You are on page 1of 7

Bioresource Technology xxx (2017) xxx–xxx

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Fabrication of engineered biochar from paper mill sludge and its


application into removal of arsenic and cadmium in acidic water
Kwangsuk Yoon a,1, Dong-Wan Cho a,1, Daniel C.W. Tsang b, Nanthi Bolan c, Jörg Rinklebe a,d, Hocheol Song a,⇑
a
Department of Environment and Energy, Sejong University, Seoul 05006, Republic of Korea
b
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, China
c
Global Centre for Environmental Remediation, The University of Newcastle, Callaghan, NSW 2308, Australia
d
Soil- and Groundwater-Management, Institute of Foundation Engineering, Water- and Waste-Management, School of Architecture and Civil Engineering, University of Wuppertal,
Pauluskirchstrabe 7, 42285 Wuppertal, Germany

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Fabrication of engineered biochar


with paper mill sludge in a single
step.
 High fraction of Fe- and Ca solid
minerals in the biochar.
 pH neutralization of acidic
wastewater with addition of biochar.
 Simultaneous adsorption capability of
the biochar for As(V) and Cd(II).

a r t i c l e i n f o a b s t r a c t

Article history: An engineered biochar was fabricated via paper mill sludge pyrolysis under CO2 atmosphere, and its
Received 31 May 2017 adsorption capability for As(V) and Cd(II) in aqueous solution was evaluated in a batch mode. The char-
Received in revised form 29 June 2017 acterization results revealed that the biochar had the structure of complex aggregates containing solid
Accepted 4 July 2017
minerals (FeO, Fe3O4 and CaCO3) and graphitic carbon. Adsorption studies were carried out covering var-
Available online xxxx
ious parameters including pH effect, contact time, initial concentrations, competitive ions, and desorp-
tion. The adsorption of As(V) and Cd(II) reached apparent equilibrium at 180 min, and followed the
Keywords:
pseudo-second-order kinetics. The highest equilibrium uptakes of As(V) and Cd(II) were 22.8 and
Paper mill sludge
Magnetic biochar
41.6 mg g1, respectively. The adsorption isotherms were better described by Redlich-Peterson model.
Arsenic The decrease in As(V) adsorption was apparent with the increase in PO3 4 concentration, and a similar

Cadmium inhibition effect was observed for Cd(II) adsorption with Ni(II) ion. The feasibility of regeneration was
Adsorption demonstrated through desorption by NaOH or HCl.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction

Environmental pollution by heavy metals has markedly acceler-


⇑ Corresponding author at: 209 Neungdong-Ro, Gwangjin-Gu, Seoul 143-747,
ated with rapid industrialization in recent decades. Heavy metals
Republic of Korea. can pose detrimental threats to human health since they are highly
E-mail address: hcsong@sejong.ac.kr (H. Song). toxic and carcinogenic even at very low concentration levels (Beyki
1
Both authors equally contributed. et al., 2016). Contamination of surface water and groundwater

http://dx.doi.org/10.1016/j.biortech.2017.07.020
0960-8524/Ó 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
2 K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx

