Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Colloid and Interface Science 452 (2015) 69–77

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Preparation of iron nanoparticles-loaded Spondias purpurea seed waste


as an excellent adsorbent for removal of phosphate from synthetic and
natural waters
M. Arshadi a,⇑, S. Foroughifard b, J. Etemad Gholtash c, A. Abbaspourrad d
a
Department of Chemistry, Shiraz Branch, Islamic Azad University, P.O. Box 71955-149, Shiraz, Fars, Iran
b
Department of Fishery, University of Guilan, P.O. Box 1144, Sowmesara, Iran
c
Department of Chemistry, Firozabad Branch, Islamic Azad University, Firozabad, Fars, Iran
d
School of Engineering and Applied Sciences, Department of Physics, Harvard University, Cambridge, MA 02138, United States

g r a p h i c a l a b s t r a c t

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

Article history: The synthesis and characterization of nanoscale zerovalent iron particles (NZVI) supported on Spondias
Received 5 March 2015 purpurea seed waste (S-NaOH-NZVI) was performed for the adsorption of phosphate (P) ions from waste
Accepted 10 April 2015 waters. The effects of various parameters, such as contact time, pH, concentration, reusability and tem-
Available online 16 April 2015
perature were studied. The adsorption of phosphate ions has been studied in terms of pseudo-first- and
-second-order kinetics, and the Freundlich, and Langmuir isotherms models have also been used to the
Keywords: equilibrium adsorption data. The adsorption kinetics followed the mechanism of the pseudo-second-
Spondias purpurea
order equation. The thermodynamic parameters (DG, DH and DS) indicated that the adsorption of phos-
Adsorption
NZVI
phate ions were feasible, spontaneous and endothermic at 25–80 °C. No significant loss of activity was
Phosphate observed; confirming that the S-NaOH-NZVI has high stability during the adsorption process even after
12th runs. The suggested adsorbent in this paper was also implemented to remove P from the Persian
Gulf water. XRD, FTIR and EDX analysis indicated the presence of Fe3 (PO4)28H2O (vivianite) on the
S-NaOH-NZVI@P surface.
Ó 2015 Elsevier Inc. All rights reserved.

⇑ Corresponding author.
E-mail addresses: m-arshadi@ch.iut.ac.ir, mohammadarshadi@yahoo.com (M. Arshadi).

http://dx.doi.org/10.1016/j.jcis.2015.04.019
0021-9797/Ó 2015 Elsevier Inc. All rights reserved.
70 M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77

1. Introduction magnetic characteristics of nano iron accelerate the easy separa-


tion of NZVI from soil and water, via a magnetic field. Bare NZVI
The fast growth of advanced economy, the surplus liberation of has many benefits such as high reactivity and selectivity, since in
the phosphate ions including agricultural, industrial and domestic bare NZVI every single catalytic entity can act as a single active
wastewater originates eutrophication, and eventually reducing the site. Despite their advantages, in some cases bare NZVI have not
water quality, causing the bloom of aquatic plants, depletion of dis- been commercialized because of difficulty encountered when the
solved oxygen and growth of algae [1,2]. Therefore, it is vital to bare NZVI should be separated and cleaned from the reaction mix-
purify phosphorus from the phosphate contaminated waste. ture. However, many industrial problems such as corrosion and
Many methods for treatment are available such as precipitation, deposition on reactor wall are associated with these bare NZVI.
adsorption, membrane processes, and ionic exchange. However, These drawbacks were moderate if bare NZVI was supported on
these techniques have inherent limitations, such as poor efficiency, the insoluble solid supports with excellent chemical and thermal
sensitive operating conditions and the production of a secondary stability and easily accessibility. Immobilization is robust enough
sludge which requiring further costly disposal [3]. These disadvan- to endure the harsh reaction conditions and the NZVI can be used
tages, together with the need for more economical and effective for many times [5–7]. This work reports the preparation of nanos-
methods for recovering the contaminants, have resulted in the cale zerovalent iron supported on S. purpurea seed waste as an
development of alternative separation technologies. One such inexpensive and potent nano-biosorbent in removing phosphate
alternative is biosorption, in which certain types of biomass are ions from aqueous media. The effect of contact time and initial
able to bind and concentrate metals from even very dilute aqueous concentration, temperature, reusability, coexisting ions and pH
solutions. However, the materials with biological origin present on the removal of phosphate ion has been studied. The mechanism
several advantages with respect to activated carbon, including bet- of P ions removal was also studied by EDS, FTIR and XRD
ter mechanical stability, diversity of chelating groups on the sur- techniques.
face, and they are often inexpensive, easily available and likely
easy to produce in bulk quantities in third world. 2. Experimental
In this study the seed of Spondias purpurea (sineguelas) as a
waste biomaterial was applied. S. purpurea is a species of flowering 2.1. Materials and analytical techniques
plant in the cashew family, Anacardiaceae, that is native to tropical
regions of the Americas (Fig. 1). It is most commonly known as All reagents (A.R.) were purchased from Merck or Aldrich and
Jocote, the other common names include Red Mombin, Purple were used without further purification, except that solvents were
Mombin, Hog Plum, Ciruela Huesito (Venezuela), and Sineguelas. treated according to the standard methods. The pH of solution
It was selected because of its high cellulose content. Additionally, was adjusted using 0.1 M HCl/NaOH using a pH meter (Metrohm,
sineguelas is a fruit that grows in tropical climate. The fruit is 827 pH Lab). The concentrations of phosphate solution were mea-
indigenous to Southern Mexico and the Northern part of Peru. sured by atomic absorption spectrophotometry using a Perkin–
Some of its species is believed also to originate from tropical parts Elmer 3030 instrument. The pH at the point of zero charge
of America. Sineguelas grows in shrub and does not actually (pHPZC) of different adsorbents was obtained by the solid addition
require delicate cultivation [4]. method [8]. IR spectra (Jasco FT/IR-680 plus spectrophotometer
One of the best ways of using biomass effectively for removal and KBr pellets) were used for the characterization of the adsor-
and desorption for recycling of contaminations from aqueous bents. XRD (Philips X’PERT MPD diffractometer) was performed
media is to be modified by inorganic compounds such as NZVI. for the crystalline structure of the adsorbents. The XRD patterns
NZVI is particularly attractive for separation purposes due to their were recorded in the 2h range of 10–100°. Transmission electron
remarkable surface area to weight ratio leading to a greater density microscopy (TEM) was carried out on the powder samples with a
of reactive sites and phosphate removal capacity. Moreover, the Tecnai F30 TEM operating at an accelerating voltage of 300 kV. In

