Вы находитесь на странице: 1из 11

Applied Clay Science 162 (2018) 507–517

Contents lists available at ScienceDirect

Applied Clay Science


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

Research Paper

Highly-effective phosphate removal from aqueous solutions by calcined T


nano-porous palygorskite matrix with embedded lanthanum hydroxide

Lingchao Kong, Yu Tian , Ning Li, Yang Liu, Jian Zhang, Jun Zhang, Wei Zuo
State Key Laboratory of Urban Water Resource and Environment, School of Environment, Harbin Institute of Technology, Harbin 150090, China

A R T I C LE I N FO A B S T R A C T

Keywords: Efficient phosphate removal from wastewater is critical for the safety of natural water bodies against eu-
Palygorskite trophication and for the replenishment of unrenewable phosphorus resources. In order to enhance phosphate
Lanthanum hydroxide capture efficiency and exploit raw clay for high value-added products, HPAL-LaOH was fabricated by embedding
Phosphate adsorption lanthanum hydroxide onto calcined nano-porous palygorskite clay through a simple-green hydrothermal method
Surface complexation
where the host palygorskite with cross-linked networks providing abundant binding sites for La(OH)3 in-
corporation, allowing the accessibility of phosphate for the adequate crystallization of lanthanum phosphate
without triggering pore blockage. Excellent phosphate adsorption capacity (109.63 mg/g) was achieved by
HPAL-LaOH, over 13 times higher than that of raw palygorskite, also much higher than commercial pure La
(OH)3 (69.64 mg/g) in batch runs. Interestingly, for solutions below 100 mg/L in a wide pH range of 3–11,
almost complete phosphate sequestration (< 0.01 mg/L) was achieved. The presence of high level competing
anions (sulfate, nitrate, bicarbonate and chloride) merely exhibited an insignificant effect. Notably, HPAL-LaOH
demonstrated satisfactory recyclability, settleability and excellent stability with negligible lanthanum leaching
even under ultrasonic challenge. Mechanism analysis revealed that the impregnated La(OH)3 exerted specific
phosphate adsorption where inner-sphere complexation by ligand-exchange played a major role. As compared to
several La-based adsorbents developed recently for phosphate sequestration, HPAL-LaOH exhibited great com-
petiveness in terms of adsorption capacity and La usage. The results indicated the potential utilization of HPAL-
LaOH as a highly cost-effective adsorbent for phosphate removal from wastewater.

1. Introduction et al., 2009), diatomite (Xie et al., 2014a), zeolite (Xie et al., 2014b;
Goscianska et al., 2018), aluminum oxide clay (Tanada et al., 2003; Yan
Excessive phosphorus (P) from drained farmlands, detergents and et al., 2010), and iron oxide clay (Pan et al., 2009; Acelas et al., 2015),
industrial activities washing into lakes and streams causes eutrophica- etc. However, their relatively low adsorption capacity and complex
tion, which leads to hazardous algae bloom, dissolved oxygen depletion preparation procedures play restricting roles on governing the running
and fish death (Zamparas and Zacharias, 2014), severely disturbing the cost of their utilization in practical wastewater treatment and in-lake
ecological balance of organisms present in water, and consequently, eutrophication control (Yin et al., 2016).
impairing public health. Thus, feasible techniques are required to Palygorskite, a raw hydrated aluminum-magnesium silicate clay,
suppress P loading into natural aqueous bodies. Adsorption is re- has been regarded as a promising adsorbent for contaminant removal.
commended as a promising option for aqueous pollutant removal due to The unique structural and adsorptive properties of palygorskite have
its simplicity of operation and feasibility of economics (Zhong et al., also been drawn upon for developing aqueous phosphate remover as
2014; Hu et al., 2017; Zhou et al., 2018), which also provides the matrix (Carazo et al., 2018; Yang et al., 2018). Ye et al. (2006) modified
possibility of P recovery (Loganathan et al., 2014; Mitrogiannis et al., palygorskite by hydrochloric acid or thermal treatment and rendered a
2017; Zeng et al., 2017; Zhou et al., 2017). To this end, low-cost clay selective phosphate capture with adsorption capacity of 10.20 mg/g in
materials have attracted much attention to develop effective adsorbents complex solutions including Cl−, NO3−, CO32– and SO42−. Then, Gan
for enhanced phosphorus removal, including montmorillonite (Tian et al. (2009) further activated palygorskite by calcination at 700 °C and

Abbreviations: P, phosphorus; La, lanthanum; PAL, raw palygorskite; HPAL, calcined palygorskite at 700 °C for 4 h; HPAL-LaOH, lanthanum hydroxide-impregnated
HPAL; HPAL-NaOH, NaOH solution treated HPAL without introducing lanthanum

Corresponding author.
E-mail address: hit_tianyu@163.com (Y. Tian).

https://doi.org/10.1016/j.clay.2018.07.005
Received 9 March 2018; Received in revised form 8 June 2018; Accepted 2 July 2018
0169-1317/ © 2018 Elsevier B.V. All rights reserved.
L. Kong et al. Applied Clay Science 162 (2018) 507–517

