Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Inorganic Chemistry Communications 132 (2021) 108804

Contents lists available at ScienceDirect

Inorganic Chemistry Communications


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

Tuning adsorption capacity of metal–organic frameworks with Al3+ for


phosphorus removal: Kinetics, isotherm and regeneration
Zaifu Yang *, Tong Zhu , Meiyu Xiong , Anran Sun , Yujuan Xu , Yansheng Wu , Wenjun Shu ,
Zhinan Xu
College of Environmental Science and Engineering, Donghua University, Shanghai, China

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

Keywords: The separation and recycling of phosphate from sewage are indispensable to be explored due to eutrophication
Phosphate adsorption and phosphorus deficiency. Herein, iron-aluminum metal organic framework (Fe-Al-MOF) was successfully
Water treatment synthesized via solvothermal one-step treatment. The Fe-Al-MOF was measured using scanning electron mi­
Eutrophication
croscopy (SEM), infrared spectrometry (IR), thermogravimetric analysis (TGA) and X-ray powder diffractometer
Fe-Al-MOF
Kinetics
(XRD). Fe-Al-MOF was selected to explore phosphate adsorption performance and its mechanism. The results
Isotherm showed that Fe-Al-MOF adsorbent demonstrated an excellent phosphate sorption capacity of 38.33 mg P/L,
strong interference immunity in the presence of co-existing ions and optimal pH adapted to real wastewater. The
experimental data was in accordance with the Langmuir model and the pseudo-second-order model, indicating
monolayer adsorption as well as physical and chemical adsorption. More importantly, Fe-Al-MOF could reserve
over 80% after four continuous regeneration cycles in comparison with the first adsorption capacity, showing its
economics. A series of characterizations declared that chemical adsorption, ligand exchange and electrostatic
attraction were main adsorption mechanisms by Fe-Al-MOF. Overall, Fe-Al-MOF has high stability, good reus­
ability and high economy to apply for controlling wastewater and eutrophication.

1. Introduction There are many factors(e.g. organic extent, nitrate, ammonium, and so
on) influencing the efficiency of biological treatment [5]. Membrane
Phosphate is both hazards and resources. For one thing, phosphate separation has many limitations because it is pollution-prone, high-cost
plays a significant role during the growth of living organisms. Nitrogen and has specific treatment objects. To sum up, adsorption is a better
fertilizer can be produced by using components in the atmosphere, while choice for phosphate removal and recovery because it is undemanding
phosphate fertilizer can only be obtained by consuming phosphate to operate and low cost [6].
rocks. Phosphorite reserves are predicted to be exhausted in nearly fifty So far, lots of adsorbents have been investigated to adsorb phos­
years [1]. Nevertheless, phosphate resources can only be exploited phate, including red mud [7], graphene-nanoparticle aerogel compos­
about ten years in China. For another thing, excessive phosphate brings ites [8]. MOFs are better favorable choices because of specific binding
about eutrophication, which can cause the overgrowth of cyanobacteria, between metal sites and phosphate ions, abundant metal active sites and
diatoms and aquatic plants as well as water quality deterioration [2]. large surface areas [9–11]. So far, the shortcomings of MOFs are their
Microcystin, a secondary metabolite of cyanobacteria, causes non- low phosphate sorption capacity and their actual application is limited
alcoholic fatty liver by interfering with fat metabolism and has cancer- [12]. Plenty of researches illustrate materials about Fe and Al are
promoting effects. All in all, it is particularly important to reclaim applied for adsorbing inorganic oxyanions because it is environmentally
phosphate from wastewater. friendly and its price is low. Fe-MOF has the potential to deal with
In the past, there are various approaches applying for phosphate contaminants in water [13–15]. For example, MIL-101(Fe) has influence
removal, which include chemical precipitation, biological treatment, on phosphate adsorption but its phosphorus uptake capacity is low,
and membrane separation [3,4]. The chemical precipitation processes which is only 9.25 mg P/g [12]. Under equal conditions, Chan et al.
can consume plenty of chemicals and produce huge sludge production. found that Al(III)-SiO2 coprecipitates were easier to attract anionic

* Corresponding author.
E-mail address: zzfyang@dhu.edu.cn (Z. Yang).

https://doi.org/10.1016/j.inoche.2021.108804
Received 12 May 2021; Received in revised form 14 July 2021; Accepted 17 July 2021
Available online 26 July 2021
1387-7003/© 2021 Elsevier B.V. All rights reserved.
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

contaminants due to more positive charge in comparison with iron-


(C0 − Ce ) × V
based adsorbents [16]. Consequently, it can be a valid way to accel­ Qe = (1)
ma
erate the phosphate adsorption performance by adding Al into Fe-MOF.
Therefore, it is indispensable to develop MOFs with high adsorption
where Qe is equilibrium uptake capacity of adsorbent for phosphate (mg
capacity and selectivity for phosphates.
P/g); C0 and Ce are the initial and equilibrium concentrations (mg P/L),
In this work, Fe-Al-MOF was synthesized by solvothermal one-step
respectively; V represents the volume of the solution(L); ma is the dry
method and studied the adsorption performance of phosphate in simu­
weight of the adsorbent.
lated wastewater. Batch experiments were taken to explore the
Adsorption isotherms including the Langmuir and Freundlich iso­
adsorption kinetics and isotherms, as well as the influence of initial pH
therms were used to fit the uptake capacity of adsorbent for phosphate
and co-existing ions on phosphate adsorption. SEM and FT-IR spec­
[18]:
troscopy were used to explain the potential phosphate removal mecha­
nism of Fe-Al-MOF. The impact of phosphate adsorption in real water by Ce ×
1
= Ce ×
1 1
+ ×
1
(2)
Fe-Al-MOF was evaluated. Finally, the regeneration of Fe-Al-MOF was Qe Qmax KL Qmax
made a thorough inquiry.
1
lnQe = × lnCe + lnKF (3)
2. Materials and methods n

