Você está na página 1de 13

Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

High removal efficiency of Hg(II) and MeHg(II) from aqueous solution


by coconut pith—Equilibrium, kinetic and mechanism analyses
Norasikin Samana , Khairiraihanna Joharib , Shiow-Tien Songa , Helen Konga ,
Siew-Chin Cheua , Hanapi Mata,c,*
a
Advanced Materials and Process Engineering Laboratory, Faculty of Chemical and Energy Engineering, Universiti Teknologi Malaysia, 81310 UTM Skudai,
Johor, Malaysia
b
Department of Chemical Engineering, Universiti Teknologi PETRONAS, 32610 Bandar Seri Iskandar, Perak, Malaysia
c
Advanced Materials and Separation Technologies (AMSET) Research Group, Health and Wellness Research Alliance, Universiti Teknologi Malaysia, 81310 UTM
Skudai, Johor, Malaysia

A R T I C L E I N F O A B S T R A C T

Article history:
Received 15 February 2016 This study was conducted to investigate the high efficiency of coconut pith (CP) adsorbent in removing
Received in revised form 11 April 2016 Hg(II) and MeHg(II) ions from aqueous solution. The CP adsorbent was characterized using a scanning
Accepted 30 April 2016 electron microscope (SEM), nitrogen adsorption-desorption (NAD) analysis, determination of pH at zero
Available online 2 May 2016 point charge (pHpzc), Fourier transform infrared (FTIR) spectroscopy and X-ray photoelectron
spectroscopy (XPS). Adsorption performance of the CP adsorbent at various parameters was conducted
Keywords: by varying the initial pH of solution, adsorbent dosage, initial pH, temperature, and contact time. It was
Coconut pith found that adsorption capacity, adsorption rate and enthalpy of the Hg(II) adsorption were higher
Adsorption
compared to the MeHg(II). The adsorption capacity of Hg(II) was 2.60 mmol/g, which was five times
Mercury
higher than MeHg(II). The adsorption isotherm analysis showed that the Hg(II) and MeHg(II) adsorption
Equilibrium
Kinetics data fitted to the Langmuir and Freundlich models, respectively. The overall mechanism of both mercury
Mechanism adsorptions is a combination of physical and chemical processes in which the film diffusion was the rate
controlling-step. The adsorbent regenerability study results showed that the Hg(II) adsorption remained
stable up to five adsorption cycles, which was better than MeHg(II). The selectivity studies reveal the
potential application of the CP adsorbent for the treatment of oilfield produced water (OPW) and natural
gas condensate (NGC) that are rich in mercury ions as well as other cations.
ã 2016 Elsevier Ltd. All rights reserved.

1. Introduction wastewater and soil promotes the easy transformation of the


inorganic mercury to a more toxic organic mercury such as
The presence of mercury in the environment has become of a methylmercury [2,3]. Since various transformations of mercury in
great public concern due to its toxicity in which poses a significant the environment are generated from ionic mercury, it is necessary
threat to the public health and ecological systems [1]. In the to control the low level of mercury discharge into the environment.
environment, mercury may exist in three different forms namely Numerous treatment methods such as chemical precipitation,
elemental mercury (Hg ), organic mercury (e.g. methylmercury, conventional coagulation, adsorption, ion exchange, and reverse
MeHg) and inorganic mercury (e.g. HgCl2). Organic mercury osmosis have been applied to control the discharge of mercury
especially methylmercury is recognized as the most toxic mercury from various effluents [4]. Among these methods, adsorption is
form to humans [2]. The inorganic mercury is less adsorbed by more preferred due to its simplicity, economical and high quality of
living organisms as compared to the elemental mercury and treated effluents [5]. In an adsorption process, the removal
organic mercury. However, in the presence of organic matters in efficiency depends on the properties of the adsorbate and surface
chemistry of the adsorbent [6]. Materials such as silicas, carbons,
zeolites, synthetic polymers, agricultural residues, biomass, peats,
and industrial residues, are the most common adsorbents that
* Corresponding author at: Advanced Materials and Process Engineering
have been used to adsorb mercury in aqueous solutions. However,
Laboratory, Faculty of Chemical and Energy Engineering, Universiti Teknologi
Malaysia, 81310 UTM Skudai, Johor, Malaysia. some adsorbents are expensive; thus, persistent efforts have been
E-mail address: hbmat@cheme.utm.my (H. Mat). made for developing highly effective, low-cost and locally available

http://dx.doi.org/10.1016/j.jece.2016.04.033
2213-3437/ ã 2016 Elsevier Ltd. All rights reserved.
2488 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

adsorbents [7]. The waste materials from agricultural industry are mercury ions solutions, which were prepared using deionized
a common source of low-cost adsorbents due to their abundance water produced by using the Purite Water System (U.K).
and less expensive as compared to other materials [8].
The agricultural materials like other lignocellulosic materials 2.2. Methods
consist of lignin, hemicellulose, cellulose, lipids, proteins, simple
sugar and other organic matter [7]. These components have variety 2.2.1. Adsorbent preparations
of oxygen functional groups (e.g. hydroxyl, carboxyl, phenol and The CP was ground, sieved into particle size range of between
ketone) that facilitate the adsorption of various pollutants. Many 75 mm and 100 mm, and washed repeatedly using double-distilled
researchers have published and highlighted the use of agricultural water to eliminate water-soluble impurities. After washing, the
residues as low-cost materials to adsorb mercury ions such as sample was dried in an oven at 50.0  1.0  C overnight and further
coconut fruit/tree residues [9–11], grains residues [12,13], sago dried in a vacuum oven at 30.0  1.0  C for 5 h. The sample was then
residues [14] fruits and vegetables residues [15–17]. These stored in a desiccator for characterizations and adsorption
agricultural residues have been utilized as adsorbents either in experiments.
pristine or modified forms.
The utilization of unmodified agricultural residues (usually 2.2.2. Adsorbent characterizations
used after washing, drying and grinding) can reduce the Moisture and ash contents were determined using a proximate
production and energy costs associated with chemical and thermal chemical analysis of biomass, according to ASTM E1756-08 [19]
treatments of the agricultural residues in adsorbent synthesis [8]. and ASTM E1755-01 [20], respectively. It involved a measurement
Although variety of residues have been studied as potential of weight loss, following the combustion of about 1 g of sample in a
adsorbent for targeting mercury ions, however, most of them have ceramic crucible at 105.0  1.0  C and 575  1.0  C, respectively. The
only focused on inorganic mercury, Hg(II) removal. Adsorbents pH at point zero charge (pHpzc) was determined using a
such as desiccated coconut waste [10], coconut fiber [11], rice potentiometric titration method. A series of 25 mL of 0.1 M
straw [12,13], and Carica papaya wood [16] show a high adsorption KNO3 having an initial pH values between 2.0 and 12.0 were
capacity towards Hg(II). However, they cannot guarantee to have a prepared. The CP adsorbent (50 mg) was placed in each flask, then
high adsorption capacity towards organic mercury ions such as was agitated at 200 rpm in the controlled temperature shaker for
methyl mercury and phenyl mercury. The development of 48 h; which was sufficient to achieve equilibrium. The initial pH
adsorbents that is capable of removing both organic ions is values (pHi) of the solutions were adjusted using either HCl or
required since in many cases both mercury ions may exist together NaOH solution. The equilibrium pH (pHe) and final pH (pHf) were
in the waste solutions. The various mercury ions in waste solution measured using a Mettler Toledo Delta 320 pH meter. The pHpzc
may also result from transformations that occur within the waste was obtained from the plot of DpH (pHi–pHf) against pHi; the
solutions [18]. intersection point at which the DpH is equal to zero (DpH = 0).The
In this study, the performance of coconut pith adsorbent in the cation-exchange capacity (CEC) of the CP adsorbent was deter-
removal of inorganic and organic mercury ions (Hg(II) and MeHg mined via the barium chloride method [21].
(II)) was investigated. The coconut pith was used widely as The surface morphology of the CP adsorbent was analyzed
adsorbents for removing various heavy metals such as Cr(IV), Cd using a JEOL model JSM-6390LV scanning electron microscope
(II) and Ni(II). So far, there has been no study conducted to (SEM). The adsorbent was coated with a thin gold layer and
investigate the comparative adsorption behavior of both mercury scanned at an accelerating voltage of 10 kV. The surface area and
ions. Thus, the batch adsorption of both mercury ions at various porosity of the adsorbent were determined by using a nitrogen
adsorption conditions such as pH, concentration, loading, contact adsorption-desorption (NAD) isotherm analysis measured using a
time, and temperature was investigated. The adsorption data surface area analyzer model Micromeritics ASAP 2020 (USA). The
were analyzed using the existing models to evaluate the sample was first degassed at 120  C overnight and the NAD analysis
mechanisms of the adsorption process. The mercury-loaded was measured at 196.15  C. A Fourier transform infrared (FTIR)
adsorbent was characterized using various analytical techniques spectrophotometer (PerkinElmer Model 2000) was used to
in order to obtain a better understanding on the nature of the identify the functional groups on the adsorbent surfaces before
mercury ion interactions and their adsorption mechanisms with and after mercury adsorption. The FTIR spectra was analyzed using
the surface of the CP adsorbent, which could lead to an a KBr disc method and was scanned at wavenumber range from
enhancement in mercury adsorption. The adsorption selectivity 4000 cm1 to 400 cm1. The X-ray photoelectron spectroscopy
of both mercury ions over other metal ions as well as adsorbent (XPS) analysis was conducted using an AMICUS XPS instrument
regenerability was also investigated due to the industrial (Kratos Analytical Inc., UK) equipped with a monochromatic Al Ka
importance, found in many industries such as the removal of radiation at 1486.8 eV, operated at 15 kV and 10 mA. The FTIR and
mercury and other metals from natural gas condensate and XPS analyses were carried out for the adsorbents before and after
produced water from oil and gas exploration activities. Hg(II) and MeHg(II) adsorptions.