Fig. 1. Sample of Spondias purpurea fruit (left) and interior schematic of the sineguelas seed (right).
M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77 71

addition, energy dispersive X-ray analysis was conducted on each solute concentration determined using a inductively coupled
sample. Nitrogen (99.999%) adsorption experiments have been plasma atomic emission spectroscopy (ICP-AES) using a
performed at 196 °C using a volumetric apparatus (Quanta chrome Shimadzu ARL 34000 instrument.
NOVA automated gas sorption analyzer). Before the adsorption The amount adsorbed was calculated as:
experiments, the sample is out-gassed at 120 °C for 16 h. The speci-
fic surface areas were calculated using BET method. Chemical anal- qe ¼ V  ðC 0  C e Þ=m ð1Þ
yses are measured by inductively coupled plasma atomic emission
where C0 and Ce are the initial and equilibrium liquid-phase concen-
spectroscopy (ICP-AES) using a Shimadzu ARL 34000 instrument
trations (mg L1) of adsorbates; V is the volume of the solution (L);
(Spectro flamed); typically, 30 mg of sample was dissolved in
and m is the amount of adsorbent (g). This equation assumes that
500 lL of 40% HF solution, 4 mL of 1:4 H2SO4 solution and 45 mL
the change in volume of the bulk liquid phase is negligible as the
H2O.
solute concentration is small and the volume occupied by the adsor-
bent is also small. The amount of the phosphate adsorbed on the
2.2. Preparation of the modified S. purpurea
sample was calculated using a previously determined calibration
curve.
S. purpurea fruit was obtained from a local store and was cut
and then its pit (seed) was removed. An easy way to do this is to
slice into the S. purpurea and cut it all around the seed to remove 3. Results and discussion
its flesh cleanly. After separating the seeds and their flesh, all seeds
include seed coats or testa, embryo and endosperm (Fig. 1) were The chemical compositions of the pristine and modified S by
put in the oven for 72 h in 70 °C, and after drying they were ground NaOH and NZVI were characterized by CHN and ICP. The S seed
and were passed through different sieve size. The fraction of parti- waste contains relevant amounts of plant nutrients like C, N, Ca,
cle between 250 and 400 lm (geometric mean size: 305 lm) was P, and Na (Table S1). In addition, the biomaterial contains a number
selected. Fresh S. purpurea seed waste was washed thoroughly with of elements that are vital to plants in small doses, i.e., micronutri-
hot distilled water and was dried at 65 °C. The sorbent thus ents, but also generally contain small amounts of undesirable
obtained was designated pristine S. purpurea (S). In the next step heavy metals (data not presented). CHN analysis of the S shows
S treated with base to optimize the sorption of metal ions. the high C content originates from the organic compounds adhe-
Pristine S was treated with 0.1 M NaOH solution at reflex for 2 h. sive the framework of pristine S seed waste. NaOH treatment
A typical experimental procedure is as follows: 25 g of the pristine removes organic compound as a change in the color of the extrac-
biomaterial is dispensed in 0.5 L of distilled water. Then a certain tion liquid can also be visualized (the extraction liquid is dirty
amount of 0.1 M NaOH is added and the suspension is subjected brown) during the modification reactions and increasing the inor-
to mechanical stirring for 2 h on heater. The final material is sepa- ganic phase (the carbon content of pristine S after NaOH treatment
rated by centrifugation and washes with distilled water. Excess of decreased from 48.8 to 30.5 wt.% and the calcium and phosphor
NaOH was removed with distilled water and the material was contents increased). Also, the NaOH treatment may induce nega-
dried at 50 °C. Alkaline treated of pristine S was designated as S- tively charged by converting functional groups into their Na salts,
NaOH. which increasing the active sites of the modified biomaterial (the
sodium content of S-NaOH after modification increased). The car-
2.3. Preparation of the biomaterial-supported NZVI (S-NaOH-NZVI bon content of S-NaOH-NZVI significantly decreased, which could
nanobiocomposite) be due to the removing of the organic compounds by alkali, which
resulted from the reaction of NaBH4 and H2O, during the prepara-
The NZVI-biomaterial sample is synthesized based on the fol- tion of NZVI. The result in Table S1 also indicates that the calcium
lowing procedure: FeCl24H2O was dissolved in a 4/1 (v/v) etha- phosphorous contents increased as the NZVI is immobilized, it is
nol/water mixture (72 mL ethanol + 18 mL deionized water), then believed that Fe ions are immobilized over the created active site
the S-NaOH is added to this solution and the mixture is left in an of the S-NaOH-NZVI (-Ca-ONa, and -P-ONa).
ultrasonic shaker for 25 min in order to disperse the biomaterial The surface morphology of the S-NaOH-NZVI was analyzed by
grains. Meanwhile, sodium borohydride solution is prepared by transmission electron microscope (TEM) (Fig. 2). The TEM image
dissolving NaBH4 in 100 mL of deionized water. The borohydride was recorded in backscatter mode to increase the contrast between
solution is then added drop wise to the aqueous Fe(II)–S-NaOH the nanoparticles and the biomaterial, i.e. the iron appears brighter
mixture while stirring continuously on a magnetic stirrer. Black than the support. NZVI with higher electron contrast on the
solid particles of NZVI appeared immediately following the addi- S-NaOH indicates the immobilization and distribution of nanoiron
tion of the first drop of NaBH4 solution. After the full addition of on the S-NaOH-NZVI surface without aggregation, thereby, NZVI
the borohydride solution, the mixture is left for a further 20 min was rough with a round shape, which was very different from bare
of stirring and then filtered. Immobilization of NZVI on the Fe(0), where chain-like structure was observed [9]. The coarse and
S-NaOH was designated as S-NaOH-NZVI. rough morphology of the NZVI could provide more reactive sites
than the smooth morphology. However, the NZVI particles (ca.
2.4. Adsorption measurements 5–70 nm) were clearly discrete and well dispersed on the modified
S, without aggregation even after aging time of 640 days. This con-
Adsorption experiments were carried out in batch conditions: firms why NZVI are commonly stabilized on support materials
0.05 g of modified biomaterial was shaken with 50 ml of the phos- such as resin, starch, clay and zeolite [10] because this strategy
phate solution, at a concentration of between 0.75 and 1000 mg/L, suppresses the aggregation of NZVI and increases mechanical
at a controlled temperature of 25 °C. Each experiment was per- strength. Therefore, S-NaOH could retain the NZVI particles from
formed three times and the averaged values were reported that aggregating together.
displayed a relative standard deviation lower than 1.65%. The time Zeta potentials of the pristine S and S-NaOH-NZVI (0.1 mg/ml)
required to reach equilibrium conditions was determined by pre- were measured in 103 M NaCl aqueous solution at different pHs.
liminary kinetic measurements. No significant variation in sorption The solution pH was adjusted by NaOH or HCl (Fig. S1). The pH
capacity was observed after 24 h of contact. After centrifugation at value of point of zero charge (pHpzc) of the pristine S is about 7.8
3000 r.p.m. for 5 min, the liquid phase was separated and the while after being modified with NaOH and NZVI, the pHpzc have
72 M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77