the highest phosphate adsorption capacity reached 42.0 mg/g. Un- Table 1
fortunately, using sole raw palygorskite clay for phosphate removal Chemical composition of the raw palygorskite (wt
usually present limited active sites and inhomogeneous distribution of %).
inherent components, severely restraining the adsorption efficiency. Constituent Value (%)
Thus, developing new generation palygorskite-based cost-effective hy-
brid adsorbents for enhanced phosphate removal is of much demand. SiO2 53.897
Al2O3 15.326
Attractively, an abundant rare earth element, lanthanum (La), is a
Fe2O3 10.905
promising candidate for the modification of palygorskite due to its MgO 8.027
excellent affinity with phosphate, along with inexpensive and en- CaO 7.660
vironmental-friendly nature comparing with Cerium and Scandium. K2O 2.224
TiO2 1.961
Firsching and Kell (1993) reported the solubility product, kps, of lan-
thanum phosphate as 10−25 at 298 K, implying lanthanum phosphate
tended to be the most insoluble compound within all the rare earth-
study. The bulk chemical constitution was revealed in Table 1. The
phosphate complexes. He et al. (2015) also found lanthanum phosphate
phosphate stock solution was obtained by dissolving analytically pure
was able to form even in solutions with low concentrations and low pH
potassium dihydrogen phosphate (KH2PO4) purchased from Benchmark
values. However, La particle agglomeration and pore blockage are still
Chemistry Co. Ltd. (Tianjin, China) in deionized water. Lanthanum
major obstacles in the application of current La-based phosphate ad-
nitrate (La(NO)3), sodium nitrate (NaNO3), sodium bicarbonate
sorbents. Notably, Yang et al. (2012) reported host materials with
(NaHCO3), sodium chloride (NaCl), sodium sulfate (Na2SO4), sodium
meso/macro-pores and cross-linked framework were ideal candidates
hydroxide (NaOH), hydrochloric acid (HCl), ascorbic acid and mo-
for lanthanum incorporation as the macro-porous structure allowed the
lybdenum acid ammonium were purchased from Guangfu Fine Che-
accessibility of phosphate for the adequate crystallization of lanthanum
micals Co. Ltd. (Tianjin, China). Lanthanum hydroxide (La(OH)3) was
phosphate without triggering pore blockage. In that case, lots of effort
purchased from Macklin Co. Ltd. (Shanghai, China). All chemical re-
has been put into embedding lanthanum oxide/hydroxide onto syn-
agents used in this study were analytical grade.
thetic meso/macro-porous polymers, such as polystyrene and carbon
nanofiber (Zhang et al., 2012, 2016). Interestingly, palygorskite clay
has immanent meso/macro-pores constructed by cross-linked crystal-
2.2. Preparation and characterization of adsorbents
line fibrous networks, providing ideal structure for lanthanum loading
without extra structural modification. In addition, compared with those
The raw palygorskite (PAL), was calcined in an electric furnace at a
synthetic organic meso/macro-porous materials, the reserves of paly-
range of set temperatures (500 °C, 600 °C, 700 °C, 800 °C and 900 °C)
gorskite clay are abundant in plenty of regions around the world and
over a series of retention periods (2 h, 3 h, 4 h and 5 h). Among them,
the price is extremely cheap (Alvarezayuso and Garcíasánchez, 2003).
the calcined sample with the best phosphate sorption capacity was se-
Moreover, the abundant porosity provided by palygorskite ensure the
lected for the following modification and was named as HPAL. La(OH)3
uniform distribution of lanthanum loaded on the surface and inside the
suspension was prepared by dropwise addition of 1 mol/L NaOH solu-
networking pores, avoiding lanthanum particle agglomeration. As ac-
tion into 50 mL 0.1 mol/L La(NO)3 solution until pH value reached 10
cessible functional sites for phosphate are introduced by lanthanum
(Zhang et al., 2012). Afterwards, the suspension was immediately
incorporation, the adsorption efficiency of lanthanum palygorskite
treated with 130 W ultrasonic power for 5 min for sufficient dispersion.
hybrid adsorbent is expected to be prominent. Thus, a La-based layer
Then, 1 g HPAL was added followed by 10 min ultrasonic treatment at
coated on palygorskite surface would be a tailor-designed approach for
130 W. Subsequently, the mixture was heated at 60 °C for 6 h in a water
the preparing a novel hybrid adsorbent for P removal.
bath to ensure the loaded La effectively bound to the host. Afterwards,
As few previous studies have reported raw porous palygorskite clay
the La(OH)3 saturated palygorskite was filtered out and rinsed with
supported lanthanum hydroxide for advanced phosphate removal, this
deionized water until neutral pH was reached, and then air-dried at
study aims at fabricating a novel palygorskite-based adsorbent HPAL-
106 °C for 48 h. Finally, the hybrid adsorbent HPAL-LaOH was ob-
LaOH with high phosphate adsorption capacity and fast kinetics. The
tained. For comparison, a sample without introducing La (denoted
lanthanum hydroxide is expected to exert the dominant specific ad-
HPAL-NaOH) was prepared using the similar procedures, but in the
sorption to phosphate while the palygorskite plays as a host to provide
absence of La(NO)3, where the HPAL was only treated in NaOH aqueous
abundant binding sites for the deposit. It is well known that thermal
solution.
treatment has a significant effect on the structure of palygorskite,
Surface morphologies of the adsorbents were analyzed using a field
leading to the release of both more calcium and magnesium ions for
emission scanning electron microscope (FESEM, Quanta FEG 250,
enhanced phosphate capture and more active groups for the firm
USA). Crystalline phases were characterized by employing X-ray dif-
combination of the lanthanum hydroxide shell layer and palygorskite
fraction (XRD, Rigaku D/MAX-2000, Japan) using a monochromator
core via chemical bonds. Thus, the raw palygorskite was modified by
with the irradiation source of Cu Kα (40 kV, 25 mA) and wavelength of
calcination activation before embedding with La(OH)3. Therefore, the
1.54060. The scanning range and step were 10–80° and 0.02°, respec-
objectives of this study include: (1) synthesizing lanthanum hydroxide-
tively. Fourier transformation infrared spectrometer (FT-IR, Perkin
impregnated palygorskite (HPAL-LaOH); (2) characterizing the raw
Elmer, USA) was employed to detect the specific functional groups of
palygorskite (PAL), calcined palygorskite (HPAL) and HPAL-LaOH; (3)
adsorbents by using dry potassium bromide for background subtraction.
revealing the phosphate adsorption behavior and mechanism of HPAL-
The nitrogen N2 adsorption-desorption test was conducted at liquid
LaOH as a function of solution pH and co-existing ions; (4) evaluating
nitrogen temperature using a gas adsorption analyzer (ASAP 2020M,
the settleability and stability of HPAL-LaOH after phosphate adsorp-
Micromeritics, USA), and then the specific surface area and average
tion.
pore diameter were analyzed accordingly. The interactions between
HPAL-LaOH and phosphate were probed from X-ray photoelectron
2. Materials and methods
spectroscopy (XPS, PHI 5700 ESCA System, USA) using a monochro-
matic rotating anode Al Kα excitation source with pass energy of 30 eV.
2.1. Materials
The La content of HPAL-LaOH was measured by X-ray fluorescence
spectrometer (XRF, PW4400, Netherland).
Raw palygorskite clay was procured from Anbang Minerals Co., Ltd.
(Anhui, China), and the < 200 mesh particles were selected for this

508
L. Kong et al. Applied Clay Science 162 (2018) 507–517

2.3. Batch phosphate adsorption and desorption experiments

An aliquot of phosphate stock solution (1000 mg P/L) was mixed


with a volume of deionized water to obtain the phosphate solutions
with expected concentrations. In the batch phosphate adsorption pro-
cess, 0.2 g adsorbent was added into 100 mL Erlenmeyer flasks con-
taining 50 mL KH2PO4 solution with a series of initial concentrations (5,
10, 20, 50, 100, 150, 200, 300, 400, 500, 600, 800 and 1000 mg/L).
These flasks were then shaken in thermostatic air bath shaker under
200 rpm for 48 h. Afterwards, the equilibrium phosphate concentration
was analyzed after filtering through a 0.45 μm membrane filter. The
adsorption capacity of individual sample was obtained due to the var-
iation of phosphate concentration as described in Eq. S1.
In the adsorption kinetics studies, 0.5 g adsorbent was added into
200 mL of 1000 mg/L phosphate solution, the following adsorption
process was performed as aforementioned procedure. At designated
sampling times (2 min, 5 min, 10 min, 30 min, 1 h, 2 h, 4 h, 6 h, 8 h,
10 h, 12 h, and 24 h), a 2 mL of solution was taken out of the flask for
the analysis of phosphate concentration. After adsorption equilibrium,
30 mL sample was collected each time at designated intervals (0 min,
Fig. 1. Effect of calcination temperature and holding time on the phosphate
30 min, 1 h, 2 h, and 3 h) by using a syringe sucking from the surface of
adsorption capacity of raw palygorskite (C0 = 1000 mg P/L, adsorbent
the solution to investigate the settling behaviors of diverse adsorbents.
dose = 4 g/L, pH = 6.5–7.0, contact time = 24 h, T = 25 °C).
Subsequently, the turbidity and zeta potential of each sample were
measured by using a turbidimeter (Turb 550, WTW, Germany) and
Zetasizer (Nano-Z, Malvern, UK), respectively. The contaminated ad- 3. Results and discussion
sorbents were regenerated by 50 mL of 0.1 M NaOH solution under
strong stirring for 30 min, then washed with deionized water several 3.1. Thermal activation of raw palygorskite
times and dried for the next adsorption test. Five consecutive adsorp-
tion-desorption experiments were conducted to investigate the reusa- Raw palygorskite was calcinated in diverse conditions and the effect
bility of the as-prepared adsorbents. of calcination on phosphate uptake was shown in Fig. 1. The phosphate
The pH effect on the phosphate sequestration was evaluated by adsorption capacity of the raw palygorskite (PAL: 7.87 mg/g) was in-
adjusting the initial pH of 50 mL 500 mg/L phosphate solution in the creased once heated over 600 °C. The sample calcined at 700 °C ex-
range of 3–11 with the aid of 1 mol/L HCl and 1 mol/L NaOH. The pH hibited the best adsorption performance, followed by the samples he-
values were determined by a high performance pH meter (FE20, Mettler ated at 800 °C and 900 °C. The results were consistent with the findings
Toledo, Switzerland). A mixture composed of KH2PO4, NaNO3, Na2SO4, of previous study (Gan et al., 2009). Apart from the calcining tem-
NaCl and Na2CO3 was prepared to investigate the phosphate seques- perature, the holding time also had a significant impact on the ad-
tration performance of HPAL-LaOH in the presence of competitive ionic sorption performance. For the samples calcined at 700 °C, the increase
at different strengths. of phosphate adsorption capacity kept pace with the extension of the
The concentration of phosphate was measured by the method of holding time from 2 h to 4 h, which was mainly attributed to the en-
molybdenum-blue ascorbic acid using a UV–vis spectrophotometer (T6, hanced effect of irreversible dehydration and dehydroxylation in the
PGENERAL, China). All the batch adsorption results are the average crystal structure of palygorskite, causing more calcium, magnesium and
values of three replicate experiments. aluminum cations with high affinity for phosphate releasing from the
crystal matrix (Gan et al., 2009). As the holding time increased over 4 h,
the crystal structure of palygorskite collapsed substantially, the de-
2.4. Evaluation of stability of La loaded on HPAL-LaOH composition of lattice led to the destruction of porous structure, which
might interpret the decreasing adsorption capacity as adsorbate was
The stability of La loaded on the HPAL-LaOH was evaluated before restricted to pass though the blocked pores to access interior adsorption
and after phosphate adsorption process. By digesting 0.2 g HPAL-LaOH sites. To this end, the sample heated at 700 °C for 4 h with the highest
with 50 mL of HNO3 (30%) solution for one day to ensure complete phosphate adsorption capacity of 33.13 mg/g was selected for the
lanthanum dissolution, the total amount of La impregnated on HPAL- subsequent experiments and named as HPAL.
LaOH was determined. La leaching from HPAL-LaOH was obtained by
adding the same amount HPAL-LaOH in 50 mL deionized water and 3.2. Characterization of as-prepared adsorbents
treating it with ultrasound challenge for 30 min at the power of 300 W.
Then a sucking filter was utilized to separate the possible released La The morphology of PAL, HPAL, and the resultant La-based hybrid
(OH)3 particles with a 1 μm filtration membrane under a filtration adsorbent HPAL-LaOH was shown in Fig. 2. The raw palygorskite de-
pressure of 0.1 bar (T-50, Jingteng, China). The mass of particles in the monstrated a reticular structure constructed by rod-like fibers with the
filtrate was analyzed after filtering through a 0.22 μm membrane filter diameter of 5–40 nm (Fig. 2a). The closely-packed arrangement of the
and air drying. The amount of La dissolution after adsorption was in- elementary fiber contributed to the abundant meso/macro-porosity of
vestigated by detecting La concentration in the residual phosphate so- the clay. When the raw palygorskite was heated under 700 °C for 4 h, no
lution. The concentrations of La in these solutions were measured evident changes in the structure of individual fibers of palygorskite
through inductively coupled plasma-optic emission spectrometer (ICP- occurred except for the presence of sintering in some places (Fig. 2b).
OES, PerkinElmer Optima, USA). However, the space between the layers was reduced because of irre-
versible dehydration and dihydroxylation (Gan et al., 2009). After
loaded with La(OH)3, the individual fibers of palygorskite matrix were
coated with La(OH)3 particles leading to the reduction of the proportion
of macropores. However, the abundant porosity remained recognizable

