Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Cleaner Production 324 (2021) 129270

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Production of biochar from lignocellulosic biomass with acidic deep


eutectic solvent and its application as efficient adsorbent for Cr (VI)
Yali Zhang a, b, Yahui Meng a, Li Ma a, b, Hairui Ji a, b, Xianqin Lu a, Zhiqiang Pang a,
Cuihua Dong a, b, *
a
State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology (Shandong Academy of Sciences), Jinan, 250353, China
b
School of Light Industry and Engineering, Qilu University of Technology (Shandong Academy of Sciences), Jinan, 250353, China

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

Handling editor. M.T. Moreira Deep eutectic solvents (DES) as potential green solvents have gained tremendous attention in biorefinery. Herein,
we presented a simple one-pot method for biochar production in acidic DES (p-Toluenesulfonic acid
Keywords: monohydrate-choline chloride) at lower temperature (140 ◦ C) using lignocellulose as feedstock. The results
Deep eutectic solvents showed that the residual solid biomass treated in acidic DES were featured as biochar with excellent properties.
Biochar
The biochar exhibited the high content of carbon, plentiful oxygen-containing functional groups and porous
Cr (VI)
structure, which endowed biochar with a superior adsorption capability on Cr (VI) up to 270.3 mg/g at 30 ◦ C.
Adsorption
Recyclability The pseudo-second order model and Langmuir model can be used to perfectly describe the adsorption process,
and adsorption behavior was assigned to chemical adsorption. In addition, the prepared biochar could be readily
recycled via alkali-treated desorption. This study offered an environmental-friendly and cost-effective method for
lignocellulose-derived biochar production.

1. Introduction consumption, peculiar heating rates and rigorous design of reaction


vessel (Chen et al., 2019), inferior yield and less surface functional
Lignocellulosic biomass has become the main raw material for groups. To address those issues, many studies were committed to explore
massive fuels, chemicals and functional product for its advantages of novel methods for efficient, cheap and environmental-friendly biochar
sustainability, renewability, bio-degradability and availability (Yu et al., preparation.
2019). Among those products, biochar as a facile-preparation, low-cost, Essentially, suitable solvent is the vital factor to fabricate a system
and environmentally friendly carbonaceous material could be produced for biochar production. Recently, Lu groups have reported that biochar
by thermochemical conversion of biomass under anoxic conditions could be synthesized by hydrolysis crop residues in acidic concentrated
(Lehmann, 2009). Typically, biochar was wealthy in specific surface lithium bromide at suitable conditions (Lu et al., 2019). The biochar
area, steady pore structure and surface functional groups (carboxylic, obtained by this approach exhibited excellent adsorbent ability to heavy
carbonylic and aldehydic, etc.) (Ma et al., 2019). Noteworthy, biochar metal Cr (VI). Therefore, we expected that production of biochar with
can be drove to ameliorate soil fertility, alleviate environmental pollu­ excellent properties could be potentially achieved at lower temperature
tion and lessen greenhouse gas emissions. by choosing appropriate solvent. As we known, DES was an ideal
The prevailing techniques for manufacture of biochar involved candidate, with remarkable advantages, such as extensive applications
gasification, hydrothermal carbonization, pyrolysis, and flash carbon­ in lignin removal (Thulluri et al., 2021), modified reverse osmosis
ization of carbon feedstock at inert environments (Pecha et al., 2019). membrane (Shahabi et al., 2020), ultrasound-assisted liquid phase
Comparatively, pyrolysis is the primary method for biochar production, microextraction separation (Menghwar et al., 2018), and efficient pre­
and many biomass species could be favorable as feedstock. High tem­ treatment of lignocellulosic biomass (Ling et al., 2021). Deep eutectic
peratures (600–900 ◦ C) and pressures was the key guarantee for pyrol­ solvents (DES) have been explored at the beginning of this century as
ysis of the biomass. Those severe conditions induced to high energy ionic liquid inspired green solvents and gained rapid attention for its

* Corresponding author. State Key Laboratory of Biobased Material and Green Papermaking, Qilu University of Technology (Shandong Academy of Sciences),
Jinan, 250353, China.
E-mail address: xiaodong771111@163.com (C. Dong).

https://doi.org/10.1016/j.jclepro.2021.129270
Received 21 April 2021; Received in revised form 14 September 2021; Accepted 3 October 2021
Available online 4 October 2021
0959-6526/© 2021 Elsevier Ltd. All rights reserved.
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