bodies by heavy metals originates from various industrial activities ions. Desorption kinetics experiment was conducted to investigate
including mining, metal plants, textile, batteries, pulp/paper, and the regeneration possibility of used biochar.
cloth industries (Chang et al., 2016). Among these contamination
sources, acid mine drainage (AMD) arising from mining activities 2. Materials and methods
is of a significant concern, and treatment of AMD is one of the
greatest environmental challenges in the world (Hedrich and 2.1. Chemical reagents
Johnson, 2014). The AMD has a very low pH value (<3), and con-
tains diverse inorganic contaminants such as anionic metalloids Paper mill Sludge (PMS) was obtained from a wastewater treat-
and cationic metals (Gurung et al., 2017). ment plant of Moorim Paper Co., South Korea. Sodium arsenate
Arsenic (As) and cadmium (Cd) are anionic and cationic heavy dibasic heptahydrate (Na2HAsO47H2O), cadmium nitrate tetrahy-
metals commonly found in AMD, respectively. Arsenic occurs as drate (Cd(NO3)24H2O), sodium phosphate dibasic (HNa2O4P),
two species of arsenate (As(V)) and arsenite (As(III)), and is sodium bicarbonate (NaHCO3), sodium chromate (Na2CrO4), nickel
mainly found in mining regions rich in arseno-metal sulfide min- nitrate (Ni(NO3)2), calcium nitrate (Ca(NO3)2) and ammonium
erals (Le Pape et al., 2017). Due to the high toxicity and adverse nitrate (NH4NO3) were purchased form Sigma Aldrich, USA. Induc-
effects of arsenic, the recommended standard in drinking water tively coupled plasma (ICP) standard solution (multi-element stan-
level is set up at 10 ppb in the WHO guideline (Mahimairaja dard solution XVI, 100 mg L1 of As, Cd, Ca, Cr, Fe, Pb etc.) in diluted
et al., 2005). Cadmium has been also reported to be very toxic, nitric acid (2%) was purchased from Merck Millipore Co., USA.
and the exposure to Cd can lead to fatal damages to human organs
including kidney, nervous system, and blood vessel (Bolan et al.,
2.2. Preparation of engineered biochar
2013a,b). Various technologies such as chemical precipitation
(Guo et al., 2016), membrane separation (Arevalo et al., 2013),
A batch type tubular reactor (TR) was used to fabricate engi-
and adsorption (Streat et al., 2008) have been employed to treat
neered biochar. A 25.4 mm of stainless Ultra Torr Vacuum Fitting
waters contaminated with As and Cd. Among these technologies,
(Swagelok SS-4-UT-6-400) was used to connect the both ends of
adsorption has gained much recognition as a viable method
the TR that was housed in a quartz tubing (25.4-mm outer diame-
because of its simplicity, low operation cost, technical flexibility
ter and 610 mm in length, Chemglass CGQ-0900T-13, USA). The
(Jiang et al., 2016).
atmospheric gas of CO2 used in the experiments was obtained from
In the treatment of acidic metal-contaminated water by adsorp-
Daesung Gas Co., South Korea and was ultra-high purity (UHP)
tion, it is highly desirable to restore the quality of water by increas-
grade. The gas flow rate was controlled with mass flow controller
ing the solution pH to neutral values. In this regard, the use of
(5850E, Brooks, USA). The flow rate was increased from
alkaline adsorbent that can neutralize the acidity is preferable.
500 mL min1 set up in the previous study (Cho et al., 2017a) to
For example, limestone (CaCO3) or steel slag containing Fe, Mn,
600 mL min1 to provide enough amount of CO2 in the production
Ca, and Si have been used as treatment medium for acidic wastew-
of a large amount of biochar. A split-hinged programmable furnace
ater due to their great potential in increasing the solution pH as
(FT-830, DAIHAN Scientific) was used to manipulate the pyrolysis
well as adsorbing/precipitating the metals (Lee et al., 2016). How-
temperature. The loaded amount of PMS was 40 g inside the TR
ever, liming is not suitable for the treatment of As-contaminated
and the heating rate of 10 °C min1 ranged from 270 to 720 °C
wastewater because limestone has little ability to adsorb anionic
(45 min). This temperature condition has been fully justified by
metalloids, whereas steel slag has demerits such as leaching prob-
our previous work to generate highly porous biochar from diverse
lem of heavy metals in acidic conditions.
biomass feedstocks (Cho et al., 2017 b). The resulting biochar is
Most biochars are alkaline due to the release of alkaline miner-
denoted as paper mill sludge-derived biochar (PMSB).
als (i.e., Ca, Mg) from thermal degradation of biomass feedstocks
during pyrolysis. Biochar has received considerable interests as a
treatment medium for environmental remediation because of its 2.3. Characterization of the biochar
advantageous physical properties associated with high surface area
and porous structures (Ahmad et al., 2014). Biochar prepared from The morphology and elemental information of the biochar were
poultry litter has been applied to treat acidic wastewater, and syn- examined using a field emission-scanning electron microscope (FE-
ergistic effect for metals removal was demonstrated (Oh and Yoon, SEM, Hitachi S-4700, Japan) and energy-dispersive X-ray spec-
2013). However, the issues associated with separation of used bio- troscopy (EDS). Brunauer–Emmett–Teller (BET) analysis was con-
char and low adsorption ability toward anionic metalloids of bio- ducted with a surface analyzer (Belsorp-mini II, USA). X-ray
char were not fully resolved in the study. diffraction (XRD, D8 Advance, Bruker-AXS) analysis was performed
Paper mill sludge (PMS) is generated during wastewater treat- to indentify the mineral phases on the biochar. Thermo-
ment in pulp and paper industry, and contains high concentrations gravimetric analysis (TGA, Q600, TA Instrument) was carried out
of inorganics (i.e., Ca and Fe species) due to addition of those chem- to identify the chemical composition. High-performance X-ray
icals in the treatment process (Kunhikrishnan et al., 2012). Our photoelectron spectrometer (XPS, K-ALPHA+, Thermo Scientific)
previous work has used PMS as a feedstock of pyrolysis to obtain was used to investigate the composition and chemical state of ele-
renewable energy (i.e., syngas) in the presence of CO2 (Cho et al., ments on the surface of the biochar. Raman spectra were obtained
2017a). It also produced magnetic biochar rich in Fe/Ca minerals using a Dimension P1 Raman spectroscopy (Lambda Solution Co.,
with high alkalinity. These characteristics of PMS-derived biochar USA).
could be advantageous to the treatment of acidic wastewater con-
taining cationic metals and anionic metalloids. 2.4. Adsorption experiments
In this work, engineered biochar was fabricated with PMS via
pyrolysis under CO2 atmosphere, and applied for the adsorption All the adsorption experiments were performed using 25 mL
of As(V) and Cd(II) in strongly acidic solution. The physicochemical high-density polyethylene (HDPE, Fisher Scientific, USA) vials in
properties of the biochar were characterized using various micro- duplicate at room temperature (25 ± 2 °C), and the HDPE caps were
scopic/spectroscopic instruments. The adsorption ability of the used to seal the vials. The pH values of solution were adjusted with
biochar was evaluated under various parameters such as solution 5 M HNO3 and 5 M NaOH. The solution was shaken at 200 rpm
pH, contact time, initial concentrations, and effect of competitive using an orbital shaker (Stuart, UK). The samples were filtered with

