Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Environmental Chemical Engineering 7 (2019) 103430

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Adsorption of Sr(II) cations onto phosphated mesoporous titanium dioxide: T


Mechanism, isotherm and kinetics studies
Ivan Mironyuka, Tetiana Tatarchukb,c,⁎, Hanna Vasylyevad, Mu. Naushade, Igor Mykytyna
a
Department of Chemistry, Vasyl Stefanyk Precarpathian National University, 57 Shevchenko Street, 76018, Ivano-Frankivsk, Ukraine
b
Educational and Scientific Center of Material Science and Nanotechnology, Vasyl Stefanyk Precarpathian National University, Ivano-Frankivsk, 76018, Ukraine
c
Faculty of Chemical Technology and Engineering, UTP University of Science and Technology, 3, Seminaryjna str., 85-326, Bydgoszcz, Poland
d
Uzhhorod National University, 3 Narodna Square, 88000, Uzhhorod, Ukraine
e
Department of Chemistry, College of Science, Building #5, King Saud University, Riyadh, 11451, Saudi Arabia

ARTICLE INFO ABSTRACT

Keywords: In this work, phosphate-modified titanium dioxide nanoparticles with mesoporous structure were synthesized
Titania and utilised as an efficient nanoadsorbent for Sr(II) removal from aqueous environment. The globular titania
Sr(II) particles with a 3–4 nm diameter with chemosorbed on their surfaces phosphate groups = Ti(O2POOH) have
Adsorption been obtained by liquid-phase synthesis. The characterization of titania nanoparticles were performed using
Phosphate
XRD, TEM, IR-spectroscopy and Brunauer-Emmett-Teller analyses. The synthesized modified adsorbents ex-
Radionuclide
hibited SBET = 396–410 m2 g−1 and mesopore volume of 0.262 – 0.275сm3 g−1. The isotherm and kinetic
Aqueous environment
models were used for the description of Sr(II) adsorption. The sample 4P-TiO2 showed the best adsorption ability
for strontium ions removal. The adsorption capacity of Sr(II) onto modified 2P-TiO2, 4P-TiO2 and 8P-TiO2
samples was found to be 94.1, 172.5, 128.9 mg/g, respectively. An adsorption mechanism for Sr(II) removal was
proposed. The adsorption/desorption studies were conducted to find the reusability of phosphated adsorbents. A
titanium dioxide-based mesoporous material possesses a significant adsorption activity due to high content of
surface acid sites. All the outcomes revealed that the phosphate modified titania showed great potential in Sr(II)
removal from aqueous environment and nuclear effluents.

1. Introduction with atmospheric precipitations. Heavy metals are oxidized in the soils;
washed out by rainwater and fall into the reservoirs. Different treat-
Water is an essential and important substance on Earth, because it ment techniques like coagulation, precipitation, ultrafiltration, filtra-
provides the vital functions of living organisms. However, due to nat- tion and adsorption have been used for the removal of toxic pollutants
ural and technogenic impacts, the quality of our water resources is [8–10]. Among them, the adsorption technology has some more bene-
constantly deteriorating because of the accumulation of inorganic and fits – simplicity, accessibility and efficiency [11].
organic harmful substances [1,2]. Particularly serious danger to the Strontium-90 originates from nuclear power plants and is well
biosphere causes water pollution by toxic heavy metals such as stron- known hazardous element for the environment due to its high harmful
tium(II), lead(II), cadmium(II), zinc(II), cupper(II), as well as arsenate, effect and radioactivity [12]. The accumulation of strontium in the
selenate, and fluoride anions [3–6]. Excessive presence of these pollu- body leads to the destruction of the whole body (general toxic effect).
tants in the water can causes carcinogenic, mutagenic or teratogenic The most typical form of this disease is the development of dystrophic
effects on living organisms. changes in the osteo-articular system during the period of body growth
The main origins of environmental pollution are thermal power (there is a lag in growth, exhaustion, baldness, etc.). This disease is
plants, ferrous metallurgy enterprises, non-ferrous metallurgy en- accompanied also by a pronounced disturbance of the calcium-phos-
terprises, and chemical industry. The pollutants are accumulating in the phorus ratio in the blood, intestinal dysbiosis. In the presence of
earth near the enterprises of ferrous and nonferrous metallurgy [7]. strontium, iodine becomes inaccessible to the human organism; as a
More than 95% of them entering into the soils as technogenic dust or result, the internal iodine deficiency occurs. In this regard, the problem

Corresponding author at: Educational and Scientific Center of Material Science and Nanotechnology, Vasyl Stefanyk Precarpathian National University, Ivano-

Frankivsk, 76018, Ukraine.


E-mail addresses: myrif555@gmail.com (I. Mironyuk), tetiana.tatarchuk@pu.if.ua, tatarchuk.tetyana@gmail.com (T. Tatarchuk),
h.v.vasylyeva@hotmail.com (H. Vasylyeva), mnaushad@ksu.edu.sa (M. Naushad).

https://doi.org/10.1016/j.jece.2019.103430
Received 10 July 2019; Received in revised form 14 September 2019; Accepted 20 September 2019
Available online 29 September 2019
2213-3437/ © 2019 Elsevier Ltd. All rights reserved.
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