where Qmax represents the maximum adsorption capacity of phosphate


2.1. Materials
(mg P/g); KL(L/mg) and KF(mg/g (mg/L)− 1/n) are constant of Langmuir
and Freundlich model, respectively; n is the Freundlich index.
In addition to potassium dihydrogen phosphate (KH2PO4, Guaran­
teed reagent), all the drugs were of analytical purity and were used
2.4.2. Adsorption kinetics
without further purification. Iron nitrate nonahydrate (Fe(NO3)3⋅9H2O),
Batch experiments were performed to capture the kinetic model of
aluminum chloride (AlCl3), terephthalic acid (H2BDC), N,N-
phosphate by the Fe-Al-MOF. Apart from setting the initial concentra­
dimethylformamide (DMF), sodium nitrate (NaNO3), sodium sulfate
tion to 50 mg P/L and measuring the concentration of phosphate at a
(NaSO4), potassium persulfate (K2S2O8) were provided by Sinopharm
special time, the other conditions were the same as the determination of
Chemical Reagent Co. Ltd. (Shanghai, China). All solutions are prepared
the adsorption isotherm. The kinetic rate equations can be transformed
by dissolving in deionized water.
as follows:

2.2. Preparation of various adsorbents Qt = Qe × (1 − e− k1 t


) (4)

H2BDC (1.24 mmol), Fe(NO3)3⋅9H2O (2.45 mmol) and AlCl3 (2.45 t


=
t 1
+ ×
1
(5)
mmol) were dissolved in DMF (15 mL). The mixed solution was stirred Qt Qe k2 Q2e
for 1 h and then transferred to an autoclave and reacted at 110 ◦ C for 20
h. After that, the solutions were cooled down to room temperature where Qt is the adsorbed amount at time t (mg/g); k1(min− 1) and k2(g/
naturally. Finally, the light brown precipitate was collected by centri­ (mg.min)) represent rate constants of pseudo-first-order and pseudo-
fugation, washed several times with DMF and ethanol, then dried at second-order sorption, respectively.
60 ◦ C. Fe-MOF and Al-MOF were also synthesized by a similar method
and only metal salt was single. 2.4.3. Effects of initial solution pH
To investigate effects of initial solution pH, 0.2 g of sorbents were
2.3. Characterization added in 100 mL of 10 mg P/L phosphate solutions, the initial pH of
suspensions was adjusted to specific value (3–11). The pH was adjusted
FT-IR spectra of products were performed by a BRUKER OPTICS with 0.1 M HCl or NaOH solution.
Infrared Spectrometer Tensor27 with KBr as the reference. The mor­
phologies and microstructures of the products were observed on Field 2.4.4. Effects of co-existing anions
Emission Scanning Electron Microscope (FE-SEM). The crystalline Experiments were taken to examine effects of different interference
structures of adsorbents were discussed using an X-ray powder diffrac­ ions (e.g. Ca2+, Mg2+, NO−3 , SO2−4 ) on the efficiency of phosphorus up­

tion spectrometer (XRD, rigaku Ultima IV) equipped with Cu Kα radia­ take. Ca2+, Mg2+, NO−3 and SO2− 4 were derived from CaCl2, MgCl2,

tion for 2θ ranging from 10 to 80◦ . Specific surface area (SBET) was NaNO3 and Na2SO4, respectively. Two different concentrations of co-
measured by using Brunauer-Emmett-Teller (BET) method. existing anions were selected to low (50 mg/L) and high (100 mg/L)
concentration. Eventually, 0.2 g of the adsorbent was mixed with 100
2.4. Batch adsorption experiments mL of a phosphate solution (10 mg P/L) for 24 h (25 ◦ C, 120 rpm).

2.4.1. Adsorption isotherms 2.5. Adsorbent regeneration


Adsorption kinetics are applied to investigate the adsorption capac­
ity of Fe-Al-MOF for phosphate at 25 ◦ C. Phosphate ions (PO3− 4 ) stock The Fe-Al-MOF was prepared in the same manner as the adsorption
solution was prepared by dissolving 0.2197 g KH2PO4 in 1 L deionized isotherm in a 50 mg P/L phosphate solution. The Fe-Al-MOF was
water. First of all, 0.2 g of sorbent was added into 100 mL of different collected by centrifugation, washed and dried. It was then put into a
primary concentrations of phosphate simulation solution (2–100 mg P/ NaOH solution (0.1 mol/L) for desorption by ultrasonic for 30 min. Next,
L). Afterwards, the initial pH of all solutions was adjusted to 7 by adding the desorbed Fe-Al-MOF adsorbent was washed to neutral pH and dried.
HCl or NaOH solutions. The suspension was agitated at 120 rpm for 24 h After these steps, the regenerated adsorbent was used in the following
at 25 ◦ C. After complete sorption, all samples were preprocessing by phosphate adsorption experiment. The process was repeated four times.
filtering through 0.45 μm membrane syringe filters, phosphate in su­ The desorption and adsorption efficiency were calculated as fol­
pernatants was determined by ammonium molybdate spectrophoto­ lowed:
metric method (Chinese standard GB 11893–89) with a UV–vis
desorption(%) = (Cd × V)/(ma × Qe ) (6)
spectrometer (HACH, DR1900, America). The phosphate uptake ca­
pacity was figured up using Eq. (1)[17].