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx 3

0.45 lm syringe filter (Whatman, USA). For pH effect experiments, the Freundlich model constant, and n is a constant related to the
0.02 g PMSB was added into 20 mL of solution containing strength of adsorption. The Redlich–Peterson model is a three-
22.7 mg L1 As(V) and 33.0 mg L1 Cd(II) in binary adsorbate sys- parameter adsorption model combining features of the Langmuir
tem, and the suspension was reacted for 6 h. The adsorption kinet- and Freundlich models (Foo and Hameed, 2010). The Redlich-
ics experiments were conducted by adding 0.02 g PMSB into 20 mL Peterson isotherm model is represented by the following equation,
solution containing 21.8 mg L1 As(V) and 33.0 mg L1 Cd(II) solu-
tion over the time course of 240 min. The deviation of the values kR C e
qe ¼ ð5Þ
from the concentrations in control samples was less than 1 þ aR C be
0.05 mg L1. Further experiments were conducted in a single
where kR, aR, and b (0 < b < 1) are the Redlich-Peterson model con-
adsorbate system. Adsorption isotherm experiments were carried
stants. kR is the Redlich–Peterson constant related to the adsorption
out for 24 h by changing initial concentrations of As(V) and Cd(II)
capacity, aR is the Redlich–Peterson constant related to the affinity
over a range of 20.9–189.5 mg L1 and 21.0–281.7 mg L1, respec-
of the binding sites and b is the Redlich–Peterson constant related
tively, at a constant amount of 1g L1 PMSB.
to the adsorption intensity.
For the experiments to investigate the effect of competitive
ions, As(V) adsorption tests were performed in the presence of
varying concentration of phosphate (PO3 4 , 0.5–3.0 mM) at initial 3. Results and discussion
As(V) concentration of 30 mg L1. For the Cd(II) adsorption, the
effect of nickel (Ni2+) in the ragne of 0.5–3.0 mM was investigated 3.1. Physicochemical properties of PMSB
at initial Cd(II) concentration of 30 mg L1. Desorption kinetics
experiments based on ionic exchange [anion exchange – OH vs The physicochemical properties of PMSB were characterized by
As(V); cation exchange – H+ vs Cd(II)] were performed by adding microscopic/spectroscopic analyses. The results of FE-SEM/EDS
5 M NaOH or 5 M HCl solution into the solutions in which 2 g L1 indicate that PMSB possessed the morphology of irregularly
PMSB was reacted with 10.2 mg L1 As(V) or 11.4 mg L1 Cd(II) shaped microscale aggregates consisting of 12.43 wt% C, 33.85 wt
for 36 h. The concentrations of As(V) and Cd(II) in the solutions % O, 2.8 wt% S, 11.27 wt% Ca, and 38.73 wt% Fe. The N2
were determined by ICP-optical emission spectrometry (ICAP adsorption-desorption isotherm result of PMSB showed ‘‘Type III-
6000 series, Thermo Scientific, USA). The adsorbed amounts at like” isotherm and BET surface area of 13 m2 g1, which indicates
equilibrium (qe) were calculated using the equation, that it possessed macroporous structure. This observation may be
qe ðmg g1 Þ ¼ ðC 0 C
W
e ÞV
, where C0 and Ce are initial and equilibrium due to that the formed solid mineral aggregates during pyrolysis
concentrations of adsorbates (mg L1), respectively, W is the dry (shown in FE-SEM image) can block small pores (micropores and
mass of the adsorbent (g), and V is the volume of the solution (L). mesopores) in biochar matrix. The similar phenomenon has been
observed in the BET results of Ruan et al. (2015), which showed
macroporous structure and relatively low surface are
2.5. Data analysis
(10.6 m2 g1) of bentonite/hematite-embedded biochar. The XRD
patterns of PMSB showed that the peak patterns assigned to calcite
The obtained kinetics data was evaluated by pseudo-first-order
(CaCO3) (2h = 29.4°, 36.2°, 39.4°, 43.3°, 47.4°, 48.5°, and 57.4°,
(PFO) and pseudo-second-order (PSO) kinetic models. A linear form
JCPDS Card No. 05-0586) were apparent. Calcite added as a filler
of PFO kinetics can be expressed as follows:
for paper making process remained as calcite during the pyrolysis
lnðqe  qt Þ ¼ lnqe  k1 t ð1Þ process because of the high partial pressure of CO2. Calcite might
1 be embedded into PMSB as a filler during paper making process.
In this equation, qt (mg g ) is the amount adsorbed at time t
There were a few peaks observed at 2h = 30.0°, 35.4°, 38.2°, 43.1°,
(min), and k1 (min1) is the rate constant of pseudo-first-order
53.5°, 57.0°, and 62.6°, which correspond to magnetite (Fe3O4)
adsorption. The values of ln (qe  qt) are calculated from the
(JCPDS Card No. 89-3854). The impregnated Fe species onto PMS
adsorption kinetics data. The PSO based on the adsorption equilib-
could be converted into magnetite during pyrolysis. The magnetic
rium capacity is expressed as follows:
property of PMSB mainly results from the presence of magnetite.
t 1 t These observations are consistent with the XRD and SQUID results
¼ þ ð2Þ
qt k2 q2e qe of pyrolyzed PMS in our previous work (Cho et al., 2017a). In addi-
tion, four distinctive peaks were observed at 2h = 36.2°, 42.0°,
In this equation, k2 (g mg1 min1) is the rate constant of 61.1°, 73.2°, and 76.9°, which are related to the presence of wüstite
pseudo-second-order adsorption.
(FeO) (JCPDS Card No. 80-0686). The FeO phase detected in the
The obtained isotherm data were fitted with Langmuir, Fre- XRD spectrum is differentiated with the XRD pattern that showed
undlich, and Redlich-Peterson isotherm models. Among three iso- only Fe3O4 as Fe mineral phase in our previous work (Cho et al.,
therm models, the Langmuir and Freundlich isotherm models are 2017a). The occurrence of FeO could be a result arising from
commonly used isotherm models. The Langmuir model assumes incomplete oxidation of reduced intermediate (i.e., FeO and Fe0)
adsorption occurs at specific homogeneous sites with equal to Fe3O4 by CO2. Zhang et al. (2015) suggest that ferrous salts or
adsorption energy, and the Freundlich isotherm assumes that ferrous-based minerals (e.g., FeSO4, FeCl2, and FeCO3) that were
adsorbent surface sites have a spectrum of active sites with differ- used as additives are converted into metallic Fe such as FeO and
ent binding energies (Deng and Ting, 2005). The Langmuir and Fre- Fe0 during pyrolysis of coal, followed by subsequent re-oxidation
undlich isotherm models are expressed as follows: to Fe3O4 with CO2 at high temperature. In this work, the loaded
1 1 1 amount of PMS (40 g) was larger compared to the case of our pre-
¼ þ ð3Þ vious work (3 g) (Cho et al., 2017a), although the CO2 flow rate
qe kL C e qm qm
increased from 500 to 600 mL min1. The relatively high content
of FeO phase could be incompletely oxidized to Fe3O4 at the pre-
qe ¼ kf C e1=n ð4Þ
sent condition, thereby showing the co-existence of FeO and
1
In these equations, qm (mg g ) is the maximum amount of Fe3O4 in the PMSB.
adsorption, kL (L mg1) is the Langmuir model constant, Ce (mg L1) One peak (D-band, 1355 cm1) originated from sp3 hybridiza-
is the concentration of contaminant in the solution, kf (L mg1) is tion and another peak (G-band, 1564 cm1) induced by crystalline

