You are on page 1of 11

Ecological Engineering 95 (2016) 101–111

Contents lists available at ScienceDirect

Ecological Engineering
journal homepage: www.elsevier.com/locate/ecoleng

Removal of nitrate from aqueous solution by modified sugarcane


bagasse biochar
Laleh Divband Hafshejani a,∗ , Abdolrahim Hooshmand a , Abd Ali Naseri a ,
Amir Soltani Mohammadi a , Fariborz Abbasi b , Amit Bhatnagar c,∗
a
Department of Irrigation and Drainage Engineering, Faculty of Water Sciences Engineering, Shahid Chamran University of Ahvaz, Khuzestan, Iran
b
Agricultural Engineering Research Institute (AERI), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran
c
Department of Environmental and Biological Sciences, University of Eastern Finland, P.O. Box 1627, FI-70211 Kuopio, Finland

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

Article history: In the present study, chemically modified biochar (developed from sugarcane bagasse) was used for
Received 17 March 2016 the nitrate removal from aqueous solution. The physico-chemical properties of modified biochar such as
Received in revised form 9 May 2016 morphology, surface functional groups, elemental composition, cation exchange capacity, anion exchange
Accepted 15 June 2016
capacity and surface area were analyzed. The effects of various operational parameters such as solution
pH, adsorbent dosage, contact time, initial concentration of nitrate, co-existing anions and temperature
Keywords:
were examined on nitrate adsorption by modified biochar. The experimental data were fitted to dif-
Adsorption
ferent adsorption kinetic models (pseudo-first-order, pseudo-second-order, intraparticle diffusion and
Modification
Biochar
Avrami models) and adsorption isotherms models (Langmuir, Freundlich, Sips and Dubinin–Raduskovich
Nitrate removal models). The obtained results showed that the maximum percentage of nitrate adsorption attained at
Sugarcane bagasse equilibrium pH 4.64, after 60 min of contact time and with an adsorbent dose of 2 g L−1 . In competing
anions experiments, carbonate and chloride ions have shown maximum and minimum influence on
the adsorption of nitrate by modified biochar sugarcane bagasse. Pseudo-second-order kinetic model
and Langmuir isotherm model showed the best fit to the experimental adsorption data. The maximum
adsorption capacity of modified biochar for nitrate removal was found to be 28.21 mg g−1 . The values of
H◦ , G◦ and S◦ indicated that the nature of adsorption was endothermic, spontaneous and feasible.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction sis, ion exchange, biological denitrification, chemical denitrification


and adsorption (Bhatnagar and Sillanpää, 2011). However, some
Contamination of water resources by nitrate has become an disadvantages in these methods have been reported in the liter-
important problem globally in recent decades. The wide use of ature, for example, ion exchange requires disposal of waste brine
fertilizers in agriculture is one of the major sources of nitrate pol- and renewal of costly ion exchange resins, the electrodes are eas-
lution. Nitrate ion is highly soluble in water and cannot bind well ily inactive in electrochemical reduction, reverse osmosis has high
with the soil particles and as a result, nitrate ions can easily enter operational cost, catalytic reduction could produce the hazardous
into the groundwater and surface water (Bhatnagar et al., 2010). by-products such as nitrite and ammonia and biological processes
Increasing concentration of nitrate in drinking water can cause are in need of addition of carbon source and disposal of large
harmful effect on human health especially in infant, such as blue- amounts of biomass waste (Hu et al., 2015; Mukherjee and De,
baby syndrome (Bhatnagar et al., 2010; Moradzadeh et al., 2014). 2014). Among several methods, adsorption has been suggested as
The U.S. Environmental protection Agency (U.S. EPA) has set the one of the most effective methods (Olgun et al., 2013; Xu et al.,
allowable concentration for nitrate as 10 mg L−1 in drinking water. 2013). The adsorption process may be even more encouraging
Different processes have been applied for the removal of nitrate when natural or agro-industrial wastes are utilized as adsorbents
from water and wastewater, such as reverse osmosis, electrodialy- because these are renewable, low cost, highly available, and in
many cases, their disposal pose a serious environmental problem.
In addition, these materials can be converted into high value-
added adsorbents e.g., biochar or activated carbons, making them
∗ Corresponding authors.
a cleaner source and a highly effective material for environmen-
E-mail addresses: mdivband@gmail.com (L. Divband Hafshejani),
amit.bhatnagar@uef.fi (A. Bhatnagar). tal applications (removal of pollutants from air, soil and water)

http://dx.doi.org/10.1016/j.ecoleng.2016.06.035
0925-8574/© 2016 Elsevier B.V. All rights reserved.
102 L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111

(Mohamad Nor et al., 2013). Numerous materials have been used 2. Materials and methods
for the adsorption of nitrate from water such as sugarcane bagasse,
rice hull, zeolite, chitosan, clays, slag and fly ash (Bhatnagar and 2.1. Preparation of modified biochar
Sillanpää, 2011).
In recent years, the use of biochar in many environmental appli- The sugarcane bagasse was collected from Amir Kabir sugar fac-
cations has gained wide attention (Cheng et al., 2016; Li et al., tory of Khuzestan in Iran. Sugarcane bagasse was chosen for this
2016; Miller-Robbie et al., 2015; Mohammadi et al., 2016; Peng study because it is produced in large quantities as a waste in sugar
et al., 2016; Trakal et al., 2016; Zhang and Wang, 2016). Agricultural industries. Sugarcane bagasse was air dried and was converted to
wastes have been preferred for the production of biochar because small size with electric mills. Air dried material was heated in
these are available abundantly, free of cost and are non-toxic. a carbonization furnace at five different pyrolysis temperatures
Use of agricultural wastes as feedstocks for biochar production (200–600 ◦ C), with a fixed residence time of 4 h. Biochar yield was
can improve the waste management and protect the environment calculated by comparing the weight of obtained biochar to the
(Ding et al., 2016; Tan et al., 2015). Biochar is produced from the weight of sugarcane bagasse (dry basis) used for pyrolysis. Oxidis-
biomass pyrolysis under limited or in absence of oxygen conditions able organic carbon content (OC) was determined by the method of
(Chintala et al., 2013). Due to its porous structure, biochar has been potassium dichromate oxidation. Loss on ignition (LOI) of resultant
used in many studies as an impressive adsorbent to remove differ- biochar was obtained by heating of produced biochar in a muffle
ent pollutants from aqueous solution. The production of biochar, furnace at 750 ◦ C for 6 h (ASTM method, D-1762-84) (Masto et al.,
in comparison with activated carbon, is fast, cheaper and simple 2013). Stable organic matter (SOM) was determined by (Eq. (1))
(Ahmad et al., 2012; Tan et al., 2015). Furthermore, some biochars (Masto et al., 2013).
have shown better adsorption capacity than commercial activated
carbons for the removal of toxic pollutants (Tan et al., 2015). Some SOM = LOI − (OC × 1.724) (1)
researchers have reported that biochar after adsorption may con-
where 1.724 is the factor for converting organic carbon to organic
tain a lot of nutrients such as, nitrate, ammonium and phosphate.
matter. Stable organic matter yield index (SOMYI) was calculated
Therefore, it can also be used as fertilizer for increasing the soil
by the following equation:
fertility (Yao et al., 2013; Yao et al., 2011). Also, biochar is rich in
 BY 
carbon and after adsorption, it can be applied to soil for carbon
SOMYI = × SOM (2)
sequestration (Fornes et al., 2015; Xiang et al., 2015). 100
It has been reported by various researchers that adsorption
In this study, chemical modification method was selected to
capacity of unmodified adsorbents is low for different pollutants
improve the adsorption capacity of biochar according to the
as compared to the modified adsorbents (Bhatnagar et al., 2013;
method of Xu et al. (2010). Ten grams of biochar was added
Chintala et al., 2013; Xu et al., 2010). Various physical or chemical
in 6 ml of epichlorohydrin (Merck, Germany) and 5 ml of N, N-
modification methods have been found to improve the perfor-
dimethylformamide (Merck, Germany) was further added to it. The
mance of adsorbents (Bhatnagar and Sillanpää, 2011; Chintala et al.,
mixture was keep at 85 ◦ C. After 60 min, 2 ml of ethylenediamine
2013). Also, results of a previous study showed that unmodified
(Merck, Germany) was combined with it and the solution was agi-
adsorbents were converted into anion exchangers after reaction
tated for 45 min at 85 ◦ C. After this time, 5 ml of 40% trimethylamine
with epichlorohydrin and dimethylamine in the presence of N,N-
(Merck, Germany) was added and the suspension was shaken for
dimethylformamide (DMF) and pyridine as catalyst (Orlando et al.,
120 min at 85 ◦ C. For elimination of residual chemicals, the solid
2002). The results of previous studies have also shown that adsorp-
part was washed with 500 ml of distilled water and dried at 60 ◦ C
tion rate of modified adsorbents is higher than unmodified ones
for 12 h (Xu et al., 2010).
(Iakovleva et al., 2015; Rangabhashiyam and Selvaraju, 2015).
In the present work, the potential of modified sugarcane bagasse
biochar has been examined for the removal of nitrate from aque- 2.2. Characterization of the adsorbent
ous solution under batch conditions. One of the positive aspects of
the study is the possibility of using a residue of agribusiness, gen- The morphology of modified biochar was investigated using
erated in large quantities worldwide, with little or no use, for the a scanning electron microscopy (SEM, Leo 1455 VP model, made
removal of nitrate from aqueous solution. The aim of this study in Germany). The functional groups in unmodified and modified
was to assess the parameters that influence the nitrate adsorp- biochar were analyzed with a Fourier transform infrared spec-
tion process using modified biochar as an adsorbent, obtained troscopy (FTIR, Spectrum GX, and Perkin-Elmer). The concentration
from sugar cane bagasse. Physico-chemical properties of modi- of different elements (C, H, N, S, O) in modified biochar was deter-
fied biochar were determined using different analytical techniques mined by the CHNSO analyzer (vario ELIII- elementar- made in
such as scanning electron microscope (SEM), Fourier transform Germany). Surface area was obtained using methylene blue method
infrared (FTIR) and CHNSO analyzer. The effects of different param- (Kaewprasit et al., 1998). The cation exchange capacity (CEC) and
eters such as solution pH, adsorbent dosage, contact time, initial anion exchange capacity (ACE) were measured by the method
concentration of nitrate, co-existing (competing) anions and tem- reported elsewhere (Chintala et al., 2013).
perature on adsorption capacity of modified biochar toward nitrate
removal were investigated. The equilibrium and kinetic data of 2.2.1. Determination of pHpzc of modified biochar
nitrate adsorption were evaluated using adsorption kinetic models For the determination of pH value at the point of zero charge of
(pseudo-first-order, pseudo-second-order, intraparticle diffusion modified biochar, 50 ml 0.1 M NaCl solution was added into a series
and Avrami) and the adsorption isotherm models (Langmuir, Fre- of 100 ml capped glass tubes. The initial pH (pHi ) of solutions was
undlich, Sips and Dubinin–Raduskovich). adjusted between 2 and 11 by the addition of 0.1 M HCl or 0.1 M
NaOH. After pH adjustment, 0.05 g modified biochar was added to
the solutions and suspension was shaken with a speed of 120 rpm
at 22 ± 1 ◦ C. After 24 h, the solutions were filtered by 0.45 ␮m cel-
lulose acetate membrane filters (Sartorius AG) and final pH (pHf ) of
the filtrates were measured. The difference between the initial and
final pH values (pH = pHi − pHf ) was plotted versus pHi . The pH at
L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111 103

