Вы находитесь на странице: 1из 6

Desalination 255 (2010) 78–83

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / d e s a l

Surface modification of coconut-based activated carbon by liquid-phase oxidation


and its effects on lead ion adsorption
Xiaolan Song a,⁎, Hongyan Liu a, Lei Cheng a, Yixin Qu b
a
Department of Inorganic Materials, School of Resources Processing and Bioengineering, Central South University, Changsha 410083, China
b
Department of Chemical Engineering, Beijing University of Chemical Technology, Beijing 100029, China

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

Article history: In this study, two series of functionalized samples were prepared by means of oxidation of a commercially
Received 18 August 2009 available activated carbon with either HNO3 or H2O2. The effect of the liquid-phase oxidation on the Pb2+
Received in revised form 10 January 2010 adsorption capacities of activated carbon was investigated. The porous structure of the functionalized
Accepted 16 January 2010
activated carbon was characterized using N2 adsorption at 77 K. The surface functional group characteristics
Available online 6 February 2010
were examined by Fourier transform infrared (FTIR) spectroscopy, acid/base titrations, Zeta potential as well
as the point of zero charge (pHPZC) measurement. The adsorption capacity for lead ions in aqueous solution
Keywords:
Activated carbon was found to depend on the amount of acidic oxygen functional groups. The isothermal adsorption data were
Modification measured and were fitted with the Langmuir and Freundlich isotherm model. The adsorption capacity of
Liquid-phase oxidation activated carbon oxidized with HNO3 (10 mol L− 1) at 363 K was about 2.5 times higher than that of the
Adsorption original activated carbon. The increase in the adsorption capacity of the oxidized activated carbons is
Lead ions attributed to the increased oxygen groups, which enhanced the hydrophilicity of the activated carbon,
lowered the pHPZC, made the surface more negatively charged, and increased the amount of homogeneous
active sites available for lead ions.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction aqueous solution is complex and may involve ion exchange,


electrostatic effects and coordination with functional groups on
Lead contamination in water and wastewater, which resulted from activated carbon surfaces. Several factors, such as specific surface
many industries such as metal plating, mining and tanneries, is a area, pore-size distribution, pore volume, surface charge, and
serious problem due to the poisoning effect on humans and to the presence of surface functional groups, affect the adsorption of metal
other adverse effects on receiving water [1–4]. A number of ions on activated carbon. So, special modification on activated carbon
techniques have been developed for removing lead pollutant from is necessary in order to enable activated carbons to develop affinity for
aqueous effluents to minimize its impact. These methods include metal ions in aqueous solution and remove them effectively.
precipitation, electroplating, evaporation, ion exchange, membrane The surface of activated carbons consists of hydrophobic graphene
separation, adsorption and various biological processes [1,4–9]. layers and hydrophilic functional groups. Organic compounds are
Among these methods, adsorption, which can remove trace amount adsorbed on the former, whereas polar species are adsorbed on the
of lead pollutant from aqueous solution [7,10,11], is highly effective later [18]. The treatment with oxidizing agent, either in gas phase or in
and economical. Activated carbons have well developed porous solution, introduces a large amount of oxygen-containing surface
structures, large surface area and a wide spectrum of surface complexes on the surface of activated carbons, which makes the
functional groups which can be modified by chemical and thermal activated carbon more hydrophilic and acidic. Although chemical
treatment procedures. Thus, activated carbons have been proven to be oxidation results in the fixation of both oxygen and nitrogen
effective adsorbents for the removal of a wide variety of organic and functional groups on the surface of activated carbon, it does not
inorganic pollutants from both aqueous and gaseous media [12–17]. significantly modify the textural properties of activated carbon. This
As an effective adsorbent for the adsorption of trace metal ions, treatment enhances its adsorption and modifies its selectivity to
activated carbon has the largest market share. However, the aqueous metal cation species [19,20]. The oxygen functional groups
mechanism of the adsorption of metal ions on activated carbon is on activated carbon can act as acids or bases, showing properties of
still not very clear. Adsorption of metal ions by activated carbon from ion exchange and coordination. For the oxidation treatment, various
reagents including concentrated nitric or sulfuric acid, sodium
hypochlorite, permanganate, bichromate, hydrogen peroxide, transi-
⁎ Corresponding author. Tel.: +86 731 8877 203; fax: +86 731 8710 804. tion metals and ozone-based gas mixtures have been used as oxidizers
E-mail address: xlsong365@126.com (X. Song). [21]. However, it is difficult to obtain optimal oxidation treatment