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
4 K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx

graphitic/sp2 carbon atoms were found in the Raman spectra,


respectively. In addition, the existence of calcite and magnetite
was validated by the bands at 318 cm1, 502 cm1, and 719 cm1
(Liu et al., 2017).
To approximately examine the fractions of carbonized carbon
and inorganics in the PMSB, TGA was conducted by completely
burning PMSB at the heating rate of 10 °C min1 in oxygen-rich
atmosphere (air flow rate = 150 mL min1). Interestingly, the slight
increase in weight (2.49 wt%) was observed in the temperature
range from 138 to 356 °C, which is possibly attributed to the oxida-
tion of Fe(II) species present in magnetite surface to Fe(III) species
(Sun et al., 2004; Yoon et al., 2011). Afterwards, the weight was
gradually decreased with increasing temperature to 800 °C. The
total reduced weight of PMSB measured at 800 °C was found to
be about 10 wt%, which suggests that PMSB mainly consists of
inorganic compounds (i.e., Fe- and Ca minerals) in addition to the
small amount of carbon that can be burnt at high temperature.
The relatively low fraction of residual carbon can be explained by
the inherently high level of inorganic minerals in the PMS (i.e., high
ash content). Another possibility is the depletion of carbon due to
greater production of syngas in CO2 atmosphere compared to
pyrolysis in reducing atmosphere (N2). The origin of high-yield
syngas (especially CO) in the presence of CO2 is mostly condens-
able hydrocarbon generated during pyrolysis, which has been fully
validated by our previous studies (Cho et al., 2017 b; Kwon et al.,
2015). However, solid-phase carbon (C) could be also consumed Fig. 1. Solution pH effect on the As(V) and Cd(II) adsorption by PMSB in binary
adsorbate system; inset figure of initial pH vs final pH (contact time = 6 h,
for the generation of syngas during pyrolysis at this experimental adsorbent dosage = 1 g L1, final pH range = 2.6–8.6, initial As(V) conc.
condition due to the catalytic effect of Fe/Ca species, thereby = 22.7 mg L1 and initial Cd(II) conc. = 33.0 mg L1).
decreasing the fraction of carbon in PMSB.
To gain insight into the elemental information on the surface of
PMSB, XPS analysis was conducted. The XPS spectra revealed peaks could show the mutual cooperation in adsorption via electrostatic
assigned to C1s, O1s, N1s and Fe2p3. The C1s spectrum shows attraction (formation of a ternary cation–anion-surface complex)
peaks at two binding energies (284.68 eV and 290.1 eV), which cor- and co-precipitation, especially in the presence of Fe3O4 (Jiang
responds to C-C and CO2 3 , respectively (Rahman et al., 2014). The et al., 2013).
presence of carboxyl groups in the biochar was found with the
large peak observed at 531.88 eV in the O1s spectrum. This obser- 3.3. Adsorption kinetics
vation is consistent with the peak observed at 513.9 eV assigned to
C@O group in the XPS analysis of carbon composite (Koilraj and Adsorption kinetics experiments of As(V) and Cd(II) using
Sasaki, 2017). Two peaks observed at binding energy at 398.58 1 g L1 of PMSB were conducted at final pH 6.2, and the results
and 400.58 eV in the N1s spectrum represents CAN and C@N, are presented in Fig. 2. The time required to reach apparent equi-
respectively, which is consistent with a previous study (Ma et al., librium for As(V) and Cd(II) was found to be 240 min. The experi-
2012). The peaks at 711.18 eV and 724.88 eV were observed in mental data was evaluated by PFO and PSO kinetic models as
the Fe2p XPS spectrum, which corresponds to the 2p3/2 binding mentioned in Section 2.5. The kinetics parameters for As(V) and
energies of Fe3O4 (Zhao et al., 2012). Cd(II) adsorption are presented in Table 1. The experimental
adsorption capacities of As(V) and Cd(II), qe (cal), were 9.63 and
3.2. Effect of solution pH 16.57 mg g1, respectively, which are close to the calculated
adsorption capacities from PSO model (9.71 and 16.58 mg g1).
Addition of PMSB into the solution containing 22.7 mg L1 As(V) The correlation coefficients (R2) of PSO for As(V) and Cd(II) adsorp-
and 33.0 mg L1 Cd(II) increased the solution pH from 2.0–4.0 to tion (R2 = 0.9993 and 0.9995) were higher compared to those of
2.5–8.6 due to the pH buffering effect of calcite as well as the alka- PFO (R2 = 0.8966 and 0.9219), which may suggest that adsorption
linity of ash minerals (Fig. 1), which has been observed in our pre- of both ions is governed by chemisorption (Dinu and Dragan,
vious work (Cho et al., 2017a). This observation clearly indicates 2010). This estimated adsorption mechanism is consistent with
the effectiveness of PMSB in the neutralization of acidic the suggested modes of As(V) and Cd(II) adsorption involving
wastewater. specific surface interactions (Jiang et al., 2013).
The adsorption of Cd(II) sharply decreased from 31.2 to
1.1 mg g1 with the decrease of final pH from 8.6 to 2.4, which is 3.4. Adsorption isotherm
attributed to the increase of hydrogen ions or protons (H+) that
act as a competitor for the adsorption sites for Cd(II) (Zhu et al., The adsorption isotherm experiments were conducted to evalu-
2016). Interestingly, the removal of As(V) was slightly enhanced ate the adsorption capacities of As(V) and Cd(II) in the initial con-
at the high pH condition. This trend is not matched with the pH centration ranges of 20.9–189.5 mg L1 and 21.0–281.7 mg L1,
edge of As(V) anion adsorption in single adsorbate system (Guo respectively. The highest adsorption capacities of As(V) and Cd(II)
et al., 2015). In general, the adsorption of As(V) would be decreased at this experimental condition were found to be 23.1 and
due to electrostatic repulsion between negatively charged As(V) 41.6 mg g1, respectively (Fig. 3). For a comparison, meso-iron oxy-
ions and hydroxyl ion (OH) in the single adsorbate system. The hydroxide impregnated bead exhibited greater As(V) adsorption
opposite trend shown in the binary adsorbate system can be attrib- capacity (71.9 mg g1) than Cd(II) (9.9 mg g1) (Chung et al.,
uted possible interactions between As(V) and Cd(II). Both ions 2014), while modified multi-wall carbon nanotubes showed the

