVA SEM was to be at 392 °C, which is associated with the decomposition
used for the analysis where a beam of high-energy elec- of higher lignin and lipids [41]. The last stage, up to 800 °C,
trons scanned across the surface of a sample to obtain high is attributed to solid decomposition, where the weight loss
magnification and high-resolution images. rate is slower, and a slight weight loss could be due to char
X-ray photoelectron spectroscopy (XPS) analysis was consolidation [42]. For instance, the final stage weight loss
carried out to reveal the elemental composition and oxi- for SCGs was 1.95%.
dation states of biochar surfaces and adsorbed silver. A
Kratos Axis Supra connected with a monochromated Al 3.2 Kinetics analysis of biomass pyrolysis
Kα X-ray source and a helium lamp for UPS measurements
were used. The binding energy was calibrated by the C 1 s 3.2.1 Effect of heating rate
peak at 284.6 eV. XPS peak processing software was used
to deconvolute the peaks and identify different species of TGA and DTGA curves of SCGs were recorded at three
the same element. different heating rates of 5, 10, and 15 °C/min. Figure 1a–b
13
Biomass Conversion and Biorefinery
shows the effect of heating rates on TGA and DTGA curves. energy of a reaction means that this reaction requires a long
The decomposition behaviour was similar at three different residence time or high reaction temperature to start the
heating rates. It can be observed from Fig. 1b that the DTGA reaction. Ozawa-Flynn-Wall (OFW) and Kissinger–Aka-
peaks reached the maximum value at 10 °C/min, and the hira–Sunose (KAS) methods were used to obtain kinetic
DTGA curves shifted to a higher temperature zone when plots; results are shown in Fig. 1 (c–d), and the calcula-
increasing the heating rate. These findings indicate the une- tion details are reported in Supplementary data (Table S8).
ven heat transfer through the particles of SCG biomass. For The activation energy for SCGs using the OFW method
instance, the temperature in the centre of the SCG particles varied from 143.80 to 269.74 kJ.mol−1 with an average
is lower than the temperature on the surface [43]. The heat of 188.59 kJ.mol−1, and by the KAS method varied from
transfer efficiency was poor at a high heating rate compared 142.69 to 271.98 kJ.mol−1 with an average of 188.38 kJ.
to a low heating rate [44]. This heat transfer limitation and mol−1. In this study, the average values of activation energy
shifting of the DGTA curves towards the higher heating rate calculated by both the OFW and the KAS methods were
was also observed during the pyrolysis of coffee wastes and similar to previous studies where the activation energy val-
other biomasses in previous investigations [45]. ues were 196.08 and 197.26 kJ.mol−1 for SCGs from OFW
and KAS methods [47]. Overall, all kinetic parameters cal-
3.2.2 Determination of activation energy culated by the two methods were in good compliance with
each other and could be used to simulate the decomposition
TGA and DTGA data at three different temperatures were process under a specific condition.
used to derive the kinetic parameters using model-free meth- According to Figure S4, the required activation energy for
ods, as shown in Fig. 1 (c–d). To calculate the iso-conversion the removal of extra moisture was low. A gradual increase
plot of activation energy, Eα, with a degree of decomposition of the activation energy up to 0.5 degrees of conversion
α, the degree of decomposition was considered up to 90% was reported for the pyrolysis of SCGs. However, a slight
with a 10% size of each step according to the recommenda- decrease in activation energy from 0.5 to 0.7 degrees of
tion by the Kinetics Committee of the International Confed- conversion was noticed. This gradual increase suggests
eration for Thermal Analysis and Calorimetry (ICTAC) [46]. that SCG decomposition initiation requires high activation
Activation energy, Eα, is the minimum energy that is energy, and a slight decrease indicates the completion of
required to initiate a chemical reaction. The high activation the decomposition of cellulose and hemicellulose [48]. The
13
Biomass Conversion and Biorefinery
final stage, after 0.7 degree of conversion, required activa- Gibbs free energy mainly indicates the total energy increase
tion energy sharply increased to complete the decomposition during the decomposition reactions [50], where a gradual
of lignin which required extra energy [49]. increase of ∆G was reported with the degree of decomposi-
tion. High Gibbs free energy changes (∆G) of coffee grounds
3.3 Pyrolysis thermodynamics analysis were reported to be 173.09 kJ·mol−1 in a previous study
[54]. This relatively high Gibbs free energy change indi-
The apparent activation energies, Eα, calculated from the cates that the decomposition reactions consume significant
KAS method were used to derive the thermodynamic param- energy, which makes the pyrolysis process thermodynami-
eters, such as pre-exponential factor (A), the kinetic reac- cally unfavourable [52].
tion rate constant, k, enthalpy (ΔH), entropy (ΔS), and the Entropy (∆S) represents the degree of disorderness of
Gibbs free energy (ΔG) as shown in supplementary data the reaction system. The smaller value of ∆S indicates the
(Table S9). Pre-exponential factor (A) values exhibited wide system is near its thermodynamic equilibrium, which also
variations for a range of conversion degrees from 1015 to suggests the system is less reactive. In contrast, a high value
1021, which was attributed to the complex composition of of ∆S suggests the higher reactivity of the system to form an
SCGs as well as complex decomposition reactions during activated complex [50]. According to Table S9, the average
the thermal conversion. The low pre-exponential factors value of entropy was reported as 0.1 k Jmol−1 K−1 which is
(A < 109 min−1) also imply surface reaction, whereas the similar to the previous study for bamboo biomass [52]. Posi-
high pre-exponential factors (A ≥ 109 min−1) indicate the tive values of ∆S in SCGs imply a high degree of disorderli-
formation of the activated complex, which probably restricts ness in the system after decomposition and formation of less
rotation compared to the initial decomposition [50]. How- stable pyrolysis products than the biomass [55].