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.01.011
X. Song et al. / Desalination 255 (2010) 78–83 79

condition for a specific activated carbon only on the basis of results transform spectrophotometer (Nicolet Model Nexus 670, U.S.A) using
from literature without experimental determination because the pellets of KBr containing about 0.5% finely ground activated carbon
activated carbons reported are derived from different precursors and samples. These pellets were dried overnight at 393 K before the
their physical and chemical properties vary considerably. spectra were recorded and the spectra were the results of averaging
The objective of this work is to investigate the effect of surface 32 scans with a spectral resolution of 2 cm− 1 in the range of 400–
modification on the adsorption of trace lead ions from aqueous 4000 cm− 1.
solution. Two series of samples were prepared by means of oxidation
of a commercially obtained activated carbon with either HNO3, or 2.4. Acid/base titrations
H2O2. Prior to oxidation treatment, the activated carbon was deashed
with HCl because ash constituents usually give hydrophilic sites on The amphoteric characteristics of the treated activated carbon
the activated carbon surface. The changes of physicochemical samples were characterized by measuring the amount of the surface
properties of the activated carbons upon the treatment were functional groups using the acid–base titration method proposed by
investigated by N2 adsorption, Fourier transform infrared (FTIR) Boehm [23]. According to this method, activated carbon samples each
spectroscopy, acid/base titrations, Zeta potential and point of zero with 0.2 g were placed into 50 mL conical flasks containing 25 mL
charge (pHPZC). An optimal treatment condition was obtained based aqueous solution of the following material respectively: sodium
on the analysis of the relationship between the properties of surface hydroxide, sodium carbonate, sodium bicarbonate and hydrochloric
functional groups and adsorption capacity of the activated carbons. acid. The concentration of the aqueous solution was 0.05 M. The
containers were then sealed with rubber caps and shaken for 24 h at
2. Experimental section 298 K, after which the resulting suspensions were filtered and 10 mL
of accurately withdrawn aliquot of filtrates were titrated up to pH 4.5
2.1. Liquid-phase oxidation treatment by 0.05 M hydrochloric acid in order to estimate the residual base or
up to pH 11.5 by 0.05 M sodium hydroxide for the residual acid. The
A commercial coconut-shell derived activated carbon (obtained titration value was measured three times for each sample and then
from Shanghai Xinhuoli Activated Carbon Co., Ltd, China) was used as the amount of acidic/basic functional groups was calculated using the
the precursor. This activated carbon was ground and sieved to a final average of the three titration data.
particle between 50 and 60 mesh. Then, a fraction of it was purified
(mainly for elimination of ash impurities) by treatment with 2.5. pH at the point of zero charge (pHPZC)
HCl (2 mol L− 1) at room temperature for 12 h. The amount of HCl
solution used was 2 ml g− 1. After the HCl-purification, the activated The pH at the point of zero charge (pHPZC) of the activated carbon,
carbon was washed with deionized water until no further change in namely the pH value required to give zero net surface charge, was
pH could be detected, and then it was dried at 383 K for 24 h (denoted measured using mass titration method proposed by Noh and Schwarz
as AC0). The HCl-treated activated carbon was subsequently subjected [22]. For the measurement of the pHPZC, three solutions with different