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx 5

overlapped as shown in Fig. 3. The correlation coefficients (R2) of


the Redlich-Peterson isotherm (As(V): 0.9989 and Cd(II): 0.9935)
are relatively larger than those of the Langmuir isotherm (As(V):
0.8995 and Cd(II): 0.7714) and the Freundlich isotherm (As(V):
0.9925 and Cd(II): 0.9589). When the b value obtained using the
Redlich-Peterson isotherm is close to 1, the adsorption curve
resembles the curve of the Langmuir isotherm. The b values
obtained in this study were 0.72 and 0.74, indicating weak confor-
mity to the Langmuir model. Also, the values of kf in the Freundlich
parameters (As(V): 5.59 and Cd(II): 8.16 L mg1) were close to the
values of kR/aR in the Redlich-Peterson parameters (As(V): 5.55 and
Cd(II): 8.69 L mg1). These results indicate that the isotherms are
more approaching to the Freundlich isotherm form than the Lang-
muir isotherm form, indicating the contribution of heterogeneous
surface active sites for adsorption of these ions (Zhang et al., 2011).

3.5. Effect of competitive ions

The effect of co-existing ions on the removal efficiencies of As


(V) and Cd(II) by PMSB was examined at different concentrations
from 0.5 to 3.0 mM at final pH 7.2. The results are presented in
Fig. 2. Adsorption kinetics of (a) As(V) and (b) Cd(II) in binary adsorbate system Fig. 4. The co-existence of phosphate (PO34 ) gave a negative effect
(adsorbent dosage = 1 g L1, final pH = 6.2, initial concentrations of As(V) and Cd(II) on the adsorption of As(V), showing the sharp decrease of As(V)
= 21.8 mg L1 and 33.0 mg L1) *plots of adsorption were fitted by pseudo-second-
removal efficiency from 61.0 to 10.4% with the increase of PO3 4
order kinetics model (PSO).
from 0.5 to 3.0 mM. The substantial reduction in As(V) may be
attributed to the strong competition for the adsorption sites of
PMSB between PO3 4 and As(V). The similar chemical properties
opposite results (As(V): 13.0 mg g1 at pH 4 and Cd(II): of PO3 3
4 and As(V) could increase the inhibition effect of PO4 on
77.6 mg g1 at pH 8) (Veličković et al., 2013). the adsorption of As(V) (Bolan et al., 2013a,b). The reduction in
These experimental data were fitted with Langmuir, Freundlich, Cd(II) was apparent with the increase in Ni(II) concentration from
and Redlich-Peterson isotherm models as mentioned in Section 2.5. 0.5 to 3.0 mM. The removal of Cd(II) decreased from 95.8 to 64.3%
The relevant constants for the Langmuir, Freundlich, and Redlich- in the presence of 3.0 mM Ni(II) because of the competition of Ni
Peterson isotherm are presented in Table 2. First, the Freundlich (II) for the adsorption sites on PMSB. The observed effect of Ni(II)
constants (n) for As(V) and Cd(II) (3.60 and 3.55) are in the range can be explained by the difference in ionic radius of Ni(II) and Cd
of 1–10, which suggests that the adsorption is favorable. The fitting (II). Aliabadi et al. (2013) suggested that greater adsorption of Ni
lines of Freundlich and Redlich-Peterson isotherm were almost (II) occurs onto adsorbent compared to Cd(II) due to the relatively

Table 1
Adsorption kinetics model parameters for the adsorption of As(V) and Cd(II) by PMSB.

