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

Journal of Molecular Liquids 266 (2018) 373–380

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Adsorption and thermodynamic mechanisms of manganese removal


from aqueous media by biowaste-derived biochars
Muhammad Idrees a,b,⁎, Saima Batool a,b, Hidayat Ullah b, Qaiser Hussain c,d,⁎⁎, Mohammad I. Al-Wabel c,
Mahtab Ahmad e, Amjad Hussain f, Muhammad Riaz g, Yong Sik Ok h, Jie Kong a
a
MOE Key Laboratory of Space Applied Physics and Chemistry, Shaanxi Key Laboratory of Macromolecular Science and Technology, School of Natural & Applied Sciences, Northwestern
Polytechnical University, Xi'an 710072, PR China
b
Institute of Chemical Sciences, Gomal University, Dera Ismail Khan 29220, Pakistan
c
Soil Sciences Department, College of Food & Agricultural Sciences, King Saud University, P.O. Box 2460, Riyadh 11451, Saudi Arabia
d
Department of Soil Science & Soil Water Conservation, PMAS Arid Agriculture University, Rawalpindi 46300, Pakistan
e
Department of Environmental Sciences, Faculty of Biological Sciences, Quaid-i-Azam University, Islamabad 45320, Pakistan
f
Research and Development Department, Higher Education Commission, Islamabad, Pakistan
g
Department of Environmental Sciences & Engineering, Government College University, Faisalabad, Pakistan
h
Korea Biochar Research Center, Divison of Environmental Science and Ecological Engineering, Korea University, Seoul, Republic of Korea

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

Article history: In the present investigation, poultry manure and farmyard manure-derived biochars were applied as cost-
Received 25 April 2018 effective adsorbents for manganese (Mn) removal from aqueous media. Effects of functional parameters such
Received in revised form 9 June 2018 as solution pH, contact time, temperature and concentration on the Mn removal efficiency of biochars were eval-
Accepted 12 June 2018
uated. Poultry manure-derived biochar exhibited greater adsorption efficiency than farmyard manure-derived
Available online 13 June 2018
biochar due to its porosity and surface functionality. The maximum adsorption was achieved at pH 6, tempera-
Keywords:
ture 298 K and contact time of 3 h. The adsorption isotherm data was well fitted to the Freundlich model indicat-
Sorption ing multilayer adsorption onto heterogeneous surfaces of the biochars. Thermodynamics calculations affirmed
Manure that Mn adsorption onto biochars was spontaneous and exothermic process governed by hydrogen bonding
Char type of electrostatic interaction. Post-adsorption spectroscopic analysis of Mn-loaded biochars evidenced the
Water treatment binding of Mn with active surface functionalities of biochars.
Equilibrium © 2018 Elsevier B.V. All rights reserved.

1. Introduction groundwater and surface water, is one of the main sources of Mn toxic-
ity in humans. The world health organization (WHO) has established a
Heavy metals enter into the environment through natural and an- safe drinking water concentration of 0.05 mg L−1 for Mn. Relatively
thropogenic sources. The anthropogenic emission of heavy metals into less attention has been given to Mn removal from water because of its
the environment is deteriorating largely the soil and water resources importance as an essential micronutrient. However, it is reported that
[1]. Manganese (Mn) is the second most abundant among these heavy continuous administration of Mn may increase the neurotoxicity risk
metals in nature. It has been considered as an essential micronutrient in humans [3].
for plants and animals' growth. However, at high concentrations, Mn Various water treatment techniques are in practice, particularly fo-
becomes toxic to humans causing the Parkinson illness, pulmonary cusing on heavy metals removal, such as adsorption, coagulation-
track disorder, bronchitis, and hindering the intellectual development flocculation, chemical precipitation, ion-exchange, electrocoagulation,
and normal growth of infants [2]. Drinking water, including etc. [4]. Among these techniques, adsorption has been shown to be the
simplest, economical and efficient technique of removing heavy metals
from aqueous media [5]. Different types of materials including carbona-
⁎ Correspondence to: M. Idrees, MOE Key Laboratory of Space Applied Physics and ceous materials (activated carbon, charcoal, carbon-nano-tubes), clay
Chemistry, Shaanxi Key Laboratory of Macromolecular Science and Technology, School minerals, synthetic polymers, activated alumina, and silica gel have
of Natural & Applied Sciences, Northwestern Polytechnical University, Xi'an 710072, been successfully applied as adsorbents for metals removal [6–8]. In-
People's Republic of China. creasing global demand of efficient adsorbents has resulted in their
⁎⁎ Correspondence to: Q. Hussain, Soil Sciences Department, College of Food &
Agricultural Sciences, King Saud University, P.O. Box 2460, Riyadh 11451, Saudi Arabia.
cost elevation; hence there is a surge of interest in developing the alter-
E-mail addresses: m.idrees8223@gmail.com (M. Idrees), qaiseruaf@gmail.com native low-cost adsorbents capable of removing heavy metals from
(Q. Hussain). aqueous solution [9].

https://doi.org/10.1016/j.molliq.2018.06.049
0167-7322/© 2018 Elsevier B.V. All rights reserved.
374 M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380

