You are on page 1of 6

Chemosphere 190 (2018) 97e102

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Chromium (Ⅵ) removal from aqueous solutions through powdered


activated carbon countercurrent two-stage adsorption
Wenqiang Wang
College of Chemistry and Chemical Engineering, Dezhou University, Dezhou 253023, China

h i g h l i g h t s

 The PAC countercurrent two-stage adsorption process was performed.


 The Cr(VI) removal efficiency was improved by > 50% by the process.
 A calculation method correlating the effluent Cr(VI) concentration and PAC dose was feasible.

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

Article history: To exploit the adsorption capacity of commercial powdered activated carbon (PAC) and to improve the
Received 1 June 2017 efficiency of Cr(VI) removal from aqueous solutions, the adsorption of Cr(VI) by commercial PAC and the
Received in revised form countercurrent two-stage adsorption (CTA) process was investigated. Different adsorption kinetics
29 July 2017
models and isotherms were compared, and the pseudo-second-order model and the Langmuir and
Accepted 25 September 2017
Freundlich models fit the experimental data well. The Cr(VI) removal efficiency was >80% and was
Available online 28 September 2017
improved by 37% through the CTA process compared with the conventional single-stage adsorption
Handling Editor: Xiangru Zhang process when the initial Cr(VI) concentration was 50 mg/L with a PAC dose of 1.250 g/L and a pH of 3. A
calculation method for calculating the effluent Cr(VI) concentration and the PAC dose was developed for
Keywords: the CTA process, and the validity of the method was confirmed by a deviation of <5%.
Chromium © 2017 Published by Elsevier Ltd.
Powered activated carbon
Countercurrent two-stage adsorption

1. Introduction A variety of techniques have been developed to remove Cr(VI),


including adsorption (Di Natale et al., 2015; Sun et al., 2015;
Cr(VI) is classified as a carcinogen by the International Agency Demiral et al., 2008), precipitation (Gheju and Balcu, 2011; Golder
for Research on Cancer due to its strong mutagenic and teratogenic et al., 2011), membrane separation (Gherasim and Bourceanu,
properties (De Flora, 2000), and the US Environmental Protection 2013; Bao et al., 2015) or solvent extraction (Wionczyk et al.,
Agency has established 0.1 mg/L as the maximum total Cr con- 2011) combined with adsorption and ion exchange on clays
centration in drinking water (Costa, 2003). Currently, the wide- (Bhattacharyya and Gupta, 2008), biomaterials (Zhu et al., 2016;
spread use of Cr(VI) and its compounds in different industries, such Albadarin et al., 2011; Mohan et al., 2011; Mohanty et al., 2006),
as mining operations, electroplating, leather tanning, water cool- nanocomposites (Kara and Demirbel, 2012; Gu et al., 2012), ex-
ing, power generation and petroleum refining, has led to numerous change resins (Polowczyk et al., 2016; Kowalczyk et al., 2013), and
diseases and serious environmental pollution (Cao et al., 2014). It biological processes (Chen and Gu, 2005). Most of these techniques
has been reported that long-term exposure to Cr(VI) ions increases have limitations such as high capital and operational costs, as well
the risk of dermatitis; damage to the liver, kidney circulation, and as issues with the disposal of residual metal sludge (Sharma and
nerve tissue; and death (Kotas and Stasicka, 2000). Therefore, there Forster, 1994). In contrast, adsorption technologies for Cr(VI)
is a need to develop a more effective methodology to control the removal have been investigated in both laboratory and industrial
discharge of Cr(VI) into the environment. settings due to the high adsorption capacities, operational conve-
nience, versatility, and range of the available compounds (Min et al.,
2015). Natural minerals (Swas and Uysal, 2014); artificial meso-
porous silica materials; various non-commercial, self-synthesized
E-mail address: wwqzs@126.com.

https://doi.org/10.1016/j.chemosphere.2017.09.141
0045-6535/© 2017 Published by Elsevier Ltd.
98 W. Wang / Chemosphere 190 (2018) 97e102