509
L. Kong et al. Applied Clay Science 162 (2018) 507–517

Fig. 3. XRD patterns of (a) PAL, (b) HPAL, (c) HPAL-LaOH and (d) HPAL-LaOH
after phosphate adsorption (*: palygorskite, ◇: quartz, ●: dolomite, △: cal-
cium magnesium silicate, ■: La(OH)3, ▼: LaPO4·xH2O).

adsorbed water dwindled (Fig. 3b) (Shuali, 1990), and the reflections of
dolomite all disappeared. However, moderate reflections related to
calcium–aluminum silicate and magnesium oxide emerged (JCPDS
No.45–0946), which was attributed to the decomposition of dolomite
under high temperature and the interaction of the released calcium
oxide with crystalline SiO2. In the pattern of HPAL-LaOH (Fig. 3c),
distinct peaks were observed at 15.6°, 27.9°,39.4° and 48.6° (d-value of
5.65 Å, 3.18 Å, 2.28 Å and 1.87 Å, respectively), revealing characteristic
diffraction of hexagonal phase La(OH)3 (JCPDS No.36–1481), which
implied the effective impregnation of La(OH)3 on the substrate. In ad-
dition, the peaks belonging to palygorskite were dramatically weakened
along with substantial reduction of intensities of crystalline quartz
phases which used to remain the same after calcination, which was
probably attributed to the interaction with NaOH in the preparing so-
lution.
FT-IR spectra of different adsorbents were presented in Fig. 4. PAL
demonstrated typical bands at 3615, 3584, 3550, 3418 and 1648 cm−1
corresponding to the structural OeH and physically adsorbed water of
palygorskite (Fig. 4a) (Chahi et al., 2002; Mckeown et al., 2002).
However, in the patterns of HPAL (Fig. 4b), no residues of hydroxyl

Fig. 2. SEM photos of (a) PAL, (b) HPAL, (c) HPAL-LaOH.

indicating La(OH)3 particles were well dispersed on the surface of


HPAL-LaOH (Fig. 2c). No apparent aggregation was observed owing to
the strong shearing of ultrasonic action. Most of La was distributed in
the outer region of HPAL-LaOH, which was favorable for effective
phosphate ionic adsorption.
The XRD patterns of PAL, HPAL, HPAL-LaOH were shown in Fig. 3.
The raw sample was comprised of major mineral consortia of paly-
gorskite, dolomite and quartz (Fig. 3a). The diffraction at 2θ = 8.3°
represented the characteristic diffraction of the raw palygorskite, cor-
responding to the (110) plane of the crystal (d-value = 10.47 Å). Some
moderate diffractions were also detected at 4.48, 4.15, 3.65 and 3.22 Å,
which coincided with the characteristic patterns in previous studies (Ye
et al., 2006; Ma et al., 2014). After calcined under 700 °C, the position Fig. 4. FT-IR spectra of (a)PAL, (b) HPAL, (c) HPAL-LaOH, and (d) HPAL-LaOH
of the 110 spacing of palygorskite exhibited to alter as the content of after phosphate adsorption.

510
L. Kong et al. Applied Clay Science 162 (2018) 507–517

groups related bands, along with the ones related to CO32– groups of
dolomite phase at 2531, 1819 and 1452 cm−1, could be observed.
These results imply that under calcination conditions, coordinated
water molecules in the pores were all removed and dolomite was de-
composed completely and converted to CaO and MgO (Suárez and
Garcia, 2006). However, the sharp bands at 1024 and 467 cm−1 con-
tributed by the SieO and Si–O–Si bending, respectively, remained their
intensities (Blanco et al., 1989). These observations were in agreement
with the findings in the XRD spectra indicating that crystallinity of
quartz was scarcely influenced by heating. For the spectrum of HPAL-
LaOH (Fig. 4c), the characteristic band at 3609 cm−1 belonging to the
hydroxyl groups of lanthanum hydroxide was detected (González-
Rovira et al., 2008). The noticeable sharp band at 1385 cm−1 was
corresponded to the vibration mode of NO3− anions which were in-
tercalated in the deposit structure during the chemical preparation
process (Zhang et al., 2012). Notably, the absorption bands centered at
3428 and 1632 cm−1 appeared again corresponding to water molecules
(coordinated and zeolitic water). In addition, the bands at 1484 cm−1
were attributed to the carbonate group originating from the reaction of
La(OH)3 with CO2 from air. The distinct band, located at 658 cm−1, was
assigned to characteristic La-OH bond in La(OH)3 (Aghazadeh et al.,
2011, 2014). It is worth noting that the as-prepared HPAL-LaOH reveals
a good coincidence in bibliographies with respect to the La(OH)3 re-
lated peaks (Aghazadeh et al., 2011; Zhang et al., 2012, 2016). The
results of XRD and FT-IR analyses imply that the as-prepared La(OH)3
has been potently incorporated onto the palygorskite matrix.
Nitrogen adsorption–desorption isotherms of the samples before and
after modification were compared in Fig. 5a. The N2 adsorption curves
for three samples all exhibited a type-IV isotherm with a H3 hysteresis
loop as described by IUPAC, revealing that palygorskite has evident slit-
shaped pores originating from aggregation of plate-like particles
(Pierotti and Siemieniewska, 1985). In the BJH desorption pore volume
plots (Fig. 5b), the pore volume inflected sharply in the pore diameter
range of 3–5 nm, these sharp peaks implied that mesopores were pro-
minent in those samples. It was observed that encapsulation of La(OH)3
resulted in reduction of the proportion of both < 3 nm and 5–10 nm
pores, while those of 3–5 nm almost remained the same. Additionally,
according to the results of pore structure parameters (Table 2), micro-
pore area and volume dramatically increased from 10.54 to 19.65 m2/g
and from 0.004 to 0.009 cm3/g, respectively. The creation of micro-
porosity probably originated from the opening of blocked pores during
the ultrasonic and chemical process, as well as the dispersion of La
Fig. 5. Nitrogen adsorption-desorption data of PAL, HPAL, and HPAL-LaOH. (a)
(OH)3 nano-scale particles inside the meso/macro-pores of paly-
N2 adsorption-desorption isotherms (b) pore volume distribution plots.
gorskite. The specific surface area of HPAL-LaOH (73.32 m2/g) re-
mained almost the same as that of the HPAL sample. It might be in-
terpreted that the chemical process of La(OH)3 impregnation did not Table 2
result in the collapse of the pore structure of palygorskite. According to Pore structure parameters calculated from the N2 adsorption-desorption iso-
the results of characterization of the three samples, their structural therms (a: BET, Brunauer, Emmett, and Teller; b: t-Plot; c: BJH, Barrett, Joyner,
and Halenda).
changes were depicted in Fig. 6.
Samples Specific surface Micropore Pore Micropore
3.3. Phosphate adsorption and desorption behavior area(a) (m2/g) area(b) volume(c) volume(b)
(m2/g) (cm3/g) (cm3/g)