which pH = 0 is as pH value at the point of zero charge (Hafshejani q2e k2 t


qt = (5)
et al., 2015; Tangsir et al., 2016). 1 + qe k2 t

qt = Kp t ⁄2 + I
1
2.3. Batch adsorption studies (6)
 nAV 
qt = qe 1 − e(−kAV t) (7)
2.3.1. Effect of solution pH
The pH of the solution is an important factor in adsorption where qt and qe and are the amounts of nitrate ions adsorbed on the
studies as it not only affects surface charges and dissociation of modified biochar (mg g−1 ) at time t and equilibrium time, respec-
functional groups of the adsorbent, but also chemical speciation tively; k1 is the rate constant of the pseudo-first-order (min−1 ), k2
and diffusion rate of solute (Rahmani et al., 2010). For determi- is the rate constant of the pseudo-second-order (g mg−1 min−1 ); Kp
nation of the pH effect on the nitrate adsorption by modified is the intraparticle diffusion rate constant (mg g−1 min1/2 ) and I
sugarcane bagasse biochar, initial pH of nitrate solutions with (mg g−1 ) is the intercept of intraparticle diffusion model and kAV
concentration of 50 mg L−1 were adjusted at different pH values (min−1 ) and nAV are the exponential of the Avrami model.
(2–11) by using 0.1 M HCl and 0.1 M NaOH. Then desired amount of
modified sugarcane bagasse biochar (1 g L−1 ) was added to 50 ml 2.3.4. Effect of co-existing (competing) anions
solution. The mixture was shaken at 120 rpm at 22 ± 1 ◦ C. After The adsorption capacity of modified biochar for the nitrate
24 h, solid phase was separated by using centrifugation (4000 rpm) adsorption in the presence of co-existing (competing) anions (e.g.,
for 20 min and then solutions were filtered by 0.45 ␮m cellu- phosphate, carbonate, sulfate and chloride), which are commonly
lose acetate membrane filters (Sartorius AG). Nitrate concentration present in real ground water, was also investigated. Experiments
in the filtered solution was analyzed immediately by UV spec- were conducted by adding varying concentrations of co-existing
trophotometer (model Hach, DR5000) at wavelength of 220 nm (UV anions (10, 50 and 100 and 200 mg L−1 ) in nitrate solution (fixed
Screening Method 10049). The amount of nitrate adsorbed per unit nitrate concentration of 50 mg L−1 ), optimum pH, and an optimum
weight of adsorbent (qe in mg g−1 ) was calculated as follows: adsorbent dosage of 2 g L−1 . The mixtures were shaken with a speed
(Ci − Ce )V of 120 rpm at 22 ± 1 ◦ C for 60 min. Then solid phase was separated
qe = (3) from solution by filtration using 0.45 ␮m cellulose acetate mem-
m
brane filters (Sartorius AG) and residual nitrate concentration in
where Ci and Ce are initial and equilibrium concentrations of nitrate the filtered solution was analyzed. The amount of nitrate adsorbed
in solution (mg L−1 ), V is the volume of solution (L) and m is mass per unit weight of adsorbent (qe in mg g−1 ) was calculated by Eq.
of the modified biochar (g). Reproducibility of the measurements (3).
was determined in triplicate and the average values are reported.
2.3.5. Adsorption equilibrium studies
2.3.2. Effect of adsorbent dosage Equilibrium adsorption experiments were conducted by tak-
The effect of adsorbent dosage on nitrate adsorption was stud- ing a series of solutions with initial nitrate concentrations of
ied by adjusting the solution pH at optimum (pH 4.64) with an 1–100 mg L−1 , optimum adsorbent dosage of 2 g L−1 , at 22 ± 1 ◦ C
initial nitrate concentration of 50 mg L−1 and by adding the vary- with an optimum solution pH (equilibrium pH 4.64). The mix-
ing adsorbent dosage ranging from 1 to 40 g L−1 . The mixtures tures were shaken at 120 rpm and after equilibration, samples were
were shaken with a speed of 120 rpm using a shaking assembly at filtered by 0.45 ␮m cellulose acetate membrane filters (Sartorius
22 ± 1 ◦ C for 24 h. Then solid phase was separated by using centrifu- AG) and were analyzed by spectrophotometer for the determi-
gation (4000 rpm) for 20 min and then solutions were filtered by nation of residual nitrate concentration. Equilibrium data are the
0.45 ␮m cellulose acetate membrane filters (Sartorius AG). Resid- important requirements for the successful modeling of adsorption
ual nitrate concentration in the filtered solution was determined systems. There are different theoretical and empirical relation-
by spectrophotometer. The amount of nitrate adsorbed per unit ships for the modeling of adsorption process (Rahmani et al.,
weight of adsorbent (qe in mg g−1 ) was calculated by Eq. (3). 2010). In this study, Langmuir (Eq. (8)) (Langmuir, 1916), Fre-
undlich (Eq. (9)) (Freundlich, 1909), Sips (Eq. (10)) (Sips, 1948) and
2.3.3. Adsorption kinetic studies Dubinin–Radushkevich (Eq. (11)) (McEnaney, 1987) models were
Kinetic adsorption experiments were conducted by a series of applied to describe the adsorption process.
solutions with initial nitrate concentration of 50 mg L−1 , optimum (bCe qm )
equilibrium pH of 4.64, optimum adsorbent dosage of 2 g L−1 , at qe = (8)
(1 + bCe )
22 ± 1 ◦ C by varying the contact time between 5–180 min with
1
agitation speed of 120 rpm. After completing predetermined time qe = KF Cen (9)
interval, solids phase was removed from the solutions by filtration
1/n
using 0.45 ␮m cellulose acetate membrane filters (Sartorius AG) (qm Ks Ce )
qe = (10)
and nitrate concentrations were analyzed by spectrophotometer. 1/n
(1 + Ks Ce )
The amount of nitrate adsorbed per unit weight of adsorbent (qe    2 
in mg g−1 ) was calculated by Eq. (3). Rate of adsorption can be pre- 1
qe = qm exp −ˇ RTln 1 + (11)
dicted by adsorption kinetic parameters. This factor is important ce
and useful in predicting the adsorption mechanism (Rahmani et al.,
2010). The dynamics of the nitrate adsorption process by modi- where qe (mg g−1 ) is the equilibrium adsorption capacity of modi-
fied biochar was evaluated with the pseudo-first-order (Eq. (4)) fied biochar, Ce (mg L−1 ) is the equilibrium concentration of nitrate
(Lagergren, 1898), pseudo-second-order (Eq. (5)) (Ho and McKay, in solution, qm (mg g−1 ) is the maximum adsorption capacity of
1999), intraparticle diffusion (Eq. (6)) (Weber and Morris, 1963) modified biochar; K (mg g−1 ) (L mg−1 )1/n and n (g L−1 ) are the
and Avrami (Eq. (7)) (Avrami, 1939) kinetic models. The non-linear Freundlich equilibrium constant and exponent, respectively; b
forms of these models are given below: (L mg−1 ) in Langmuir model is related to the energy of adsorption
(L mg−1 ), Ks in Sips model is the affinity constant for adsorption; R
qt = qe (1 − e−k1 t ) (4) is the universal gas constant (8.314 kJ K−1 mol−1 ), T is the absolute
104 L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111