to liquid-phase oxidation treatment as follows: (i) oxidation of initial pH values, pH = 3, 6, 11, were prepared with addition of 0.05 M
sample AC0 with 10 mol L− 1 H2O2 at room temperature and at 363 K aqueous solution of HNO3 or NaOH. NaNO3 was used as the
for 24 h (samples labeled as AC1 and AC2, respectively); (ii)oxidation background electrolyte. For each initial pH, ten containers were filled
of sample AC0 with 10 mol L− 1 HNO3 at room temperature and at with 20 ml of the solution and different amounts of activated carbon
363 K for 4 h (samples labeled as AC3 and AC4 respectively) and with were added (0.05, 0.1, 0.25, 0.5, 1, 1.5, 2.5, 4, 6, and 10% by weight).
5 mol L− 1 HNO3 at 363 K for 4 h (labeled as AC5). The oxidation was The containers were sealed under N2 and placed on a thermostatic
conducted under stirring in a three-necked flask equipped with a swing shaker for 24 h. Then the carbon was separated and the pH was
stirrer and a reflux condenser and heated in a thermostat water bath. measured. A plot of the equilibrium pH versus mass fraction yields a
The amount of the oxidizing solution used was 10 ml g− 1. The curve showing a plateau and the pHPZC is identified as the point at
oxidized activated carbons were washed repeatedly with deionized which the change of pH is zero. The pHPZC is then taken as the average
water until the pH of the effluent water was constant, dried at 383 K of the three asymptotic pH values.
for 24 h and stored in a desiccator for later use. For the determination
of pHPZC and pHIEP, the measurement was carried out 10 min after the 2.6. Zeta potential measurement
oxidation treatments in order to avoid atmospheric oxidation as much
as possible. Zeta potential was obtained using a Zeta Potential Analyzer
(Coulter Delsa440SX, U.S.A). Samples prepared for determination of
2.2. Textural characterization pHPZC were also used for the zeta potential measurements. Approx-
imately 0.5 g of fresh sample with a particle size below 0.038 mm
The textural characterization of the activated carbons was carried (400 mesh) was added to a plastic bottle containing 1 L of freshly
out using N2 adsorption at 77 K with a Quantachrome Autosorb-1 outgassed CO2-free deionized water, sealed under N2 and hand-
apparatus. Prior to the N2 adsorption measurement, samples were shaken periodically during 24 h to ensure complete wetting of the
outgassed under vacuum at 423 K for 6 h. The cross sectional area of sample. The suspension was dispersed by ultrasonic before being
nitrogen at 77 K was taken as 0.162 nm2. Specific surface areas were transferred to the electrophoresis cell. The pH of the slurries measured
calculated via the BET model at relative pressures of P/P0 = 0.01–0.1. by a pHS-3C pH meter was adjusted by adding either HCl or NaOH. In
The total pore volume was estimated from the uptake of nitrogen at a each case, the zeta potential was measured three times for each pH to
relative pressure of P/P0 = 0.99. Surface area and pore volume of provide an average value.
micro-pores were calculated via t-plot analyses as a function of
relative pressure using the Carbon Black model for thickness curves 2.7. Adsorption of lead ions
measured between 4.5 and 7.0 Å.
The stock solution of Pb2+ was prepared from lead nitrate
2.3. Fourier transform infrared (FTIR) spectroscopy (analytical reagent grade) at a concentration of 1 g L− 1. Experimental
solutions of lead were prepared by appropriately diluting a different
The surface functional groups of some selected samples were volume of stock solution to achieve the desired concentration. The pH
investigated by transmission infrared spectra obtained from a Fourier of the working solutions was adjusted to 4 by adding different
80 X. Song et al. / Desalination 255 (2010) 78–83