Phosphate adsorption isotherms were summarized in Fig. 7 and PAL 95.21 14.58 0.203 0.006
Table 3. These five samples all demonstrated enhanced phosphate ad- HPAL 73.01 10.54 0.246 0.004
sorption with the increase of initial concentrations of phosphate solu- HPAL-LaOH 73.32 19.65 0.215 0.009
tions and gradually reached saturation. However, HPAL-LaOH ex-
hibited the highest adsorption capacities (109.63 mg/g), over 13 times
while the encapsulated lanthanum hydroxide allows more active ad-
of that of raw palygorskite (7.87 mg/g) and 3 times of that of HPAL
sorption sites and more suitable conditions for enhanced phosphate
(33.13 mg/g). Note that alkaline activation only slightly improved the
removal. The price of raw palygorskite clay is much cheaper than that
capacity of HAPL by about 30% as the adsorption capacity of HPAL-
of La(OH)3 (about 1:500), so coating La(OH)3 on the surface of paly-
NaOH turned out to be 43.73 mg/g. The pure La(OH)3 sample showed a
gorskite provides an elegant approach for the utilization of La-based
much lower phosphate retention capacity (69.64 mg/g) than that of
material for phosphate removal. Note that for the solutions with low
HPAL-LaOH, verifying the synergistic effect of palygorskite matrix on
phosphate concentrations (5, 10, 20, and 50 mg/L) HPAL-LaOH could
La(OH)3 incorporation for better phosphate capture performance.
achieve almost complete phosphate uptake with the effluent con-
These results indicated that the palygorskite matrix with large surface
centration of < 0.01 mg/g, which could adequately meet the effluent
area exhibits a positive effect on the distribution of La(OH)3 particles

511
L. Kong et al. Applied Clay Science 162 (2018) 507–517

Fig. 6. Schematic illumination of the structural evolution of PAL, HPAL and HPAL-LaOH.

Table 4
Comparison of phosphate adsorption capacity of different adsorbents in the last
decade.
Adsorbent qm (mg/g) References

Raw palygorskite clay 10.9 (Gan et al., 2009)


Raw palygorskite clay 7.87 This study
Thermal activated raw palygorskite 42.0 (Gan et al., 2009)
Fe(III)-modified bentonite 11.5 (Zamparas et al., 2012)
Modified bentonite Bephos® 26.5 (Zamparas et al., 2013)
Ferrihydrite-modified diatomite 37.3 (Xiong and Peng, 2008)
Magnesium hydroxide modified 45.7 (Mitrogiannis et al., 2017)
Diatomite
Coal fly ash treated with NaOH 57.1 (Pengthamkeerati et al.,
2008)
Fe loaded skin split waste 68 (Huang et al., 2009)
La(III)-modified pillared 10 (Tanada et al., 2003)
montmorillonite
La(III)-modified bentonite 14 (Kuroki et al., 2014)
La(III) loaded orange waste 14.0 (Biswas et al., 2007)
Lanthanum hydroxide-doped 15.3 (Zhang et al., 2012)
activated carbon fiber
Fig 7. Phosphate adsorption isotherms of PAL, HPAL, HPAL-NaOH, La(OH)3 Lanthanum loaded mesoporous silica 45.63 (Yang et al., 2011)
and HPAL-LaOH at 25 °C with results fitted to the Langmuir and Freundlich SBA-15
equations (Ce: equilibrium concentration, T: 25 °C, pH: 6.5–7.0). La(OH)3 modified exfoliated 79.60 (Huang et al., 2014)
vermiculites
HPAL-LaOH 109.63 This work
discharge standard of China GB18918–2002 (0.5 mg P/L) and the reg-
ulation of Water Framework Directive in European Union where the
permissible concentration of P in discharge was limited to 0.1 mg/L (Li Table 3. By comparing the R2 and SE values, the adsorption data of
et al., 2018). It should be remarkable that the phosphate adsorption HPAL-LaOH agreed well with the Freundlich isotherm, indicating that
capacity of HPAL-LaOH was fairly competitive with several low-cost multilayer adsorption on heterogeneous surfaces takes into account in
clay mineral adsorbents synthesized recently and the values were the sorbent-sorbate interactions between metallic ions and phosphate
summarized in Table 4. Notably, the raw palygorskite clay sample used (Mitrogiannis et al., 2017).
in this study showed inferior adsorption capacity comparing with that Phosphate adsorption kinetics were depicted in Fig. 8 and Table 5.
in previous study (Gan et al., 2009), probably due to the relative im- Obviously, there was a great gap between the palygorskite-based and
purity and low palygorskite phase content of the raw palygorskite La-based adsorbents on adsorption rate, namely, HPAL-LaOH and La
material used in this study. It suggests that by employing the proposed (OH)3 sample reached equilibrium within 2 h contact time while it took
preparation method in this study, the phosphate adsorption capacity of HPAL and PAL over 4 h to do so. The kinetic data was also fitted well to
palygorskite-based adsorbent has the potential to be further advanced. pseudo-first-order and pseudo-second-order models (Table 5). The
The adsorption behaviors were described by Langmuir and Freundlich higher correlation coefficients (R2) for a pseudo-second-order equation
isotherm models using non-linear regression method. The relevant for HPAL-LaOH implied that chemisorption played a dominant role in
formulas and fitting results were summarized in Eqs. S2, S3 and the adsorption process (Zhao et al., 2016). The rate constants (k2) of La

Table 3
Parameters of Langmuir and Freundlich isotherms for each sample.
Sample Langmuir isotherm Freundlich isotherm

2
KL Qm R SE Kf n R2 SE
(L/mg) (mg/g)

PAL 0.01 13.94 0.9908 0.0709 0.09 1.54 0.9962 0.0454


HPAL 0.02 33.57 0.8950 1.0791 7.63 4.49 0.9807 0.4628
HPAL-LaOH 0.08 99.01 0.9143 3.4189 32.17 5.15 0.9652 2.1797
La(OH)3 0.17 66.17 0.9489 1.7884 17.04 4.31 0.8603 2.9555
HPAL-NaOH 0.02 44.24 0.9421 1.0892 9.13 4.16 0.9843 0.5667

512
L. Kong et al. Applied Clay Science 162 (2018) 507–517

Fig. 8. Phosphate sorption kinetics of PAL, HPAL, HPAL-NaOH, La(OH)3 and HPAL-LaOH with results fitted to pseudo-first-order and pseudo-second-order models
(C0: 1000 mg/L, T: 25 °C, pH: 6.5–7.0).

