You are on page 1of 8

Chemical Engineering Journal 184 (2012) 205–212

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Adsorption of fluoride, arsenate and phosphate in aqueous solution by cerium


impregnated fibrous protein
Hui Deng a,∗ , Xili Yu b
a
Department of Environmental Science and Engineering, University of Liaoning Petroleum and Chemistry, Fushun 113001, China
b
School of Chemistry Chemical and Biological Engineering, University of Donghua, Shanghai 021, China

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

Article history: This study successfully synthesized a new adsorbent cerium impregnating fibrous protein (CeFP) to
Received 22 November 2011 remove fluoride, arsenate and phosphate from aqueous solutions. The sorbent was characterized by the
Received in revised form 2 January 2012 SEM, FTIR, BET specific surface area and other physical measurements. FTIR showed that the hydroxyl
Accepted 3 January 2012
group of hydrous metal oxide (Ce-OH) of CeFP occurred at 1122 cm−1 , and the pH of zero point charge
(pHpzc ) increasing up to 9.0 after Ce impregnation was benefit for electrostatic attraction for anions.
Keywords:
Batch experiments results indicated that sorption properties of three anions were strongly dependent
Cerium impregnating fibrous protein (CeFP)
on solution pH. The maximum sorption capacity of fluoride was evaluated to be 5.59 mmol/g at pH = 3.0,
Adsorption
Fluoride
and the optimum adsorption capacity for arsenate reached 2.3 mmol/g in the pH range of 3.0–7.0 and
Arsenate 2.11 mmol/g for phosphate in pH 4.0–7.0 at 303 K respectively, better than those of most sorbents avail-
Phosphate able in the market. And the increase of solution pH after adsorption revealed a possible ligand exchange
mechanism and the surface hydroxyl group of CeFP played an important role in the adsorption. The
Langmuir isotherm equation described all the three anions’ sorption equilibrium well at 303 K, 313 K and
323 K, and the thermodynamic parameters such as Go , Ho and So indicated that the nature of fluoride
sorption was spontaneous and exothermic while arsenate and phosphate adsorption were endothermic.
The pseudo-second-order rate model fitted the adsorption kinetics of three anions perfectly. The com-
petitive adsorption experiment indicated that CeFP had a stronger affinity to arsenate than to fluoride
and phosphate.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction water bodies is higher than 0.02 mg/L [7], and wastewater con-
taining phosphates must meet the discharge limit of 0.5–1.0 mg/L
Fluoride, arsenate and phosphate exert a significant impact on [8].
human health and other living organisms when discharged into the Various treatment techniques have been employed to remove
environment. Excess intake of fluoride usually has adverse effect on fluoride, arsenic and phosphates, including precipitation [9,10] and
general body metabolism [1] and leads to dental and skeletal fluo- coagulation [11], adsorption [12] or ion exchange process [13,14],
rosis, lesions of endocrine glands, thyroid and liver [2]. The World etc. Among various commercially available methodologies, adsorp-
Health Organization (WHO) has established the maximum accept- tion is relatively simple, economical, more appropriate for small
able concentration of fluoride in drinking water to be 1.5 mg/L [3]. communities and considered as one of the most popular separation
Exposure to arsenate can cause hyperkeratosis and hyper pigmen- methods. Common adsorbents involve activated alumina [15–17],
tation of skin, as well as carcinogenic diseases of skin, lungs, blood activated carbon [18–20], iron and iron oxide [21–23] and other
and kidneys, bladder cancers even ultimately death [4]. In 2001, materials [24–26].
the U.S. Environmental Protection Agency (EPA) lowered the limit Recently, metal ions including aluminum, iron, titanium, zirco-
of arsenic in drinking water from 50 to 10 ␮g/L [5]. Phosphate is nium and manganese especially rare earth elements [27] such as
an essential nutrient for organism growth in most ecosystems, but lanthanum and cerium are known to be have a specific affinity
excessive phosphate in water is one of the main factors leading to for fluoride, arsenic and phosphates. Various rare earth com-
eutrophication and deterioration of water bodies [6]. The eutroph- pounds have been extensively developed as the adsorbents for
ication would happen when the concentration of phosphate in hazardous anions from aqueous solutions and shown promising
result [28–31]. Tokunaga et al. [32] compared the fluoride removal
by multivalent metals like AI(III), Y(III), La(III), Ce(III/IV), and Zr(IV)
in the form of oxides, hydrous oxide and basic carbonates and found
∗ Corresponding author. Tel.: +86 431 6860759. that hydrous Ce(IV) oxide was most effective without significant
E-mail address: denghui0310@yahoo.cn (H. Deng). dissolution even at pH 2.0.

1385-8947/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2012.01.031
206 H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212

Fig. 1. SEM images of FP (a) and CeFP (b).