2
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

(C0 − Ce ) 3. Results and discussion


adsorption(%) = × 100% (7)
C0
3.1. Fe-Al-MOF characterization and mechanism of phosphorus removal
where desorption(%) and adsorption(%) are on behalf of phosphate
desorption and adsorption efficiency; Cd (mg P/L), C0(mg P/L) and Ce According to Fig. 1(a) and Fig. 1(b), it was approximately spherical
(mg P/L) represent the desorption, initial and equilibrium concentration and had a rough surface. The diameter of Fe-Al-MOF was about 500 nm.
of phosphate, respectively; Qe is equilibrium uptake capacity of adsor­ To demonstrate the phosphate adsorption mechanism of Fe-Al-MOF, FT-
bent for phosphate (mg P/g); V represents the volume of the solution(L); IR measurement was used for distinct chemical structure analysis before
ma is the dry weight of the adsorbent (g). and after phosphate adsorption (Fig. 1(c)). The peak at 2926.38 cm− 1
was observed because of –CH2 group. The peak at 591.35 cm− 1 was
from Fe-O stretching vibration [19]. Characteristic carboxylate peaks
appeared in the range of 1610–1550 cm− 1 and 1420–1300 cm− 1[20].
The tensile vibration peak of carboxylate at 1601.12 cm− 1 and 1412.72

Fig. 1. Fe-Al-MOF: (a) and (b) SEM images; (c) FT-IR spectra; (d) Thermal stability; (e) XRD pattern before and after phosphate adsorption; (f) N2 adsorp­
tion–desorption isotherms.

3
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

cm− 1 had a large offset, which may mean that this group was coordi­
nated with Al3+ to obtain the target compound [21–23]. It was specu­
lated that the carboxylate participated in the uptake process on account
of receding the strength of the carboxylate before and after the phos­
phate adsorption [24]. 757.32 cm− 1 and 476.12 cm− 1 referred to the
C–H bending vibrations of benzene [25,26]. The band at 757.32 cm− 1
could be attributed to the bending vibration of Fe/Al–O in the Fe-Al-
MOF. These not only indicated the successful coordination of carbox­
ylates of organic ligands with metal ions, but also provided strong evi­
dence for the simultaneous introduction of Fe and Al into the prepared
framework. Compared with vacant adsorbents, Fe-Al-MOF loaded with
P showed new peaks at 1050.30 and 1276.55 cm− 1. The two new peaks
could be attributed to P–O asymmetric vibration and bending vibration
of P–O [27]. In addition, the Fe/Al-O bond also shifted from 757.32
cm− 1 to the lower frequency of 756.28 cm− 1, which could be attributed
to the strong coordination between Fe/Al and P. Therefore, ligand ex­
change and chemical adsorption are possible mechanisms because ad­
sorbents provide a large number of metal adsorption sites in FT-IR
spectra of Fe-Al-MOF [27].
In addition, the thermal stability of Fe-Al-MOF was examined. It
could be seen from Fig. 1(d) that as the concentration of Al in the pre­
cursor solution increased, the decomposition temperature of the syn­
thesized metal–organic framework also gradually increased. According
to the turning temperature points about Fe-Al-MOF, different stages of
adsorbents could be separated into three phases, including phase I
(30–144 ◦ C), phase II (144–212 ◦ C) and phase III (441–550 ◦ C). External
water and residual DMF led to weight loss and mass loss was about 15%
in phase I. The evaporation of water molecules inside the material
caused the change of phase II. At 441–550 ◦ C, Fe-Al-MOF starts to be
strongly decomposed. The special structure and high thermal decom­
position temperature of Fe-Al-MOF suggested that it would be an
excellent material for the removal of phosphate in water.
XRD (Fig. 1(e)) analyzed the crystalline phases of Fe-Al-MOF. The
diffraction peak of Fe-Al-MOF was consistent with that of Fe-MIL-101
and was similar to that previously reported for Al-MIL-101 because of
the adding of Al3+[28]. Compared to Fe-MIL-101, the crystallinity of Fe-
Al-MOF increased obviously. The typical characteristic peaks of Fe-Al-
MOF still existed, indicating that the typical skeletal topology struc­
ture of Fe-Al-MOF remained after phosphate adsorption. The location
and intensity of the characteristic peaks had shifted to a certain extent, Fig. 2. (a) Phosphate adsorption isotherms. Experimental conditions for 2 mg
resulting from phosphate adsorption [29]. P/L, 0.5 g/L, pH 7.0 ± 0.1 , 25 ◦ C, 24 h; (b) Phosphate adsorption kinetics.
Compared with single metal organic framework, the BET surface Experimental conditions for 50 mg P/L, 0.2 g of sorbent dosage, pH 7.0 ± 0.1,
area of Fe-Al-MOF was 533.0809 (m2/g) (Table 1). N2 adsorp­ contact time 0–480 min.
tion–desorption isotherm (Fig. 1(f)) showed that pore size was about
5.4212 nm, indicating as the mesoporous structure. The higher BET L to 100 mg/L, the adsorption capacity remained basically stable. There
surface area and pore size larger than phosphate diameter provided the are two reasons explaining these phenomena. On the one hand, the
possibility to enhance the phosphate adsorption. active sites on the Fe-Al-MOF adsorbent are sufficient and the adsorption
capacity increases rapidly as the phosphate concentration increases
3.2. Phosphate sorption by Fe-Al-MOF when the phosphate concentration is low. On the other hand, the
adsorption capacity shows a gentle increase trend due to a large number
3.2.1. Adsorption isotherms of sites occupied by the adsorbent and reaches equilibrium states until
The figure above (Fig. 2(a)) showed the isotherms models of the adsorption site is completely occupied when the phosphate con­
adsorption in order to represent the equilibrium studies of adsorption. centration is high [30].
The Langmuir and Freundlich models are suitable choices to analyze the The Langmuir model demonstrated a better fitting with high corre­
experimental data. The Langmuir and Freundlich models were displayed lation coefficient (i.e. R2 = 0.998) than that of Freundlich isotherm (i.e.
in Fig. 2(a) and Table 2. With the increasing concentration of the R2 = 0.988). Langmuir separation constant (RL) between 0 and 1 is a sign
phosphate solution, the adsorption capacity increased from 0.976 mg/g of good adsorption capacity [31]. 1/n between 0 and 1 represents
to 21.914 mg/g. When the phosphate concentration ranged from 30 mg/ favorable adsorption capacity in the Freundlich model [32]. The
conclusion demonstrates that the adsorption occurs as monolayer over
Table 1
the homogeneous surface that have a restricted number of adsorptive
Comparison of Particle Size and BET of different MOFs. sites. Furthermore, the value of Qm calculated by Langmuir isotherm
was quite closer to the actually determined adsorption capacity (38.33
Average Particle Size(nm) BET Surface Area(m2/g)
mg/g) at an adsorbent dose of 0.5 g/L (pH = 7).
Fe-MOF 624.4231 22.3624 Table 4 summarized previously reported phosphate adsorption ca­
Fe-Al-MOF 526.66 533.0809
pacities and equilibration times for phosphate adsorbents. It could be
Al-MOF 700.16 151.3880