Biochar, a carbon-rich material obtained from thermal decompo- combusting the oven-dried samples at 700 °C for 2 h in open top ce-
sition of organic waste, has recently gained popularity as a universal ramic crucibles.
and environmental green sorbent for organic and inorganic contam-
inants removal from water. Biochar has also shown its potential for 2.3. Sorption equilibrium experiment
mitigating climate change, carbon sequestration, soil quality im-
provement, enhanced crop production, and environmental manage- To test the capability of manure-derived BCs as sorbents, batch type
ment of waste recycling [10]. Specifically, during the last years, sorption equilibrium experiments were carried out with aqueous Mn
biochar has proved its tendency of removing heavy metals such as solutions. A sorbent dose of 10 g L−1 was applied to each Mn solution
cadmium and copper from water [11–13]. However, to our knowl- of initial concentration ranging from 2 to 50 mg L−1. The pH of solutions
edge, no study has yet been conducted on removal of Mn from was adjusted between 2 and 6 with 1 N NaOH and H2SO4 solutions. The
water using biochar. High surface area, the presence of numerous mixtures were equilibrated on an incubator shaker (MSC-100, China)
surface functional groups, and microporous structure are the for 24 h at 180 rpm. To determine the effect of temperature on Mn sorp-
distinguishing characteristics of biochar making it an efficient adsor- tion onto BCs, the sorption experiments were conducted at three differ-
bent for metals removal from water. Additionally, biochar is ~6 times ent temperatures of 298, 308 and 318 K at a fixed pH of 4. After
cheaper than the most widely used activated carbon, thus fascinating equilibrium time of 24 h, the suspensions were filtered through
it as a low-cost sorbent too [14]. Whatman 42 filter paper, and the solution was analyzed for Mn concen-
Manure waste is widely used as organic soil amendment. However, tration on an atomic absorption spectrometer (AAS, Perkin-Elmer 214,
there are several environmental issues related to manure waste such Norwalk, CT). All the batch experiments were carried out in triplicate.
as odor, methane production, and eutrophication of groundwater. Pyro-
lyzing the manure waste into biochar could provide an alternative 2.4. Sorption isotherms
waste management technology, thereby reducing the associated envi-
ronmental issues of manure application to soil. The present study The retention of Mn on manure-derived BCs was investigated by
aimed at (1) conversion of poultry and farmyard manure into biochar, drawing the sorption isotherms between Mn concentration at equilib-
(2) assessment of manure-derived biochar for Mn removal from rium and amount of Mn sorbed onto the BCs. The sorbed amount of
water, and (3) prediction of sorption phenomenon of Mn onto two dif- Mn was calculated using the following (Eq. (1)) [15]:
ferent biochars.
V
Qe ¼ ðC0 −Ce Þ ð1Þ
2. Materials and methods m

where Qe is the amount of Mn sorbed onto BCs (mg g−1) at equilibrium,


2.1. Chemical reagents
V is the volume of solution (L), C0 is the initial Mn concentration
(mg L−1), Ce is the Mn concentration at equilibrium (mg L−1), and m
The chemical reagents used were of analytical grade. The deionized
is the mass of sorbent (g). The isotherm data was then subjected to
water of 18 MΩ·cm resistivities was used for solutions preparation.
the most widely used empirical models of Freundlich, Langmuir, and
Manganese sulfate (MnSO4·H2O), sulfuric acid (H2SO4) and sodium hy-
Temkin, the linearized forms are presented in (Eqs. (2), (3) and (4), re-
droxide (NaOH) were purchased from Sigma. A stock solution of
spectively.
1000 mg Mn L−1 was prepared in deionized water, and further working
solutions of desired strength were prepared by diluting the stock solu- 1
tion in deionized water. ln Q e ¼ lnK F þ lnCe ð2Þ
n
 
2.2. Biochar production and characterization 1 1 1 1
¼ þ ð3Þ
Qe Qm KL Q m Ce
Poultry manure and farmyard manure were selected as feedstock for
the production of biochar (BC). Poultry manure and farmyard manure RT
Qe ¼ ln ðACe Þ ð4Þ
samples were collected from the poultry farm and livestock farm of b
the University campus, respectively. The manure samples were oven-
RT
dried at 60 °C. The oven-dried manure samples were packed in the ce- B¼ ð4aÞ
b
ramic container covered with a lid and then placed in a muffle furnace
at 450 °C for 5 h under oxygen-limited condition. After the pyrolysis, where KF is the Freundlich constant [(mg g−1)·(L g−1)n], n is a di-
containers were taken out of the furnace and allowed to cool at room mensionless constant, Qm is the Langmuir maximum adsorption ca-
temperature. Finally, the poultry manure-derived BC (PBC) and farm- pacity (mg g−1), KL is the Langmuir affinity constant (L mg−1), R is
yard manure-derived BC (FBC) were passed through the 0.18 mm the universal gas constant (8.314 J mol −1 ), T is the temperature
sieve to get the uniform particle size. (K), b is the Temkin isotherm constant, B is the heat of sorption
The BCs' surface structural morphology was analyzed by scanning (J mol−1 ), and A is the Temkin equilibrium binding constant
electron microscopy (SEM; INCAX-ACT, 58794; Oxford Instruments, (L g−1 ). The removal efficiency of each BC for Mn was calculated
China), while their surface functionality was identified by Fourier trans- using the following (Eq. (5)):
form infrared (FTIR) spectroscopy (PerkinElmer, USA). The thermal
 
transformation of the prepared BCs' was determined by a thermogravi- C0 −Ce
Removal efficiency ð%Þ ¼  100 ð5Þ
metric analyzer (TGA; NETZSCH STA449F3, Germany) with a heating C0
rate of 10 °C min−1 within 40–1000 °C in Argon. The surface electronic
structure of BCs' was checked by X-ray photoelectron spectroscopy. El-
emental (C, H, and N) analyses were performed using a CHN elemental 2.5. Sorption kinetics experiment
analyzer (Perkin Elmer, USA), while the surface area of BCs' was deter-
mined via N2 adsorption on a BET Surface Area Analyzer (Gemini VII To evaluate the sorption rate of Mn on manure-derived BCs, batch
2390 Series-Micromeritics, USA). The pH of BCs was determined using type sorption kinetics experiments were performed. The aqueous Mn
a pH meter (Mettler Toledo Delta 320) in a suspension of 1:10 BC/deion- solution of 50 mg L−1 initial concentration was mixed with each BC at
ized water after 1 h shaking. Ash contents of BCs were determined by a rate of 10 g L−1. The mixtures were shaken on an incubator shaker
M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380 375