2. Materials and methods 2.2.3. Adsorption studies


The adsorption performance of Hg(II) and MeHg(II) was studied
2.1. Materials using a batch procedure. The stock solutions of 5 mmol/L of each
Hg(II) and MeHg(II) were prepared from Hg(NO3)2H2O and
Cocos nucifera pith (or coconut pith) was purchased from the CH3HgCl salts, respectively. The stock solutions were diluted as
T&H Coconut Fiber Sdn. Bhd. (Johor, Malaysia). Methylmercury (II) required using double-distilled water.
chloride (CH3HgCl, 13% Cl), sodium hydroxide (NaOH, 99%), nitric The batch adsorption experiment was carried out at equilibri-
acid (HNO3, 65%), hydrochloric acid (HCl, 37%), potassium nitrate um and at different contact times. In equilibrium adsorption
(KNO3, 99.0%), potassium iodide (KI, 99.5%), were purchased from experiments, effects of solution pH, adsorbent dosages, initial
Merck (Germany), while mercury (II) nitrate monohydrate (Hg mercury concentrations, and temperatures were investigated. For
(NO3)2H2O, 99%) was purchased from Sigma-Aldrich (USA). All each experiment, 50 mg of adsorbent was placed in a 125 mL
chemicals were used as received. The double-distilled water was polycarbonate Erlenmeyer flask having 50 mL of aqueous mercury
used for all experimental studies except for the preparing of stock solutions of known concentration and shaken two days at 200 rpm
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2489

using a controlled temperature shaker at 30.0  1.0  C. After two desorption solution and the mixture was shaken overnight at
days of shaking, which was long enough to achieve equilibrium 30.0  1.0  C using a controlled temperature shaker. The mixture
conditions, the mixture was filtered using nylon membrane filters was then filtered and the CP adsorbent was washed repeatedly
(0.80 mm) and the residual mercury concentration in the using double-distilled water. The CP adsorbent was dried in an
supernatant was determined using an atomic absorption spectro- oven at 50.0  1.0  C, followed by drying in a vacuum oven at
photometer (AAS, Perkin-Elmer precisely HGA 900). For low 30.0  1.0  C for five hours to further eliminate water vapor. The
mercury concentrations (< 0.2 ppm), the cold vapour-atomic regenerated CP adsorbent was reused in the next cycle of
absorption spectrophotometer (CV-AAS) was used. All analyses adsorption experiment. The adsorption and desorption cycles
were carried out in triplicate, and the mean values were reported. were repeated for five times. It must be noted that the adsorbent
No results were accepted if the standard relative error was greater dosage used was constant at 1 g/L. The adsorption performance of
than 3%. The adsorption capacity at equilibrium, Qe (mmol/g) and each adsorption cycle was evaluated by its regenerated adsorption
removal efficiency, hr (%) of mercury were calculated using Eqs. (1) capacity, Qe,r (mmol/g) and regeneration efficiency, hreg (%). The
and (2), respectively: regeneration efficiency was calculated using Eq. (4),
Q e ¼ ½ðCo  Ce ÞV=m ð1Þ hreg ð%Þ ¼ 100Q e;r =Q e;o ð4Þ

where Qe,o is the adsorption capacity of the virgin adsorbent


hr ð%Þ ¼ ½ðCo  Ce Þ=Co   100 ð2Þ (mmol/g).

where Co and Ce are the solution mercury concentrations 2.2.5. Adsorbent selectivity
respectively at initial and equilibrium (mmol/L), m is the adsorbent The adsorption selectivity of mercury ion onto the CP adsorbent
mass (g) and V is the mercury solution volume (L). was studied by using oilfield produced water (OPW) and natural
The effects of contact time or time dependent adsorption were gas condensate (NGC) which were supplied by a local petroleum
studied by varying the initial mercury concentrations (0.25 mM, company. The concentrations of mercury ion in both produced
0.5 mM and 1.0 mM) and temperatures (30  C, 45  C and 60  C). For water and natural gas condensate were low, therefore, 0.25 mM of
a selected time interval, a small amount of sample was collected for Hg(II) and MeHg(II) were spiked into the OPW and NGC
mercury concentration determination. The adsorption capacity at respectively, in order to increase the mercury concentration to
time t, Qt (mmol/g) was calculated using Eq. (3), detectable levels. The adsorption study was carried out using the
same procedure as described in Section 2.2.3. After the adsorption
Q t ¼ ½ðCo  Ct ÞV=m ð3Þ
process, the mixture was filtered and the filtrate was collected
where Co and Ct are the mercury concentrations respectively at t = 0 for metal concentration determination. The concentration of heavy
(initial) and time t, respectively (mmol/L). metals in produced water was directly measured by using an
inductively coupled plasma-mass spectrometry (ICP-MS, Perki-
2.2.4. Adsorbent reusability studies nElmer model ELAN 6100). The ICP-MS is capable of measuring
The reusability of the CP adsorbent was investigated by carrying multi-elements at once, having detection limit between 0.1 ppt
out the mercury adsorption-desorption experiment. The mercury and 100 ppb.
loaded-CP adsorbent obtained from the first adsorption experi- In the case of NGC, after two days of adsorption process, the
ment was initially washed repeatedly using double-distilled water mixture was filtered and the filtrate was collected for metal
and dried. Thereafter, the mercury-loaded CP was mixed with digestion process according to the method reported by Zettlitzer

Fig. 1. SEM image of CP adsorbent.


2490 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

et al. [22]. The metal digestion procedure was required to recover


heavy metals in the NGC so that the concentration of heavy metals
could be determined using the ICP-MS. For this purpose, 25 mL of
the NGC was mixed with 250 mL of aqua-regia solution and stirred
for 3 h at 160  C. The aqua-regia solution was prepared by mixing
100 mL of hydrochloric acid (HCl, 37%) with 35 mL of nitric acid
(HNO3, 65%) and diluted with double-distilled water to the final
volume of 500 mL in a volumetric flask. After three hours of
digestion, the mixture was cooled to room temperature, and
thereafter, the organic NGC and aqueous solution were separated
using a separating funnel. The aqueous solution containing eluted
metal ions was analyzed using the ICP-MS. The amount of metals
adsorbed onto the CP adsorbent was calculated based on a mass
balance. The experiments were conducted in duplicate and the ICP-
MS measurements were done in triplicate. For both cases, the
mean data were reported.
The adsorption selectivity of metal ions in OPW and NGC was
evaluated by calculating the adsorption capacity, Qe (mmol/g) and
distribution coefficient, Kd (L/g) of the respective metal ions. The Kd
can be determined by using Eq. (4).
Kd ¼ ðCo  Ce ÞV=mCe ð5Þ