ever, similar values of pre-exponential factors, 1019 to 1021,
were reported for the KAS method in previous studies [49]. 3.4 Production and characterisation of SCG biochar
The enthalpy changes (∆H) demonstrate the energy dif-
ference between the reactant (biomass) and decomposition Table 1 shows SCG biochar yields and textural properties
products [51]. Positive enthalpy indicates an endothermic of the produced biochars at the studied temperatures of
reaction, while negative enthalpy indicates an exother- 500 to 1000 °C. Results revealed that increasing the tem-
mic reaction [52]. The enthalpy changes (∆H) at various perature gradually decreased the biochar yield. The highest
decomposition stages are presented in Table S9, with an SCG conversion into biochar was observed at 500 °C pyroly-
average value of ∆H for SCGs being 183.29 kJ·mol−1. It is sis, which produced 23.5% biochar, and the lowest was at
important to note that Py-GCMS analysis (Figures S2 and 1000 °C, producing 21.8% biochar. However, the reduction
S3, Tables S4 and S5) in this study and previous research in biochar yield from 500 to 1000 °C was not significant
showed that higher molecular weight compounds were and was expected according to TGA analyses [56]. These
identified in SCGs, and require higher energy to decompose results from our study are consistent with previous stud-
[41]. In addition, a small difference (~ 5 kJ·mol−1) in average ies that demonstrated the decrease in biochar yield with an
activation energy derived from both methods was observed, increase in temperature [57]. For instance, SCGs pyrolysed
which demonstrated that thermal decomposition is a favour- at 400 °C and 700 °C produced 43% and 26%, respectively
able conversion process [53]. [58]. Therefore, it could be suggested that lower pyrolysis
The Gibbs free energy changes (∆G) are also presented temperature could be helpful to achieve higher biochar yield
in Table S9, and the average value is 133.15 kJ·mol−1. The and could be comparatively more economical since less
Table 1 SCG-derived biochar yields at different pyrolysis temperatures and their physicochemical properties
Pyrolysis Elemental analysis (%) Biochar yield (%) pH Surface Pore Pore sizec (nm) Silver adsorption Silver
temperature areaa volumeb capacity (mg/g) removal
(°C) C H N Ox (m2/g) (cc/g) (%)
500 70.50 2.87 3.69 22.94 23.48 10.29 40.10 0.019 3.28 49.00 99.90
600 75.82 2.89 4.63 16.66 22.74 10.40 31.30 0.012 5.18 47.30 96.00
700 76.29 2.61 3.23 17.87 22.14 11.02 33.50 0.013 6.01 48.40 99.70
800 77.73 1.16 3.26 17.85 22.33 10.38 5.20 0.003 8.14 28.70 57.70
900 80.08 0.87 3.83 15.22 21.86 11.27 3.40 0.005 16.04 25.30 51.50
1000 83.53 2.01 2.04 12.42 21.78 10.53 7.50 0.005 7.13 23.60 48.10
x = calculated by the difference; aBET surface area; bMP (micropore analysis proposed by Mikhail) method; cBJH (Barrett, Joyner & Halenda)
method
13
Biomass Conversion and Biorefinery
energy would be used for pyrolysis. However, in addition summarised in Table 1. As expected, carbon content was
to biochar yield, it is critical to obtain biochar with suitable found to increase with the pyrolysis temperature, while
textural properties, such as high surface area, pore volume, oxygen content decreased. The maximum carbon content
and physicochemical properties like high acidic functional of 83.53% was found in the biochar produced at 1000 °C,
groups to achieve high adsorption capacity. and the minimum carbon content of 70.5% was found in the
Specific surface area (BET), pore volume, and pore size biochar produced at 500 °C. In contrast, the oxygen content
results of each SCG biochar sample are presented in Table 1. of 22.94% was the highest in biochar produced at 500 °C and
Specific surface areas of SCG biochar were reported as 40.1 the lowest at 12.42% in the biochar produced at 1000 °C.
to 3.4 m2/g which are very low compared to other ligno- These findings demonstrated that oxygen-containing groups
cellulosic origin-derived biochar, such as biochar produced decomposed at high temperatures either in the form of pyro-
from sugarcane bagasse and wood pyrolysed at 500 °C that lytic gases like C O2, CO, or N
Ox and eventually decreased
reported surface area of 202 and 316 m 2/g, respectively [59], the oxygen content and increased the carbon proportion in
whereas SCGs pyrolysed at 500 °C previously reported only the residual biochar [40]. Similar elemental composition of
1.46 m2/g [31]. One of the reasons might be the smaller SCG biochar was observed in previous studies [61].
number of micropores in the biochar matrix. For instance, Table 1 presents the pH results of SCG biochar produced
in this study, SCG biochar produced at 500° and 800 °C pro- at various pyrolysis temperatures. Based on the experimen-
vided a pore volume of 0.019 and 0.003 cc/g, respectively, tal results, all SCG biochars were alkaline in nature as the
but BET surface area of 40.1 and 5.2 m 2/g, respectively. A pyrolysis process led to the formation of alkaline miner-
maximum BET surface area of 40.1 m2/g was found for the als from micronutrients in the biomass [64] with a slightly
biochar produced at 500 °C, and the minimum was 3.4 m2/g non-linear change with temperature. However, alkalinity was
at 900 °C. slightly increased with the increase in pyrolysis temperature,
Additionally, biochar produced at 500 °C had a mini- which indicates a decrease in acidic functional groups, such
mum pore size of 3.28 nm with a maximum pore volume of as – OH, and – COOH [65].
