Accepted Manuscript Fabrication of granular activated carbons derived from spent coffee grounds by entrapment in calcium alginate beads for adsorption of acid orange 7 and methylene blue Kyung-Won Jung, Brian Hyun Choi, Min-Jin Hwang, Tae-Un Jeong, Kyu-Hong Ahn PII: DOI: Reference:
S0960-8524(16)31077-X http://dx.doi.org/10.1016/j.biortech.2016.07.098 BITE 16860
To appear in:
Bioresource Technology
Received Date: Revised Date: Accepted Date:
23 June 2016 21 July 2016 22 July 2016
Please cite this article as: Jung, K-W., Choi, B.H., Hwang, M-J., Jeong, T-U., Ahn, K-H., Fabrication of granular activated carbons derived from spent coffee grounds by entrapment in calcium alginate beads for adsorption of acid orange 7 and methylene blue, Bioresource Technology (2016), doi: http://dx.doi.org/10.1016/j.biortech.2016.07.098
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1
Fabrication of granular activated carbons derived from spent coffee
2
grounds by entrapment in calcium alginate beads for adsorption of
3
acid orange 7 and methylene blue
4
Kyung-Won Junga, Brian Hyun Choia, Min-Jin Hwangb, Tae-Un Jeonga, Kyu-Hong
5
Ahna,*
6
a
7
Technology, 5 Hwarang-ro 14-gil, Seongbuk-gu, Seoul 02792, South Korea
8
b
9
Munsan, Jinju, Gyeongsangnam-do 52834, South Korea
Center for Water Resources Cycle Research, Korea Institute of Science and
Future Environmental Research Center, Korea Institute of Toxicology, 17 Jeigok-gil,
10
*
11
958-5832; fax: +82-2-958-6854)
12
Abstract
Corresponding author: Kyu-Hong Ahn (E-mail address:
[email protected], Tel: +82-2-
13
Biomass-based granular activated carbon was successfully prepared by entrapping
14
activated carbon powder derived from spent coffee grounds into calcium-alginate beads
15
(SCG-GAC) for the removal of acid orange 7 (AO7) and methylene blue (MB) from
16
aqueous media. The dye adsorption process is highly pH-dependent and essentially
17
independent of ionic effects. The adsorption kinetics was satisfactorily described by the
18
pore diffusion model, which revealed that pore diffusion was the rate-limiting step
19
during the adsorption process. The equilibrium isotherm and isosteric heat of adsorption
20
indicate that SCG-GAC possesses an energetically heterogeneous surface and operates
21
via endothermic process in nature. The maximum adsorption capacities of SCG-GAC
22
for AO7 (pH 3.0) and MB (pH 11.0) adsorption were found to be 665.9 and 986.8 mg/g
23
at 30 oC, respectively. Lastly, regeneration tests further confirmed that SCG-GAC has
1
promising potential in its reusability, showing removal efficiency of more than 80%
2
even after seven consecutive cycles.
3 4
Keywords
5
Activated carbon; spent coffee grounds; calcium-alginate beads; adsorption; acid orange
6
7; methylene blue
7 8 9
1. Introduction The presence of dyestuff in effluents from multiple industries such as food, paper,
10
carpets, rubbers, cosmetics, and textiles is a major concern because it can directly
11
damage the entire ecosystems due to its resistance to biodegradation and high stability
12
towards light and/or oxidizing agents (Doğan et al., 2009). It is estimated that roughly
13
7×10 5 tons of dyestuff are produced annually and approximately 15% of the produced
14
dyes is released into the environment without proper treatment (Rafatullah et al., 2010).
15
The presence of minuscule concentrations of dyestuff in wastewater (~1 mg/L) is easily
16
visible and undesirable, as the color of wastewater is recognized as the primary factor
17
when determining the degree of contaminant (Lee et al., 2006). Moreover, the major
18
concern of widespread dyes in water systems is that they are toxic to aquatic organisms,
19
as well as carcinogenic and mutagenic to humans (Ma et al., 2015). Hence, the
20
development of industrially applicable and efficient treatments for dye containing
21
wastewater is imperative before disposal to aquatic ecosystems.
22
Among numerous approaches towards dye removal from wastewater (i.e., physical,
23
chemical, biological, advanced oxidation processes, electrochemical), activated carbon-
24
based adsorption processes has been found to be superior than other methods because of
25
its attractive features such as high specific surface area, surface reactivity, efficiency,
1
and convenience in removing targets from aqueous solution (Li et al., 2016; Sun et al.,
2
2015). The major drawback of the conventional activated carbon-based adsorption
3
process is the use of non-renewable and relatively expensive raw materials (i.e., wood
4
and coal), and therefore, it is still considered to be a costly process with restricted
5
applicability. To overcome such drawbacks and to attain better economic viability,
6
adsorption processes using non-conventional low-cost activated carbon derived from
7
renewable and inexpensive materials such as naturally abundant biomass, agricultural,
8
and industrial wastes have been examined (Li et al., 2016; Rafatullah et al., 2010; Sun et
9
al., 2015).
10
Coffee is one of the most abundant agricultural products, as well as one of the most
11
widely consumed beverages in the world. Approximately six million tons of solid
12
residue known as spent coffee grounds (SCG)/exhausted coffee residue is produced
13
during the brewing process (Franca et al., 2009). In addition to further conversion of
14
SCG into valuable products such as enzymes, organic acids, fuels, ethanol, dyes,
15
compost, and antioxidant phenolic compounds (Murthy and Naidu, 2012), preparation
16
of SCG-based activated carbon as an effective adsorbent for the removal of pollutants
17
(e.g., heavy metals, dyes, and organic contaminants) from aqueous solutions also has
18
been proposed and scrutinized (Franca et al., 2009; Ma and Ouyang, 2013; Reffas et al.,
19
2010; Safarik et al., 2012). From these studies, SCG-based activated carbon has been
20
considered as an effective and promising adsorbent for removing pollutants from
21
aqueous solution. However, in some cases, its use is still considered undesirable due to
22
its difficulty in recovering or separating the powdered SCG-based activated carbon from
23
the mixture during the post-adsorption stage, which can substantially obstruct its
24
beneficial utility on an industrial scale.
25 26
To bypass this critical disadvantage, alginate (alginic acid), a naturally occurring polysaccharide extracted from brown seaweed, has recently received considerable
1
attention as researchers have sought to immobilize powdered activated carbon by cross-
2
linking the powder to the matrix. One of the most significant properties of alginate
3
solution is the ease to form hydrogels (egg-box structure) by selective ionic interactions
4
with divalent metals such as Ca (Ai et al., 2011). Taking this factor into consideration,
5
the objective of this work is to investigate a convenient and economical method for
6
dyestuff removal from an aqueous solution by adsorption using SCG-based granular
7
activated carbon (SCG-GAC), which is prepared by entrapping powdered SCG-based
8
activated carbon in porous calcium-alginate beads. In this study, a mono-azo anionic
9
dye-acid orange 7 (AO7) and cationic dye-methylene blue (MB) were chosen as model
10
pollutants because of their versatility and wide application in multiple industries such as
11
medical processing, paper colorization, cotton dying, and so on. In addition to
12
monitoring the physicochemical properties of SCG-GAC, its adsorption properties
13
including effects of additives, kinetics and equilibrium isotherms were evaluated at
14
different temperatures.