of Sr(II) removal from water is very important for Ukraine and other 2.3. Adsorption of Sr(II)
counties, which had accidents at nuclear power plants [13].
The different types of sorbents have been intensively investigated The adsorbed amount of Sr(II) was estimated from the difference
for toxic pollutants removal from aquatic environment such as sand, between the initial concentration of Sr(II) (from 0.001 to 0.1 mol·L−1
clays, kaolin, dolomite, etc. However, such adsorbents have low effi- SrCl2) and the equilibrium concentration of Sr(II) after its contact with
ciency, since the adsorption process is very slow and their adsorption the adsorbent. The adsorbent with certain mass was added to a model
capacity is small [14–16]. In addition, the ZrO2 microspheres [17], solution of metal at 293 K, after 4 h it was centrifuged and the equili-
magnetic graphene oxide [18], activated carbon [19], titanate nano- brium concentration of Sr(II) (Ce) was determined using a com-
tubes [20], TiO2 nanoparticles [21], hydroxyapatite [22], niobate na- plexometry [33] by the following equation:
nofibers [23] have also been used for Sr(II) adsorption. In could be seen
(Ci (Sr ) Ce (Sr ))* V (solution)
that the titanate adsorbents possess much better adsorption capacity qe =
mads (1)
towards strontium ions compared to the known adsorbents. It was
noted that the adsorptive properties of titanante nanomaterials depend −1
where Ci(Sr) and Ce(Sr) (mg·L ) are initial and equilibrium con-
on the synthesis methods. The controlled synthesis creates the oppor- centration of Sr(II) respectively, V(solution) (L) is volume of model
tunities for designing the micro- and mesoporous adsorbents with de- solution, and mads (g) is mass of the titania adsorbent.
sired surface, morphology and structural properties which increase Kinetics of Sr(II) adsorption onto raw and phosphated titania was
their activity in various environmental processes. studied at pH: 7.0 and temperature: 298 K in the solution with Co(Sr)
It was shown that titanium(IV)-based compounds had a capability to =0.01 mol·L−1; m(ads) = 0.05 g and contact time 140 min. At certain
adsorb heavy metals [24–27], and the surface modification increased intervals, the solution was separated from the adsorbent and analyzed
their adsorption properties [28,29]. For example, according to the [25], for strontium content. All the data were the average values of three
the layered titanium phosphate was investigated as promising ad- parallel adsorption tests.
sorbent due to chemosorbed phosphate groups located at its surface, In addition, the point of zero charge (pHPZC) of titania adsorbents
which can bind the metal cations in the aqueous environment. How- was also measured by drift method [34] and the adsorption/desorption
ever, its adsorption capacity was low due to small porosity and low experiments were performed to estimate the reusability of titania ad-
surface area. Earlier our scientific group proposed a new way to obtain sorbents.
mesoporous titania [30] with chemisorbed orthophosphate groups on
its surface. The high porosity of the material and the existence of
3. Results and discussion
chemisorbed phosphate groups on the pores surface could substantially
increase its ability to bind Sr(II) cations, as well as cations of other
3.1. Modification of titania adsorbent
heavy metals. The promising phosphatized adsorbents for the removal
of Sr(II) ions in water treatment were also reported by Ivanets et al.
The titanium aquacomplex [Ti(H2O)6]3+·3Cl− as a precursor allows
[31], which were prepared by the soft non-acidic method using ther-
to obtain a titania nanoparticles with desired physic-chemical proper-
mally activated dolomite and NaH2PO4,Na2HPO4 and Na3PO4 as
ties by liquid-phase route [30]. The NaOH addition leads to the for-
phosphating reagents.
mation of molecules Ti(OH)4·2H2O according to the scheme (2):
Therefore, the aim of this work was to obtain mesoporous titania
with chemosorbed phosphate groups and to investigate their impact on
the sorption ability to eliminate Sr(II) from aquatic environment. The
mesoporous TiO2 was obtained by the liquid-phase route using a tita-
nium aquacomplex and a modifying reagent Na3PO4 as a precursors.
(2)
The introduction of Na3PO4 modifier into the aquacomplex pre-
2. Experimental cursor changes the reaction equilibrium and causes the oxidation of
Ti3+ cations. The anions OH−, PO43- form intermediate complex [Ti
2.1. Synthesis of unmodified and modified titania adsorbent (O2POOH)(OH2)5]3+·3Cl−. After NaOH addition, the pH is increased
and intermediate complexes are converted into the molecules Ti
The titanium aquacomplex precursor [Tі(H2O)6]3+·3Cl– was ob- (O2POOH)3(OH)3·2H2O, which become as centers of origin and growth
tained as described in [32]. The modification of TiO2 was performed by of primary oxide particles (Eq. 3):
sodium orthophosphate (2, 4 and 8% (wt.) of Na3PO4). The NaOH was
added in a drop wise manner to obtain the pH = 1–2 at temperature
70 °C during 30–40 min. After that, the pH of the solutions was in-
creased to around 6-7. The obtained nanoparticles were detached from
solutions by the vacuum filtration, washed and dried at temperature
120–140 °C.
(3)
The peculiarity of the titania particles growth process is that
2.2. Characterization methods the = Ti(O2POOH) groups are focused on the primary particles surface.
Thereafter, the condensation process is changed in such a way that the
The phase analysis was performed by XRD (STOE STADI P, Cukα surface with chemosorbed phosphate groups forms the interparticle
anode (λ =0.154 nm)). The calculation of crystallite sizes was carried pores between the particles, increase the surface area and pore volume
out by methods of integral width of diffraction reflexes with the help of in the titania adsorbents.
WinPLOTR software. The structural and adsorption characteristics were
carried out using BET-surface area analyzer (Quantachrome Autosorb, 3.2. Characterization
Nova 2200e, at 77 K). The IR-spectra were recorded using SPECORD
M80 spectrophotometer in the range of 4000-300 cm−1. Transmission 3.2.1. X-ray diffraction analysis
electron microscopy (TEM) micrographs were performed using the The diffractograms of the TiO2 samples are depicted in Fig. 1 and
JEM-100 CX II and JSM 2100 F microscopes. proved that the samples of the unmodified titania (TiO2) and the

2
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

(Fig. 2b). The drying of the water dispersion lets to grow up of spherical
aggregates and formed a porous xerogel product. It could be seen that
most of particles of the xerogel were crystallographically disoriented
(Fig. 2c).