0.019 cc/g. These results were significantly higher than in Peaks identified in FTIR spectra of different SCG biochar
the previous study where the specific surface area and pore samples are summarised in Table S3. The results revealed
volume were 11.0 m 2/g and 0.009 cc/g, respectively, under that lower pyrolysis temperatures produced biochar with
the same pyrolysis conditions for SCG biochar [60]. The a larger number of functional groups, demonstrating the
pore size increased with the increase in pyrolysis tempera- presence of a variety of organic compounds in the biochar.
ture. Biochar produced at 900 °C had the largest pore size For example, biochar produced at 500 °C showed peaks at
of 16.04 nm diameter, providing the lowest surface area. 1050, 1470, and 1890 cm−1, which can be attributed to the
This might result from collapsing of pores, which leads to presence of oxygen-containing alcoholic groups, aliphatic
the formation of larger pores, but at 1000 °C, these pores C − H groups, and aromatic C − H stretching, respectively.
might shrink again. Pore size increased gradually till 900 °C, Conversely, increasing pyrolysis temperature started consist-
then again decreased to 7.13 nm at 1000 °C. These results ently decreasing the number of functional groups in the bio-
indicate the weakness of the BET surface area measurement char, suggesting the volatilisation of organic compounds into
technique, such as the N 2 adsorption–desorption process, pyrolytic gases and their condensation into the bio-oil. For
where complete desorption might not be able to be achieved instance, biochar formed at 600 °C showed peaks at 1470,
and, consequently, is unable to accurately measure the sur- 1900, and 2100 cm−1, which can be ascribed to the occur-
face area of SCG biochar. For instance, BET analysis of rence of organic compounds with aliphatic C − H, aromatic
these biochars took a long time, and the results provided C − H, and C≡C, respectively. In addition, a temperature of
a non-linear BET curve and a non-equilibrium system of 1000 °C produced biochar with only aromatic compounds
adsorption–desorption of N 2, which might be responsible C − H. The analyses show that higher pyrolysis temperature
for less accurate results [30]. Degassing the pore and high enhances the decomposition of organic compounds into
burn-off degrees or long residence time might be responsi- pyrolytic gases and bio-oil, leaving the biochar with a very
ble for widening the micropores, thus providing low surface few less or a negligible number of functional groups that
area [61]. In addition, the BET technique is not valid for are necessary for the adsorption of heavy metals. There-
macroporous materials, and SCG biochar might be one of fore, lower pyrolysis temperatures should be applied for
these materials based on SEM images [62]. This observation biochar production to generally result in a variety of func-
was similar to the previous study which reported the inac- tional groups in biochar, which could play a pivotal role as
curate surface area from the BET method for the material adsorption sites.
having mesopores larger than 2 nm [63]. The XRD analysis spectra of different biochars
Elemental analysis results of SCG biochars produced produced under different pyrolysis temperatures are
at different temperatures between 500° and 1000 °C are presented in Fig. 2a. XRD patterns showed nearly
13
Biomass Conversion and Biorefinery
amorphous structures of all biochars. The broad peaks where high pyrolysis temperature (> 500 °C) reduced the
at 2θ of 24.5° and 43° regions indicate amorphous carbo- functional groups of biochar and thereby the adsorption
naceous structures. However, amorphous carbonaceous capacity for Cd reduced significantly [68]. On the other
structures rearranged at higher temperatures to form a hand, biochars that contained a significant number of func-
graphitic structure are characterised by broad peaks at tional groups favoured the chemisorption mechanism for
2θ of 24.5° [66]. silver adsorption [69].
Table 2 presents the comparison of silver adsorption
3.5 Silver adsorption study by SCG biochar capacities by biochars derived from different biomass
wastes. According to Table 2, the maximum silver adsorp-
The results of the silver adsorption performance of SCG- tion capacity achieved was 137.4 mg/g with a 2000 mg/L
derived biochar are shown in Fig. 2c and Table 1, which of Ag+ solution, which was high compared to other studies.
revealed that biochar produced at 500 °C showed a maxi- This difference can be explained by the direct impact of
mum adsorption capacity of 49.0 mg/g and 99.9% sil- initial high concentration and different types of biomasses
ver removal performance. In contrast, biochar produced [70]. However, for a similar initial concentration of Ag+,
at 1000 °C provided the lowest adsorption capacity of SCGs in this study offered the highest adsorption capacity.
23.6 mg/g and 48.1% silver removal. In addition, SCG bio- Despite having a low specific surface area, the SCG bio-
char produced at 600 and 700 °C provided the adsorption char showed a higher adsorption capacity (49.0 mg/g after
capacity of 47.3 mg/g and 48.4 mg/g with 96% and 99.7% 24 h) than biochar produced from biosolids (43.9 mg/g),
silver removal, respectively. The decrease in adsorption which had a higher specific surface area (38 to 151 m2/g)
capacity and % of silver removal with increasing pyrolysis [13]. Similarly, wood biochar having a specific surface
temperatures can be attributed to textural properties, such area of 83.6 m2/g provided the maximum silver adsorption
as low surface area, pore volume, and less acidic groups capacity of 19.1 mg/g which was half of the SCG biochar
in biochars. Since the biochars with a low surface area [71]. This observation demonstrated that SCG biochar has
and negligible functional groups showed low adsorption substantial potential for Ag removal from aqueous waste
capacity, it can be estimated that physisorption (that might mediums, mainly because of the presence of various oxy-
include Van der Waal forces) played a major role in silver gen-containing functional groups and other metals that
removal [67]. A previous study confirmed similar findings participated in different Ag removal mechanisms.
13
Biomass Conversion and Biorefinery
Table 2 Comparison of silver adsorption capacity of different bio- adsorbed by the adsorption sites on the outer surface of the
mass-derived biochars biochar matrix, and after the first 2 h, silver ions diffused
Type of biomasses Adsorption Initial silver Reference into the pores slowly. This second step could be explained
capacity concentration by intraparticle diffusion, which reduced the adsorption rate.
(mg/g) (mg/L) The results confirm the contribution of both chemical (slow)
Biofuel residue 90.06 50–100 [25] and physical (rapid) adsorption phenomena.