concentrations of NaOH solution. Fresh dilutions were used for each


experiment. 0.2 g of activated carbon (<60 mesh) was added to
100 mL of the aqueous solution within a conical flask, which was
partially immersed in a water bath and was hand-shaken periodically
to equilibrate at 298 K. The time of 48 h was sufficient to achieve
equilibrium according to the kinetic experiments showed by Thomas
[24]. The amounts of lead ions adsorbed on the activated carbon were
calculated from the mass balance expression according to Eq. (1).

q = ðC0 −Ce ÞV = m ð1Þ

Where q is the amount of lead ions adsorbed onto unit amount of


adsorbent (mg g− 1); C0 and Ce are the initial and equilibrium
concentration of lead ions (mg L− 1), respectively; V is the volume of
aqueous phase (L) and m is the mass of the adsorbent (g). The
aqueous lead ions concentration in solution was determined by
atomic absorption spectrophotometer (SHIMADZU AA-6800, Japan).
The pH values of solutions after adsorption were also measured.

3. Results and discussion

3.1. Textural structure of the activated carbons

All activated carbons presented quite similar nitrogen adsorption


isotherms with a saturation plateau at high relative pressures (not Fig. 1. FTIR spectra of the activated carbons.
shown here), which are of type I according to the IUPAC classification.
The results of the pore structural characterization of samples are listed
in Table 1. From these data, it can be seen that treatment with HNO3 carbons and produces new IR adsorption peaks. Strong bands at
and H2O2 at room temperature does not significantly modify the ∼3430, which are assigned to the O_H stretching vibration due to the
textural properties of activated carbon. Oxidation treatment at 363 K existence of surface hydroxylic groups and chemisorbed water, are
gives rise to a slight decrease in the surface area and the pore volume. observed for all the samples. The bands at 1717 cm− 1 observed for
The decrease in surface area and volume brought about by stronger AC2, AC3, AC4 and AC5 are the characteristic of C=O group,
oxidation is generally explained by the restriction of the pore volume demonstrating that C=O groups are produced in these samples
available for N2 adsorption due to the formation of oxygen-containing upon oxidation. The difference in the intensity of the 1717 cm− 1
groups at the entrance and/or on the walls of micro-pores and by the between AC1 and AC2 as well as between AC3 and AC4 indicates that
possible destruction of the pore walls and its collapse when oxidation of the activated carbon with either H2O2 or HNO3 at higher
oxygenated terminal groups are created [18,25,26], especially for temperature produces more C=O groups. The difference in the
H2O2 oxidation treatment which results in a severe damage of pore intensity of the 1717 cm− 1 between AC4 and AC5 indicates that
walls. Besides the above discussed textural changes, treatment with oxidation of the activated carbon with higher concentration of HNO3
HNO3 at higher temperature, e.g. AC4 and AC5, results in the fixation also produces more C=O groups. Other than this, the IR spectra of
of substantial oxygen-containing groups (will be discussed later) on sample AC4 and AC5 show shoulder peaks at 1690 cm− 1. This band is
the walls of micro- and meso-pores, leading to a decrease of the pore- probably due to the presence of ketone groups. It has been reported
size, and therefore to a decrease in the meso-pore volume and surface that the C=O stretching vibration of carboxyl acid group in an aromatic
area. ring generally appear at 1700–1680 cm− 1 [27]. At 1645 cm− 1, the IR
spectra of samples AC1 and AC3 show a defined peak which can be
3.2. Fourier transform infrared (FTIR) spectroscopy attributed to the stretching vibration of C=O moieties in quinine and/or
ion radical structures, while samples AC2, AC4 and AC5 only show a
Fig. 1 shows the FTIR spectra of the activated carbons. It can be shoulder [28–32].
seen that liquid-phase oxidation of the activated carbons modifies the The bands in the region of 1600 cm− 1 (at 1586 cm− 1) has been
intensity of IR adsorption peaks presented in the original activated observed by many authors and has not been interpreted unequivo-
cally. This band is assigned to aromatic ring stretching coupled to
highly conjugated hydrogen-bonded carbonyl groups (C=O) [32].
Table 1
Some small weak peaks in the region of 1520–1410 cm− 1 consist of a
Textural characterization of the activated carbons.
series of overlapping absorptions that can be ascribed to carboxyl–
Activated Surface area (m2 g− 1) Pore volume (cm3 g− 1) d
APD carbonates structures [31,32] and to the deformation vibrations of
carbon a b c b (nm) surface hydroxyl groups [31], or to the in-plane C–H vibrations in
Total Micropore Total Micropore
various surface C=C–H structures [28].
AC0 1245 1126 0.5517 0.4482 1.817
AC1 1199 1107 0.5405 0.4453 1.803 Assignment of the bands in the region of 1000–1200 cm− 1 is
AC2 1062 970 0.5120 0.3837 1.928 difficult because there is superposition of a number of broad
AC3 1162 1085 0.5251 0.4426 1.808 overlapping bands. They cannot, therefore, be described in terms of
AC4 1053 1036 0.4359 0.4120 1.657
simple motion of specific functional groups or chemical bonds [33].
AC5 1070 1017 0.4588 0.4044 1.716
Some authors assign this overlapping band to ether (symmetrical
a
Specific surface area was determined by BET method within relative pressures of P/ stretching vibrations), epi-oxide and phenolic structures existing in
P0 = 0.01–0.1.
b
Surface area and pore volume of micro-pores were calculated by the t-plot method.
different structural environments [31–33]. The peak at 680 cm− 1 can
c
Total pore volume measured at a relative pressure of P/P0 = 0.99. be assigned to out-of-plane deformation vibrations of C–H groups
d
APD: average pore diameter. located at the edges of aromatic planes.
X. Song et al. / Desalination 255 (2010) 78–83 81