2+
15 16
2. Materials and methods
17
2.1. Chemicals and materials
18
Raw-SCG, a by-product collected from coffees houses, was used as a precursor.
19
Raw-SCG was washed with distilled deionized water (Milli Q plus, Merck Millipore
20
Co., Germany) to remove impurities such as dust and water soluble substances from the
21
surface. After drying at 105 oC for 24 h, the fraction passing through a 425 µm sieve
22
(mesh No. 40) was collected prior to its activation. Anionic AO7 (C16H11N2NaO4S,
23
MW= 350.32 g/mol) was supplied by Tokyo Chemical Industry (Japan). Cationic MB
24
(C16H18CIN3S·3H2O, MW= 373.90 g/mol) was purchased from Samchun Pure
25
Chemical Co., Ltd. (South Korea). Each stock solution was prepared by dissolving 2.0 g
1
of AO7 and MB in 1.0 L of deionized distilled water, respectively, without further
2
purification and was diluted to appropriate concentrations. The measurement of AO7
3
and MB concentrations was performed using a UV Spectrophotometer (DS5000, HACH
4
Company, Netherlands). The maximum wavelength (λmax) for AO7 and MB was found
5
to be 485 and 663 nm, respectively. After the adsorption, the respective concentrations
6
were calculated using the calibration curves in the range of 0-20 mg/L (Abs485 at 20
7
mg/L= 1.465) for AO7 and 0-10 mg/L (Abs663 at 10 mg/L= 2.012) for MB. The removal
8
efficiencies were expressed as percent ratio of removed dye concentrations to their
9
initial concentrations.
10 11 12
2.2. Preparation of SCG-based granular activated carbon (SCG-GAC) First, in the carbonization step, the dried raw-SCG was heated in a horizontal electric
13
furnace ((TF-80-6, Woori Science Instrument Co., South Korea) with a quartz tubular
14
reactor up to 450 oC with a heating rate of 7 oC/min. After dwelling for 2 h under a
15
nitrogen atmosphere (flow rate= 25 mL/min), the product was cooled to room
16
temperature under a nitrogen atmosphere. Then, KOH was impregnated into the
17
carbonized product in 5:1 ratio w/w (g KOH/g carbonized-SCG) for 1 h, and the
18
mixture was oven-dried at 105 oC until it was completely dried. Thereafter, activation of
19
the impregnated SCG was carried out via pyrolysis. The impregnated-SCG was placed
20
in a horizontal electric furnace at 700 oC for 1 h with a ramp rate of 7 oC/min under
21
nitrogen protection (flow rate= 25 mL/min). After cooling to room temperature, the
22
samples were transferred to a bottle containing 200 mL of 0.1 M HCl and stirred for 1 h.,
23
followed by filtering and washing repeatedly with distilled deionized water until the
24
filtrate pH became neutral (pH 6.5-7.0). The samples were dried at 105 oC for 24 h, and
25
then stored in sealed bottles before their use. As shown in Fig. S1 (supplementary data),
1
the percentage removals of AO7 and MB at 200 mg/L of the initial concentration were
2
99.9 and 100%, respectively, after 24 h of contact time using powdered SCG-based
3
activated carbon, because the prepared activated carbon has more suitable textural
4
properties for AO7 and MB adsorption than those of raw-SCG and carbonized-SCG, as
5
provided in the supplementary data (Fig. S2 and Table S1). The N2
6
adsorption/desorption isotherms was determined as a typical type I isotherm according
7
to the international union of pure and applied chemistry (IUPAC) classification and a
8
remarkable uptake at low relative pressure (Fig. S3) further confirmed that the
9
micropores were predominant in the prepared activated carbon. Thus, the experimental
10
data suggest that the activated carbon was successfully prepared and could be used as a
11
highly effective adsorbent to remove AO7 and MB from aqueous solutions.
12
SCG-GAC was prepared from the powdered SCG-based activated carbon according
13
to the procedure described by the Jung research group (Jung et al., 2016). In detail,
14
2.0% (w/v) sodium alginate solution was prepared by mixing sodium alginate in 500
15
mL of distilled deionized water with stirring for 1 h at room temperature. Subsequently,
16
2.0% (w/v) of prepared powdered SCG-based activated carbon was added to the
17
alginate solution, and the mixture was then stirred for 10 h. After the mixture became
18
homogeneous, it was dropped through a burette into 2.0% (w/v) calcium chloride to
19
form granules under gentle stirring. The excess unbounded calcium chloride was
20
removed by washing several times with distilled deionized water, and the product was
21
then oven-dried at 45 oC for 24 h. From here on, the prepared SCG-based granular
22
activated carbon is denoted as SCG-GAC.
23 24
2.3. Characterization of SCG-GAC
1
The morphology and the microstructure of the prepared SCG-GAC were investigated
2
by scanning electron microscopy (SEM, S-4200, Hitachi Co., Japan). Fourier transform
3
infrared (FTIR) spectra were recorded using a FTIR spectrometer in the 4000-600 cm-1
4
range (NICOLET iS10, Thermo Scientific, USA) to identify existing functionalities
5
within the prepared material. The porosity of SCG-GAC were characterized by N2
6
vapor adsorption/desorption studies at 77K using a nano porosity system (NP-XQ,
7
Mirae Scientific Instruments, South Korea). Based on the results of N2-mediated
8
adsorption/desorption isotherms, surface area, pore diameter, and total pore volume of
9
the product were measured using the Brunauer-Emmett-Teller (BET), the Barrett-
10
Joyner-Halenda (BJH), and the Horvath-Kawazoe methods. In addition, the pore size
11
distribution was estimated by the BJH method from the acquired desorption branches.
12
To determine the point of zero charge (pHpzc), the solution pH was adjusted in a range of
13
2.0-10.0 (with intervals of “1.0”) with 0.1 M HCl and NaOH using a LabQuest2
14
portable meter (LQ2-LE, Vernier, USA); the SCG-GAC was crushed into powder and
15
added to the reaction vessels. After shaking for 1 h at 30 oC for homogeneity, zeta
16
potential was measured using a zeta potential analyzer (Zetasizer Nano ZS, Malvern
17
Instrument. Ltd., UK).
18 19 20
2.4. Batch adsorption experiments The influence of solution pH on the adsorption capacities of AO7 and MB were
21
monitored to determine optimal conditions for further investigations. The effect of
22
background salt (ionic strength) on the adsorption performance also evaluated. Based on
23
the results of preceding tests, to evaluate the adsorption properties of SCG-GAC for
24
AO7 and MB and elucidate the main adsorption mechanisms, the adsorption kinetics
25
and equilibrium isotherms were carried out at three different temperatures (10, 20, and
1
30 oC). Lastly, to assess the reusability of saturated SCG-GAC, regeneration tests were
2
carried out. The detailed experimental conditions for each test are described below. All
3
experiments were carried out in duplicate and the experimental errors were found to be
4
within ±1% of the presented average values.