Biosolids 43.9 100–1000 [13] Compared to all kinetic models, the pseudo-second-order
Sunflower husk 22.9 100–300 [71] model fits well with the experimental data, having R2 val-
Rapeseed 26.9 ues of 0.9990, followed by the pseudo-first-order model
Wood waste 19.1 (Table S6). Experimental data also showed that 61% sil-
Gracilaria R. algae 4.7 50–100 [72] ver was adsorbed in the initial 5 min. The fitting of results
Pomelo peel 137.4 2000 [70] with the pseudo-second-order model suggests that the silver
Pine sawdust and 0.60–5.25 170 [73] adsorption was mainly dominated by chemisorption, carried
bagasse
out by the functional groups and minerals present on the bio-
SCGs 46.2 50 [34]
char surface, and physisorption that mainly involves the Van
SCGs 49.0 50 This study
der Waals forces. Additionally, a low k2 value of 0.0010 g/
(mg h) indicated that the adsorption process was driven by
the number of unoccupied sites in biochar [74]. The pseudo-
3.6 Silver adsorption kinetics second-order model also explained the additional complex
mechanisms, such as surface adsorption, external liquid film
The physical and chemical properties of the biochar can diffusion, and precipitation [13]. Our results are consistent
affect the adsorption behaviour of Ag. Kinetic studies can with previous studies that showed that biochar-driven heavy
provide vital information to understand the physicochemi- metal removal follows the pseudo-second-order and pseudo-
cal mechanism of silver adsorption on the SCG biochar, first-order kinetic models. For example, a study carried out
involving chemical binding and mass transport. Based on the application of sludge-based biochar for Pb(II) adsorption
silver adsorption performance, biochar produced at 500 °C and showed that the adsorption kinetics followed the pseudo-
was selected and used to evaluate the kinetics and isother- second-order and pseudo-first-order models since the values
mal behaviour of silver adsorption from an aqueous solu- of R2 were greater than 0.99 [75]. In another study, bio-
tion. The adsorption capacity was measured by varying the char was applied for the adsorption of Ni(II) and Co(II) and
contact time from 5 min to 24 h. Finally, the experimental the results revealed that both metals primarily followed the
data were compared with different kinetic models, such as pseudo-second-order and then pseudo-first-order kinetics,
pseudo-first-order (eq. S16), pseudo-second-order (eq. S17), suggesting the adsorption of Ni(II) and Co(II) was controlled
intraparticle diffusion (Weber and Morris) (eq. S18), and the by chemisorption process [76].
Elovich equation (eq. S19); results are shown in Fig. 3a and
Table S6. The results suggest that the silver adsorption on 3.7 Silver adsorption isotherms
the biochar surface occurred in two steps. For instance, the
adsorption of silver ions was rapid at the beginning (up to Adsorption isotherms were studied to understand the affinity
2 h contact time) and reached equilibrium after around 8 h. between silver ions and biochar, and the results are shown in
These findings demonstrate that silver ions were rapidly Fig. 3b and Table S7. The adsorption isotherms showed the
13
Biomass Conversion and Biorefinery
impact of the initial silver concentration in the solution on recognised as pure silver based on the face-centered cubic
silver removal capacity by biochar under a constant pH and structure (JCPDS, file No. 04–0783). These findings dem-
isothermal condition (22 ± 2 °C). The same biochar showed onstrate that the silver ions were reduced during adsorption
an adsorption capacity of 28.1 mg/g in 25 mg/L silver solu- and produced Ag NPs on the biochar surface.
tion and 67.1 mg/g in 400 mg/L silver solution with the same The presence of silver particles and their oxidation state
retention time of 5 h. These findings indicate that higher in biochar samples was further confirmed by XPS analy-
concentration increases the driving force of mass transfer sis. Figure 2d shows the results of XPS analysis for SCG
during the interaction between the solid and liquid phases biochar after Ag+ adsorption, showing the electron bind-
and increases adsorption capacity [13]. ing energies of Ag 3d5/2 and Ag 3d3/2 orbitals. A separate
The experimental data were compared with different iso- table of XPS analysis can be found in supplementary data
therm models, such as the Langmuir model (eq. S20), Fre- Table S10. For silver ions, Ag 3 d5/2 and Ag 3d3/2 orbitals can
undlich isotherm (eq. S21), and the Temkin isotherm (eqs. be further divided into 368.56/369.34 and 374.54/375.29 eV,
S24 and S25). The Freundlich model assumes mathemati- respectively [81]. The peaks at the binding energy of
cally unlimited adsorption sites on heterogeneous adsorp- 368.56/369.34 eV can be attributed to the presence of
tion surfaces and the Langmuir model is an ideal monolayer Ag+ in the biochar, while peaks at 374.54/375.29 eV can
adsorption model, assuming that all the adsorption sites have be ascribed to metallic A g0 [82]. According to Table S10,
equivalent adsorption energies and there are no mutual inter- 34.60% and 6.54% of the atomic concentration of Ag+ were
actions between adsorbed molecules [77]. Among all iso- reduced to metallic A g0 and Ag2+, respectively. These find-
thermal models, the Langmuir model fits the experimental ings confirmed the chemisorption phenomenon, which
data better, as the R2 value of the Langmuir model was the was revealed by the kinetic and isothermal studies of silver
highest at 0.987, whereas Freundlich and Temkin’s mod- adsorption by SCG biochar.
els provided R2 values of 0.8589 and 0.9131, respectively. Figure 4a–c illustrate the SEM images of SCG biochar
Langmuir’s model indicates monolayer adsorption of Ag produced at 500 °C before adsorption and show that biochar
on a homogeneous biochar surface, which is in line with had a rough, rigid, uneven, and irregular cracked surface. Fig-
the adsorption kinetic analysis. However, previous studies ure 4d–f show SEM images of biochar after silver adsorption.
showed that the Langmuir model was also used to simulate Significant changes can be seen in the biochar surface com-
the adsorption process by chemical precipitation [78]. In pared to the untreated biochar samples, showing the homo-
addition, the separation factor, RL, indicates the nature of geneous distribution of silver particles as white spots on the
the adsorption process. Generally, RL > 1 indicates unfavour- biochar surface. These white spots are silver nanoparticles that
able; RL < 1 indicates favourable, RL = 1 indicates linear; and were confirmed by SEM–EDS analysis [67]. Small and large
RL = 0 indicates irreversible adsorption phenomenon. In this white spots on the surface indicated silver particles less than
study, RL was 0.1031, which indicates that the adsorption 100 nm in diameter. Large white clusters are the agglomerates
was favourable; thus, the affinity or bond strength was strong of Ag NPs, which might be produced after the surface reduc-
between the adsorbate and adsorbent [31]. The strong affin- tion of silver ions (Ag+) [78].
ity can be attributed to the high pore volume in the biochar In addition, EDS mapping (shown in Fig. 4g) confirms
that provided better accessibility and increase sorption sites the adsorption of silver by biochar, confirming the presence
for Ag species. In addition, the presence of oxygen-contain- of 3.18 wt% silver on SCG biochar. Other major elements
ing functional groups on the biochar surface is also respon- such as carbon, oxygen, and trace amounts of other ele-
sible for enhanced affinity for Ag ions [79]. ments, such as magnesium, were also found in the biochar.