polymer or copolymer adsorbents (Illanes et al., 2008; Sun et al., where V (L) is the volume of Cr(VI) solution, C0 and Ce (mg/L) are
2013; Wang et al., 2015); many agricultural wastes; and different the initial and final concentrations of Cr(VI) in solution, respec-
modified activated carbons had been used as sorbents for Cr(VI) tively, and W (g) is the weight of adsorbent. All results were re-
adsorption from wastewater. However, the largest obstacle to the ported as the average values for duplicate experiments.
use of these materials is their inability to meet commercial demand
due to high consumption. As a result, industries have depended on 2.3. CTA experiments
commercial activated carbons (Anupam et al., 2011). Hence,
another way to reduce the cost of Cr(VI) removal from wastewater The CTA experiments were batch processed using a magnetic
through adsorption is to explore the adsorption capacity and stirring pressure cup with an MF membrane disc; the adsorption
reduce the dose of these carbons. device is shown in Fig. S1. To ensure that the MF membrane was
It is commonly recognized that countercurrent adsorption can immersed, a portion of the equilibrium effluent remained in the
improve the adsorption capacity of activated carbon, but powdered pressure cup.
activated carbon (PAC) cannot be used in a traditional column Fig. 1 shows a schematic of the CTA process. During the initiation
countercurrent adsorption operation, which requires granular stage, the V ¼ 0.2 L Cr(VI) solution with an initial Cr(VI) concen-
activated carbon. However, when combining PAC adsorption with a tration of C0 ¼ 50 mg/L and a fresh PAC of 0.2 g corresponding to a
microfiltration (MF) membrane, countercurrent adsorption opera- PAC dose of m=V ¼ 1.0 g/L were sequentially added into the pres-
tion can be achieved. The effectiveness of a PAC countercurrent sure cup for 30 min of adsorption while stirring at 293 K. Next,
multi-stage adsorption/MF hybrid process has been demonstrated 0.8Vequilibrium effluent was discharged, and 0.2Veffluent
by the removal of organics from a reverse osmosis concentrate remained to ensure MF membrane immersion. Afterward, the
(Zhao et al., 2012; Wang et al. (2013b); Wei et al., 2014). operational stage experiments were repeated.
The objective of this study was to explore the countercurrent During the operational stage, the 0.8VCr(VI) solution was added
two-stage adsorption (CTA) process for the removal of Cr(VI) from into the pressure cup to form a mixture with the 0.2Veffluent
aqueous solutions using commercial PAC. At the same time, the (Cr(VI) concentration of Ce2 ) remaining after the last effluent
method for calculating the PAC dose and effluent Cr(VI) concen- cycle. The mixture, which had a Cr(VI) concentration of
tration was determined and validated. =
C0 ¼ 0:2Ce2 þ 0:8C0 , stayed in contact with the loaded PAC
remaining in the pressure cup after the last effluent cycle for 30 min
2. Materials and methods
to achieve a Cr(VI) concentration of Ce1 ; this was considered second
stage adsorption, during which, fresh PAC was not added. After the
2.1. Materials
second adsorption, the fresh PAC of m=V ¼ 1.0 g/L was added into
the pressure cup for 30 min of adsorption to further reduce the
The materials used for the experiments were potassium di-
Cr(VI) concentration from Ce1 to Ce2 ; this was considered first stage
chromate (K2Cr2O7), sulphuric acid, hydrochloric acid, phosphoric
adsorption. Ce2 was the final effluent Cr(VI) concentration of the
acid, sodium hydroxide, and acetone obtained from Sinopharm
CTA process.
Chemical Reagent Co. (Shanghai, China) and sodium hydroxide
During the CTA process, the PAC experienced two-stage
supplied by Hengxing Chemical Reagent Co. (Tianjin, China). The
adsorption, and its adsorption capacity was sufficiently explored.
commercial PAC (200 mesh) was purchased from the Luda Co.
For the same PAC dose, the CTA process can be expected to remove
(Zunhua, China), and more information about the PAC is described
more Cr(VI) than the conventional single-stage adsorption (CSA)
in the literature (Zhao et al., 2012). Prior to use, the PAC was baked
process. The average values for two parallel experiments were
in an oven at 105  C for 2 h to remove any moisture. All chemical
reported.
regents were of analytical grade, the Cr(VI) solution with an initial
concentration of 50 mg/L was prepared by dissolving potassium
dichromate with distilled water. 3. Results and discussion

2.2. Batch adsorption experiment 3.1. Adsorption kinetics

The influences of contact time (0e90 min, at a pH of 4 and a The adsorption kinetics provide valuable information regarding
PAC dose of 1.0 g/L), initial pH (2.0e11.0, for a contact time of the equilibrium time and the adsorption mechanism along with the
30 min and a PAC dose of 1.0 g/L), and PAC dose (0.2e2.2 g/L, at a reaction pathways. Fig. 2 depicts the effluent Cr(VI) concentration
pH of 3 and a contact time of 30 min) on Cr(VI) removal in a as a function of adsorption time; the effluent concentration of
batch mode of operation under constant conditions were recor- Cr(VI) rapidly decreased at the initial stage with a high adsorption
ded (a shaker speed of 200 rpm, a temperature of 293 K and an rate due to the availability of large number of adsorption sites
initial Cr(VI) concentration of 50 mg/L). The pH values of the which saturates with time and tended to remain almost constant
adsorbate solutions were controlled by adding 0.1 M hydro- after an optimum adsorption time of 30 min. Therefore, the equi-
chloric acid or sodium hydroxide. Once equilibrium was attained, librium time for Cr(VI) removal was 30 min.
the adsorbent was separated from the solution though a 0.45 mm The pseudo-first-order kinetics is one of the most used models
filter membrane, and the equilibrium concentration was deter- to describe adsorption from aqueous solution by solid adsorbent.
mined at the wavelength for maximum adsorbance The pseudo-second-order model based on the assumption that the
(lmax ¼ 540 nm) by the 1,5-diphenyl carbazide method using a rate limiting step might be chemical adsorption involving valence
UVevis spectrophotometer (752, Shanghai Jinghua Instruments forces through sharing or exchange of electrons between adsorbate
Co., China). and adsorbent. The intra-particle diffusion model is of major
The amount of Cr(VI) per unit weight of adsorbent, qe (mg/g) concern because it is the rate-determining step in the liquidesolid
was calculated by the following equation: adsorption systems. During the batch mode of operation, there is a
possibility of transport of adsorbate species into the pores of
VðC0  Ce Þ adsorbent, which is often a rate-controlling step. Table S1 displays
qe ¼ (1) the parameters calculated from the three models and the kinetic
W
W. Wang / Chemosphere 190 (2018) 97e102 99