4
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

Table 2
Isothermal constants of Fe-Al-MOF adsorption phosphate.
Langmuir Isothermal Model Freundlich Isothermal Model
2 1/n
T(K) KL(L/mg) Qmax (mg/g) R Kf (mg/g (mg/L)− ) 1/n R2
298 0.00938 31.849 0.998 0.0334 0.693 0.988

appreciated that Fe-Al-MOF showed a relatively high phosphate


Table 4
adsorption capacity.
The maximum adsorption capacities (mg/g) of phosphate by some other re­
ported previously in literature.
3.2.2. Adsorption kinetics
Sample Qe(mg/g) Reference
As displayed in Fig. 2(b), as the time ranging from 10 min to 130 min,
phosphate adsorption capacity increased significantly. When the reac­ Bauxite residue 0.4–2.7 [7]
tion time exceeded 130 min, the capacity for adsorbing phosphate Low-grade iron ore 11.4 [33]
Modified lithium silica fume 24.10 [34]
multiplied slowly. Eventually, adsorption obtained equilibrium at 240
Modified pumice 9.7 [35]
min. Fe-Al-MOF 38.33 This work
To explore phosphate adsorption kinetics, two models were applied
to fit the adsorption process. Table 3 showed that the pseudo-second-
order model (R2 = 0.9849) fitted the time-dependent sorption curve lower than pHPZC, the surface charge of Fe-Al-MOF can be positive,
much better than the pseudo-first-order model (R2 = 0.9065). The propitious to adsorbing anion [44]. Therefore, electrostatic attraction is
dominant mechanism in the adsorption process was physical and speculated to one of the main mechanisms because the final pH is lower
chemical adsorption [36]. The larger the value of the adsorption rate than the initial pH [45]. In summary, the mechanisms associated with
constant k2, the faster the adsorption rate. The results showed that the phosphate binding onto Fe-Al-MOF included electrostatic attraction,
Fe-Al-MOF adsorbent could quickly adsorb phosphate ions and the ligand exchange and chemical adsorption, as presented in Fig. 4.
adsorption capacity was larger [37]. To estimate the stability of Fe-Al-MOF, reaction solutions after
phosphate uptake in different pH were tested by ICP-OES. The conse­
3.2.3. Effects of initial solution pH quence showed that Fe and Al were almost negligible. Therefore, Fe-Al-
The initial solution pH is one of primary factors that affect the MOF in different pH solutions had strong stability.
adsorption capacity of the adsorbent, which not only affects the exis­
tence form of phosphate, but also affects the surface charge of the 3.2.4. Effects of co-existing anions
adsorbent [39]. As shown in Fig. 3(a), as the initial pH increased from 3 Exploring the influence of coexisting ions can provide an effective
to 7, the capacity of Fe-Al-MOF rose continuously. When the pH was basis for use in actual water. The concentration of coexisting ions was set
above 7, the adsorption efficiency declined gradually. The Fe-Al-MOF to low (50 mg/L) and high (100 mg/L) concentration. Each experiment
adsorbent exhibits excellent phosphorus adsorption performance on was repeated three times. For multiple comparisons, one-way analysis of
account that the pH of most actual water is between 6.0 and 8.5[40]. variance (ANOVA, p < 0.05) analyses the significant difference among
The phosphate dissociation equilibrium in the aqueous medium treatments. According to analysis results (Fig. 5), there was no signifi­
could be stated as follows [41]: cant difference in the effects of these four coexisting ions on the
adsorption capacity.
H3 PO4 + H2 O⇌H2 PO−4 + H3 O+ pK1 = 2.15 When the hydration radius of the interfering ion closely resembled
that of the phosphate radical, the higher the hydration energy, the easier
H2 PO−4 + H2 O⇌HPO2−4 + H3 O+ pK2 = 7.20 it was to contend with the phosphate radical. Accompanied by the
increment of ion concentration, it was capable to strengthen electro­
HPO2−4 + H2 O⇌PO3−4 + H3 O+ pK3 = 12.33 static attraction due to reducing distance between molecules. This
Fig. 3(b) shows the ionic form of phosphorus under different pH maybe one reason for promoting the phosphate adsorption. In general,
conditions simulated by the Visual Miniteq model. Under strong acidic the influence of the co-existing ions studied on the phosphorus adsorp­
pH conditions, the adsorption is poor because phosphate mainly exists in tion efficiency of Fe-Al-MOF is consistent with the previously reported
the form of H3PO4. When the initial pH is between 5 and 7, phosphate results of similar phosphorus removal materials [27]. The strong selec­
mainly exists in the form of H2PO−4 . H2PO−4 is facilely adsorbed by Fe-Al- tivity is suitable for the recovery of high-purity phosphate by the
MOF under strong electrostatic interaction. Therefore, Fe-Al-MOF has regeneration.
good adsorption performance under this condition. As the pH was above
7, the adsorption had bad trends on account that phosphate exists as 3.3. Desorption and regeneration of Fe-Al-MOF
HPO2−4 . Under alkaline conditions, the electrostatic repulsion between
adsorbent and adsorbate is not conducive to the adsorption process. Considering the cost of phosphate uptake process, the Fe-Al-MOF
Competing between OH− and phosphorus anions led to the consequence was subjected to four adsorption–desorption cycles. The results illus­
[38,39,42,43]. Moreover, ligand exchange is speculated to one of the trated that the Fe-Al-MOF adsorption capacity for phosphorus showed a
main mechanisms because the finial pH value tends to decrease relative gradual downward trend after the adsorption–desorption cycle (Fig. 6).
to the initial pH value (Fig. 3(a))[40]. Compared with fresh Fe-Al-MOF, the adsorption capacity of Fe-Al-MOF
Fig. 3(c) illustrates the Zeta potentials of Fe-Al-MOF at different pH reduced by only 15% after four cycles. Nevertheless, the adsorption
conditions. The pHPZC value calculated is 10.32. When the solution pH is efficiency of Fe-Al-MOF for phosphorus is still 85%. These phenomena