The biochar samples successfully demonstrated the adsorp-
tion of silver, suggesting that biochar could be an efficient
3.8 Characterisation of silver‑loaded biochar adsorbent for silver removal from wastewater.
13
Biomass Conversion and Biorefinery
Fig. 4 a–c SEM images of SCG biochar produced at 500 °C before f refers to the silver distribution on the biochar; g EDS mapping spec-
silver adsorption experiments; d–f SEM images of SCG biochar after tra showing the elemental quantification in biochar after Ag adsorp-
silver adsorption; d backscattering image where the silver nanoparti- tion
cles are white spots; e refers to the carbon distribution in the biochar;
Ag adsorption on SCG biochar can be proposed. Figure 5 that Ag ions can be physically adsorbed to the biochar via
shows the possible mechanisms of Ag adsorption on SCG the Van der Waals forces. Ag ions can be adsorbed on the
biochar. The most common Ag adsorption mechanism can surface of the biochar or can diffuse into the pores of the
be physical adsorption. SEM and BET analyses indicate the biochar. The strength of physical adsorption is generally con-
porous structure of the biochar; therefore, it can be predicted sidered weak and may depend on the surface area and pore
13
Biomass Conversion and Biorefinery
volume. The BET analysis in our study suggests that the high constructing electrostatic interactions with positively charged Ag
temperature above 800 °C produced biochar with low sur- ions and resulting in the enhanced adsorption capacity.
face area and pore volume, while 500 °C generated biochar Another possible Ag removal mechanism may involve cation
with a maximum surface area of 40.10 m 2/g and pore volume exchange. The functional groups on the biochar surface play an
of 0.019 cc/g, which suggests that the latter biochar would important role in the cation exchange mechanism. FTIR and
favour physical adsorption mechanism for the Ag removal. EDS analyses confirm the presence of –COOH and elements
Physical adsorption of heavy metals on biochar has been like Ca, Na, and K. Thus, it can be predicted that C a2+ and K+
demonstrated in previous studies. For instance, Zhang et al. can be replaced on biochar with positively charged Ag ions like
(2019) demonstrated Pb removal from an aqueous solution Ag2+ or A
g+. The cation exchange of heavy metals on biochar is
using sludge-based biochar and revealed from isothermal and well known in the literature [83]. Complexation, which involves
kinetics analyses that the main mechanism for Pb removal the formation of complex structures with specific metal–ligand
was physical adsorption rather than chemical adsorption [75]. interactions and precipitation that includes the formation of sol-
The main mechanism for Ag removal that can be estimated ids, is another known mechanism contributing to the removal
from our analyses is electrostatic interactions between Ag spe- of heavy metals from biochar [84]. XPS analysis confirmed
cies and surface-charged biochar. EDS analysis showed the pres- the presence of metallic Ag species in the biochar, which sug-
ence of various elements which could be presented in positively gests that the reduction process also took place during the Ag
or negatively charged forms and have the tendency to interact adsorption.
with oppositely charged Ag ions. The literature suggests that low
pH offers a high concentration of H+ in the solution, and thus
the biochar surface groups combine with H+, which reduces the 5 Conclusions
negative charge, consequently resulting in the reduced adsorp-
tion capacity for heavy metal cations [27]. On the other hand, at This study investigated the SCG biomass pyrolysis and the
the high pH of the aqueous solution, biochar surface groups lose effect of pyrolysis temperature from 500 to 1000 °C on the
protons, which increases the negative charge, and, subsequently, properties of SCG biochar and their implication for silver
the adsorption capacity for heavy metal cations increases [27]. adsorption from water solution. Results revealed that 10 °C/
Since our study was conducted in high pH solution, the bio- min was the optimum heating rate for the pyrolysis of SCGs.
char is predicted to contain more negatively charged species, Py-GCMS and pyrolysis kinetic studies showed that SCG
13
Biomass Conversion and Biorefinery
biomass had a slightly different composition than lignocel- Consent for publication All authors agreed on the publication of this
lulosic biomasses (Figures S2 and S3, Tables S4 and S5). research work.
Fluctuation of activation energy and higher molecular weight Competing interests The authors declare no competing interests.