Fibrous protein (FP) usually comes from skin, bone and con- 2.3. Characterization
nective tissue of animals with protein structure composed of three
polypeptide chains of triple helical structure. There are abundant The Ce content of CeFP was obtained by measuring the resid-
functional groups such as carboxyl group of the side chains and car- ual Ce amount in the loading reaction solution using inductively
bonyl group of polypeptide chains in its molecule structure, which coupled plasma-atomic emission spectroscopy (ICP-AES) (ICP,
exhibit excellent chemical properties reacting with many kinds of Perkin-Elmer Optima 2100DV, USA). The specific area of CeFP was
metal ions (such as Al(III),Zr(IV), Fe(III)) [33]. FP has been used as determined by a surface area and porosity analyzer (Tristar3000,
a carrier of metal ions and applied for metal ions recovery [34]. Micromeritics, USA). The pH of zero point charge measurement
La3+ impregnated cross-linked gelatin was prepared for fluoride (pHpzc ) of the adsorbent was detected by the solid addition method
removal from drinking water or agroindustrial wastewater [35]. [37]. The surface morphology of the adsorbent was observed by
It was proved that fluoride on the Zr(IV) and Ti(IV) immobilizing scanning electron microscope (JSM-5900LV). Fourier transform
gel-phase was adsorbed from aqueous solution by both the ligand infrared (FTIR) spectra of CeFP adsorbent was recorded on Perkin-
exchange mechanism and the addition mechanism, while arsen- Elmer (USA) system 2000 spectrophotometer with a resolution of
ate and phosphate and selenite were adsorbed only by the ligand 2 cm−1 . And other physical properties were determined by com-
exchange mechanism [36]. However, reports about fibrous protein mon methods.
loaded with rare earth metal ion cerium for anions adsorption are
few. 2.4. Sorption experiments
The aim of this study was devoted to synthesize a new adsorbent
(CeFP) by bonding Ce onto fibrous protein and examine its ability Batch experiments were conducted in 250 mL flasks contain-
towards removal of fluoride (F), arsenate (As(V)) and phosphate ing 100 mL anions solution. Each experiment 0.100 g amount of
(P). Its characteristics such as surface morphology, microstructure, CeFP was suspended in a flask and shaken at 120 rpm in a ther-
surface charge and FTIR spectrum were described and sorption mostatic shaker for 24 h. The total adsorbed anions on the CeFP
behaviors for fluoride (F), arsenate (As(V)) and phosphate (P) were calculated from the difference between the initial and resid-
was evaluated by means of adsorption experiments including ual concentration of anions in solution. The fluoride concentration
adsorption kinetics, adsorption isotherm, solution pH effect and was analyzed by ion selective electrode method, and the arsen-
coexistence anions effect on adsorption. ate and phosphate concentration were determined by ICP-AES (ICP,
Perkin-Elmer Optima 2100DV, USA).
The investigation evaluating the pH effect was conducted at a
2. Materials and methods series of different initial solution pH varying from 2.0 to 12.0 by
means of 0.1 M HNO3 or 0.1 M NaOH. The procedures of adsorp-
2.1. Materials tion kinetics were kept similar to the previous pH experiments but
at the desired pH value and the anions concentrations were ana-
The fibrous protein was provided by China Academy of For- lyzed at a regular interval during adsorption. The sorption isotherm
est, Nanjin, China. Fluoride, phosphate, and arsenate stock solution experiments were studied at the desired pH value but varying initial
(10 mmol/L) were prepared separately by dissolving appropriate concentration of anions and carried out in 303 K, 313 K and 323 K
quantities of analytical grade sodium fluoride, potassium dihydro- respectively. Competitive adsorption effect was investigated using
gen phosphate, and dibasic sodium arsenate in deionized water. a mixed anions solution containing 2 mM fluoride, 1 mM arsenate
Other chemicals were of reagent grade. and 1 mM phosphate, and other sorption condition was as same as
that in sorption kinetic.

2.2. Loading of fibrous protein with cerium 3. Results and discussion

The sulfate of cerium and FP were mixed by pH 1.5 (adjusted by 3.1. CeFP characterization
1 M H2 SO4 solution) solution in the ratio of 1.19:1 (w/w) with con-
tinuously agitating for 4 h at room temperature. Then the solution SEM was used to observe the surface morphologies of raw mate-
pH was gradually increased to 3.0–3.5 by addition of sodium bicar- rial (FP) and prepared adsorbent (CeFP), and the obtained photos
bonate and constantly stirred for another 4–5 h. Finally the fibrous were shown in Fig. 1. The surface of FP was smooth and no promi-
product was rinsed by distilled water several times until no Ce ion nent coating was observed but in Fig. 1(b) it was clearly visible
was tested and vacuum-dried below 60 ◦ C for 7 h. that the CeFP fibers were covered by large quantities of fine grains,
H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212 207

Table 1
90 Physical properties of o-fibrous protein and Ce-fibrous protein.

80 Property items Limits

70 o-Fibrous protein Ce-fibrous protein


Adsorption percentage(%)

60 Particle size 11.71–16.16 nm 18.6–21.28 nm


Bulk density (dry) 0.051–0.053 g/cm3 0.519–0.521 g/cm3
50 Carbon surface area 2.11–2.15 m2 /g 3.57–3.65 m2 /g
Pore volume 0.0192–0.0196 cm3 /g 0.0046–0.0054 cm3 /g
40
Moisture 1.83–2.22 g H2 O/g 1.14–1.50 g H2 O/g
30 Metal ion content – 0.8122–0.8126 g/g
pHpzc 5.4 8.6–9.0
20