3. Results and discussions

3.1. Adsorbent characterizations

The moisture and ash contents of CP-Pure were 12.04 (wt.%) and
3.16 (wt.%), respectively. The ash content suggests that the coir pith
contains low extracts with little or no waxes and resin, when
compared with other non-woody lignocellulosic materials [23].
The value of pH at point zero was 6.2  1.48, which was slightly
lower than other reported data [24]; and the cationic exchange
capacity (CEC) was 0.44  0.03 meq/g, which was higher than
previously measured [24]. Fig. 2. FTIR spectra of CP and mercury loaded CP adsorbents.
The surface morphology of the CP adsorbent as given by the
SEM image was not smooth and some nodes were observed (Fig. 1).
The particles had irregular shapes and surfaces. The surface area,
total pore volume, and average pore diameter were 3.275 m2/g,
1.315  102 cm3/g, and 3.60  103 cm3/g, respectively. These
values were much lower compared to other CP adsorbent data
as reported by previous researchers [24].
The FTIR analysis was conducted to identify the functional
groups responsible for mercury adsorption. The vibrations were
contributed by the structures of cellulose, hemicellulose and
lignin. The FTIR spectrum of CP adsorbent is given in Fig. 2. A broad
and strong peak at 3399 cm1 is attributed to the O H stretching.
The CH stretching and bending vibrations from CH3 were
observed at wavelength of 2918 cm1 and 1447 cm1, respectively.
The C¼O stretching peak of ketone, aldehyde and carboxylic from
lignin and hemicellulose were observed at 1727 cm1. The peak at
1440 cm1 is attributed to the C O stretching and C OCH3
deformation, while the peak at 1375 cm1 is attributed to the C O
and C C stretching vibrations.
Fig. 3. Effect of pH on adsorption capacity of Hg(II) and MeHg(II) adsorption onto CP
adsorbent.
3.2. Equilibrium mercury adsorption
presented in Fig. 3. Generally, a high adsorption capacity was
3.2.1. Effect of pH observed for Hg(II) as compared to the MeHg(II). The Hg(II)
The pH of metal aqueous solution has been identified as the adsorption capacity was found to be constant, around 0.25 mmol/
most important variable governing the metal ions adsorption on g over the pH range studied. The adsorption by chelation and
the adsorbents. The pH of the solutions will change the metal complexes seemed to be dominant for the Hg(II) adsorption
physicochemical state of the mercury species in the solution, because no significant difference towards pH was observed. This
ionization state of the functional groups, as well as the total result was in contrast to that of the MeHg(II), in which the initial
charge present on the adsorbent surfaces. The effect of pH on pH greatly affected the MeHg(II) adsorption capacity. The Qe of
mercury ions adsorption by the CP adsorbent is graphically MeHg(II) was found to be the lowest at pHi = 2, but increased as
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2491

the pH increased to a maximum (0.16 mmol/g) in between pH However, the hr of the Hg(II) removal reached almost 100%
6 and 7, and thereafter the adsorption decreased. regardless of adsorbent dosages, which was clearly higher as
At pH < pHpzc, the surface of the adsorbent is positively charged compared to the MeHg(II). The hr of MeHg(II) increased
due to protonation of the acidic functional groups. At lower pH, significantly from (26 to 75) % when using (0.4 to 3.0) g/L of
mercury species with positive charge (i.e., Hg2+ and CH3Hg+) are adsorbent dosage, respectively.
dominant, attributed to electrostatic repulsion with the positively The increase of adsorbent dosage reduced the amount of
charged surface adsorbent, thus lower adsorption capacity was mercury adsorbed onto a unit mass of adsorbent and the effect was
observed. In addition, the presence of high concentration of H+ at more pronounced at higher adsorbent dosage. This was because a
lower pH also provides a strong competition with mercury ions for fixed mass of adsorbent can only adsorb a certain amount of
the available adsorption sites. With increasing pH, deprotonated adsorbate, which means that with more adsorbent dosage, the
sites on the adsorbent surfaces increased, resulting in stronger larger the quantity of adsorbate can be adsorbed by the adsorbent.
attraction for positively charged metal ions. However, for Hg(II) In addition, because of the high dosage, the particle interaction
adsorption, the physicochemical state of mercury species in the may desorb some of the metal ions, which are only loosely and
solution and ionization state of the functional groups at various pH reversely bound to the adsorbent surfaces [9].
values did not affect the adsorption performance. This result might
be due to a strong affinity of the Hg(II) ion towards active sites 3.2.3. Effect of mercury concentration
compared to MeHg(II) ion. The effect of mercury ion concentrations was studied by using
As can be seen in the small figure inside Fig. 3, the equilibrium the adsorbent dosage of 1.0 g/L. It was observed that more than 90%
pH tended to change from the original pH. The pHe of Hg(II) of Hg(II) was adsorbed for the entire concentration ranges (Fig. 5).
adsorption became less acidic or more basic, which might be due to However, the hr of the MeHg(II) reduced significantly, as its initial
the adsorption of H+ alongside with the Hg2+. On the other hand, concentration increased. At lower concentration, the ratio of
for MeHg(II) adsorption, the pHe of the MeHg(II) did not show any mercury molecules available to adsorbent active sites is low,
significant changes as compared to that of the initial pH (pHi). The therefore there is a higher probability for the mercury ions to be
optimum adsorption pH was obtained at the pHi of 6, thus it was adsorbed. At higher concentration, the available active sites for
chosen for the subsequent experiments of both mercury species. adsorption become fewer hence reducing the removal perfor-
mance, since substantial mercury molecules remain in the
3.2.2. Effect of adsorbent dosage solution. It was found that the Qe of MeHg(II) was lower than
The adsorption of Hg(II) and MeHg(II) at different dosages using Hg(II) which contradicted the report that the adsorption capacity

their concentration of 0.25 mM was studied. The hr of mercury ions of metal ion was higher in the presence of Cl thanNO 3 [25]. Thus,
increased with the adsorbent dosage from (0.4 to 3.0) g/L (Fig. 4). the lower Qe of MeHg(II) might be due to the inability of the bulkier
group of the MeHg(II) to diffuse inside the adsorbent, thereby
reducing the adsorption capacity [15].

3.2.4. Adsorption isotherm


An adsorption isotherm is critical in optimizing the usage of
adsorbent. The adsorption isotherms of Hg(II) and MeHg(II) were
obtained by using 0.2 g/L and 1.0 g/L of adsorbent dosage,
respectively. The adsorption isotherms of Hg(II) and MeHg(II)
are shown in Fig. 6. The adsorption isotherm of the Hg(II) using
1.0 g/L of adsorbent dosage did not reach the plateau Qe, which
means that, the CP adsorbent was still capable to adsorb Hg(II), but
at a higher Hg(II) concentration, it could not be carried out because
it resulted in the precipitation of mercury ion. The Qe at this
condition was only 1.6 mmol/g. Therefore, in order to achieve the
saturation or the maximum adsorption capacity (Qe max), the

Fig. 4. Effect of adsorbent dosage of (a) Hg(II) and (b) MeHg(II) adsorption onto CP Fig. 5. Effect of initial concentration on adsorption capacity of Hg(II) and MeHg(II)
adsorbent (solid and dashed lines represent Qe and h, respectively). adsorption onto CP adsorbent.
2492 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

the adsorption process was favorable as the dimensionless


constant separation factor, RL (RL = 1/[1 + bCo]) over the initial
concentration range used was between 0 and 1. This result
indicates that the adsorption is favorable. The values of n from the
Freundlich isotherm model also supported the favorability of the
Hg(II) and MeHg(II) adsorption onto the CP adsorbent. The kF
values for Hg(II) indicated that the CP adsorbent had a higher
adsorption capacity compared to MeHg(II). The mean free energy
calculated from the D–R model (EDR) of Hg(II) and MeHg(II) was
17.033 kJ/mol and 11.396 kJ/mol, respectively. These results indi-
cated that the Hg(II) and MeHg(II) adsorption fell within the
chemisorption/ion exchange mechanisms [26]. The binding energy
calculated from the Temkin isotherm model also showed that the
absorption of Hg(II) and MeHg(II) was of chemisorption or ion
exchange mechanism. This result was only acceptable at the
temperature of 30  C. If the adsorption occurs at other temper-
atures, the mechanism might be different due to the different
Fig. 6. Adsorption isotherm of Hg(II) and MeHg(II) onto CP adsorbent. nature of the thermodynamic properties of the adsorbent-
adsorbate interactions.
adsorbent dosage for Hg(II) adsorption was reduced to 0.2 g/L and In order to identify the most fitted isotherm models to the
the Qe max was found to be 2.60 mmol/g. The Qe max of the MeHg(II) experimental data, the model having a high coefficient of
was 0.50 mmol/g. Table 1 shows the comparison of Qe max of CP determination (r2) value was normally chosen as the best fit to
adsorbent with other materials towards Hg(II) and MeHg(II) describe the adsorption isotherm data. However, Ho [27] reported
adsorption. It was found that the CP adsorbent still give a higher that the use of non-linear method would be better for comparison
adsorption capacity towards both mercury ions despite its low to obtain the best isotherm model. The isotherm fitting was plotted
surface area and simple preparation procedure. based on the non-linear equations, using the model constant
The adsorption isotherms of both mercury ions were further parameters obtained from the linear equation plot analysis [14]. In
analyzed using the Langmuir, Freundlich, Dubinin–Radushkevich addition, another three error functions have been used together to
(D–R) and Temkin isotherm models. The isotherm equations and determine the most fitted model namely, average relative error
its linearization forms are shown in Table 2. The constant (ARE), Marquardt’s percent standard deviation (MPSD) and
parameters of the isotherm models were calculated using a linear normalized standard deviation (DQ) [28]. The r2 of the Hg(II)
plot of the isotherm equations. Table 3 shows the adsorption adsorption was the highest (r2 > 0.99) for the Langmuir isotherm
isotherm parameters of the Langmuir, Freundlich, D–R and Temkin model and a good fitting was observed for the whole concentration
models for the adsorption of Hg(II) and MeHg(II) onto the CP studied (Fig. 6). Other isotherm models show lower values
adsorbent. (r2 < 0.80) and other error functions (ARE, MPRD, and DQ) were
The Langmuir isotherm model assumes that the adsorption bigger than the Langmuir error functions. In the case of MeHg(II)
occurs on a homogeneous surface by monolayer adsorption adsorption, the r2 of the Freundlich and D-R models exceeding
without interaction between adsorbed ions, uniform energies of 0.99 and other error functions (e.g. ARE, MPSD, and DQ) were the
adsorption onto the surface, and no transmigration of adsorbate in lowest compared to that of the Langmuir and Temkin models.
the surface plane. As can be seen in Table 3, the monolayer However, from the fittings (Fig. 6), the Freundlich matched the
adsorption capacity, Qm L was found to be close to the experimental experimental data better than those of the Langmuir, D-R, and
Qe max. The parameter b is related to the affinity of active sites with Temkin. Therefore, the isotherm of Hg(II) and MeHg(II) were
the adsorbate and it can be seen that the Hg(II) adsorption concluded to follow the Langmuir and Freundlich isotherm
presented a stronger affinity than MeHg(II). It was also found that models, respectively.