3.2.3. BET analysis


Fig. 3 depicts the N2 adsorption/desorption isotherms by TiO2
samples and the associated parameters given in Table 2. The pore size
distributions are depicted in Fig. 4. The textural characteristics of the
adsorbents porous structure indicated that the surface-chemical mod-
ification of samples resulted in a rise in their specific surface area,
volume of pores, mostly in the mesopores volume (Vmeso), and the
samples possess micro/mesoporous structure.
The Table 2 exhibits that the specific surface area of mesopores
Smeso for 2P-TiO2, 4P-TiO2, 8P-TiO2 was 354, 368, and 378 m2 g−1,
respectively, which was 2.55, 2.65 and 2.72 times greater than that of
raw titania. The Vmeso for modified samples was 0.262, 0.271 and
0.275сm3 g−1 for 2P-TiO2, 4P-TiO2, 8P-TiO2, respectively which was in
Fig. 1. The X-ray diffraction analysis of raw and modified TiO2.
2.67, 2.76 and 2.80 times greater than that of raw titania.
The pore size distribution for unmodified titania showed that most
Table 1 of mesopores had a radius less than 2.2 nm (Fig. 4). Mesopores of this
The structural characteristics of anatase phase for the raw and modified TiO2. sample had a bimodal structure. The first maximum on the PSD curve
Sample Lattice constants, Å Lattice volume, Å3 Crystallite sizes, nm corresponded to pores with a radius around 1.3 nm, and second max-
imum on the PSD curve corresponded to pores with a radius 1.7 nm.
TiO2 a = 3.786, с = 9.430 135.17 ± 0.05 4.7 The introduction of 2% (wt.) of anions PO43− into TiO2 leds to an
2P-TiO2 a = 3.791, с = 9.430 135.52 ± 0.05 3.4
increase in the quantity of mesopores with the size 1.7 nm. Moreover,
4P-TiO2 a = 3.795, с = 9.470 136.38 ± 0.05 3.2
8P-TiO2 a = 3.809, с = 9.500 137.83 ± 0.05 3.0 an increase in the amount of chemosorbed phosphate groups in the
sample 4P-TiO2 led to the shift of the second maximum on the PSD
curve from 1.7 nm to 2.0 nm, and to the increase in the number of those
modified titania (4P-TiO2) were single-phase. Their crystalline structure pores. Further increase in the amount of phosphate groups in the 8P-
belongs to the anatase structure (space group I41/amd). Structural TiO2 sample provided an increase in Vmeso and Smeso, which leds to a
parameters of synthesized TiO2 samples are specified in Table 1. It reduction in mesopore size. The second maximum on the PSD curve was
could be concluded that there was a tendency to decrease the crystallite shifted to 1.7 nm. The surface area of phosphated samples increased in
sizes with increasing the amount of chemosorbed phosphate groups. 2–2.5 times compared to raw TiO2 due to increase of pore volume. This
The crystallite sizes were calculated from the XRD data and it could be can be attributed to role of surface phosphate groups, which prevent the
seen that they are actually match the dimensions of primary particles condensation of small TiO2 particles.
formed due to the condensation of the Ti(OH)4·2H2O and Ti(O2POOH)
(OH)2·2H2O molecules. 3.2.4. Infrared (IR) spectroscopy
The data presented in Table 1 shows that the lattice parameters of the IR-spectra of the raw titania sample and modified samples are depicted
modified TiO2 samples are greater than the lattice parameters of the raw in Fig. 5. It can be concluded, that the bands near 340, 460, 575 and
TiO2 sample. There is a tendency to increase the cell volume with the 730 cm−1 corresponded to the vibrations of Ti-O bonds in the TiO6
increase in phosphate groups amount. Specifically, in the sample 8P-TiO2, polyhedrons [32,37]. The bands at 1630 cm−1 corresponded to vibrations
the lattice volume was 137.83 Å3, which were 1.7% higher than the lattice of water [38–40]. According to [41–44], the asymmetric νasym and sym-
volume of the 2P-TiO2 sample and 1.97% higher than the lattice volume of metric νsym stretching vibrational modes of the PeO and PeOH bonds in
the unmodified titania sample. This dimensional effect was due to the chemisorbed phosphate groups for modified samples are observed at 780,
decrease in the dimensions of the crystallite sizes in the investigated TiO2 972, 1048 cm−1 (Fig. 5). These data indicated that covalent attachment of
samples. There is a tendency to reduce the size of the primary TiO2 NPs PO43- anions to cations Ti4+ from titanium aqua complex precursor was
with the increase of amount of chemosorbed phosphate groups on the carried out by dentate bonds and leads to the formation of chemi-
titania surface. The necessity of the lattice parameters of the oxide mate- sorbed = Ti(O2POOH) groups on the nanoparticles surface.
rials from the particle size was due to the Laplace pressure, which could
grow to 1.96 × 106 kPa in the nanoparticles. The convergence of Ti4+
3.3. Adsorption of strontium(II)
cations and the reduction of Ti-O-Ti valence angle between neighbor TiO6
octahedrons, due to pressure, causes a counteraction that manifests in the
3.3.1. Surface charge (pHPZC)
increase of the interatomic distance in the Ti-O bonds. This is due to the
The magnitude of pHPZC is a useful characteristic of the surface,
spacial effect caused by Laplace pressure, which was detected also in the
since it indicates the range of pH in which the adsorbent is positive or
fumed silica nanoparticles [35,36].
negative charged. If рH < рHPZC, the negatively charged nucleic centers
≡TiO− are appear on the TiO2 surface:
3.2.2. Transmission electronic microscopy ≡TiOH+OH− ↔ ≡TiO− + H2O (4)
The TEM observations displayed that the primary particles were
globular in shape with diameter 3–6 nm (Fig. 2a). In the acidic reaction If рH < рHPZC, so the positively charged nucleic centers ≡TiOH2+
medium, the primary particles do not coagulate, since they formed the appear on the TiO2 surface:
colloidal micelles with the electrolyte ions. The double electric layer
≡TiOH + H+ ↔ ≡TiOH2+ (5)
determines the micelle’s stability. During the NaOH addition, the ag-
gregate stability of the primary particles was lost and the aggregates The hydroxylated surface of TiO2 behaves like an amphoteric elec-
with spherical shape and a diameter of 20–60 nanometers were formed trolyte and it can exhibit both acidic and basic properties.

3
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

Fig. 2. TEM: (a) Primary particles of 4P-TiO2; (b) Aggregates of 4P-TiO2 nanoparticles; (c) Nanocrystallites of 4P-TiO2 xerogel.

Fig. 4. The PSD with respect to pore radii of raw and modified TiO2.

Fig. 3. Nitrogen adsorption/desorption isotherms for TiO2 samples.


qmax are in good consent with experimental obtained qmax and this
conclusion indicates that the adsorption of Sr(II) cations onto titania
The point of zero charge of each sample was measured by drift
samples is monolayer process. The rooting of titanium-phosphate
method. The measured magnitude of the pHPZC for raw TiO2 was 5.35,
groups = Tі(O2POOH) onto the titania surface increases the proton-
while for 2P-TiO2, 4P-TiO2, and 8P-TiO2 samples the magnitudes of the
donor ability of surface hydroxyls and they become able to bind the Sr(II)
pHPZC were 3.4, 3.1 and 3.1 respectively. The chemisorption of = Ti
cations. The 4P-TiO2 sample has highest value of KL = 0.0013 L/mg. The
(O2POOH) groups on the titania particles surface displaces the pHPZC
qmax, obtained from Langmuir isotherm, shown the following capacity
into the acidic area and extend the functionality of the adsorbent.
order: 4P-TiO2 > 8P-TiO2 > 2P-TiO2 > TiO2. It is well known that the
interactions between individual components can significantly alter
3.3.2. Adsorption isotherms sorption characteristics. The results obtained in the current study showed
The adsorption of Sr(II) cations were conducted for raw and mod- that Sr(II) removal by titania-based adsorbents involved monolayer ad-
ified TiO2 samples in order to know the maximum adsorption ability. sorption on the well-distributed active sites of the adsorbent with no
The adsorption isotherms are displayed in Fig. 6a. The results specified interactions between the neighboring adsorbate ions. The Freundlich
that the adsorption capacity of the raw titania reached to 70.9 mg/g model does not fitt well adsorption data due to lower R2 (0.9531-0.9799)
compared to phosphated samples 2P-TiO2, 4P-TiO2, and 8P-TiO2, which (Table S1). The parameters obtained from Dubinin-Radushkevich model,
demonstrated much higher adsorption capacity 94.1 mg/g, 172.5 mg/g which was utilised in order to get conclusion about physical and che-
and 128.9 mg/g, respectively. The modified adsorbent 4P-TiO2 had mical adsorption, are represented in Table S1. The data indicated the
significantly high adsorption capacity among the modified specimens high R2 value (more than 0.9606). The application of Dubinin-
and thus more effective in comparison with unmodified sample TiO2. Radushkevich model gave the mean adsorption energies of 11.04, 10.78,
The Langmuir, Freundlich and Dubinin-Radushkevich isotherm 12.31 and 11.79 kJ·mol−1 for TiO2, 2P-TiO2, 4P-TiO2, and 8P-TiO2 re-
models [45–48] were involved to determine the mechanism of stron- spectively. These values indicated that the Sr(II) adsorption on the raw
tium(II) adsorption by raw and phosphate-modified titania adsorbents and phosphated titania is ion-exchange process.
(Fig. 6b-d) and the model parameters are displayed in Table S1. Fig. 7 shows the removal efficiencies of raw and phosphated titania
It was established that adsorption data fitted well by Langmuir model adsorbents for 0.001 M Sr(II) solution at 298 K. The removal efficiency
due to the high R2 values (0.9763–0.9912). The calculated magnitudes of

