Вы находитесь на странице: 1из 7

Journal of Environmental Management 114 (2013) 225e231

Contents lists available at SciVerse ScienceDirect

Journal of Environmental Management


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

Arsenic removal by modified activated carbons with iron hydro(oxide)


nanoparticles
Alma Veronica Vitela-Rodriguez, Jose Rene Rangel-Mendez*
Division of Environmental Sciences, Instituto Potosino de Investigación Científica y Tecnológica, A.C. Camino a la Presa San José 2055, Col. Lomas 4ta. Sección,
C.P. 78216 San Luis Potosí, S.L.P., Mexico

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

Article history: Different activated carbons modified with iron hydro(oxide) nanoparticles were tested for their ability to
Received 21 March 2012 adsorb arsenic from water. Adsorption isotherms were determined at As (V) concentrations < 1 ppm,
Received in revised form with varying pH (6, 7, 8) and temperature (25 and 35  C). Also, competition effect of anions on the As (V)
6 September 2012
adsorption capacity was evaluated using groundwater. The surface areas of the modified activated
Accepted 3 October 2012
Available online 10 November 2012
carbons ranged from 632 m2 g1 to 1101 m2 g1, and their maximum arsenic adsorption capacity varied
from 370 mg g1 to 1250 mg g1. Temperature had no significant effect on arsenic adsorption; however,
arsenic adsorption decreased 32% when the solution pH increased from 6 to 8. In addition, when
Keywords:
Activated carbon
groundwater was used in the experiments, the arsenic adsorption considerably decreased due to the
 
Iron hydro(oxide) presence of competing anions (mainly SO2 4 , Cl and F ) for active sites. The data from kinetic studies
Arsenic fitted well to the pseudo-second-order model (r2 ¼ 0.98e0.99). The results indicated that sample CAZ-M
Adsorption had faster kinetics than the other two materials in the first 10 min. However, sample F400-M was only
Kinetics 5.5% slower than CAZ-M. The results of this study show that iron modified activated carbons are efficient
adsorbents for arsenic at concentrations lower than 300 mg L1.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction reverse osmosis, membrane filtration, and adsorption processes.


From these, adsorption has been widely used to remove arsenic due
Contamination of groundwater with arsenic (As), a highly toxic to its low cost and high efficiency (Han et al., 2002; DeMarco et al.,
element, is a serious problem around the world that affects millions 2003; USEPA, 2007; Mohan and Pittman, 2007; Vaclavikov et al.,
of people (Smedley and Kinniburgh, 2002). Chronic exposure to this 2008). Iron oxides, such as hematite and goethite, have shown
pollutant is associated with several health problems such as good performance as arsenic adsorbents due to their high selec-
arsenicosis and different kinds of cancer (Hughes, 2002; Kapaj tivity for this element. However, iron oxides have low mechanical
et al., 2006; Ferreccio et al., 2000). The World Health Organiza- resistance, which does not permit their application in fixed bed
tion (WHO) guideline value for As in drinking water is 10 mg L1 columns (Guo et al., 2007; Solozhenkin et al., 2003; Raven et al.,
(WHO, 2008). Nevertheless, approximately 50 million people 1998). Modified adsorbents, such as activated carbon with iron
worldwide are exposed to higher arsenic concentrations in oxides, have been reported in the literature to improve adsorption
drinking water. Well-known As contaminated countries include capacities and mechanical properties, indicating that arsenic
Bangladesh, India, China, Hungary, Chile, Argentina, Mexico, and adsorption processes are becoming more efficient (Chen et al.,
Vietnam (Smedley and Kinniburgh, 2002). In Mexico, As contami- 2007; Fierro et al., 2009; Jang et al., 2008). It is known that
nated aquifers exceeding the maximum permissible level of arsenic is adsorbed on iron hydro(oxides) by inner-sphere surface
25 mg L1 have been found in Coahuila, Chihuahua, Durango, complexes (Goldberg and Johnston, 2001; Vaughan and Reed,
Guanajuato, San Luis Potosi and Zacatecas (CAN, 1999; Del Razo 2005), so the available surface area of iron loaded on the carbo-
et al., 1990; Camacho et al., 2011; Armienta and Segovia, 2008). naceous matrix determines the maximum adsorption capacity of
Several technologies have been reported to remove arsenic from the adsorbent, rather than the amount of iron contained in the
drinking water, including coagulation-flocculation, ion exchange, activated carbon. Because of this, the objective of our previous
publications (Nieto-Delgado and Rangel-Mendez, 2012; Arcibar-
* Corresponding author. Tel.: þ52 (444) 834 20 00; fax: þ52 (444) 934 20 10. Orozco et al., 2012) was to determine the optimal parameter to
E-mail address: rene@ipicyt.edu.mx (J.R. Rangel-Mendez). anchor iron nanoparticles onto activated carbon.