5 6 7
2.4.1. Influences of solution pH and ionic strength The influence of solution pH on the adsorption capacity of the adsorbent was
8
evaluated by mixing 0.1 g of SCG-GAC with 50 ml of stock solutions containing 200
9
mg/L of AO7 or MB in the reaction vessel. The pH of the solutions were adjusted to
10
3.0-11.0 (with intervals of “2.0”) using 0.1 M HCl and NaOH; the reaction vessels were
11
then homogeneously shaken in an orbit shaking incubator at 200 rpm with controlled
12
temperature at 20 oC (WIS02011, WiseCube®, Germany). After 24 h of contact time, the
13
concentrations of the dyes within the solutions were analyzed immediately. In the same
14
manner, the effect of ionic strength on the adsorption performance was also evaluated
15
with varying salt concentration (0.5, 1.0, 2.0, and 3.0 g/L) because salts are often found
16
in dye containing wastewater, particularly NaCl as it is the most common salt existent in
17
the wastewater (Zille et al., 2003). The pH of the solutions was fixed at 3.0 and 11.0 for
18
AO7 and MB, respectively, while other experimental procedures and conditions were
19
unaltered. The removal efficiencies and the amount of AO7 and MB adsorbed onto
20
SCG-GAC were calculated using the following equations:
Removal rate (%) =
Qe =
(Ci - C f )
V ( Ci − C e ) M
Ci
× 100
(1)
(2)
1
where Ci and Cf are the AO7 and MB concentration (mg/L) of the initial and final
2
solution, respectively. Qe is the adsorbed amount of AO7 and MB per unit weight of
3
adsorbent (mg/g) at a given time, M is the mass of the adsorbent (g), V is the volume of
4
dye solutions (L), and Ci and Ce (mg/L) are the initial and equilibrium concentrations of
5
the target compounds, respectively.
6 7 8 9
2.4.2. Adsorption kinetics The adsorption kinetics were investigated to evaluate the adsorption properties of SCG-GAC by mixing 0.5 g of SCG-GAC in a reaction vessel containing 250 mL of
10
2000 mg/L AO7 or MB. Based on the preceding experiment, the solution pH was
11
adjusted to pH 3 for AO7, and pH 11 for MB using 0.1 M of HCl and NaOH. The
12
solution temperature was controlled at 10, 20, and 30 oC. Aliquots were taken
13
periodically at designated time intervals; the amounts of AO7 and MB adsorbed onto
14
SCG-GAC were then calculated as described earlier. Using the obtained experimental
15
data, the pore diffusion model was employed to fully identify the diffusion mechanisms
16
of dye molecules onto SCG-GAC. The pore diffusion model equations and numerical
17
method for solving the equations (initial and boundary conditions) are given as the
18
following, respectively (Jung et al., 2016): Pore diffusion model:
εp
∂C p
∂t
+ ρ
p
∂q = ε ∂t
p
∂C p 1 ∂ 2 r D p r 2 ∂ r ∂ r
(3)
Initial condition:
C p (r, t = 0) = 0
(4)
Boundary conditions: At the center of the particle:
At the surface of the particle:
∂C p ∂r
(5)
=0 r=0
ε p Dp
∂C p ∂r
= k f ( C − Cs )
(6)
r =Rp
1
where εp is the porosity of the particle, ρp is the particle density of the adsorbent (kg/L),
2
rp is the radial coordinate of the particle (m), Dp is the internal pore diffusivity (m2/s),
3
Cp is the solution concentration in pores (mg/L), q is the adsorbate concentration in
4
particle phase (mg/g), t is time (s), kf is the external-film transfer coefficient (m/s), and
5
C and Cs are the concentrations in the bulk aqueous phase and the external surface
6
(mg/L), respectively. The particle porosity, density, and radius of the prepared SCG-
7
GAC were measured as 0.65, 0.281 kg/L, and 2.0316 mm, respectively.
8 9
2.4.3. Adsorption equilibrium isotherms
10
Adsorption equilibrium isotherm studies were carried out by subjecting 0.1 g of
11
SCG-GAC in 50 mL of AO7 and MB solutions with different concentrations, ranging
12
from 50 to 5000 mg/L at three different temperatures (10, 20, and 30 oC). Contact time
13
of 24 h was chosen based on the results of kinetics studies; other experimental
14
procedures were set constant as before. The experimental results were analyzed using
15
three adsorption isotherm models, Langmuir, Freundlich, and Sips models, which are
16
expressed as the following, respectively (McKay et al., 2014):
Qe =
Q m K LC e 1 + K LC e
(7)
Qe = K F C e1 / n
(8)
Q m K S C e1 / n Qe = 1 + K S C e1 / n
(9)
1
where Qe is the adsorbed amounts of AO7 and MB per unit weight of adsorbent (mg/g)
2
at an equilibrium concentration of adsorbate in bulk solution (Ce, mg/L). KL, KF, and KS
3
are the Langmuir, Freundlich, and Sips constants, respectively. Qm denotes the
4
maximum adsorption capacity and n is the heterogeneity factor.
5
Using the obtained results from equilibrium isotherms, the isosteric heat of
6
adsorption and its variation with surface loading were determined by the Clausius-
7
Clapeyron equation, expressed as follows (Jung et al., 2016):
Qst ∂ ln Ce = RT 2 ∂T N
(10)
8
where Qst is the isosteric heat of adsorption (kJ/mol), R is the gas constant (8.3145 J/mol
9
K), Ce is the equilibrium concentration (mol/L), T is temperature (K), and N is the moles
10
adsorbed.
11 12
2.4.4. Evaluation of experimental data
13
In order to evaluate the applicability of equilibiurm isotherm models for fitting the
14
experimental data, the coefficient of determination (R2) and error functions including
15
the residual root mean square error (RMSE), the sum of the squares of the errors (SSE),
16
and the chi-square test (χ2) were employed, expressed as the following, respectively:
n
∑
1− R2 =
(Q e.n − Q m .n ) 2
n =1
(11)
n
∑
(Q e.n − Q m .n )
2
n =1
RMSE =
1 n (Q e . n − Q m . n ) 2 ∑ n − 1 n =1
n
SSE = ∑ ( Qm ,n − Qe ,n )
2
(12)
(13)
n =1
2
χ =
n
∑ n =1
(Q e . n − Q m . n ) 2 Q e.n
(14)
1
where Qe and Qm are the adsorbed amount per unit weight of adsorbent (mg/g) at
2
equilibrium and the calculated value at equilibrium from the model, respectively. n is
3
the number of observation.
4 5 6
2.4.5. Regeneration test To assess the recyclability potential of the used SCG-GAC, regeneration tests were
7
repeated for seven cycles. In this test, after the equilibrium conditions were reached,
8
each saturated SCG-GAC by AO7 and MB was taken immediately from the reaction
9
vessels; the samples were then put into 100 mL of methanol. Subsequently, the mixtures
10
were shaken homogeneously at 200 rpm for 24 h with controlled temperature at 20 oC.