3.3. Acid/base titrations basic character such as ACO will readily adsorb protons from solution,
whereas those with acidic character and lower pHPZC value will
The acid/base titration results of the samples are shown in Table 2. release protons to form net negative surface charges, which is
Taking into account that NaOH titrates carboxyl, lactone and phenolic beneficial for the aqueous lead cation adsorption.
groups, Na2CO3 titrates carboxyl and lactone, and NaHCO3 titrates
only carboxyl groups, one can therefore obtain the amount of the
different acid groups present on the activated carbons [26]. The data 3.5. Zeta potential measurement
presented in Table 2 indicate that the original carbon AC0 does not
react with Na2CO3 and NaHCO3 due to the absence of strongly acidic As mentioned above the activated carbons show amphoteric
surface groups such as carboxylic groups. Oxidation treatment results property, therefore, it is possible to evaluate their isoelectric point
in an increase in the total surface acidity and a decrease in the total (pHIEP), i.e. the pH at which the zeta potential is zero. Fig. 2 gives the
surface basicity. This phenomenon is due to the presence of different results of zeta potential measurements and Table 2 gives the pHIEP
types and concentrations of functional groups which give the values of the original and the oxidized carbons. The pHIEP value of the
amphoteric characteristics to activated carbons in aqueous solution. original carbon is 3.5. Oxidation results in a decrease of the pHIEP
The greatest surface acidity was obtained after drastic nitric acid values, especially those of AC4 and AC5, which pHIEP could not be
oxidation treatment. An intermediate surface acidity and basicity was detected in the pH range of 2–12 due to the presence of weakly acidic
obtained with a treatment by concentrated hydrogen peroxide at carboxylic functional groups with lower dissociation pH values. Thus,
363 K. The concentration of lactone and carboxylic groups on the dissociation of the groups in the pH range of 2–5 renders the
activated carbon surface after different oxidation treatment follows surface of AC4 and AC5 to be negatively charged.
the sequence AC4 > AC5 > AC2 > AC3 > AC1 > AC0. It has been suggested that the isoelectric point value is represen-
It is generally accepted that oxidization treatment leads to the tative of the external charges of carbon particles in solution, whereas
incorporation of a great number of acidic oxygen-containing groups the point of zero charge varies in response to the net total (external
on the surface of the activated carbon. However, as shown in Table 2, and internal) surface charge of the particles [23,35]. The difference
the oxidized activated carbons always contain basic surface sites in between pHPZC and pHIEP values would give an indication of the
addition to the acidic sites, though their number is usually smaller surface charge distribution of the porous activated carbon. As shown
than that of the original carbon (AC0). The basic sites are of Lewis type in Table 2, when the original sample was oxidized with HNO3 and
and are associated with the presence of basic oxygen and nitrogen H2O2, both the pHPZC and pHIEP values became lower than those
functional groups, such as pyrone-like structures, and with π electron- observed for the original sample, indicating an increase in acidic
rich regions within the basal planes of activated carbon [23,34]. In oxygen-containing surface groups [36–38]. However, for these
contrast, the acidic oxygen functional groups, incorporated by liquid- samples, as pHPZC > pHIEP, the internal surface is always more basic
phase oxidation show characteristics of Brønsted acid. than the external surface. Upon oxidation, the pHPZC values, all in the
acid range, decrease to a greater extent than the pHIEP values,
although pHPZC is still higher than pHIEP and the difference between
3.4. pH at the point of zero charge (pHPZC) pHPZC and pHIEP decreases sharply from the original carbon to the
oxidized carbons. This means that the external surface is somewhat
The point of zero charge describes the condition when the more acidic than the internal surface, but the introduction of the
electrical charge density on a surface is zero. The pH of point of zero acidic oxygen-containing groups essentially takes place on the
charge (pHPZC) of an activated carbon depends on the chemical and internal surface, showing a uniform oxidation. In the cases of AC4
electronic properties of the functional groups on its surface and is a and AC5, the lowest {pHpzc–pHIEP} difference shows that more
good indicator of these properties. Table 2 gives the pHPZC values of uniform oxidation took place, and will render more acidic sites to
the original and the oxidized activated carbons. The original carbon release protons for exchange with lead ions in the adsorption process.
AC0 has a pHPZC value of 10.1. Upon oxidation treatment, the pHPZC of
the activated carbons decreased markedly due to an increase in the
amount of the surface acidic oxygen-containing groups. Carbons
oxidized with nitric acid at 363 K (AC4 and AC5) give the lowest pHPZC
values about 3.1. The pHPZC of AC3 oxidized with nitric acid at ambient
temperature is 4.9, which is significantly higher than that of AC4 and
AC5. This means that oxidation of the activated carbon at higher
temperature produce higher amount of acidic groups. This is also the
case for the carbons oxidized with H2O2 (AC1 and AC2).
As shown in Table 2, the amount of the basic sites on AC1, AC2, and
AC3 is much more than that on AC4 and AC5. Activated carbons with