(OH)3 and HPAL-LaOH at 0.28 and 0.19 g/(mg·min), respectively, were


much greater than those of PAL, HPAL, and HPAL-LaOH at 0.10, 0.07
and 0.06 g/(mg·min), respectively, revealing that the introduction of La
contributed to the dramatic increase of adsorption speed of phosphate.
Five consecutive adsorption-desorption experiments were con-
ducted for the contaminated HPAL-LaOH. The good reusability of
HPAL-LaOH was realized since the phosphate removal rate only de-
creased by 10% after four adsorption-desorption cycles comparing with
the initial removal efficiency (Fig. S1). In addition, the concentrations
of La released from HPAL-LaOH were < 0.02% during the adsorption
(Table 7), indicating the satisfactory regeneration and stability of
HPAL-LaOH for phosphate removal from wastewater.
The pH value effect on HPAL-LaOH was revealed in Fig. 9. With
respect to the phosphate solutions below 100 mg/L, HPAL-LaOH could
achieve almost complete phosphate removal, regardless of pH variation
in a large given scope of 3–11, which is fairly favorable for the appli-
cation of HPAL-LaOH to treating extremely acid and alkaline waste-
water even at low phosphate concentration. The point of zero charge
(PZC) of HPAL-LaOH was around 10.5, which can interpret the wide
pH-independent phosphate removal because HPAL-LaOH was posi- Fig. 9. Effect of initial pH on phosphate adsorption efficiency of HPAL-LaOH
(adsorbent dose: 2 g/L, contact time: 24 h, T: 25 °C).
tively charged at pH < 10.5, favoring the adsorption of negatively
charged phosphate ions. Concerning the 200 mg/L phosphate solution,
HPAL-LaOH exhibited a moderate evolution of phosphate adsorption at adsorbents (Kuroki et al., 2014; Zhang et al., 2016). The pH de-
acidic pH of 3–7 and the highest phosphate adsorption (91.78 mg/g) pendency of phosphate adsorption could be illuminated in terms of the
occurred at pH value of 6. Then, the phosphate sequestration dropped dissolution behavior of immanent metal cations from internal paly-
gradually until the ending stage of pH growth. Similar tendency was gorskite and the interaction of encapsulated lanthanum active groups
found from previous reports about phosphate adsorption on other with the phosphate ions in aqueous solution. It was well known that

Table 5
Parameters of pseudo-first-order and pseudo-second-order models for each sample.
Samples Pseudo-first-order Pseudo-second-order

2
k1 qe R SE k1 qe R2 SE
(min−1) (mg/g) (min−1) (mg/g)

PAL 0.62 7.21 0.9839 0.1104 0.10 8.08 0.9972 0.1104


HPAL 1.46 30.80 0.9731 0.5711 0.07 32.81 0.9912 0.5711
HPAL-LaOH 12.25 105.48 0.9676 1.7117 0.19 108.32 0.9967 1.7117
La(OH)3 12.67 63.77 0.9481 1.4345 0.28 65.87 0.9373 1.4345
HPAL-NaOH 1.46 36.97 0.9731 0.6853 0.06 39.37 0.9912 0.6853

513
L. Kong et al. Applied Clay Science 162 (2018) 507–517

precipitation played a significant role in the retention of phosphate disappeared, and two distinct diffractions representing LaPO4 were
through combination with calcium and magnesium cations from solid observed, fairly consistent with the data of previous studies
colloids (Ruixia et al., 2002). However, the concentrations of Ca and Mg (Mendelovici, 1973). As there were no evident changes of crystal phases
ions released from the surface of adsorbents dwindled with the growth related to metal compounds from palygorskite, it could be concluded
of the pH of the solution (Gan et al., 2009). In that case the effect of that La(OH)3 on the surface of HPAL-LaOH was dominantly converted
immobilization of phosphate by metallic cations was alleviated, leading in situ into LaPO4. FT-IR spectra also provided an insight into the
to the reduction of the phosphate removal efficiency. Meanwhile, at the phosphate species formed inside HPAL-LaOH (Fig. 4d). After phosphate
solid–liquid interfaces, the phosphate ions adsorbed via electrostatic adsorption saturation, the band at 3609 cm−1 belonging to hydroxyl
attraction and ligand exchange were susceptible to the variation of the groups of La(OH)3 disappeared, implying the dissolution of La(OH)3 in
solution pH (Zhang et al., 2012). The species of hydroxide groups the adsorption process. In addition, the NO3− and CO32−anions which
loaded on HPAL-LaOH altered as a function of pH due to protonation used to respond at 1484 and 1385 cm−1, respectively, were both dis-
and deprotonation effects as described in the following equations: placed from the adsorbent as a result of ligand exchange by PO43−. In
contrast, the sharp bands at 614 and 539 cm−1 corresponding to O-P-O
≡ La − OH + H+ ↔ La − OH+2 pH < 8 (1) bend vibration were detected (Li et al., 2008), along with the strong
absorption peak at 1051 and 1008 cm−1 attributed to vas and vs
≡ La − OH − H+ ↔ La − O− pH > 8 (2)
stretching vibration of PeO of HPO42− and PO43− group, respectively
At acid pH, ligand exchange between the hydroxide groups loaded (Xie et al., 2014c). These results indicated La(OH)3 species on the
on HPAL-LaOH and phosphate played a dominant role in the highly- composite adsorbent HPAL-LaOH offered prominent specific adsorption
efficient adsorption process. Simultaneously, the acidic medium trig- toward phosphate through surface LaeP complexation.
gered HPAL-LaOH to protonate (Eq. (1)), and –OH2+ was easier to be More evidence about the mechanism of phosphate adsorption onto
replaced than –OH on the surface HPAL-LaOH by phosphate ions HPAL-LaOH was gained through XPS analysis. The survey spectra of
(Zhang et al., 2012). The positively charged surface could also facilitate HPAL-LaOH after phosphate saturation were shown in Fig. 10(a), the
electrostatic attraction of anionic phosphate (Chubar et al., 2005). presence of a P 2p peak revealed phosphate was adsorbed onto the
Thus, it is reasonable that acid pH is favorable for phosphate seques- surface of HPAL-LaOH. Two sharp peaks corresponding to La 3d3/2 and
tration. However, at alkaline pH region, the deprotonation effect of La- La 3d5/2 exhibited much stronger intensity than those moderate peaks
OH was strengthened (Eq. (2)), leading to the enhancement of elec- related to Ca, Mg, Al and Fe elements, which verified that lanthanum
trostatic repulsion between the negatively charged surface of ad- played a significant role among all the metals on HPAL-LaOH in the
sorbents (LaeOe) with the phosphate anion. Meanwhile, competition phosphate removal. The P 2p spectrum after phosphate adsorption sa-
adsorption between phosphate and hydroxyl anions was also ag- turation can be divided into two representative peaks (Fig. 10b). The
gravated, which exacerbated the low adsorption performance. Lewis binding energy of the phosphate captured by La(OH)3 and palygorskite
acid–base interaction, which induced coordination complexes between were centered at 132.6 eV and 131.9 eV, respectively, consistent with
the oxygen anions of phosphate and the lanthanum active sites became the values in previous studies (Zhang et al., 2016; Koilraj and Sasaki,
dominant in this condition (Zhang et al., 2012). 2017). The higher binding energy of LaeP interactions revealed that
Concomitant anions such as SO4−, NO3−, HCO3– and Cl− are often the loaded La(OH)3 had a stronger affinity to phosphate than paly-
found in natural water bodies, which can influence phosphate uptake gorskite. Additionally, the corresponding area fractions of phosphate
by competing for adsorption sites. Thus, the phosphate adsorption adsorbed by La(OH)3 and palygorskite were 54.1% and 38.9%, re-
performance of HPAL-LaOH in mixed solutions in the presence of SO4−, spectively, which was close to the ratio of the phosphate adsorption
NO3−, HCO3– and Cl− was investigated, and the results were illustrated capacity in batch adsorption experiment of Section 3.3. Furthermore,
in Table 6. It demonstrated that increasing the concentrations of in- the species of phosphate on HPAL-LaOH were also investigated. The
dividual competing ion from 5 mg/L (molar ratio: 1:0.1) to 500 mg/L species of phosphate ions present at different pH were based on the
(molar ratio: 1:10) only resulted in a slight interfering effect on HPAL- following formula (Xie et al., 2014c).
LaOH's phosphate removal efficiency and over 90% of phosphate anions H3 PO4 ↔ H2 PO−4 + H+ pK1 = 2.13 (3)
were removed even in the co-existence of four kinds of high con-
centration competing ions. This results indicate the high preference H2 PO−4 ↔ HPO24− + H+ pK2 = 7.20 (4)
toward phosphate sequestration of HPAL-LaOH in complexed solutions.
HPO24− ↔ PO34− + H+ pK3 = 12.33 (5)
3.4. Adsorption mechanism
According to the given pH of the experiment conditions, HPO42−
and PO43− were available in the solution. In that case, LaPO4 and
A combination of detection methods including XRD, FT-IR and XPS
La2(HPO4)3 tended to form during the adsorption process. The high
were employed for structural diagnosis of HPAL-LaOH before and after
resolution spectra of La 3d5/2 showed that LaPO4 and La2(HPO4)3 were
the phosphate adsorption. The XRD pattern of HPAL-LaOH after phos-
centered at 835.3 eV and 839.1 eV (Fig. 10c), with their satellite peaks
phate sorption was given in Fig. 3d, the crystal state of HPAL-LaOH
at 838.6 eV and 836.0 eV, respectively, indicating the fact that the in-
changed such that four diffractions belonging to La(OH)3 all
teractions between surface La(OH)3 on the HPAL-LaOH and adsorbed
phosphate could in situ form the amorphous structure consortia of
Table 6
LaPO4 and La2(HPO4)3.
Effect of concomitant anions on phosphate adsorption onto HPAL-LaOH (25 °C,
C0 = 50 mg P/L, adsorbent dosage = 0.6 g/L, reaction time = 24 h).
From the above analysis, the removal of phosphate by HPAL-LaOH
is dominantly attributed to strong inner-sphere complexation by ligand-
Molar ratio of phosphate Concentration of individual Phosphate removal exchange between highly dispersed La(OH)3 layer on the surface of
to competing ions competing ion (mg/L) ratio (%)
palygorskite and phosphate in solutions. Starting with the adding of
0 0 99.45 HPAL-LaOH with high specific surface area into the phosphate solution,
1:0.1 5 97.20 the surface La(OH)3 layer could continuously capture phosphate from
1:1 50 97.16 solutions by ligand-exchange, then more hydroxide groups were re-
1:2 100 96.35
leased resulting the increase of pH of the solution. Due to the electro-
1:5 250 94.37
1:10 500 90.62 static repulsion, the negatively charged adsorbent surface sites do not
favor the adsorption of phosphate anion. The pH dependent adsorption