availability, nontoxicity, biodegradability, designability and recycla­ pressure-resistant flask, which was put into an oil bath at 140 ◦ C under
bility (Paiva et al., 2014; Menghwar et al., 2018; Ling et al., 2021). magnetic stirring (200 rpm). After 1 h of treatment, the reaction slurry
Typically, DES is a complexation formed by admixing a hydrogen-bond was cooled down to room temperature, and then filtration of the solu­
donor (HBD) with a hydrogen-bond acceptor (HBA) at appropriate ratio, tion was performed by ash-free filter paper on a Buchner funnel. The
and the resultant DES exhibited a melting point below that of its indi­ trapped black solid biochar was washed to neutral pH then freeze-dried.
vidual component (Garcia et al., 2015; Shahabi et al., 2020; Al-Risheq The biochar using poplar, pine and corn stalk as feedstock were desig­
et al., 2021). And one of the most popular salt acted as HBA in DES is nated as poplar biochar (PB), pine biochar (CB) and corn stalk biochar
choline chloride for its availability and biodegradability. Contrasted to (SB), respectively.
traditional ionic liquids, choline chloride-derived DESs were endowed
many superior properties including low price, inertness with water, easy 2.4. Batch adsorption experiment
prepare, etc. (Dannie et al., 2017; Aydin et al., 2017; Thulluri et al.,
2021). Due to these remarkable advantages, DESs were able to be used in The aqueous Cr (VI) solution for adsorption experiment was obtained
biochar modification for enhanced properties. For example, DES (glyc­ by adding K2Cr2O7 reagent into distilled water. Additionally, the pH of
erol-hexadecyltrimethylammonium bromide) was used to modify the aqueous Cr (VI) solution was regulated with HCl (12.0 mol/L) or
clay-biochar hybrid adsorbent, which showed high adsorption capa­ NaOH (40.0 w/v%) solution prior to use. The adsorption experiments
bility on organic pollutants (Lawal et al., 2019). In addition, fiber sep­ were implemented in 50 mL flasks on a thermostatic oscillator with a
aration and the simultaneous nitrogen doping could be obtained in speed of 200 rpm and a set temperature. After adding biochar (50 mg)
ChCl-urea DES. This DES modification improved the catalytic perfor­ into the flask containing of the Cr (VI) solution (25 mL), the Cr (VI)
mance of biochar, and an environmentally friendly high-performance adsorption was initiated. The biochar and spent liquor were separated
biochar could be prepared (Shujing et al., 2021). with a sand core funnel after each batch run. The concentration of re­
Moreover, considering principal of carbonization to form biochar, sidual Cr (VI) was judged with an ultraviolet visible spectrophotometer.
acid condition was more beneficial to promote dehydration of biomass. The quantity of Cr (VI) loaded on the biochar at adsorption equilibrium
Hence, the higher acidity value of DES was a key factor in the design of (qe) and removal efficiency (η) were counted on basis of the following
DES once the target product was settled as biochar. Generally, the equations (Eq. (1), Eq. (2)).
chemical nature of the HBDs has a vigorous effect on the acid or basic
strength of the forming DES (Zhang et al., 2012). As we known, sulfonic qe =
(C0 − Ce ) × V
(1)
acid has strong acidity as an organic acid. Therefore, we reported a DES m
combined by choline chloride and p-toluenesulfonic acid monohydrate
C0 − Ce
toward prepare biochar using lignocellulosic biomass in view of its η= × 100% (2)
C0
outstanding acidity in this study (Wang et al., 2020). The study showed
that the biochar could be prepared by one-pot method at low tempera­ Where qe is the amount of Cr (VI) adsorbed on biochar at equilibrium
ture and used as an efficient adsorbent for Cr (VI) removal. The time; C0 and Ce represent the initial and equilibrium concentrations of Cr
adsorption mechanism of biochar on Cr (VI) was discussed via exploring (VI) in the solution, respectively; V is the volume of the Cr (VI) solution;
the effect of contact time and initial pH on the process of adsorption and m denotes the dosage of biochar added into the solution. Where η is
transformation. The adsorption process was studied by fitting adsorp­ removal efficiency.
tion isotherm model, adsorption kinetic model and adsorption thermo­ The experiment results were analyzed using Langmuir (Eq. (3)),
dynamic model. The physicochemical properties of biochar were Freundlich (Eq. (4)) and Temkin models (Eq. (5)) to evaluate the Cr (VI)
characterized and its adsorption ability to heavy metal were investi­ adsorption behavior.
gated, which confirmed the adsorption mechanism of Cr (VI). Further­
more, this study is expected that an environmentally friendly strategy qe =
qm × KL × Ce
(3)
was proposed for fabricating biochar materials and applications in 1 + KL × Ce
heavy metal removal. 1

(4)
/
qe = KF × Ce n
2. Experimental
RTln(αTem × Ce )
qe = (5)
2.1. Materials βTem

Hybrid poplar (P) and caribbean pine (C) chips were gotten from a Where qm (mg/g) is the maximum adsorption capacity, KL (L/mg) is
pulping mill in Shandong Province, China. Corn stalk (S) was gathered indicator of adsorption capacity for Langmuir model. Where n is
from a farm in Hebei Province. The three kinds of biomass were ground empirical index, KF is indicator of adsorption capacity for Freundlich
into a size of 20–40 mesh using a Wiley mill (IKA MF10 Basic). Choline model. Where R, αTem and βTem are indicators of adsorption capacity for
chloride (ChCl, AR, 98.0%) and p-Toluenesulfonic acid monohydrate (p- and Temkin model, T (K) is the absolute temperature.
TsOH, AR, 99.0%) were provided by Shanghai Macklin Biochemical Co., The effect of contact time on Cr (VI) removal was carried out by
Ltd. Potassium dichromate (K2Cr2O7, AR, 99.9%) were obtained from pseudo-first order model (Eq. (6)) and pseudo-second order model (Eq.
Laiyang Economic Development Zone Chemical Plant. (7)). The diffusion mechanism was simulated by the intra-particle
diffusion model (Eq. (8)).
2.2. Preparation of deep eutectic solvents K1 × t
log(qe − qt ) = log qe − (6)
2.303
DES was prepared by simply mixing p-TsOH and ChCl (molar ratio
1:1) under magnetic stirring (200 rpm) at 70 ◦ C. The stirring was t 1 1
= + t (7)
stopped, once a homogeneous colorless liquid (DES) happened. The qt K2 q2e qe
liquid was cooled to room temperature and stored in a desiccator.
(8)
1
/
qt = Kid × t 2 + C
2.3. Preparation of biochar
Where qt (mg/g) and qe (mg/g) are the adsorption amounts at time t (h)
A mixture of 5 g biomass and 50 g DES were added into a 500 mL and equilibrium state, respectively. K1 (1/h) represents the rate constant

