banner

News

Sep 01, 2023

Synthesis of double

Scientific Reports volume 13, Article number: 8092 (2023) Cite this article

683 Accesses

2 Altmetric

Metrics details

Herein, we report the synthesis of double-shelled periodic mesoporous organosilica nanospheres/MIL-88A-Fe (DSS/MIL-88A-Fe) composite through a hydrothermal method. To survey the structural and compositional features of the synthesized composite, a variety of spectroscopic and microscopic techniques, including FT-IR, XRD, BET, TEM, FE-SEM, EDX, and EDX-mapping, have been employed. A noteworthy point in this synthesis procedure is the integration of MOF with PMO to increase the adsorbent performance, such as higher specific surface area and more active sites. This combination leads to achieving a structure with an average size of 280 nm and 1.1 μm long attributed to DSS and MOF, respectively, microporous structure and relatively large specific surface area (312.87 m2/g). The as-prepared composite could be used as an effective adsorbent with a high adsorption capacity (250 mg/g) and quick adsorption time (30 min) for the removal of Pb2+ from water. Importantly, DSS/MIL-88A-Fe composite revealed acceptable recycling and stability, since the performance in Pb2+ removal from water remained above 70% even after 4 consecutive cycles.

For all human activities, safe water is an essential material, but the presence of pollutants is one of the most significant human challenges to obtaining clean water. The rapid development of industrialization leads to increased heavy metal pollution in the environment1,2,3. To reduce water pollution from the discharge of heavy metals into nature, the environmental protection agency (EPA) has set specified allowed limits for this purpose. The maximum acceptable concentration of Pb2+ in industrial discharge and drinking water according to the guidelines of WHO and the EPA was determined to approach 0.01 and 0.015 mg/L, respectively4,5. This value for Pb (II) in wastewater is 0.05 mg/L, based on the EPA guidance6,7. Furthermore, lead-ion concentrations in industrial wastewaters are about 200–500 mg/L. It should be noted that this value is remarkably higher than the standard quality of water, hence, before discharging the wastewater to waterways or sewage systems, the lead-ion concentration must be reduced to a level of 0.05–0.10 mg/L6,8,9. Heavy metal ions including lead, hurt human health and the environment. These heavy metals can lead to many diseases and complications in the body10,11. Therefore, removing heavy metals, including Pb2+, from water and wastewater is of great importance not only for protecting water resources but also for the permanent survival of humans. According to this essential issue, scientists have focused on new technologies that will enable heavy metals to be eliminated from environmental supplies12,13,14,15. Generally, some treatment methods utilized to purify water by entrapment of heavy metal and radioactive ions have been focused on adsorption-16, membrane-17, chemical-18, electric-18, and photocatalytic-19,20 based treatments. Among them, the adsorption process is one of the most efficient methods to deal with heavy metal ions because of its simple performance, various sources of adsorbent, affordable cost, simple operation, high efficiency, and regenerative ability of adsorbents21. In recent years, conventional adsorbent materials including, metal oxides22,23,24,25,26,27, active carbon28,29,30,31,32,33,34, and carbon nanotubes35,36,37,38,39,40 have gained enormous attention. Undoubtedly these materials represent great adsorbent behaviour; nevertheless, some of them still suffer from some disadvantages including, small sizes and pore volumes, down adsorption kinetics, elaborate preparation, difficult renewal, and low adsorption efficiency. Hence, there is an urgent need to develop new adsorption materials.