565.21
(a) 2831.93 798.49 (b) 798.49
2955.66 1031.77
3399.66 2955.68 1031.77
1418.06
3377.92 1090.30

1432.27

FBC
BC-FYM
PBC
BC-PM

1585.28
1592.80
2510.86

4000 3500 3000 2500 2000 1500 1000 500 4000 3500 3000 2500 2000 1500 1000 500
-1 -1
Wavenumbers (cm ) Wavenumbers (cm )

(c) (d)

1429.43 3422.82
3436.37 1606.27
1075.75
1062.20

Mn-loaded
Mn FYM-BC
loaded FBC

Mn loaded PBC
Mn-loaded PM-BC

4000 3600 3200 2800 2400 2000 1600 1200 800 400 4000 3600 3200 2800 2400 2000 1600 1200 800 400
-1 -1
Wavenumbers (cm ) Wavenumbers (cm

Fig. 1. Fourier transform infrared (FTIR) spectra of (a) farmyard manure derived biochar (FBC), (b) poultry manure derived biochar (PBC), (c) Mn-loaded FBC and (d) Mn-loaded PBC.

at 180 rpm. Samples were withdrawn after 1, 2, 3, 4, 5, and 6 h, filtered experiments were carried out in triplicate. The amount of Mn sorbed
through Whatman 42 filter paper and analyzed for aqueous Mn concen- at different time intervals was calculated following the (Eq. (6)).
tration by using AAS. The kinetics sorption experiments were conducted
at two different temperatures of 308 and 318 K. All the batch V
Qt ¼ ðC0 −Ct Þ ð6Þ
m

Fig. 2. Scanning electron micrographs (SEM) of (a) farmyard manure-derived biochar (FBC) and (b) poultry manure-derived biochar (PBC).
376 M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380

where Qt is the amount of Mn sorbed (mg g−1) onto BCs at time t (h), the BCs spectra. The presence of aromatic C shows the stability of BC
and Ct is the remaining Mn concentration (mg L−1) in aqueous solution while aliphatic C may indicate the labile component of BC [17].
at time t. The V, m, and C0 have already been described in Section 2.4. The Fig. 2 shows the SEM images of FBC and PBC. A tessellated car-
bon surface is seen in crater-like structures. Pores and channels were
2.6. Thermodynamics calculations formed in both the BCs owing to the removal of volatiles during pyroly-
sis at 450 °C. Nevertheless, it should be noted that structural features in
To determine the effect of temperature on Mn sorption on manure- the heterogeneous range were observed in both these BCs.
derived BCs, Gibb's free energy change (ΔG°), enthalpy change (ΔH°), The TGA curves of FBC and PBC are presented in Fig. 3(c). The curves
and entropy change (ΔS°) were calculated. The apparent equilibrium indicate weight loss with increasing temperature. Greater weight loss
constant (K) is given by the following (Eq. (7)) [16]:

K ¼ KnF  M  1000  55:5 ð7Þ (a) C O


5
3x10
where M is the molecular weight of Mn (54.94 g mol−1), and 55.5 is the
molar concentration of water. The obtained value of K was then used
into the (Eq. (8)) to calculate the ΔG°.
5
PBC

Intensity/cps
2x10
0
ΔG ¼ −RT lnK ð8Þ Si
O
where R is the universal gas constant (8.314 J mol−1 K−1), and T is the C
temperature (K). The values of ΔH° and ΔS° were respectively obtained
5
1x10
from the slope and intercept of the plot between lnK and 1/T (van't Hoff FBC
plot) by considering the following (Eqs. (9) and (10)):

0
ΔGo ¼ ΔHo −TΔSo ð9Þ
  0 100 200 300 400 500 600 700 800 900
ΔHo 1 ΔSo
lnK ¼ − þ ð10Þ Binding Energy/eV
R T R 1.5x10
5

(b) O Mn-loaded PBC


C

5
2.7. Statistical analysis 1.2x10 Mn 2p

An average of three replicates from each batch experiment was 4


9.0x10
employed to plot sorption equilibrium and kinetics isotherms. Linear re-
Intensity/cps

MnO, Mn2O3
gression fittings were used for empirical and kinetic model equations.
O 640.56 eV
Slope and intercept values were used to calculate different models 6.0x10
4

parameters. Mn 2p
The standard errors of the mean were calculated for the experimen- C
tal data interpretation in all batch type sorption experiments shown in 3.0x10
4

the Table S1.