2
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

for the pseudo-first order adsorption kinetics. Where K2 (g/mg/h) is the 3. Results and discussion
rate constant for pseudo-second order adsorption kinetics. Where Kid
(mg/g/min1/2) is the intra-particle diffusion rate constant. C is a con­ 3.1. Properties of biochar from acidic deep eutectic solvents treatment
stant related to the thickness of the boundary layer.
The thermodynamics parameters of the adsorption processes were The element compositions of solid residue after treatment in DES at
calculated with the following equations (Eq. (9), Eq. (10), Eq. (11), Eq. different temperatures for residence time of 1 h were determined
(12)). (Table 1). In respect of temperature less than 110 ◦ C, the predominant
role of acidic DES treatment was extraction, hydrolysis and dissolution
ΔG = − R⋅T⋅lnKC (9)
of the lignocellulosic components (Kim et al., 2018), and morphology of
qe the original biomass was roughly conserved without formation of bio­
Kc = (10) char. Therefore, production of biochar requires higher temperature in
Ce
DES. As temperature went up to 110 ◦ C, the treated solid residue
ΔG = ΔH − T⋅ΔS (11) (Fig. S1.) possesses the morphology of biochar and the carbon content
was similar to the products using pyrolysis process (Lu et al., 2019).
lnKC = −
ΔH ΔS
+ (12) Specifically, carbon content of biochar enhanced obviously with
R⋅T R increasing processing temperature, which implied high temperature
facilities dehydration and carbonization reactions. Meanwhile, oxygen
Where R (J/(mol⋅K)) is the gas constant. T (K) is the absolute tempera­
content of biochar dropped gradually. Comparatively, the compositional
ture, and KC (L/g) is the adsorption equilibrium constant. Where qe is the
results of biochar varied slightly with biomass species, and this acidic
amount of Cr (VI) adsorbed on biochar at equilibrium time; Ce is equi­
DES treatment has excellent flexibility on feedstock. Compared with
librium concentrations of Cr (VI) in the solution. All experiments were
biochar by pyrolysis (Chen et al., 2019), the biochar in this study was
conducted in triple.
endowed with high oxygen content, which possibly ascribed to more
oxygen-rich functional groups in the final products.
2.5. Biochar characterization
The specific surface area together with the pore structure of the
biochar were judged using the N2 adsorption-desorption isotherm. The
Elemental content in biochar was analyzed using an elemental
BET (Fig. S2.) surface area of biochars produced from PB, CB and SB
analyzer (Unicube, Elementar, Germany). All samples were tested three
were 7.1 m2/g, 6.9 m2/g and 7.1 m2/g respectively. The biomass species
times and the mean values were calculated. BET surface areas of the
have negligible effect on surface area of biochars using the DES treat­
biochar prepared at different conditions were measured by N2 adsorp­
ment. Moreover, the surface area of this DES-derived biochar was rela­
tion isotherm using a nano-porous system (ASAP 2460, Micromeritics,
tively low as an adsorbent, and the pore structure of biochar has a small
USA). Scanning electron microscopy (Regulus 8220, Hitachi, Japan)
contribution to the adsorption capability. The enhancement on the
images were gotten by scanned sample at accelerating voltage of 5 kV.
properties of this biochar is still possible by forming more micro-pore
Gold sputtering of the samples was performed, then the samples were
structure, which will be elucidated in our following study. The
observed and pictures were captured. The X-ray diffraction (XRD)
morphology of biochar was watched by SEM (Fig. S3.). In agreement
spectra of the biochars were acquired by an X-ray diffractometer
with the specific surface area result, the morphology of biochars derived
(Rigaku D/Max 2400, Bruker, Germany). FTIR spectra with wave­
from different species negligibly changed. Some biochar fragments at
number ranged from 500 to 4000 cm− 1 were scanned at a resolution of 4
varying size could be detected on the surface. A small number of pores
cm− 1 in a fourier transform-infrared spectrometer (Alpha-T, Bruker,
and interstices were formed on the surface, which accounts for its low
Germany). Surface functional groups were determined through X-ray
specific surface area.
photoelectron spectroscopy (XPS; ESCALAB 250Xi). The determination
XRD pattern of the biochar were also detected (Fig. S4.). Peaks
of Cr (VI) was primarily based on colorimetry combined with UV–Vis
located at around 2θ of 25◦ and 45◦ are attributed to the (002) and (100)
spectrophotometry (GENESYS 10s UV–Vis Spectrophotometer) at 540
plane, respectively. This is characteristic diffraction pattern of porous
nm.
carbon materials (Sun et al., 2016). In our study, only one broad char­
acteristic peak at around 2θ of 25◦ (002) was detected, indicating that
2.6. Desorption experiments
the structure of the biochar is essentially in disordered state. The almost
disappearance of peak (100) in the biochar suggested that the defectives
Cr (VI)-loaded biochar was dispersed in 1 M and 2 M NaOH solution
periodically stacked (Sun et al., 2016).
at 37 ◦ C with a stirring speed of 200 rpm for 24 h. After separation, the
FTIR spectra of PB, CB and SB were measured (Fig. S5.). The broad
collected biochar was washed with distilled water to neutral pH, and
band at 3372 cm− 1 was designated to O–H stretching vibration deriving
then dried at 105 ◦ C for 8 h. The adsorption-desorption process was
from carboxyl, alcohol and phenol structure (Aboulkas et al., 2017). The
circulated four recycles. In order to provide a theoretical basis for
absorption peak at 2935 cm− 1 and 1452 cm− 1 belonged to stretching
practical engineering, this study adopted 300 mg/L chromium con­
vibration of C–H (Xiang et al., 2019). The peaks at about 1695 cm− 1
taining solution as the pollution model, which is very rare in the pre­
represented C– – O stretching vibrations attributed by to ketones, car­
vious studies.
boxylic acidic and aldehydes groups (Shao et al., 2018). The wide band
at 1119 cm− 1 was C–O stretching vibrations (Zou et al., 2009). The FTIR
analysis suggested that the biochar prepared in DES possess abundant

Table 1
Elemental analysis of biochar produced at different conditions.
Sample Hybrid poplar Caribbean pine Corn stalk

110 ◦ C 140 ◦ C 170 ◦ C 110 ◦ C 140 ◦ C 170 ◦ C 110 ◦ C 140 ◦ C 170 ◦ C

C 48.38 61.11 63.28 51.36 62.06 64.25 49.85 58.99 63.72


H 5.12 4.78 4.53 4.92 4.57 4.38 6.14 5.23 4.94
O 42.46 29.10 28.35 40.85 28.97 27.01 38.87 30.28 27.96
N 0.69 0.33 0.25 0.63 0.38 0.29 1.12 0.89 0.73