11
After washing with distilled deionized water and thoroughly drying at 45 oC for 24 h,
12
the adsorption tests were carried out at initial dye concentrations of 200 mg/L following
13
the same procedures as those used for the equilibrium isotherm experiments. The
14
removal efficiencies and the adsorbed amount of AO7 and MB onto SCG-GAC were
15
then calculated by using Eqs. (1) and (2), respectively.
1 2
3. Results and discussion
3
3.1. General characteristics of SCG-GAC
4
Fig. S4 (see supplementary data) depicts the general characteristics of the prepared
5
SCG-GAC, including morphological, textural, and surface chemical properties. As
6
shown in Fig. S4(A), SCG-GAC is identified as a black sphere with an average
7
diameter of 2.03±0.14 mm, an indicative of successful immobilization of the powdered
8
SCG-based activated carbon into calcium-alginate matrix. The SEM images in Fig.
9
S4(B), taken at a magnitude of ×40 (inset) and ×100, delineate that SCG-GAC bears an
10
egg-like structure with wrinkled and rough surface. Such heterogeneous surface may be
11
beneficial for the uptake of dye molecules in an aqueous solution (Ai et al., 2011).
12
N2 adsorption/desorption isotherms at 77 K for the prepared SCG-GAC were
13
analyzed and core textural properties, including BET surface area, total pore volume,
14
and an average pore size, are presented in Fig. S4(C). It can be seen that the isotherm is
15
classified as a typical type I isotherm, indicating a typical microporous surface
16
according to the IUPAC classification. A remarkable uptake of the N2 vapor at low
17
relative pressure (P/P0) is due to the micropore filling effect. The determined BET
18
surface area and the total pore volume for SCG-GAC were 704.23 m2/g and 0.2926
19
cm3/g, respectively. The pore size distribution estimated by the BJH method from the
20
desorption branch (Fig. S4(C)-inset) further confirmed that the average pore size of
21
SCG-GAC was 2.1966 nm. Even though the estimated average pore size exceeded the
22
micropore range (< 2.0 nm), the prepared SCG-GAC can be considered a microporous
23
material because it had a peak at 1.7754 nm and this is very close to the upper limit of
24
the micropore range (Ma et al., 2015). The overall structural analysis suggest that the
25
prepared SCG-GAC possesses suitable properties as an adsorbent for removing AO7
1
and MB, from an aqueous solution because the molecular dimensions of AO7 and MB
2
in water are 0.544 × 1.003 × 1.567 nm and 0.400 × 0.793 × 1.634 nm, respectively
3
(Zhao et al., 2013).
4
The FTIR spectra of both naked and exhausted SCG-GAC were obtained from 4000
5
to 600 cm-1, shown in Fig. S4(D). The peak observed at 3339.92 cm-1 is caused by
6
residual water or hydrogen-bonded O−H stretching vibrations existent in the alginate
7
matrix. The peaks at 3000-2800 cm-1 (2924.54 and 2853.66 cm-1) are axial and
8
equatorial sp3 C-H stretching vibrations (Bedin et al., 2016) from the alginate. The
9
peaks in the wave number range of 1600-1300 cm-1 (1599.61 and 1421.68 cm-1)
10
correspond to aromatic C=C stretching vibrations and COO- stretching vibrations (Liou,
11
2010). The peak observed at 1028.08 cm-1 represents C-O stretching vibrations and the
12
existence of multiple shoulders adjacent to the main peak is attributed via the positional
13
differences among these functional groups (OH and COO-) within the alginate matrix
14
(Papageorgiou et al., 2010). The peak observed in 800-600 cm-1 region can be attributed
15
to different types (aromatic and non-aromatic) of C=C bending vibrations (Unur, 2013).
16
After the adsorption of the substrates, the major peaks identified from the naked
17
adsorbent were slightly shifted due to the adsorbent’s interaction with adsorbate
18
molecules. However, the values fall within the range of the assigned functional groups,
19
indicating that they are involved in the interaction with the adsorbate, but no chemical
20
alterations regarding the functional groups occurred. In addition, the shift of the bands
21
and their decrease of corresponding intensities observed in 800-500 cm-1 region can be
22
attributed to the presence of π- π interactions between dye molecules and aromatic
23
regions of SCG-GAC (Bedin et al., 2016).
24 25
3.2. Influences of solution pH and ionic strength
1
The solution pH is one of the most influential factors controlling the adsorption
2
process due to its close relations with the surface charge (protonated or deprotonated) of
3
the adsorbent surface and the degree of ionization degree of dye molecules (Elmoubarki
4
et al., 2015; Franca et al., 2009). In this study, the influence of solution pH on the
5
adsorption of AO7 and MB was analyzed by varying the solution pH from 3.0 to 11.0,
6
and the results are shown in Fig. 1(A). The adsorption capacities were strongly
7
dependent on the solution pH and diametric trends were observed between AO7 and
8
MB, where the highest adsorption performance was found to be 97.4% (73.5 mg/g) at
9
pH 3.0 for AO7 and 99.6% (75.2 mg/g) at pH 11.0 for MB. Similar observations were
10
reported by several groups and demonstrated that such phenomena occur due to the
11
possession of strong ionic characters by AO7 and MB under acidic and alkaline
12
conditions, respectively (Ai et al., 2011; Benhouria et al., 2015; Doğan et al., 2009;
13
Elmoubarki et al., 2015; Nourmoradi et al., 2015; Reddy et al., 2013; Uddin et al., 2009).
14
In detail, as depicted in Fig. 1(A), the adsorbed amount of AO7 dwindled with
15
increasing solution pH from 3.0 to 11.0, which may be due to the following reasons:
16
First, the formation of negatively charged electrical double layer near the surface of the
17
SCG-GAC may induce an electrostatic repulsion between dissociated AO7 (pKa1: 11.4,
18
pKa2: 1.0) and the negatively charged layer near the surface of SCG-GAC. Another
19
possible explanation may be due to the increased degree of tautomerization within the
20
dye molecule with an increase of pH condition. The proton in the hydroxyl group can
21
go under intramolecular transfer to one of the nitrogen atoms in the azo group as shown
22
in Fig. S5 (see supplementary data). Because the predominant species of AO7 in water
23
is the hydrazine tautomer at relatively higher pH condition (Alosmanov, 2016), the ketal
24
group in the hydrazine tautomer may undermine the adsorption capacity due to its lower
25
degree of interaction with the adsorbent compared to the azo tautomer due to the lack of
1
hydrogen bonding at the oxygen atom of the substrate. The transferred hydrogen may be
2
less prone to hydrogen bonding as it is located in the inner core of the dye molecule.
3
Lastly, the point of zero charge (pHpzc) for this study was 7.6 (Fig. S6), and as a result,
4
the surface charge of SCG-GAC will be negatively charged at pH > pHpzc suppressing
5
the adsorption capacity of AO7 (60.5% and 45.8 mg/g at pH 11).