Fig. 1. Schematic diagram of the CTA process.

plots (Fig. S2). Conformity between experimental data and the adsorption sites and repulsive forces between the Cr(VI) ions on the
model-predicted values was expressed by the correlation coeffi- solid and bulk phases had a low diffusion rate.
cient (R2 ). A relatively high R2 value indicated that the model
successfully described the kinetics of Cr(VI) adsorption. Through 3.2. Effects of solution pH on Cr(VI) adsorption
comparing R2 of these three models, it was determined that the
pseudo-second-order model fit the kinetics data better than the The pH of the aqueous solution is an important parameter for
other two models. The theoretical qe values calculated from the the removal of heavy metals because it affects both the surface
first- and second-order kinetic models were 12.74 mg/g and charge of an adsorbent and the speciation of a heavy metal, so the
20.37 mg/g, respectively, and the calculated adsorption capacity heavy metal adsorption process differs with the pH level (Liu et al.,
(qe;cal ) from the pseudo-second-order-kinetic model was very close 2012). The effluent Cr(VI) concentration, with an initial value of
to the experimental adsorption capacity (qe;exp ) of 20.13 mg/g, 50 mg/L, over a pH range of 2.0e11.0 is depicted in Fig. 3, it is clear
indicating that the adsorption process followed the pseudo- that the adsorption of Cr(VI) onto the commercial PAC is signifi-
second-order kinetic equation. This result also demonstrates that cantly influenced by the solution pH and that the Cr(VI) concen-
the chemical process controls the adsorption of Cr(VI) onto the tration decreases as pH decreases. There was a sharp decrease in
commercial PAC, which occurs at high Cr(VI) concentrations and the effluent concentration of Cr(VI) when pH decreased from 5.0 to
when the interactions of the ions and the active sites on the surface 2.0, but the effects were negligible over a pH range of 6.0e11.0.
are effective. The intra-particle-diffusion model is always used to Cr(VI) exists in three different ionic forms (HCrO 2
4 , Cr2O7 , and
comprehensively illustrate the adsorbate diffusion mechanism and CrO2
4 ), and the species ratio depends on the pH of the solution
determine whether intra-particle-diffusion is the rate-limiting system (Singh and Tiwari, 1997). In an acidic solution, Cr(VI) pre-
factor controlling the adsorption process especially in a solid- dominately exists as HCrO 2
4 and is transformed into CrO4 when
liquid adsorption system. R2 < 0.96 indicated that the intra- the pH increases. The concentration of Cr(VI) decreased in the
particle-diffusion model was not the only rate-limiting step, and lower pH ranges due to a high electrostatic force of attraction. The
a multi-step process is expected to work during the entire time surfaces of the adsorbent were negatively charged at pH > pHpzc
range. The large kP1 value of the intra-particle diffusion model (pzc-point of zero charge) and positively charged at pH < pHpzc
suggests that the commercial PAC initially adsorbed the Cr(VI) (Sun et al., 2014). At lower pH (pH < pHpzc), the surface of the
though the external surface adsorption caused by capillary effect or adsorbent was positively charged, and at the same time more Hþ
the rate-controlled intra-particle diffusion process at a high diffu- ions are neutralized by the negative charge on the adsorbent sur-
sion rate, but the small kP2 value indicates that the final equilibrium face. Due to the high electrostatic force of attraction between the
stage where intra-particle diffusion slows down due to limited positively charged adsorbent and the negatively charged chromate