Mn-loaded FBC
3. Results and discussion 0.0
0 100 200 300 400 500 600 700 800 900
3.1. Biochar characteristics Binding Energy/eV

The FBC had high pH (8.2), total carbon (47.47%) and surface area FBC
(10.11 m2 g−1), and fewer ash contents (27.18%) than PBC. Relatively 100 PBC
low surface area of PBC (8.61 m2 g−1) compared to FBC was attributed
to its high ash contents (31.44%). In fact, high ash contents could fill 80
and block the micropores of biochar, resulting in a relatively low surface
area [12, 13].
Mass %

The FTIR spectral analysis predicted several functional groups on the 60


surfaces of manure-derived BCs (Fig. 1a&b). The wave numbers
3399.66 cm−1 and 3377.92 cm−1 for FBC and PBC, respectively, were
40
assigned to O\\H stretching vibrations, while the bands from 2955.68
to 2510.86 cm−1 indicated the C\\H aliphatic stretching vibration. The
band numbers (1418.06 cm−1, 1432.27 cm−1), (1090.30 cm−1– 20
1031.77 cm−1) and (798.49 cm−1–565.21 cm−1) in both the BCs (c)
could be attributed to CO−23 , C\\O and aromatic C\\H groups, respec-
0
tively. The bands at 1592.80 cm−1 in FBC and 1585.28 cm−1in PBC
200 400 600 800 1000
were assigned to C_O stretching of conjugated ketones and quinones. o
Temperture( C)
The bands at 1418.06 cm−1 and 1432 cm−1 for FBC and PBC, respec-
tively, represented the aromatic C_C stretching, while the band at
Fig. 3. X-ray photon spectra (XPS) of (a) farmyard manure-derived biochar (FBC) and
798.49 cm−1 in both BCs indicated out-of-plane deformation by aro- poultry manure-derived biochar (PBC), (b) Mn-loaded FBC and Mn-loaded PBC and
matic C\\H groups that might be caused by carbonates. The sharp (c) Thermogravimetric analysis (TGA) of farmyard manure-derived biochar (FBC) and
peaks at 1031.77 cm−1 were assigned to phosphate (PO3− 4 ) in both poultry manure-derived biochar (PBC).
M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380 377

(a) 100
(b) 100

80 80
Removal efficiency (%)

Removal efficiency (%)


60 60

40 40

20 20
FBC FBC
PBC PBC
0 0
2 4 6 298 308 315

pH Temperature (K)

(c) 100
(d) 100

80 80

Removal efficiency (%)


Removal efficiency (%)

60 60

40 40

20 FBC-298 20 PBC-298
FBC-308 PBC-308
FBC-318 PBC-318
0 0
2 4 8 15 30 50 2 4 8 15 30 50

Initial concentration (mg Mn L-1) Initial concentration (mg Mn L-1)

Fig. 4. Effects of (a) pH, (b) temperature and (c & d) metal initial concentration on Mn removal efficiency of manure-derived biochars.

was observed for PBC, which predicted its less stability than FBC. Gener- 3.2. Effects of pH, temperature and initial metal ion concentration on Mn re-
ally, poultry manure contains greater contents of organic matter than moval efficiency of BCs
farmyard manure causing its less thermal stability [18]. Specifically, a
sharp weight loss in FBC was observed at 750–800 °C corresponding The acidity of solution (pH) is one of the important parameters that
to the decomposition of CO−2 3 to CO2 [19], whereas no sharp weight affect the adsorbent surface charge, degree of ionization and speciation
loss was observed for PBC. These results are consistent with the FTIR re- of the metal ions in solution. Mn adsorption onto FBC and PBC was car-
sults showing a more intense band of CO−2 3 in FBC (1418 cm−1) than ried out at varying range of pH (2.0, 4.0 and 6.0), an adsorbent dose of
PBC (1432 cm−1) (Fig. 1a&b). 0.25 g, contact time of 24 h, 180 rpm shaking speed and adsorbate con-
The surface elemental composition of the BCs was analyzed by XPS. centration of 50 mg L−1. Results indicated that Mn removal efficiency
For both the BCs, it could be seen that the main peaks at binding ener- increased with increasing solution pH (Fig. 4a), regardless of BC type.
gies of 283.58 eV and 530.22 eV were attributed to the C and O atoms, High removal efficiency (N80%) at less acidic pH could be attributed to
respectively (Fig. 3a). Moreover, compared to FBC, the intensity of C low H+ ion concentration, providing less competition with Mn2+ for
and O peaks was more in case of PBC, indicating the relatively high pro- adsorption onto negatively charged surfaces of BCs [20]. In other
portion of these atoms on the surface of PBC. A short peak at a binding words, an excess of H+ ions at low pH of 2.0 and 4.0 surround the bind-
energy of 99.25 eV of Si appeared in PBC, while it was absent in FBC. ing sites of BCs, thus making sorption relatively unfavorable.

Table 1
Constant parameters and correlation coefficients of isothermal models for Mn sorption onto manure-derived biochars.

Adsorbents Temperature Freundlich Langmuir Temkin


(K)
KF N R2 KL Qm R2 B A R2
[(mg−1) (L mg−1) (mg g−1) (J mol−1) (L g−1)
(L g−1)n]

FBC 298 0.520 1.306 0.984 0.092 6.652 0.998 2942 9.634 0.922
308 0.258 1.501 0.975 0.292 1.341 0.975 4817 2.950 0.773
318 0.131 1.278 0.956 0.118 1.342 0.980 5068 0.675 0.686
PBC 298 1.103 1.863 0.809 1.056 2.403 0.479 3603 285.6 0.949
308 0.322 1.451 0.984 0.338 1.527 0.972 4166 4.520 0.805
318 0.167 1.390 0.982 0.057 2.842 0.979 5671 1.199 0.867
378 M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380