Recently, porous materials such as metal–organic frameworks (MOF)41,42,43,44,45,46, and hollow periodic mesoporous organosilica (PMOs)47,48,49,50,51,52,53,54,55,56 are attracted an inclusive application outlook in the field of adsorption. Metal–organic frameworks (MOF), which are also known as a new class of hybrid and crystalline materials in the field of metal–organic materials (MOMs)57,58,59 are constructed by metal centers or clusters that bridged through strong coordination bonds with organic linkers60,61,62,63,64. The synthesis procedure for obtaining MOF NPs is classified into two major categories, including hydrothermal65 and solvothermal techniques66,67. In both procedures, two solutions containing the metal ions in their stable oxidation states, i.e., alkaline, alkaline earth, transition metal, and rare earth elements68 were mixed with the organic linkers such as poly-carboxylic molecules and poly-azaheterocycles69 to achieve a wide range of crystalline and stable MOFs structures. In recent years, these materials have experienced rapid and extensive growth of attention owing to their attractive features such as high surface area70, well-defined pore architectures71, and tunable structural features72. Their unique properties make them a great candidate for many applications, including gas storage73, purification74, molecular sensing75, drug delivery76, organic catalysts, and water purification77. In the case of water treatment applications, using MOFs with small-sized powder is associated with some risks due to the high affinity of these compounds toward the water which leads to increase the possibility of agglomeration and therefore difficult recovery45. As a result of this problem, MOF nanoparticles can directly enter drinking water78, and cause long-term environmental nano-toxicity, heavy metal pollution problems, and affect human health79. On the other hand, some of the MOF structures are sensitive to moisture and water which leads to a structural collapse in metal–organic frameworks (MOFs), which is a serious disadvantage in applied usage80. The main reason for this phenomenon can be explained by the structure of MOFs (the nature of the linker and the metal cluster) and the activation process which leads to the bridging effect that occurs in water adsorption46. For example, in the activation process of HKUST-1 MOF (= (Cu3(BTC)2) (BTC = benzene-1,3,5-tricarboxylate)) which is composed of copper ions capped by axial water ligand and BTC linkers, the axial water ligands were removed, resulting a new arrangement in the geometry of the cupric center toward a relatively stable square planar coordination81. However, according to the reported articles46,81, HKUST-1 has a high water adsorption affinity and showed no long-term stability when directly contacting with water. Against this group, some of MOF materials, such as MIL-101(Fe), exhibit excellent water stability. They can introduce as great candidates for promising materials for water adsorption applications such as the removal of heavy metals82,83,84,85. In order to increase the advantages of MOF compounds, incorporating excellent adsorbents such as PMOs with these materials leads to forming a composite with enhanced mechanical properties.

So-called periodic mesoporous organosilica (PMOs) as advanced hybrid organic–inorganic materials have been investigated with great interest owing to their unique physicochemical properties such as porous channels54, robust porous organic–inorganic framework86, adjustable pore size organization87,88, biocompatibility89, and the highest organic content in the (nano)material90. These excellent properties render PMOs attractive for applications in many fields, such as adsorption47,48,49,50,51,52, catalysis91,92,93,94,95,96,97, light-harvesting98,99, electronics100,101,102, drug release studies in simulated biological media103,104,105,106,107,108, chromatography109, enzyme immobilization110,111, and bactericides112,113. PMOs have a lot of applications as catalysts in many fields such as the synthesis of dihydropyrano[3,2-c] chromene derivatives114, sonogashira reaction115, Chan-Lum coupling116, condensation of a variety of different aldehydes with malononitrile117, Knoevenagel condensation118, oxidation of alcohols119,120, clean production of polyhydroquinolines121, and Heck reaction122. Also, the application of these materials as adsorbents for selective adsorption and separation of heavy metals (like lead) is considered one of the most attractive research hotspots. It is interesting to note that the removal of heavy metal ions using PMOs is sensitive to functional silane precursors in their frameworks123. In the other words, according to the basis of the research purpose and the type of metal ions to be removed, the PMO precursors are changed and selected very carefully (according to the hard-soft acid–base (HSAB) theory). In this regard, hard acids (such as Mg2+, Ca2+, and Cr3+) have an affinity for hard bases with high electronegativity; soft acids (such as Hg2+, Pd2+, and Hg2+) react with soft bases and have a powerful affinity to soft N and O-donor ligands and boundary acids (such as Zn2+, Fe2+, and Cr2+) react with boundary bases124,125,126. PMOs which comprised siloxane units bridged by organic groups, are traditionally synthesized through both sol–gel and grafting methods119. The sol–gel process involves mixing the organo-bridged alkoxysilanes (RʹO)3Si–R–Si(ORʹ)3 with copolymers or surfactants as a blocking agent56,127,128. While, in the grafting method, a non-bridged organosilica precursor was attached to a pre-prepared PMO material119. By inserting some specific organic groups into the pore walls of PMOs, new silica (nano) materials with both advantages of the organic and inorganic units were obtained56,87,127,128,129,130,131,132,133,134,135. Interestingly, by modifying the internal and external surface of PMOs, the hydrophilicity and hydrophobicity of the pores adjust, resulting in to control the properties of these materials in solvents136,137,138. Moreover, according to the reported literature139, under certain conditions such as pH, redox, photochemical, or biochemical conditions, the PMO materials may changes. Hereupon, modification of these compounds with other materials (such as MOFs) can be a great solution to solve their problems.