3
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

functional groups, such as hydroxyl and carbonyl groups. These groups exploited as adsorbent. To further understand the adsorption nature of
decorated biochar with mighty ability to adsorb heavy metals. our biochar product, we carried out a comparison with biochar prepared
The surface analysis of P, PB was conducted using XPS (Fig. 1a). by other methods (Table 3). According to the data listed in Table 3, we
Typical C1s XPS spectrum of the samples can be divided into three peaks can conclude that the biochar in this paper had superior adsorption
(Fig. 1b). The fitting results of C1s of P were as follows: C–C (284.8 eV), properties to Cr (VI), which was based on the maximum adsorption
C–O (286.0 eV) and C–N (287.5 eV) (Shao et al., 2018; Dong-Wan et al., capability from biochar prepared with different method. Hence, the
2018; Da et al., 2019). The peaks of PB represented C–C, C–O, and method we provided is a promising route for the industrial scale pro­
–COOH (289.0 eV) (Da et al., 2019). After DES treatment, C–C content of duction of biochar in the future.
the biochar increased while C– – O and C–N content decreased signifi­
cantly. Therefore, the biochar had high carbon content. These results 3.3. Effect of pH on the adsorption capability
indicated that the C–O bonds could be converted into –COOH. The
surface oxygen of biochar was composed of various functional groups Usually, adsorption of metal ion was much controlled by pH of
(Fig. 1c). The percentage of C– – O (531.2 eV) and C–O (532.4 eV) in the environment. The effect of pH on adsorption ability of biochar to Cr (VI)
corresponding ketones and carboxylic acids increased greatly was estimated by changing pH range from 1 to 6. As displayed in Fig. 2A
(Table S1), and the percentage of C–OH (533.0 eV) decreased duo to (the concentration of Cr (VI) was 325.0 mg/L), the adsorption properties
oxidization of C–OH bond into C = O bond during DES treatment (Lu of biochar fluctuated widely with the pH of the solution. At low pH value
et al., 2021). of 1, all the biochar dated from poplar, pine, and corn stalk, could almost
completely remove the heavy metal Cr (VI). However, the color of so­
3.2. The adsorption properties of biochar to Cr (VI) lution appeared brown and was not transparent, which was induced by
the release of low molecular substance from biochar at the high acid
To check the adsorption capacity of the biochar, we chose biochar conditions. As the pH was boosted to 2, biochar prepared from different
prepared at 140 ◦ C as adsorbent for Cr (VI) removal. As shown in biomass also could remove Cr (VI) thoroughly and the solution became
Table 2, the biochar showed wonderful adsorption capacity for Cr (VI) clear. Meanwhile, the adsorption capability decreased dramatically with
and the maximum adsorption capacity up to 270.3 mg/g. For all biochar increasing pH value and the remove rate reduced to only 22.3%–31.1%
from poplar, pine, and corn stalk, they could thoroughly get rid of the Cr at pH 6.
(VI) at the Cr (VI) concentration of 325.0 mg/L. Even at high concen­ In case of the concentration of Cr (VI) multiplied to 525.0 mg/L
tration of 525.0 mg/L, the removal efficiency was remarkable. Removal (Fig. 2B), the biochar still exhibited excellent adsorption performance,
rate of Cr (VI) was 98.2%, 96.8% and 96.4% for biochar from poplar, and the removal rates of Cr (VI) using biochar from poplar, pine, corn
pine, and corn stalk, respectively. Meanwhile, the biochar from poplar stalk, were 98.2%, 96.8% and 96.4%, separately. Effect of the solution
exhibited a little higher adsorption ability on Cr (VI) than those from the pH on adsorption ability behaved similarly as that of 325.0 mg/L Cr (VI)
other two feedstocks. solution. Therefore, the optimal pH of solution should be controlled
In terms of removing heavy metals is concerned, adsorption is widely around 2 for high adsorption efficiency.
applied for its low cost and renewability, and various biochar was The adsorption capacity of biochar on Cr (VI) significant increased as

Fig. 1. (a) XPS spectra of P, PB before and after Cr(VI) adsorption. (b) C1s XPS spectra of P and PB. (c)O1s XPS spectra of P and PB. (d) Cr2p XPS spectra of PB after
Cr(VI) adsorption.

4
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

Table 2
The concentrations (Ce) and removal efficiencies (Rf) of Cr (VI) with biochar.
C0 (mg/L) Hybrid poplar Caribbean pine Corn stalk

Ce Rf (%) Qmax Ce Rf (%) Qmax Ce Rf (%) Qmax

325 0 100 270.3 0 100 256.2 0 100 256.2


525 9.4 98.2 16.6 96.8 19.0 96.4

Note: the pH of Cr (VI) solution was 2; The C0 meant the initial concentrations of Cr (VI); The Ce(mg/l) meant the concentrations of Cr (VI) after adsorption; Qmax(mg/l)
mean the measured maximum adsorption capacity at 30 ◦ C.