0301-4797/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jenvman.2012.10.004
226 A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231

Since a detail study of As (V) adsorption by granular activated 2.3. Materials characterization
carbons containing iron hydro(oxide) nanoparticles has not been
reported, the main objective of this research is to correlate surface 2.3.1. Surface area and porosity
area, iron content and the materials’ pH to the As (V) adsorption Surface area and pore size distribution of the activated carbons
capacity and kinetics of these iron modified activated carbons in were determined from the corresponding nitrogen adsorptione
synthetically and naturally contaminated water. desorption isotherms at 77 K using a Micromeritics ASAP 2020
system. The surface area was determined by the Brunauer, Emmett
2. Materials and methods and Teller (BET) equation and the pore size distribution was
calculated by using the density functional theory (DFT) method.
2.1. Materials
2.3.2. Iron content determination
All chemicals used were reagent grade. Iron solutions were The iron content was determined after acid digestion of the
prepared from FeCl3$6H2O salt (Fermont). An arsenic stock solution activated carbons. 40 mg of sample were added to 20 mL of a HNO3:
of 5 mg L1 was prepared from Na2HAsO4$7H2O salt (Sigmae H2SO4 (5:1) solution. The samples were digested for 1 h at 150  C
Aldrich). Appropriated dilutions were prepared to conduct using an advanced digestion system by microwave (Milestone,
adsorption experiments with initial concentrations of As (V) Ethos 1). The obtained solution was analyzed by ICP-OES (Varian
ranging from 25 mg L1 to 1500 mg L1, in accordance with those 730-ES) at a wavelength of 271.9 nm.
found in contaminated waters (Smedley and Kinniburgh, 2002;
CAN, 1999; Del Razo et al., 1990; Camacho et al., 2011; Armienta and 2.3.3. Slurry pH
Segovia, 2008). Deionized water was used to prepare all solutions. The point of zero charge (PZC) was determined as following.
The granular activated carbons used in this research were: bitu- 100 mg of activated carbon were added to 5 mL of deionized water,
minous Filtrasorb-400 (F400) from Calgon, CAZ and CAP, both previously aerated with nitrogen. The containers were sealed and
produced from Agave salmiana bagasse and chemically activated constantly stirred for 72 h. After that, the slurry pH was measured
with ZnCl2 and H3PO4, respectively (Nieto-Delgado and Rangel- and considered to be equal to the material’s pH.
Mendez, 2011). Natural groundwater was collected from a deep
well located in the city of San Luis Potosi, Mexico, following the 2.3.4. X-ray diffraction studies
NOM-014-SSA1-1993 methodology (NOM, 1993). The characteris- XRD patterns were obtained in a Bruker D8 diffractometer using
tics of this water are given in Table 1. a CuKa radiation (l ¼ 1.5418  A). Iron modified samples were
reduced to a mesh size <90 in an agate mortar, and then the
powder was placed in the XRD sample port. The patterns were
2.2. Activated carbons modification
obtained with a step size of 0.02 2q at 10 s per step.
It should be mentioned that scanning electron micrographs of
Granular activated carbons (F400, CAZ and CAP) were modified
these samples have been recently reported some were else (Nieto-
with iron hydro(oxide) nanoparticles by forced hydrolysis of FeCl3
Delgado and Rangel-Mendez, 2011), however, the main findings are
solution following the established methodology by Nieto-Delgado
mentioned in the discussion Section 3.1 of this manuscript.
and Rangel-Mendez (2011): (1) 200 mg of activated carbon of U.S.
mesh no. 100/140 (148e105 mm) were placed in contact with 10 mL
2.4. Adsorption isotherms
of FeCl3 solution (FeCl3$6H2O, 97% ACS grade from Fermont) inside
a sealed glass container; (2) the mixture was mixed for 24 h at 25  C
The As (V) adsorption capacity of modified activated carbons
to allow the diffusion of iron into the carbon pores; (3) thermal
and control samples (unmodified activated carbon) was deter-
hydrolysis was carried out by placing the containers in a preheated
mined in duplicate at pH 7 and 25  C and the average was reported.
furnace (70e120  C) during a selected time; (4) after heating, the
15 mg of activated carbon were added to 20 mL of As (V) solution at
samples were rinsed with distilled water until the elimination of
different initial concentrations. The experiments were kept under
all the soluble iron and the iron particles that were not anchored
constant stirring. The pH of each sample was adjusted daily with
to the carbon surface: this was determined by analyzing the
0.1 N NaOH and/or 0.1 N HNO3 until the equilibrium was reached:
supernatant. The activated carbons reported in this manuscript
this was determined once the solution pH and the As concentration
were treated at the optimum established conditions (Nieto-
remained constant. The initial and the final arsenic concentration
Delgado and Rangel-Mendez, 2011): F400 (96  C for 56 h), CAZ
was measured by atomic emission spectroscopy (AES) in an ICP-AES
(110  C for 6.8 h) and CAP (58  C for 6.8 h).
(Varian 730-ES) at a wavelength of 188.980 nm. The adsorption
capacity of the activated carbons was calculated by using the
following equation:
Table 1
Physicochemical parameters of well water used in adsorption experiments. q ¼ VðCo  Ce Þ=m (1)