Table 1
A comparison of mercury ion adsorption onto low-cost adsorbents as reported in the literature.

Adsorbents Surface area (m2/g) Hg species Qe.max References


(mmol/g)
CP 3.275 Hg(II) 2.60 This study
3.275 MeHg(II) 0.50
CP-Char 464.28 Hg(II) 0.23
MeHg(II) 0.88
CP-AC 505.00 Hg(II) 0.71
MeHg(II) 0.42
Garlic (Allium sativum L.) powder 0.126 Hg(II) 0.003 Eom et al. [17]
RPSR  Hg(II) 0.288 Saman et al. [14]
 MeHg(II) 0.21 3
Coconut milk waste 9.67 Hg(II) 2.23 Johari et al. [10]
Rice straw 8.46 Hg(II) 0.44
Rice straw-MPTES 11.46 Hg(II) 0.78
Lemna minor  Hg(II) 0.14  103
 MeHg(II) 0.13  103
Carica papaya wood 0.955 Hg(II) 0.35 Basha et al. [16]
Palm shell powder 2.23 Hg(II) 0.006
CP-poy(hydroxyethylmethacrylate) 93.5 Hg(II) 0.80 Anirudhan et al. [9]
CP-polyacrylamide 42.3 Hg(II) 1.27 Anirudhan et al. [24]
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2493

Table 2
Non-linear and linear adsorption isotherm and chemical reaction kinetic model equations.

Models Non-linear equation Linear equation


Isotherm models
Langmuir Qe = Qm LbCe/(1 + bCe) Ce/Qe = 1/Qm L b +Ce/Qm L
Freundlich Qe = kfCe1/n ln Qe = ln kf + 1/n ln Ce
Dubinin-Radushkevich (D-R) Qe = Qm exp(bDRe2) ln Qe = ln Qm  2bDRe2
where e = RT ln (1 1/Ce)
Temkin Qe = RT/bT [ln (ATCe)] Qe = (RT/bT) ln AT + (RT/bT) ln Ce

Chemical reaction kinetic models


Pseudo-first order (PFO) Qt = Qe [1  exp(k1t)] ln (Qe  Qt) = ln Qe  k1t
Pseudo-second order (PSO) Qt = t/(1/k2Qe2 + t/Qe) t/Qt = 1/(k2Qe2) + (1/Qe)t
Elovich – Qt = (1/b) ln(ab) + (1/b) ln t

3.2.5. Effect of temperature sign of DG for both Hg(II) and MeHg(II) adsorption indicates that
The effect of temperature on the adsorption of Hg(II) and MeHg the adsorption was spontaneous. The magnitude of DG increased
(II) was studied by varying the adsorption temperatures at with temperature, implying an increased degree of spontaneity at
(30, 45 and 60)  C. It was observed that the adsorption of Hg(II) high temperature. The positive value of DH indicates that the
and MeHg(II) increased with temperature, thereby suggesting the adsorption process was endothermic in nature [29]. The DH value
endothermic nature of the adsorption process, where the net heat could be used to measure the interaction force between adsorbent
is adsorbed during the adsorption process. and adsorbate, giving an indication of the bonding strength of the
The thermodynamic parameters namely the standard Gibbs interactions. The DH value of the Hg(II) and MeHg(II) adsorption
free energy (DG ), standard enthalpy change (DH ) and standard was found to be in the range of chemisorption mechanism. The
entropy change (DS ) are calculated using Eqs. (6)–(8), respective- positive values of DS suggest that the randomness of the solid-
ly. solution interface has increased [29], which indicate a good affinity
of CP adsorbent towards mercury ions.
DGo ¼ RTlnKb ð6Þ
3.3. Mercury adsorption kinetics

RTlnKb ¼ TDSo  DHo ð7Þ The adsorption behavior of Hg(II) and MeHg(II) onto the CP
adsorbent against time was studied by varying the initial mercury
concentrations (0.25 mM, 0.5 mM and 1.0 mM) and temperatures
D So DHo (30  C, 45  C and 60  C). The adsorption kinetics of Hg(II) and MeHg
lnKb ¼  ð8Þ
R RT (II) at various temperatures and initial mercury concentrations are
where R is the universal gas constant (8.314 J/mol K) and T is the presented in Fig. 7. The adsorption rate of the mercury ions
adsorption temperature (K). Parameter Kb is the equilibrium adsorption was initially rapid, then gradually approaching
constant, which relates the ratio of adsorbate concentration on the equilibrium and finally achieved equilibrium at contact time of
solid adsorbent (mmol/g) to the adsorbate concentration in the 360 min. The rapid adsorption rate at the beginning of adsorption
solution (mmol/L) at equilibrium, Kb = Qe/Ce. process could be explained by the high number of active-binding
The thermodynamic parameters of the Hg(II) and MeHg(II) ions sites available on the CP adsorbent surfaces and faster interactions,
adsorption onto the CP adsorbent are given in Table 4. The negative thus creating a bigger mercury concentration gradient between the
solution and adsorbent surfaces. The adsorption rate of Hg(II) was
more rapid than that of MeHg(II) for all conditions studied.
Table 3 Adsorption of metal ions onto the adsorbent surfaces is
Adsorption isotherm parameters of mercury adsorption onto CP adsorbent. basically governed by four consecutive steps: (a) transport of
Parameters Hg(II) MeHg(II)
metal ions from the bulk solution to adsorbent surfaces (external
mass transfer); (b) transfer metal ions from the film to the external
Qe max (mmol/g) 2.60 0.50
surface of the adsorbent (film diffusion); (c) transfer of metal ions
Langmuir from external surface to the internal structure of the adsorbent
Qm L (mmol/g) 2.60 0.53 (internal pore diffusion or intra-particle diffusion); and (d)
b (l/mmol) 40.513 5.362 adsorption of metal ions on active sites via physical or chemical
r2 0.994 0.956
interactions. The internal pore diffusion involves pore diffusion
Freundlich (diffusion within the pore volume) and surface diffusion (diffusion
kf (lnm1n/g) 0.011 0.009 along the surface of pores) [30]. The first three steps are
n 5.184 2.329 categorized as physical steps (or diffusion process), while the
r2 0.707 0.991

Dubinin-Radushkevich (D-R) Table 4


Qm DR (mmol/g) 4.66 1.46 Thermodynamic properties of mercury adsorption onto CP adsorbent.
bDR  109 (J/mol) 1.723 3.850
EDR (kJ/mol) 17.033 11.396 Parameters Temperature ( C) Hg(II) MeHg(II)
r2 0.729 0.993 DH (kJ/mol)  110.713
DS (J/mol K) 374.604
Temkin DG (kJ/mol) 30  4.832  63.196
bT  103 (kJ/mol) 7.428 26.023 45  8.467  66.326
AT (l/mmol) 2821.957 87.836 60  14.086  69.456
r2 0.767 0.948
2494 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

where V is the volume of the solution (mL), m is the mass of


adsorbent (g), S is the external surface area per mass of adsorbent
(m2/g), bL is the external mass transfer coefficient in liquid phase
(cm/s), CAO and CA are the initial concentration and concentration
of adsorbate (mmol/mL), respectively. Several models have been
developed to estimate the value of bL [31,32]. In this study, the
value of bL was estimated according to the method suggested by
Ocampo-Pérez et al. [31]. The term on the right of Eq. (10) is the
slope of the concentration-decaying curve. The first two data
points at t = 0 and t = 5 min, were used to estimate the slope of the
concentration decaying curve, in which the adsorption can be
assumed to be exclusively controlled by the external mass transfer.
It was found that the external mass transfer coefficients of Hg
(II) were higher than MeHg(II) (Table 5). The MeHg(II) has a larger
molecular size than Hg(II), it thus diffused slower than Hg(II). At
the same concentration, the bL values increase with temperature
for both mercury ions. At higher temperatures, mercury ions
diffused faster because more energy is available for the diffusion
process thus increasing the bL value. The relationships of bL, the
diffusivity of the adsorbate molecule (DM) and temperature are
given in Eqs. (11) and (12):
   
DM ¼ 7:4  108 ðwMB Þ0:5 T = mf V0:6 A ð11Þ

bL ¼ DM =d ð12Þ

where d is the thickness of the film layer (cm), f is the associate


parameter of water = 2.60, MB is the molecular weight of water
(18 g/mol), mf is the viscosity of fluid (water) = 0.89 cp, and VA is the
liquid molar volume of adsorbate at normal boiling temperature.