KL is the constant of Langmuir isotherm, C0 (mg/L) is the original


Table 3 concentration of the Cr (VI). Based on the value of RL, adsorption ten­
Comparison of adsorption performance of various biochars.
dency could be revealed: irreversible (RL = 0), favorable (0 < RL < 1),
Biochar feedstock Adsorption Qmax (mg/ Reference linear (RL = 1), or unfavorable (RL > 1) (Pezoti et al., 2015). Further­
pH g)
more, the smaller is the RL, the higher affinity between the Cr (VI) and
Marine Chlorella 2 15.9 Amin and Chetpattananondh the biochar is (Webi and Chakravort, 1974). The value of RL was
sp. (2019) deduced to be 0.0004 by Equation (13), indicating that the adsorption
Melia azedarach 3 25.3 Zhang et al. (2018a)
Siderite 2 81.0 Bibi et al. (2018)
reaction was favorable and the affinity between Cr (VI) and biochar was
Corn residue 3 55.2 Lu et al. (2019) very strong. The high intensity of adsorption between adsorbate and
Tobacco petiole 7.4 99.4 Zhang et al. (2018b) metals can be effective restraint for desorption of metals from biochar
Sweet lime peel 2 100.0 Shakya et al. (2019) during the removal of heavy metals from effluent.
Peanut hull 2 142.9 Cai et al. (2019)
Corn straw 3 175.4 Ma et al. (2019)
Douglas fir 2 150 Amali Herath al. (2021) 3.5. Adsorption kinetics
Acidic vinegar 2 236.81 Kaili Ding al. (2021)
residue
Hybrid poplar 2 270.3 This paper The adsorption kinetics of the biochar to Cr (VI) were probed by
three different models, namely the pseudo-first order, the pseudo-second
Qmax mean maximum adsorption capacity.
order and intra-particle diffusion model. The linear fitted results were
displayed in Fig. 3 and Table 5 separately. The adsorption process of Cr
the lowering in pH value, which mainly attributed to that Cr2O72− and (VI) onto the biochar had highly fitted consistency with pseudo-second
CrO42− were converted into HCrO4− in acidic solution (Lu et al., 2021). order model. Moreover, the calculated adsorption capacity (145.4 mg/
Compared with Cr2O72− and CrO42− , the HCrO4− needs lower adsorp­ g) was nearly identical to the practical equilibrium adsorption capacity
tion energy and is more easily adsorbed by biochar. Moreover, proton­ (144.6 mg/g). Therefore, from the results of Langmuir and pseudo-
ation of the surface functional groups and the electrostatic attraction of second order model, the adsorption category of Cr (VI) onto the bio­
biochar to Cr (VI) were superior at low pH, which promoted the char was predominantly governed by chemisorption of single molecular
adsorption on Cr (VI). Thus, the proportion of HCrO4− decreased grad­ layer with oxygen-containing functional groups (-OH, –COOH, –C– – O).
ually with pH increased, thus resulted to the lower adsorption capacity Hydrogen bonding and π-π bonding may be the main possible in­
on Cr (VI). teractions between biochar and Cr (VI). The reduction of Cr (VI) by
XPS analysis was carried out for PB after Cr(VI) adsorption and the C–OH bonds leads to the formation of Cr (III) (Valentín-Reyes et al.,
spectra is illustrated in Fig. 1a and 1d. Fig. 1d shows the Cr2p spectrum 2019; Lu et al., 2021), which was consistent with results of the XPS
of PB-Cr, corresponding to the Cr2p3/2 and Cr2p1/2 orbitals. The analysis. As well as, microporous adsorption of Cr(VI) and electrostatic
binding energy values at 576.9 and 586.4 eV were assigned to Cr(III) attraction were also the factors contributing to the adsorption as shown
(32.5%), 578.2 and 587.5 eV to Cr(VI) (67.5%). It is concluded that the in Fig. 4.
adsorption process for Cr (VI) by biochar was partly through the
reduction of Cr(VI) to Cr(III).
3.6. Adsorption thermodynamic

3.4. Adsorption isotherm The nature and feasibility of adsorption process was tested by ther­
modynamics at different temperature and the results were shown in
Different isotherm model including Langmuir, Freundlich and Table 6. The adsorption capacity of biochar to Cr (VI) elevated from
Temkin model are employed to elucidate the adsorption type of the 270.4 mg/g to 281.2 mg/g as temperature changed from 303 K to 323 K.
prepared biochar on Cr (VI). Here, the biochar using SB as feedstock was Consequently, the adsorption process is proclaimed to be endothermic
chosen. The fitting results and correlation coefficients are separately as a result of high temperature favoring the adsorption. In addition,
listed in Fig. 3 and Table 4. The correlation coefficient calculated with adsorption process was spontaneous because of the negative ΔG values
Langmuir model (R2 = 0.99897) is fantastically higher than that of determined at all temperatures. The absolute value of ΔG enhancing
Freundlich model (R2 = 0.77990) and Temkin model (R2 = 0.77173). with temperature further verified that the adsorption readily proceeds at
Therefore, Langmuir model are suitable to the adsorption process. The high temperature.
adsorbed Cr (VI) was uniformly distributed and formed a monolayer on The endothermic adsorption of biochar on Cr (VI) can be further
the biochar surface. In addition, there were almost no difference be­ corroborated by the positive value of ΔH (28.5633 kJ/mol). As the ΔH
tween the calculated adsorption capacity (256.4 mg/g) from Langmuir value in the range of 2.1–20.9 kJ/mol, the nature of adsorption is
model and the measured maximum adsorption capacity (256.2 mg/g). ascribed to physical adsorption. As the ΔH value belong to the range of
Separation factor (RL) is an important parameter of the Langmuir 20.9–418.4 kJ/mol, the adsorption process was regarded as chemical
isotherm, which can be used to make sure if the adsorption process is adsorption (Sun et al., 2012). Therefore, the adsorption process of bio­
favorable. RL could be calculated by the following equation (Eq. (13)): char to Cr (VI) is dominated by chemical adsorption.
1 The positive value of ΔS revealed that the degree of dislocation at the
RL = (13) solid-liquid interface augmented and the adsorption process was feasible
1 + KL ⋅C0
(Borah et al., 2015).

5
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

Fig. 2. Effect of pH on the adsorption properties of biochar. A: the initial concentration of Cr (VI) is 325.0 mg/L; B: the initial concentration of Cr (VI) is 525.0 mg/L;
the dosage of biochar was 0.05 g and the temperature was 30 ◦ C; the volume of the solution was 25 mL.