Where q is the adsorption capacity (mg g1), V is the volume of


Parameter Concentration Maximum permissible
detected limit (NOM-127-SSA1-1994)
solution (L), Co is the initial As concentration (mg L1), Ce is the
pH 7.55 6.5e8.5
Total hardness 20.24 500
equilibrium As concentration (mg L1), and m is the mass of
(mg L1 CaCo3) adsorbent (g).
Cl (mg L1) 14.55 250
F (mg L1) 4.03 1.50
1 2.5. Adsorption kinetics
PO3
4 (mg L ) 0.44c NIa
1
SO2
4 (mg L ) 12.70 400
As (mg L1) NDb 25 1 g of carbon sample was added to a rotating basket that was
a
Not indicated.
then submerged in a 1 L reactor: it should be mentioned that the
b
Not detected. adsorber design fulfilled the criteria for a vertical-shaft turbine. The
c
Determined as total phosphorous. reactor was partially immersed in a water bath that operated at
A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231 227

constant temperature. The adsorption kinetics were conducted


θ
under the following conditions: 25  C, 300 min1, initial pH 7, 1 L Hematite
solution, and at an initial arsenic concentration of 50 mg L1. 5 mL of
* Akaganeite
+
θ Graphite
solution were taken after 1, 3, 5, 7, 10, 20, 40, 60, 80, and 100 min of
the start of the experiment. The samples were analyzed for arsenic
as previously described.
*

Intensity (a.u.)
* +
3. Results and discussion * +
*
3.1. Activated carbons characterization

Table 2 shows that the iron content in the activated carbons


increased almost 2% after modification. For example, for F400 this +
increased from 0.25% to 1.67%, and for CAZ and CAP it went from
0.08% to 1.67% and to 1.81%, respectively. X-ray diffraction analysis
of F400-M (Fig. 1) reported the presence of hematite (a-Fe2O3) at 30 35 40 45 50 55 60 65 70
33 , 36 , 41 and 50 and akaganeite (b-FeOOH) at 35 , 39 and 68 .
2 theta
It should be mentioned that under oxic conditions, gohetite and
hematite are thermodynamically the most stable compounds in Fig. 1. XRD patterns of iron modified activated carbon with a step size of 0.02 2q at
this iron hydro(oxide) system (Cornell and Schwertmann, 2003; 10 s per step. This spectrum shows the presence of hematite and akaganeite on the
Hristovski et al., 2009), therefore, it is expected that the arsenic surface of modified activated carbons.

adsorption capacity of the modified materials reported herein will


From these results, it is clear that surface area, iron content, and
remain with time. Moreover, the surface area and pore volume also
the materials’ pH are factors that influence the arsenic adsorption
slightly decreased after modification with iron hydro(oxide)
capacity of adsorbents (Table 2). Some authors (Viraghavan et al.,
nanoparticles due to the obstruction of some micro and mesopores
2001) have reported an improvement of arsenic adsorption
as shown in Table 2. CAZ-M and CAP-M had a surface area decrease
of around 12% while that of F400-M reduced by only 4.5%: CAZ-M 1200
and F400-M have the highest surface area, 1101 and 1045 m2 g1 A F400-SM
CAZ-SM
respectively. The reduction in pore volume of micropores, from
1000 CAP-SM
about 5 to 12% (Table 2), suggested that the iron hydro (oxides) ____ Langmuir model
particles anchored on activated carbons are smaller than 20  A. The
micrographs of iron modified activated carbons recently reported 800
by our research group (Nieto-Delgado and Rangel-Mendez, 2011)
Qe (μg g-1 )

showed iron hydro(oxide) particles from 3 to 36 nm inside pores


600
and amorphous agglomerates at the external surface of carbon
grains, which support the results reported in this manuscript.
400
3.2. Adsorption isotherms
200
Arsenic adsorption isotherms of activated carbons at pH 7 are
shown in Fig. 2. Experimental data were fitted with the Langmuir
0
model (Table S.1, see Supplementary material) since this adjusted 0 100 200 300 400 500 600
better than the Freundlich model. It can be observed that the As (V)
Ce (μg L-1 )
adsorption capacity of untreated CAZ and CAP is very low. However,
the As (V) adsorption capacity of these carbons considerably 1200
increased after modification with iron hydro(oxide) nanoparticles B F400-M
CAZ-M
from 250 mg g1 to 526 mg g1 for CAZ-M, and from 167 mg g1 to 1000 CAP-M
370 mg g1 for CAP-M at 500 ppb, which is related to the approxi- ___ Langmuir model
mately 21 fold increase in iron content. In contrast, commercial
800
F400 removed a considerable amount of As (V), which was attrib-
Qe (μg g-1 )

uted to its 0.25% of iron. However, its adsorption capacity increased


almost 25% after modification from 1010 mg g1 to 1250 mg g1. 600

400
Table 2
Physical characteristics and iron content of unmodified and treated activated
carbons. 200
2 1 3 1
Sample % Fe Slurry pH SBET (m g ) Pore volume (cm g )