temperature (K), ␤ is activity coefficient related to mean adsorp- Table 1


Effect of different temperatures on the properties of sugarcane bagasse biochar.
tion energy per mole. Modeling calculations were conducted using
Microsoft Office Excel 2007. Isotherm parameters were determined Temperature (◦ C) 200 300 400 500 600
by minimizing the Sum of the Squares of the Errors (ERRSQ) func- Biochar yield (%) 29.3 26.1 17.1 15.2 12
tion across the concentration range studied: Oxidizable organic carbon content (%) 41.4 34 29.5 28.9 21
Loss on ignition (%) 89.1 86 77.3 74.8 59.3

n
Stable organic matters (%) 17.7 27.4 26.5 24.9 23
2
ERRSQ = (Oi − Pi ) (12) Stable organic matter yield index 5.2 7.2 4.6 3.8 2.8
i=1

where: Table 2
n: number of observations. Elemental contents of sugarcane bagasse biochar produced at 300 ◦ C after modifi-
Oi : ith value of the observed measurement. cation (percentage).
Pi : ith value of the predicted measurement. Elemental C H N S O

Sugarcane bagasse biochar 56.2 3.8 4.1 0.8 21.1


2.3.6. Thermodynamic studies Sugarcane bagasse 43.1 6.2 0.3 0.2 39.4
The effect of temperature on nitrate adsorption by modified
biochar was studied at different temperatures (viz., 10 ◦ C, room
temperature (22 ◦ C) and 30 ◦ C). Thermodynamics parameters were high amounts by decomposing the biomass (Mašek et al., 2013).
calculated using the following equations: Oxidizable organic carbon content (OC) and loss on ignition (LOI)
decreased with increasing pyrolysis temperature, which is due to
G0 = −RTlnK (13)
increase in ash content in the biochar at high pyrolysis tempera-
K1 H 0
1 1

tures. A comparison between the stable organic matters (SOM) at
ln = − (14)
K2 R T1 T2 different temperatures showed that SOM increased with increase
in pyrolysis temperature up to 300 ◦ C, thereafter it decreased which
G0 = H 0 − TS 0 (15)
might be due to the increase in the ash content. The results of the
where G◦
is standard free energy change R is the(kJ mol−1 ), present study showed that maximum stable organic matter yield
universal gas constant (8.314 J K−1 mol−1 ), T is the absolute tem- index (SOMYI) occurred at 300 ◦ C (Table 1). Biochar is a carbon rich

perature (in Kelvin), H is standard enthalpy (kJ mol−1 ), S◦ is material and carbon content can significantly influence the adsorp-
standard entropy (J mol K−1 ), and K is the equilibrium constant,
−1 tion capacity. Thus, the stable organic matter yield index of biochar
related to the Langmuir constant ‘b’ (Eq. (16)). is an important parameter (Chintala et al., 2013; Kumar et al., 2013).
Therefore, biochar produced at 300 ◦ C was selected for the chemi-
K = b × 55.5 (16)
cal modification and nitrate adsorption experiments. The elemental
where the value 55.5 corresponds to the molar concentration of the analyses (CHNSO) of modified sugarcane bagasse biochar are given
solvent (water, in this study). in Table 2.
It was found that carbon was present in higher amounts as
2.4. Error analysis compared to other elements which might be due to the pres-
ence of carbon containing functional groups such as carboxylic.
Two Error functions were used to evaluate the goodness of fit The biochar has been graded into three classes according to car-
of the applied adsorption models (kinetics and isotherms) to the bon content. The class 1 contains 60% carbon or more, class 2
experimental data. possesses between 30 and 60% carbon in biochar and class 3
1. The coefficient of determination (R2 ) contains between 10% and 30% carbon in biochar (Mohan et al.,
 n 2 2014). Carbon content (56.2%) in modified sugarcane bagasse
(Oi − Oave ) . (Pi − Pave ) biochar shows that produced biochar is in the upper range of
2
R = n i=1
n (17) class 2. The surface area of modified biochar was determined as
i=1 (Oi
− Oave )2 . i=1 (Pi
− Pave )2
41.67 m2 g−1 using methylene blue method. Also, cation exchange
2. Average relative error (ARE) capacity (CEC) and anion exchange capacity (ACE) were obtained
as 12.90 cmol kg−1 and 10.21 cmol kg−1 , respectively. SEM pho-
100

n O − P 
i i tographs of modified biochar before and after nitrate adsorption
ARE = (18)
n Oi are presented in Fig. 1a–b.
i=1
SEM image of biochar (before nitrate adsorption) shows
where Oi and Pi are ith of the observed and predicted value, respec- that modified biochar has a porous structure (honeycomb like
tively. Oave and Pave are average of the observed and predicted value structure), which indicates that modified biochar consists a car-
and n is number of observations. bonaceous skeleton (Ghani et al., 2013). Some tiny fine pores are
also visible in the SEM image of modified biochar (Fig. 1a). After
3. Results and discussion nitrate adsorption, carbonaceous skeleton was broken and pores
were blocked. Fourier transform infrared spectroscopy is often used
3.1. Characterization of the adsorbent to examine the surface functional groups available on the surface of
adsorbents. Fig. 2 shows FTIR spectrums of modified biochar before
The properties of biochars produced under different tempera- and after nitrate adsorption. The comparison between results of
tures in this study are listed in Table 1. FTIR spectrum before and after nitrate adsorption shows that wave
It can be seen that the biochar yield decreased as the pyroly- numbers and the intensity of some peaks were shifted or sub-
sis temperature increased. Similar results have also been found by stantially lowered than those before adsorption suggesting the
other researchers where a decrease in biochar yield was observed involvement of some functional groups of biochar in nitrate adsorp-
at increased pyrolysis temperature (Kumar et al., 2013; Masto et al., tion. An absorption band around 3262.48 cm−1 corresponding to
2013). It can be explained due to the fact that high pyrolysis tem- N H groups is shifted to 3331.79 cm−1 after adsorption of nitrate.
perature increases the release of water and volatile compounds in This group (N H) has positive charge, which may increase the
L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111 105

Fig. 3. pHpzc of modified sugarcane bagasse biochar.

ability of nitrate adsorption with negative charge. The peaks at


1743.52 cm−1 may be attributed to C O stretching vibration, thus
confirms the presence of carbon in biochar structure (Ghani et al.,
2013). The peak at 1505.92 cm−1 in the spectrum shows C C groups
(Kim et al., 2014). The peak at 1306.48 cm−1 may be attributed to
C N groups with positive charge, therefore, it might involve in the
adsorption of nitrate and has shifted to 1270.84 cm−1 after adsorp-
tion (Luo et al., 2014). Peaks at 856.48 cm−1 should correspond to
C H stretching (Chen et al., 2015).