10

0
of other porous adsorbents such as activated carbon (1044 m2 /g
-10
[39]), which implied that the mass transfer of the adsorption pro-
-20 cess may mainly be attributed to chemical reaction between metal
fluoride Arsenate phosphate
ion Ce and anions rather than physical adsorption dependent on
FP Anions high specific area. Simultaneously, it was suggested that modifi-
CeFP cation of FP with Ce (pHpzc of hydrous cerium oxide is 6.7 [38])
significantly increased the pHpzc from pH 5.4 to pH 8.6–9.0. Thus
Fig. 2. Sorption of fluoride, arsenate and phosphate on the FP and CeFP adsorbent.
(1 g/L adsorbent, initial concentration of anions is 2 mmol/L, initial pH of fluoride the potential of the CeFP surface was positive when solution pH was
solution is 3.0, initial pH of arsenate and phosphate solution is 5.0, 303 K.) below 9.0, and expected to exert better adsorption performance for
anionic adsorbate than FP since the positive surface charge range
of FP was widened and surface electrostatic interaction between
maybe prone to the formation of uniform Ce complexes on the adsorbent and anions in the solution was enhanced.
surface of FP during the impregnation procedure.
Sorption capabilities of CeFP and FP for F, As(V) and P were 3.2. Effect of pH on adsorption
compared in Fig. 2. It was found that FP nearly has no adsorption
ability for anions, after modification by Ce its adsorption capaci- The adsorption amount of fluoride, arsenate and phosphate as a
ties for anions removal were drastically improved and the sorption function of initial pH values was given in Fig. 4. It was suggested that
percentages of fluoride, arsenate and phosphate were higher than the fluoride removal increased dramatically with an pH increase
84%. below pH 3.0 and then decreased sharply when solution pH was
Fig. 3 showed the FTIR spectra of raw material and the adsor- beyond 3.0.The adsorption extent of arsenate reached the adsorp-
bent. It was illustrated that Ce has been bonded onto the surface of tion maximum at pH 3.0 and then remained a constant level from
fibrous protein. The great band at 3330–3600 cm−1 was assigned to 3.0 to 7.0 but decreased with an further increasing pH, and phos-
the stretching vibration of OH on carboxyl, and the strong band phate maintained the highest adsorption capacity during the pH
at around 1122 cm−1 of CeFP (b) attributing to the Ce OH bond value range of 4.0–7.0. An examination of the final solution pH val-
stretching [38] unable to be observed in FTIR spectra of FP. The ues after the adsorption was undertaken and shown in Fig. 5(a)–(c).
appearance of intensified band of Ce OH group may indicate that It was observed that similar profiles for all three elements were pre-
the metal hydroxyl group of CeFP would play an important role in sented and the final solution pH value rose compared with the pH
anions’ adsorption. value before adsorption when the initial solution pH was below
The physical properties of FP and CeFP were listed in Table 1.
The specific area of CeFP adsorbent was much smaller than those

80 2.0
(a) fluoride
(b) Arsenate
70 (a) FP
(c) pkosphate
Adsorption capacity(mmol/g)

(b) CeFP
1.5
60

50 1.0
Transmittance

40

0.5
30

20
0.0

10
0 2 4 6 8 10 12
0 pH
4500 4000 3500 3000 2500 2000 1500 1000 500 0
wavelength (cm-1) Fig. 4. Effect of initial pH on the adsorption of fluoride (a), arsenate (b), and phos-
phate (c). (1 g/L adsorbent, initial fluoride concentration is 2 mmol/L, initial arsenate
Fig. 3. FTIR spectra of (a) raw material (FP) and (b) prepared adsorbent (CeFP). and phosphate concentration is 1 mmol/L, 303 K.)
208 H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212

a special pH value close to the pHpzc of CeFP, but dropped when


12 (a)
before adsorption solution pH was above this pH value.
after adsorption Anions could be adsorbed on the adsorbents through nonspecific
10 and/or specific adsorption [38]. Nonspecific adsorption involves
the coulombic forces and is mainly limited to pH-dependent sites
below pHpzc of the adsorbent [40]. The specific adsorption, how-
8
ever, involves ligand exchange reactions where the anions displace
OH and/or H2 O from the surface [41]. Furthermore, it is essen-
pH

6 tial to consider the speciation diagram of the adsorbate expected


for the characters of functional groups on the adsorbent related
to pH in adsorption course. HF is weakly ionized (pK1 = 3.2) in
4
solution at lower pH values [42]. Arsenate exists in solution in
the forms of H3 AsO4 , H2 AsO4 − , HAsO4 2− , and AsO4 3− (pK1 = 2.20,
2 pK2 = 6.67, pK3 = 11.53) [43] in different pH solution, and phos-
phate exists in the species of H3 PO4 , H2 PO4 − , HPO4 2− , and PO4 3−
0
(pK1 = 2.2, pK2 = 7.2, pK3 = 12.2) [44]. The fluoride adsorption capac-
0 2 4 6 8 10 12 14 16 ities were over 0.95 mmol/g in pH range 1.5–4.0 and the maximum
Number of samples of 1.97 mmol/g was at pH = 3.0, but the adsorption efficiency was
below 50% when initial solution pH was above 4.0, which sug-
gested that fluoride adsorption by CeFP was slightly influenced by
12 (b) before adsorption its existing speciation maybe due to deoxygenated root. As far as
after adsorption oxyacid groups of arsenate and phosphate were concerned, adsorp-
10 tion behaviors were a little different. The maximum adsorption
capacity 0.989 mmol/g for arsenate and 0.997 mmol/g for phos-
phate were obtained respectively in pH 3.0–7.0 and pH 4.0–7.0, and
8
this moment the predominant species of arsenate and phosphate
were H2 AsO4 2− and H2 PO4 − . At lower pH value below 3.0 (for arse-
pH

6 nate) and 4.0 (for phosphate) or higher pH above 7.0, the adsorption
efficiency decreased probably since the neutral species existing in
H3 AsO4 , and H3 PO4 were difficult to attach the adsorptive sites of
4
the sorbent surface.
In addition, the solution pH values increased after adsorp-
2 tion shown in Fig. 5(a)–(c), and similar adsorption behaviors have
also been observed by Zhang et al. [38,45]. As a metal-ion-based
0 adsorbent, CeFP adsorbent possibly contained the hydroxyl group
0 2 4 6 8 10 12 (Ce-OH) in aqueous environment [38] which have been certificated
Number of sample in Fig. 3. Then it was implied that ligand exchange might take place
in adsorption and the surface complexes between anions and CeFP
(c) might form, proposed as in the following scheme:
12
FPMOH + H+ + R− ⇔ FPMR + H2 O (1)
before adsorption
10
after adsorption FPMOH + R− ⇔ FPMR + OH− (2)
FPMOH + H+ + R− ⇔ FPMOH+
2
· · ·R− (3)
8 where M represents the metal ion Ce(IV), and R represents the
anions of F− , H2 AsO4 2− , and H2 PO4 − respectively. According to Eqs.
pH