 Film diffusion

The film diffusion is a transport of adsorbate across the liquid


film surrounding the adsorbent particles. According to Eq. (12), the
mass transfer rate is inversely proportional to the film thickness.
The thicker the liquid film, the slower the transfer rate because the
film becomes a diffusion barrier.
Fig. 7. Adsorption kinetics of Hg(II) and MeHg(II) onto CP adsorbent at different (a) Many models have been developed to determine the film
temperatures and (b) initial concentrations. Value in the bracket (p;q) in the legend
diffusion coefficient (Df) [33]. One of such models is given by
represents (p) temperature ( C) and (q) initial mercury concentration (mM).
Eq. (13) which was developed based on Fick’s law [34]. The film
diffusion coefficient (Df) is calculated from the slope of the Qt/Qe
against t0.5 at the early stage of kinetic data by assuming the
fourth step is a chemical bonding interaction. Several authors have adsorbent was spherical in shape with radius Rp (cm) [34].
proposed different models to describe these adsorption kinetic  0:5
steps using either physical or chemical based models in order ðQ t =Q e Þ ¼ 6 Df =pR2P t0:5 ð13Þ
obtain an insight of the mechanism of metal ion adsorption on the
adsorbent surfaces. In the present study, the adsorption data were
analyzed consecutively following the physical and chemical steps
involved in the adsorption process. Table 5
Diffusion model parameters of Hg(II) and MeHg(I) adsorption onto CP adsorbent.

(a) External mass transfer [C]o T bL  105 Df  1012 Deff  1010


(mM) ( C) (cm/s) (cm2/min) (cm2/min)
When adsorbent particles are in contact with a solution Hg(II) adsorption
containing mercury ions, the mercury molecules first migrate 0.25 30 4.349 2.914 7.518
from the bulk solution to the surface of liquid film. The 45 4.792 4.265 14.118
60 5.620 4.765 20.155
concentration of adsorbate decreases from the value of CA at bulk
0.50 30 5.415 5.082 15.558
to Ce,A at the particle surface due to the external film resistance. 1.00 30 5.595 5.969 19.305
The rate of external mass transport (Nt) is presented by Eq. (9).
 MeHg(II) adsorption
Nt ¼ VdCt =dt ¼ mbL S CA  Ce;A ð9Þ 0.25 30 1.804 1.061 2.941
45 2.716 1.516 3.251
When t ! 0 then Ce,A ! 0 and CA ! CA0, thus Eq. (9) can be 60 3.143 2.105 4.919
simplified to: 0.50 30 1.004 2.497 4.093
1.00 30 0.742 2.719 5.004
½dðCA =CAO =dt ¼ mSbL =V ð10Þ
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2495

The value of Df estimated using Eq. (13) is given in Table 5. The second order (PSO), and Elovich kinetic models were used to
Df increased with temperature and initial concentration. The Df of analyze the Hg(II) and MeHg(II) adsorption onto CP adsorbent. The
Hg(II) adsorption values range from 2.914  1012 cm2/min to mathematical expressions of those reaction kinetic models are
5.969  1012 cm2/min, while lower Df was observed for MeHg(II) shown in Table 2.
with the value range from 1.061 1012 cm2/min to 2.719  1012 The calculated parameters of the reaction kinetic modeling
cm2/min. analysis of Hg(II) and MeHg(II) adsorption onto the CP adsorbent
are tabulated in Table 6. The adsorption capacity calculated from
 Internal diffusion the PSO kinetic model (Qe2) is very close to the experimental Qe (Qe.
exp). The increase of mercury initial concentrations resulted in
The internal diffusion is a transport of adsorbate from the higher Qe2 and initial adsorption rate (ho2). A similar trend was
external surface of adsorbent (at film boundary) to the internal obtained for the initial adsorption rate calculated from the PFO
surfaces of the adsorbent particles. The transport process has been (ho1) and Elovich (a) kinetic models. However, the rate of
assumed to occur by pore diffusion and surface/solid phase adsorption analyzed from the PSO (k2) was found to decrease
diffusion [33]. The pore diffusion is the transport of adsorbent with increasing temperature from 30  C to 60  C.
within the pores of the adsorbents. Surface diffusion is the By comparing the r2, ARE and MPSD values of all adsorption
transport of molecules adsorbed on the interior surface of solid kinetic models, it was found that the PSO kinetic model best fitted
materials from one site to another direction of decreasing to the experimental data. These results indicated that the
concentration [30]. By this definition, the ‘lumped’ effective
diffusion coefficient, Deff is adopted incorporating the effect of
pore diffusion and surface diffusion. The mass transport at the Table 6
Reaction based kinetic model parameters of Hg(II) and MeHg(II) adsorption onto CP
internal surface is given by Eq. (14). adsorbent.
Nt ¼ Dp @C=@r þ rs Ds @Q=@r ð14Þ Co (mM) 0.25 0.50 1.00

Thus, the Deff is defined as (Ho et al. [30]): T ( C) 30 45 60 30 30


Hg(II) adsorption
Deff ¼ Dp þ rs Ds @Q=@C ð15Þ
Qe exp (mmol/g) 0.217 0.232 0.245 0.495 1.012
where Deff is the effective pore diffusion coefficient, Dp is the pore
PFO
diffusivity, and Ds is the surface diffusivity. Qe1 (mmol/g) 0.118 0.099 0.094 0.193 0.316
Several models have been developed to determine the Deff. One k1 (1/min) 0.028 0.032 0.038 0.034 0.039
of such models is the Boyd plot, which can be obtained by plotting ho1 103 (mmol/g min) 3.172 3.185 3.578 6.632 1.213
Bt against time, where the Bt can be expressed by the following r2 0.8636 0.8677 0.8729 0.9055 0.8228
ARE 61.6154 68.8921 70.8609 71.5395 75.9665
equation [35].
MPSD 71.3835 80.7258 81.2244 83.5692 86.6670
Bt ¼ 0:4977  lnð1  FÞ; Fvalue > 0:85 ð16Þ
PSO
Qe2 (mmol/g) 0.220 0.232 0.250 0.497 1.026
k2 (mmol/min) 0.790 0.936 1.006 0.518 0.320
p p ho2  102 (mmol/g min)
Bt ¼ ð p  ½p  ðp2 F=3ÞÞ2 ; Fvalue < 0:85 ð17Þ 3.811 5.038 6.289 12.795 33.667
r2 0.9979 0.9987 0.9983 0.9997 0.9989
F is defined as the ratio of amount adsorbed at time t to amount at ARE 2.6008 2.2246 2.2494 1.2532 2.7455
equilibrium (F = Qt/Qe). The slope (s) of the Boyd plot is used to MPSD 4.0356 3.9387 4.3360 2.3882 5.6795

calculate the Deff value according to the following relationship Elovich


(Eq. (18)). b (g/mmol) 25.128 25.518 25.517 12.859 6.487
2
a (mmol/g min) 0.122 0.201 0.331 0.674 2.054
2
s ¼ p Deff =Rp ð18Þ r2 0.9545 0.9181 0.8859 0.9146 0.8175
ARE 11.3625 12.0871 12.1332 10.8130 14.2657
It should be noted that the value of Bt becomes less reliable as MPSD 16.2829 17.3956 16.8036 15.8741 19.0528
the equilibrium is approached [35], therefore, the plot is only
examined at the initial period. The Deff value of Hg(II) was much MeHg(II) adsorption
Qe exp (mmol/g) 0.126 0.169 0.202 0.168 0.202
higher than MeHg(II). This is due to fact that Hg(II) ion has a
PFO
smaller radius than MeHg(II), and the smaller ion will diffuse faster Qe1 (mmol/g) 0.089 0.115 0.124 0.119 0.134
than the larger one. The Deff values of Hg(II) and MeHg(II) were k1 (1/min) 0.018 0.024 0.027 0.021 0.025
found in the range of 13.258  1010 cm2/min to 19.305  1010 ho1 103 (mmol/g min) 1.587 2.763 3.339 2.500 3.352
cm2/min and 2.941 1010 cm2/min to 5.004  1010 cm2/min, r2 0.9563 0.9940 0.9578 0.9271 0.9465
ARE 51.1318 50.1476 55.3214 49.7557 51.1839
respectively.
MPSD 64.3212 62.8488 7.9632 62.3630 63.3802