the alkaline range. A plausible explanation is that more oxygen


containing functional groups (the dominant active sites of
S-NaOH-NZVI are FeAOH, APAOH and „CaAOAH species) in the
S-NaOH-NZVI became deprotonated and negatively charged as
pH increased, consequently S-NaOH-NZVI uptakes lower P anions
due to the electrostatic repulsion interaction. At lower pH, the con-
centration of H+ ion is high, causing an electrostatic interaction
between the acid active sites ion and P anions (H2PO 4 , Fig. S2).
The active sites are closely associated with hydroxyl ions OH which
restricted the approach of P ions, due to the repulsive force, and so
fewer groups are approachable for the removal of P.
The effect of adsorption time (0–30 min) on the removal of P
(19, 45 and 150 mg L1) by S-NaOH-NZVI (0.05 g) at 25 °C, in a
solution with a pH of 4.0 was studied (Fig. 4A), where the most
of P was uptake within the first 5 min of contact time. The black
iron nanoscale loaded on S-NaOH-NZVI were instantly solubilized
(the system turned reddish-brown when interacts with oxygen),
that is, the S-NaOH-NZVI reacted at once with oxidative matters
in the bulk solution. Percentage of maximum removal of P ions
was 99.9%, 88.3% and 86.0% at initial concentration of 19, 45 and
150 mg L1, respectively, after 5 min. Indeed, the fast adsorption
during the initial time is attributed to the high concentration gra-
dient between the adsorbate in solution and that on the adsorbent
as there are a high number of vacant sites accessible through this
Fig. 2. TEM image of S-NaOH-NZVI after aging time of 640 days.
process, thereby the plateau after 7 min relates to a slow rate of
removal that related to the aggregation of P ions on the S-NaOH-
NZVI active sites. The removal of P ions from solution was com-
plete within 5 min. To optimize the removal process, the adsorp-
been shifted to 7.1 and 6.75, respectively, displaying that the NZVI tion isotherms for the rest of initial concentrations were studied
was successfully loaded on S-NaOH. Indeed, pure iron oxides typ- for a time of 5.0 min.
ically have zero point charges (ZPC) in the pH range 7–9 [9]. The empirically obtained kinetic sorption data were fitted to the
Therefore, the modified samples yield acidic surface since pHpzc pseudo-first-order, pseudo-second-order, and intra-particle
of them are lower than that of the pristine one and this surface
acidity is due to the introduction of several oxygen-containing
functional groups.
The effect of the pH on the removal capacity of S-NaOH-NZVI
were designed at initial concentrations of 70, 250 and 800 mg L1 A 140
and pH range between 2.0 and 12.0 (Fig. 3). The S-NaOH-NZVI dis- 130
Amount adsorbed, qe (mg/g)