(1)–(3) the adsorption of the three anions on CeFP can result in an


6
increase of solution pH due to consumption of H+ and discharge
of OH− in the adsorption process. When the solution pH value
4 was above 7.0, the OH− content in solution became higher, which
may compete with anions for exchangeable sites on the fibers. At
the higher pH beyond pH = 9.0, the CeFP surface was electrically
2 negative with a decrease of adsorption capacity of anions due to
the strong electrostatic repulsion and a decrease of the solution
0 2 4 6 8 10 12 pH values after adsorption was resulted from the consumption of
Number of samples hydroxyl ions.

Fig. 5. The pH of anions solution before and after adsorption: (a) fluoride, (b) arse- 3.3. Adsorption kinetics
nate, and (c) phosphate. (1 g/L adsorbent, initial fluoride concentration is 2 mmol/L,
initial arsenate and phosphate concentration is 1 mmol/L, 303 K.)
The adsorption kinetic experiments were carried out at 25 ◦ C in
order to design appropriate adsorption technologies. As illustrated
in Fig. 6(a), these three anions showed different time-dependent
behavior. The adsorption of arsenate took place rapidly and adsorp-
tion amount attained a stable value fast within 2 h compared with
some porous adsorbents [46], while the uptakes of fluoride and
phosphate were slow and approached equilibrium gradually.
H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212 209

1.5 Table 2
(a) Pseudo-second-order rate constants (k) of anions on the CeFP.
1.4 Phosphate
Arsnate Anions Conc. qeexp k2 qe R2
1.3
Fluoride (mmol/L) (mmol/g) (g/mmol min) (mmol/g)
1.2
F−
Adsorption capacity(mmol/g)

2 1.3451 0.09956 1.337 0.9994


1.1 3 1.9562 0.13926 1.940 0.9998
4 2.5759 0.12997 2.587 0.9999
1.0
AsO4 3− 1 0.9051 0.19695 0.9010 0.9996
0.9 2 1.6202 1.36581 1.6219 0.9999
0.8 PO4 3− 1 0.8715 0.06500 0.8817 0.9975
2 1.4157 0.09950 1.4074 0.9940
0.7

0.6

0.5
correlation coefficient R2 for the pseudo-second-order rate model
0.4
was extremely high (>0.99), which suggested that the pseudo-
0.3 second-order model could describe the adsorption of these three
0 100 200 300 400 500 600 anions on CeFP perfectly. Additionally it was found that the theo-
Time t(minutes) retical adsorption capacities (qe ) calculated according to the model
were significantly close to the experimental values (qeexp ). From the
600 (b) accuracy of the model it was reasonably assumed in the sorption
Phosphate process, concentrations of both sorbate and sorbent were involved
500 Arsenate in rate-determining step, which maybe a chemical sorption or
Fluoride chemisorption [48], as reported by Liu et al. [31,49].
400
t/qt (min.g/mmol)

3.4. Adsorption isotherm


300

Adsorption isotherm describes how adsorbates interact with


200 adsorbents and is important in optimizing the use of adsorbents.
The adsorption equilibrium of fluoride, arsenate and phosphate was
100 investigated over a range of anions concentration at 303 K, 313 K
and 323 K presented in Fig. 7(a)–(c), which gave a series of plots of
0
the adsorption capacity (qe , mmol/g) versus the equilibrium anion
concentration (Ce , mmol/L). The adsorption capacities were found
0 100 200 300 400 500 600 to increase with increasing equilibrium concentrations of anions
Time t(min) and increasing temperatures. To determine the maximum sorp-
tion capacities of anions on CeFP, the Langmuir isotherm equation
Fig. 6. Adsorption kinetics of fluoride, arsenate and phosphate on CeFP (a). Pseudo- was used to describe the equilibrium experimental results given as
second-order rate model fitting of adsorption kinetics data (b). (1 g/L adsorbent, follows:
initial pH of fluoride solution is 3.0, initial pH of arsenate and phosphate solu-
tion is 5.0; initial fluoride concentration is 2 mmol/L, initial arsenate and phosphate
bqmax Ce
concentration is 1 mmol/L, 303 K.) qe = (5)
1 + bCe