3.7. Evaluation recyclability of biochar and deep eutectic solvents with fresh DES and DES recovered in the 1st, 2nd and 3rd cycle was
88.81±1.24%, 84.48±1.48%, 83.29±1.28% and 75.94±2.41% respec­
Evaluation of biochar reusability is of vital importance for devel­ tively. The decrease in adsorption capacity with recycled DES possibly
oping a cost-effective technique for biochar production, and the per­ attributed to the impurities in the recovered DES inhibited carbonization
formance of the recycled biochar via adsorption-desorption cycles was reaction during DES treatment. Notably, the removal efficiency of bio­
investigated (Fig. 5). 2 M and 1 M NaOH solution were used in the char prepared with recovered DES in the 3rd cycle still reached
regeneration experiment. The removal efficiency of biochar to Cr (VI) 75.94±2.41%, which manifested the biochar prepared by the recovered
reserved well enough (93.67% and 89.56%) after undergoing the first DES still had excellent adsorption capacity on Cr (VI).
adsorption-desorption cycle. After two cycles, the regeneration per­
centage of biochar was 83.04% and 60.41%, respectively. The removal 4. Conclusions
efficiency of recycled biochar still able to reach 59.42% and 40.01%
after the fourth cycle. These above results indicated that more concen­ In summary, a facile method for biomass-derived biochar prepara­
tration NaOH solution was better for improving the regeneration of tion using acidic DES (p-TsOH-ChCl) as solvent was fabricated. The
biochar. Conceivably, the reduced adsorption ability on Cr (VI) are due biochar could be prepared from hybrid poplar, caribbean pine and corn
to loss of microspores and surface-active adsorption sites. However, the stalk in DES with relatively low temperature (140 ◦ C), and featured high
biochar could still hold high performance on Cr (VI) removal after four content of carbon and oxygen-containing functional groups. The
adsorption-desorption cycles. adsorption capacity of the biochar was affected by adsorption temper­
The p-TsOH-ChCl DES was recovered and reused and its recyclability ature and feedstock origins. The adsorption process of biochar on Cr (VI)
was investigated. The regeneration process of DES was as following: the could be fitted into pseudo-second order model and Langmuir model,
solid residue after DES treatment was thoroughly washed with water, and the adsorption behavior was assigned to chemical adsorption of
and the collected mixed solution (aqueous DES solution) was rotary single molecular layer. In addition, both the DES effluent and biochar
evaporated, then reused in further three successive pretreatments. The could be easily recycled. This biochar preparation using acidic DES
removal efficiency of Cr(VI) (694.12 ppm) with biochar that prepared method provides a novel green pathway toward production of adsorbent

6
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

Fig. 3. Adsorption isotherm and Kinetic models for adsorption of Cr (VI) on biochar. (a) Langmuir model; (b) Freundlich model; (c) Temkin model. (d) pseudo-first
order model; (e) pseudo-second order model; (f) intra-particle diffusion model.

Table 4
Parameters and determination coefficients of the isotherm model.
Langmuir Freundlich Temkin

qm, cal (mg/g) KL (L/mg) R2 1/n KF R2 RT/βTem αTem R2

256.4 4.1 0.99900 0.0939 178.3 0.77990 19.2 12854.2 0.77170

7
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

Table 5
Parameters and determination coefficients of the kinetic models for Cr (VI) adsorption.
Pseudo-first order model Pseudo-second order model Intra-particle diffusion model
2 2
qe, cal (mg/g) K1 (1/h) R qe, cal (mg/g) K2 (1/h) R Kid1 (mg/g/min1/2) R12 Kid2 (mg/g/min1/2) R22

141.8 1.2 0.94850 145.4 0.1 0.99980 76.1 0.94360 1.2 0.91270

Fig. 4. Schematic mechanism diagram of Cr (VI) removal by biochar.

Fig. 5. Recyclability of biochar as an adsorbent for Cr (VI) removal. (Adsorp­


tion conditions: the initial Cr (VI) concentration of 300 mg/L, 37 ◦ C, pH of 2,
Table 6 biochar dosage of 100 mg and of Cr (VI) solution of 50 mL).
Thermodynamic parameters for adsorption of Cr (VI) on biochar.
T (K) ΔG (KJ/ ΔH (KJ/ ΔS (KJ/ R2 Qm (mg/g) Acknowledgements
mol) mol) (mol⋅K))

303 − 4.4940 28.5633 0.1091 0.99990 270.3724 The authors are grateful for the support of the Project Supported by
313 − 5.5850 277.2115 the International Cooperation Funding of Qilu University of Technology
323 − 6.6759 281.2010
(QLUTGJHZ2018030; QLUTGJHZ2018027).
Note: The initial concentration of Cr (VI) was 586.2 mg/L. Qm mean maximum
adsorption capacity at different temperatures. Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.


material. org/10.1016/j.jclepro.2021.129270.

CRediT authorship contribution statement References

Aboulkas, A., Hammani, H., El Achaby, M., Bilal, E., Barakat, A., El Harfi, K., 2017.
Yali Zhang: Methodology, Investigation, Writing – original draft, Valorization of algal waste via pyrolysis in a fixed-bed reactor: production and
preparation, Formal analysis, Data curation, Data Curing. Yahui Meng: characterization of bio-oil and bio-char. Bioresour. Technol. 243, 400–408. https://
Methodology, Investigation. Li Ma: Investigation. Hairui Ji: Investiga­ doi.org/10.1016/j.biortech.2017.06.098.
Al-Risheq, D., Nasser, M., Qiblawey, H., Hussein, I., Al-Ghouti, M., 2021. Influence of
tion, Visualization. Xianqin Lu: Data curation, Methodology. Zhiqiang
choline chloride based natural deep eutectic solvent on the separation and
Pang: Writing – review & editing, Formal analysis. Cuihua Dong: rheological behavior of stable bentonite suspension. Separ. Purif. Technol. 270,
Conceptualization, Methodology, Resources, Supervision, Formal anal­ 118799. https://doi.org/10.1016/j.seppur.2021.118799.
ysis, Project administration, Writing – review & editing. Amali, Herath, Reid, Claudia, Perez, Felio, Pittman, Charles U., Todd, E., Mlsna, 2021.
Biochar supported polyaniline hybrid for aqueous chromium and nitrate adsorption.
J. Environ. Manag. 296, 113186. https://doi.org/10.1016/j.jenvman.2021.113186.
Declaration of competing interest Amin, M., Chetpattananondh, P., 2019. Biochar from extracted marine Chlorella sp.
residue for high efficiency adsorption with ultrasonication to remove Cr(VI), Zn(II)
and Ni(II). Bioresour. Technol. 289, 121578. https://doi.org/10.1016/j.
The authors declare that they have no known competing financial biortech.2019.121578.
interests or personal relationships that could have appeared to influence Aydin, F., Yilmaz, E., Soylak, M., 2017. A simple and novel deep eutectic solvent based
the work reported in this paper. ultrasound-assisted emulsification liquid phase microextraction method for
malachite green in farmed and ornamental aquarium fish water samples.
Microchem. J. 132, 280–285. https://doi.org/10.1016/j.microc.2017.02.014.