In this work, we focused on synthesizing a new nanocomposite through the integration of double-shelled periodic mesoporous organosilica (DSS) with MIL-88A-Fe aimed at enhancing water stability and adsorption capacity. In the following, the performance of DSS/MIL-88A-Fe as adsorbents will be examined to eliminate of lead from water. It provides a theoretical basis for integration of PMO with MOFs in the treatment of heavy metal ions in water.

The FTIR spectra were recorded in transmission mode 4000–400 cm−1 on Thermo Nicolet Avatar 370 spectrometer equipped at room temperature. X-ray powder diffraction (XRD) was performed on a PANalytical Company X’Pert Pro MPD diffractometer with Cu Kα (λ = 0.154 nm) radiation. Nitrogen adsorption isotherms were measured on a Quantachrome Instruments version 2.2 using N2 as the adsorbate at − 196 °C. Transmission electron microscopy (TEM) was carried out using an EM10C-100 kV microscope (ZEISS Company). FE-SEM images, EDX, and EDX-mapping were recorded by TESCAN (model: Sigma VP) scanning electron microscope operating at a low accelerating voltage of 15.00 kV and resolution of about 500 nm (ZEISS Company). Inductively coupled plasma optical emission spectroscopy (ICP-OES) was accomplished with a Varian Vista Pro CCD (Australia). UV–Vis spectra of the samples were obtained on a Hitachi UV-2910 spectrophotometer.

All reagents and chemicals were used without further purification. Absolute ethanol (EtOH, 99.9%), concentrated ammonia (28 wt%), cetyltrimethylammonium bromide (CTAB, ≥ 98%), hydrochloric acid (HCl, 38%), tetraethyl orthosilicate (TEOS, 98%), 1,2-Bis(triethoxysilyl)ethane (BTEE, 97%), and Fumaric acid (HO2CCH = CHCO2H) used in this study were purchased from Sigma-Aldrich. Ferric chloride hexahydrate (FeCl3·6H2O) was bought from PubChem. Deionized water with a resistivity of 18.2 MΩ cm−1 was used in all experiments.

Double-shelled PMO nanospheres were obtained via a sol–gel process based on literature reports140. In a typical synthesis, 0.16 g of CTAB was combined with a mixed solution of ethanol (30 mL), concentrated ammonia (1.0 mL), and deionized water (75 mL) at 40 °C for half-hour. Afterward, a mixture containing BTSE (0.119 g, 0.33 mmol) and TEOS (0.116 g, 0.56 mmol) was quickly added to the above mixture under vigorous stirring (1100 rpm) at 40 °C and kept for 24 h. To obtain two-layered mesostructured organosilica spheres, a mixture of TEOS and BTSE with an initial molar ratio was added to the mixture of the previous step. After further stirring for 24 h at 40 °C, the slurry was collected by centrifugation and washed with ethanol. The periodic mesostructured organosilica spheres were re-dispersed in 360 mL of deionized water and then transferred to a Teflon-lined stainless-steel autoclave, which was heated in an airflow electric oven at 140 °C for 5 h. After cooling the autoclave to room temperature, the product was collected by centrifugation. Subsequently, by the solvent-extraction process containing a solution containing 180 mL of ethanol and 360 µL of concentrated HCl, CTAB templates were removed from the product. Finally, double-shelled ethane-bridged PMO nanospheres were obtained after washing with ethanol three times and drying under a high vacuum at 80 °C overnight.