compounds demonstrate the complex decomposition behav-
iour of SCG biomass. In addition, Py-GCMS analysis opened Open Access This article is licensed under a Creative Commons Attri-
the possibility of collecting bio-oil during pyrolysis to off- bution 4.0 International License, which permits use, sharing, adapta-
tion, distribution and reproduction in any medium or format, as long
set the biochar production cost. Biochar produced at 500 °C as you give appropriate credit to the original author(s) and the source,
provided the maximum yield of 23.48% biochar, the highest provide a link to the Creative Commons licence, and indicate if changes
surface area of 40.1 m2/g, and pore volume of 0.019 cc/g. were made. The images or other third party material in this article are
Also, the silver adsorption performance showed that the bio- included in the article's Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
char produced at 500 °C achieved a maximum adsorption the article's Creative Commons licence and your intended use is not
capacity of 49.0 mg/g with 99.9% silver removal efficiency. permitted by statutory regulation or exceeds the permitted use, you will
In contrast, biochars produced at 1000 °C achieved the lowest need to obtain permission directly from the copyright holder. To view a
adsorption capacity of 23.6 mg/g with 48.1% silver removal copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
efficiency. Low adsorption efficiency by biochars produced at
higher temperatures could be attributed to their low surface
area and pore volume and the smaller number of functional
groups. SEM–EDS results showed an even distribution of sil- References
ver on the biochar surface. XPS analysis confirmed the pres-
1. Jones ER et al (2021) Country-level and gridded estimates of
ence of metallic silver (Ag0) on the biochar surface, indicat- wastewater production, collection, treatment and reuse. Earth
ing the partial chemical transformation of silver ions during System Sci Data 13(2):237–254
adsorption. The kinetic study validated that silver adsorption 2. Islam MA et al (2022) Silver ions and silver nanoparticles
by SCG biochar supported the pseudo-second-order model, removal by coffee derived biochar using a continuous fixed-
bed adsorption column. J Water Process Eng 48:102935
which revealed the presence of chemisorption. Additionally, 3. Mahmoud ME, El-Bahy SM, Elweshahy SM (2021) Decorated
isothermal analysis established that the batch process of silver Mn-ferrite nanoparticle@ Zn–Al layered double hydroxide@
removal was dominated by homogeneous monolayer adsorp- Cellulose@ activated biochar nanocomposite for efficient reme-
tion. This study indicates that biochar production from SCGs diation of methylene blue and mercury (II). Biores Technol
342:126029
can be used as an adsorbent that is a sustainable approach to 4. Agoro MA et al (2020) Heavy metals in wastewater and sewage
mitigate silver pollution while avoiding the hazardous con- sludge from selected municipal treatment plants in eastern cape
sequences of landfilling of SCGs. However, further studies province, south africa. Water 12(10):2746
are required to evaluate the impact of the activation of SCG 5. Mahmoud, M.E., A.M. El-Ghanam, and S.R. Saad, Sequential
removal of chromium (VI) and prednisolone by nanobiochar-
biochar on the silver adsorption performance with co-existing enriched-diamine derivative. Biomass Conversion and Biorefin-
pollutants. A pilot scale using a continuous adsorption column ery, 2022: p. 1–20.
is required to better understand the capacity to handle large 6. Mahmoud ME et al (2021) Doping starch-gelatin mixed hydrogels
volumes of wastewater contaminated with silver species. with magnetic spinel ferrite@ biochar@ molybdenum oxide as
a highly efficient nanocomposite for removal of lead (II) ions. J
Supplementary Information The online version contains supplemen- Environ Chem Eng 9(6):106682
tary material available at https://d oi.o rg/1 0.1 007/s 13399-0 22-0 3491-0. 7. Yan N, Wang W-X (2021) Novel imaging of silver nanoparti-
cle uptake by a unicellular alga and trophic transfer to Daphnia
Author contribution Md Anwarul Islam and Elsa Antunes proposed magna. Environ Sci Technol 55(8):5143–5151
the research and designed the experiments. Md Anwarul Islam and 8. Mahmoud ME et al (2021) Novel immobilized fibrous natural cot-
Tewodros Kassa Dada conducted the experiments. Md Anwarul Islam, ton on Corchorus olitorius stalks biochar@ diethylenetriamine@
Tewodros Kassa Dada, Mst Irin Parvin, and Ravinder Kumar analysed feroxyhyte@ diethylenetriamine composite for coagulative
experimental results and reviewed and edited the manuscript. Elsa removal of silver quantum dots (Ag-QDs) from water. Cellulose
Antunes supervised, reviewed, and edited the manuscript. All authors 28(18):11397–11416
read the manuscript and approved it for publication. 9. Bhuyar P et al (2020) Synthesis of silver nanoparticles using marine
macroalgae Padina sp. and its antibacterial activity towards pathogenic
Funding Open Access funding enabled and organized by CAUL and bacteria. Beni-Suef Univ J Basic Appl Sci 9(1):1–15.
its Member Institutions 10. Bhuyar P et al (2021) Removal of nitrogen and phosphorus from
agro-industrial wastewater by using microalgae collected from
Data availability All data are available upon request. coastal region of peninsular Malaysia. Afr J Biol Sci 3(1):58–66
11. Mahmoud ME et al (2022) Adsorption behavior of silver quan-
tum dots by a novel super magnetic CoFe2O4-biochar-polymeric
Declarations nanocomposite. J Colloid Interface Sci 606:1597–1608
12. Zhang W et al (2019) Fate and toxicity of silver nanoparticles in
Ethics approval and consent to participate Not applicable. freshwater from laboratory to realistic environments: a review.
Environ Sci Pollut Res 26(8):7390–7404
13
Biomass Conversion and Biorefinery
13. Antunes E et al (2017) Silver removal from aqueous solution by 34. Jeon C (2017) Adsorption of silver ions from industrial wastewater
biochar produced from biosolids via microwave pyrolysis. J Envi- using waste coffee grounds. Korean J Chem Eng 34(2):384–391
ron Manage 203(Pt 1):264–272 35. Antunes E et al (2017) Biochar produced from biosolids using
14. Silva-Medeiros FV et al (2016) Kinetics and thermodynamics a single-mode microwave: characterisation and its potential for
studies of silver ions adsorption onto coconut shell activated car- phosphorus removal. J Environ Manage 196:119–126
bon. Environ Technol 37(24):3087–3093 36. Bardestani R, Patience GS, Kaliaguine S (2019) Experimental
15. Cantuaria ML et al (2016) Adsorption of silver from aqueous methods in chemical engineering: specific surface area and pore
solution onto pre-treated bentonite clay: complete batch system size distribution measurements—BET, BJH, and DFT. Can J
evaluation. J Clean Prod 112:1112–1121 Chem Eng 97(11):2781–2791
16. Zhang M, Helleur R, Zhang Y (2015) Ion-imprinted chitosan gel 37. Meng F et al (2019) The contribution of oxygen-containing
beads for selective adsorption of Ag+ from aqueous solutions. functional groups to the gas-phase adsorption of volatile organic
Carbohyd Polym 130:206–212 compounds with different polarities onto lignin-derived activated
17. Whangchai K et al (2021) Biomass generation and biodiesel pro- carbon fibers. Environ Sci Pollut Res 26(7):7195–7204
duction from macroalgae grown in the irrigation canal wastewater. 38. Romero Millán LM, Sierra Vargas FE, Nzihou A (2017) Kinetic
Water Sci Technol 84(10–11):2695–2702 analysis of tropical lignocellulosic agrowaste pyrolysis. BioEn-