 Surface interaction step PSO


Qe2 (mmol/g) 0.109 0.161 0.198 0.158 0.200
k2 (mmol/min) 0.899 0.645 0.639 0.584 0.474
The surface interaction step involves the adsorption of
ho2 (mmol/g min) 1.073 1.674 2.506 1.454 1.893
adsorbate by active sites of adsorbents, in which the adsorption r2 0.9967 0.9948 0.9988 0.9996 0.9998
rate can be represented in the same manner as the rate of chemical ARE 13.3716 9.4859 7.9632 9.8493 8.0146
reaction. Several chemical reaction models have been reported in MPSD 18.2221 12.5184 10.0961 12.7513 9.8286
the literature to describe the chemical reaction step. The modeling
Elovich
normally assumed the overall rate of adsorption is exclusively
b (g/mmol) 44.356 33.369 25.269 28.828 23.320
controlled by the adsorption rate of the adsorbate on the surface of a (mmol/g min) 0.025 0.042 0.061 0.040 0.046
adsorbent assuming that both external and internal mass transport r2
0.9854 0.9971 0.9754 0.9900 0.9903
are neglected. In this study, three most common chemical reaction ARE 5.4299 4.2087 8.2554 9.1467 7.5580
MPSD 7.7115 5.9463 10.7545 14.5184 11.520
kinetic models namely the pseudo-first order (PFO), pseudo-
2496 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

adsorption of Hg(II) and MeHg(II) followed the PSO kinetic model, it is controlled by the chemical process [30]. The calculated values
suggesting that the adsorption is controlled by the chemisorption of Ea of Hg(II) and MeHg(II) adsorption were 16.370 kJ/mol and
process. It might involve valency forces through sharing or 23.640 kJ/mol, respectively, which indicate the diffusion process
exchange of electrons between the active sites of the CP adsorbent was the rate-limiting step.
and mercury ions.
3.4. Mechanism of mercury adsorption
 Rate-limiting process
Throughout the kinetic adsorption studies, two mechanism
During the adsorption process, the kinetics of adsorption processes have been proposed to interpret the adsorption
process may be controlled by either diffusion or the surface behaviors of Hg(II) and MeHg(II) namely physical and chemical
reaction process. The features of these two processes are often processes. The physical process mechanism involves diffusion
difficult. According to the chemical reaction kinetic based model, steps namely external mass transfer, film diffusion, pore diffusion,
the kinetic data analysis of the Hg(II) and MeHg(II) adsorption onto and surface diffusion [38]. The chemical process mechanism refers
the CP adsorbents followed the PSO kinetic model, suggesting that to the chemical interaction between the adsorbate and active sites
the chemical process is a rate-limiting step, without considering of the adsorbents. The interaction can be classified as either
the physical (or diffusion) process. physisorption or chemisorption interaction depending on the
For a diffusion step, there are three main steps of diffusion nature of the bonding formation. The physisorption has a weak
process, which are external mass transfer, film diffusion and chemical interaction that could be due to the intermolecular forces
intraparticle diffusion (including pore and surface diffusion). Since of attractions between adsorbent active site and the adsorbate
the rapid stirring was conducted during the adsorption process, such as the van der Waals forces, London dispersion forces, H-
bulk diffusion effect was assumed negligible and film diffusion bonding, and classical electrostatic forces. The chemisorption
becomes a dominant role during the external mass transfer interactions such as ion exchange, chelation and metal complex are
process. Therefore, in order to determine the rate-limiting step of however identified as strong interactions.
these two diffusion processes (i.e. film diffusion and intraparticle The CP adsorbent had a low surface area and low porosity,
diffusion), the kinetic data were quantitatively analyzed by using therefore, the mechanism of adsorption tend to be that of surface
the Weber-Morris plot (Qt versus t0.5). If the plot gives a straight covering rather than pore filling. The FTIR result showed that the
line passing through the origin, the intraparticle particle diffusion CP adsorbent surfaces most possibly contained oxygen active
is the sole-limiting step; otherwise, other diffusion processes can groups from phenol, ether, aldehyde, carboxylic, ketone, and
be the limiting step [36]. The Weber-Morris plot did not pass hydroxyl compounds. These active groups play an important role
through origin; thereby the intraparticle diffusion was not the rate- that promotes chemical interactions with mercury ions during the
limiting step. There were three-line regions observed for both Hg adsorption process. The monolayer adsorption capacities of the
(II) and MeHg(II) adsorption, which indicated the presence of more Langmuir isotherm model of Hg(II) and MeHg(II) were 2.60  103
than one step diffusion process. An additional confirmation to mol/g and 5.30  104 mol/g, respectively, which were higher than
identify the rate-limiting step between film and internal particle the CEC value (0.44 meq/g or 4.4  104 mol/g). This result showed
diffusion could be obtained from the Boyd plot [37]. The Boyd plot that the ionic exchange process was not a dominant mechanism of
of Bt versus time in the initial period of Hg(II) and MeHg(II) the adsorption. As previously discussed in Section 3.2.1, it can be
adsorption onto the CP adsorbent showed a straight line that concluded that the adsorption of Hg(II) was mainly through metal-
crosses at the Y-axis. These results indicated that the film diffusion complexion, because there was no significant changes in Qe at
was the rate-limiting step. If a straight line of Boyd plot passes various initial pH values was observed. On the contrary, 41% of the
through the origin, it indicated that the adsorption process was total MeHg (II) adsorbed was found to be pH dependent, which was
governed by intraparticle diffusion [37]. In order to confirm that assumed to be due to the ion-exchange process. The remaining 59%
the diffusion coefficient of the two steps was calculated and was pH independent, which might be due to the chelation or
compared. The value of Df was smaller than Deff, thus diffusion metal-complexes process. The mean free energy from the D-R
through the film is slower than the intraparticle diffusion, isotherm model (ED-R) and binding energy from the Temkin (bT)
confirming the film diffusion as the rate-limiting step. This isotherm model indicate that strong chemical interactions were
conclusion was made when considering only the diffusion step present during the adsorption of Hg(II) and MeHg(II). Adsorption at
processes. various temperatures found that high temperature favors an
It is noted that in the above discussion that the rate limiting- endothermic reaction. It can be deduced that the chemical reaction
step was chosen according to the classification of the process occurs due to the formation of stronger bonds at high temper-

mechanism. Ho et al. [30] stated that one of the general guidelines atures. The DH values of both mercury adsorptions fall in the
to estimate the controlled process is according to the time region of strong chemisorption interactions. Therefore, it can be
equilibrium is achieved. If the equilibrium is achieved within three concluded that the chelation and metal-complexion might be a
hours, the process is reaction kinetic controlled and above twenty- dominant process for the Hg(II) and MeHg(II) adsorption. In order
four hours, it is a diffusion-controlled process. If the equilibrium to obtain more insight into the chemistry and reaction state of
achieved is in between three to twenty four hour period, either or adsorbate and its interaction with binding sites, the FTIR and XPS
both reaction and diffusion process may be the rate-limiting step analyses of the fresh and mercury-loaded CP were carried out.
[30]. In this study, it was found that the adsorption of Hg(II) and The intensity of some FTIR peaks was reduced and other peaks
MeHg(II) achieved equilibrium at the maximum contact time of appeared apparently because of the mercury ions adsorption
6 h. Therefore, either or both reaction and diffusion processes can (Fig. 2). The OH bending vibrations at 1609 cm1 shifted to a
be the rate-limiting process. A more appropriate quantitative higher wavenumber for the CP-Hg(II) and CP-MeHg(II). The peak
approach to determine the rate-limiting step is by their activation intensity at 1382 cm1 from the C H bending vibration strongly
energy (Ea) value. The Ea value was determined from the plot of kid sharpened and shifted to 1395 and 1372 cm1 after Hg(II) and
against the reciprocal temperature (Arrhenius plot). The kid is the MeHg(II) adsorption, respectively. The peaks at 1507 and
intraparticle diffusion rate by the Weber-Morris plot. If the Ea value 1447 cm1 of the C H bending vibration appeared strongly after
is less than 25–30 kJ/mol, the adsorption is controlled by a the MeHg(II) adsorption. However, these peaks almost diminished
diffusion process; otherwise, for energy greater than 25–30 kJ/mol, after the Hg(II) adsorption. Fig. 8 shows a wide scan of the CP
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2497

adsorbent before and after the Hg(II) and MeHg(II) adsorptions.