3.2. Point of zero charge (pHpzc)

pHPZC represents the pH at which the net surface charge of the


adsorbent is zero and cation exchange capacity and anion exchange
capacity are equal on adsorbent’s surface (Chintala et al., 2013).
Results of this study (Fig. 3) shows that the pHpzc of biochar is about
5.35. When pH of the solution is lower than the pHpzc, net sur-
face charge of modified biochar is positive due to the adsorption
of excess H+ . In this situation, modified biochar has high ability for
the adsorption of anionic species. Also, when pH of the solution is
higher than the pHpzc, the net surface charge on surface modified
biochar is negative due to desorption of H+ . In this situation, surface
Fig. 1. SEM images of modified sugarcane bagasse biochar. (a) Before nitrate adsorp- of biochar becomes suitable for desorption of cations.
tion. (b) After nitrate adsorption.

3.3. Effect of solution pH

Fig. 4 shows the effect of solution pH (initial and final pH) on


adsorption capacity of modified sugarcane bagasse biochar towards
nitrate in the solution with initial concentration of 50 mg L−1 ,
adsorbent dosage of 1 g L−1 and contact time of 24 h. It was observed
that at initial pH 2, nitrate adsorption capacity is 28.3 mg g−1
and with change of initial pH from 3 to 11, final (equilibrium)
pH changes from 4.64 to 9.11 and the adsorption capacity of
modified sugarcane bagasse biochar decreased from 30.7 mg g−1
to 10.3 mg g−1 . Thus, maximum adsorption capacity for nitrate
(30.7 mg g−1 ) was obtained at initial pH 3 (final pH 4.64). It is clear
that by increasing the pH, the adsorption of nitrate decreased. It can
be due to higher competition between nitrate and hydroxide ions
for same sites on adsorbent’s surface (Cengeloglu et al., 2006). Also
at low pH, surface of modified biochar has positive charge due to
protonation reactions, which increased the electrostatic attraction
between modified biochar surface and negatively charged nitrate
ions (Chintala et al., 2013; Demiral and Gündüzoğlu, 2010; Olgun
Fig. 2. FTIR spectra of modified sugarcane bagasse biochar before and after nitrate et al., 2013). Similar results have been reported by other researchers
adsorption. who reported that adsorption of nitrate by other adsorbents was
106 L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111

Fig. 4. Effect of solution pH on adsorption capacity of modified sugarcane bagasse


biochar for nitrate adsorption (adsorbent dosage = 1 g L−1 , initial nitrate concentra- Fig. 6. Effect of contact time on adsorption capacity of modified sugarcane bagasse
tion = 50 mg L−1 , contact time = 24 h, temperature = 22 ± 1 ◦ C, initial pH = 2-11). biochar for nitrate adsorption (adsorbent dosage = 2 g L−1 , initial nitrate concentra-
tion = 50 mg L−1 , contact time = 5–180 min, temperature = 22 ± 1 ◦ C, initial pH = 3).

adsorption efficiency increased upto 90% and adsorption capac-


ity of modified sugarcane bagasse biochar was decreased. Further
increase in adsorbent dosage beyond 20 g L−1 did not significantly
effect the nitrate adsorption capacity. Similar results were also
reported by other researchers where nitrate adsorption efficiency
from aqueous solution was found to increase up to the optimum
dosage and with further increase in adsorbent dose, the adsorp-
tion efficiency remained constant (Katal et al., 2012; Öztürk and
Bektaş, 2004). Therefore, based on the results of this study, adsor-
bent dosage of 2 g L−1 was selected for further experiments. With
adsorbent dosage of 2 g L−1 , there was a balance between adsorp-
tion capacity and adsorption efficiency and adsorption capacity of
modified sugarcane bagasse biochar towards nitrate was found to
be 18.40 mg g−1 .

3.5. Effect of contact time


Fig. 5. Effect of adsorbent dosage on adsorption capacity of modified sugar-
cane bagasse biochar for nitrate adsorption (adsorbent dosage = 1–40 g L−1 , initial
The effect of contact time on nitrate adsorption capacity by mod-
nitrate concentration = 50 mg L−1 , contact time = 24 h, temperature = 22 ± 1 ◦ C, initial ified biochar was conducted with the initial nitrate concentration of
pH = 3). 50 mg L−1 , adsorbent dosage 2 g L−1 and at optimum pH. The results
are shown in Fig. 6. It was found that the adsorption capacity of
high at acidic pH and decreased at higher pH values (Chintala et al., modified sugarcane bagasse biochar for removing nitrate was rapid
2013; Ganesan et al., 2013; Olgun et al., 2013). Results of this study in the beginning stages of contact time. After 15 min, the rate of
(Fig. 3) show that the pHpzc of biochar is about 5.35. Thus, it is rea- adsorption decreases with the time and adsorption process reached
sonable that maximum adsorption of nitrate onto biochar occurred to equilibrium within 60 min. This phenomenon might be due to
at final pH 4.64. At pH < 5.35, surface of biochar is positively charged the presence of greater number of active sites on modified biochar
and due to coulombic attraction, biochar could adsorb negative for the adsorption of nitrate ions during the initial stages. Similar
charge species like nitrate. results were also reported by other researchers where the nitrate
adsorption was found to increase with increase in contact time
3.4. Effect of adsorbent dosage and after reaching to equilibrium time, adsorption remained con-
stant (Ganesan et al., 2013; Olgun et al., 2013; Öztürk and Bektaş,
Adsorbent dosage is an important factor that determines the 2004). According to the obtained results, a contact time of 60 min
optimum adsorbent dose which is required to remove a definite was chosen for further experiments. The results show that the pro-
amount of pollutants from the solution (Du et al., 2012). In gen- cess of nitrate adsorption by modified biochar in this study is faster
eral, increase in adsorbent dosage increases adsorption of adsorbate than other adsorbents reported/studied by other researchers which
from the solution which is due to the availability of more active sites might be due to modification of biochar in this work (Bhatnagar
and increase in surface area at higher dosage (Hu et al., 2015; Öztürk et al., 2008; Chatterjee et al., 2009).
and Bektaş, 2004). Fig. 5 shows the results of nitrate adsorption
efficiency and adsorption capacity of modified sugarcane bagasse 3.6. Effect of initial nitrate concentration
biochar versus different adsorbent dosage in the solution with
initial concentration of 50 mg L−1 , optimum pH (initial pH 3 and The effect of initial nitrate concentrations on adsorption capac-
final pH 4.64) and contact time of 24 h. It was evident that with ity of modified sugarcane bagasse biochar was studied by varying
the increase in adsorbent dosage from 1 to 40 g L−1 , the nitrate different initial nitrate concentrations in the range of 1–100 mg L−1 ,
L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111 107

Fig. 7. Effect of initial nitrate concentration on adsorption capacity of modi- Fig. 8. Effect of solution temperature on adsorption capacity of modified sugar-
fied sugarcane bagasse biochar for nitrate adsorption (adsorbent dosage = 2 g L−1 , cane bagasse biochar for nitrate adsorption (adsorbent dosage = 2 g L−1 , initial nitrate
initial nitrate concentration = 1–100 mg L−1 , contact time = 60 min, tempera- concentration = 50 mg L−1 , contact time = 60 min, temperature = 10, 22 and 30 ± 1 ◦ C,
ture = 22 ± 1 ◦ C, initial pH = 3). initial pH = 3).

adsorbent dosage of 2 g L−1 , contact time of 60 min and at opti-


mum pH. The results are presented in Fig. 7. It was evident from the
results that with the increase in initial nitrate concentrations from
1 to 80 mg L−1 , adsorption capacity of modified biochar increased
from 0.5 mg g−1 to 23.7 mg g−1 and further increase in initial nitrate
concentration (more than 80 mg L−1 ) did not effect on adsorption
capacity of modified biochar towards nitrate. This phenomenon can
be attributed to the number of available active adsorption sites on
modified biochar for the nitrate adsorption (Milmile et al., 2011;
Olgun et al., 2013). Similar results were also reported by other
researchers where adsorption capacity of adsorbents increased
with increasing initial concentration of nitrate (Milmile et al., 2011;
Olgun et al., 2013).