18. Saman N et al (2015) Silver adsorption enhancement from aque- ergy Res 10(3):832–845.
ous and photographic waste solutions by mercerized coconut fiber. 39. Kabir G, Hameed B (2017) Recent progress on catalytic pyroly-
Sep Sci Technol 50(7):937–946 sis of lignocellulosic biomass to high-grade bio-oil and bio-
19. Chu C-Y et al (2021) High performance of biohydrogen produc- chemicals. Renew Sustain Energy Rev 70:945–967
tion in packed-filter bioreactor via optimizing packed-filter posi- 40. Kumar R et al (2019) Bio-oil upgrading with catalytic pyroly-
tion. Int J Environ Res Public Health 18(14):7462 sis of biomass using copper/zeolite-Nickel/zeolite and Copper-
20. Wajima T (2016) Synthesis of zeolitic material from green tuff Nickel/zeolite catalysts. Bioresour Technol 279:404–409
stone cake and its adsorption properties of silver (I) from aqueous 41. Campos-Vega R et al (2015) Spent coffee grounds: a review on
solution. Microporous Mesoporous Mater 233:154–162 current research and future prospects. Trends Food Sci Technol
21. Liu Y et al (2019) Understanding the high adsorption-reduction 45(1):24–36
performance of triethanolamine modified graphene oxide for sil- 42. Rodriguez, C. and G. Gordillo, Adiabatic gasification and pyrol-
ver ions. Colloids Surf, A 567:96–103 ysis of coffee husk using air-steam for partial oxidation. Journal
22. Mahmoud ME, Mohamed AK, Salam MA (2021) Self-decoration of Combustion, 2011.
of N-doped graphene oxide 3-D hydrogel onto magnetic shrimp 43. Milosavljevic I, Suuberg EM (1995) Cellulose thermal decom-
shell biochar for enhanced removal of hexavalent chromium. J position kinetics: global mass loss kinetics. Ind Eng Chem Res
Hazard Mater 408:124951 34(4):1081–1091
23. Mahmoud ME, Abou-Ali SA, Elweshahy SM (2021) Efficient and 44. Gonnella G et al (2022) Thermal analysis and kinetic modeling
ultrafast removal of Cd (II) and Sm (III) from water by leaves of of pyrolysis and oxidation of hydrochars. Energies 15(3):950
Cynara scolymus derived biochar. Mater Res Bull 141:111334 45. Rijo B et al (2021) Catalyzed pyrolysis of coffee and tea wastes.
24. Islam MA, Jacob MV, Antunes E (2021) A critical review on Energy 235:121252
silver nanoparticles: from synthesis and applications to its mitiga- 46. Mishra G, Kumar J, Bhaskar T (2015) Kinetic studies on the
tion through low-cost adsorption by biochar. J Environ Manage pyrolysis of pinewood. Biores Technol 182:282–288
281:111918 47. Lee XJ et al (2021) Solid biofuel production from spent cof-
25. Yao Y et al (2015) Engineered biochar from biofuel residue: fee ground wastes: process optimisation, characterisation and
characterization and its silver removal potential. ACS Appl Mater kinetic studies. Fuel 292:120309
Interfaces 7(19):10634–10640 48. Burnham AK, Zhou X, Broadbelt LJ (2015) Critical review of
26. Dong X, Ma LQ, Li Y (2011) Characteristics and mechanisms of the global chemical kinetics of cellulose thermal decomposition.
hexavalent chromium removal by biochar from sugar beet tailing. Energy Fuels 29(5):2906–2918
J Hazard Mater 190(1–3):909–915 49. He Q et al (2019) Effect of torrefaction on pinewood pyrolysis
27. Qiu B et al (2021) Biochar as a low-cost adsorbent for aqueous kinetics and thermal behavior using thermogravimetric analysis.
heavy metal removal: a review. J Anal Appl Pyrolysis 155. Biores Technol 280:104–111
28. Armstrong DL et al (2016) Temporal trends of perfluoroalkyl sub- 50. Yuan X et al (2017) Cattle manure pyrolysis process: kinetic and
stances in limed biosolids from a large municipal water resource thermodynamic analysis with isoconversional methods. Renew-
recovery facility. J Environ Manage 165:88–95 able Energy 107:489–496
29. Li X, Strezov V, Kan T (2014) Energy recovery potential analysis 51. Xu Y, Chen B (2013) Investigation of thermodynamic param-
of spent coffee grounds pyrolysis products. J Anal Appl Pyrol eters in the pyrolysis conversion of biomass and manure to
110:79–87 biochars using thermogravimetric analysis. Biores Technol
30. Tala W, Chantara S (2019) Use of spent coffee ground biochar as 146:485–493
ambient PAHs sorbent and novel extraction method for GC-MS 52. Pattanayak S et al (2021) Experimental investigation on pyrolysis
analysis. Environ Sci Pollut Res 26(13):13025–13040 kinetics, reaction mechanisms and thermodynamic parameters of
31. Zhang X et al (2020) Characterization and sulfonamide antibiotics biomass and tar in N2 atmosphere. Sustain Energy Technol Assess
adsorption capacity of spent coffee grounds based biochar and 48:101632
hydrochar. Sci Total Environ 716:137015–137015 53. Loy ACM et al (2018) Thermogravimetric kinetic modelling of
32. Jayakumar S et al (2021) Effects of light intensity and nutrients on in-situ catalytic pyrolytic conversion of rice husk to bioenergy
the lipid content of marine microalga (diatom) Amphiprora sp. for using rice hull ash catalyst. Biores Technol 261:213–222
promising biodiesel production. Sci Total Environ 768:145471 54. Fu J et al (2022) Torrefaction, temperature, and heating rate
33. Ma’arof NANB et al (2021) Exploitation of cost-effective renew- dependencies of pyrolysis of coffee grounds: its performances,
able heterogeneous base catalyst from banana (Musa paradisiaca) bio-oils, and emissions. Biores Technol 345:126346
peel for effective methyl ester production from soybean oil. Appl 55. Mallick D et al (2018) Discernment of synergism in pyrolysis of
Nanosci 2021:1–12. biomass blends using thermogravimetric analysis. Biores Technol
261:294–305
13
Biomass Conversion and Biorefinery
56. He J et al (2019) Effect of temperature on heavy metal(loid) 72. Naga Babu A et al (2021) Mathematical investigation into the
deportment during pyrolysis of Avicennia marina biomass sequential adsorption of silver ions and brilliant green dye using
obtained from phytoremediation. Bioresour Technol 278:214–222 biochar derived from Gracilaria Rhodophyta algae. Biomass Con-