8
Y. Zhang et al. Journal of Cleaner Production 324 (2021) 129270

Bibi, I., Niazi, N.K., Choppala, G., Burton, E.D., 2018. Chromium(VI) removal by siderite Pecha, M.B., Arbelaez, J.I.M., Garcia-Perez, M., Chejne, F., Ciesielski, P.N., 2019.
(FeCO3) in anoxic aqueous solutions: an X-ray absorption spectroscopy investigation. Progress in understanding the four dominant intra-particle phenomena of
Sci. Total Environ. 640–641, 1424–1431. https://doi.org/10.1016/j. lignocellulose pyrolysis: chemical reactions, heat transfer, mass transfer, and phase
scitotenv.2018.06.003. change. Green Chem. 21 (11), 2868–2898. https://doi.org/10.1039/C9GC00585D.
Borah, L., Goswami, M., Phukan, P., 2015. Adsorption of methylene blue and eosin Pezoti, O., Cazetta, A.L., Bedin, K.C., Yamazaki, D.A.S., Bandoch, G.F.G., Asefa, T.,
yellow using porous carbon prepared from tea waste: adsorption equilibrium, Visentainer, J.V., Almeida], V.C., 2015. Removal of tetracycline by NaOH-activated
kinetics and thermodynamics study. J. Environ. Chem. Eng. 3 (2), 1018–1028. carbon produced from macadamia nut shells: kinetic and equilibrium studies. Chem.
https://doi.org/10.1016/j.jece.2015.02.013. Eng. J. 260, 291–299. https://doi.org/10.1016/j.cej.2014.09.017.
Cai, W., Wei, J., Li, Z., Liu, Y., Zhou, J., Han, B., 2019. Preparation of amino- Shahabi, S.S., Azizi, N., Vatanpour, V., 2020. Tuning thin-film composite reverse osmosis
functionalized magnetic biochar with excellent adsorption performance for Cr(VI) by membranes using deep eutectic solvents and ionic liquids toward enhanced water
a mild one-step hydrothermal method from peanut hull. Colloids Surf., A 563, permeation. J. Membr. Sci. 610, 118267. https://doi.org/10.1016/j.
102–111. https://doi.org/10.1016/j.colsurfa.2018.11.062. memsci.2020.118267.
Chen, W., Meng, J., Han, X., Lan, Y., Zhang, W., 2019. Past, present, and future of Shakya, A., Núñez-Delgado, A., Agarwal, T., 2019. Biochar synthesis from sweet lime
biochar. Biochar 1, 75–87. https://doi.org/10.1007/s42773-019-00008-3. peel for hexavalent chromium remediation from aqueous solution. J. Environ.
Da, Ouyang, Chen, Yun, Yan, Jingchun, Qian, Linbo, Han, Lu, Chen, Mengfang, 2019. Manag. 251, 109570. https://doi.org/10.1016/j.jenvman.2019.109570.
Activation mechanism of peroxymonosulfate by biochar for catalytic degradation of Shao, J., Zhang, J., Zhang, X., Feng, Y., Zhang, H., Zhang, S., Chen, H., 2018. Enhance
1,4-dioxane: important role of biochar defect structures. Chem. Eng. J. 370, SO2 adsorption performance of biochar modified by CO2 activation and amine
614–624. https://doi.org/10.1016/j.cej.2019.03.235. impregnation. Fuel 224, 138–146. https://doi.org/10.1016/j.fuel.2018.03.064.
Dannie, J.G.P., Osch, V., Laura, J.B.M., Kollau, Adriaan, Bruinhorst, V.D., Shujing, Ye, Xiong, Weiping, Liang, Jie, Yang, Hailan, Wu, Haipeng, Zhou, Chengyun,
Sari, Asikainen, Marisa, A.A., Rocha, 2017. Ionic liquids and deep eutectic solvents Du, Li, Guo, Jiayin, Wang, Wenjun, Xiang, Ling, Zeng, Guangming, Tan, Xiaofei,
for lignocellulosic biomass fractionation. Phys. Chem. Chem. Phys. 19 (4), 2021. Refined regulation and nitrogen doping of biochar derived from ramie fiber by
2636–2665. https://doi.org/10.1039/c6cp07499e. deep eutectic solvents (DESs) for catalytic persulfate activation toward non-radical
Dong-Wan, Cho, Kim, Sohyun, Tsang, Daniel C.W., Bolan, Nanthi S., Kim, Taejin, organics degradation and disinfection. J. Colloid Interface Sci. 601, 544–555.
Kwon, Eilhann E., Yong Sik, Ok, Song, Hocheol, 2018. Contribution of pyrolytic gas https://doi.org/10.1016/j.jcis.2021.05.080.
medium to the fabrication of co-impregnated biochar. J CO2 UTIL. 26, 476–486. Sun, C., Li, C., Wang, C., Qu, R., Niu, Y., Geng, H., 2012. Comparison studies of
https://doi.org/10.1016/j.jcou.2018.06.003. adsorption properties for Hg(II) and Au(III) on polystyrene-supported bis-8-
Garcia, G., Aparicio, S., Ullah, R., Atilhan, M., 2015. Deep eutectic solvents: oxyquinoline-terminated open-chain crown ether. Chem. Eng. J. 200–202, 291–299.
physicochemical properties and gas separation applications. Energy Fuel. 29, https://doi.org/10.1016/j.cej.2012.06.007.
2616–2644. https://doi.org/10.1021/ef5028873. Sun, F., Gao, J., Liu, X., Yang, Y., Wu, S., 2016. Controllable nitrogen introduction into
Kaili, Ding, Zhou, Xinyun, Hadiatullah, Hadiatullah, Lu, Yulin, Zhao, Guozhong, porous carbon with porosity retaining for investigating nitrogen doping effect on SO2
Jia, Shiru, Zhang, Rongfei, Yao, Yunping, 2021. Removal performance and adsorption. Chem. Eng. J. 290, 116–124. https://doi.org/10.1016/j.
mechanisms of toxic hexavalent chromium (Cr(VI)) with ZnCl2 enhanced acidic cej.2015.12.044.
vinegar residue biochar. J. Hazard Mater. 420, 126551. https://doi.org/10.1016/j. Thulluri, C., Balasubramaniam, R., Velankar, H.R., 2021. Generation of highly amenable