Fig. 2. Effect of adsorption time on Cr(VI) removal by commercial PAC (PAC


dose ¼ 1.0 g/L; pH ¼ 4; initial Cr(VI) concentration ¼ 50 mg/L; and Fig. 3. Effect of pH on Cr(VI) removal by the commercial PAC (PAC dose ¼ 1.0 g/L;
temperature ¼ 293 K). initial Cr(VI) concentration ¼ 50 mg/L; and temperature ¼ 293 K).
100 W. Wang / Chemosphere 190 (2018) 97e102

ions, the diffusion of chromate ions into the bulk of the adsorbent 3.4. CTA process
increases (Rao et al., 2002). In contrast, at alkaline pH ranges
(pH > pHpzc), the surface of the adsorbent was negatively charged, 3.4.1. Calculation method
and the electrostatic repulsion between the negatively charged A calculation method to predict the effluent Cr(VI) concentra-
surface of the adsorbent and the negatively charged Cr(VI) ions, as tion according to the PAC dose of the CTA process was developed
well as the competition between the Cr(VI) ions and OH ions in based on the adsorption isotherm equation and the accumulative
the aqueous solution for the adsorption sites, reduced the removal adsorption principle (Zhang et al., 2009). If PAC desorption did not
of Cr(VI) (Sun et al., 2014). The lowest concentration of Cr(VI) occur, then the adsorption capacity would be cumulative in
occurred at a pH of 2.0, but there was only a minute decrease in the accordance with the isotherm equation. The multi-stage adsorption
concentration of Cr(VI) from pH 3.0 to pH 2.0. Therefore, a pH of 3.0 capacity of PAC can be calculated appropriately using Eq. (2).
was considered the optimal value for further studies of Cr(VI)
adsorption on the commercial PAC. C0  Cej
qej ¼ qeðj1Þ þ ð j > 1Þ (2)
m=V
3.3. Adsorption isotherms
where qej and qeðj1Þ are the PAC adsorption capacity for j stages and
Isotherms for the adsorption of Cr(VI) onto the commercial PAC j  1 stages (mg/g), respectively; C0 and Cej are the initial and
showed how Cr(VI) interacted with carbon as well as the rela- equilibrium adsorbate concentration (mg/L); m=V is the PAC dosage
tionship between the amounts of adsorbed Cr(VI) (qe ) and its (g/L). In addition, the variations in isotherm parameters due to the
equilibrium concentration (Ce ) in solution; these parameters were different initial adsorbate concentrations were ignored to simplify
significant for the application of adsorbents. Three commonly used the calculation (Zhao et al., 2012). The adsorption isotherm equa-
isotherm models, the Langmuir, Freundlich, and Temkin isotherms, tions obtained with the initial Cr(VI) concentration of C0 were used
were applied for the data analysis. The Langmuir model assumes to calculate the equilibrium Cr(VI) concentration with the initial
that the adsorption of adsorbate molecules occurs on a homoge- =
Cr(VI) concentration of C0 or Ce1.
neous surface by monolayer adsorption without any interaction
As shown in Fig. 1 for first stage adsorption, when 1.0 g/L fresh
between the adsorbed molecules (Deng et al., 2009). The Freund-
PAC was added into the pressure cup to adsorb Cr(VI) for 30 min,
lich model is widely applied in multi-layer sorption caused by the
the Cr(VI) concentration was reduced from Ce1 to Ce2 . According to
presence of different functional groups on heterogeneous surfaces,
the Freundlich equation and the parameters (Table 1), Eq. (3) was
and it incorporates various adsorbent-adsorbate interactions to