2.5 2.5
(a) (b)
1.5 1.5

0.5 0.5

e
-0.5 e
-0.5

-1.5 FBC-298
-1.5
LnQ LnQ
PBC-298
FBC-308 PBC-308
-2.5 FBC-318 -2.5 PBC-318
Linear (FBC-298) Linear (PBC-298)
-3.5 Linear (FBC-308) -3.5 Linear (PBC-308)
Linear (FBC-318) Linear (PBC-318)
-4.5 -4.5
-2.0 -1.0 0.0 1.0 2.0 3.0 4.0 -3.0 -1.0 1.0 3.0 5.0
LnCe LnCe

9.0
(c) (d) 10.0 PBC-298
PBC-308
PBC-318
7.0 8.0 Linear (PBC-298)
Linear (PBC-308)
Linear (PBC-318)
5.0 6.0
e e

1/Q

3.0 FBC-298
1/Q

4.0
FBC-308
FYM-BC-318
1.0 Linear (FBC-298) 2.0
Linear (FBC-308)
Linear (FBC-318)
-1.0 0.0
0.0 1.0 2.0 3.0 4.0 0.0 2.0 4.0 6.0 8.0 10.0
1/Ce 1/Ce

Fig. 5. Linear regression fittings of the Mn equilibrium sorption data to the Freundlich (a&b), and Langmuir (c&d) models at three different temperatures of 298, 308 and 318 K.

The Fig. 4b shows that higher temperature did not favor the Mn ad- model are presented in Table 1. Based on the correlation coefficient
sorption; rather it decreased the adsorption onto BCs. This could be due (R2) values, the Freundlich and Langmuir models well described the
to the reason that higher temperature leads to the higher kinetic energy equilibrium adsorption data (Fig. 5). Freundlich model was the ade-
of the Mn2+ ions, therefore weakening the forces of attraction between quate model fitting to the adsorption data at all the temperatures. The
the Mn2+ ions and the BCs. This also indicated that the process was exo- n value, representing the adsorption intensity, was N1 in both adsor-
thermic. The similar decreasing trend of Mn removal efficiency with in- bents and at all the temperatures, which indicated beneficial adsorption
creasing temperature was observed for both the BCs, and no significant of Mn [22]. The KF values were higher at 298 K (0.520 mg g−1 for FBC
difference in removal efficiencies was noted for FBC and PBC. and 1.103 mg g−1for PBC) and gradually decreased with increasing
Initial metal ion concentration gives an impelling cause to overcome temperature. This indicated that the relative adsorption capacity of
all metal transfer resistances in the solid and aqueous media and this PBC was higher than FBC at 298 K. The Langmuir predicted maximum
leads to a collision of higher probability between the active sites of adsorption capacity (Qm) was also high at 298 K for FBC
BCs and Mn. The adsorption sites at some point of time become (6.652 mg g−1) than other temperatures. Likewise, the heat of sorption
exhausted and reach a constant value where further adsorption from (B) values, calculated from Temkin model, was also low at low temper-
aqueous solution is not possible. Following these principles, Mn re- ature and gradually increased with increasing temperature in both the
moval efficiency decreased with increasing initial concentration, in BCs. This implied greater adsorption of Mn onto BCs at 298 K as the
this study (Fig. 4c&d). At an initial Mn concentration of 2 mg L−1, the re- heat of adsorption decreases with coverage [23].
moval efficiency ranged from 59.25% to 85.5% for FBC and 59.5% to 91.5% Overall, the Mn adsorption data followed the Freundlich isotherm
for PBC at different temperatures, while at 50 mg L−1 of Mn initial con- model favorably at all the temperatures, indicating heterogeneity of
centration the removal efficiencies were 47.82% to 69.95% and 40.26% to the BCs and multilayer adsorption system. The maximum adsorption
70.58% for FBC and PBC, respectively. Temperature showed a significant capacities of BCs in this study were compared with other carbonaceous
effect on Mn removal efficiency at different initial metal ion concentra- adsorbents (Table 2). The adsorption capacities of BCs reported in this
tions (Fig. 4c&d). Low temperature (298 K) favored the Mn sorption study are comparable with other adsorbents [24]. This makes the BCs
onto both the BCs. Specifically, at 298 K and very low initial concentra-
tion of 2 mg Mn L−1, the Mn removal efficiency was 85.5% and 91.5% for
FBC and PBC, respectively, while at 318 K and high initial concentration
Table 2
of 50 mg Mn L−1, the Mn removal efficiency decreased to 47.82% and Comparison of adsorption capacity of Mn with other adsorbents.
40.26% for FBC and PBC, respectively. The decrease in Mn removal effi-
ciency with increasing temperature may be due to the weakening of Adsorbents qm (mg g−1) References

electrostatic forces between BC active sites and metal ion causing de- Pithacelobium dulce carbon 0.415 [30]
sorption from the interphase to the solution [21]. Granular activated carbon 6.94 [24]
Steam activated carbon 9.52 [31]
Activated carbon modified by nitric acid 10.00 [31]
3.3. Adsorption isotherms Activated carbon modified by persulfate 6.66 [31]
Activated carbon modified by iron oxide 14.49 [24]
The equilibrium adsorption isotherm data were subjected to linear Activated carbon immobilized by tannic acid 1.73 [32]
forms of the Freundlich, Langmuir and Temkin models. The calculated Farmyard manure derived biochar 6.652 This study
Poultry manure derived biochar 2.842 This study
constant parameters and correlation coefficients of each isotherm
M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380 379