where Ce was equilibrium concentration (mmol/L), qe was the


amount adsorbed under equilibrium (mmol/g), qmax was the the-
The pseudo-second-order model has been widely applied to oretical maximum adsorption capacity, and b (L/mmol) was a
the adsorption of pollutants from aqueous solution [47] and used Langmuir constant related to energy of adsorption. The model
to describe chemisorption. In order to interpret the experimental parameters of the Langmuir isotherm were summarized in Table 3.
result, the time-dependent sorption data has been analyzed using It could be seen that the Langmuir model fitted the experimental
the linear form of the pseudo-second-order kinetic equation. As data well according to the high correlation coefficients R2 (>0.99).
shown in Fig. 6(b) the plots were found to be linear over the entire The predicted maximum adsorption capacities (qmax ) of fluoride
sorption period and the pseudo-second-order model was given as decreased slightly with increasing temperature, while those of
follows: arsenate and phosphate increased, probably due to the oxyacid
t 1 t groups of arsenate and phosphate. The calculating adsorption
= + (4) capacities of fluoride, arsenate, and phosphate of CeFP at 303 K were
qt k2 q2e qe
up to 17.5 mg/g, 48.73 mg/g, and 82.65 mg/g at the equilibrium con-
where k2 was the constant of the pseudo-second-order rate centration of 1 mg/L respectively, higher than those of commonly
(g/mmol min), on the basis of the experimental data the equilib- used adsorbents such as activated carbon, activated alumina and
rium adsorption capacity qe , qt and k2 could be concluded from the ferric hydroxide reported in the literatures in Table 4 [31,50–60]. In
slope and intercept respectively. addition, the satisfactory fitting of the Langmuir isotherm for anions
The fitting parameters of kinetic studies for anions adsorption on adsorption suggested that chemical adsorption mechanism might
CeFP by the pseudo-second-order rate model were listed in Table 2. be involved in the adsorption process and anions might be adsorbed
The goodness of the linear plot of the kinetic model could be judged in the formation of monolayer coverage of fluoride, arsenate and
from the value of the coefficient of the plot, which could also be phosphate on the surface of CeFP.
regarded as the criterion in the determination of adequacy of a To evaluate the thermodynamic feasibility and further analyze
kinetic model. Remarkably, it can be seen from Table 2 that the the nature of the adsorption process three basic thermodynamic
210 H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212

5 (a) 100

4 80
Adsorption capacity(mmol/g)

Adsorption percentage(%)
3
60
323K
313K
2
303K
40

1
20

0
0.0 0.5 1.0 1.5 2.0 2.5 F As P
Equilibrium concentration of fluoride(mmol/L) individal Anion
mixture

2.5
(b) Fig. 8. Competitive anions adsorption on CeFP. (1 g/L adsorbent, initial fluoride con-
centration is 2 mmol/L, initial arsenate and phosphate concentration is 1 mmol/L,
initial pH is 3.0, 303 K.)
2.0
Adsorption Capacity(mmol/g)

parameters, standard free energy (Go ), standard enthalpy change


1.5 (Ho ) and standard entropy change (So ) were calculated [61].
323K
313K Go = −RT ln K0 (6)
1.0 303K
H o S o
ln K0 = + (7)
RT R
0.5
Go was free energy of sorption (kJ/mol), K0 was the sorption
equilibrium constant, T was the absolute temperature in Kelvin, R
0.0 was the universal gas constant (8.314 J/mol K), Ho was the stan-
dard enthalpy change (kJ/mol), and So was the standard entropy
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 change (J/mol K). The values of the thermodynamic parameters
Equilibrium concentration of Arsenate(mmol/L) were shown in Table 3. The K0 equaled to the fluoride adsorbed
3.0
per kg sorbent divided by the equilibrium concentration of the
(c) solution, and came from the equilibrium experiment data at ini-
tial concentration of anions 2 mmol/L at different temperatures.
2.5 Ho and So could be calculated from the slope and intercept of
the plot of ln K0 versus 1/T, respectively. The negative value Go
Adsorption capacity(mmol/g)

2.0 confirmed the feasibility and the spontaneous nature of the anions
sorption. The enthalpy change (Ho ) of fluoride was found to be
negative while Ho values of arsenate and phosphate were posi-
1.5
tive probably due to being oxyacid groups. The positive value of the
323K enthalpy change (Ho ) for arsenate and phosphate indicated that
313K
1.0 the sorption process on the adsorbent was definitely endothermic
303K
in nature. The positive value of the entropy change (So ) suggested
that the adsorption on the adsorbent surface occurred with increas-
0.5
ing entropy possible due to the release of lots of molecular water
at the solid/liquid interface during the sorption process [62].
0.0

3.5. Competitive adsorption


-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Equilibrium concentration of phosphate(mmol/L)
The results of competitive adsorption experiments were shown
Fig. 7. Sorption isotherm of fluoride (a), arsenic (b) and phosphate (c) on the CeFP in Fig. 8 Among fluoride, arsenate and phosphate, the adsorption
(1 g/L adsorbent, initial pH of fluoride solution is 3.0, initial pH of arsenate and efficiency of fluoride dropped most drastically from 92% to 57%
phosphate solution is 5.0). when three anions were mixed, which indicated that fluoride was
most susceptible to be interfered by the presence of the other
two anions. Simultaneously CeFP appeared the highest selectiv-
ity to arsenate since the removal percentage of arsenate decreased
least. Furthermore, it can be seen that CeFP appeared to have the
strongest affinity to arsenate based on the maximum mass uptake
percentage in mixed solution.
H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212 211

Table 3
Parameters modeling by the Langmuir adsorption isotherm and thermodynamic constants for fluoride, arsenate and phosphate sorption on CeFP.

Anion Langmuir parameters Thermodynamic constants

Temperature (K) qmax (mmol/g) b (L/mmol) R2 G (kJ/mol) Ho (kJ/mol) So (kJ/mol K)

Fluoride 303 5.5975 3.2630 0.9932 −10.856 −4.008 0.0226


313 5.4722 2.8010 0.9972 −11.082
323 5.3422 2.1583 0.9992 −11.308

Arsenate 303 2.3012 21.459 0.9951 −8.1544 2.1076 0.0339


313 2.3389 31.100 0.9967 −8.4931
323 2.4772 47.940 0.9981 −8.8317

Phosphate 303 2.1110 40.416 0.9961 −8.3377 0.9187 0.0305


313 2.3588 52.081 0.9986 −8.6431
323 2.1945 198.13 0.9997 −8.9486

Table 4
Comparison of adsorption capacity of some conventional sorbents for fluoride, arsenate and phosphate.