514
L. Kong et al. Applied Clay Science 162 (2018) 507–517

indicates that the adsorption is dominated by surface complexation.


Under the experimental conditions, H2PO42− and PO43− can be
formed. These surface complexed PO4 species onto HPAL-LaOH could
in situ react with La(OH)3 and convert to the noncrystal structure
compound of LaPO4 and La2(HPO4)3. The proposed sorption me-
chanism of HPAL-LaOH for specific phosphate adsorption from aqueous
solution is shown in Fig. 11 and Eqs. 6, 7.
HPAL ≡ La − (OH)3 + PO34− → ≡LaPO4 + 3OH− (6)

HPAL ≡ La − (OH)3 + HPO24− → ≡La2 (HPO4 )3 + OH− (7)


According to the results of competitive adsorption, HPAL-LaOH's
high adsorption capacity was inconspicuously affected when co-existing
anions were present in the mixture. These observations suggest that
electrostatic attraction only plays an insignificant role in the phosphate
capture process.

3.5. Settleability of modified palygorskite adsorbents

The particle settleability of PAL, HPAL and HPAL-LaOH was eval-


uated by quantifying the turbidity of the phosphate solutions after sa-
turation as a function of 3 h settling time and the results were illustrated
in Fig. 12. Once the settling started, the rapid settling behaviors of three
palygorskite samples were all observed. After 0.5 h settling, HPAL and
HPAL-LaOH demonstrated much better settling performance with the
turbidity of 62.8 and 72.9 NTU, respectively, against that of PAL (170.2
NTU). During the following settling process, turbidity of HPAL and
HPAL-LaOH samples remained superior to that of untreated sample. It
was observed that the settling evolution of HPAL and HPAL-LaOH
ended within 2 h, while the settling process of PAL was still undergoing
after 3 h due to the continuous downtrend of turbidity.
In order to elucidate the settling behaviors of modified samples, zeta
potentials were measured to provide information regarding the particle
charges. As shown in Fig. 13, zeta potentials of PAL ranged between
−18.7 and −20.53 mV during the entire 3 h settling period, which
remained greater than those of HPAL (−13.37 to −19.03 mV) and
HPAL-LaOH (−12.37 to −15.1 mV). The above observations in both
turbidity and zeta potentials in the supernatant suggest that negative
surface charges of palygorskite were reduced after heating, indicating
the electrostatic repulsion between these negatively charged particles
was weakened, which left them in a relatively unstable state. Thus, it
was easy for more aggregation or the formation of large domains to
occur in solution so that the settling was accelerated. A possible reason
is that after thermal treatment, the layers of the palygorskite collapsed,
and a lot of inner calcium and aluminum active groups with positive
charge, which used to be absorbed in the pores of the octahedral and
tetrahedral sheets of palygorskite fibers, emerged on the surface (Gan
et al., 2009). Of note, dosing of La(OH)3 did not alter the favorable
settleablity of heated palygorskite while the negative charged surface
was further neutralized probably due to the deprotonation of water
molecules on lanthanum active sites.

3.6. Stability of lanthanum loaded on HPAL-LaOH

The La loaded on HPAL-LaOH demonstrated an excellent stability


even after suffering from a strong ultrasonic challenge. The percentage
of La(OH)3 particles leaching from HPAL-LaOH was < 1% after ultra-
sound treatment for 30 min. As shown in Table 7, the La dissolution in
Fig. 10. XPS analysis of HPAL-LaOH after phosphate adsorption. (a) survey
the phosphate solution after adsorption saturation for 24 h was also
scan (b) P 2p spectrum (c) La 3d5/2 spectrum (adsorbent dose: 4 g/L, initial negligible. These results suggest that La loading layer was stably con-
phosphate concentration: 500 mg/L, contact time: 24 h, pH:6.5 ± 0.5, T: fined onto the host palygorskite, which was possible due to the meso/
25 °C). macro-porous palygorskite providing proper pore space for the ac-
commodation of nano-scale La particles with strong bonding strength.
The concentration of La in typical river is about 30–50 ng/L (Ma et al.,
2016), while the La leaching from HPAL-LaOH was below 20 ng/L, so it
is safe enough for the application of HPAL-LaOH for real wastewater

515
L. Kong et al. Applied Clay Science 162 (2018) 507–517

Fig. 11. Mechanism of specific phosphate uptake onto HPAL-LaOH (a) before phosphate adsorption, (b) after phosphate adsorption.

Table 7
The concentrations of La leaching from HPAL-LaOH.
Samples Concentrations (mg/L)

Total amount of La loaded on HPAL-LaOH 106.3 ± 0.5


La leaching suffering untrasonic challenge 0.02 ± 0.005
La dissolution after phosphate saturation 0.02 ± 0.005

treatment or inlake nutrient pollutant control.