6
In contrast, it was found that the adsorption capacity of MB remained almost
7
constant, but it significantly increased from 7.0 to 11.0 (pH > pHpzc). This is due to the
8
stronger electrostatic attraction between the positively charged MB (pKa= 3.8) and
9
more negatively charged surface of SCG-GAC when the solution pH was 11.0, and
10
thereby the adsorption capacity of MB was higher than that at pH 7.0 (95.8% and 70.8
11
mg/g) and 9.0 (98.9% and 74.8 mg/g). At lower solution pH values (pH < pHpzc),
12
protonation at the surface of the adsorbent takes place, which causes electrostatic
13
repulsion with the MB molecules. Additionally, high concentration of H+ will protonate
14
the carboxylate groups within the alginate matrix, as well as form a positively charged
15
double layer at the surface, repelling the MB from approaching the active sites, and
16
ultimately plummeting the adsorption capacity. These results confirmed that the
17
adsorption of the prepared SCG-GAC for AO7 and MB were pH dependent; therefore,
18
the remaining tests (ionic strength, kinetic and equilibrium isotherm) were investigated
19
at constant pH conditions of 3.0 and 11.0 for AO7 and MB, respectively.
20
Followed by the study on the influence of pH, the influence of ionic strength on the
21
adsorption capacities of AO7 and MB onto SCG-GAC also was investigated by varying
22
the NaCl concentration from 0.5 to 3.0 g/L because the adsorption process in some
23
cases is markedly affected by the existence of electrolytes due to their interference on
24
the electrostatic interaction between the adsorbent and absorbent, drastically impacting
25
the adsorption process (Doğan et al., 2009; Özdemir et al., 2006). However, contrary to
1
expectations, superficial change of the adsorption capacities for AO7 and MB were
2
observed when increasing the ionic strength even to a NaCl concentration of 3.0 g/L
3
(Fig. 1(B)). Similar results were reported for the adsorption of dyes such as AO7, MB,
4
and reactive red 24 using zinc oxide coated activated carbon and paper mill sewage
5
sludge derived activated carbon, respectively (Li et al., 2011; Nourmoradi et al., 2015).
6
These results suggest that the binding affinity between SCG-GAC and dye molecules is
7
not significantly affected by background salt concentrations; thus, the adsorption of
8
AO7 and MB predominantly takes place at the internal-layer of SCG-GAC (ionic bonds,
9
strongly bonding) rather than the external-layer (electrostatic or Van der Waals
10
interactions, weakly bonding) because the latter could be more prone to the influence
11
from ionic intervention than the former (Zhu et al., 2015).
12 13 14
3.3. Adsorption kinetics Figs. S7(A) and (B) (see supplementary data) delineate that the adsorption kinetics
15
appear to involve three phases over the course of reaction, insinuating that the
16
adsorption is comprised of multiple steps. Generally, adsorption process can be
17
categorized into three consecutive diffusion steps, followed by an equilibrium step: (1)
18
bulk diffusion: transport of the adsorbate toward the external surface of the adsorbent;
19
(2) film diffusion or boundary layer diffusion: diffusion of the adsorbate to the external
20
surface of the adsorbent via the boundary layer; and (3) pore diffusion or intraparticle
21
diffusion: diffusion of the adsorbate within the pores and interior surface of the
22
adsorbent except for a small amount of adsorption (Sheha and El-Khouly, 2013). As
23
previously mentioned, the pore size of SCG-GAC is larger than the molecular sizes of
24
AO7 and MB. Thus, it can be conjectured that pore diffusion is the primary mechanism
25
for the investigated adsorption process. Despite the prevalence of reaction-based models
1
such as pseudo-first and pseudo-second models in kinetic analysis, these models are
2
insufficient to utterly interpret and elucidate the diffusion mechanisms involved in
3
adsorption processes. To overcome such limitations, a numerical simulation model, pore
4
diffusion model (Eq. (3)), was used to describe the adsorption kinetics of AO7 and MB
5
onto SCG-GAC and determine the external-film mass transfer (kf) and the pore
6
diffusion coefficients (Dp) at three different temperatures.
7
Figs. 2(A) and (B) represent the concentration decay curves of AO7 and MB,
8
respectively, along with their corresponding theoretical predictions for SCG-GAC at
9
three different temperatures. The results reveal that the employed pore diffusion model
10
demonstrates an excellent agreement with the experimental data, suggesting that the
11
model accurately describes the dynamics of the adsorption. The calculated parameters in
12
terms of the external-film transfer coefficient and the pore diffusion coefficient are
13
listed in Table 1 with the coefficient of determination values (R2). The obtained values
14
propose that the pore diffusion is the main rate-limiting step during the adsorption of
15
AO7 and MB. As seen, the external-film transfer coefficients for AO7 and MB vary
16
with respect to temperature and were found to be in the range of 2.228 × 10-6 to 3.101 ×
17
10 -6 m/sec and 2.280 × 10 -7 to 4.291 × 10-6 m/sec, respectively. The pore diffusion
18
coefficients were found to be in a range of 6.12 × 10 -10 to 1.36 × 10 -9 m/sec for AO7 and
19
7.823 × 10 -10 to 2.443 × 10-9 m/sec for MB, with a high R2 values around 0.99, which
20
reflects high accuracy.
21
The simulation for the concentration gradients of particle phase AO7 and MB
22
demonstrated that the pore diffusion rate of dye molecules as a function of contact time
23
gradually increased with increasing temperature, as shown in Figs. S8 and S9 (see
24
supplementary data). These phenomena may be caused by not only the increased
25
mobility and kinetic energy of the adsorbate with lowered boundary layer thickness, but
1
also decreased intraparticle diffusion resistance for adsorbing dye molecules with
2
increasing temperature. In addition, the adsorbates may penetrate deeper and faster at
3
higher temperatures via swelling of the internal structure (Baccar et al., 2013). These
4
findings confirmed that the adsorption mechanism is an endothermic process.
5
Comparing AO7 and MB, the relatively lower external-film transfer and pore diffusivity
6
of AO7 are likely due to its heavier molecular weight (350.32 g/mol for AO7 vs. 319.85
7
g/mol for MB) and larger dimensions compared to those of MB.
8 9 10
3.4. Adsorption equilibrium isotherms Figs. 3(A) and (B) show the plots of equilibrium concentration (Ce) versus the
11
adsorption capacities of the AO7 and MB at different temperatures (10-30 oC),
12
respectively. The plotted curves indicated that SCG-GAC has high binding affinity
13
towards the substrates. As seen, the adsorption capacities of AO7 and MB onto SCG-
14
GAC gradually increased with increasing temperature, and reached from 460.5 mg/g at
15
10 oC to 634.6 mg/g at 30 oC for AO7 and 465.5 mg/g at 10 oC to 710.2 mg/g at 30 oC
16
for MB. These findings are consistent with the obtained kinetic results, suggesting the
17
enhancement of adsorbent-adsorbate interaction via simultaneous increase of both
18
kinetic energies of the two matters and the reactivity at the surface of the adsorbent
19
(Chowdhry et al., 2011). To quantitatively evaluate the respective adsorption capacities
20
of SCG-GAC for AO7 and MB and to unearth it adsorption behaviors, three isotherm
21
models, Langmuir, Freundlich, and Sips models, were employed and the obtained fitting
22
curves are shown in Figs. 3(A) and (B). Also, the determined isotherm model constants
23
along with R2, RMSE, SSE, and χ2 for the three isotherm models, are summarized in
24
Table 2. Both AO7 and MB exhibited the largest R2 values and the lowest error
25
functions values were estimated by the Sips isotherm model. Generally, a larger R2 and
1
smaller error functions values reflect greater coherence between the model and
2
experimental results. Out of the employed models, the Sips model is the most accurate
3
representation of the experimetnal adsorption of the dye molecules. Therefore, the
4
adsorption of AO7 and MB onto SCG-GAC possibly took place on an energetically
5
heterogeneous surface in a multi-step fashin (Jung et al., 2016).