Table 2
Morphological characteristics of titania samples.
Adsorbent S, m2·g−1 Smicro m2·g−1 Smeso m2·g−1 Smeso/S, % V cm3·g−1 Vmicro сm3·g−1 Vmeso сm3·g−1 Vmeso/V, %

TiO2 239 100 139 58.2 0.152 0.054 0.098 64.5


2P-TiO2 408 54 354 86.8 0.287 0.025 0.262 91.3
4P-TiO2 410 42 368 89.8 0.290 0.019 0.271 93.4
8P-TiO2 396 18 378 95.5 0.281 0.006 0.275 97.9

4
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

Fig. 7. The removal efficiencies of Sr(II) from 0.001 mol/L solution of SrCl2
onto unmodified and modified titania.

strontium on the modified titania samples (Fig. 8).


Compared to other known adsorbents (Table 3), the adsorption
capacity of raw and phosphated mesoporous titania adsorbents towards
strontium ions is much better and demonstrate the effectiveness of
synthesized in current study titania adsorbents for the Sr(II) elimination
from aquatic environment.
Fig. 5. IR spectra of raw and modified TiO2.

3.3.3. Kinetics studies


for Sr(II) adsorption onto adsorbents followed the order: 4P-TiO2 In order to estimate the rate of adsorption, which is one of the useful
(83%) > 8P-TiO2 (78%) > 2P-TiO2 (65%) > TiO2 (58%) suggesting a parameter in the adsorbent activity evaluation [50], the adsorption
larger Sr(II) removal efficiency for 4P-TiO2 sample. The EDS analysis kinetics studying was carried out and the adsorption mechanism was
was involved to evidence the presence of phosphorus and adsorbed proposed. The adsorption capacities (qt, mg/g) of unmodified and

Fig. 6. (a) Sr(II) adsorption data on the raw and phosphated titania adsorbents, fitted to (b) Langmuir, (c) Freundlich, (d) Dubinin-Radushkevich isotherm models
(pH 7.0, V = 5 mL, m(ads) = 50 mg, T =293 K, contact time 4 h).

5
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

Fig. 8. EDS of 4P-TiO2 sample after Sr(II) adsorption.

Table 3 where k2 (g·mg−1 min−1) is the rate constant of the PSO model; q e
Comparing of the adsorption ability for reported adsorbents for the Sr(II) (mg·g−1) is the equilibrium amount of adsorbate. It could be seen that
elimination. obtained values of initial sorption rate h are 2.95; 4.58; 23.53 and
Adsorbent qe (mg·g−1) References 19.65 mg g−1 min-1 for TiO2, 2P-TiO2, 4P-TiO2, and 8P-TiO2, respec-
tively. These data suggested that the titania modified by 4%(wt.) of
magnetic chitosan beads 11.58 [16] phosphate revealed the maximum h. It should be indicated that the PSO
phosphate-modified montmorillonite 12.5 [15]
model fit well the processes means that it involved more than one stage
Egyptian soils 12.53 [14]
TiO2 NPs 25.00 [21]
for adsorption from aquatic medium between the adsorbent and ad-
Titanate nanotubes 66.72 [20] sorbate species [51], what is taking place in the current study (please
Hydroxyapatite 24.5 (0.28 mmol·g−1) [49] see the explanation below).
TiO2 70.9 This study The Elovich model was applied to explore the adsorption initial rate
2P-TiO2 94.1 This study
and the desorption constant . From Fig. 10c it is clear that plot qt vs.
8P-TiO2 128.9 This study
4P-TiO2 172.5 This study ln(t) show good linear correlation. The slope and intercept give us the
Elovich constants, which are tabulated in Table S2. The highest mag-
nitude of is for 8P-TiO2 sample ( = 2008.11), and desorption con-
modified titanium adsorbents for the Sr(II) adsorption vs. the contact stant for phosphated titania are in three-four times smaller ( = 0.17-
time (t, min) depicted in Fig. 9 (for 0.01 mol/L initial strontium(II) 0.28) than that of raw sample ( = 0.69 for TiO2).
concentration and for contact time 140 min). The adsorption was fast at In order to forecast the rate-controlling step of adsorption the in-
the beginning and reached the equilibrium after 60 min. traparticle diffusion model (IPD) [48] has been involved to fit the ex-
A pseudo-first-order model, a pseudo-second-order model, an perimental data and the estimated kinetic parameters were shown in
Elovich model, and an intraparticle diffusion model [48] were applied Table S2. The analysis has shown that plot qt vs. t0.5 has one linear
to explain the kinetics of adsorption. All kinetics parameters, estimated ranges for TiO2 and two linear ranges for phosphated samples
from experimental data and the above-mentioned models, are pre- (Fig. 10d). It can be concluded that the adsorption onto raw titania
sented in Table S2 and depicted in Fig. 10(a–d). Fitting degree of the includes surface adsorption and micropore diffusion, while the ad-
kinetic models was evaluated by correlation coefficient R2. sorption process onto phosphated adsorbents includes the transporta-
The results indicated that the obtained data well described by the tion of strontium cations to the external surface and diffusion of Sr(II)
PSO kinetic model (R2 = 0.9980 – 0.9999). The calculated from PSO into the mesopores of adsorbent. In addition, the plots did not pass
model equilibrium adsorption capacities qe(calc) are in the good agree- through the origin (Fig. 10d), and constants C1, and C2 were not zero,
ment with experimentally obtained adsorption capacity qe(exp). The demonstrating that diffusion through boundary layer also occurred in
initial sorption rate h (mg·g−1 min−1) was estimated by the Eq. 6: the uptake of Sr(II) onto adsorbents [48]. It can be noticed, the con-
stants C2 are higher than constants C1 indicating that the contribution
h = k2 qe2 of boundary layer to adsorption rate increased from stage 1 to stage 2.
(6)

3.3.4. The influence of pH


The pH is a parameter that affect the adsorption ability of any ad-
sorbent [52–56]. Fig. 11 displayed the influence of pH on Sr(II) ions
adsorption onto raw and phosphated titania samples.
As can be seen from Fig.11, the phosphated sorbents exhibited a
significantly higher adsorption ability compared to raw titania in acidic
and neutral mediums. In the alkaline solution (pH = 8–9) the raw TiO2
and phosphated adsorbents possess better adsorption ability. For ex-
ample, the adsorption ability of 4P-TiO2 sample was increased from
51.7 mg/g (at neutral pH) to 73.9 mg/g (at pH = 9), while the ad-
sorption ability of raw titania was increased from 21.09 mg/g to
70.4 mg/g at neutral pH and pH = 9, respectively.
It was shown that the phosphate groups = Tі(O2POOH) and groups
≡TіOHδ+ (Brønsted acid centers) are the active centers for Sr(II)
binding on the surface of modified TiO2 samples, because they are
transforming into negatively charged centers = Tі(O2POO–) and
Fig. 9. Kinetics of Sr(II) adsorption onto raw and phosphated titania (pH = 7.0, ≡TіO− in a medium with pH > pHPZC. The structural characteristics
298 K, m(ads) = 0.05 g, Co(Sr)=0.01 mol·L−1, V = 5 ml). of the TiO2 samples were taken into account to calculate the average