obtained:
describe the non-ideal and reversible adsorption (Deng et al.,
2009). Finally, the Temkin model describes the uniform distribu- 1=n Ce1  Ce2 1=2:523
tion of binding energies, and the adsorption heat of all the mole- qe ¼ kF Ce ¼ ¼ 7:237Ce2 (3)
1:0
cules in the layer decreases linearly rather than logarithmically
with coverage (Foo and Hameed, 2010). For second stage adsorption, the remaining adsorption capacity
The calculated parameters of the models by the isotherms plots of the loaded PAC after first stage adsorption during the last effluent
=
(Fig. S3) are presented in Table 1. For the commercial PAC, the cycle reduced the concentration of Cr(VI) from C0 ¼ 0:2Ce2 þ 0:8C0
correlation coefficients fit by the Freundlich and Langmuir iso- to Ce1 . The surplus adsorption capacities of the dual-loaded PAC
therms were higher than those fit by the Temkin isotherm, and the were ignored when compared to the fresh PAC. Therefore, the total
correlation coefficient for the Freundlich isotherm was slightly adsorption capacity of the PAC at a dose of m=V ¼ 1.0 g/L for the CTA
higher than that for the Langmuir isotherm. Similar results were process was calculated by Eq. (4), which was deduced from
also reported by Sun (Sun et al., 2014). It is also evident from the Freundlich equation and Eq. (2) following the accumulative
correlation coefficient values that the Freundlich isotherm fitted adsorption principle (Zhang et al., 2009).
the experimental data particularly well, which indicates the het-
erogeneous carbon surface and the multi-layer coverage of Cr(VI). A 0:2Ce2 þ 0:8C0  Ce1 Ce1  Ce2 1=2:523
þ ¼ 7:237Ce1 (4)
value of n larger than 1 indicates that Cr(VI) is favourably adsorbed 1:0 1:0
by the commercial PAC. The maximum adsorption capacity (qm ) Eq. (4) describes the CTA process of the operational stage. The
value, as calculated from the Langmuir adsorption isotherm for second item on the left-hand side of Eq. (4) represents fresh PAC
Cr(VI), was 43.29 mg/g, which was also considerably higher than adsorbing low-concentration Cr(VI), and the first item represents
the maximum adsorption capacity of most reported adsorbents for loaded PAC adsorbing high-concentration Cr(VI). By solving Eqs. (3)
a given pH (Table 2). It appeared that the commercial PAC has and (4), the values of Ce1 and Ce2 were 29.14 mg/L and 10.65 mg/L,
desirable qualities for the removal of Cr(VI). respectively.
Similarly, the calculation method could be based on the Lang-
muir equation and Eq. (2) as shown by Eqs. (5) and (6):
Table 1
Adsorption isotherm model constants for Cr(VI) adsorption onto the commercial
PAC (initial Cr(VI) concentration ¼ 50 mg/L; temperature ¼ 293 K; pH ¼ 3; and Ce1  Ce2 0:0617  43:29Ce2
¼ (5)
adsorption time ¼ 30 min). 1:0 1 þ 0:0617Ce2
Isotherms Parameters Values
0:2Ce2 þ 0:8C0  Ce1 Ce1  Ce2 0:0617  43:29Ce1
Langmuir qm (mg/g) 43.29 þ ¼ (6)
kL (L/mg) 0.0617 1:0 1:0 1 þ 0:0617Ce1
R2 0.9774
By solving Eqs. (5) and (6), the values of Ce1 and Ce2 were
Freundlich kF (mg/g (l/mg)1/n) 7.237
n 2.523 28.08 mg/L and 10.79 mg/L, respectively.
R2 0.9779
Temkin kT (L/mg) 0.5475 3.4.2. Validation of the calculation method
B 9.997
The effluent Cr(VI) concentration could be calculated using the
R2 0.9622
Freundlich or Langmuir equations for the CSA process and by the
W. Wang / Chemosphere 190 (2018) 97e102 101