6
In the Sips isotherm model, KS and 1/n represent the affinity of the binding sites and
7
adsorption intensity (heterogeneity factor), respectively. A favorable adsorption is
8
predicted when KS is large and 1/n is small. As shown in Table 2, KS values for AO7
9
and MB increased from 0.0007 to 0.0134 and from 0.0114 to 0.0387, while the 1/n
10
values decreased from 1.1328 to 0.8012 and from 0.7035 to 0.4851, respectively,
11
indicating that at higher temperature, adsorptoin of the substrates are favored. As also
12
illustrated in Table 2, all values of 1/n in both AO7 and MB are less than 1 (0 < 1/n < 1),
13
indicating favorable adsorption. In addition, the lowest 1/n values at 30 oC reveal that
14
the adsorbate-adsorbent interactions might be strong (Ai et al., 2011; Hameed et al.,
15
2007). The adsorption mechanism is believed to be endothermic because the
16
interactions between the adsrobent and the adsrobate are enhanced at higher
17
temperatures. The maximum adsorption capacities of AO7 and MB obtained from the
18
Sips model gradually increased from 500.5 to 665.9 mg/g and 553.2 to 986.8 mg/g,
19
respectively. These values are superior to those of many other adsorbents and even
20
comparable to powdered adsorbents reported in the literature (Table 3).
21 22 23
3.5. Isosteric heat of adsorption Owing to its usefulness for evaluating the strength of interaction between the
24
adsorbate molecules and adsorbent lattice atoms, which is viable as a measure of the
25
energetic heterogeneity of a solid surface, the isosteric heat of adsorption and its
1
variation with surface loading has been generally adopted (Himeno et al., 2005). Thus,
2
in order to identify the underlying factors in the increasing trend in adsorption capacities
3
for AO7 and MB, the isosteric heat of AO7 and MB adsorptions were evaluated by the
4
Clausius-Clapeyron equation (Eq. 10) as functions of the adsorbed amount
5
(0.00068669-9.0789 mol/kg for AO7 and 0.0000057825-14.0364 mol/kg for MB) at
6
different temperatures. As depicted in Figs. 4(A) and (B), isosteric heat of adsorption
7
(Qst) varies with an increase in adsorbed amount, indicating that SCG-GAC has an
8
energetically heterogeneous surfaces. It is also observed that the isosteric heat of
9
adsorption values decreased from 130.5050 to 58.2439 kJ/mol for AO7 and 174.5457 to
10
57.0930 kJ/mol for MB at lower surface loading, whereas a substantial increase was
11
detected as the loading was increased, ranging from 58.2442 to 206.0887 kJ/mol for
12
AO7 and 57.0947 to 281.3331 kJ/mol for MB, respectively. From these observations, it
13
can be concluded that surface energetic heterogeneity and/or adsorbate-adsorbate
14
interaction initially took place at low surface coverage when adsorbing AO7 and MB
15
from aqueous solutions, while the interactions between adsorbate-adsorbent become
16
dominant as the loading increased, resulting in high isosteric heat of adsorption values
17
(Himeno et al., 2005).
18 19 20
3.6. Regeneration test Since reusability of a spent adsorbent is one of the most crucial factors when
21
assessing the cost-effectiveness of the overall processes, the regeneration of exhausted
22
SCG-GAC was carried out. As shown in Fig. 5, the adsorption performance of SCG-
23
GAC gradually decreased with increasing regeneration cycles as a result of incomplete
24
and/or partial recovery of available active sites after the regeneration process.
25
Nevertheless, it is clear that the removal efficiencies or the adsorption capacity were
1
maintained at certain levels (70.2% and 51.6 mg/g for AO7 and 75.4% and 57.1 mg/g
2
for MB) even after seven cycles. These values are comparable to those of various fresh
3
adsorbents (Table 3), demonstrating that SCG-GAC offers sustainability for treatment
4
of dye-containing wastewater. Generally, owing to their simplicity and effectiveness,
5
continuous flow column-type adsorption processes have been employed to ensure
6
practical applicability of adsorbents in the real field. From this point of view, research
7
on the adsorption of AO7 and MB onto SCG-GAC under a continuous flow fixed-bed
8
column condition with a detailed investigation of various operating parameters (solution
9
pH, bed height, flow rate, initial dye concentration) and their prediction for column
10
breakthrough is currently underway.
11 12 13
4. Conclusion This study describes the preparation of SCG-GAC and its application to AO7 and
14
MB adsorption from an aqueous solution. The results indicated that adsorption was
15
highly dependent on the solution pH with respect to the binding affinity between SCG-
16
GAC and dye molecules, while it was independent of ionic effects. Furthermore, the
17
acquired experimental results of dye adsorption suggest that the surface of SCG-GAC is
18
energetically heterogeneous and the adsorption process is governed by an endothermic
19
mechanism. Finally, regeneration results clearly indicated that SCG-GAC is a highly
20
potent adsorbent that can simultaneously meet the requirements of environmental and
21
economic benefits.
22 23
Acknowledgement
24
This work was supported by grants from the National Research Council of Science and
25
Technology (Project No. Asia-02-002).
1 2
References
3
1. Ai, L., Li, M., Li. L., 2011. Adsorption of methylene blue from aqueous solution
4
with activated carbon/cobalt ferrite/alginate composite beads: Kinetics, isotherms,
5
and thermodynamics. J. Chem. Eng. Data 56, 3475-3483.
6
2. Alosmanov, R.M., 2016. Adsorption of arsenazo (III) by phosphorus-containing
7
polymer sorbent. J. Serb. Chem. Soc. Doi: 10.2298/JSC151008042A.
8
3. Baccar, R., Blánquez, P., Bouzid, J., Feki, M., Attiya, H., Sarrà, M., 2013.
9
Modeling of adsorption isotherms and kinetics of a tannery dye onto an activated
10
carbon prepared from an agricultural by-product. Fuel Process. Technol. 106, 408-
11
415.
12
4. Bedin, K.C., Martins, A.C., Cazetta, A.L., Pezoti, O., Almeida, V.C., 2016. KOH-
13
activated carbon prepared from sucrose spherical carbon: Adsorption equilibrium,
14
kinetic and thermodynamic studies for Methylene Blue removal. Chem. Eng. J. 286,
15
476-484.
16
5. Benhouria, A., Islam, Md. A., Zaghouane-Boudiaf, H., Boutahala, m., Hameed,
17
B.H., 2015. Calcium alginate-bentonite-activated carbon composite beads as highly
18
effective adsorbent for methylene blue. Chem. Eng. J. 270, 621-630.