6
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

Fig. 10. Kinetics studies: (a) PFO; (b) PSO; (c) Elovich; (d) intraparticle diffusion model (pH = 7.0, 298 K, mads = 50 mg, V =20 mL).

number of ≡TіOH groups on the titania globules surface [28]. The Table 4
calculations (Table 4) have shown that the average amount of those The adsorption centers on the phosphated titania surface.
groups is 128 units per 10 nm2 of surface area. According to [57] there Adsorbent The amount of adsorption centers (per 10 nm2) The
are 120–140 ≡TіOH groups per 10 mn2 on the TiO2 surface area and percentage of
those groups can be neutral ≡TіOH, basic ≡TіOHδ– or acidic adsorption
≡TіOHδ+. centers (%)
=Tі(O2POOH) ≡TіOHδ+ ≡TіOHδ−
The results show that the rise in the amount of chemosorbed
phosphate groups leads to an increase of titania activity regarding to TіO2* 0 27.0 0 21.1
strontium adsorption. From Table 4 it can be seen that there are 27 2P-TіO2 1.9 29.7 0.7 25.2
≡TіOHδ+ active centers per 10 mn2 for non-modified titania ad- 4P-TіO2 3.7 54.1 1.8 45.2
8P-TіO2 8.0 36.8 1.0 35.8
sorbent, what is equal to 21.1% of total number of ≡TіOH groups. The
2P-TіO2 adsorbent surface contain 1.9 phosphate groups and 29.7 ac- * the average amount of hydroxyl groups is 128 species per 10 nm2 [28].
tive centers ≡TіOHδ+ per 10 mn2, while the chemisorption of 3.7
phosphate groups per 10 mn2 in the 4P-TіO2 sample leads to formation centers, what is less than in the 4P-TіO2 sample. The decreasing of the
of 54.1 acidic centers ≡TіOHδ+ (45.2%) on its surface. In the 8P-TiO2 active centers number in this adsorbent is probably due to higher ag-
sample the eight phosphate groups cause the formation of 36.8 acidic gregation of the primary particles, which limits the access of Sr(II)

Fig. 11. The impact of pH on Sr(II) ions adsorption onto phosphated titania.

7
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

cations to the active centers. The data presented in the Table 4 allow us
to calculate the ability of chemisorbed titanium-phosphate groups to
form acidic centers ≡TіOHδ+ in the vicinity. In particular, in the 4P-
TiO2 adsorbent the one = Tі(O2POOH) group generates 14.6 Brønsted
centers which are able to bind the strontium cations.
The rooting of titanium-phosphate groups = Tі(O2POOH) into the
TiO2 surface structure increases the proton-donor ability of surface
hydroxyls and they become able to bind the Sr(II) cations. The higher
electronegativity of phosphorus atoms compared to titanium atoms
causes the displacement of the electron density to the phosphorus atoms
in the = Tі(O2POOH) groups. The inductive effect of the electron
density redistribution is also taking place in the Ti-O-Ti bridges at-
tached to the titanium-phosphate groups, therefore, the neighboring
hydroxyl groups get proton-donor properties and become as Brønsted
acid centers (Scheme (7)): Fig. 12. Reusability of phosphated titania adsorbents (experimental conditions:
0.0025 mol·L−1 Sr(II) solution, temperature 25 °C, pH = 7.0).

3.3.5. Regeneration studies


The adsorption/desorption experiments were conducted to explore
the reusability of phosphated sorbents. The 0.001 M HNO3 solution was
used to investigate the desorption of Sr(II) cations from the surface of
(7) sorbents. The obtained results are shown in Fig. 12 that indicated that
adsorbents have good reusability. The adsorbents (TiO2, 2P-TiO2 and
Thus, in phosphated TiO2 samples, the developed structure of me-
4P-TiO2) did not lose their adsorption capacity during the four cycles of
sopores, their large specific surface, and also the ionogenic nature of
adsorption/desorption, while 8P-TiO2 lose its adsorption capacity from
chemosorbed groups = Ti(O2POO−) provided the high adsorption ac-
10.45 mg/g to 2.4 mg/g. Typically, all adsorbed amount of Sr ions was
tivity for Sr(II) cations. The adsorption of cations by phosphated sor-
desorbed by eluent. However, the 8P-TiO2 sample lose its adsorption
bents in acidic and neutral media (pH = 2–7) was carried out according
capacity. The probable reasons are pore blocking and nanoparticle
to the following scheme (8):

(8)

In a low alkaline medium (pH = 8.0), the formation of cations


SrOH+ led to an increase in their adsorption (Scheme 9):

(9)

Based on these schemes, Sr adsorption mechanism was mostly


electrostatic interaction and Sr(II) cations could be adsorbed on the aggregation. The decrease of adsorption capacity in the 8P-TiO2 sample
phosphate modified TiO2 because of negatively charge surface caused can also be attributed to loss of some phosphate groups from the titania
by deprotonation of TiOH and phosphate groups. However, in a surface resulting in decrease of active adsorption centers.
strongly alkaline medium (pH ≥ 9.0), chemisorbed phosphate groups
were split off from the adsorbent surface. The breaking of Ti-O-P bridge 4. Conclusions
bonds and the loss of chemosorbed groups aligned the adsorption ac-
tivity of the unmodified and modified sorbents. The globular titania particles with a 3–4 nm diameter with chemo-
The adsorption of iodide anions from potassium iodide solution were sorbed on their surfaces phosphate groups = Ti(O2POOH) have been
performed to calculate the amount of basic centers ≡TіOHδ− onto ad- obtained by liquid-phase route. During the gelation and dispersion
sorbents surface. The experiments have shown that non-modified TiO2 does drying stage, the sites with the chemosorbed phosphate groups blocked
not adsorb anions while the adsorption of iodide anions by the 2P-TiO2, 4P- the accretion of the primary particles and led to the formation of a
TiO2, 8P-TiO2 adsorbents is 0.052, 0.126 and 0.066 mmol·g-1 respectively. xerogel material with a homogeneous mesoporous structure. Such

8
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

liquid-phase synthesis with the use of 4% (wt.) of modifying reagent 1039/c7cp03770h.