Adsorbate Pseudo-first-order kinetics model Pseudo-second-order kinetics model


qe (exp) k1 (L min1) qe (cal) (mg g1) R2 k2 (g mg1 min1) qe (cal) (mg g1) R2
As(V) 9.63 0.0161 2.646 0.8966 0.0194 9.71 0.9993
Cd(II) 16.57 0.0138 4.204 0.9219 0.0122 16.58 0.9995

Fig. 3. Adsorption isotherm of PMSB for (a) As(V) and (b) Cd(II) (contact time = 24 h, adsorbent dosage = 1 g L1, final pH = 6.5, initial concentration ranges of As(V) and Cd(II)
= 20.9–189.5 mg L1 and 21.0–281.7 mg L1).

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
6 K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx

Table 2
Adsorption isotherm model parameters for the removal of As(V) and Cd(II) by PMSB.

Adsorbate Langmuir isotherm Freundlich isotherm Redlich-Peterson isotherm


qm (mg g 1
) kL (L mg 1
) R2
kf L mg 1
) n R2
kR (L mg1) aR b R2
As(V) 22.76 0.087 0.8995 5.59 3.60 0.9925 1986 357 0.72 0.9989
Cd(II) 41.55 0.088 0.7714 8.16 3.55 0.9589 1690 194 0.74 0.9935

Fig. 4. Effect of co-existing ions on the adsorption of (a) As(V) and (b) Cd(II) by PMSB [As(V): 0.5–1.5 mM PO3
4 ; Cd(II): 0.5–1.5 mM Ni(II)].

small ionic radius of Ni(II) in binary adsorbate system. In other


words, adsorbates having small ionic radius can be more accessible
and adsorbed onto available active sites. As a result, the high
adsorption affinity toward Ni(II) could contribute to the substantial
decrease of Cd(II) adsorption.

3.6. Desorption studies

To further investigate the binding mechanisms of As(V) and Cd


(II) onto PMSB, the desorption kinetics experiments were carried
out using 0.5 M NaOH and 0.5 M HNO3 solutions, respectively.
The desorption efficiencies of As(V) and Cd(II) were calculated by
C f C t
using the following equation, Dð%Þ ¼ 100  Cf
, where D is the des-
orption efficiency (%), Cf is the concentration of adsorbate mea-
sured at final contact time (mg L1), Ct is the concentration of
adsorbate measured at interval time (mg L1). For As(V) adsorp-
tion/desorption experiments, the complete removal of As(V) was
achieved within 300 min (Fig. 5). In the desorption experiment,
73.4% of the totally adsorbed As(V) was rapidly desorbed from
the used PMSB at 60 min, and the desorption efficiency increased
gradually, followed by 100% desorption efficiency achieved at
36 h. The slow desorption rate evidences the chemical bonding Fig. 5. Adsorption/desorption kinetics of As(V) and Cd(II) of PMSB in 0.5 M NaOH or
between As(V) and Fe-solid mineral surfaces (Mostafa et al., 0.5 M HNO3 (adsorbent dosage = 2 g L1, initial concentrations of As(V) and Cd(II)
2010). Meanwhile, the adsorption process of Cd(II) reached 100% = 10.2 mg L1 and 11.4 mg L1 final pH for adsorption process = 7.11).
at 300 min (Fig. 5). One interesting observation is that the desorp-
tion process of Cd(II) was completely finished within 240 min, con-
observed (Ma et al., 2015). Thus, the time required to regenerate
trary to the desorption of As(V) (100% desorption at 36 h). The
the used PMSB is entirely dependent on the target contaminants.
relatively rapid desorption of Cd(II) is ascribable to the relatively
weak sorption of cations by surface functional groups in PMSB. In
other words, almost all Cd(II) ions could be easily desorbed in 4. Conclusions
the short period of time via protonation of surface functional
groups (COO- ? COOH). The rapid desorption of Cd(II) from spent A paper mill sludge-derived biochar was prepared by pyrolyz-
carbon-based adsorbent using acidic solution has been previously ing paper mill sludge in CO2 atmosphere. The resulting biochar

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020
K. Yoon et al. / Bioresource Technology xxx (2017) xxx–xxx 7