Micro Meso Total 0


0 100 200 300 400 500 600
F400-SM 0.25 8.62 1045 0.329 0.073 0.402
F400-M 1.67 6.05 998 0.312 0.064 0.376 Ce (μg L-1 )
CAZ-SM 0.08 3.28 1249 0.394 0.076 0.470
CAZ-M 1.67 4.04 1101 0.346 0.071 0.417 Fig. 2. As (V) adsorption isotherms of unmodified activated carbons (A) and modified
CAP-SM 0.08 2.40 726 0.211 0.185 0.396 activated carbons (B) at pH 7 and 25  C. These figures show that the adsorption
CAP-M 1.81 3.57 632 0.185 0.162 0.347 experimental data follow the Langmuir model at low and high concentration of arsenic
that is in agreement with the adsorption mechanism proposed.
228 A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231

capacity as the iron content increases. In this study, although the 1000
modified activated carbons have similar iron content, their As (V)
adsorption capacity is quite different, due to the influence of both
surface area and slurry pH. Even though CAP-M has the highest iron 800
content (1.81%), its As (V) adsorption capacity is the lowest because
it has the least surface area (632 m2 g1) and slurry pH (3.57). On

Qe (μg g-1 )
the other hand, CAZ-M has 1.67% of iron, the highest surface area 600
(1101 m2 g1) and a slurry pH of 4.04, and its As (V) adsorption
capacity is higher than for CAP-M: this is obviously related to
400
surface area but also to the effect of the materials pH that is
explained in section 3.3. Moreover, F400-M has 1.67% of iron,
pH 6
a surface area (998 m2 g1) slightly lower than CAZ-M, and both the 200 pH 7
highest slurry pH (6.05) and As (V) adsorption capacity. These pH 8
results suggest that the materials pH is a determinant parameter. ___ Langmuir model
According to the speciation diagram (Vaclavikov et al., 2008), As (V) 0
is present as negative species (H2AsO 2
4 and HAsO4 ) at pH 7, at
0 50 100 150 200 250 300 350 400
-1
which the adsorption experiments were conducted. This means Ce (μg L )
that electrostatic interactions partially contribute to the adsorption
process. F400-M has a less negative surface at pH 7 and hence the Fig. 3. As (V) adsorption isotherms of F400-M at different pH and 25  C. An increase in
pH slightly decreases the adsorption capacity and the Langmuir model adequately
arsenic anions experience less repulsion than with CAZ-M, which adjusts the experimental data suggesting that the adsorption mechanism does not
has a more negative surface at the same conditions. Hence, F400-M change.
has a higher As (V) adsorption capacity than CAZ-M despite having
a higher surface area. Comparing these results with others previ- electrostatic repulsion between As (V) and the activated carbons
ously reported in the literature (Fierro et al., 2009; Badruzzaman surface increases, the inner-sphere surface complexes will diminish
et al., 2004; Chuang et al., 2005), ranging from 4 to 28 mg L1, for and consequently the adsorption capacity decreases.
similar materials at the same experimental conditions (25  C, pH 7
and 300 mg L1 at equilibrium), the As (V) adsorption capacity of the
modified activated carbons reported herein, 847 mg L1 the highest, 3.4. Effect of temperature
is much higher (see Table 3). Something to remark is that the
anchorage of iron hydro(oxide) nanoparticles onto activated As (V) adsorption isotherms (Fig. 4) were conducted at 25  C and
carbons without considerably decreasing surface area is a very 35  C and fitted with the Langmuir model (Table S.2, see
important factor for the removal of arsenic from water. Supplementary material). The maximum As (V) adsorption capacity
Since F400-M had the highest As (V) adsorption capacity, it was of F400-M at 400 mg L1 was 588 mg g1 at 35  C and 526 mg g1 at
selected to study the effect of pH, temperature, and competing 25  C. These results indicate that temperature did not significantly
anions on adsorption capacity. affect adsorption capacity, and suggest that adsorption takes place
by chemical bonds between As (V) and iron hydro(oxides). Several
authors have reported similar results. For instance, Banerjee et al.
3.3. Effect of pH
(2008), Mondal et al. (2007) and Solozhenkin et al. (2003) re-
ported that when changing the temperature from 25  C up to 60  C
Adsorption experiments were conducted at pH 6, 7 and 8
during the removal of As (V) from water, the adsorption capacity of
(Fig. 3). Result show that the As (V) adsorption capacity of F400-M
iron modified carbons and/or iron hydro(oxides) does not change
at 300 mg L1 at equilibrium was: 880 mg g1 at pH 6, 810 mg g1 at
significantly.
pH 7, and 590 mg g1 at pH 8, respectively. The As (V) adsorption
capacity decreased 32% when the pH increased from 6 to 8. This
effect can be attributed to the electrostatic repulsion between As 600
(V) and activated carbon (Streat et al., 2008; Ona-nguema et al.,
2005). The surface area of F400-M becomes more negative as the 500
solution pH increases (slurry pH 6.05), hence, the attractive force
toward As (V) anionic species decreases. Several authors
(Solozhenkin et al., 2003; Sherman and Randall, 2003; Mondal 400
Qe (μg g-1)