19
6. Doğan, M., Abak, H., Alkan, M., 2009. Adsorption of methylene blue onto
20
hazelnut shell: Kinetics, mechanisms and activation parameters. J. Hazard. Mater.
21
164, 172-181.
22
7. Elmoubarki, R., Mahjoubi, R.Z., Tounsadi, H., Moustadraf, J., Abdennouri, M.,
23
Zouhri, A., Albani, A.E., Barka, N., 2015. Adsorption of textile dyes on raw and
24
ecanted Moroccan clays: Kinetics, equilibrium and thermodynamics. Water Resour.
25
Ind. 9, 16-29.
1 2 3
8. Franca, A.S., Oliveira, L.S., Ferreira, M.E., 2009. Kinetics and equilibrium studies of methylence blue adsorption by spent coffee grounds. Desalination 249, 267-272. 9. Hameed, B.H., Din, A.T.M., Ahmad, A.L., 2007. Adsorption of methylene blue
4
onto bamboo-based activated carbon: Kinetics and equilibrium studies. J. Hazard.
5
Mater. 141, 819-825.
6
10. Himeno, S., Komatsu, T., Fujita, S., 2005. High-pressure adsorption equilibria of
7
methane and carbon dioxide on several activated carbons. J. Chem. Eng. Data 50,
8
369-376.
9
11. Jin, X., Jiang, M.Q., Shan, X.Q., Pei, Z.G., Chen, Z., 2008. Adsorption of
10
methylene blue and orange II onto unmodified and surfactant-modified zeolite. J.
11
Colloid Interf. Sci. 328, 243-247.
12
12. Jung, K.W., Jeong, T.U., Kang, H.J., Ahn, K.H., 2016. Characteristics of biochar
13
derived from marine macroalgae and fabrication of granular biochar by entrapment
14
in calcium-alginate beads for phosphate removal from aqueous solution. Bioresour.
15
Technol. 211, 108-116.
16
13. Jung, K.W., Hwang, M.J., Jeong, T.U., Chau, D.M., Kim, K., Ahn, K.H., 2016.
17
Entrapment of powdered drinking water treatment residues in calcium-alginate
18
beads for fluoride removal from actual industrial wastewater. J. Ind. Eng. Chem. In
19
press.
20
14. Lee, J.W., Choi, S.P., Thirubenkatachari, R., Shi, W.G., Moon, H., 2006.
21
Evaluation of performance of adsorption and coagulation processes for the
22
maximum removal of reactive dyes. Dyes Pigments 69, 196-203.
23
15. Li, G., Zhu, W., Zhang, C., Zhang, S., Liu, L., Zhu, L., Zhao, W., 2016. Effect of a
24
magnetic field on the adsorptive removal of methylene blue onto wheat straw
25
biochar. Bioresour. Technol. 206, 16-22.
1
16. Li, W.H., Yue, Q.Y., Gao, B.Y., Ma, Z.H., Li, Y.J., Zhao, H.X., 2011. Preparation
2
and utilization of sludge-based activated carbon for the adsorption of dyes from
3
aqueous solutions. Chem. Eng. J. 171, 320-327.
4
17. Liou, T.H., 2010. Development of mesoporous structure and high adsorption
5
capacity of biomass-based activated carbon by phosphoric acid and zinc chloride
6
activation. Chem. Eng. J. 158, 129-142.
7
18. Ma, J., Huang, D., Zou, J., Li, L., Kong, Y., Komarneni, S., 2015. Adsorption of
8
methylene blue and Orange II pollutants on activated carbon prepared from banana
9
peel. J. Porous Mater. 22, 301-311.
10
19. Ma, X., Ouyang, F., 2013. Adsorption properties of biomass-based activated carbon
11
prepared with spent coffee grounds and pomelo skin by phosphoric acid activation.
12
App. Surf. Sci. 268, 566-570.
13
20. McKay, G., Mesdaghinia, A., Nasseri, S., Hadi, M., Aminabd, M.S., 2014.
14
Optimum isotherms of dyes sorption by activated carbon: Fractional theoretical
15
capacity & error analysis. Chem. Eng. J. 251, 236-247.
16 17 18
21. Murthy, P.S., Naidu, M.M., 2012. Sustainable management of coffee industry byproducts and value addition - A review. Resour. Conserv. Recy. 66, 45-58. 22. Nourmoradi, H., Ghiasvand, A.R., Noorimotlagh, Z., 2015. Removal of methylene
19
blue and acid orange 7 from aqueous solutions by activated carbon coated with zinc
20
oxide (ZnO) nanoparticles: equilibrium, kinetic, and thermodynamic study. Desalin.
21
Water Treat. 55, 252-262.
22 23
23. Özdemir, Y., Doğan, M., Alkan, M., 2006. Adsorption of cationic dyes from aqueous solutions by spepiolite. Microporous Mesoporous Mater. 96, 419-427.
1
24. Papageorgiou, S.K., Kouvelos, E.P., Favva, E.P., Sapalidis, A.A., Romanos, G.E.,
2
Katsaros, F.K., 2010. Metal-carboxylate interactions in metal-alginate complexes
3
studied with FTIR spectroscopy. Carbohyd. Res. 345, 469-473.
4 5 6
25. Rafatullah, M., Sulaiman, O,m hashim, R., Ahmad, A., 2010. Adsorption of methylene blue on low-cost adsorbents: A review. J. Hazard. Mater. 177, 70-80. 26. Reddy, P.M.K., Mahammadunnisa, Sk., Ramaraju, B., Sreedhar, B.,
7
Subrahmanyam, Ch., 2013. Low-cost adsorbents from bio-waste for the removal of
8
dyes from aqueous solution. Environ. Sci. Pollut. Res. 20, 4111-4124.
9
27. Reffas, A., Bernardet, V., David, B., Reinert, L., Lehocine, M.B., Dubois, M.,
10
Batisse, N., Duclaux, L., 2010. Carbons prepared from coffee grounds by H3PO4
11
activation: Characterization and adsorption of methylene blue and Nylosan Red N-
12
2RBL. J. Hazard. Mater. 175, 779-788.
13
28. Safarik, I., Horska, K., Svobodova, B., Safarikova, M., 2012. Magnetically
14
modified spent coffee grounds for dyes removal. Eur. Food Res. Technol. 234, 345-
15
350
16 17 18
29. Sheha, R.R., El-Khouly, S.H., 2013. Adsorption and diffusion of cesium ions in zirconium (IV) iodomolybdate exchanger. Chem. Eng. Res. Des. 91, 942-954. 30. Silva, J.P., Sousa, S., Rodrigues, J., Antunes, H., Porter, J.J. Gonçalves, I., Ferreira-
19
Dias, S., 2004. Adsorption of acid orange 7 dye in aqueous solutions by spent
20
brewery grains. Separ. Purif. Technol. 40, 309-315.
21
31. Sun, L., Chen, D., Wan, S., Yu, Z., 2015. Performance, kinetics, and equilibrium of
22
methylene blue adsorption on biochar derived from eucalyptus saw dust modified
23
with citric, tartaric, and acetic acids. Bioresour. Technol. 198, 300-308.