allowed obtaining the adsorbent with a mesopore volume of [10] M. Bououdina, T.S. Alwqyan, L. Khezami, B. Al-Najar, M.N. Shaikh, R. Gill,
A. Modwi, K.K. Taha, O.M. Lemine, Fabrication and characterization of nanos-
0.275 cm3 g−1 and specific surface of 368 m2 g−1, which in 2.8 and 2.6 tructured MgO·Fe2O3 composite by mechanical milling as efficient adsorbent of
times exceed the unmodified TiO2. heavy metals, J. Alloys. Compd. 772 (2019) 1030–1039, https://doi.org/10.1016/
The developed mesopore structure, a large specific surface of the J.JALLCOM.2018.09.010.
[11] I.-G. Yi, J.-K. Kang, S.-C. Lee, C.-G. Lee, S.-B. Kim, Synthesis of an oxidized meso-
sorbent, as well as the ionogenic nature of chemosorbed groups pro- porous carbon-based magnetic composite and its application for heavy metal re-
vided its high adsorption ability for Sr(II) removal. Since, chemisorbed moval from aqueous solutions, Microporous Mesoporous Mater. 279 (2019) 45–52,
phosphate groups shifted the surface charge pHPZC from 5.35 to https://doi.org/10.1016/J.MICROMESO.2018.12.016.
[12] T. Wen, X. Wu, M. Liu, Z. Xing, X. Wang, A.W. Xu, Efficient capture of strontium
3.1–3.4, which expanded the functionality of the adsorbent, and pro- from aqueous solutions using graphene oxide-hydroxyapatite nanocomposites,
vided an effective adsorption of metal cations in the acidic medium. Dalton Trans. 43 (2014) 7464–7472, https://doi.org/10.1039/c3dt53591f.
The maximum equilibrium adsorption capacity obtained from the [13] K. Hirose, Fukushima Dai-ichi nuclear power plant accident: summary of regional
radioactive deposition monitoring results, J. Environ. Radioact. 111 (2012) (2011)
Langmuir model for modified sorbent 4P-TiO2 was noted 172.5 mg.g−1
13–17, https://doi.org/10.1016/J.JENVRAD.2011.09.003.
and the adsorption of Sr(II) by phosphated adsorbent was exceeded [14] N.H.M. Kamel, Adsorption models of 137Cs radionuclide and Sr (II) on some
1.3–2.4 times compared to unmodified TiO2. Sr adsorption mechanism Egyptian soils, J. Environ. Radioact. 101 (2010) 297–303, https://doi.org/10.
was mostly electrostatic interaction and Sr(II) cations adsorbed on the 1016/J.JENVRAD.2010.01.001.
[15] B. Ma, S. Oh, W.S. Shin, S.-J. Choi, Removal of Co2+, Sr2+ and Cs+ from aqueous
phosphate modified TiO2 because of negatively charge surface caused solution by phosphate-modified montmorillonite (PMM), Desalination 276 (2011)
by deprotonation of TiOH and phosphate groups. 336–346, https://doi.org/10.1016/J.DESAL.2011.03.072.
[16] Y. Chen, J. Wang, Removal of radionuclide Sr2+ ions from aqueous solution using
synthesized magnetic chitosan beads, Nucl. Eng. Des. 242 (2012) 445–451, https://
Declaration of Competing Interest doi.org/10.1016/J.NUCENGDES.2011.10.059.
[17] H. Tel, Y. Altaş, M. Eral, Ş. Sert, B. Çetinkaya, S. İnan, Preparation of ZrO2 and
The authors declare that they have no known competing financial ZrO2–TiO2 microspheres by the sol–gel method and an experimental design ap-
proach to their strontium adsorption behaviours, Chem. Eng. J. 161 (2010)
interests or personal relationships that could have appeared to influ- 151–160, https://doi.org/10.1016/J.CEJ.2010.04.053.
ence the work reported in this paper. [18] D. Li, B. Zhang, F. Xuan, The sequestration of Sr(II) and Cs(I) from aqueous solu-
tions by magnetic graphene oxides, J. Mol. Liq. 209 (2015) 508–514, https://doi.
org/10.1016/J.MOLLIQ.2015.06.022.
Acknowledgement [19] E. Kaçan, C. Kütahyalı, Adsorption of strontium from aqueous solution using acti-
vated carbon produced from textile sewage sludges, J. Anal. Appl. Pyrolysis 97
This work was supported by the Ministry of Education and Science (2012) 149–157, https://doi.org/10.1016/J.JAAP.2012.06.006.
[20] S. Kasap, S. Piskin, H. Tel, Titanate nanotubes: preparation, characterization and
of Ukraine (Project number MESU 0117U002408). One of the authors
application in adsorption of strontium ion from aqueous solution, Radiochim. Acta
(Mu. Naushad) acknowledges the Research Supporting Project number 100 (2012) 925–929, https://doi.org/10.1524/ract.2012.1981.
(RSP-2019/8), King Saud University, Riyadh, Saudi Arabia for the [21] S. Kasap, H. Tel, S. Piskin, Preparation of TiO2 nanoparticles by sonochemical
support. method, isotherm, thermodynamic and kinetic studies on the sorption of strontium,
J. Radioanal. Nucl. Chem. 289 (2011) 489–495, https://doi.org/10.1007/s10967-
011-1090-2.
Appendix A. Supplementary data [22] S.S. Metwally, I.M. Ahmed, H.E. Rizk, Modification of hydroxyapatite for removal
of cesium and strontium ions from aqueous solution, J. Alloys. Compd. 709 (2017)
438–444, https://doi.org/10.1016/J.JALLCOM.2017.03.156.
Supplementary material related to this article can be found, in the [23] W. Mu, Q. Yu, X. Li, H. Wei, Y. Jian, Niobate nanofibers for simultaneous adsorptive
online version, at doi:https://doi.org/10.1016/j.jece.2019.103430. removal of radioactive strontium and iodine from aqueous solution, J. Alloys.
Compd. 693 (2017) 550–557, https://doi.org/10.1016/J.JALLCOM.2016.09.200.
[24] N.I. Chubar, V.A. Kanibolotskyy, V.V. Strelko, G.G. Gallios, V.F. Samanidou,