Table 2
Comparison of maximum Cr(VI) adsorption capacity (qm ) with various adsorbents.

Adsorbents Initial concentrations (mg/L) pH qm (mg/g) References

AC derived from olive bagasse 50e500 2.0 88.59 Demiral et al. (2008)
AC-Fe 5.7e63.3 5.38 34.39 Sun et al. (2014)
Coconut shell carbon 5e25 4.0 10.88 Babel and Kurniawan (2004)
AC derived from waste sawdust 100 4.0 13.85 Malwade et al. (2016)
Agricultural waste-based AC 20e60 1.5 38.5 Chaudhuri and Khairi Bin Azizan (2012)
PAC 250-50,000 4.0 46.9 Jung et al. (2013)
AC prepared from peanut shell 10e100 4.0 16.26 AL-Othman et al. (2012)
AC prepared from crofton weed 100 3.0 36.22 Wang et al. (2013a)
AC developed from tamarind wood 10e50 6.5 28.019 Acharya et al. (2009)
AC developed from date palm seed 20e125 1.0 120.48 Nemr et al. (2008)
PAC 50 3.0 43.29 this study

Fig. 4. Comparison of experimental and calculated effluent Cr(VI) concentration values: (a) Freundlich; (b) Langmuir.

above calculation method for the CTA process. 1.250 g/L. The calculated average effluent Cr(VI) concentrations
The validation of the calculation method and the improvement were 10.65 mg/L (Freundlich) and 10.79 mg/L (Langmuir) for the
in Cr(VI) removal are shown in Fig. 4. The abbreviations used in this CTA process, corresponding to removal efficiencies of 78.70% and
figure, including cal and exp stand for calculated and experimental, 78.42%, respectively. The experimental average Cr(VI) removal ef-
respectively. The PAC dose of m=V ¼ 1.0 g/L in the Cr(VI) mixed ficiency was slightly higher than the calculated averages because
liquor could be converted to the PAC dose of raw Cr(VI) solution by the negligible third or even fourth stage adsorptions were not
m=V  1:0V=0:8V ¼ 1.250 g/L. It is clear that Cr(VI) removal was considered in this study. However, this finding did not affect the
significantly improved by the CTA process compared with the CSA validity of the above calculation method. In terms of removal effi-
process at the same PAC dose of 1.250 g/L. The calculated effluent ciency, deviations of <5% between the experimental average
Cr(VI) concentrations were 20.22 mg/L (Freundlich) and 20.07 mg/L removal efficiency and the calculated averages demonstrated the
(Langmuir) for the CSA process, which corresponded to Cr(VI) validity of the calculation method.
removal efficiencies of 59.56% and 59.86% by isotherm equations at
the PAC dose of 1.250 g/L, respectively. For the CTA process, the
3.4.3. Comparison of PAC consumption
Cr(VI) concentration of the first effluent cycle was higher than that
The advantage of the countercurrent adsorption process is the
of the subsequent effluent cycles due to the absence of loaded PAC
PAC saving compared with sing-stage adsorption at the same
in the pressure cup at the initiation stage, which also affected the
removal efficiency. Therefore, from another perspective, to achieve
Cr(VI) concentration during the second effluent cycle. The experi-
the same effluent quality as the CTA process, i.e., an effluent Cr(VI)
mental effluent Cr(VI) concentration data of the first and second
concentration of 9.81 mg/L, the PAC dose required for the CSA
effluent cycles were excluded, and the average value of the other
process could be calculated using Langmuir or Freundlich equation.
seven effluent cycles was employed. The experimental average
Table 3 shows that the PAC adsorption capacity was more effec-
effluent Cr(VI) concentration for the CTA process was 9.81 mg/L,
tively exploited by the CTA process, so the PAC doses were 49.23%
corresponding to a removal efficiency of 81.95%, that was approx-
(Langmuir) and 44.37% (Freundlich), which are lower than those of
imately 37% higher than that of the CSA for the same PAC dose of
the CSA process. For the same effluent quality, the PAC dose could
be reduced under the CTA process compared to that of the CSA.

Table 3
PAC dose and adsorption capacity for the CSA and CTA processes. 4. Conclusions
Parameters CSA CTA
The present study investigated the removal of Cr(VI) from
Langmuir Freundlich
aqueous solutions by adsorption over commercial PAC, and the
PAC dose (g/L) 2.462 2.247 1.250 adsorption of Cr(VI) was determined to be highly dependent on pH.
Relative PAC dose 1.970 1.798 1.000 The kinetic studies demonstrated that a pseudo-second-order
qe (mg/g) 16.32 17.89 32.15
model could provide a realistic description of Cr(VI) adsorption
102 W. Wang / Chemosphere 190 (2018) 97e102