Table 3
Parameters for the adsorption kinetics model of Fe-Al-MOF.
pseudo-first-order model pseudo-second-order model
− 1 − 1 2
T (K) Qe(mg g ) k1(min ) R Qe(mg g− 1) k2 × 103 R2
298 23.72 0.0366 0.9065 26.08 2.1 0.9849

5
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

Fig. 4. Main mechanisms for phosphate adsorption by Fe-Al-MOF.

Fig. 5. Effects of co-existing anions. Experimental conditions for 2 mg P/L, 0.5


g/L, pH = 7, 24 h.

Fig. 3. (a) Effects of initial solution pH and pzc of Fe-Al-MOF adsorbents by pH


drift method. Experimental conditions for 2 mg P/L, 0.5 g/L, pH 3.0–11.0, 24 h;
(b) the phosphorus species distribution curve under different pH level [38]; (c)
Zeta potential of Fe-Al-MOF.

indicated that the Fe-Al-MOF composite material is a high-efficiency


adsorbent for phosphorus removal. Fe and Al were measured by ICP-
OES in each adsorption–desorption cycle. The tested concentration of Fig. 6. The reusability of Fe-Al-MOF under four consecutive sorption cy­
Fe and Al remained negligible. cles solution.

6
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

4. Conclusion media, Environ. Sci. : Wat. Res. 1 (1) (2015) 96–107, https://doi.org/10.1039/
C4EW00020J.
[7] P.B. Cusack, M.G. Healy, P.C. Ryan, I.T. Burke, L.M.T. O’ Donoghue, É. Ujaczki,
In this work, Fe-Al-MOF was synthesized successfully by sol­ R. Courtney, Enhancement of bauxite residue as a low-cost adsorbent for
vothermal one-step method. As we suspected, the Fe-Al-MOF exhibited a phosphorus in aqueous solution, using seawater and gypsum treatments, J. Clean.
remarkable absorption capacity of 38.33 mg P/g and fast sorption ki­ Prod. 179 (2018) 217–224, https://doi.org/10.1016/j.jclepro.2018.01.092.
[8] D.N.H. Tran, S. Kabiri, L. Wang, D. Losic, Engineered graphene–nanoparticle
netics. Phosphate uptake equilibrium could be reached within 130 min aerogel composites for efficient removal of phosphate from water, J. Mater. Chem.
at an initial phosphate concentration of 50 mg P/L and sorbent dosage of A 3 (13) (2015) 6844–6852, https://doi.org/10.1039/C4TA06308B.
0.2 g. Moreover, Fe-Al-MOF sorbent showed an excellent sorption [9] A.A. Alqadami, M.u. Naushad, Z.A. Alothman, A.A. Ghfar, Novel Metal-Organic
Framework (MOF) Based Composite Material for the Sequestration of U(VI) and Th
selectivity against various disturbing ions and was able to work effi­ (IV) Metal Ions from Aqueous Environment, ACS Appl. Mater. Inter. 9 (41) (2017)
ciently at a pH of real water as well as preferable reusability. Phosphate 36026–36037, https://doi.org/10.1021/acsami.7b10768.
desorption experiments indicated that exhausted sorbent could be [10] N. Wang, J. Feng, J. Chen, J. Wang, W. Yan, Adsorption mechanism of phosphate
by polyaniline/TiO2 composite from wastewater, Chem. Eng. J 316 (2017) 33–40,
effectively regenerated by NaOH solution under ultrasonic for 30 min, https://doi.org/10.1016/j.cej.2017.01.066.
high amount of phosphate in reclaimed eluate was economically [11] R. Xu, M. Zhang, R.J.G. Mortimer, G. Pan, Enhanced Phosphorus Locking by Novel
rewarding for recovery. In addition, chemical adsorption, ligand ex­ Lanthanum/Aluminum–Hydroxide Composite: Implications for Eutrophication
Control, Environ. Sci. Technol. 51 (6) (2017) 3418–3425, https://doi.org/
change and electrostatic interaction are the main mechanisms for Fe-Al- 10.1021/acs.est.6b05623.
MOF to adsorb phosphate from water. In summary, Fe-Al-MOF adsor­ [12] Q. Xie, Y. Li, Z. Lv, H. Zhou, X. Yang, J. Chen, H. Guo, Effective Adsorption and