References T.O. Shaposhnikova, V.G. Milgrandt, I.Z. Zhuravlev, Adsorption of phosphate ions
on novel inorganic ion exchangers, Colloids Surf. A Physicochem. Eng. Asp. 255
[1] T. Tatarchuk, M. Bououdina, B. Al-Najar, R.B. Bitra, Green and ecofriendly mate- (2005) 55–63, https://doi.org/10.1016/J.COLSURFA.2004.12.015.
rials for the remediation of inorganic and organic pollutants in water, in: [25] V.V. Strelko, New sol-gel processes in the synthesis of inorganic sorbents and ion
M. Naushad (Ed.), A New Gener. Mater. Graphene Appl. Water Technol. Springer, exchangers based on nanoporous oxides and phosphates of polyvalent metals, J.
Cham, 2019, pp. 69–110, , https://doi.org/10.1007/978-3-319-75484-0_4. Solgel Sci. Technol. 68 (2013) 438–446, https://doi.org/10.1007/s10971-013-
[2] E. Da’na, Adsorption of heavy metals on functionalized-mesoporous silica: a review, 2990-0.
Microporous Mesoporous Mater. 247 (2017) 145–157, https://doi.org/10.1016/J. [26] M. Kapnisti, F. Noli, P. Misaelides, G. Vourlias, D. Karfaridis, A. Hatzidimitriou,
MICROMESO.2017.03.050. Enhanced sorption capacities for lead and uranium using titanium phosphates;
[3] A.A. Basheer, New generation nano-adsorbents for the removal of emerging con- sorption, kinetics, equilibrium studies and mechanism implication, Chem. Eng. J.
taminants in water, J. Mol. Liq. 261 (2018) 583–593, https://doi.org/10.1016/J. 342 (2018) 184–195, https://doi.org/10.1016/J.CEJ.2018.02.066.
MOLLIQ.2018.04.021. [27] M. Sharma, D. Choudhury, S. Hazra, S. Basu, Effective removal of metal ions from
[4] L. Sellaoui, F.E. Soetaredjo, S. Ismadji, A. Bonilla-Petriciolet, C. Belver, J. Bedia, aqueous solution by mesoporous MnO2 and TiO2 monoliths: kinetic and equili-
A. Ben Lamine, A. Erto, Insights on the statistical physics modeling of the adsorp- brium modelling, J. Alloys. Compd. 720 (2017) 221–229, https://doi.org/10.1016/
tion of Cd2+ and Pb2+ ions on bentonite-chitosan composite in single and binary J.JALLCOM.2017.05.260.
systems, Chem. Eng. J. 354 (2018) 569–576, https://doi.org/10.1016/J.CEJ.2018. [28] I. Mironyuk, T. Tatarchuk, M. Naushad, H. Vasylyeva, I. Mykytyn, Highly efficient
08.073. adsorption of strontium ions by carbonated mesoporous TiO2, J. Mol. Liq. 285
[5] M. Vafaeifard, S. Ibrahim, K.T. Wong, P. Pasbakhsh, S. Pichiah, J. Choi, Y. Yoon, (2019) 742–753, https://doi.org/10.1016/J.MOLLIQ.2019.04.111.
M. Jang, Novel self-assembled 3D flower-like magnesium hydroxide coated gran- [29] S.-F. Zhou, J.-J. Wang, L. Gan, X.-J. Han, H.-L. Fan, L.-Y. Mei, J. Huang, Y.-Q. Liu,
ular polyurethane: Implication of its potential application for the removal of heavy Individual and simultaneous electrochemical detection toward heavy metal ions
metals, J. Clean. Prod. 216 (2019) 495–503, https://doi.org/10.1016/J.JCLEPRO. based on L-cysteine modified mesoporous MnFe2O4 nanocrystal clusters, J. Alloys.
2018.12.135. Compd. 721 (2017) 492–500, https://doi.org/10.1016/J.JALLCOM.2017.05.321.
[6] A. Maleki, Z. Hajizadeh, V. Sharifi, Z. Emdadi, A green, porous and eco-friendly [30] L.I. Myronyuk, I.F. Myronyuk, V.L. Chelyadyn, V.M. Sachko, M.A. Nazarkovsky,
magnetic geopolymer adsorbent for heavy metals removal from aqueous solutions, R. Leboda, J. Skubiszewska-Zięba, V.M. Gun’ko, Structural and morphological
J. Clean. Prod. 215 (2019) 1233–1245, https://doi.org/10.1016/J.JCLEPRO.2019. features of crystalline nanotitania synthesized in different aqueous media, Chem.
01.084. Phys. Lett. 583 (2013) 103–108, https://doi.org/10.1016/J.CPLETT.2013.07.068.
[7] M. Barczak, K. Michalak-Zwierz, K. Gdula, K. Tyszczuk-Rotko, R. Dobrowolski, [31] A. Ivanets, N. Kitikova, I. Shashkova, Y. Matrunchik, L. Kul’bitskaya, M. Sillanpää,
A. Dąbrowski, Ordered mesoporous carbons as effective sorbents for removal of Non-acidic synthesis of phosphatized dolomite and its sorption behaviour towards
heavy metal ions, Microporous Mesoporous Mater. 211 (2015) 162–173, https:// Pb 2+, Zn2+, Cu2+, Cd2+, Ni2+, Sr2+ and Co2+ ions in multicomponent
doi.org/10.1016/J.MICROMESO.2015.03.010. aqueous solution, Environ. Technol. Innov. 6 (2016) 152–164, https://doi.org/10.
[8] A. Shahat, M.R. Awual, M. Naushad, Functional ligand anchored nanomaterial 1016/J.ETI.2016.09.001.
based facial adsorbent for cobalt(II) detection and removal from water samples, [32] I. Mironyuk, T. Tatarchuk, H. Vasylyeva, V.M. Gun’ko, I. Mykytyn, Effects of che-
Chem. Eng. J. 271 (2015) 155–163, https://doi.org/10.1016/J.CEJ.2015.02.097. mosorbed arsenate groups on the mesoporous titania morphology and enhanced
[9] L. Sellaoui, F. Edi Soetaredjo, S. Ismadji, É. Cláudio Lima, G.L. Dotto, A. Ben adsorption properties towards Sr(II) cations, J. Mol. Liq. 282 (2019) 587–597,
Lamine, A. Erto, New insights into single-compound and binary adsorption of https://doi.org/10.1016/J.MOLLIQ.2019.03.026.
copper and lead ions on a treated sea mango shell: experimental and theoretical [33] D.C. Harris, Quantitative Chemical Analysis, Seventh, New York, (2007).
studies, Phys. Chem. Chem. Phys. 19 (2017) 25927–25937, https://doi.org/10. [34] T. Tatarchuk, N. Paliychuk, R.B. Bitra, A. Shyichuk, M. Naushad, I. Mironyuk,