kinetics. Both the Langmuir and Freundlich models fitted the performance and reaction stoichiometry. Sep. Purif. Technol. 76, 345e350.
Gu, H., Rapole, S.B., Sharma, J., Huang, Y., Cao, D., Colorado, H.A., Luo, Z.,
isotherm data well compared to the Temkin model. With CTA, the
Haldolaarachchige, N., Young, D.P., Walters, B., Wei, S., Guo, Z., 2012. Magnetic
Cr(VI) removal efficiency improved by approximately 37% polyaniline nanocomposites toward toxic hexavalent chromium removal. RSC
compared to the CSA at the same PAC dose of 1.250 g/L. The Adv. 2, 11007e11018.
calculation method for correlating the Cr(VI) removal efficiency and Illanes, C.O., Ochoa, N.A., Marchese, J., 2008. Kinetic sorption of Cr(VI) into solvent
impregnated porous microspheres. Chem. Eng. J. 136, 92e98.
the PAC dose base on the Langmuir and Freundlich models was Jung, C., Heo, J., Han, J., Her, N., Lee, S.-J., Oh, J., Ryu, J., Yoon, Y., 2013. Hexavalent
deduced and validated, and deviations of <5% between the exper- chromium removal by various adsorbents: powdered activated carbon, chito-
imental average removal efficiency and the calculated averages san, and single/multi-walled carbon nanotubes. Sep. Purif. Technol. 106, 63e71.
Kara, A., Demirbel, E., 2012. Kinetic, isotherm and thermodynamic analysis on
demonstrated that the calculation method was acceptable. adsorption of Cr(VI) ions from aqueous solutions by synthesis and character-
ization of magnetic-poly (divinylbenzene-vinylimidazole) microbeads. Water
Air Soil Pollut. 223, 2387e2403.
Acknowledgements
Kotas, J., Stasicka, Z., 2000. Chromium occurrence in the environment and methods
of its speciation. Environ. Pollut. 107, 263e283.
The authors are grateful for the financial support from the Talent Kowalczyk, M., Hubicki, Z., Kołodyn  ska, D., 2013. Removal of heavy metal ions in the
Introduction Subject of Dezhou University (No. 2015kjrc04). presence of the biodegradable complexing agent of EDDS from waters. Chem.
Eng. J. 221, 512e521.
Liu, W., Zhang, J., Zhang, C., Ren, L., 2012. Preparation and evaluation of activated
Appendix A. Supplementary data carbon-based iron-containing adsorbents for enhanced Cr(VI) removal: mech-
anism study. Chem. Eng. J. 189e190, 295e302.
Malwade, K., Lataye, D., Mhaisalkar, V., Kurwadkar, S., Ramirez, D., 2016. Adsorption
Supplementary data related to this article can be found at of hexavalent chromium onto activated carbon derived from Leucaena leuco-
https://doi.org/10.1016/j.chemosphere.2017.09.141. cephala waste sawdust: kinetics, equilibrium and thermodynamics. Int. J. En-
viron. Sci. Technol. 13, 2107e2116.
Min, K.K., Sundaram, K.S., Lyengar, G.A., Lee, K.P., 2015. A novel chitosan functional
References gel included with multiwall carbon nanotube and substituted polyaniline as
adsorbent for efficient removal of chromium ion. Chem. Eng. J. 267, 51e64.
Acharya, J., Sahu, J.N., Sahoo, B.K., Mohanty, C.R., Meikap, B.C., 2009. Removal of Mohan, D., Rajput, S., Singh, V.K., Steele, P.H., Pittman Jr., C.U., 2011. Modeling and
chromium(VI) from wastewater by activated carbon developed from Tamarind evaluation of chromium remediation from water using low cost bio-char, a
wood activated with zinc chloride. Chem. Eng. J. 150, 25e39. green adsorbent. J. Hazard. Mater 188, 319e333.
Albadarin, A.B., Al-Muhtaseb, A.A.H., Al-laqtah, N.A., Walker, G.M., Allen, S.J., Mohanty, K., Jha, M., Meikap, B.C., Biswas, M.N., 2006. Biosorption of Cr(VI) from
Ahmad, M.N.M., 2011. Biosorption of toxic chromium from aqueous phase by aqueous solutions by Eichhornia crassipes. Chem. Eng. J. 117, 71e77.
lignin: mechanism, effect of other metal ions and salts. Chem. Eng. J. 169, Nemr, A.E., Khaled, A., Abdelwahab, O., El-Sikaily, A., 2008. Treatment of wastewater
20e30. containing toxic chromium using new activated carbon developed from date
AL-Othman, Z.A., Ali, R., Naushad, M., 2012. Hexavalent chromium removal from palm seed. J. Hazard. Mater 152, 263e275.
aqueous medium by activated carbon prepared from peanut shell: adsorption Polowczyk, I., Urbano, B.F., Rivas, B.L., Bryjak, M., Kabay, N., 2016. Equilibrium and
kinetics, equilibrium and thermodynamic studies. Chem. Eng. J. 184, 238e247. kinetic study of chromium sorption on resins with quaternary ammonium and
Anupam, K., Dutta, S., Bhattacharjee, C., Datta, S., 2011. Adsorptive removal of N-methyl-D-glucamine groups. Chem. Eng. J. 284, 395e404.
chromium (VI) from aqueous solution over powdered activated carbon: opti- Rao, M., Parwate, A.V., Bhole, A.G., 2002. Removal of Cr6þ and Ni2þ from aqueous
misation through response surface methodology. Chem. Eng. J. 173, 135e143. solution using bagasse and fly ash. Waste Manag. 22, 821e830.
Babel, S., Kurniawan, T.A., 2004. Cr(VI) removal from synthetic wastewater using Sharma, D.C., Forster, C.F., 1994. A preliminary examination into the adsorption of
coconut shell charcoal and commercial activated carbon modified with hexavalent chromium using low-cost adsorbents. Bioresour. Technol. 47,
oxidizing agents and/or chitosan. Chemosphere 54, 951e967. 257e264.