120
plays higher P adsorption capacity in the acidic values of pHs, 110 19 mg/L
where no significant influence of pH between 2.0 and 6.0 were 100 45 mg/L
observed while the ionization state of P is pH dependent according 90
to Fig. S2 (pKa1 = 2.1, pKa2 = 7.2, and pKa3 = 12.3). As the initial pH 80 150 mg/L
70
increased 6.0–12.0, the adsorption efficiency sharply decreased 60
from 72.4% to 17.1%, respectively, at initial concentration of 50
70 mg/L. The lowest P adsorption capacity of S-NaOH-NZVI 40
30
(17.1%) was found at an initial pH solution of 12.0, which was
20
caused by the role of deprotonation of the hydroxyl species on 10
the surface of iron nanoparticles at this solution pH. The removal 0
efficiency of P is higher in the acidic pH range of 2–4.0 than in 0 10 20 30 40 50 60 70 80 90 100 110 120
t (min)

B 6.5 19 mg/L
300 70 mg/L 6 45 mg/L
280 5.5 150 mg/L
Amount adsorbed, qe (mg/g)

250 mg/L
260 5
800 mg/L
t/qt (min g mg-1)

240 4.5
220 4
200
180 3.5
160 3
140 2.5
120 2
100 1.5
80
1
60
40 0.5
20 0
0 0 10 20 30 40 50 60 70 80 90 100 110 120
1 2 3 4 5 6 7 8 9 10 11 12 t (min)
pH
Fig. 4. The adsorption kinetics (A) and pseudo-second order plot (B) of the S-NaOH-
Fig. 3. The effect of pH on the adsorption of P by S-NaOH-NZVI. NZVI for the adsorption of P at room temperature.
M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77 73

diffusion models (Table S2) in order to determine and interpret the Table 2
mechanisms and major parameters governing sorption kinetics of Fitting of the parameters of the experimental results to the Langmuir and Freundlich
equation parameters.
P ions uptake processes over S-NaOH-NZVI. The obtained kinetics
parameters for removal of P ions on the S-NaOH-NZVI, at various qm KL KF n R2 Sorption model
initial concentrations are gathered in Table 1. The pseudo-sec- S-NaOH-NZVI
ond-order equation seem to be the best-fitting model than those 1719 1.73  103 0.9910 Langmuir
for the other two equations (the correlation coefficient is extre- 0.451 0.292 0.9621 Freundlich