jhazmat.2021.126551. cellulose-iβ via selective delignification of rice straw using a reusable cyclic ether-
Kim, Kwang, Dutta, Ho, Sun, Tanmoy, Simmons, Jian, Singh, Blake, 2018. Biomass assisted deep eutectic solvent system. Sci. Rep. J. 11, 1591. https://doi.org/
pretreatment using deep eutectic solvents from lignin derived phenols. Green Chem. 10.1038/s41598-020-80719-x.
20, 809–815. https://doi.org/10.1039/C7GC03029K. Valentín-Reyes, J., García-Reyes, R.B., García-González, A., Soto-Regalado, E., Cerino-
Lawal, I.A., Klink, M., Ndungu, P., 2019. Deep eutectic solvent as an efficient modifier of Córdova, F., 2019. Adsorption mechanisms of hexavalent chromium from aqueous
low-cost adsorbent for the removal of pharmaceuticals and dye. Environ. Res. 179, solutions on modified activated carbons. J. Environ. Manag. 236, 815–822. https://
108837. https://doi.org/10.1016/j.envres.2019.108837. doi.org/10.1016/j.jenvman.2019.02.014.
Lehmann, J., 2009. Terra preta nova – where to from here? In: Woods, W.I., Teixeira, W. Wang, W., Gu, F., Zhu, J.Y., Sun, K., Jin, Y., 2020. Fractionation of herbaceous biomass
G., Lehmann, J., Steiner, C., WinklerPrins, A., Rebellato, L. (Eds.), Amazonian Dark using a recyclable hydrotropic p–toluenesulfonic acid (p–tsoh)/choline chloride
Earths: Wim Sombroek’s Vision. Springer Netherlands, Dordrecht, pp. 473–486. (chcl) solvent system at low temperatures. Ind. Crop. Prod. 150, 112423. https://doi.
https://doi.org/10.1007/978-1-4020-9031-8_28. org/10.1016/j.indcrop.2020.112423.
Ling, Z., Tang, W., Su, Y., Shao, L., Yong, Q., 2021. Promoting enzymatic hydrolysis of Webi, T.W., Chakravort, R.K., 1974. Pore and solid diffusion models for fixed-bed
aggregated bamboo crystalline cellulose by fast microwave-assisted dicarboxylic adsorbers. AIChE J. 20 (2), 228–238. https://doi.org/10.1002/aic.690200204.
acid deep eutectic solvents pretreatments. Bioresour. Technol. 333, 125122. Xiang, W., Zhang, X., Chen, K., Fang, J., Gao, B., 2019. Enhanced adsorption
Lu, X., Liu, X., Zhang, W., Wang, X., Wang, S., Xia, T., 2019. The residue from the acidic performance and governing mechanisms of ball-milled biochar for the removal of
concentrated lithium bromide treated crop residue as biochar to remove Cr (VI). volatile organic compounds (VOCs). Chem. Eng. J. 385, 123842. https://doi.org/
Bioresour. Technol. 296, 122348. https://doi.org/10.1016/j.biortech.2019.122348. 10.1016/j.cej.2019.123842.
Lu, Z., Zhang, H., Shahab, A., Zhang, K., Ullah, H., 2021. Comparative study on Yu, I.K.M., Xiong, X., Tsang, D.C.W., Wang, L., Hunt, A.J., Song, H., Shang, J., Ok, Y.S.,
characterization and adsorption properties of phosphoric acid activated biochar and Poon, C.S., 2019. Aluminium-biochar composites as sustainable heterogeneous
nitrogen-containing modified biochar employing eucalyptus as a precursor. J. Clean. catalysts for glucose isomerisation in a biorefinery. Green Chem. 21, 1267–1281.
Prod. 12, 127046. https://doi.org/10.1016/j.jclepro.2021.127046. https://doi.org/10.1039/C8GC02466A.
Ma, H., Yang, J., Gao, X., Liu, Z., Liu, X., Xu, Z., 2019. Removal of chromium (VI) from Zhang, Q., De Oliveira Vigier, K., Royer, S., Jérôme, F., 2012. Deep eutectic solvents:
water by porous carbon derived from corn straw: influencing factors, regeneration syntheses, properties and applications. Chem. Soc. Rev. 41 (21), 7108–7146. https://
and mechanism. J. Hazard Mater. 369, 550–560. https://doi.org/10.1016/j. doi.org/10.1039/C2CS35178A.
jhazmat.2019.02.063. Zhang, X., Chen, J., Ai, Z., Zhang, Z., Lin, L., Wei, H., 2018a. Targeting glycometabolic
Menghwar, P., Yilmaz, E., Sherazi, S., Soylak, M., 2018. A sensitive and selective deep reprogramming to restore the sensitivity of leukemia drug-resistant K562/ADM cells
eutectic solvent-based ultrasound-assisted liquid phase microextraction procedure to adriamycin. Life Sci. 215, 1–10. https://doi.org/10.1016/j.lfs.2018.10.050.
for separation-preconcentration and determination of copper in olive oil and water Zhang, X., Fu, W., Yin, Y., Chen, Z., Qiu, R., Simonnot, M.O., Wang, X., 2018b.
samples. Separ. Sci. Technol. 54 (15), 2431–2439. https://doi.org/10.1080/ Adsorption-reduction removal of Cr(VI) by tobacco petiole pyrolytic biochar: batch
01496395.2018.1547317. experiment, kinetic and mechanism studies. Bioresour. Technol. 268, 149–157.
Paiva, A., Craveiro, R., Aroso, I., Martins, M., Reis, R.L., Duarte, A.R.C., 2014. Natural https://doi.org/10.1016/j.biortech.2018.07.125.
deep eutectic solvents - solvents for the 21st century. Acs. Sustain. Chem. Eng. 2 (5), Zou, S., Wu, Y., Yang, M., Li, C., Tong, J., 2009. Thermochemical catalytic liquefaction of
1063–1071. https://doi.org/10.1021/sc500096j. the marine microalgae dunaliella tertiolecta and characterization of bio-oils. Energy
Fuel. 23 (4), 3753–3758. https://doi.org/10.1021/ef9000105.

You might also like