3.7. Effect of temperature

The adsorption experiments were conducted at different tem-


peratures viz. 10 ◦ C, room temperature (22 ◦ C) and 30 ◦ C to
investigate the effect of temperature on the adsorption capac- Fig. 9. Effect of co-existing anions on adsorption capacity of modified sugarcane
ity of modified biochar for the removal of nitrate with an initial bagasse biochar for nitrate adsorption (adsorbent dosage = 2 g L−1 , initial nitrate con-
nitrate concentration of 50 mg L−1 , contact time of 60 min, with centration = 50 mg L−1 , initial concentrations of co-existing anions = 0–200 mg L−1 ,
contact time = 60 min, temperature = 22 ± 1 ◦ C, initial pH = 3).
optimum pH and an optimum adsorbent dosage of 2 g L−1 . It is
observed in Fig. 8 that with increase in temperature from 10 ◦ C
to 30 ◦ C, nitrate adsorption capacity increases from 10.9 mg g−1 presence of co-existing (phosphate, carbonate, sulfate and chlo-
to 20.1 mg g−1 , which might be due to the increased interaction ride) in the water. The carbonate and chloride ions have shown
between nitrate ions and active site on surface of modified biochar. maximum and minimum effects on the adsorption of nitrate. It
Also, these results (increasing adsorption capacity with increasing could be explained by the fact that in the multi element solution,
temperature) indicate that the adsorption process is endothermic the electrostatic interaction of co-existing anions with adsorption
in nature. Similar results were also reported by other researchers sites of modified biochar was much stronger than nitrate ions
where the nitrate adsorption capacity was found to increase with species. It has been reported that multivalent anion with higher
an increase in temperature (Ganesan et al., 2013). charge density was adsorbed more readily than monovalent anion
(Rezaei Kalantary et al., 2015). Similar results have been reported
3.8. Effect of co-existing (competing) anions by other researchers where multivalent and mono charge anions
have shown more and less adsorption trend, respectively (Rezaei
The effect of co-existing (competing) anions on the nitrate Kalantary et al., 2015; Yang et al., 2015).
removal by modified biochar was studied by varying initial con-
centrations of co-existing anions (10, 50, 100 and 200 mg L−1 ) in a 3.9. Adsorption kinetics
nitrate solution with a constant nitrate concentration of 50 mg L−1 ,
at optimum pH, an optimum adsorbent dosage of 2 g L−1 , contact Study of adsorption kinetics is one of the important parameters
time of 60 min at room temperature (22 ◦ C). The results are shown because it provides good information about the reaction path-
in Fig. 9. It was found that the adsorption capacity of modified sug- way and mechanism of the adsorption process (Xu et al., 2013).
arcane bagasse biochar towards nitrate was reduced due to the In this study, adsorption kinetics of nitrate ions onto modified
108 L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111

Table 3
Kinetic parameters of nitrate adsorption by modified sugarcane bagasse biochar.

Pseudo-first-order qe(exp) (mg g−1 ) k1 (min−1 ) qe(cal) (mg g−1 ) ARE R2


18.44 0.14 17.90 4.61 0.99

Pseudo-second-order qe(exp) (mg g−1 ) k2 (g mg−1 min−1 ) qe(cal) (mg g−1 ) ARE R2
18.44 0.01 19.07 1.15 0.99

Intraparticle diffusion I (mg g−1 ) kP (mg g−1 min−0.5 ) – ARE R2


7.68 1.05 – 12.01 0.81

Avrami qe(exp) (mg g−1 ) kAV (min−1 ) nav qe(cal) (mg g−1 ) ARE R2
18.44 0.09 1.45 17.90 4.76 0.99

Table 4
Isotherm parameters of nitrate adsorption by modified sugarcane bagasse biochar.

Freundlich KF (mg g−1 ) (L mg−1 )1/n n (g L−1 ) – ARE R2


5.82 2.59 – 62.35 0.98

Langmuir b (L mg−1 ) qm (mg g−1 ) RL ARE R2


0.13 28.21 0.07 5.90 0.99

Sips KS (L mg−1 ) qm (mg g−1 ) n ARE R2


0.11 25.94 0.82 7.14 0.99

D–R ␤ qm (mg g−1 ) nav E (kJ mol−1 ) ARE R2


1.81 25.84 0.53 18.41 0.99

biochar of sugarcane bagasse was studied by using pseudo-first- of R2 and lower values of RAE (R2 = 0.99 and RAE = 1.15) has pro-
order, pseudo-second-order, intraparticle diffusion and Avrami vided a better fit as compared to the other models for adsorption of
models. Pseudo-first order model assumes that the sorption rate nitrate ions by modified biochar (Table 3). A good fit of experimen-
decreases linearly with the increase of adsorption capacity (Zhen tal data with the pseudo-second-order model would indicate the
et al., 2015). The pseudo-second order kinetic model assumes that chemical adsorption is the rate controlling mechanism (Zhen et al.,
the rate-limiting step is the interaction between two reagent par- 2015). Similar results have also been reported by other researchers
ticles and it is usually used to describe a chemical adsorption (Ho where pseudo-second-order kinetic model has shown good agree-
and McKay, 1999; Zhen et al., 2015). In the liquid–solid system, the ment with the experimental data for adsorption of nitrate ions
intraparticle diffusion model has been used to identify an adsorp- (Demiral and Gündüzoğlu, 2010; Olgun et al., 2013). In addition
tion controlled process by diffusion mechanism, where the rate of to that, qe(cal) value determined from pseudo-second-order model
adsorption process depends on the diffusion speed of adsorbate agreed very well with the experimental value (qe(exp) ).
towards adsorbent (Xu et al., 2013). Avrami model is the revision In this study, intraparticle diffusion model (qt vs. t1/2 plot)
of kinetic thermal decomposition modeling and Avrami exponen- showed three distinct linear portions (Fig. 10b). This would indicate
tial shows probable variations of the adsorption mechanism that the existence of more than one kinetic stage in adsorption process
occur throughout the adsorption process. The results of adsorption (Olgun et al., 2013). The overall rate of nitrate adsorption by intra-
kinetic models are shown in Table 3, Fig. 10a–b. The rapid equilib- particle diffusion model can be described by the following three
rium process (ca. 60 min) in Fig. 10a shows that nitrate interacts steps: (1) the first stage can be indicated to the diffusion of nitrate
with the outermost surface of modified sugarcane bagasse biochar ions to the external surface of the modified biochar (fluid trans-
layers. Pseudo-second-order kinetic model with the higher values port), (2) the second stage explains the gradual adsorption (film

Table 5
Comparison of the efficiencies of various adsorbents for nitrate removal from water.