mely high for the pseudo-second-order equation of S-NaOH- Spondias purpurea seed
NZVI; R2 > 0.999). The value of qe,cal also displayed to be very close 4.48 2.17  102 0.9438 Langmuir
1.46  102 2.189 0.3715 Freundlich
to the experimentally observed value of qe,exp. The plot of linear
form of the pseudo-second-order for the uptake of P ions was
depicted Fig. 4B.
The adsorption equilibrium data for P on the bioadsorbents
The correspondence of the experimental data with the pseudo-
were fitted by applying the Langmuir and Freundlich isotherm
second-order kinetic model (the pseudo-second order equation is
models (Table S2) that are common models utilized for aqueous-
based on the removal capacity on the solid phase) displays that
phase adsorption. These adsorption models give a representation
the adsorption of P by S-NaOH-NZVI performed by chemisorption of the adsorption equilibrium between an adsorbate in solution
(as the rate-limiting step of the removal mechanism and no partic-
and the active sites of the adsorbent.
ipation of a mass transfer in solution) including valence forces The Langmuir equation relates the coverage of molecules on a
through sharing or exchange electrons between adsorbent and
solid surface to the concentration of a medium above the solid sur-
adsorbate. The removal of the P on the supported NZVI could be face at a fixed temperature and adsorption is limited to monolayer
composed of two processes with initial removal rate of 126.5,
coverage, and intermolecular forces decrease with the distance
80.6 and 417 mg (g min)1 for 19, 45 and 150 mg/L, respectively from the adsorption surface. The Freundlich model supposes that
(Table 1) (the adsorption rate was related to the content and type
the adsorption surface is heterogeneous, that interactions between
of active adsorption site on the matrix of adsorbent). More advan- adsorbed molecules can occur, and that multilayer adsorption is
tage of the pseudo-second-order model is that it anticipates the
possible. The Langmuir and Freundlich adsorption isotherms exhi-
behavior over the all ranges of the removal process. bit an approximately linear relationship for used adsorbent (see
The effect of various initial P ion concentrations (0.75–
Table 2 and Fig. 5A). The results acquired from the studied systems
1000 mg L1) over the S and S-NaOH-NZVI are shown in Fig. S3,
showed that Langmuir isotherm correlated better (R2 = 0.943) than
which demonstrated that S-NaOH-NZVI was the most efficient
Freundlich one.
adsorbent in that more than 80% of P was removed after 5 min,
Also, the equilibrium removal efficiency of P on the S-NaOH-
while only 3% of P was removed by S for initial concentration of
NZVI was studied at diverse temperatures (15, 25, 60 and 80 °C)
110 mg L1. The adsorption was initially fast (i.e., between 0.75
in pH = 4.0 (Fig. S4). The enhancement in the temperature of P
and 250 mg L1), decreased progressively and finally approached
solutions from 15 to 80 °C results an increase in the removal poten-
equilibrium.
tial of the S-NaOH-NZVI. Therefore, adsorption of the P on active
When the initial P concentrations were increased in the pres-
sites of the S-NaOH-NZVI is endothermic and could be assessed
ence of S-NaOH-NZVI, the adsorption of the P ions reached to
by obtainable of more active sites of adsorbent, the expansion
91.4%, 91% and 46% for initial P concentrations of 22.5, 110.0 and
and activation of the adsorbent surface at higher temperatures.
550 mg L1, respectively. However, NZVI had poor removal effi-
That results in the easy mobility of P ions from the bulk solution
ciency when it was used to remove P with a concentration greater
toward the adsorbent surface and increased the attainable to the
than 100 mg L1, which may have been caused by the aggregation
S-NaOH-NZVI active sites.
of NZVI decreasing its specific surface area and reaction activity.
The thermodynamics parameters related to the adsorption of P
Immobilized NZVI on the S is a well dispersed and stable material
ions on S-NaOH-NZVI, namely the changes in Gibbs free energy
in aqueous solution that increasing the activity of NZVI. The rela-
(DG°), enthalpy (DH°) and entropy (DS°) where therefore calcu-
tive increase in the loading capacities of sorbents with increasing
lated using the following equation:
P concentrations is attributed to the interaction between the P
and NZVI which enhancing the vital driving forces to defeat the DG ¼ RT ln K ð2Þ
resistances to the mass transfer of P ions between the bulk solution
where R (8.314 J/mol K) is the ideal gas constant, T (K) is the tem-
and S-NaOH-NZVI. The increase of P uptake with rising the initial P
perature and K (L/g) is the distribution coefficient of the adsorbate
ions concentration may be attributed to an increase in electrostatic
(qe/C), and the van’t Hoff equation:
interactions (relative to covalent interactions), which includes
active sites of lower affinity for P up to saturation point. While, ln K ¼ DS =R  DH =RT ð3Þ
at low P concentration the removal efficiency of S-NaOH-NZVI
was high which may be related to the high concentration of initial The DH° and DS° values were obtained from slope and intercept of
mole numbers of P to the accessible active sites on the surface of the linear plot of ln K vs 1/T (Fig. 5B). The results of thermodynamic
the adsorbent. parameters are shown in Table 3. The results show a spontaneous
and favorable removal process over the whole temperature range

Table 1
Kinetic parameters for the adsorption of P ions by S-NaOH-NZVI at different initial concentrations.

C0 (mg L1) qe,exp (mg g1) Pseudo first order constants Pseudo second order constants Intra-particle diffusion
constants
k1 (min1) q1 (mg g1) R2 k2 (g(mg min)1) q2 (mg g1) h (mg/(g min)) R2 Kint (mg/(g min1/2)) R2
19 19 0.381 56.17 0.6756 0.349 19.04 126.5 1.0 1.015 0.3157
45 40 0.350 137 0.6332 0.040 45.0 80.6 0.9999 2.489 0.3908
150 129 0.415 370 0.7017 0.023 134 417 1.0 7.238 0.3241
74 M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77

1000
The interference of anions in real water may affect the removal
A of P ions from aqueous solution through the block of the active
adsorption sites or the complexation with of P ions. Therefore, five
800 typical anion species (chloride, nitrate, bicarbonate, sulfate and
citrate) at concentrations of 0.01 and 0.1 M were used to assess
their effect on P removal by S-NaOH-NZVI at concentration of
600
q e (mg/g)

S-NaOH-NZVI 150 mg L1 (Fig. 6). In the absence of any competitive anions, the
S percent of P removal by S-NaOH-NZVI is as high as 96.4%. The
400 Langmuir results depicted in Fig. 6 confirmed that only small interference
Freundlich
for the P removal was observed at 0.01 M sodium chloride, sodium
Langmuir
sulfate and Sodium citrate. Dissolved sodium bicarbonate com-
200 petes strongly with phosphate at 0.1 M, decreasing the uptake rate
of P from aqueous solution. It can be concluded that S-NaOH-NZVI
0 keeps its reactivity in the presence of the used interferences, while
NaHCO3 > Na3C6H5O7 > NaNO3 display the most preventing effect
0 200 400 600 800 on the removal of P ions. It is believed that metalated adsorbent
Equilibrium concentration, C e (mg/L) can bind the anions of higher base strength more strongly and lar-
gely than those with lower base strength. Moreover, the used salts
5.5 70 mg/L could create the effect of electric double layer compression on the
B surface of S-NaOH-NZVI which blocking the active sites on the sur-
4.5 250 mg/L faces, and lead to the freedom of adsorbed P. The decreased of
removal capacity of S-NaOH-NZVI toward P ions in the presence
800 mg/L
3.5 of NaHCO3 implied that NZVI choose to bind with HCO 3 and
decrease affinity of the active sites of S-NaOH-NZVI toward P,
2.5 which eventually reducing the removal of P ion from the bulk solu-
ln (K)