(a) 3 observed at 308 K by both the BCs (FBC and PBC). However, at 318 K,
PBC performed slightly better than FBC in Mn sorption, particularly
2.5 after 3 h of contact time. Rapid adsorption within 1 h of contact time in-
dicated the fast rate of reaction between BC active surfaces and Mn ions,
2 which is owed to the high availability of free reactive sites on BC sur-
Qt (mg g-1)

faces [25]. The PBC showed significantly high Mn sorption (2.460 ±


1.5 0.006 mg g−1 at 308 K and 2.146 ± 0.002 mg g−1 at 318 K) than FBC
(2.304 ± 0.001 mg g−1 at 308 K and 2.018 ± 0.001 mg g−1 at 318 K)
FBC-308
1 after one hour of contact time. The relatively high reaction rate of PBC
FBC-318 could be related to its high O contents (as indicated by XPS data in
PBC-308 Fig. 3a), which may have resulted in the deposition of Mn as MnO2
0.5
PBC-318
onto the surface of PBC [26].
Achieving the sorption equilibrium in short contact time indicates
0
the efficiency of the material. In wastewater treatment facilities, short
0 1 2 3 4 5 6 7
contact time is favored because less operational time will enhance the
t (h) process efficiency and minimize the operational cost of metal removal
[27]. Therefore, the BCs used in this study could serve as low-cost adsor-
(b) 16 bents for Mn removal from water. The fast initial sorption rate also sug-
gests the high reactivity between sorbent and sorbate, which
15 consequently determines the interaction between metal ions and active
surface groups of the adsorbent material. Moreover, it implies that
chemical adsorption could be dominant than physical adsorption taking
14 less time of reaction [28].
LnK

13 3.5. Thermodynamic studies

FYM-BC The results of calculated thermodynamic parameters for Mn adsorp-


12 PM-BC tion onto BCs are shown in Fig. 6(b) and Table 3. Negative values of ΔG°
Linear (FYM-BC) and ΔH° indicated that the adsorption process is spontaneous and exo-
Linear (PM-BC) thermic in nature, respectively. Likewise, the negative ΔS° values indi-
11
cated that the adsorption process is favorable for both BCs without
0.00312 0.0032 0.00328 0.00336 0.00344
causing a structural change at the solid-liquid interphase [29]. A direct
1/T (K) proportion was observed between ΔG° and temperature, demonstrat-
ing favorable adsorption at low temperature [33]. Mechanistic assump-
Fig. 6. (a) Effect of contact time on sorption of Mn onto manure-derived biochars tions can be made from the ΔG° values. If ΔG° is 4–10 kJ mol−1, the
(adsorbent dose 10 g L−1, temperature 308 and 318 K), (b) Van't Hoff plot between LnK dominant interaction is van der Waals force; for 2–40 kJ mol−1, hydro-
and 1/T for Mn sorption onto manure-derived biochars. gen bonding is dominant; for 2–29 kJ mol−1, dipole force of interaction
is operable; for N60 kJ mol−1, the chemical bonding is the main force of
interaction [34]. In this study, the ΔG° values range from −32.61 to
derived from manure as cost-effective adsorbents since no activation or
−37.44 kJ mol−1, predicting that hydrogen bonding could be the dom-
modification steps were involved, unlike other activated carbons.
inant interactive force between Mn ions and BCs.
3.4. Effect of contact time
3.6. Post-adsorption analyses
Contact time is another important factor affecting the removal of
metal ions in aqueous solution. The effect of contact time on Mn adsorp- Changes in the structural surface chemistry of BCs after Mn adsorp-
tion is shown in Fig. 6(a). The results showed that in the first hour Mn tion were determined by FTIR and XPS spectroscopic analyses to vali-
adsorption was very rapid, and after which the removal rate of Mn date the adsorption phenomenon. The FTIR spectra of Mn loaded BCs
slowed down and attained equilibrium as adsorption time increases were distinguishable from the original BC spectra (Fig. 1c&d). The
up to 6 h. In general, maximum Mn adsorption was achieved within \\OH band at 3399.66 cm−1 in FBC shifted towards 3436 cm−1 in
first 3 h. Temperature showed a negative effect on Mn adsorption, sim- Mn-loaded FBC indicating the binding of \\OH with Mn. Similar band
ilar to as observed in adsorption isotherm experiments. Low tempera- shifting was also observed in PBC where the\\OH band became sharp
ture (308 K) was more favorable for Mn adsorption onto BCs at all the owing to bonding with Mn. The band associated with CO2– 3 at
contact time intervals. No significant difference in sorption of Mn was 1418 cm−1 and 1432 cm−1 also shifted to 1429 cm−1 and 1606 cm−1
in FBC and PBC spectra, respectively, assuming due to Mn-carbonate for-
−1
Table 3 mation. The PO3−4 band at 1032 cm in both the BCs became less in-
−1
Thermodynamic parameters of Mn sorption onto manure-derived biochars. tense, and shifted to 1076 cm and 1062 cm−1 in FBC and PBC,
Adsorbents Temperature Thermodynamic parameters respectively, which evidenced the formation of stable Mn-phosphate.
(K) XPS spectra of Mn loaded BCs are shown in Fig. 3(b). Compared to
ΔG° ΔH° ΔS°
(kJ mol−1) (kJ mol−1) (J mol−1 K−1) the original XPS spectra of BCs, a new peak at 640.56 eV was observed
for Mn 2p region, which shows different binding energies when com-
FBC 298 −34.88 −68.92 −115.0
308 −33.03 bined with oxygen. Metallic Mn shows 638.7 eV binding energy and
318 −32.61 when combined with oxygen the binding energy is shifted up to
PBC 298 −37.44 −105.6 −229.8 640.56 eV depending on the oxidation state of oxygen. It can be noted
308 −34.02 that in Mn-loaded BCs, the main peak appearing at 640.56 eV confirmed
318 −32.90
the formation of Mn2O3 [6].
380 M. Idrees et al. / Journal of Molecular Liquids 266 (2018) 373–380