Adsorbent qm (mg g−1 ) pH Contact time Reference

Untreated coconut granular activated carbon 1.7 5.5 2h Bhatnagar et al. (2008)
ZnCl2 treated coconut granular activated carbon 10.2 5.5 2h Bhatnagar et al. (2008)
Cross-linked and quaternized chinese reed 7.55 5.8 10 min Namasivayam and Höll (2005)
Chemically modified sugar beet bagasse 9.14–27.55 6.58 – Demiral and Gündüzoğlu (2010)
Commercial activated carbon 1.22 – 10 min Mishra and Patel (2009)
Wheat straw charcoal 1.10 – 10 min Mishra and Patel 2009
Mustard straw charcoal 1.30 – 10 min Mishra and Patel (2009)
Nano-alumina 4 3.1 60 min Bhatnagar et al. (2010)
Modified rice husk 55.55 7 90 min Katal et al. (2012)
Non-activated biochar of corn stover 11.77 4 – Chintala et al. (2013)
Chemically activated biochar of corn stover 13.33 4 – Chintala et al. (2013)
Non-activated biochar of pine wood chips 3.49 4 – Chintala et al. (2013)
Chemically activated biochar of pine wood chips 13.22 4 – Chintala et al. (2013)
Non-activated biochar of switchgrass 11.87 4 – Chintala et al. (2013)
Chemically activated biochar of switchgrass 15.19 4 – Chintala et al. (2013)
Zero-valent iron nanoparticles 20.95 original pH 60 min Wang et al. (2014)
Fe3 O4 nanoparticles 2.435 original pH 10 min Wang et al. (2014)
Poly-o-toluidine zirconium (IV) ethylenediamine 20.09 4.5 90 min Rahman and Khan (2015)
Granular chitosan-Fe3+ complex 8.35 5.9 ± 0.2 90 min Hu et al. (2015)
Sugarcane bagasse 2.13 3 – This study
Biochar of sugarcane bagasse 11.56 3 4h This study
Modified biochar of sugarcane bagasse 28.21 3 60 min This study
L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111 109

diffusion), (3) the third stage shows the final equilibrium due to
reduction of nitrate concentration in the solution and decrease of
active sites of modified biochar (surface diffusion). As can be seen
from Fig. 10b, the intercept of the plot does not pass through the
origin, which indicates that adsorption kinetics may be controlled
by film diffusion and intra-particle diffusion concurrently. Similar
results have also been reported by other researchers where linear
plots of intra-particle diffusion model do not pass through the ori-
gin for nitrate adsorption (Demiral and Gündüzoğlu, 2010; Olgun
et al., 2013).

3.10. Adsorption isotherms

Adsorption isotherms show the relationship between the


amount adsorbed by a unit weight of adsorbent distribution of
adsorbable solute between the liquid and solid phases at various
equilibrium concentrations and at constant temperature. The main
assumptions for Freundlich adsorption isotherm model are that the
surface containing the adsorption sites is heterogeneous, and each
site can hold several molecule in thickness (multilayer) and the
enthalpy of adsorption decrease with an increase of occupied sites
(Keränen et al., 2015). In other words, Freundlich model expresses
reversible adsorption and is not restricted to the formation of the
monolayer (Zhen et al., 2015). The Langmuir adsorption isotherm
model is based on the assumption that the surface containing
the adsorption sites is homogeneous, and each site can hold at
most one molecule in thickness (monolayer adsorption) (Keränen
et al., 2015). Dubinin–Raduskovich adsorption isotherm model
(D–R isotherm) is based on the assumption that the characteristic
adsorption curve is related to the porous structure of adsorbent (Ho
et al., 2002). This model can be used for determination of nature of
adsorption process (physical or chemical) (Bhatnagar et al., 2010;
Olgun et al., 2013). The Sips adsorption isotherm model is a com-
bination of Langmuir and Freundlich isotherms. At low adsorbate
concentrations, it changes to a Freundlich isotherm and at high
adsorbate concentrations, it reduces to a Langmuir isotherm (Ho
et al., 2002).
The results of Langmuir, Freundlich, Sips and
Dubinin–Raduskovich (D-R) isotherms for the adsorption of
Fig. 10. (a) Pseudo-first-order, pseudo-second-order and Avrami kinetic models. (b) nitrate by modified biochar are presented in Table 4 and Fig. 11.
Intraparticle diffusion model for nitrate adsorption by modified sugarcane bagasse In this study, Langmuir model with the higher values of coeffi-
biochar.
cient of determination (R2 ) (0.99) and lower values of average
relative error (ARE) (5.91) presented a better fit for the adsorption
of nitrate as compared to other isotherm models. Therefore, a
good fit of experimental data with the Langmuir model would
indicate monolayer adsorption of nitrate on homogenous surface
of modified biochar of sugarcane bagasse.
Also, it shows that the energy of adsorption for all sites on
surface of modified biochar of sugarcane bagasse is equal. Lang-
muir fitting has been found applicable for nitrate adsorption onto
various adsorbent by other researchers as well (Cengeloglu et al.,
2006; Demiral and Gündüzoğlu, 2010). The value of the maximum
adsorption capacity was found to be 28.21 mg g−1 as calculated by
the Langmuir isotherm model.
The value of separation factor or equilibrium parameter, RL , can
be used to estimate whether adsorption is favorable. It is calculated
from the Langmuir isotherm model parameter as follows:
1
RL = (19)
1 + bc0
According to Table 4, the RL value obtained is 0.07 (between 0
and 1) which indicates that adsorption isotherm is favorable in this
study. The n value in Freundlich model is 2.59 (between 1 and 10)
Fig. 11. Different isotherm models for nitrate adsorption by modified sugarcane which indicates favorable adsorption of nitrate by modified sug-
bagasse biochar. arcane bagasse biochar. The value of mean adsorption energy, E,
is very useful in predicting the type of adsorption and if the value
110 L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111