57. Kan T, Strezov V, Evans TJ (2016) Lignocellulosic biomass vers Biorefinery 2021:1–20.
pyrolysis: a review of product properties and effects of pyrolysis 73. Peng H et al (2021) Reduction of silver ions to silver nanoparticles
parameters. Renew Sustain Energy Rev 57:1126–1140 by biomass and biochar: mechanisms and critical factors. Sci Total
58. Ktori R, Kamaterou P, Zabaniotou A (2018) Spent coffee grounds Environ 779:146326
valorization through pyrolysis for energy and materials production 74. Hussain N et al (2020) Cadmium (II) removal from aqueous
in the concept of circular economy. Materials Today: Proceedings solution using magnetic spent coffee ground biochar: kinetics,
5(14):27582–27588 isotherm and thermodynamic adsorption. Mater Res Express
59. Lee Y et al (2013) Comparison of biochar properties from biomass 7(8):085503
residues produced by slow pyrolysis at 500 C. Biores Technol 75. Zhang J et al (2019) Sludge-based biochar activation to enhance
148:196–201 Pb(II) adsorption. Fuel 252:101–108
60. Shin J et al (2020) Effects of physicochemical properties of bio- 76. Kılıç M et al (2013) Adsorption of heavy metal ions from aqueous
char derived from spent coffee grounds and commercial activated solutions by bio-char, a by-product of pyrolysis. Appl Surf Sci
carbon on adsorption behavior and mechanisms of strontium ions 283:856–862
(Sr2+). Environ Sci Pollut Res Int 77. Hashem A et al (2020) Non-linear adsorption characteristics of
61. Chwastowski J, Bradło D, Żukowski W (2020) Adsorption of cad- modified pine wood sawdust optimised for adsorption of Cd (II)
mium, manganese and lead ions from aqueous solutions using from aqueous systems. J Environ Chem Eng 8(4):103966
spent coffee grounds and biochar produced by its pyrolysis in the 78. Zhou Y et al (2014) Biochar-supported zerovalent iron reclaims
fluidized bed reactor. Materials 13(12):2782 silver from aqueous solution to form antimicrobial nanocompos-
62. Sing K (2001) The use of nitrogen adsorption for the characterisa- ite. Chemosphere 117:801–805
tion of porous materials. Colloids Surf, A 187:3–9 79. Kołodyńska D et al (2012) Kinetic and adsorptive characterization
63. Ambroz F et al (2018) Evaluation of the BET theory for the char- of biochar in metal ions removal. Chem Eng J 197:295–305
acterization of meso and microporous MOFs. Small Methods 80. Anandalakshmi K, Venugobal J, Ramasamy V (2016) Charac-
2(11):1800173 terization of silver nanoparticles by green synthesis method using
64. Zhao B et al (2018) Effect of pyrolysis temperature, heating rate, Pedalium murex leaf extract and their antibacterial activity. Appl
and residence time on rapeseed stem derived biochar. J Clean Prod Nanosci 6(3):399–408
174:977–987 81. Khan NA et al (2006) Alumina supported model Pd–Ag cata-
65. Cha JS et al (2016) Production and utilization of biochar: a review. lysts: a combined STM, XPS TPD and IRAS study. Surface Sci
J Ind Eng Chem 40:1–15 600(9):1849–1853
66. Jagdale P et al (2019) Waste coffee ground biochar: a material for 82. Ren M et al (2018) Titanium phosphate nanoplates modified with
humidity sensors. Sensors (Basel, Switzerland) 19(4):801 AgBr@Ag nanoparticles: a novel heterostructured photocatalyst
67. Gicheva G, Yordanov G (2013) Removal of citrate-coated silver with significantly enhanced visible light responsive activity. Front
nanoparticles from aqueous dispersions by using activated carbon. Chem 6:489
Colloids Surf, A 431:51–59 83. Ambaye TG et al (2020) Mechanisms and adsorption capaci-
68. Zhang L et al (2020) Preparation of biochar by mango peel and ties of biochar for the removal of organic and inorganic pol-
its adsorption characteristics of Cd (II) in solution. RSC Adv lutants from industrial wastewater. Int J Environ Sci Technol
10(59):35878–35888 18(10):3273–3294
69. Shaikh WA et al (2022) Biochar-based nanocomposite from waste 84. Komkiene J, Baltrenaite E (2015) Biochar as adsorbent for
tea leaf for toxic dye removal: from facile fabrication to functional removal of heavy metal ions [cadmium(II), copper(II), lead(II),
fitness. Chemosphere 291:132788 zinc(II)] from aqueous phase. Int J Environ Sci Technol
70. Zhao T et al (2018) Facile low-temperature one-step synthesis 13(2):471–482
of pomelo peel biochar under air atmosphere and its adsorption
behaviors for Ag (I) and Pb (II). Sci Total Environ 640:73–79 Publisher's note Springer Nature remains neutral with regard to
71. Tomczyk A, Sokołowska Z, Boguta P (2020) Biomass type effect jurisdictional claims in published maps and institutional affiliations.
on biochar surface characteristic and adsorption capacity relative
to silver and copper. Fuel (Guildford) 278:118168
13