Table 2
Acid/base titration and pHPZC and pHIEP of the activated carbons used in this study.

Activated pHIEPa pHPZC Surface functional groups (mmol g− 1)


carbon
NaOH NaHCO3 Na2CO3 HCl

AC0 3.5 10.1 0.07 0.00 0.00 0.67


AC1 2.4 5.2 0.86 0.25 0.39 0.31
AC2 2.1 4.1 1.28 0.36 0.56 0.24
AC3 2.3 4.9 1.02 0.29 0.43 0.28
AC4 ND 3.1 2.15 1.08 1.62 0.09
AC5 ND 3.2 1.96 0.84 1.37 0.12
a
ND = not detected. Fig. 2. ζ-potential–pH curves for different activated carbons.
82 X. Song et al. / Desalination 255 (2010) 78–83

Freundlich isotherm models. The isotherms are mathematically


represented as:

Q0 KL Ce
qe = ð2Þ
1 + KL Ce

1=n
qe = KF Ce ð3Þ

Where

qe (mg g− 1) is the equilibrium adsorbate concentration on the


adsorbent,
Q0 (mg g− 1) is the monolayer adsorption capacity of the adsorbent,
KL (L mg− 1) is the Langmuir constant related to the free energy of
adsorption,
Ce (mg ml− 1) is the equilibrium concentration of the adsorbate,
KF (mg g− 1 (L mg− 1)1/n) and 1/n are the Freundlich constants related
to the adsorption capacity and intensity, respectively.

Fig. 3 shows the experimentally obtained adsorption isotherms for


the uptake of lead ions in aqueous solution and the fitted lines
according to Langmuir and Freundlich models. The values of
regression parameters and the value of the regression coefficient R2,
evaluated by least square method, are listed in Table 3. Results
presented in Fig. 3 and the R2 values given in Table 3 indicate that the
isotherm of all the adsorbents can be well fitted using Langmuir
model because the adsorption isotherm is concave to the equilibrium
concentration axis and tends to define a plateau [39].
Table 3 also gives the maximum adsorption capacity of the original
and the oxidized activated carbons. It can be seen that the liquid-
phase oxidation with HNO3 at 363 K greatly enhances the adsorption
capacity of activated carbon. The adsorption capacity of oxidized
activated carbons follows the order of AC4 > AC5 > AC2 > AC3 > AC1 >
AC0. It correlates neither with the specific surface area nor pore
volume, but is in agreement with the sequence of the amount of the
surface acidic oxygen-containing groups. This means that the acid
groups on the activated carbons are responsible for the adsorption of
Fig. 3. Adsorption isotherms of Pb2+ on the activated carbons oxidized with H2O2 (A) the lead ions.
and HNO3 (B). (Solid line: Langmuir; dashed line: Freundlich). For AC0, AC1, AC2 and AC3, the pH of the solution increases since
the activated carbons not only adsorb lead cations from solution but
also adsorb H+(aq). Moreover, competitive adsorption exists between
H+(aq) ions and Pb2+(aq) ions, and the protonation of basic
3.6. Pb2+ equilibrium adsorption isotherms functional groups reduces the sites available for the adsorption of
Pb2+(aq) ions [18]. On the other hand, the presence of acid oxygen
The equilibrium adsorption isotherms are of fundamental impor- groups on the surface of activated carbons lead to a very low point of
tance in the study and design of adsorption systems. The adsorption of zero charge values [40]. When pH of the solution is higher than pHPZC,
aqueous species on activated carbons may be described by various the surface of the activated carbon are negatively charged and the
isotherms including Langmuir, Freundlich, Sips and Redlich–Peterson adsorption of cations is favored, and vice versa. For AC4 and AC5, the
isotherm equations [40]. In this study, the regression analysis of pH of the solution is lower than that of the initial aqueous solution
experimental data has been carried out using both the Langmuir and (4.0), which can be attributed to the ion exchange reaction between