Table 6 References
Thermodynamic parameters for the adsorption of nitrate ions by modified sugarcane
bagasse biochar. Öztürk, N., Bektaş, T.E., 2004. Nitrate removal from aqueous solution by adsorption
◦ ◦ ◦ onto various materials. J. Hazard. Mater. 112, 155–162.
Temperature (◦ C) G (kJ mol−1 ) H (kJ mol−1 ) S (J mol−1 K−1 )
Ahmad, M., Lee, S.S., Dou, X., Mohan, D., Sung, J.K., Yang, J.E., Ok, Y.S., 2012. Effects
10 −32.00 54.50 293.09 of pyrolysis temperature on soybean stover-and peanut shell-derived biochar
22 properties and TCE adsorption in water. Bioresour. Technol. 118, 536–544.
Avrami, M., 1939. Kinetics of phase change: I general theory. J. Chem. Phys. 7,
1103–1112.
Bhatnagar, A., Sillanpää, M., 2011. A review of emerging adsorbents for nitrate
is from 1 to 8 kJ mol−1 , then the adsorption is physical in nature removal from water. Chem. Eng. J. 168, 493–504.
and if it is from 8 to 16 kJ mol−1 , then the adsorption is chemical in Bhatnagar, A., Ji, M., Choi, Y.H., Jung, W., Lee, S.H., Kim, S.J., Lee, G., Suk, H., Kim,
nature. It can be calculated from Dubinin–Raduskovich parameter H.S., Min, B., 2008. Removal of nitrate from water by adsorption onto zinc
chloride treated activated carbon. Sep. Sci. Technol. 43, 886–907.
as follows: (Bhatnagar et al., 2010; Olgun et al., 2013): Bhatnagar, A., Kumar, E., Sillanpää, M., 2010. Nitrate removal from water by
1 nano-alumina: characterization and sorption studies. Chem. Eng. J. 163,
E=  (20) 317–323.
2ˇ Bhatnagar, A., Hogland, W., Marques, M., Sillanpää, M., 2013. An overview of the
modification methods of activated carbon for its water treatment applications.
The E value obtained in present study is 0.53 kJ mol−1 , indicating Chem. Eng. J. 219, 499–511.
Cengeloglu, Y., Tor, A., Ersoz, M., Arslan, G., 2006. Removal of nitrate from aqueous
the adsorption of nitrate by modified biochar is physical in nature.
solution by using red mud. Sep. Purif. Technol. 51, 374–378.
Adsorption capacities of modified sugarcane bagasse biochar Chatterjee, S., Lee, D.S., Lee, M.W., Woo, S.H., 2009. Nitrate removal from aqueous
from the present study were compared with other adsorbents stud- solutions by cross-linked chitosan beads conditioned with sodium bisulfate. J.
ied by other researchers (Bhatnagar et al., 2008, 2010; Chatterjee Hazard. Mater. 166, 508–513.
Chen, L., Wang, X., Yang Lu Li, H.Q.D., Yang, Q., Chen, H., 2015. Study on pyrolysis
et al., 2009; Chintala et al., 2013; Hassan et al., 2010; Katal et al., behaviors of non-woody lignins with TG-FTIR and Py-GC/MS. J. Anal. Appl.
2012; Mishra and Patel, 2009; Wang et al., 2014) for nitrate removal Pyrolysis 113, 499–507.
(Table 5). The results presented in Table 5 suggest that modified Cheng, Q., Huang, Q., Khan, S., Liu, Y., Liao, Z., Li, G., Ok, Y.S., 2016. Adsorption of Cd
by peanut husks and peanut husk biochar from aqueous solutions. Ecol. Eng.
sugarcane bagasse biochar has comparable adsorption efficiency 87, 240–245.
for the nitrate removal from aqueous solution. Chintala, R., Mollinedo, J., Schumacher, T.E., Papiernik, S.K., Malo, D.D., Clay, D.E.,
Kumar, S., Gulbrandson, D.W., 2013. Nitrate sorption and desorption in
biochars from fast pyrolysis. Microporous Mesoporous Mater. 179, 250–257.
3.11. Thermodynamic parameters Demiral, H., Gündüzoğlu, G., 2010. Removal of nitrate from aqueous solutions by
activated carbon prepared from sugar beet bagasse. Bioresour. Technol. 101,
The knowledge of thermodynamic parameters is of fundamental 1675–1680.
Ding, Z., Hu, X., Wan, Y., Wang, S., Gao, B., 2016. Removal of lead, copper, cadmium,
importance to test the spontaneous occurrence of a given process
zinc, and nickel from aqueous solutions by alkali-modified biochar: batch and
as well as the feasibility of operation at a given temperature (Muthu column tests. J. Ind. Eng. Chem. 33, 239–245.
Prabhu and Meenakshi, 2014, 2015). In this study, adsorption Du, Y., Zhu, L., Shan, G., 2012. Removal of Cd2+ from contaminated water by
experiments were compared at two temperatures (10 ◦ C and 22 ◦ C) nano-sized aragonite mollusk shell and the competition of coexisting metal
ions. J. Colloid Interface Sci. 367, 378–382.
in order to obtain the information on thermodynamic nature of the Fornes, F., Belda, R.M., Lidón, A., 2015. Analysis of two biochars and one hydrochar
adsorption process. The results of thermodynamic parameters are from different feedstock: focus set on environmental, nutritional and
listed in Table 6. The negative value of G◦ (−32.00 kJ mol−1 ) indi- horticultural considerations. J. Clean. Prod. 86, 40–48.
Freundlich, H., 1909. Kolloidchemiev. Akademischer Verlagsgeselschaft, Leipzig.
cates that the nitrate adsorption process is spontaneous in nature. Ganesan, P., Kamaraj, R., Vasudevan, S., 2013. Application of isotherm, kinetic and
The positive value of H◦ (54.50 kJ mol−1 ) shows adsorption of thermodynamic models for the adsorption of nitrate ions on graphene from
nitrate by modified biochar bagasse sugarcane is endothermic in aqueous solution. J. Taiwan Inst. Chem. Eng. 44, 808–814.
Ghani, W.A.W.A.K., Mohd, A., da Silva, G., Bachmann, R.T., Taufiq-Yap, Y.H., Rashid,
nature. Other researchers have also reported that nitrate adsorp- U., Ala’a, H., 2013. Biochar production from waste rubber-wood-sawdust and
tion by different adsorbents is endothermic in nature (Bhatnagar its potential use in C sequestration: chemical and physical characterization.
et al., 2010; Liu and Zhang, 2015). The positive value of S◦ Ind. Crops Prod. 44, 18–24.
Hafshejani, L.D., Nasab, S.B., Gholami, R.M., Moradzadeh, M., Izadpanah, Z.,
(293.09 J mol−1 K−1 ) indicates the affinity of the modified biochar Hafshejani, S.B., Bhatnagar, A., 2015. Removal of zinc and lead from aqueous
for nitrate ions. The G◦ value obtained in this study for nitrate ions solution by nanostructured cedar leaf ash as biosorbent. J. Mol. Liq. 211,
is lower than −10 kJ mol−1 , suggesting that the physical adsorption 448–456.
Hassan, M.L., Kassem, N.F., Abd El-Kader, A.H., 2010. Novel Zr (IV)/sugar beet pulp
occurs during the adsorption process (Liu and Zhang, 2015). These
composite for removal of sulfate and nitrate anions. J. Appl. Polym. Sci. 117,
results are in agreement with the results of Dubinin–Raduskovich 2205–2212.
isotherm model. Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes.
Process Biochem. 34, 451–465.
Ho, Y., Porter, J., McKay, G., 2002. Equilibrium isotherm studies for the sorption of
4. Conclusions divalent metal ions onto peat: copper, nickel and lead single component
systems. Water Air Soil Pollut. 141, 1–33.
The results of this study suggest that the modified sugarcane Hu, Q., Chen, N., Feng, C., Hu, W., 2015. Nitrate adsorption from aqueous solution
using granular chitosan Fe3+ complex. Appl. Surf. Sci. 347, 1–9.
bagasse biochar is an effective adsorbent for the nitrate adsorption Iakovleva, E., Mäkilä, E., Salonen, J., Sitarz, M., Wang, S., Sillanpää, M., 2015. Acid
from aqueous solution. Nitrate adsorption efficiency was higher mine drainage (AMD) treatment: neutralization and toxic elements removal
under acidic pH as compared to basic pH which might be due with unmodified and modified limestone. Ecol. Eng. 81, 30–40.
Kaewprasit, C., Hequet, E., Abidi, N., Gourlot, J.P., 1998. Application of methylene
to the electrostatic attraction between nitrate ions and the mod- blue adsorption to cotton fiber specific surface area measurement: part I.
ified biochar surface. The obtained results showed that adsorption Methodology. J. Cotton Sci. 2, 164–173.
equilibrium was reached within 60 min. The experimental data fit- Katal, R., Baei, M.S., Rahmati, H.T., Esfandian, H., 2012. Kinetic, isotherm and
thermodynamic study of nitrate adsorption from aqueous solution using
ted well with pseudo-second-order kinetic model and Langmuir modified rice husk. J. Ind. Eng. Chem. 18, 295–302.
isotherm model. The maximum adsorption capacity of modified Keränen, A., Leiviskä, T., Hormi, O., Tanskanen, J., 2015. Removal of nitrate by
biochar for nitrate ions was obtained as 28.21 mg g−1 . modified pine sawdust: effects of temperature and co-existing anions. J.
Environ. Manage. 147, 46–54.
Kim, P., Hensley, D., Labbé, N., 2014. Nutrient release from switchgrass-derived
Acknowledgment biochar pellets embedded with fertilizers. Geoderma 232, 341–351.
Kumar, S., Masto, R.E., Ram, L.C., Sarkar, P., George, J., Selvi, V.A., 2013. Biochar
preparation from Parthenium hysterophorus and its potential use in soil
The authors would like to thank Shahid Chamran University of
application. Ecol. Eng. 55, 67–72.
Ahwaz, Iran, for the financial and other supports.
L. Divband Hafshejani et al. / Ecological Engineering 95 (2016) 101–111 111