24 25
32. Uddin, M.T., Islam, M.A., Mahmud, S., Rukanuzzaman, M., 2009. Adsorptive removal of methylene blue by tea waste. J. Hazard. Mater. 164, 53-60.
1
33. Unur, E., 2013. Functional nanoporous carbons from hydrothermally treated
2
biomass for environmental purification. Microporous Mesoporous Mater. 168, 92-
3
101.
4
34. Zhao, X., Bu, X., Wu, T., Zheng, S.T., Wang, L., Feng, P., 2013. Selective anion
5
exchange with nanogated isoreticular positive metal-organic frameworks. Nat.
6
Commun. 4:2344, doi: 10.1038/ncomms3344.
7
35. Zille, A., Tzanov, T., Guebitz, G.M., Cavaco-Paulo, A., 2003. Immobilized laccase
8
for decolourization of Reactive Black 5 dyeing effluent, Biotechnol. Lett. 25, 1473-
9
1477.
10
1 2
Fig. 1. Effects of solution pH (A) and ionic strength (B) on the adsorption of AO7 and
3
MB onto SCG-GAC. Initial AO7 and MB concentrations, 200 mg/L; solution pH range,
4
3.0-11.0 (ionic strength: pH 3.0 and 11.0 for AO7 and MB, respectively); NaCl
5
concentration range, 0.5-3.0 g/L; temperature, 20 oC; SCG-GAC dosage, 0.1 g; working
6
volume, 50 mL; contact time, 24 h.
Fig. 2. Concentration decay curves of AO7 (A) and MB (B) on the SCG-GAC at different temperatures. Initial dyes concentrations, 2000 mg/L, solution pHs, 3.0 (AO7) and 11.0 (MB); temperature range, 10-30 oC; SCG-GAC dose, 0.5 g; working volume, 250 mL; reaction time, 24 h.
Fig. 3. Adsorption equilibrium isotherms for AO7 (A) and MB (B). Initial dyes concentration range, 50-5000 mg/L; solution pHs, 3.0 (AO7) and 11.0 (MB); temperature range, 10-30 oC; SCG-GAC dosage, 0.1 g; working volume, 50 ml; contact time, 24 h.
Fig. 4. Isosteric heat of adsorption for AO7 (A) and MB (B).
1 2
Fig. 5. Regeneration tests for the AO7 and MB adsorptions onto SCG-GAC. Initial dyes
3
concentration, 200 mg/L; adsorption pHs, 3.0 (AO7) and 11.0 (MB); temperature, 20 oC;
4
adsorbents dose, 0.1 g; working volume, 50 mL; contact time, 24 h; desorption
5
conditions, 200 mL of methanol at 20 oC for 24 h.
6 7
1
Table 1 Kinetic parameters for AO7 and MB adsorptions onto SCG-GAC. Temperature (oC) Parameters
AO7
MB
Units 10
20
30
kf
m/sec
2.228 × 10 -6
2.381 × 10 -6
3.101 × 10 -6
Dp
m2/sec
6.124 × 10-10
9.432 × 10-10
1.356 × 10 -9
R2
-
0.991
0.992
0.991
kf
m/sec
2.280 × 10 -7
2.414 × 10 -6
4.291 × 10 -6
Dp
m2/sec
7.823 × 10-10
1.243 × 10 -9
2.443 × 10 -9
R2
-
0.989
0.990
0.988
Table 2 Adsorption isotherm parameters for AO7 and MB adsorptions onto SCG-GAC. AO7 Temperature (oC)
Parameters
Langmuir
Freundlich
Sips
MB
10
20
30
10
20
30
Qm
525.9763
573.7681
622.7023
471.5177
556.184
667.8028
KL
0.0014
0.0024
0.0058
0.0029
0.0047
0.0062
R2
0.9968
0.9962
0.9938
0.9861
0.984
0.9685
RMSE
1.093E+01
1.295E+01
1.953E+01
2.056E+01
2.721E+01
4.645E+01
SSE
1.133E+02
1.682E+02
3.814E+02
4.226E+02
7.401E+02
2.158E+03
χ2
9.832E+02
1.668E+03
3.723E+03
3.272E+03
5.783E+03
1.830E+04
KF
19.3872
32.6082
63.075
34.0668
57.417
79.3159
1/n
0.3784
0.3370
0.2785
0.3091
0.2737
0.2603
R2
0.9282
0.9354
0.9396
0.9571
0.9585
0.9801
RMSE
5.005E+01
5.368E+01
6.104E+01
3.613E+01
4.384E+01
3.691E+01
SSE
2.505E+03
3.325E+03
3.725E+03
1.306E+03
1.921E+03
1.362E+03
χ2
2.157E+04
2.488E+04
3.229E+04
1.451E+04
2.064E+04
1.509E+04
Qm
500.4862
588.8762
665.8587
553.196
653.1532
986.8108
KS
0.0007
0.0033
0.0134
0.0114
0.0197
0.0387
1/n
1.1328
0.9362
0.8012
0.7035
0.6653
0.4851
R2
0.9979
0.9966
0.9968
0.9972
0.9951
0.994
RMSE
8.581E+00
1.239E+01
1.395E+01
9.300E+00
1.508E+01
2.027E+01
SSE
7.335E+01
1.598E+02
1.946E+02
8.646E+01
5.204E+02
4.106E+02
χ2
8.012E+02
1.495E+03
1.894E+03
1.517E+03
2.506E+03
4.257E+03
1
Table 3 Comparison of maximum dye adsorption capacity obtained in this study with
2
previous data. Adsorbent
Adsorption capacity (mg/g)
Dye
References
SCG-GAC
665.9
This study
Spent brewery grains
30.5
Silva et al., 2004
Surfactant-modified zeolite
0.63
Banana peel-activated carbon
333
Activated carbon coated with ZnO
66.22
SCG-GAC
986.8
This study
Surfactant-modified zeolite
15.68
Jin et al., 2008
Tea waste
85.16
Activated carbon derived from coffee
AO7
Ma et al., 2015 Nourmoradi et al., 2015
Uddin et al., 2009 Reffas et al.,
181.8
grounds Activated carbon/cobalt
33.58
ferrite/alginate composite beads Banana peel-activated carbon
1,263
Calcium alginate-bentonite-activated
994.06
carbon composite beads Activated carbon coated with ZnO
66.66
Activated carbon prepared from
704.2
sucrose spherical carbon 3 4
36
Jin et al., 2008
2010 MB
Ai et al., 2011 Ma et al., 2015 Benhouria et al., 2015 Nourmoradi et al., 2015 Bedin et al., 2016
1
Graphical Abstract
2 3
37
1
- We studied adsorption of acid orange 7 and methylene blue from aqueous solutions.
2
- Spent coffee grounds has been converted into granular activated carbon (SCG-GAC).
3
- Pore diffusion and Sips models fit the adsorption data of the SCG-GAC.
4
- Maximum adsorption capacities were 665.9 mg/g for AO7 and 986.8 mg/g for MB.
5
- Regeneration tests show the saturated SCG-GAC has excellent reusability potential.
6 7 8
38