4. Conclusion

The resultant HPAL-LaOH exhibited an excellent phosphate ad-


sorption capacity (109.63 mg/g) and a dramatically wide pH-in-
dependent phosphate sequestration in the pH range of 3–11, which is
favorable for purifying special high phosphorus-containing wastewater
with extreme acidity or alkalinity. Additionally, the presence of the
high-level (up to molar ratio of 10) competing ions (sulfate, nitrate,
Fig. 12. Turbidity of supernatant of phosphate solution treated by PAL, HPAL bicarbonate and chloride) was probed to have little effect on the su-
and HPAL-LaOH with different settling time. perior phosphate adsorption performance of HPAL-LaOH. The surface
complexation between the embedded lanthanum hydroxide and phos-
phate made main contribution to effective phosphate removal from
solutions. Moreover, the distinctive interwoven fibrous latticework of
raw palygorskite ensured the satisfactory settleablity and negligible
lanthanum leaching even under ultrasonic challenge. In summary, this
study first drew upon raw palygorskite with unique cross-linked fibrous
crystal structure for the stable immobilization of lanthanum hydroxide
to synthesize a novel highly-effective HPAL-LaOH for enhanced phos-
phate removal. Comparing the phosphate adsorption capacity with
several low-cost or lanthanum-based adsorbents synthesized recently,
HPAL-LaOH is fairly competitive (Table 4). Therefore, HPAL-LaOH can
serve as a promising cost-effective adsorbent aiming for phosphorus
removal from wastewater. Further studies are expected to conduct on
removing phosphate from real wastewater in fixed-bed reactors and
exploiting the utilization of the P saturated HPAL-LaOH as soil condi-
tioner.

Acknowledgements

This study was supported by the State Key Laboratory of Urban


Fig. 13. Zeta potentials of supernatant of phosphate solution treated by PAL,
Water Resource and Environment, Harbin Institute of Technology
HPAL and HPAL-LaOH with settling time.
(2017DX02) and the Applied Technology Research and Development
Program of Harbin (2017AB4AS035).