Lagergren, S., 1898. About the theory of so-called adsorption of soluble substances Peng, P., Lang, Y.-H., Wang, X.-M., 2016. Adsorption behavior and mechanism of
Zur Theorieder Sogenannten. Kungliga Svenska Vetenskap sakademiens pentachlorophenol on reed biochars: pH effect, pyrolysis temperature,
Handlingar 24, 1–39. hydrochloric acid treatment and isotherms. Ecol. Eng. 90, 225–233.
Langmuir, I., 1916. The constitution and fundamental properties of solid and Rahman, N., Khan, M.F., 2015. Development of poly-o-toluidine zirconium (IV)
liquids. Part I. J. Am. Chem. Soc. 38, 2221–2295. ethylenediamine as a new adsorbent for nitrate: equilibrium modelling and
Li, H., Ye, X., Geng, Z., Zhou, H., Guo, X., Zhang, Y., Zhao, H., Wang, G., 2016. The thermodynamic studies. J. Ind. Eng. Chem. 25, 272–279.
influence of biochar type on long-term stabilization for Cd and Cu in Rahmani, A., Mousavi, H.Z., Fazli, M., 2010. Effect of nanostructure alumina on
contaminated paddy soils. J. Hazard. Mater. 304, 40–48. adsorption of heavy metals. Desalination 253, 94–100.
Liu, X., Zhang, L., 2015. Removal of phosphate anions using the modified chitosan Rangabhashiyam, S., Selvaraju, N., 2015. Efficacy of unmodified and chemically
beads: adsorption kinetic, isotherm and mechanism studies. Powder Technol. modified Swietenia mahagoni shells for the removal of hexavalent chromium
277, 112–119. from simulated wastewater. J. Mol. Liq. 209, 487–497.
Luo, S., Li, X., Chen, L., Chen, J., Wan, Y., Liu, C., 2014. Layer-by-layer strategy for Rezaei Kalantary, R., Dehghanifard, E., Mohseni-Bandpi, A., Rezaei, L., Esrafili, A.,
adsorption capacity fattening of endophytic bacterial biomass for highly Kakavandi, B., Azari, A., 2015. Nitrate adsorption by synthetic activated carbon
effective removal of heavy metals. Chem. Eng. J. 239, 312–321. magnetic nanoparticles: kinetics, isotherms and thermodynamic studies.
Mašek, O., Brownsort, P., Cross, A., Sohi, S., 2013. Influence of production conditions Desalin. Water Treat., 1–11.
on the yield and environmental stability of biochar. Fuel 103, 151–155. Sips, R., 1948. On the structure of a catalyst surface. J. Chem. Phys. 16, 490–495.
Masto, R.E., Kumar, S., Rout, T., Sarkar, P., George, J., Ram, L., 2013. Biochar from Tan, X., Liu, Y., Zeng, G., Wang, X., Hu, X., Gu, Y., Yang, Z., 2015. Application of
water hyacinth (Eichornia crassipes) and its impact on soil biological activity. biochar for the removal of pollutants from aqueous solutions. Chemosphere
Catena 111, 64–71. 125, 70–85.
McEnaney, B., 1987. Estimation of the dimensions of micropores in active carbons Tangsir, S., Hafshejani, L.D., Lähde, A., Maljanen, M., Hooshmand, A., Naseri, A.A.,
using the Dubinin-Radushkevich equation. Carbon 25, 69–75. Moazed, H., Jokiniemi, J., Bhatnagar, A., 2016. Water defluoridation using Al2 O3
Miller-Robbie, L., Ulrich, B.A., Ramey, D.F., Spencer, K.S., Herzog, S.P., Cath, T.Y., nanoparticles synthesized by flame spray pyrolysis (FSP) method. Chem. Eng. J.
Stokes, J.R., Higgins, C.P., 2015. Life cycle energy and greenhouse gas 288, 198–206.
assessment of the co-production of biosolids and biochar for land application. Trakal, L., Veselská, V., Šafařík, I., Vítková, M., Číhalová, S., Komárek, M., 2016. Lead
J. Clean. Prod. 91, 118–127. and cadmium sorption mechanisms on magnetically modified biochars.
Milmile, S.N., Pande, J.V., Karmakar, S., Bansiwal, A., Chakrabarti, T., Biniwale, R.B., Bioresour. Technol. 203, 318–324.
2011. Equilibrium isotherm and kinetic modeling of the adsorption of nitrates Wang, T., Lin, J., Chen, Z., Megharaj, M., Naidu, R., 2014. Green synthesized iron
by anion exchange Indion NSSR resin. Desalination 276, 38–44. nanoparticles by green tea and eucalyptus leaves extracts used for removal of
Mishra, P., Patel, R., 2009. Use of agricultural waste for the removal of nitrate in aqueous solution. J. Clean. Prod. 83, 413–419.
nitrate-nitrogen from aqueous medium. J. Environ. Manage. 90, 519–522. Weber, W., Morris, J., 1963. Intraparticle diffusion during the sorption of
Mohamad Nor, N., Lau, L.C., Lee, K.T., Mohamed, A.R., 2013. Synthesis of activated surfactants onto activated carbon. J. Sanit. Eng. Div. Am. Soc. Civ. Eng. 89,
carbon from lignocellulosic biomass and its applications in air pollution 53–61.
control—a review. J. Environ. Chem. Eng. 1, 658–666. Xiang, J., Liu, D., Ding, W., Yuan, J., Lin, Y., 2015. Effects of biochar on nitrous oxide
Mohammadi, A., Cowie, A., Anh Mai, T.L., de la Rosa, R.A., Kristiansen, P., Brandão, and nitric oxide emissions from paddy field during the wheat growth season. J.
M., Joseph, S., 2016. Biochar use for climate-change mitigation in rice cropping Clean. Prod. 104, 52–58.
systems. J. Clean. Prod. Xu, X., Gao, B.Y., Yue, Q.Y., Zhong, Q.Q., 2010. Preparation and utilization of wheat
Mohan, D., Sarswat, A., Ok, Y.S., Pittman, C.U., 2014. Organic and inorganic straw bearing amine groups for the sorption of acid and reactive dyes from
contaminants removal from water with biochar, a renewable, low cost and aqueous solutions. J. Hazard. Mater. 182, 1–9.
sustainable adsorbent—a critical review. Bioresour. Technol. 160, 191–202. Xu, X., Gao, B., Yue, Q., Li, Q., Wang, Y., 2013. Nitrate adsorption by multiple
Moradzadeh, M., Moazed, H., Sayyad, G., Khaledian, M., 2014. Transport of nitrate biomaterial based resins: application of pilot-scale and lab-scale products.
and ammonium ions in a sandy loam soil treated with potassium Chem. Eng. J. 234, 397–405.
zeolite—evaluating equilibrium and non-equilibrium equations. Acta Ecol. Sin. Yang, Z., Zhang, L., Xu, P., Zhang, X., Niu, X., Zhou, S., 2015. The adsorption of nitrate
34, 342–350. from aqueous solution onto calcined Mg/Fe hydrotalcite. Desalin. Water Treat.
Mukherjee, R., De, S., 2014. Adsorptive removal of nitrate from aqueous solution by 54, 3400–3411.
polyacrylonitrile?alumina nanoparticle mixed matrix hollow-fiber membrane. Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X., Pullammanappallil, P., Yang,
J. Membr. Sci. 466, 281–292. L., 2011. Removal of phosphate from aqueous solution by biochar derived from
Muthu Prabhu, S., Meenakshi, S., 2014. Synthesis of surface coated hydroxyapatite anaerobically digested sugar beet tailings. J. Hazard. Mater. 190, 501–507.
powders for fluoride removal from aqueous solution. Powder Technol. 268, Yao, Y., Gao, B., Chen, J., Yang, L., 2013. Engineered biochar reclaiming phosphate
306–315. from aqueous solutions: mechanisms and potential application as a
Muthu Prabhu, S., Meenakshi, S., 2015. Chemistry of defluoridation by one-pot slow-release fertilizer. Environ. Sci. Technol. 47, 8700–8708.
synthesized dicarboxylic acids mediated polyacrylamide–zirconium complex. Zhang, J., Wang, Q., 2016. Sustainable mechanisms of biochar derived from
Chem. Eng. J. 262, 224–234. brewers’ spent grain and sewage sludge for ammonia–nitrogen capture. J.
Namasivayam, C., Höll, W.H., 2005. Quaternized biomass as an anion exchanger for Clean. Prod. 112 (Part 5), 3927–3934.
the removal of nitrate and other anions from water. J. Chem. Technol. Biot. 80, Zhen, Y., Ning, Z., Shaopeng, Z., Yayi, D., Xuntong, Z., Jiachun, S., Weiben, Y., Yuping,
164–168. W., Jianqiang, C.A., 2015. pH-and temperature-responsive magnetic composite
Olgun, A., Atar, N., Wang, S., 2013. Batch and column studies of phosphate and adsorbent for targeted removal of nonylphenol. ACS Appl. Mater. Interfaces 7,
nitrate adsorption on waste solids containing boron impurity. Chem. Eng. J. 24446–24457.
222, 108–119.
Orlando, U.S., Baes, A.U., Nishijima, W., Okada, M., 2002. Preparation of agricultural
residue anion exchangers and its nitrate maximum adsorption capacity.
Chemosphere 48, 1041–1046.

You might also like