bents are expected to be a preferred choice for phosphate uptake in Removal of Phosphate from Aqueous Solutions and Eutrophic Water by Fe-based
MOFs of MIL-101, Sci. Rep.-UK 7 (1) (2017), https://doi.org/10.1038/s41598-
natural water.
017-03526-x.
[13] X. Liu, Y. Zhou, J. Zhang, L. Tang, L. Luo, G. Zeng, Iron Containing Metal-Organic
Funding Frameworks: Structure, Synthesis, and Applications in Environmental
Remediation, ACS Appl. Mater. Inter. 9 (24) (2017) 20255–20275, https://doi.org/
10.1021/acsami.7b02563.
This study was funded by Shanghai Science and Technology Devel­ [14] J.W. Jun, M. Tong, B.K. Jung, Z. Hasan, C. Zhong, S.H. Jhung, Effect of Central
opment Foundation (033919457). Metal Ions of Analogous Metal-Organic Frameworks on Adsorption of
Organoarsenic Compounds from Water: Plausible Mechanism of Adsorption and
Water Purification, Chem. – Eur. J. 21 (1) (2015) 347–354, https://doi.org/
Code availability 10.1002/chem.201404658.
[15] M.G. Sujana, S. Anand, Iron and aluminium based mixed hydroxides: A novel
sorbent for fluoride removal from aqueous solutions, Appl. Surf. Sci. 256 (23)
Not applicable
(2010) 6956–6962, https://doi.org/10.1016/j.apsusc.2010.05.006.
Author Statement [16] Y.T. Chan, Y.T. Liu, Y.M. Tzou, W.H. Kuan, R.R. Chang, M.K. Wang, Kinetics and
All of the authors have read and approved the manuscript. This work equilibrium adsorption study of selenium oxyanions onto Al/Si and Fe/Si
has not been published previously, nor is it being considered by any coprecipitates, Chemosphere 198 (2018) 59–67, https://doi.org/10.1016/j.
chemosphere.2018.01.110.
other peer-reviewed journal. [17] W. Zhang, L. Zhang, T. Hua, Y. Li, X. Zhou, W. Wang, Z. You, H. Wang, M. Li, The
mechanism for adsorption of Cr(VI) ions by PE microplastics in ternary system of
natural water environment, Environ. Pollut. 257 (2020) 113440, https://doi.org/
CRediT authorship contribution statement
10.1016/j.envpol.2019.113440.
[18] R.-y. Wang, W. Zhang, L.-Y. Zhang, T. Hua, G. Tang, X.-Q. Peng, M.-H. Hao, Q.-
Zaifu Yang: Conceptualization, Data curation, Writing - review & T. Zuo, Adsorption characteristics of Cu(II) and Zn(II) by nano-alumina material
editing, Supervision, Funding acquisition. Tong Zhu: Conceptualiza­ synthesized by the sol-gel method in batch mode, Environ. Sci. Pollut. R. 26 (2)
(2019) 1595–1605, https://doi.org/10.1007/s11356-018-3453-5.
tion, Methodology, Software, Investigation, Writing – original draft. [19] L. Ai, L. Li, C. Zhang, J. Fu, J. Jiang, MIL-53(Fe): A Metal-Organic Framework with
Meiyu Xiong: Software, Resources, Validation, Visualization. Anran Intrinsic Peroxidase-Like Catalytic Activity for Colorimetric Biosensing, Chemistry
Sun: Software, Resources, Validation, Visualization. Yujuan Xu: Re­ – A, European Journal 19 (45) (2013) 15105–15108, https://doi.org/10.1002/
chem.201303051.
sources, Formal analysis, Data curation. Yansheng Wu: Resources, [20] F.-H. Wei, Q.-H. Ren, Z. Liang, D. Chen, Synthesis of Graphene Oxide/Metal-
Formal analysis, Data curation. Wenjun Shu: Supervision, Project Organic Frameworks Composite Materials for Removal of Congo Red from
administration. Zhinan Xu: Supervision, Project administration. Wastewater, ChemistrySelect 4 (19) (2019) 5755–5762, https://doi.org/10.1002/
slct.v4.1910.1002/slct.201900363.
[21] D. Yu, M. Wu, Q. Hu, L. Wang, C. Lv, L. Zhang, Iron-based metal-organic