Bao, Y., Yan, X., Du, W., Xie, X., Pan, Z., Zhou, J., Li, L., 2015. Application of amine- Singh, V.K., Tiwari, P.N., 1997. Removal and recovery of chromium(VI) from indus-
functionalized MCM-41 modified ultrafiltration membrane to remove chro- trial waste water. J. Chem. Technol. Biotechnol. 69, 376e382.
mium (VI) and copper (II). Chem. Eng. J. 281, 460e467. Sun, Y.J., Chae, J.W., Jamg, H.D., Cho, K., 2015. Role of chemical hardness in the
Bhattacharyya, K.G., Gupta, S.S., 2008. Adsorption of a few heavy metals on natural adsorption of hexavalent chromium species onto metal oxide nanoparticles.
and modified kaolinite and montmorillonite: a review. Adv. Colloid Interface Chem. Eng. J. 273, 401e405.
Sci. 140, 114e131. Sun, Y., Yue, Q., Guo, B., Gao, Y., Li, Q., Wang, Y., 2013. Adsorption of hexavalent
Cao, J., Wu, Y., Jin, Y., Yilihan, P., Huang, W., 2014. Response surface methodology chromium on Arundo donax Linn activated carbon amine-crosslinked copol-
approach for optimization of the removal of chromium (VI) by NH2-MCM-41. ymer. Chem. Eng. J. 217, 240e247.
J. Taiwan Inst. Chem. Eng. 45, 860e868. Sun, Y., Yue, Q., Mao, Y., Gao, B., Gao, Y., Huang, L., 2014. Enhanced adsorption of
Chaudhuri, M., Khairi Bin Azizan, N., 2012. Adsorptive removal of chromium(VI) chromium onto activated carbon by microwave-assisted H3PO4 mixed with Fe/
from aqueous solution by an agricultural waste-based activated carbon. Water Al/Mn activation. J. Hazard. Mater 265, 191e200.
Air Soil Pollut. 223, 1765e1771. Swas, H., Uysal, T., 2014. Removal of heavy metal ions from aqueous medium using
Chen, Y., Gu, G., 2005. Preliminary studies on continuous chromium(VI) biological Kuluncak (Malatya) vermiculites and effect of precipitation on removal. Appl.
removal from wastewater by anaerobiceaerobic activated sludge process. Bio- Clay Sci. 95, 1e8.
resour. Technol. 96, 1713e1721. Wang, P., Zhang, R., Hua, C., 2013a. Removal of chromium (VI) from aqueous solu-
Costa, M., 2003. Potential hazards of hexavalent chromate in our drinking water. tions using activated carbon prepared from crofton weed. Desalin. Water Treat.
Toxicol. Appl. Pharm. 188, 1e5. 51, 2327e2335.
De Flora, S., 2000. Threshold mechanisms and site specificity in chromium(VI) Wang, T., Zhang, L., Li, C., Yang, W., Song, T., Tang, C., Meng, Y., Dai, S., Wang, H.,
carcinogenesis. Carcinogenesis 21, 533e541. Chai, L., Luo, J., 2015. Synthesis of CoreShell magnetic Fe3O4@poly(m-Phe-
Demiral, H., Demiral, I., Tümsek, F., Karabacakog lu, B., 2008. Adsorption of chro- nylenediamine) particles for chromium reduction and adsorption. Environ. Sci.
mium (VI) from aqueous solution by activated carbon derived from olive Technol. 49, 5654e5662.
bagasse and applicability of different adsorption models. Chem. Eng. J. 144, Wang, W., Gu, P., Zhang, G., Wang, L., 2013b. Organics removal from ROC by PAC
188e196. accumulative countercurrent two-stage adsorptioneMF hybrid process d a
Deng, H., Yang, L., Tao, G., Dai, J., 2009. Preparation and characterization of activated laboratory-scale study. Sep. Purif. Technol. 118, 342e349.
carbon from cotton stalk by microwave assisted chemical activation e appli- Wei, X., Gu, P., Zhang, G., 2014. Reverse osmosis concentrate treatment by a PAC
cation in methylene blue adsorption from aqueous solution. J. Hazard. Mater countercurrent four-stage adsorption/MF hybrid process. Desalination 352,
166, 1514e1521. 18e26.
Di Natale, F., Erto, A., Lancia, A., Musmarra, D., 2015. Equilibrium and dynamic study Wionczyk, B., Cierpiszewski, R., Mo  l, A., Prochaska, K., 2011. Studies on the kinetics
on hexavalent chromium adsorption onto activated carbon. J. Hazard. Mater and equilibrium of the solvent extraction of chromium(III) from alkaline
281, 47e55. aqueous solutions of different composition in the system with Aliquat 336.
Foo, K.Y., Hameed, B.H., 2010. Insights into the modeling of adsorption isotherm J. Hazard. Mater 198, 257e268.
systems. Chem. Eng. J. 156, 2e10. Zhang, C., Gu, P., Zhao, J., Zhang, D., Deng, Y., 2009. Research on the treatment of
Gheju, M., Balcu, I., 2011. Removal of chromium from Cr(VI) polluted wastewaters liquid waste containing cesium by an adsorption-microfiltration process with
by reduction with scrap iron and subsequent precipitation of resulted cations. potassium zinc hexacyanoferrate. J. Hazard. Mater 167, 1057e1062.
J. Hazard. Mater 196, 131e138. Zhao, C., Gu, P., Cui, H., Zhang, G., 2012. Reverse osmosis concentrate treatment via a
Gherasim, C.V., Bourceanu, G., 2013. Removal of chromium (VI) from aqueous so- PACeMF accumulative countercurrent adsorption process. Water Res. 46,
lutions using a polyvinyl-chloride inclusion membrane: experimental study 218e226.
and modelling. Chem. Eng. J. 220, 24e34. Zhu, N., Yan, T., Qiao, J., Cao, H., 2016. Adsorption of arsenic, phosphorus and
Golder, A.K., Chanda, A.K., Samanta, A.N., Ray, S., 2011. Removal of hexavalent chromium by bismuth impregnated biochar: adsorption mechanism and
chromium by electrochemical reductioneprecipitation: investigation of process depleted adsorbent utilization. Chemosphere 164, 32e40.

You might also like