et al., 2007) have reported that adsorption of arsenate onto iron


oxides occurs by inner-sphere surface complexes. However, if the 300

Table 3 200
Comparison of As (V) adsorption capacities among different adsorbents reported in
the literature at 25  C, pH 7 and 300 g L1 at equilibrium. 25°C
100 35°C
1
Adsorbent Qe (mgAs g
a
) Reference
F400-M 847 This study
CAZ-M 431 This study 0
CAP-M 181 This study 0 20 40 60 80 100 120
CAG-Fe 28 [Fierro et al.]
Ce (μg L-1)
IOCSb 18 [Viraghavan et al.]
GFHc 4 [Badruzzaman et al.]
Fig. 4. As (V) adsorption isotherms of F400-M at different temperature and pH 7.
a
300 mg L1 at equilibrium. A change in temperature does not significantly change the arsenic adsorption capacity
b
Iron oxide coated sand. and the Langmuir model (line) adequately adjusts the experimental data suggesting
c
Granular ferric hydroxide. that the adsorption mechanism does not change.
A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231 229

To determine the nature of the adsorption process, thermody- 7000


namic parameters were calculated by using the following
equations: 6000

DG ¼ RT ln K (2) 5000

ln K ¼ DH=ðRTÞ þ C

Qe (μg g-1)
(3) 4000

DG ¼ DH  T DS (4) 3000
Cinitial
As (V) 50 μgL -1
where DG is the Gibbs free energy, T the absolute temperature (K), R 2000 SO42- 12.70 mgL -1
the gas constant (8.314 J/mol K), K the Langmuir isotherm constant, F- 4.03 mgL -1
DH the enthalpy change, and DS the entropy change (Kanel et al., 1000 Cl - 14.55 mgL -1
2005). The negative DG values reported in Table S.2 (see
Supplementary material) indicated that the adsorption process was 0
spontaneous, while the positive DH value suggested that the 0 20 40 60 80 100
adsorption process was endothermic, which is in accordance with
Time (min)
other researchers (Guo et al., 2010; Guan et al., 2009). This supports
the results shown in Fig. 4, where the arsenic adsorption capacity Fig. 6. Anions adsorption kinetics in F400-M at 25  C and pH 7 using groundwater. The

slightly increased with temperature. lines represent the tendency. SO2
4 and Cl compete with arsenic for active sites.

3.5. Effect of co-ions on the As (V) removal Concentration is another very important factor: in this study, the
SO2
4 initial concentration was 20 times higher than that of As (V),
Anions such as carbonates, sulfates, phosphates, and silicates are therefore the arsenic adsorption capacity was considerably
present in groundwater and can interfere in the removal of anionic reduced.
arsenic species. As (V) adsorption isotherms (Fig. 5) were deter-
mined in groundwater from a deep well, and the results indicated
that the adsorption capacity of F400-M considerably decreased 3.6. Adsorption kinetics
from 847 mg g1 to 148 mg g1, at 300 mg L1 and pH 7, when using
groundwater. This is attributed to the competition among arsenic The pseudo-second-order equation was used to fit experimental
and anions present in groundwater (Table 1). Several authors (Guo data. This equation is represented as follows:
and Chen, 2005; Wilkie and Hering, 1996; Payne and Abdel-Fattah, . 
2005) have reported that PO3 2
4 and SO4 affect the arsenic removal t=qt ¼ 1 kq2e þ t=qe (5)
because these have a very similar chemical structure to arsenic. It
should be noted that the concentration of phosphates, 12.7 mg L1, where qt and qe is the adsorbed amount at time t and at equilibrium,
in this groundwater (determined as total phosphorus) is many respectively, and k is the rate constant (Ho and McKay, 2000). The
times higher than the arsenic concentration used in experiments. kinetic parameters and correlation coefficients (0.98e0.99) for all
On the other hand, adsorption kinetics (Fig. 6) were carried out the carbon materials studied in this research are reported in
in groundwater to determine which anion has greater effect on the Table S.3 (see Supplementary material).
arsenic removal. Results indicated that F400-M adsorbed more As seen in Fig. 7, CAZ-M removed 36% of the arsenic initial
 