Table 3
Values of the parameters for Langmuir model and Freundlich model.
a
Activated pHads Initial concentration Langmuir parameters Freundlich parameters
carbon (mg L− 1)
Q0 (mg g− 1) KL (L mg− 1) b 2
R KF (mg g− 1 (L g− 1)1/n) 1/n b 2
R

AC0 5.3–6.1 6–200 17.193 0.100 0.961 4.446 0.269 0.955


AC1 4.4–5.5 6–200 22.792 0.203 0.962 8.785 0.201 0.932
AC2 4.3–4.7 5–200 28.458 0.171 0.972 11.501 0.186 0.931
AC3 4.4–4.8 6–210 25.197 0.158 0.981 8.657 0.222 0.891
AC4 3.2–3.6 20–260 40.119 0.161 0.943 12.533 0.237 0.939
AC5 3.2–3.7 20–260 37.938 0.099 0.931 9.955 0.265 0.935
a
pHads gives the pH range variation for the isotherm.
b
R2 is the regression coefficient.
X. Song et al. / Desalination 255 (2010) 78–83 83

Acknowledgements

The authors express thanks for the financial support by the


National High Technology Research and Development Program
(“863”Program) of China (2009AA05Z436) and by the Undergraduate
Innovation Program of Central South University (LC09099).