possessed graphitic carbon containing a large amount of solid min- Guo, X., Yang, Z., Dong, H., Guan, X., Ren, Q., Lv, X., Jin, X., 2016. Simple combination
of oxidants with zero-valent-iron (ZVI) achieved very rapid and highly efficient
erals rich in Fe and Ca species. Batch adsorption experiments under
removal of heavy metals from water. Water Res. 88, 671–680.
various experimental parameters including pH effect, contact time, Gurung, S.R., Wijesekara, H., Seshadri, B., Stewart, R.B., 2017. Chapter 3 Sources and
initial concentrations, and competitive ions effect, and desorption management of acid mine drainage. In: Bolan, N.S., Kirkham, M.B., Ok, Y.S.
demonstrated great adsorption capability of biochar for As(V) (Eds.), Spoil to Soil: Mine Site Rehabilitation and Revegetation. CRC Press.
Hedrich, S., Johnson, D.B., 2014. Remediation and selective recovery of metals from
and Cd(II) in aqueous solution. The overall results showed that acidic mine waters using novel modular bioreactors. Environ. Sci. Technol. 48,
pyrolysis of paper mill sludge using CO2 as reaction medium pro- 12206–12212.
duced Fe/Ca-rich engineered biochar, and demonstrated biochar Jiang, W., Lv, J., Luo, L., Yang, K., Lin, Y., Hu, F., Zhang, J., Zhang, S., 2013. Arsenate and
cadmium co-adsorption and co-precipitation on goethite. J. Hazard. Mater. 262,
potential as pH neutralizer and adsorbent for As(V) and Cd(II). 55–63.
Jiang, T., Zhong, W., Jafari, T., Du, S., He, J., Fu, Y.-J., Singh, P., Suib, S.L., 2016. Siloxane
D4 adsorption by mesoporous aluminosilicates. Chem. Eng. J. 289, 356–364.
Conflict of interest Koilraj, P., Sasaki, K., 2017. Selective removal of phosphate using La-porous carbon
composites from aqueous solutions: Batch and column studies. Chem. Eng. J.
317, 1059–1068.
The authors declare no competing financial interest. Kunhikrishnan, A., Bolan, N.S., Müller, K., Laurenson, S., Naidu, R., Kim, W.I., 2012.
The influence of wastewater irrigation on the transformation and bioavailability
of heavy metal(loid)s in soil. Adv. Agron. 115, 215–297.
Acknowledgement Kwon, E.E., Cho, S.H., Kim, S., 2015. Synergetic sustainability enhancement via
utilization of carbon dioxide as carbon neutral chemical feedstock in the
This work was supported by the National Research Foundation thermo-chemical processing of biomass. Environ. Sci. Technol. 49, 5028–5034.
Le Pape, P., Battaglia-Brunet, F., Parmentier, M., Joulian, C., Gassaud, C., Fernandez-
of Korea (NRF) grant funded by the Korea government (MSIP) Rojo, L., Guigner, J., Ikogou, M., Stetten, L., Olivi, L., Casiot, C., Morin, G., 2017.
(No.2014R1A2A2A09049496). Complete removal of arsenic and zinc from a heavily contaminated acid mine
drainage via an indigenous SRB consortium. J. Hazard. Mater. 321, 764–772.
Lee, W.C., Lee, S.W., Yun, S.T., Lee, P.K., Hwang, Y.S., Kim, S.O., 2016. A novel method
Appendix A. Supplementary data of utilizing permeable reactive kiddle (PRK) for the remediation of acid mine
drainage. J. Hazard. Mater. 301, 332–341.
Liu, A., Liu, J., Han, J., Zhang, W.X., 2017. Evolution of nanoscale zero-valent iron
Supplementary data associated with this article can be found, in (nZVI) in water: Microscopic and spectroscopic evidence on the formation of
the online version, at http://dx.doi.org/10.1016/j.biortech.2017.07. nano- and micro-structured iron oxides. J. Hazard. Mater. 322, 129–135.
Ma, Y., Zhang, C., Ji, G., Lee, J.Y., 2012. Nitrogen-doped carbon-encapsulation of
020. Fe3O4 for increased reversibility in Li+ storage by the conversion reaction. J.
Mater. Chem. 22, 7845–7850.
Ma, X., Yang, S.T., Tang, H., Liu, Y., Wang, H., 2015. Competitive adsorption of heavy
References metal ions on carbon nanotubes and the desorption in simulated biofluids. J.
Colloid Interface Sci. 448, 347–355.
Ahmad, M., Rajapaksha, A.U., Lim, J.E., Zhang, M., Bolan, N., Mohan, D., Vithanage, Mahimairaja, S., Bolan, N.S., Adriano, D.C., Robinson, B., 2005. Arsenic contamination
M., Lee, S.S., Ok, Y.S., 2014. Biochar as a sorbent for contaminant management in and its risk management in complex environmental settings. Adv. Agron. 86, 1–
soil and water: A review. Chemosphere 99, 19–33. 82.
Aliabadi, M., Irani, M., Ismaeili, J., Piri, H., Parnian, M.J., 2013. Electrospun nanofiber Mostafa, M.G., Chen, Y.H., Jean, J.S., Liu, C.C., Teng, H., 2010. Adsorption and
membrane of PEO/Chitosan for the adsorption of nickel, cadmium, lead and desorption properties of arsenate onto nano-sized iron-oxide-coated quartz.
copper ions from aqueous solution. Chem. Eng. J. 220, 237–243. Water Sci. Technol. 62, 378–386.
Arevalo, J., Ruiz, L.M., Perez, J., Moreno, B., Gomez, M.A., 2013. Removal performance Oh, S.Y., Yoon, M.K., 2013. Biochar for treating acid mine drainage. Environ. Eng. Sci.