SO2
4 followed by Cl , F , and As (V). It has been reported that the concentration in the first 10 min, whereas F400-M removed 32%
SO2
4 affinity to iron hydro (oxides) is weak; however, this can and CAP-M 30% in the same time. The adsorption kinetics are
compete with As (V) for active sites (Guo and Chen, 2005). considered rapid, compared to that reported in the literature (Chen
et al., 2007; Fierro et al., 2009; Weber and Morris, 1963), because
the initial concentration decreased 85% in about 1 h and the
1200 adsorption equilibrium was reached in 80 min. This could be
Deionized water attributed to the unblocked pores of carbon materials even after
1000 Groundwater modification with iron hydro(oxides). Slower arsenic adsorption
kinetics have been reported in modified activated carbons with
800 iron: for example, Chen et al. (2007) reported an arsenic removal
Qe (μg g-1)

rate of 80% after 3 h when using iron modified activated carbon,


600 Fierro et al. (2009) showed that the arsenate adsorption by iron-
doped activated carbons was 94% after 6 h, and finally Payne and
400 Abdel-Fattah (2005) reported that the arsenate adsorption by
iron-doped activated carbon was slow because the arsenic initial
200 concentration (50 mg L1) decreased 40% after 25 h.

0 3.7. Weber and Morris model


0 100 200 300 400 500
Ce (μg L-1) Several adsorption diffusion models have been reported to
Fig. 5. As (V) adsorption isotherms of F400-M using different water at 25 C
and pH 7.
describe the process of film diffusion and/or intraparticle diffusion.
The arsenic adsorption capacity decreases in groundwater due to the present of other Because the pseudo-second-order model does not provide infor-
anions and the Langmuir model (line) adequately adjusts the experimental data. mation about the rate-controlling step, experimental data were
230 A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231

50 50
A A
40 40

30

Qe (μgL-1)
30
Qe (μg g-1)

20 20

F400-SM
F400-SM 10 CAZ-SM
10 CAZ-SM CAP-SM
CAP-SM
0
0 0 2 4 6 8 10
0 20 40 60 80 100
1/2
t (min)
Time (min)

50
B 50
B
40 40

30
Qe (μg g-1)

Qe (μg g-1)

30

20 20

F400-M
10 CAZ-M 10 F400-M
CAP-M CAZ-M
CAP-M
0 0
0 20 40 60 80 100 0 2 4 6 8 10
Time (min) t1/2 (min)

Fig. 7. As (V) adsorption kinetics at 50 mg L1, 25  C, initial pH 7, and 300 min1 using Fig. 8. WebereMorris model of unmodified activated carbons (A) and modified acti-
unmodified activated carbons (A) and modified activated carbons (B). More then 30% of vated carbons (B). According to this model, intraparticle diffusion is not the rate-
arsenic is removed in the first 10 min. The lines represent the tendency. limiting step.

fitted to the Webber-Morris model (1963), which is represented as


4. Conclusions
follows:
The iron content in the modified activated carbons increases
about 2%, and this element is present as hematite and akaganeite.
qt ¼ Kint t 1=2 (6) Also, the 5e12% reduction on the pore volume of micropores
suggests that the iron hydro(oxides) particles anchored on the
where Kint is the intraparticle diffusion rate constant. carbon surface are smaller than 20  A, which helped to produce
According to this model, if a straight line through the origin is materials with a considerably high surface area of about
obtained when t1/2 is plotted against qt, intraparticle diffusion is the 1100 m2 g1. The As (V) adsorption capacity of activated carbons
rate-limiting step. Otherwise, adsorption kinetic would be limited increases after modification with iron due to the presence of iron
by film diffusion. Fig. 8 shows that the straight lines do not pass hydro(oxide) nanoparticles. The modified activated carbons
through the origin, indicating that intraparticle diffusion is not the (F400-M, CAZ-M, and CAP-M) removed As (V) at similar concen-
rate-limiting step. However, these results are not in accordance trations to those found in drinking water (<300 mg L1), but F400-
with the experimental data. If film diffusion were the rate-limiting M was found to be the most efficient material according to
step, the adsorption equilibrium would be reached almost instantly adsorption capacity and adsorption kinetics. Temperature does not
(less than 10 min); however, it occurred after 60 min. Therefore, have a significant effect on the As (V) adsorption capacity, which is
these results suggest that since the Weber and Morris model does endothermic. However, there is an effect of pH on the As (V)
not consider the physical parameter of materials such as surface adsorption capacity, which decreases as pH increases. Anions
area, pore volume, and tortuosity, these findings are not conclusive. present in groundwater, mainly SO2 4 and Cl1, compete with
Contrary to the prediction by the Weber and Morris model, arsenic for active sites on F400-M, decreasing its As (V) adsorption
Ghanizadeh et al. (2010) determined that both film diffusion and capacity. Finally, comparing the As (V) adsorption capacity of the
intraparticle diffusion affect the As (V) adsorption kinetics in iron- modified materials reported in this manuscript with other mate-
impregnated activated carbons. rials reported in the literature, these activated carbons have
A.V. Vitela-Rodriguez, J.R. Rangel-Mendez / Journal of Environmental Management 114 (2013) 225e231 231