Adsorbents pH Equilibrium concentration, Sorption capacity, qe.cal (mg/g) Reference


Ce (mg/L)

Activated carbon 7.0 1 0.075 (fluoride) [50]


Activated carbon 6.4–7.5 – 30.48 (arsenate) [51]
Activated carbon loaded with Fe(III)oxide 3.0 1 0.46 (phosphate) [52]
Bone char 7.0 1 2.71 (fluoride) [50]
Bone char 10 – 0.7933 (arsenate) [53]
Fly ash 9.0 58.7 8.26 (phosphate) [54]
Polymeric resin 7.0 1 3.62 (fluoride) [50]
Iron(III)-loaded chelating resin 3.5 1 1.54 (arsenate) [55]
Activated alumina 7.0 1 0.96 (fluoride) [50]
Activated alumina 5.2 1 9.44 (arsenate) [56]
Aluminum oxide hydroxide 4.0 1 1.39 (phosphate) [57]
Ferric hydroxide 6.0–7.0 1 ∼3 (fluoride) [58]
Ferric hydroxide 7.0 – 68.75 (arsenate) [59]
Iron oxide tailings 3.2 1 3.59 (phosphate) [60]
Ion exchange fiber 3.0 1 0.273 (fluoride) [31]
Ion exchange fiber 3.5–7.0 1 1.062 (arsenate) [31]
Ion exchange fiber 3.0–5.5 1 1.209 (phosphate) [31]
CeFP 3.0 1 17.5 (fluoride) This study
CeFP 3.0–7.0 1 48.73 (arsenate) This study
CeFP 4.0–7.0 1 82.65 (phosphate) This study

4. Conclusions [2] A. Tor, N. Danaoglu, G. Arslan, Y. Cengeloglu, Removal of fluoride from water
by using granular red mud: batch and column studies, J. Hazard. Mater. 164
(2009) 271–278.
CeFP has been developed for removal of fluoride, arsenate and [3] WHO (World Health Organization), Guidelines for Drinking Water Quality,
phosphate by impregnating rare earth metal ion Ce onto the fibrous World Health Organization, Geneva, Switzerland, 1993.
protein. The adsorbent was found to be efficient in the adsorption [4] S. Bang, M. Patel, L. Lippincott, X.G. Meng, Removal of arsenic from groundwater
by granular titanium dioxide adsorbent, Chemosphere 60 (2005) 389–397.
of the three anions, and fluoride and the single charged species of [5] P. Kumari, P. Sharma, S. Srivastava, M.M. Srivastava, Biosorption studies on
H2 AsO4 2− and H2 PO4 − appeared to be better adsorbed in an acidic shelled Moringa oleifera Lamarck seed powder: removal and recovery of arsenic
solution. Simultaneously, the introduction of metal ion Ce on the from aqueous system, Int. J. Miner. Process. 78 (2006) 131–139.
[6] P.S. Lau, N.F.Y. Tam, Y.S. Wong, Wastewater nutrients (N and P) removal by
surface of fibrous protein also improved the pHpzc of the FP material
carrageenan and alginate immobilized Chlorella vulgaris, Environ. Technol. 18
and enhanced the adsorptive ability for anions, as fibrous protein (1997) 945–951.
has no ability for adsorbing any one sort of three anions. Addition- [7] C.Y. Wang, J.P. Zhai, R. Nie, L. Huang, Experimental study on phosphorus
removal by activated sludge process in treating wastewater of low phosphorus
ally, the adsorption mechanism might include both electrostatic
concentration, Environ. Protect. Sci. 31 (4–6) (2005) 20 (in Chinese).
attraction and the ligand exchange between the anions and the [8] A. Ugurlu, B. Salman, Phosphorus removal by fly ash, Environ. Int. 24 (1998)
metal hydroxyl group on the surface of CeFP. Kinetics fitting by 911.
the pseudo-second-order model indicated that the sorption pro- [9] C.J. Huang, J.C. Liu, Precipitate flotation of fluoride-containing wastewater from
a semiconductor manufacture, Water Res. 33 (16) (1999) 3403–3412.
cess was rate-determining step, and the adsorption patterns of CeFP [10] Y. Jia, G.P. Demopoulos, Coprecipitation of arsenate with iron(III) in aqueous
can be described well by the Langmuir isotherm with high correla- sulfate media: effect of time, lime as base and co-ions on arsenic retention,
tion coefficients greater than 0.99. The affinity preference of CeFP Water Res. 42 (3) (2008) 661–668.
[11] M.G. Mitrakas, P.C. Panteliadis, V.Z. Keramidas, R.D. Tzimou-Tstouridou, C.A.
for anions was arsenate, phosphate and fluoride in sequence, and Sikaladis, Predicting Fe3+ dose for As(V) removal at pHs and temperatures
the maximum adsorption capacity of fluoride, arsenate, and phos- commonly encountered in natural water, Chem. Eng. J. 155 (2009) 716–721.
phate at 303 K were 5.597 mmol/g, 2.30 mmol/g, and 2.181 mmol/g [12] S.H. Lee, B.C. Lee, K.W. Lee, S.H. Lee, Y.S. Cjoi, K.Y. Park, M. Iwamoto, Phospho-
rus recovery by mesoporous structure material from wastewater, Water Sci.
respectively, which showed a great potential application of CeFP in Technol. 55 (2007) 169–176.
water treatment to minimize the presence of these toxic anions. [13] Y. Ku, H.M. Chiou, W. Wang, The removal of fluoride ion from aqueous solution
by a cation synthetic resin, Sep. Sci. Technol. 37 (1) (2002) 89–103.
[14] K. Hristovoski, P. Westerhoff, T. Moller, P. Sylvester, W. Condit, H. Mash,
References Simultaneous removal or perchlorate and arsenate by ion-exchange media
modified with nanostructured iron (hydr)oxide, J. Hazard. Mater 152 (1) (2008)
[1] S. Kagne, S. Jagtap, P. Dhawade, S.P. Kamble, S. Devotta, S.S. Rayalu, Hydrate 397–406.
cement: a promising adsorbent for the removal of fluoride from aqueous solu- [15] J.L. Shupe, H.P. Peterson, N.C. Leone, Fluoride Effects on Vegetation Animals and
tion, J. Hazard. Mater. 154 (2008) 88–95. Humans, Paragon Press, Salt Lake City, 1983.
212 H. Deng, X. Yu / Chemical Engineering Journal 184 (2012) 205–212