Declaration of Competing Interest frameworks as novel platforms for catalytic ozonation of organic pollutant:
Efficiency and mechanism, J. Hazard. Mater. 367 (2019) 456–464, https://doi.org/
10.1016/j.jhazmat.2018.12.108.
The authors declare that they have no known competing financial [22] F. Wei, Q. Ren, H. Zhang, L. Yang, H. Chen, Z. Liang, D. Chen, Removal of
interests or personal relationships that could have appeared to influence tetracycline hydrochloride from wastewater by Zr/Fe-MOFs/GO composites, RSC
Adv. 11 (17) (2021) 9977–9984, https://doi.org/10.1039/D1RA01027A.
the work reported in this paper. [23] F.-H. Wei, D. Chen, Z. Liang, S.-q. Zhao, Y. Luo, Synthesis and characterization of
metal–organic frameworks fabricated by microwave-assisted ball milling for
References adsorptive removal of Congo red from aqueous solutions, RSC Adv. 7 (73) (2017)
46520–46528, https://doi.org/10.1039/C7RA09243A.
[24] F. Wei, D. Chen, Z. Liang, S. Zhao, Comparison Study on the Adsorption Capacity of
[1] Y. Zheng, B. Wang, A.E. Wester, J. Chen, F. He, H. Chen, B. Gao, Reclaiming
Rhodamine B, Congo Red, and Orange II on Fe-MOFs, Nanomaterials 8 (4) (2018)
phosphorus from secondary treated municipal wastewater with engineered
248, https://doi.org/10.3390/nano8040248.
biochar, Chem. Eng. J 362 (2019) 460–468, https://doi.org/10.1016/j.
[25] Y.-K. Seo, J.W. Yoon, J.S. Lee, U.H. Lee, Y.K. Hwang, C.-H. Jun, P. Horcajada,
cej.2019.01.036.
C. Serre, J.-S. Chang, Large scale fluorine-free synthesis of hierarchically porous
[2] D.W. Schindler, The dilemma of controlling cultural eutrophication of lakes,
iron(III) trimesate MIL-100(Fe) with a zeolite MTN topology, Micropor. Mesopor.
P. Roy. Soc. B: Biol. Sci. 279 (1746) (2012) 4322–4333, https://doi.org/10.1098/
Mat. 157 (2012) 137–145, https://doi.org/10.1016/j.micromeso.2012.02.027.
rspb.2012.1032.
[26] F.H. Wei, D. Chen, Z. Liang, S.Q. Zhao, Y. Luo, Preparation of Fe-MOFs by
[3] J. Jack, T.M. Huggins, Y. Huang, Y. Fang, Z.J. Ren, Production of magnetic biochar
microwave-assisted ball milling for reducing Cr(VI) in wastewater, Dalton T. 46
from waste-derived fungal biomass for phosphorus removal and recovery, J. Clean.
(2017) 16525–16531, https://doi.org/10.1039/c7dt03776g.
Prod. 224 (2019) 100–106, https://doi.org/10.1016/j.jclepro.2019.03.120.
[27] R. Liu, L. Chi, X. Wang, Y. Wang, Y. Sui, T. Xie, H. Arandiyan, Effective and
[4] L. Kong, M. Han, K. Shih, M. Su, Z. Diao, J. Long, D. Chen, Li’an Hou, Y. Peng,
selective adsorption of phosphate from aqueous solution via trivalent-metals-based
Nano-rod Ca-decorated sludge derived carbon for removal of phosphorus, Environ.
amino-MIL-101 MOFs, Chem. Eng. J 357 (2019) 159–168, https://doi.org/
Pollut. 233 (2018) 698–705, https://doi.org/10.1016/j.envpol.2017.10.099.
10.1016/j.cej.2018.09.122.
[5] S. Shan, W. Wang, D. Liu, Z. Zhao, W. Shi, F. Cui, Remarkable phosphate removal
[28] P. Serra-Crespo, E.V. Ramos-Fernandez, J. Gascon, F. Kapteijn, Synthesis and
and recovery from wastewater by magnetically recyclable La2O2CO3/γ-Fe2O3
Characterization of an Amino Functionalized MIL-101(Al): Separation and
nanocomposites, J. Hazard. Mater. 397 (2020) 122597, https://doi.org/10.1016/j.
Catalytic Properties, Chem. Mater. 23 (10) (2011) 2565–2572, https://doi.org/
jhazmat.2020.122597.
10.1021/cm103644b.
[6] J. Lalley, C. Han, G.R. Mohan, D.D. Dionysiou, T.F. Speth, J. Garland, M.
N. Nadagouda, Phosphate removal using modified Bayoxide® E33 adsorption