The mercury element peaks detected in XPS spectra were Hg4f and
Hg4d. The Hg4fwas observed at the region between 100 eV to
106 eV, while Hg4d was observed at binding energy between
355 eV to 380 eV. The high resolution of the Hg4f peak is shown in
Fig. 9(a). The slight but significant changes of O1s and C1s peaks
were observed for the CP adsorbent before and after adsorption.
The optimum binding energy of O1s (Fig. 9(b)) decreased from
532.6 eV to 351.8 eV and 532.1 eV after Hg(II) and MeHg(II)
adsorption, respectively. The C1s spectrum of CP adsorbent
consisted of three peaks (Fig. 9(c)). A strong peak at 285 eV might
be attributed to sp3 of CC bond and other two peaks at 286.5 eV
and 288.8, indicating the presence of CO and O C¼O,
respectively [39]. The peak at 286.5 eV might come from phenol,
hydroxyl or ether groups, and the peak at 288.8 eV might be due to
the presence of carboxylic, ketone and ester groups in the lignin,
hemicellulose and cellulose compounds. After adsorption, the
binding energy of the C O and OC¼O shifted to 285.7 eV and
288.3 eV, respectively. The mercury ions might form complexes
with the oxygen groups such as phenol-Hg, hydroxyl-Hg,

Fig. 9. XPS spectrum of CP and mercury loaded-CP adsorbents: (a) Hg4f; (b) O1s; and
(c) C1s.

carboxylate-Hg or ether-Hg, thus changing the optimum binding


energy of O1s and C1s.

3.5. Adsorbent regeneration and selectivity studies

The adsorbent regeneration is important since it will determine


the adsorption process viability. Thus, it was investigated. Since the
adsorption mechanism of Hg(II) and MeHg(II) involved chemical
reaction process, therefore, the regeneration of mercury-loaded CP
was performed using chemical solutions. In this study, two types of
desorption agents were used namely 0.1 M hydrochloric acid (HCl)
and potassium iodide (KI) solutions. The adsorption desorption
processes were carried out for five adsorption cycles. Fig. 10 shows
the adsorption performance of Hg(II) and MeHg(II) up to
Fig. 8. XPS wide scan of CP and mercury loaded-CP adsorbents.
2498 N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499

Table 7
Initial concentration (Co), adsorption capacities (Qe) and distribution coefficient
(Kd) of mercury ions and other metals adsorption in OPW and NGC by using CP
adsorbent.

Cations Co  103 (mM) Qe  103 (mmol/ hr (%) Kd (L/g)


g)
Oilfield produced water (OPW)
Hg(II) 276.41  5.03 234.02  0.47 85.26  0.05 5.74  0.03
Zn 0.44  0.09 0.42  0.0005 95.04  0.01 19.03  0.02
V 3.34  0.16 1.63  0.04 49.09  1.01 0.96  0.04
Cu 2.49  0.29 0.91  0.06 40.00  2.54 0.66  0.07
As 1.94  0.18 1.10  0.04 57.07  2.13 1.32  0.12
Ba 1.99  0.33 0.66  0.05 33.16  2.47 0.49  0.06
Ni 0.04  0.01 0.02  0.002 41.75  3.81 0.72  0.11
Mg 4556.06  125.93 1154.44  99.56 25.45  2.16 0.34  0.04

Natural gas condensate (NGC)


MeHg(II) 259.81  0.08 98.02  7.14 38.00  2.82 0.61  0.07
Zn 14.71  0.56 10.53  1.07 71.43  7.16 2.60  0.90
V 567.72  39.36 479.97  1.34 88.52  0.28 7.66  0.20
Cu 183.71  5.92 168.34  0.21 92.33  0.02 11.96  0.01
As 107.46  15.59 88.69  0.13 82.51  0.24 4.69  0.07
Ba 1.76  0.56 1.06  0.07 63.32  3.43 1.73  0.26
Ni 48.73  1.69 47.69  0.19 97.92  0.25 43.61  5.04
Mg 53.21  5.50 40.30  0.47 75.68  0.81 3.09  0.14

cations in the OPW and NGC after the addition of mercury species
are shown in Table 7. The adsorption selectivity of mercury ions
was decreased in the presence of foreign cations in OPW and NGC.
In OPW, the Zn adsorption had the highest Kd and hr, which
indicate the highest selectivity, followed by Hg(II). However, in the
case of NGC, the selectivity of Ni was the highest, followed with Cu,
V, Ad, Mg, Zn, Ba and MeHg(II). The difference of metal selectivity
trends of both OPW and NGC indicates that the chemical properties
of the adsorbate medium play an important factor in determining
the effectiveness of the cations adsorption process. In addition, the
Fig. 10. Regeneration efficiency of (a) Hg(II) and (b) MeHg(II) adsorption in multi- existing other metal ions in the OPW and NGC increased the
adsorption cycles. mercury ions competition towards the adsorption sites and
reduced their adsorption efficiency.
5 adsorption cycles. The Cycle 1 represents the adsorption onto the
fresh CP adsorbent, while the Cycles 2–5 are adsorption after 4. Conclusions
regeneration. After a sequence of 4 desorption cycles using KI
solution (KI-regenerated CP), the hreg of Hg(II) adsorption The adsorption of the Hg(II) and MeHg(II) onto the coconut pith
remained higher than 95%. In contrast, for the CP regenerated adsorbent from aqueous solution conducted in a batch adsorption
using HCl solution (HCl-regenerated CP), high hreg (i.e. hreg > 95%) system at various experimental conditions showed that both
of Hg(II) adsorption can only be achieved up to 2 desorption cycles. mercury ions have different adsorption characteristics. The Hg(II)
For the MeHg(II) adsorption, the hreg of HCl-regenerated CP adsorption capacity did not give a significant change over the pH
reduced gradually from 94.6% to 55.2% for the first to fourth cycles range studied, but it was highly affected towards the MeHg(II)
of desorption. The hreg of MeHg(II) onto KI-regenerated CP adsorption. The adsorption capacity of Hg(II) was 2.5 mmol/g,
however, remained constant around 77% after the first to third which was 5 times higher than that of MeHg(II) (0.5 mmol/g). The
cycles of desorption process. Thereafter, the hreg was reduced to Hg(II) and MeHg(II) adsorption isotherm data were best described
66.5%. In general, the performance of KI-regenerated CP was better by the Langmuir and Freundlich isotherm models, respectively. The
than HCl-regenerated CP. This is because, anion I has the greatest adsorption process of both Hg(II) and MeHg(II) was an endother-

affinity towards Hg ions as compared to anion Cl , and anion I mic and spontaneous process. The rate of Hg(II) adsorption was
tends to form a complex with Hg in solution compared to O Hg faster than MeHg(II) and increased with temperature and initial
complexes on adsorbent surfaces. Therefore, more Hg ions can be concentration. The pseudo-second order kinetic model fitted well
desorbed using KI solution, and consequently increased the to the kinetic data but the overall rate was controlled by the film
adsorption performance of the regenerated adsorbent. It can be diffusion. The enthalpy, mean energy and binding energy values
concluded that the CP adsorbent can be reused for removing Hg(II) indicated the formation of mercury-complexes was a dominant
ions without significantly decreasing its adsorption performance. mechanism for both mercury ions adsorption onto the CP
Adsorbent selectivity towards Hg(II) and MeHg(II) was carried adsorbent. It was further evidenced from the changes of the FTIR
out using OPW and NGC, respectively. In order to increase the and XPS spectra of the CP adsorbent before and after mercury
mercury ion concentration detectable levels, 0.25 mM of Hg(II) and adsorption. The regenerability study of Hg(II) adsorption onto the
MeHg(II) was spiked into the OPW and NGC, respectively. The CP adsorbent showed a better result as compared to the MeHg(II).
concentration of cations present in both OPW and NGC was A high selectivity of Zn and Hg(II) in OPW was observed, while the
measured. It was found that the Mg existed in high concentration selectivity of MeHg(II) in NG condensate was low, but a higher
in OPW, while the V, Cu and As in NGC. The concentrations of selectivity was observed for Ni and Cu.
N. Saman et al. / Journal of Environmental Chemical Engineering 4 (2016) 2487–2499 2499