of heavy metals in MBR systems and their influence in water reuse. Water Sci. 30 (10), 589–593.
Technol. 67, 894–900. Rahman, M.M., Netravali, A.N., Tiimob, B.J., Rangari, V.K., 2014. Bioderived ‘‘Green”
Beyki, M.H., Alijani, H., Fazli, Y., 2016. Poly o-phenylenediamine-MgAl@CaFe2O4 composite from soy protein and eggshell nanopowder. ACS Sustain. Chem. Eng.
nanohybrid for effective removing of lead(II), chromium(III) and anionic azo 2, 2329–2337.
dye. Process Saf. Environ. Protect. 102, 687–699. Ruan, Z.-H., Wu, J.-H., Huang, J.-F., Lin, Z.-T., Li, Y.-F., Liu, Y.-L., Cao, P.-Y., Fang, Y.-P.,
Bolan, N., Mahimairaja, S., Kunhikrishnan, A., Choppala, G., 2013a. Phosphorus- Fang, Y.-P., Xie, J., Jiang, G.-B., 2015. Facile preparation of rosin-based biochar
arsenic interactions in variable-charge soils in relation to arsenic mobility and coated bentonite for supporting a-Fe2O3 nanoparticles and its application for Cr
bioavailability. Sci. Total Environ. 463–464, 1154–1162. (VI) adsorption.
Bolan, N.S., Makino, T., Kunhikrishnan, A., Kim, P.J., Ishikawa, S., Murakami, M., Streat, M., Hellgardt, K., Newton, N.L.R., 2008. Hydrous ferric oxide as an adsorbent
Naidu, R., Kirkham, M.B., 2013b. Cadmium contamination and its risk in water treatment: Part 3: Batch and mini-column adsorption of arsenic,
management in rice ecosystems. Adv. Agron. 119, 183–273. phosphorus, fluorine and cadmium ions. Process Saf. Environ. Protect. 86, 21–
Chang, Y., Lai, J.-Y., Lee, D.-J., 2016. Thermodynamic parameters for adsorption 30.
equilibrium of heavy metals and dyes from wastewaters: Research updated. Sun, S., Zeng, H., Robinson, D.B., Raoux, S., Rice, P.M., Wang, S.X., Li, G., 2004.
Bioresour. Technol. 222, 513–516. Monodisperse MFe2O4 (M = Fe Co, Mn) nanoparticles. J. Am. Chem. Soc. 126,
Cho, D.W., Kwon, G., Ok, Y.S., Kwon, E.E., Song, H., o et al., 2017 b. Reduction of 273–279.
bromate by cobalt-impregnated biochar fabricated via pyrolysis of lignin using Veličković, Z.S., Bajić, Z.J., Ristić, M.D., Djokić, V.R., Marinković, A.D., Uskoković, P.S.,
CO2 as a reaction medium. ACS Appl. Mater. Interfaces 9, 13142–13150. Vuruna, M.M., 2013. Modification of multi-wall carbon nanotubes for the
Cho, D.W., Kwon, G., Yoon, K., Tsang, Y.F., Ok, Y.S., Kwon, E.E., Song, H., 2017a. removal of cadmium, lead and arsenic from wastewater. Dig. J. Nanomater.
Simultaneous production of syngas and magnetic biochar via pyrolysis of paper Biostruct. 8, 501–511.
mill sludge using CO2 as reaction medium. Energy Conv. Manag. 145, 1–9. Yoon, T., Chae, C., Sun, Y.K., Zhao, X., Kung, H.H., Lee, J.K., 2011. Bottom-up in situ
Chung, S.-G., Ryu, J.-C., Song, M.-K., An, B., Kim, S.-B., Lee, S.-H., Choi, J.-W., 2014. formation of Fe3O4 nanocrystals in a porous carbon foam for lithium-ion battery
Modified composites based on mesostructured iron oxyhydroxide and synthetic anodes. J. Mater. Chem. 21, 17325–17330.
minerals: A potential material for the treatment of various toxic heavy metals Zhang, L., Hong, S., He, J., Gan, F., Ho, Y.S., 2011. Adsorption characteristic studies of
and its toxicity. J. Hazard. Mater. 267, 161–168. phosphorus onto laterite. Desalin. Water Treat. 25, 98–105.
Deng, S., Ting, Y.P., 2005. Fungal biomass with grafted poly (acrylic acid) for Zhang, F., Xu, D., Wang, Y., Argyle, M.D., Fan, M., 2015. CO2 gasification of Powder
enhancement of Cu (II) and Cd (II) biosorption. Langmuir 21 (13), 5940–5948. River Basin coal catalyzed by a cost-effective and environmentally friendly iron
Dinu, M.V., Dragan, E.S., 2010. Evaluation of Cu2+, Co2+ and Ni2+ ions removal from catalyst. Appl. Energy 145, 295–305.
aqueous solution using a novel chitosan/clinoptilolite composite: Kinetics and Zhao, X., Xia, D., Zheng, K., 2012. Fe3O4/Fe/carbon composite and its application as
isotherms. Chem. Eng. J. 160 (1), 157–163. anode material for lithium-ion batteries. ACS Appl. Mater. Interfaces 4, 1350–
Foo, K.Y., Hameed, B.H., 2010. Insights into the modeling of adsorption isotherm 1356.
systems. Chem. Eng. J. 156 (1), 2–10. Zhu, X., Song, T., Lv, Z., Ji, G., 2016. High-efficiency and low-cost a-Fe2O3
Guo, L., Ye, P., Wang, J., Fu, F., Wu, Z., 2015. Three-dimensional Fe3O4-graphene nanoparticles-coated volcanic rock for Cd(II) removal from wastewater.
macroscopic composites for arsenic and arsenate removal. J. Hazard. Mater. Process Saf. Environ. Protect. 104, 373–381.
298, 28–35.

Please cite this article in press as: Yoon, K., et al. Fabrication of engineered biochar from paper mill sludge and its application into removal of arsenic and
cadmium in acidic water. Bioresour. Technol. (2017), http://dx.doi.org/10.1016/j.biortech.2017.07.020

You might also like