7
Z. Yang et al. Inorganic Chemistry Communications 132 (2021) 108804

[29] H. Zhou, Y. Tan, Y. Yang, Y. Zhang, X. Lei, D. Yuan, Application of FeMgMn layered Fibers for the Removal of Phosphate from Water, J. Hazard. Mater. (2021) 125464,
double hydroxides for phosphate anions adsorptive removal from water, Appl. Clay https://doi.org/10.1016/j.jhazmat.2021.125464.
Sci. 200 (2021) 105903, https://doi.org/10.1016/j.clay.2020.105903. [38] Ru-yi Zhou, Jun-xia Yu, Ru-an Chi, Selective removal of phosphate from aqueous
[30] Hongcheng Chen, Yuming Huang, P. Feng, Fe-Zr bimetal-organic framework solution by MIL-101(Fe)/bagasse composite prepared through bagasse size control,
material adsorbs and removes phosphate from water, J. Southwest Univ. (Natural Environ. Res. 188 (2020) 109817, https://doi.org/10.1016/j.envres.2020.109817.
Science Edition) 41 (05) (2019) 80–84, https://doi.org/10.13718/j.cnki. [39] M. Nehra, N. Dilbaghi, N.K. Singhal, A.A. Hassan, K.-H. Kim, S. Kumar, Metal
xdzk.2019.05.013. organic frameworks MIL-100(Fe) as an efficient adsorptive material for phosphate
[31] M. Hernández Rodiguez, J. Yperman, R. Carleer, J. Maggen, D. Dadi, management, Environ. Res. 169 (2019) 229–236, https://doi.org/10.1016/j.
G. Gryglewicz, B. Van der Bruggen, J. Falcón Hernández, A. Otero Calvis, envres.2018.11.013.
Adsorption of Ni(II) on spent coffee and coffee husk based activated carbon, [40] Y. Gu, D. Xie, Y. Wang, W. Qin, H. Zhang, G. Wang, Y. Zhang, H. Zhao, Facile
J. Environ Chem. Eng. 6 (1) (2018) 1161–1170, https://doi.org/10.1016/j.jece: fabrication of composition-tunable Fe/Mg bimetal-organic frameworks for
2017.12.045. exceptional arsenate removal, Chem. Eng. J 357 (2019) 579–588, https://doi.org/
[32] M. Akram, X. Xu, B. Gao, Q. Yue, S. Yanan, R. Khan, M.A. Inam, Adsorptive 10.1016/j.cej.2018.09.174.
removal of phosphate by the bimetallic hydroxide nanocomposites embedded in [41] Z. Ajmal, A. Muhmood, M. Usman, S. Kizito, J. Lu, R. Dong, S. Wu, Phosphate
pomegranate peel, J. Environ. Sci. 91 (2020) 189–198, https://doi.org/10.1016/j. removal from aqueous solution using iron oxides: Adsorption, desorption and
jes.2020.02.005. regeneration characteristics, J. Colloid. Interf. Sci 528 (2018) 145–155, https://
[33] X. Yuan, C. Bai, W. Xia, B. Xie, J. An, Phosphate adsorption characteristics of doi.org/10.1016/j.jcis.2018.05.084.
wasted low-grade iron ore with phosphorus used as natural adsorbent for aqueous [42] G. Shedrawi, E.S. Harvey, D.L. McLean, J. Prince, L.M. Bellchambers, S.J. Newman,
solution, Desalin. Water Treat. 54 (11) (2015) 3020–3030, https://doi.org/ Evaluation of the effect of closed areas on a unique and shallow water coral reef
10.1080/19443994.2014.905974. fish assemblage reveals complex responses, Coral Reefs 33 (2014) 579–591,
[34] R. Xie, Y. Chen, T. Cheng, Y. Lai, W. Jiang, Z. Yang, Study on an effective industrial https://doi.org/10.1007/s00338-014-1160-3.
waste-based adsorbent for the adsorptive removal of phosphorus from wastewater: [43] Y. Li, Q. Xie, Q. Hu, C. Li, Z. Huang, X. Yang, H. Guo, Surface modification of
equilibrium and kinetics studies, Water Sci. Technol. 73 (2016) 1891–1900, hollow magnetic Fe3O4@NH2-MIL-101(Fe) derived from metal-organic
https://doi.org/10.2166/wst.2016.021. frameworks for enhanced selective removal of phosphates from aqueous solution,
[35] G.H. Safari, M. Zarrabi, M. Hoseini, H. Kamani, J. Jaafari, A.H. Mahvi, Trends of Sci. Rep.-UK 630651 (2016), https://doi.org/10.1038/srep30651.
natural and acid-engineered pumice onto phosphorus ions in aquatic environment: [44] Abida Kausar, Kashaf Naeem, Tariq Hussain, Zill-i-Huma Nazli, Haq Nawaz Bhatti,
adsorbent preparation, characterization, and kinetic and equilibrium modeling, Farhat Jubeen, Arif Nazir, Munawar Iqbal, Preparation and characterization of
Desalin. Water Treat. 54 (11) (2015) 3031–3043, https://doi.org/10.1080/ chitosan/clay composite for direct Rose FRN dye removal from aqueous media:
19443994.2014.915385. comparison of linear and non-linear regression methods, J. Mater. Res. Technol. 8
[36] S. Tian, P. Jiang, P. Ning, Y. Su, Enhanced adsorption removal of phosphate from (1) (2019) 1161–1174, https://doi.org/10.1016/j.jmrt.2018.07.020.
water by mixed lanthanum/aluminum pillared montmorillonite, Chem. Eng. J 151 [45] P. Sirajudheen, P. Karthikeyan, S. Vigneshwaran, S. Meenakshi, Synthesis and
(1-3) (2009) 141–148, https://doi.org/10.1016/j.cej.2009.02.006. characterization of La(III) supported carboxymethylcellulose-clay composite for
[37] K.N. Palansooriya, S. Kim, A.D. Igalavithana, Y. Hashimoto, Y.-E. Choi, toxic dyes removal: Evaluation of adsorption kinetics, isotherms and
R. Mukhopadhyay, B. Sarkar, Y.S. Ok, Fe(III) Loaded Chitosan-Biochar Composite thermodynamics, Int. J. Biol. Macromol. 161 (2020) 1117–1126, https://doi.org/
10.1016/j.ijbiomac.2020.06.103.

You might also like