Acknowledgements [18] S.W. Lee, G.V. Lowry, H. Hsu-Kim, Biogeochemical transformations of mercury
in solid waste landfills and pathways for release, Environ. Sci.: Processes
Impacts 18 (2016) 176–189 (in press).
The financial supports from the Ministry of Agriculture (MOA), [19] ASTM E1756-08, Standard Test Method for Determination of Total Solids in
Malaysia (Project No. 05-01-06-SF1006/79410) and MyBrain15 Biomass, ASTM International, West Conshohocken, U.S, 2008.
scholarship from the Ministry of Higher Education (MOHE), [20] ASTM E1755-01, Standard Test Method for Ash in Biomass, ASTM International,
West Conshohocken, U.S, 2007.
Malaysia are gratefully acknowledged. [21] British Standard BS 7755-3.3, Soil Quality-Part 3: Chemical Methods- Sec-
tion 3.3: Determination of Effective Cation Exchange Capacity and Base Sat-
uration Level Using Barium Chloride Solution, British Standard BS, 1995.
References [22] M. Zettlitzer, H.F. Scholer, R. Eiden, R. Falter, Determination of elemental,
inorganic and organic mercury in North German gas condensates and for-
[1] K. Pillay, E.M. Cukrowska, N.J. Coville, Improved uptake of mercury by sulphur- mation brines, Soc. Pet. Eng. SPE 37260 (1997) 509–516.
containing carbon nanotubes, Microchem. J. 108 (2013) 124–130. [23] A.U. Israel, R. Ogali, O. Akaranta, Extraction and characterization of coconut
[2] X.B. Feng, G.L. Qiu, Mercury pollution in Guizhou, southwestern China—an (Cocos nucifera L.) coir dust, Songklanakamin J. Sci. Technol. 33 (6) (2011) 717–
overview, Sci. Total Environ. 400 (1) (2008) 227–237. 724.
[3] P. Li, X. Feng, G. Qiu, Methylmercury exposure and health effects from rice and [24] T.S. Anirudhan, M.R. Unnithan, Arsenic (V) removal from aqueous solutions
fish consumption: a review, Inter. J. Environ. Res. Public Health 7 (6) (2010) using an anion exchanger derived from coconut coir and its recovery, Che-
666–2691. mosphere 6 (2007) 60–66.
[4] U.S. EPA, Capsule Report: Aqueous Mercury Treatment, U.S. Environmental [25] S. Yu, Z.L. He, C.Y. Huang, G.C. Chen, D.V. Calvert, Effect of anions on the
Protection Agency, Washington D.C, 1997 (EPA/625/R-97/004). capacity and affinity of copper adsorption in two variable charge soils, Bio-
[5] R. Qadeer, Adsorption behavior of ruthenium ions on activated charcoal from geochem 75 (2005) 1–18.
nitric acid medium, Colloids Surf. A 293 (2007) 217–223. [26] C.S. Zhu, L.P. Wang, W.B. Chen, Removal of Cu(II) from aqueous solution by
[6] S.K. Das, A.R. Das, A.K. Guha, A study on the adsorption mechanism of mercury agricultural by product: peanut hull, J. Hazard. Mater. 168 (2009) 739–746.
on Aspergillus versicolor biomass, Environ. Sci. Technol. 41 (2007) 8281– [27] Y.S. Ho, Isotherms for the sorption of lead onto peat: comparison of linear and
8287. non-linear method, Pol. J. Environ. Stud. 15 (1) (2006) 81–86.
[7] S. Rangabhashiyam, N. Anu, N. Selvaraju, Sequestration of dye from textile [28] K.Y. Foo, B.H. Hameed, Insight into the modeling of adsorption isotherm
industry wastewater using agricultural waste products as adsorbents, J. En- systems, Chem. Eng. J. 156 (1) (2010) 2–10.
viron. Chem. Eng. 1 (2013) 629–641. [29] M. Horsfall, A.I. Spiff, Effects of temperature on the sorption of Pb2+ and Cd2+
[8] K.A. Adegoke, O.S. Bello, Dye sequestration using agricultural wastes as from aqueous solution by Caladium bicolor (Wild Cocoyam) biomass, Electron.
adsorbents, Water Resour. Ind. 12 (2015) 8–24. J. Biotechnol. 8 (2) (2005) 1–9.
[9] T.S. Anirudhan, L. Divya, M. Ramachandran, Mercury(II) removal from aqueous [30] Y.S. Ho, J.C.Y. Ng, G. McKay, Kinetics of pollutant sorption by biosorbents:
solutions and wastewaters using a novel cation exchanger derived from co- review, Sep. Purif. Technol. 29 (2) (2000) 189–232.
conut coir pith and its recovery, J. Hazard. Mater. 157 (2008) 620–627. [31] R. Ocampo-Pérez, R. Leyva-Ramos, P. Alonso-Dacila, J. Rivera-Utrilla, M.
[10] K. Johari, N. Saman, S.T. Song, H. Mat, D.C. Stuckey, Utilization of coconut milk Sanchez-Polo, Modeling adsorption rate of pyridine onto granular activated
processing waste as low-cost mercury sorbent, Ind. Eng. Chem. Res. 52 (2013) carbon, Chem. Eng. J. 165 (2010) 133–141.
15648–15657. [32] S. Karthikeyan, B. Sivakumar, M. Sivakumar, Film and pore modeling for ad-
[11] K. Johari, N. Saman, S.T. Song, J.Y.Y. Heng, H. Mat, Study on Hg (II) removal from sorption of reactive red 2 from aqueous solution on to activated carbon
aqueous solution using lignocellulosic coconut fiber biosorbents—equilibrium prepared from bio-diesel industrial waste, E-J. Chem. 7 (S1) (2010) 175–184.
and kinetic evaluation, Chem. Eng. Commun. 201 (2014) 1198–1220. [33] C.W. Hui, B. Chen, G. McKay, Pore-surface diffusion model for batch adsorption
[12] S.T. Song, N. Saman, K. Johari, H. Mat, Surface chemistry modifications of rice processes, Langmuir 19 (2003) 4188–4196.
husk toward enhancement of Hg(II) adsorption from aqueous solution, Clean [34] G.L. Dotto, L.A.A. Pinto, Adsorption of food dyes acid blue 9 and food yellow
Technol. Environ. Policy 16 (2014) 1747–1755. 3 onto chitosan: stirring rate effect in kinetics and mechanism, J. Hazard.
[13] S.T. Song, N. Saman, K. Johari, H. Mat, Biosorption of mercury from aqueous Mater. 187 (2011) 164–170.
solution and oilfield produced water by pristine and sulfur functionalized rice [35] B.H. Hameed, M.I. El-Khaiary, Batch removal of malachite green from aqueous
residues, Environ. Progress Sustain. Energy 34 (5) (2015) 1298–1310. solutions by adsorption on oil palm trunk fibre: equilibrium isotherms and
[14] N. Saman, K. Johari, S.T. Song, H. Mat, Removal of Hg(II) and CH3Hg(I) using kinetic studies, J. Hazard. Mater. 154 (2008) 237–244.
rasped pith sago residue biosorbent, Clean-Soil Air Water 42 (11) (2014) 1541– [36] F. Pagnanelli, S. Mainelli, F. Veglio, L. Toro, Heavy metal removal by olive
1548. pomace: biosorbent characterization and equilibrium modeling, Chem. Eng.
[15] D. Karunasagar, M.V. Balarama Krishna, S.V. Rao, J. Arunachalam, Removal and Sci. 58 (2003) 4709–4717.
preconcentration of inorganic and methyl mercury from aqueous media using [37] S. Basha, Z.V.P. Murthy, B. Jha, Removal of Cu(II) and Ni(II) from industrial
a sorbent prepared from the plant Coriandrum sativum, J. Hazard. Mater. B 118 effluents by brown seaweed, Cystoseira indica, Ind. Eng. Chem. Res. 48 (2009)
(2005) 133–139. 961–975.
[16] S. Basha, Z.V.P. Murthy, B. Jha, Sorption of Hg(II) onto Carica papaya: exper- [38] K.M. Shareef, Sorbents for contaminants uptake from aqueous solutions. Part
imental studies and design of batch sorber, Chem. Eng. J. 147 (2009) 226–234. 1: heavy metals, World J. Agric. Sci. 5 (2009) 819–831.
[17] Y. Eom, J.H. Won, J.Y. Ryu, T.G. Lee, Biosorption of mercury (II) ions from [39] E. Sumesh, M.S. Bootharaju, A.T. Pradeep, A practical silver nanoparticles-
aqueous solution by garlic (Allium sativum L.) powder, Korean J. Chem. Eng. 28 based adsorbent for the removal of Hg2+ from water, J. Hazard. Mater. 189
(6) (2011) 1439–1443. (2011) 450–457.

Você também pode gostar