a considerably higher adsorption capacity. Therefore, they can be Guo, X., Chen, F., 2005. Removal of arsenic by bead cellulose loaded with iron
oxyhydroxide from groundwater. Environ. Sci. Technol. 39, 6808e6818.
a viable alternative for arsenic removal on large scale and contin-
Guo, X., Du, Y., Chen, F., Park, H.S., Xie, Y., 2007. Mechanism of removal of arsenic by
uous systems. bead cellulose loaded with iron oxyhydroxide (b-FeOOH): EXAFS study.
J. Colloid Interface Sci. 314, 427e433.
Acknowledgments Han, B., Runnells, T., Zimbron, J., Wickramasinghe, R., 2002. Arsenic removal from
drinking water by flocculation and microfiltration. Desalination 145, 293e298.
Ho, Y.S., McKay, G., 2000. The kinetics of sorption of divalent metal ions onto
The authors gratefully acknowledge the financial support from sphagnum moss flat. Water Res. 34 (3), 735e742.
CONACyT through projects FMSLP-2008-C02-99664, SEP-CB-2008- Hristovski, K.D., Westerhoff, P.K., Möller, T., Sylvester, P., 2009. Effect of synthesis
conditions on nano-iron (hydr)oxide impregnated granulated activated carbon.
01-105920, and scholarship received (232499). In addition, the Chem. Eng. J. 146, 237e243.
authors appreciate the national laboratories LINAN and LANBAMA Hughes, M.F., 2002. Arsenic toxicity and potential mechanisms of action. Toxicol.
and the technical support of M.C. Rocha-Medina, D.I. Partida- Lett. 133, 1e16.
Jang, M., Chen, W., Cannon, F., 2008. Preloading hydrous ferric oxide into granular
Gutiérrez, and G. Vidriales-Escobar. activated carbon for arsenic removal. Environ. Sci. Technol. 42, 3369e3374.
Kanel, S.R., Manning, B., Charlet, L., Choi, H., 2005. Removal of arsenic (III)
Appendix A. Supplementary data from groundwater by nanoscale zero-valent iron. Environ. Sci. Technol. 39,
1291e1298.
Kapaj, S., Peterson, H., Liber, K., Bhattacharya, P., 2006. Human health effects from
Supplementary data related to this article can be found at http:// chronic arsenic poisoning-a review. J. Environ. Sci. Health A 41, 2399e2428.
dx.doi.org/10.1016/j.jenvman.2012.10.004. Mohan, D., Pittman, C.U., 2007. Arsenic removal from water/wastewater using
adsorbents-a critical review. J. Hazard. Mater. 142, 1e53.
Mondal, P., Balomajumder, C., Mohanty, B., 2007. A laboratory study for the treat-
References ment of arsenic, iron, and manganese bearing ground water using Fe3þ
impregnated activated carbon: effects of shaking time, pH and temperature.
Arcibar-Orozco, J.A., Avalos-Borja, M., Rangel-Mendez, J.R., 2012. Effect of phosphate J. Hazard. Mater. 144, 420e426.
on the particle size of ferric oxyhydroxides anchored onto activated carbon: Nieto-Delgado, C., Rangel-Mendez, J.R., 2011. Production of activated carbon from
As(V) removal from water. Environ. Sci. Technol. 46, 9577e9583. organic by-products from the alcoholic beverage industry: surface area and
Armienta, M.A., Segovia, N., 2008. Arsenic and fluoride in the groundwater of hardness optimization by using the response surface methodology. Ind. Crop
Mexico. Environ. Geochem. Health 30, 345e353. Prod. 34, 1528e1537.
Badruzzaman, M., Westerhoff, P., Knappe, D.R.U., 2004. Intraparticle diffusion and Nieto-Delgado, C., Rangel-Mendez, J.R., 2012. Anchorage of iron hydro(oxide)
adsorption of arsenate onto granular ferric hydroxide (GFH). Water Res. 38, nanoparticles onto activated carbon to remove As(V) from water. Water Res. 46,
4002e4012. 2973e2982.
Banerjee, K., Amy, G.L., Prevost, M., Nour, S., Jekel, M., Gallagher, P.M., Norma Oficial Mexicana NOM-014-SSA1-1993, 1993. Procedimientos sanitarios para
Blumenschein, C.D., 2008. Kinetic and thermodynamic aspects of adsorption of el muestreo de agua para uso y consumo humano en sistemas de abasteci-
arsenic onto granular ferric hydroxide (GFH). Water Res. 42, 3371e3378. miento de agua públicos y privados.
Camacho, L.M., Gutiérrez, M., Alarcón-Herrera, M.T., Villalba, M.L., Deng, S., 2011. Ona-nguema, G., Morin, G., Juillot, F., Calas, G., Brown, G.E., 2005. EXAFS analysis of
Occurrence and treatment of arsenic in groundwater and soil in northern arsenite adsorption onto two-line ferrihydrite, hematite, goethite and lep-
Mexico and southwestern USA. Chemosphere 83, 211e225. idocrocite. Environ. Sci. Technol. 39, 9147e9155.