[16] T.S. Singh, K.K. Pant, Equilibrium, kinetics and thermodynamic studies for [40] F.J. Hingston, R.J. Atkinson, A.M. Posner, J.P. Quirk, Specific adsorption of anions,
adsorption of As(III)on activated alumina, Sep. Purif. Technol. 36 (2004) Nature 215 (1967) 1459–1461.
139–147. [41] G. Prasad, Removal of arsenic(V) from aqueous systems by adsorption on to
[17] H. Brattebo, H. Odeggaard, Phosphorus removal by granular activated alumina, some geological materials, in: J.O. Nriagu (Ed.), Arsenic in the Environment,
Water Res. 20 (1986) 977–986. John Wiley, New York, 1994, pp. 33–152.
[18] M. Srimurali, A. Pragathi, J. Karthikeyan, A study on removal of fluorides from [42] A. Eskandarpour, M.S. Onyango, A. Ochieng, S. Asai, Removal of fluoride ions
drinking water by adsorption onto low-cost materials, Environ. Pollut. 99 from aqueous solution at low pH using schwertmannite, J. Hazard. Mater. 152
(1998) 285–289. (2008) 571–579.
[19] D. Mohan, C.U. Pittman Jr., Arsenic removal from water/wastewater using [43] S. Marity, S. Chakravarty, S. Bhattacharjee, et al., A study on arsenic adsorp-
adsorbents—a critical review, J. Hazard. Mater. 142 (2007) 1–53. tion on polymetallic sea nodule in aqueous medium, Water Res. 39 (2005)
[20] C. Namasivayam, D. Sangeetha, Equilibrium and kinetic studies of adsorption 2579–2590.
of phosphate onto ZnCl2 activated coir pith carbon, J. Colloid Interface Sci. 280 [44] N.I. Chubar, V.A. Kanibolotskyy, V.V. Strelko, G.G. Gallios, V.F. Samanido, T.O.
(2004) 359–365. Shaposhnikova, V.G. Milgrandt, I.Z. Zhuravlev, Adsorption of phosphate ions
[21] T. Hiemstra, W.H. Van Riemsdijk, Fluoride adsorption on goethite in relation to on novel inorganic ion exchangers, Colloid. Surf. A 255 (2005) 55–63.
different types of surface sites, J. Colloid Interface Sci. 225 (2000) 94–104. [45] S.B. Deng, R.B. Bai, Adsorption and desorption of humic acid on aminated poly-
[22] S. Bang, G.P. Korfiatis, X.G. Meng, Removal of arsenic from water by zero-valent acrylonitrile fibers, J. Colloid Interface Sci. 280 (2004) 36–43.
iron, J. Hazard. Mater. 121 (2005) 61–67. [46] E. Ou, Z.J. Junjie, M.C. Shaochun, W.Q. Jiaqiang, X. Fei, M. Liang, Highly efficient
[23] A. Ler, R. Stanforth, Evidence for surface precipitating of phosphate on goethite, removal of phosphate by lanthanum-doped mesoporous SiO2 , Colloids Surf. A:
Environ. Sci. Technol. 37 (2003) 2694–2700. Physicochem. Eng. Aspect 308 (2007) 47–53.
[24] M. Mahramanlioglu, I. Kizilcikli, I.O. Bicer, Adsorption of fluoride from aqueous [47] Y.S. Ho, Second-order kinetic model for the sorption of cadmium onto tree fern:
solution by acid treated spent bleaching earth, J. Fluorine Chem. 115 (2002) a comparison of linear and non-linear methods, Water Res. 40 (2006) 119–125.
41–47. [48] Y.S. Ho., G. McKay, Pseudo-second-order model for sorption processes, Process.
[25] M.J. DeMarco, A.K. Sengupta, J.E. Greenleaf, Arsenic removal using a polymet- Biochem. 34 (1999) 451–465.
ric/inorganic hybrid sorbent, Water Res. 37 (2003) 164. [49] K. Baris, O. Duygu, G. Ali, V.N. Bulut, D. Celal, S. Mustafa, Removal of fluoride ions
[26] K.M. Parida, S. Mallick, S.S. Das, Studies on manganese nodule leached residues. from aqueous solution by waste mud, J. Hazard. Mater. 168 (2009) 888–894.
2. Adsorption of aqueous phosphate on manganese nodule leached residue, J. [50] N.A. Medellin-Castillo, R. Leyva-Ramos, R. Ocampo-Perez, R.F.G. de la Cruz, A.
Colloid Interface Sci. 290 (2005) 22–27. Aragon-Pina, J.M. Matinez-Rosales, R.M. Guerrero-Coronada, L. Fuentes-Rubio,
[27] I. Hideaki, N. Junji, I. Yuzuru, K. Tokuzo, Anion adsorption behavior of rare earth Adsorption of fluoride from water solution on bone char, Ind. Eng. Chem. Res.
oxide hydrates, Chem. Soc. Jpn. 5 (1987) 807–813 (in Japanese, with English 46 (2007) 9205–9212.
abstract). [51] J. Pattanayak, K. Mondal, S. Mathew, S.B. Lalvani, A parametric evaluation of
[28] S. Tokunaga, S.A. Wasay, S.W. Park., Removal of arsenic(V) ion from aqueous the removal of As(V) and As(III) by carbon based adsorbents, Carbon 38 (2000)