tion. Nevertheless, the initial pH of the solution with 0.1 M NaHCO3


1.5
is 8.3, under which a large amount of OH is present, that com-
0.5
petes with P ions for the attractive interaction with active sites
of loaded NZVI and blocks the ligand-exchange mechanism, so
-0.5 reducing P ions removal from aqueous solution.
To determine the reusability, the used S-NaOH-NZVI was sepa-
-1.5 rated by filtration after the first run and for extraction of adsorbed
2.75 2.85 2.95 3.05 3.15 3.25 3.35 3.45 3.55 P ions various concentrations of NaOH was tested to regenerate P
-3 ion binding ability of the S-NaOH-NZVI and finally dried at 80 °C
(1/T) × 10
under vacuum that then used for the next runs under the same
Fig. 5. Equilibrium absorption of P by the pristine sineguela and S-NaOH-NZVI at conditions. Increasing ionic strength of solution significantly
25 °C and pH = 4.0 (A), and plot of ln K vs. 1/T for the P adsorption on S-NaOH-NZVI
increases desorption of P ions (Fig. 7A). At the NaOH concentration
(B).
of 0.1 M, 79.4% of P was desorbed, while that in the lower concen-
tration of NaOH was 4.5%. It was observed that the efficiency of
desorption process was generally above 98% at NaOH = 1.0 M,
Table 3
and the activity of adsorbent was almost not affected, implying
Thermodynamic parameters for the adsorption of P ions on S-NaOH-NZVI as a
function of temperature. that the S-NaOH-NZVI has vigorous solidity even after 12th runs

Initial P conc. DH° DS° DG° (kJ/mol)


(mg/L) (J/mol) (J/mol K)
258 298 333 353
100 0.1 M
70 45.12 162.5 41.98 48.49 54.18 57.43
95 0.01 M
250 32.49 114.9 29.69 34.28 38.31 40.61 90
800 6.634 16.43 4.245 4.902 5.477 5.805 85
1000 7.740 17.91 4.629 5.345 5.972 6.330 80
75
% Removal of P

70
65
60
55
(DG < 0). The standard enthalpy (DH) for the removal process is 50
45
positive, thus implying that the process is endothermic in nature. 40
The necessity of a large amount of heat to uptake the P ions from 35
30
the solution make the removal process endothermic. The positive 25
20
value of DS implied the increased randomness and an enhancement 15
in the degrees of freedom at the solid-solution interface through the 10
5
removal of the P ions on the active site of the adsorbent which 0
showed the partial dehydration of the P ions before adsorption, thus
raising the spontaneity. This is also in agreement by the positive
value of DH, when the positive value of the standard enthalpy alters
for P ions sorption displays endodermic nature of adsorption. It
could be seen that with enhancement in temperature the value of
DG reduces, which indicated that removal processes of P over
loaded NZVI was spontaneous and thermodynamically favorable
Fig. 6. The effect of chloride, nitrate, bicarbonate, sulfate and citrate on the removal
by a raise in the solution temperature (Table 3). of P by S-NaOH-NZVI; pH = 4.0, contact time = 24 h and T = 25 °C.
M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77 75

A 100 105
100
90

% Removal of P from the Persian Gulf water


95
% Desorption of P

80 90
70 85
80
60
75
50 70
40 65
60
30
55 150 mg/L, 0.5 g
20 50
150 mg/L, 0.25g
10 45
40 19 mg/L, 0.25 g
0
0.001 0.005 0.01 0.05 0.1 0.5 1 35
30
NaOH M 25
20
15
B 90
10
80 5
70 0
0 30 60 90 120 150 180 210 240 270 300 330 360
% Removal of P

60
Time (min)
50
40 Fig. 8. P removal from the Persian Gulf water after exposure to S-NaOH-NZVI for
varying times and adsorbent dosages.
30
20
10
(Fig. 9). S-NaOH-NZVI contains relevant amounts of plant nutrients
like Na, S, Ca, P and B + Fe (caused by NZVI and NaBH4) (Fig. 9A). A
0
1 2 3 4 5 6 7 8 9 10 11 12 typical EDX spectrum recorded for P-loaded S-NaOH-NZVI,
Cycle exposed to an initial concentration of 150 mg L1 at pH 4.0, is
shown in Fig. 9B. EDX analysis of S-NaOH-NZVI@P displayed that
Fig. 7. The effect of chemical reagent on desorption of P from S-NaOH-NZVI (A) and
adsorption capacity of the S-NaOH-NZVI after repeated regeneration (B), initial
concentration of P = 150 mg L1 and T = 25 °C.