The post adsorption spectral analyses of Mn-loaded BCs predicted [8] W.F. Zhao, Y.S. Tang, J. Xi, J. Kong, Functionalized graphene sheets with poly (ionic
liquid) s and high adsorption capacity of anionic dyes, Appl. Surf. Sci. 326 (2015)
the formation of Mn compounds probably with carbonates and phos- 276–284.
phates. These results are consistent with the adsorption kinetics data [9] T.A. Kurniawan, G.Y. Chan, W.H. Lo, S. Babel, Physico–chemical treatment tech-
fitting to the pseudo-second-order model, prevailing chemisorption as niques for wastewater laden with heavy metals, Chem. Eng. J. 118 (2006) 83–98.
[10] J. Lehmann, S. Joseph, Biochar for environmental management: an introduction, Bio-
the main phenomenon of Mn adsorption onto BCs. char for Environmental Management Science and Technology, Earthscans, UK 2009,
pp. 1–12.
4. Conclusions [11] S. Batool, M. Idrees, Q. Hussain, J. Kong, Adsorption of copper (II) by using derived-
farmyard and poultry manure biochars: efficiency and mechanism, Chem. Phys. Lett.
689 (2017) 190–198.
The present study evaluated the poultry manure- and farmyard [12] M. Idrees, S. Batool, T. Kalsoom, S. Yasmeen, A. Kalsoom, S. Raina, Q. Zhuang, J. Kong,
manure-derived BCs as viable, and cost-effective adsorbents for the re- Animal manure-derived biochars produced via fast pyrolysis for the removal of di-
valent copper from aqueous media, J. Environ. Manag. 213 (2018) 109–118.
moval of Mn from aqueous media. The results revealed that the sorption
[13] M. Idrees, S. Batool, T. Kalsoom, S. Raina, H.M.A. Sharif, S. Yasmeen, Biosynthesis of
was dependent on different functional parameters such as pH, temper- silver nanoparticles using Sida acuta extract for antimicrobial actions and corrosion
ature, contact time and concentration. The maximum adsorption was inhibition potential, Environ. Technol. (2018) 1–8.
achieved at pH 6, temperature 298 K and contact time of 3 h. The ad- [14] G. Aad, T. Abajyan, B. Abbott, V.J. Abdallah, S.A. Khalek, O. Abdinov, H. Abramowicz,
Electron reconstruction and identification efficiency measurements with the ATLAS
sorption isotherms were well fitted to the Freundlich model. The ther- detector using the 2011 LHC proton–proton collision data, Eur. Phys. J. B (2014)
modynamic parameters from the experimental model presented that 1–38 (C 74.7).
adsorption process was exothermic and spontaneous, involving hydro- [15] G. Limousin, J.P. Gaudet, L. Charlet, S. Szenknect, V. Barthes, M. Krimissa, Sorption
isotherms: a review on physical bases, modeling and measurement, Appl. Geochem.
gen bond type interaction between Mn ions and BC surfaces. Post ad- 22 (2007) 249–275.
sorption spectroscopic analyses predicted changes in the surface [16] H. Zhou, Q. Chen, G. Li, S. Luo, T.B. Song, H.S. Duan, Y. Yang, Interface engineering of
structural chemistry of the BCs as a result of Mn bonding with O- highly efficient perovskite solar cells, Science 345 (2014) 542–546.
[17] T.J. Purakayastha, S. Kumari, H. Pathak, Characterisation, stability, and microbial ef-
containing functional groups. The manure-derived BCs, therefore, fects of four biochars produced from crop residues, Geoderma 239 (2015) 293.
could potentially be applied as efficient adsorbents for Mn removal [18] J. Kong, T. Schmalz, G. Motz, A.H.E. Müller, Novel hyperbranched ferrocene-
from aqueous solutions. containing poly(boro)carbosilanes synthesized via a convenient “A2 + B3” ap-
proach, Macromolecules 44 (2011) 1280–1291.
[19] J.E. Lim, M. Ahmad, V.S.S. Usman, S.S.W. Lee, T. Jeon, S.E. Oh, J.E. Yang, Y.S. Ok, Effects
Acknowledgment of natural and calcined poultry waste on Cd, Pb and As mobility in contaminated
soil, Environ. Earth Sci. 69 (2013) 11–20.
[20] M. Ahmad, S.S. Lee, S.E. Lee, M.I. Al-Wabel, D.C. Tsang, Y.S. Ok, Biochar-induced
The authors extend their appreciation to the Deanship of Scientific
changes in soil properties affected immobilization/mobilization of metals/metal-
Research at King Saud University for funding this work through research loids in contaminated soils, J. Soils Sediments 17 (2017) 717–730.
group No (RG-1439-043). The authors are thankful to Analytical and [21] K. Saltali, A. Sari, M. Aydin, Removal of ammonium ion from aqueous solution by
Testing Center of NPU, Northwestern Polytechnical University, Xi'an natural Turkish (Yıldızeli) zeolite for environmental quality, J. Hazard. Mater. 141
(2007) 258–263.