9
I. Mironyuk, et al. Journal of Environmental Chemical Engineering 7 (2019) 103430

D. Ziolkovska, Adsorptive removal of toxic Methylene blue and Acid Orange 7 dyes 385–470, https://doi.org/10.1016/j.cej.2009.09.013.
from aqueous medium using cobalt-zinc ferrite nanoadsorbents, Desalin. Water [47] T.P. O’Connor, J. Mueller, Modeling competitive adsorption of chlorinated volatile
Treat. 150 (2019) 374–385, https://doi.org/10.5004/dwt.2019.23751. organic compounds with the Dubinin–Radushkevich equation, Microporous
[35] I.F. Myronyuk, V.I. Mandzyuk, T.V. Herhel, Dimensional effects in the pyrogenic Mesoporous Mater. 46 (2001) 341–349, https://doi.org/10.1016/S1387-1811(01)
silicon dioxide nanoparticles, Phys. Chem. Solid State. 6 (2005) 34. 00317-1.
[36] I.F. Myronyuk, Y.P. Voronin, V.I. Mandzyuk, N.A. Bezruka, T.V. Dmytrotsa, The [48] H.N. Tran, S.-J. You, A. Hosseini-Bandegharaei, H.-P. Chao, Mistakes and incon-
dimensional effect in trimethylsilylated silica nanoparticles, J. Nano- Electron. sistencies regarding adsorption of contaminants from aqueous solutions: a critical
Phys. 9 (2017) 1–5, https://doi.org/10.21272/jnep.9(5).05030. review, Water Res. 120 (2017) 88–116, https://doi.org/10.1016/J.WATRES.2017.
[37] V. Kotsyubynsky, I. Myronyuk, V. Chelyadyn, A. Hrubiak, V. Moklyak, 04.014.
S. Fedorchenko, Rod-like rutile nanoparticles: synthesis, structure and morphology, [49] I.L. Shashkova, A.I. Ivanets, N.V. Kitikova, Sorption of Co2+, Pb2+, and Sr2+ ions
J. Nano Res. 50 (2017) 32–40, https://doi.org/10.4028/www.scientific.net/ on hydroxyapatite, synthesized in the presence of oxyethylidenediphosphonic acid,
JNanoR.50.32. Russ. J. Appl. Chem. 92 (2019) 625–633, https://doi.org/10.1134/
[38] T. Tatarchuk, N. Paliychuk, M. Pacia, W. Kaspera, W. Macyk, A. Kotarba, S1070427219050070.
B.F. Bogacz, A.T. Pędziwiatr, I. Mironyuk, R. Gargula, P. Kurzydło, A. Shyichuk, [50] T. Tatarchuk, B. Al-Najar, M. Bououdina, M.A.A. Ahmed, Catalytic and photo-
Structure-redox reactivity relationships in Co 1−x Zn x Fe 2 O 4 : the role of catalytic properties of oxide spinels, in: L. Martínez, O. Kharissova, B. Kharisov
stoichiometry, New J. Chem. 43 (2019) 3038–3049, https://doi.org/10.1039/ (Eds.), Handb. Ecomater. Springer, Cham, 2019, pp. 1701–1750, , https://doi.org/
c8nj05329d. 10.1007/978-3-319-68255-6_158.
[39] I.F. Mironyuk, V.M. Gun’ko, H.V. Vasylyeva, O.V. Goncharuk, T.R. Tatarchuk, [51] X. Guo, J. Wang, A general kinetic model for adsorption: theoretical analysis and
V.I. Mandzyuk, N.A. Bezruka, T.V. Dmytrotsa, Effects of enhanced clusterization of modeling, J. Mol. Liq. 288 (2019) 111100, , https://doi.org/10.1016/J.MOLLIQ.
water at a surface of partially silylated nanosilica on adsorption of cations and 2019.111100.
anions from aqueous media, Microporous Mesoporous Mater. 277 (2019) 95–104, [52] M. Naushad, Surfactant assisted nano-composite cation exchanger: development,
https://doi.org/10.1016/j.micromeso.2018.10.016. characterization and applications for the removal of toxic Pb2+ from aqueous
[40] P. Tiwari, R. Verma, S.N. Kane, T. Tatarchuk, F. Mazaleyrat, Effect of Zn addition on medium, Chem. Eng. J. 235 (2014) 100–108, https://doi.org/10.1016/J.CEJ.2013.
structural, magnetic properties and anti-structural modeling of magnesium-nickel 09.013.
nano ferrites, Mater. Chem. Phys. 229 (2019) 78–86, https://doi.org/10.1016/j. [53] M. Naushad, T. Ahamad, G. Sharma, A.H. Al-Muhtaseb, A.B. Albadarin, M.M. Alam,
matchemphys.2019.02.030. Z.A. ALOthman, S.M. Alshehri, A.A. Ghfar, Synthesis and characterization of a new
[41] A. Barhoumi, J.J. Suñol, M. Belhouchet, Crystal structure, vibrational studies and starch/SnO2 nanocomposite for efficient adsorption of toxic Hg2+ metal ion,
optical properties of a new organic phosphate (C12H14N2S) (H2PO4)2, J. Mol. Chem. Eng. J. 300 (2016) 306–316, https://doi.org/10.1016/J.CEJ.2016.04.084.
Struct. 1173 (2018) 448–455, https://doi.org/10.1016/j.molstruc.2018.06.115. [54] M. Naushad, T. Ahamad, B.M. Al-Maswari, A. Abdullah Alqadami, S.M. Alshehri,
[42] A. Roguska, M. Pisarek, M. Andrzejczuk, M. Dolata, M. Lewandowska, M. Janik- Nickel ferrite bearing nitrogen-doped mesoporous carbon as efficient adsorbent for
Czachor, Characterization of a calcium phosphate-TiO2 nanotube composite layer the removal of highly toxic metal ion from aqueous medium, Chem. Eng. J. 330
for biomedical applications, Mater. Sci. Eng. C. 31 (2011) 906–914, https://doi. (2017) 1351–1360, https://doi.org/10.1016/J.CEJ.2017.08.079.
org/10.1016/j.msec.2011.02.009. [55] L. Sellaoui, D.I. Mendoza-Castillo, H.E. Reynel-Ávila, B.A. Ávila-Camacho,
[43] C. Sronsri, B. Boonchom, Synthesis, characterization, vibrational spectroscopy, and L.L. Díaz-Muñoz, H. Ghalla, A. Bonilla-Petriciolet, A. Ben Lamine, Understanding
factor group analysis of partially metal-doped phosphate materials, Spectrochim. the adsorption of Pb2+, Hg2+ and Zn2+ from aqueous solution on a lig-
Acta - Part A Mol. Biomol. Spectrosc. 194 (2018) 230–240, https://doi.org/10. nocellulosic biomass char using advanced statistical physics models and density
1016/j.saa.2018.01.034. functional theory simulations, Chem. Eng. J. 365 (2019) 305–316, https://doi.org/
[44] X. Shen, Q. Wang, Y. Liu, W. Xue, L. Ma, S. Feng, M. Wan, F. Wang, C. Mao, 10.1016/J.CEJ.2019.02.052.
Manganese phosphate self-assembled nanoparticle surface and its application for [56] A.Q. Selim, L. Sellaoui, M. Mobarak, Statistical physics modeling of phosphate
superoxide anion detection, Sci. Rep. 6 (2016) 1–9, https://doi.org/10.1038/ adsorption onto chemically modified carbonaceous clay, J. Mol. Liq. 279 (2019)
srep28989. 94–107, https://doi.org/10.1016/J.MOLLIQ.2019.01.100.
[45] I. Langmuir, The adsorption of gases on plane surfaces of glass, MICA AND [57] H.V. Lysychkyn, A.Y. Fadeev, A.A. Serdan, P.N. Nesterenko, P.H. Mynhalev,
PLATINUM, J. Am. Chem. Soc. 40 (1918) 1361–1403, https://doi.org/10.1021/ D.B. Furman, Khymyia Pryvytykh Poverkhnostnykh Soedynenyi, FYZMATLYT,
ja02242a004. Moscow, 2003.
[46] H.M.F. Freundlich, Adsorption in solution, Z. Phys. Chem. N. (N F) 57 (1906)

10

You might also like