(Fig. 7B). It is assumed that the modified S. purpurea seed waste, as


the host of NZVI, provides an environment that decreases the
aggregation of adsorbed P ions on the NZVI surface; consequently
this will retain the reactive sites on the loaded NZVI surface after
several runs. The high adsorption of P ions by the S-NaOH-NZVI
could be due to the large amount of nanoparticles (NZVI) on its sur-
face that has a strong ability to bind with P ions in aqueous media
[11,12]. Indeed, precipitation of Fe3 (PO4)28H2O onto the S-NaOH-
NZVI could block the active sites of nano-sized iron particles, so
decreasing the phosphate adsorption capacity of the immobilized
NZVI.
The suggested adsorbent in this paper was also implemented to
remove P from the Persian Gulf water as a function of time (Fig. 8).
The Persian Gulf itself is a shallow embayment of the northeastern
Arabian Sea. It is separated from the Gulf of Oman by the Strait of
Hormuz which is only 56 km wide. The surface area of the Persian
Gulf is approximately 2.39  105 km2 and a mean depth of 36 m
implies an average volume of 8.63  103 km3 [13]. The Persian
Gulf water gathered from the cost of Bushehr and was filtered using
0.45 lm filter membrane and then they were stored in polyethy-
lene bottles. The concentration of P in the real water was recorded
by atomic absorption spectrophotometric (AAS). The removal pro-
cess had been done by spiking known amount of standards (19 and
150 mg L1 P ions) in the sample matrix and then analyzing it
(Fig. 8). The removal efficiency of P from the Persian Gulf water
reached to 99.9% and 85% for 30 and 150 mg L1 P, respectively,
after 10 min. However, when the adsorbent dosage was increased
to 0.5 g adsorption capacity reached to 99.8% for 150 mg L1 P after
15 min. The results confirm that S-NaOH-NZVI has reliable and
potent potential to uptake P ions from the real water.
Energy dispersive X-ray (EDX) analysis was recorded to indicate Fig. 9. Energy dispersive absorption spectroscopy photograph of S-NaOH-NZVI
the elemental content of S-NaOH-NZVI and S-NaOH-NZVI@P before (A) and after (B) adsorption of P from aqueous solution.
76 M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77

the signal of P was stronger at about 1.95 keV which is similar to


the P ion signals observed at the same energy in the literature
[12,14]. The EDX spectrum provides the spectroscopic evidence
for the presence of P ions on the surface of the S-NaOH-NZVI,
thereby, confirming the adsorption of the P ions on the S-NaOH-
NZVI surface.
FTIR spectra of S-NaOH-NZVI and P-loaded S-NaOH-NZVI sam-
ple, before and after the adsorption process, were recorded in the
range of 400–4000 cm1 (Fig. 10). In the equilibrated sample of
the S-NaOH-NZVI with P solution, extra bands at 524 (the bend
vibration of OAPAO), 669, 1039 (bending vibration of adsorbed
phosphate PAO) and 1053 cm1 (the asymmetric stretch vibration
of PAO) were appeared confirming the presence of P anchored to
the active sites of the immobilized NZVI [15]. Furthermore, the
peak at 1357 cm1 (OAH bending vibration) displayed the exis-
tence of hydroxyls groups on the surface of immobilized NZVI,
which after removal of P the hydroxyl groups were covered by Fig. 11. Wide-range X-ray diffraction of the P loaded S-NaOH-NZVI.
PO3
4 , and this peak was shifted to 1375 cm
1
and weakened
[16]. However, the bands of 437 and 497 cm1 observed in samples
were attributed to metal oxygen stretching vibration [16]. The new
peaks appeared after phosphate adsorption onto S-NaOH-NZVI, and 34.3° corresponded to vivianite. However, Fig. 11 shows the
probably showing the presence of inner-sphere surface complex loaded NZVI oxidation after reaction with P, where the apparent
(FeAOAP) between phosphate and the active sites of NZVI. peaks at the 2h of 36.2°, 39.8° 55.6° and 44.9° indicate the presence
In order to better understanding of mechanism and interaction of magnetite/maghemite (Fe3O4/c-Fe2O3) and Fe(0), respectively
between the P ions and S-NaOH-NZVI, XRD pattern of the S-NaOH- [19]. Indeed, magnetite (a conductive iron oxide) [20], can give a
NZVI was recorded before and after shaking with the P solution conductive environment between the reactive surface species in
(Fig. 11). XRD analysis confirms the presence of P on S-NaOH- solution and the inner electron-rich core, where the existence of
NZVI@P. The XRD pattern of S-NaOH-NZVI was compared with maghemite or hematite at the surface, both less conductive than
the XRD patterns attained from standard materials, to characterize magnetite could potentially reduce the reactivity and oxidation
the apparent peaks attributable to different iron compounds. The rate of NZVI. Thus, the peak of Fe(0) could still be found and the
iron nanoparticles consisted mostly of iron (reflection at 2h of peak intensity for Fe(III) and Fe(II) increased, showing that Fe(0)
44.98°), and no signals for iron oxides with a 2h value of 36° were had not been oxidized completely on the surface of S-NaOH-NZVI.
found [6,7,17]. This indicated that the iron present on the modified Beside the efficient activity and nature-friendly aspect, eco-
S surface is mainly in its zerovalent state. Iron oxide signals were nomic cost can also determine whether a new system can be used
not detected in the XRD patterns of the freshly prepared samples to the environment or not. Therefore, we have collected the costs of
and even after six months which demonstrated that the S-NaOH several bio materials which could be used as the support. On the
could be a good stabilizer of NZVI. In fact, when the particle size fruit market there is a lot of activities [21]. Some fruit prices in pri-
of the NZVI is smaller than 5 nm, the diffraction peaks are signifi- mary receiving markets of Manila have mentioned at Table S3. As
cantly broadened and the intensity decreases The powder XRD pat- the comparison of fruit prices shows, S has the lowest price in
tern was complicated by the presence of iron oxides and compare with the other fruit, so it would be more reasonable eco-
oxyhydroxides while showed a progression of peaks consistent nomically to use S seed waste and apply it as the bio adsorbent.
with Fe3 (PO4)28H2O (vivianite) [18]. Thereby, 2h at 23.0°, 30.0° Thus S-NaOH-NZVI is inexpensive and likely easy to prepare in