516
L. Kong et al. Applied Clay Science 162 (2018) 507–517

Appendix A. Supplementary data Mitrogiannis, D., Psychoyou, M., Baziotis, I., Inglezakis, V.J., Koukouzas, N., Tsoukalas,
N., Palles, D., Kamitsos, E., Oikonomou, G., Markou, G., 2017. Removal of phosphate
from aqueous solutions by adsorption onto ca(OH)2 treated natural clinoptilolite.
Supplementary data to this article can be found online at https:// Chem. Eng. J. 320, 510–522.
doi.org/10.1016/j.clay.2018.07.005. Pan, B., Wu, J., Pan, B., Lv, L., Zhang, W., Xiao, L., Wang, X., Tao, X., Zheng, S., 2009.
Development of polymer-based nanosized hydrated ferric oxides (HFOs) for en-
hanced phosphate removal from waste effluents. Water Res. 43, 4421–4429.
References Pengthamkeerati, P., Satapanajaru, T., Chularuengoaksorn, P., 2008. Chemical mod-
ification of coal fly ash for the removal of phosphate from aqueous solution. Fuel 87,
Acelas, N.Y., Martin, B.D., López, D., Jefferson, B., 2015. Selective removal of phosphate 2469–2476.
from wastewater using hydrated metal oxides dispersed within anionic exchange Pierotti, R.A., Siemieniewska, T., 1985. Reporting physisorption data for gas/solid sys-
media. Chemosphere 119, 1353–1360. tems with special reference to the determination of surface area and porosity. Pure
Aghazadeh, M., Golikand, A.N., Ghaemi, M., Yousefi, T., 2011. A novel lanthanum hy- Appl. Chem. 57, 603–619.
droxide nanostructure prepared by cathodic electrodeposition. Mater. Lett. 65, Ruixia, L., Jinlong, G., Hongxiao, T., 2002. Adsorption of fluoride, phosphate, and ar-
1466–1468. senate ions on a new type of ion exchange fiber. J. Colloid Interface Sci. 248,
Aghazadeh, M., Arhami, B., Barmi, A.A.M., Hosseinifard, M., Gharailou, D., Fathollahi, F., 268–274.
2014. La(OH)3 and La2O3 nanospindles prepared by template-free direct electro- Shuali, U., 1990. Thermal analysis of sepiolite and palygorskite treated with butylamine.
deposition followed by heat-treatment. Mater. Lett. 115, 68–71. Clay Miner. 25, 107–119.
Alvarezayuso, E., Garcíasánchez, A., 2003. Palygorskite as a feasible amendment to sta- Suárez, M., Garcia, E., 2006. FTIR spectroscopic study of palygorskite: influence of the
bilize heavy metal polluted soils. Environ. Pollut. 125, 337–344. composition of the octahedral sheet. Appl. Clay Sci. 31, 154–163.
Biswas, B.K., Inoue, K., Ghimire, K.N., Ohta, S., Harada, H., Ohto, K., Kawakita, H., 2007. Tanada, S., Kabayama, M., Kawasaki, N., Sakiyama, T., Nakamura, T., Araki, M., Tamura,
The adsorption of phosphate from an aquatic environment using metal-loaded orange T., 2003. Removal of phosphate by aluminum oxide hydroxide. J. Colloid Interface
waste. J. Colloid Interface Sci. 312, 214–223. Sci. 257, 135–140.
Blanco, C., GonzãLez, F., Pesquera, C., Benito, I., Mendioroz, S., Pajares, J.A., 1989. Tian, S.L., Jiang, P.X., Ning, P., Su, Y.H., 2009. Enhanced adsorption removal of phos-
Differences between one aluminic palygorskite and another magnesic by infrared phate from water by mixed lanthanum/aluminum pillared montmorillonite. Chem.
spectroscopy. Spectrosc. Lett. 22, 659–673. Eng. J. 151, 141–148.
Carazo, E., Borrego-Sánchez, A., García-Villén, F., Sánchez-Espejo, R., Viseras, C., Cerezo, Xie, F., Wu, F., Liu, G., Mu, Y., Feng, C., Wang, H., Giesy, J.P., 2014a. Removal of
P., Aguzzi, C., 2018. Adsorption and characterization of palygorskite-isoniazid na- phosphate from eutrophic lakes through adsorption by in situ formation of magne-
nohybrids. Appl. Clay Sci. 160, 180–185. sium hydroxide from diatomite. Environ. Sci. Technol. 48, 582–590.
Chahi, A., Petit, S., Decarreau, A., 2002. Infrared evidence of Dioctahedral-Trioctahedral Xie, J., Wang, Z., Fang, D., Li, C., Wu, D., 2014b. Green synthesis of a novel hybrid
site occupancy in Palygorskite. Clay Clay Miner. 50, 306–313. sorbent of zeolite/lanthanum hydroxide and its application in the removal and re-
Chubar, N.I., Kanibolotskyy, V.A., Strelko, V.V., Gallios, G.G., Samanidou, V.F., covery of phosphate from water. J. Colloid Interface Sci. 423, 13–19.
Shaposhnikova, T.O., Milgrandt, V.G., Zhuravlev, I.Z., 2005. Adsorption of phosphate Xie, J., Wang, Z., Lu, S., Wu, D., Zhang, Z., Kong, H., 2014c. Removal and recovery of
ions on novel inorganic ion exchangers. Colloids Surf. Physicochem. Eng. Asp. 255, phosphate from water by lanthanum hydroxide materials. Chem. Eng. J. 254,
55–63. 163–170.
Firsching, F.H., Kell, J.C., 1993. The solubility of the rare-earth-metal phosphates in sea Xiong, W., Peng, J., 2008. Development and characterization of ferrihydrite-modified
water. J. Chem. Eng. Data 38, 132–133. diatomite as a phosphorus adsorbent. Water Res. 42, 4869–4877.
Gan, F., Zhou, J., Wang, H., Du, C., Chen, X., 2009. Removal of phosphate from aqueous Yan, L.G., Xu, Y.Y., Yu, H.Q., Xin, X.D., Wei, Q., Du, B., 2010. Adsorption of phosphate
solution by thermally treated natural palygorskite. Water Res. 43, 2907–2915. from aqueous solution by hydroxy-aluminum, hydroxy-iron and hydroxy-iron-alu-
González-Rovira, L., Sánchez-Amaya, J.M., López-Haro, M., Hungria, A.B., Boukha, Z., minum pillared bentonites. J. Hazard. Mater. 179, 244.
Bernal, S., Botana, F.J., 2008. Formation and characterization of nanotubes of La Yang, J., Zhou, L., Zhao, L., Zhang, H., Yin, J., Wei, G., Qian, K., Wang, Y., Yu, C., 2011. A
(OH)3 obtained using porous alumina membranes. Nanotechnology 19, 495305. designed nanoporous material for phosphate removal with high efficiency. J. Mater.
Goscianska, J., Ptaszkowska-Koniarz, M., Frankowski, M., Franus, M., Panek, R., Franus, Chem. 21, 2489–2494.
W., 2018. Removal of phosphate from water by lanthanum-modified zeolites ob- Yang, J., Yuan, P., Chen, H.Y., Zou, J., Yuan, Z., Yu, C., 2012. Rationally designed
tained from fly ash. J. Colloid Interface Sci. 513, 72–81. functional macroporous materials as new adsorbents for efficient phosphorus re-
He, J., Wang, W., Sun, F., Shi, W., Qi, D., Wang, K., Shi, R., Cui, F., Wang, C., Chen, X., moval. J. Mater. Chem. 22, 9983–9990.
2015. Highly efficient phosphate scavenger based on well-dispersed La(OH)3 na- Yang, R., Li, D., Li, A., Yang, H., 2018. Adsorption properties and mechanisms of paly-
norods in polyacrylonitrile nanofibers for nutrient-starvation antibacteria. ACS Nano gorskite for removal of various ionic dyes from water. Appl. Clay Sci. 151, 20–28.
9, 9292–9302. Ye, H., Chen, F., Sheng, Y., Sheng, G., Fu, J., 2006. Adsorption of phosphate from aqueous
Hu, L., Wan, J., Zeng, G., Chen, A., Chen, G., Huang, Z., He, K., Cheng, M., Zhou, C., solution onto modified palygorskites. Sep. Purif. Technol. 50, 283–290.
Xiong, W., 2017. Comprehensive evaluation of the cytotoxicity of CdSe/ZnS quantum Yin, H., Han, M., Tang, W., 2016. Phosphorus sorption and supply from eutrophic lake
dots in Phanerochaete chrysosporium by cellular uptake and oxidative stress. sediment amended with thermally-treated calcium-rich attapulgite and a safety
Environ. Sci. Nano 4. evaluation. Chem. Eng. J. 285, 671–678.
Huang, X., Liao, X., Shi, B., 2009. Adsorption removal of phosphate in industrial waste- Zamparas, M., Zacharias, I., 2014. Restoration of eutrophic freshwater by managing in-
water by using metal-loaded skin split waste. J. Hazard. Mater. 166, 1261–1265. ternal nutrient loads. A review. Sci. Total Environ. 496, 551–562.
Huang, W.Y., Li, D., Liu, Z.Q., Tao, Q., Zhu, Y., Yang, J., Zhang, Y.M., 2014. Kinetics, Zamparas, M., Gianni, A., Stathi, P., Deligiannakis, Y., Zacharias, I., 2012. Removal of
isotherm, thermodynamic, and adsorption mechanism studies of La(OH)3 -modified phosphate from natural waters using innovative modified bentonites. Appl. Clay Sci.
exfoliated vermiculites as highly efficient phosphate adsorbents. Chem. Eng. J. 236, 62–63, 101–106.
191–201. Zamparas, M., Drosos, M., Georgiou, Y., Deligiannakis, Y., Zacharias, I., 2013. A novel
Koilraj, P., Sasaki, K., 2017. Selective removal of phosphate using La-porous carbon bentonite-humic acid composite material Bephos™ for removal of phosphate and
composites from aqueous solutions: batch and column studies. Chem. Eng. J. 317, ammonium from eutrophic waters. Chem. Eng. J. 225, 43–51.
1059–1068. Zeng, G., Jia, W., Huang, D., Liang, H., Chao, H., Min, C., Xue, W., Gong, X., Wang, R.,
Kuroki, V., Bosco, G.E., Fadini, P.S., Mozeto, A.A., Cestari, A.R., Carvalho, W.A., 2014. Jiang, D., 2017. Precipitation, adsorption and rhizosphere effect: the mechanisms for
Use of a La(III)-modified bentonite for effective phosphate removal from aqueous phosphate-induced Pb immobilization in soils - a review. J. Hazard. Mater. 339–354.
media. J. Hazard. Mater. 274, 124–131. Zhang, L., Zhou, Q., Liu, J., Chang, N., Wan, L., Chen, J., 2012. Phosphate adsorption on
Li, L., Jiang, W., Pan, H., Xu, X., Tang, Y., Ming, J., Xu, Z., Tang, R., 2008. Improved lanthanum hydroxide-doped activated carbon fiber. Chem. Eng. J. 185–186,
luminescence of lanthanide(III)-doped nanophosphors by linear aggregation. J. Phys. 160–167.
Chem. C 111, 4111–4115. Zhang, Y., Pan, B., Chao, S., Xiang, G., 2016. Enhanced phosphate removal by nanosized
Li, N., Tian, Y., Zhao, J., Zhan, W., Du, J., Kong, L., Zhang, J., Zuo, W., 2018. Ultrafast hydrated La(III) oxide confined in cross-linked polystyrene networks. Environ. Sci.
selective capture of phosphorus from sewage by 3D Fe3O4 @ZnO via weak magnetic Technol. 50, 1447–1454.
field enhanced adsorption. Chem. Eng. J. 341, 289–297. Zhao, J., Liu, J., Li, N., Wang, W., Nan, J., Zhao, Z., Cui, F., 2016. Highly efficient removal
Loganathan, P., Vigneswaran, S., Kandasamy, J., Bolan, N.S., 2014. Removal and re- of bivalent heavy metals from aqueous systems by magnetic porous Fe3O4-MnO2:
covery of phosphate from water using sorption. Crit. Rev. Environ. Sci. Technol. 44, adsorption behavior and process study. Chem. Eng. J. 304, 737–746.
847–907. Zhong, S., Zhou, C., Zhang, X., Hui, Z., Hui, L., Zhu, X., Yan, W., 2014. A novel mole-
Ma, J., Jing, Z., Li, L., Chao, Y., Yong, K., Cui, B., Zhu, R., Li, D., 2014. Nanocomposite of cularly imprinted material based on magnetic halloysite nanotubes for rapid en-
attapulgite–Ag3PO4 for Orange II photodegradation. Appl. Catal. B 144, 36–40. richment of 2,4-dichlorophenoxyacetic acid in water. J. Hazard. Mater. 276, 58–65.
Ma, X., Zuo, H., Tian, M., Zhang, L., Meng, J., Zhou, X., 2016. Assessment of heavy metals Zhou, C., Lai, C., Huang, D., Zeng, G., Zhang, C., Cheng, M., Hu, L., Wan, J., Xiong, W.,
contamination in sediments from three adjacent regions of the yellow river using Wen, M., 2017. Highly porous carbon nitride by supramolecular preassembly of
metal chemical fractions and multivariate analysis techniques. Chemosphere 144, monomers for photocatalytic removal of sulfamethazine under visible light driven.
264–272. Appl. Catal. B 220, 202–210.
Mckeown, D.A., Post, J.E., Etz, E.S., 2002. Vibrational analysis of palygorskite and se- Zhou, C., Lai, C., Xu, P., Zeng, G., Huang, D., Zhang, C., Cheng, M., Hu, L., Wan, J., Liu, Y.,
piolite. Clay Clay Miner. 50, 667–680. 2018. In situ grown AgI/Bi12O17Cl2 heterojunction photocatalysts for visible light
Mendelovici, E., 1973. Infrared study of attapulgite* and HCl treated attapulgite. Clay degradation of sulfamethazine: efficiency, pathway and mechanism. ACS Sustain.
Clay Miner. 21, 115–119. Chem. Eng. 4174–4184.

517

Вам также может понравиться