References
[1] T.G. Chuah, A. Jumasiah, I. Azni, S. Katayon, S.Y. Thomas Choong, Desalination 175
(2005) 305–316.
[2] S. Çay, A. Uyanik, A. Ozasik, Sep. Purif. Technol. 38 (2004) 273–280.
[3] M. Sekar, V. Sakthi, S. Rengaraj, J. Colloid Interface Sci. 279 (2004) 307–313.
[4] B. Volesky, in: B. Volesky (Ed.), Biosorption of Heavy Metals, CRC Press, Florida,
1990, pp. 7–43.
[5] M. Kobya, E. Demirbas, E. Senturk, M. Ince, Bioresour. Technol. 96 (2005) 1518–1521.
[6] M. Helen Kalavathy, T. Karthikeyan, S. Rajgopal, Lima Rose Miranda, J. Colloid
Interface Sci. 292 (2005) 354–362.
[7] M. Madhava Rao, A. Ramesh, G. Purna Chandra Rao, K. Seshaiah, J. Hazard. Mater.
129 (2006) 123–129.
Fig. 4. Relationship between adsorption capacity and acid density. [8] K.G. Sreejalekshmia, K. Anoop Krishnanb, T.S. Anirudhana, J. Hazard. Mater. 161
(2009) 1506–1513.
[9] R. Sabry, A. Hafez, M. Khedr, A. El-Hassanin, Desalination 212 (2007) 165–175.
[10] Liliana Giraldo-Gutiérrez, Juan Carlos Moreno-Piraján, J. Anal. Appl. Pyrol. 81 (2008)
278–284.
[11] Özgül Gerçel, H. Ferdi Gerçel, Chem. Eng. J. 132 (2007) 289–297.
the hydrogen ions and lead ions and thus leads to a higher adsorption [12] M. Streat, D.J. Horner, Process Saf. Environ. Prot. 78 (2000) 363–382.
capacity. [13] I.H. Suffet, L. Brenner, J.T. Coyle, P.R. Cairo, Environ. Sci. Technol. 12 (1978) 1315–1322.
A correlation of the adsorption capacity with acid density is shown [14] B. Saha, M.H. Tai, M. Streat, Process Saf. Environ. Prot. 79 (2001) 345–351.
[15] S. Biniak, M. Pakua, G.S. Szymanski, A. Swiatkowski, Langmuir 15 (1999) 6117–6122.
in Fig. 4. A linear relationship is observed with correlation coefficient [16] P. Chingombe, B. Saha, R.J. Wakeman, Carbon 43 (2005) 3132–3143.
of 0.9926. Therefore, it can be inferred that the adsorption capacity is [17] R. Ayyappan, A. Carmalin Sophia, K. Swaminathan, S. Sandhya, Process Biochem.
proportional to the acid density. 40 (2005) 1293–1299.
[18] B. Xiao, K.M. Thomas, Langmuir 21 (2005) 3892–3902.
[19] Y.F. Jia, K.M. Thomas, Langmuir 16 (2000) 1114–1122.
4. Conclusion [20] Y.F. Jia, B. Xiao, K.M. Thomas, Langmuir 18 (2002) 470–478.
[21] A. Lisovskii, R. Semiat, C. Aharoni, Carbon 35 (1997) 1639–1643.
[22] J.S. Noh, J.A. Schwarz, J. Colloid Interface Sci. 139 (1990) 139–148.
Liquid-phase oxidation with H2O2 and HNO3 significantly in- [23] H.P. Boehm, Carbon 32 (1994) 759–769.
creased the acidic oxygen-containing groups on the surface of [24] B. Xiao, K.M. Thomas, Langmuir 20 (2004) 4566–4578.
activated carbons without leading to marked physical changes. [25] C. Moreno-Castilla, F. Carrasco-Marin, F.J. Maldonado-Hodar, J. Rivera-Utrilla,
Carbon 36 (1998) 145–151.
Activated carbon oxidized with HNO3 at 363 K shows the lowest
[26] C. Moreno-Castilla, F. Carrasco-Marin, A. Mueden, Carbon 35 (1997) 1619–1626.
{pHpzc–pHIEP} difference, implying that uniform oxidation has taken [27] S. Shin, J. Jang, S.-H. Yoon, I.A. Mochida, Carbon 35 (1997) 1739–1743.
place, the surface of such treated activated carbons becomes more [28] A. Macias-Garcia, C. Valenzuela-Calahorro, V. Gomez-Serrano, A. Espinosa-
Mansilla, Carbon 31 (1993) 1249–1255.
homogeneous and the highest amount of acid carboxyl groups is
[29] Dong Jin Suh, Park Tae-Jin, Ihm Son-Ki, Carbon 31 (1993) 427–435.
generated. These render the activated carbon having more acidic sites [30] P. Vinke, M. van der Eijk, M. Verbree, A.F. Voskamp, H. van Bekkum, Carbon 32
to release protons to exchange with lead ions in the adsorption (1994) 675–686.
process. [31] V. Gomez-Serrano, M. Acedo-Ramos, A.J. Lopez-Peinado, C. Venezuela-Calahorro,
Fuel 73 (1994) 387–395.
The maximum adsorption capacity increased from 17.19 mg g− 1 [32] Paul E. Fanning, M. Albert Vannice, Carbon 31 (1993) 721–730.
on the original activated carbon to 40.12 mg g− 1, the highest value [33] P. Painter, M. Starsinic, M. Coleman, Fourier Transform Infrared Spectroscopy,
obtained with the carbon oxidized in 10 M concentrated HNO3 at Academic Press, New York, 1985, pp. 169–189.
[34] C.A. Leony y Leon, J.M. Solar, V. Calemma, L.R. Radovic, Carbon 30 (1992) 797–811.
363 K for 4 h. The maximum adsorption capacity is well correlated [35] M.O. Corapcioglu, C.P. Huang, Carbon 25 (1987) 569–578.
with the acid density of the activated carbons and is proportional to [36] Gayle Newcombe, Rob Hayes, Mary Drikas, Colloid. Surf. A 78 (1993) 65–71.
the acid density. The adsorption equilibrium data of Pb2+ ions on the [37] Joong S. Noh, James A. Schwarz, Carbon 28 (1990) 675–682.
[38] Yoshinobu Otake, Robert G. Jenkins, Carbon 31 (1993) 109–121.
activated carbons have been measured. It was found that these data [39] Amir Fouladi Tajar, Tahereh Kaghazchi, Mansooreh Soleimani, J. Hazard. Mater.
could be well fitted by Langmuir isotherm model. Based on these 165 (2009) 1159–1164.
results, the values of the parameters of the Langmuir isotherm model [40] C.O. Ania, J.B. Parra, J.J. Pis, Fuel Process. Technol. 79 (2002) 265–271.
were estimated.

Вам также может понравиться