1375 669
1735
%T 1053 1037

B 1357
524
493

437

4000 3000 2000 1000 400


Wavenumber [cm-1]

Fig. 10. FTIR spectra of S-NaOH-NZVI before (A) and after (B) adsorption of P from aqueous solution (P = 150 mg/L, S-NaOH-NZVI = 0.05 g).
M. Arshadi et al. / Journal of Colloid and Interface Science 452 (2015) 69–77 77

large quantities, so proffering the opportunity to use this biomate- Appendix A. Supplementary material
rial in third world drinking water filtration in order to alleviate the
poisoning and suffering of millions of people worldwide. Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.jcis.2015.04.019.

4. Conclusion References

[1] M.C. Molles, Ecology: Concepts and Applications, second ed., The McGraw Hill
The synthesis and characterization of NZVI loaded on Inc., 2002.
S. purpurea seed waste (S-NaOH-NZVI) was accomplished for the [2] D. Cordell, J.O. Drangert, S. White, Global Environ. Change 19 (2009) 292–305.
removal of phosphate ions from waste waters. The adsorption of [3] L.E. de Bashan, Y. Bashan, Water Res. 38 (2004) 4222–4246.
[4] R. Boning Charles, Florida’s Best Fruiting Plants: Native and Exotic Trees,
phosphate ions has been studied in terms of pseudo-first- and Shrubs, and Vines, Pineapple Press, Inc., Sarasota, Florida, 2006.
-second-order kinetics, and the Freundlich, and Langmuir iso- [5] W.L. Yan, H.L. Lien, B.E. Koel, W.X. Zhang, Environ. Sci.: Process. Impacts 15
therms models have also been used to the equilibrium adsorption (2013) 63–77.
[6] R.A. Crane, T.B. Scott, J. Hazard. Mater. 211–212 (2012) 112–125.
data. The adsorption kinetics followed the mechanism of the
[7] M. Arshadi, M. Soleymanzadeh, J.W.L. Salvacion, F. SalimiVahid, J. Colloid
pseudo-second-order equation. The thermodynamic parameters Interface Sci. 426 (2014) 241–251.
(DG, DH and DS) indicated that the adsorption of phosphate ions [8] R. Cason, W.R. Lester, Proc. Okla. Acad. Sci. 57 (1977) 116–118.
were feasible, spontaneous and endothermic at 25–80 °C. No [9] Yuan-Pang Sun, Xiao-qin Li, Jiasheng Cao, Wei-xian Zhang, H. Paul Wang, Adv.
Colloid Interface Sci. 120 (2006) 47–56.
significant loss of activity was observed, confirming that the [10] R.A. Crane, T.B. Scott, J. Hazard. Mater. 211–212 (2012) 112–125.
S-NaOH-NZVI has high stability during the adsorption process even [11] V.A. Chernyakhovskii, Refractories 26 (1–2) (1985) 41–44.
after 12th runs The suggested adsorbent in this paper was also [12] Y. Yao, B. Gao, M. Inyang, A.R. Zimmerman, X. Cao, P. Pullammanappallil, L.
Yang, Bioresour. Technol. 102 (2011) 6273–6278.
implemented to remove P from the Persian Gulf water. XRD, FTIR [13] Hosny Ibrahim Emara, J. Persian Gulf 1 (2011) 33–44.
and EDX analysis indicated the presence of Fe3 (PO4)28H2O [14] H. Yin, M. Kong, Desalination 351 (2014) 128–137.
(vivianite) on the S-NaOH-NZVI@P surface. Thus S-NaOH-NZVI is [15] D.S. Soejoko, M.O. Tjia, J. Mater. Sci. 38 (2003) 2087–2093.
[16] S.K. Apte, S.D. Naik, R.S. Sonawane, B.B. Kale, J. Am. Ceram. Soc. 90 (2007) 412–
inexpensive and likely easy to prepare in large quantities, so prof- 414.
fering the opportunity to use this biomaterial in third world drink- [17] M. Soleymanzadeh, M. Arshadi, J.W.L. Salvacion, F. SalimiVahid, Chem. Eng.
ing water filtration in order to alleviate the poisoning and suffering Res. Des. 93 (2015) 696–709.
[18] B.C. Reinsch, B. Forsberg, R.L. Penn, C.S. Kim, G.V. Lowry, Environ. Sci. Technol.
of millions of people worldwide. 44 (2010) 3455–3461.
[19] H. Kim, H.J. Hong, Y.J. Lee, H.J. Shin, J.W. Yang, Desalination 223 (2008) 212–
220.
[20] R.M Cornell, U. Schwertmann, The Iron Oxides: Structure, Properties,
Acknowledgment Reactions, Occurrences and Uses, second ed., Wiley-VCH: Weinheim,
Germany, 2003 (pp. 129, 235, 506–508, 537–538, 506–508).
Thanks are due to the Iranian Nanotechnology Initiative for sup- [21] http://www.globalprice.info/en/ Philippines/the-cost-of-fruits-Philippines.
porting of this work.

You might also like