710072 P.R. China and Laboratory of Soil Science & Soil Water Conserva- [22] T.K. Naiya, P. Chowdhury, A.K. Bhattacharya, S.K. Das, Saw dust and neem bark as
tion, PMAS Arid Agriculture University, Rawalpindi, Pakistan for provid- low-cost natural biosorbent for adsorptive removal of Zn (II) and Cd (II) ions from
ing research opportunities, analysis and characterization of the aqueous solutions, Chem. Eng. J. 148 (2009) 68–79.
[23] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems,
sorbents. Chinese Scholarship Council (CSC No: 2015GXZ039, Chem. Eng. J. 156 (2010) 2–10.
2016GXYF32) financial support is gratefully acknowledged to pursuing [24] N.Y. Rachel, N.J. Nsami, K. Daouda, A.A. Victoire, T.M. Benadette, K.J. Mbadcam, Ad-
Ph.D. studies. sorption of manganese (II) ions from aqueous solutions onto Granular Activated
Carbon (GAC) and Modified Activated Carbon (MAC), Intern. J. Innovative Sci.,
Eng. & Technol. 2 (2015) 606–614.
Conflicts of interest [25] S. Anagho Gabche, D.R. TchuifonTchuifon, G. NcheNdifor-Angwafor, J. NdiNsami,
Nickel adsorption from aqueous solution onto kaolinite and metakaolinite: kinetic
and equilibrium studies, Int. J. Chem. 4 (2013) 1–235.
There are no conflicts to declare. [26] R. Contreras-Bustos, E. Manríquez-Reza, J. Jiménez-Becerril, M. Jiménez-Reyes, Syn-
thesis of MnO2 on activated carbon and its potential application in the adsorption of
Appendix A. Supplementary data As(V) and Pb(II) in aqueous solutions, Acta Chim. Slov. 64 (2017) 438–448.
[27] M. Shafiq, A.A. Alazba, M.T. Amin, Removal of heavy metals from wastewater using
date palm as a biosorbent: A comparative review, Sains Malaysiana 47 (2018)
Supplementary data to this article can be found online at https://doi. 35–49.
org/10.1016/j.molliq.2018.06.049. [28] J. Li, J. Hu, G. Sheng, G. Zhao, Q. Huang, Effect of pH, ionic strength, foreign ions and
temperature on the adsorption of Cu(II) from aqueous solution to GMZ bentonite,
Colloids Surf. A Physicochem. Eng. Asp. 349 (2009) 195–201.
References [29] M.P. Tavlieva, S.D. Geneiva, V.G. Georgieva, L.T. Vlaev, Thermodynamics and kinetics
of the removal of manganese (II) ions from aqueous solutions by white rice husk
[1] M. Idrees, S. Batool, Q. Hussain, H. Ullah, M.I. Al-Wabel, M. Ahmad, J. Kong, High- ash. J, J. Mol. Liq. 211 (2015) 938–947.
efficiency remediation of cadmium (Cd2+) from aqueous solution using poultry [30] K.A. Emmanuel, A.V. Rao, Adsorption of Mn (II) from aqueous solutions using
manure–and farmyard manure–derived biochars, Sep. Sci. Technol. 51 (2016) pithacelobium dulce carbon, Rasayan J. Chem. 1 (2008) 840–852.
2307–2317. [31] M.A. Akl, A.M. Yousef, S. AbdElnasser, Removal of iron and manganese in water sam-
[2] J.J. Anguille, G.M. Ona Mbega, T. Makani, J. Ketcha Mbadcam, Adsorption of manga- ples using activated carbon derived from local agro-residues, Br. Chem. Eng. Process.
nese (II) ions from aqueous solution on to volcanic ash and geopolymer based vol- Technol. 4 (2013) 154.
canic ash, Intern. J. Basic Appl. Chem. Sci. 3 (2013) 7–18. [32] A. Üçer, A. Uyanik, Ş.F. Aygün, Adsorption of Cu (II), Cd (II), Zn (II), Mn (II) and Fe
[3] J. Crossgrove, W. Zheng, NMR in biomedicine, manganese toxicity upon overexpo- (III) ions by tannic acid immobilised activated carbon, Sep. Purif. Technol. 47
sure, NMR Biomed. 17 (2004) 544–553. (2006) 113–118.
[4] F. Fu, Q. Wang, Removal of heavy metal ions from wastewaters: a review, J. Environ. [33] Y. Han, A.A. Boateng, P.X. Qi, et al., Heavy metal and phenol adsorptive properties of
Manag. 92 (2011) 407–418. biochars from pyrolyzed switchgrass and woody biomass in correlation with surface
[5] M. Emadi, E. Shams, M.K. Amini, Removal of zinc from aqueous solutions by magne- properties, J. Environ. Manage. 118 (2013) 196–204.
tite silica core-shell nanoparticles, J. Chemother. 2013 (2012) 1–10. [34] F. Huang, B.R. Leonard, X. Wu, Resistance of sugarcane borer to Bacillus thuringiensis
[6] F. Chen, W. Zhao, J. Zhang, J. Kong, Magnetic two-dimensional molecularly Cry1Ab toxin, Entomol. Exp. Appl. 124 (2007) 117–123.
imprinted materials for recognition and separation of proteins, Phys. Chem. Chem.
Phys. 18 (2016) 718–725.
[7] L.L. Meng, X.F. Zhang, Y.S. Tang, K.H. Su, J. Kong, Hierarchically porous silicon–
carbon–nitrogen hybrid materials towards highly efficient and selective adsorption
of organic dyes, Sci. Report. 5 (2015) 7910.

You might also like