Chen, W., Parette, R., Zou, J., Cannon, F., 2007. Arsenic removal by iron-modified Payne, K.B., Abdel-Fattah, T.M., 2005. Adsorption of arsenate and arsenite by iron-
activated carbon. Water Res. 41, 1851e1858. treated activated carbon and zeolites: effects of pH, temperature, and ion
Chuang, C.L., Fan, M., Xu, M., Brown, R.C., Sung, S., Saha, B., Huang, C.P., 2005. strength. J. Environ. Sci. Heal. A 40, 723e749.
Adsorption of arsenic(V) by activated carbon prepared from oat hulls. Chemo- Raven, K.P., Jain, A., Loeppert, R.H., 1998. Arsenite and arsenate adsorption on fer-
sphere 61, 478e483. rihydrite: kinetics, equilibrium, and adsorption envelopes. Environ. Sci. Technol.
Comisión Nacional del Agua, 1999. Identificación de zonas de riesgo en agua sub- 32, 344e349.
terránea por presencia de arsénico y fluoruros. Gerencia de Saneamiento y Sherman, D.M., Randall, S.R., 2003. Surface complexation of arsenic (V) to iron (III)
Calidad del Agua. (hydr)oxides: structural mechanism from ab initio molecular geometries and
Cornell, R.M., Schwertmann, U., 2003. The Iron Oxides. Structure, Properties, EXAFS spectroscopy. Geochim. Cosmochim. Acta 67, 4223e4230.
Reactions, Ocurrences and Uses. Wiley-VCH, Weinheim. Smedley, P.L., Kinniburgh, D.G., 2002. A review of the source, behavior and distri-
Del Razo, L.M., Arellano, M.A., Cebrián, M.E., 1990. The oxidation states of arsenic in bution of arsenic in natural waters. Appl. Geochem. 17, 517e568.
well-water from a chronic arsenicism area of northern Mexico. Environ. Pollut. Solozhenkin, P.M., Deliyanni, E.A., Bakoyannakis, V.N., 2003. Removal of As(V) ions
64, 143e153. from solution by akaganeite b-FeO(OH) nanocrystals. J. Min. Sci. 39, 287e296.
DeMarco, M.J., SenGupta, A.K., John, E., Greenleaf, J.E., 2003. Arsenic removal using Streat, M., Hellgardt, K., Newton, N.L.R., 2008. Hydrous ferric oxide as an adsorbent
a polymeric/inorganic hybrid sorbent. Water Res. 37, 164e176. in water treatment. Part 2. Adsorption studies. Process Saf. Environ. 86, 11e20.
Ferreccio, C., González, C., Milosavjlevic, V., Marshall, G., Sacha, A.M., Smith, A.H., US Environmental Protection Agency, 2007. Treatment Technologies for Arsenic
2000. Lung cancer and arsenic concentrations in drinking water in Chile. Removal.
Epidemiology 11 (6), 673e679. Vaclavikov, M., Gallios, G.P., Hredzak, S., Jakabsky, S., 2008. Removal of arsenic from
Fierro, V., Muñiz, G., Gonzalez-Sanchez, G., 2009. Arsenic removal by iron-doped water streams: an overview of available techniques. Clean Techn. Environ.
activated carbons prepared by ferric chloride forced hydrolysis. J. Hazard. Policy 10, 89e95.
Mater. 168, 430e437. Vaughan Jr., R.L., Reed, B.E., 2005. Modeling As(V) removal by iron oxide impreg-
Ghanizadeh, G., Ehrampoush, M.H., Ghaneian, M.T., 2010. Application of iron nated activated carbon using the surface complexation approach. Water Res. 39
impregnated activated carbon for removal of arsenic from water. Iran J. Environ. (6), 1005e1014.
Health Sci. Eng. 7, 145e156. Viraghavan, T., Thirunavukkarasu, O.S., Suramanian, K.S., 2001. Removal of arsenic
Goldberg, S., Johnston, C.T., 2001. Mechanism of arsenic adsorption on amorphous in drinking water by iron oxide-coated sand and ferrihydrite-batch studies.
oxides evaluated using macroscopic measurements, vibrational spectroscopy, Water Qual. Res. J. Can. 36, 55e70.
and surface complexation modeling. J. Colloid Interface Sci. 234 (1), 204e216. Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption of carbon from solutions.
Guan, X., Dong, H., Ma, J., Jiang, L., 2009. Removal of arsenic from water: effects of J. Sanit. Eng. Div 2, 31e59.
competing anions on As(III) removal in KMnO4-Fe(II) process. Water Res. 43, Wilkie, J.A., Hering, J.G., 1996. Adsorption of arsenic onto hydrous ferric oxide:
3891e3899. effects of adsorbate/adsorbent ratios and co-occurring solutes. Colloid Surface.
Guo, H., Li, Y., Zhao, K., 2010. Arsenate removal from aqueous solution using A 107, 97e110.
synthetic siderite. J. Hazard. Mater. 176, 174e180. World Health Organization, 2008. Guidelines for Drinking-water Quality, third ed.

Вам также может понравиться