solutions by lanthanum compounds, Water Sci. Technol. 35 (1997) 71–78. 589–596.
[29] S.A. Wasay, M.J. Haron, S. Tokunaga, Adsorption of fluoride, phosphates and [52] Z.L. Shi, F.M. Liu, S.H.Yao., Adsorptive removal of phosphate from aqueous solu-
arsenate ions on lanthanum-impregnated silica gel, Water Environ. Res. 6 tions using activated carbon loaded with Fe(III)oxide, New Carbon Mater. 26
(1996) 295–300. (4) (2011) 299–306.
[30] T.M. Suzuki, J.O. Bomani, M. Hideyuki, Y. Toshiro, Preparation of porous resin [53] Y.N. Chen, L.Y. Chai, Y.D. Shu, Study of arsenic(V) adsorption on bone char from
loaded with crystalline hydrous zirconium oxide and its application to the aqueous solution, J. Hazard. Mater. 160 (2008) 168–172.
removal of arsenic, React. Funct. Polym. 43 (2000) 165–172. [54] N.M. Agyei, C.A. Strydom, J.H. Potgieter, The removal of phosphate ions from
[31] R.X. Liu, J.L. Guo, H.X. Tang, Adsorption of fluoride, phosphates and arsenate aqueous solution by fly ash, slag, ordinary Portland cement and related blends,
ions on a new type of ion exchange fiber, J. Colloid Interface Sci. 248 (2002) Cem. Concr. Res. 32 (2002) 1889–1897.
268–274. [55] H. Matsunaga, T. Yokoyama, R.J. Eldridge, B.A. Bolto, Adsorption characteristics
[32] S. Tokunaga, M.J. Haron, S.A. Wasay, Removal of fluoride ions from aqueous of arsenic(III) and arsenic(V) on iron(III)-loaded chelating resin having lysine-
solution by multivalent metal compounds, Int. J. Environ. Stud. 48 (1995) N␣ ,N␣ -diacetic acid moiety, React. Polym. 29 (1996) 167–174.
17–18. [56] T.F. Lin, J.K. Wu, Adsorption of arsenite and arsenate within activated alumina
[33] N.A. Evans, B. Milligan, K.C. Montgomery, Collagen crosslinking: new binding grains: equilibrium and kinetics, Water Res. 35 (2001) 2049–2057.
sites for mineral tannage, J. Am. Leather Chem. Assoc. 82 (1987) 86–95. [57] S. Tanada, M. Kabayama, N. Kawasaki, T. Sakiyama, T. Nakamura, M. Araki, T.
[34] P.D. Unger, R. Rohrbach, Use a solvent impregnated crosslinked matrix for metal Tamura, Removal of phosphate by aluminum oxide hydroxide, J. Colloid Inter-
recovery, Hydrometallurgy 46 (1997) 387–390. face Sci. 257 (2003) 135–140.
[35] Z. Yuming, C. Yu, Y. Shan, Adsorption of fluoride from aqueous solution on [58] E. Kumar, A. Bhatnagar, M. Ji, W. Jung, S.H. Lee, S.J. Kim, G. Lee, H. Song, J.Y.
La3+ -impregnated cross-linked gelatin, Sep. Purif. Technol. 36 (2004) 89–94. Choi, J.S. Yang, B.H. Jeon, Defluoridation from aqueous solutions by granular
[36] Y. Akio, M. Kiyohiro, Adsorption of anions to zirconium(IV) and tita- ferric hydroxide (GFH), Water Res. 43 (2009) 490–498.
nium(IV) chemically immobilized on gel-phase, J. Chromatogr. A 1082 (2005) [59] B.J. Lafferty, R.H. Loeppert, Methyl arsenic adsorption and desorption behavior
208–213. on iron oxides, Environ. Sci. Technol. 39 (7) (2005) 2120–2127.
[37] S.S. Tripathy, S.B. Kanungo, Adsorption of Co2+ , Ni2+ , Cu2+ and Zn2+ from 0.5 M [60] L. Zeng, X.M. Li, J.D. Liu, Adsorptive removal of phosphate from aqueous solu-
NaCl and major ion sea water on a mixture of ␦-MnO2 and amorphous FeOOH, tions using iron oxide tailings, Water Res. 38 (2004) 1318–1326.
J. Colloid Interface Sci. 284 (2005) 30–33. [61] C. Jianguo, L. Aimin, S. Hongyan, F. Zhenghao, L. Chao, Z. Quanxing, Adsorp-
[38] Y. Zhang, M. Yang Min, X. Huang, Arsenic removal with a Ce(IV)-doped iron tion characteristics of aniline and 4-methylaniline onto bifunctional polymeric
oxide adsorbent, Chemosphere 51 (2003) 945–952. adsorbent modified by sulfonic groups, J. Hazard. Mater. B 124 (2005) 173–180.
[39] R. Leyva Ramos, J. Ovalle-Turrubiartes, M.A. Sanchez-Castillo, Adsorption of [62] K. Biswas., S.K. Saha, U.C. Ghosh, Adsorption of fluoride from aqueous solution
fluoride from aqueous solution on aluminum-impregnated carbon, Carbon 37 by a synthetic iron(III)–aluminum(III) mixed oxide, Ind. Eng. Chem. Res. 46
(1999) 609–617. (2007) 5346–5356.

You might also like