To prepare the MIL-88A-Fe by the ultrasonic method, 1 mmol of FeCl3·6H2O (0.27 g) was dissolved in 10 mL mixture of the DMF and ethanolic solution previously prepared with a volume ratio of 4.5:1 while maintaining a constant molar ratio (NaOH: Fe = 0.8:1). In an equimolar ratio of FeCl3·6H2O, fumaric acid (1 mmol, 0.116 g) was dissolved in 5 mL of DMF. Afterward, both solutions were mixed and sonicated for a duration 10 min with a probe using 20 W and 10 kHz in continuous wave mode. Then, the as-synthesized MIL-88A-Fe rods were centrifuged and washed with DMF and ethanol several times and then dried under vacuum at 85 °C overnight141.

As demonstrated in Scheme 1, the synthesis procedure of DSS/MIL-88(A)-Fe composite has proceeded through hydrothermal treatment. Firstly, a certain amount of the double-shelled ethane-bridged PMO powder (10 wt%) was dispersed into 20 mL ultrapure water in a bath sonicator at room temperature for 1 h. Afterward, 1 mmol of FeCl3·6H2O (0.27 g) and 1 mmol fumaric acid (0.116 g) were dissolved in a mixture of DMF and the ethanolic solution previously prepared with a proportion in volume 4.5:1 obtaining a 0.8:1 NaOH to Fe ratio in the reaction media. Then, double-shelled ethane-bridged PMO solution was gradually dropped into the above solution and sonicated for 15 min under the same conditions. Then the mixture was transferred into a 100 mL autoclave, sealed and heated to 65 °C for 12 h. The DSS/ MIL-88A-Fe composite was obtained by centrifugation and washed with ethanol several times and then dried under vacuum at 85 °C for 12 h.

Preparation of the double-shelled periodic mesoporous organosilica nanospheres/MIL-88A-Fe composite.

Various initial concentrations of Pb2+ solution (10, 20, 40, 60, 80, and 100 mg/L) were prepared by dissolving Pb(OAc)2 salt into deionized. Subsequently, DSS/MIL-88(A)-Fe composite (15 mg) was added in Pb2+ solution (50 mL) to adsorb Pb (II) at pH 6. The suspension was stirred during the test (5–150 min), and the temperature was controlled by a thermostat water bath (25 ± 2 °C). The adsorption kinetic data were obtained by sampling 5 mL of suspension at different periods during the experiment's progress. The solid adsorbent composite was separated from the solution by filtration.

The adsorption capacity (\({q}_{e}\), mg g−1) of Pb2+ by the solid adsorbent composite at equilibrium and as well as the removal efficiency (%) (R) of Pb2+ was calculated by the following formula: where C0 (mg/L) is the initial concentration of Pb (II), and Ce (mg/L) is the equilibrium concentration in the liquid phase. V is the volume of solution (mL), and m is the amount of the adsorbent (mg).

The preparation of the DSS/MIL-88(A)-Fe composite is demonstrated in Scheme 1. At the outset, monodispersed double-shelled periodic mesoporous silica nanospheres were obtained via sol–gel polymerization process140. SiO2 nanospheres two layers were achieved via a cetyltrimethylammonium bromide (CTAB) surfactant-directed sol–gel method by a mixture of silane precursors including tetraethyl orthosilicate (TEOS) and 1,2-bis(triethoxysilyl)ethane (BTSE) in water–ethanol solution. Double-shelled SiO2 nanospheres (DSS) are attained by adding a mixture of TEOS and BTSE into the reaction solution, with step-by-step addition at 24-h intervals. The successively grown DSS spheres were hydrothermally treated at 120 °C for 5 h for the transformation of the solid state to a hollow structure. SiO2 nanospheres two layers with the ordered radial mesochannels, could be obtained after the extract of CTAB surfactants from the shells by acidic ethanol. The preparation of DSS/MIL-88(A)-Fe composite adsorbent to remove Pb2+ from solution has proceeded through hydrothermal treatment. To a certain amount of FeCl3·6H2O in ultrapure water, fumaric acid was added under stirring. Then, the DSS nanospheres solution was gradually dropped into the above the solution and stirred forcefully for 2 h. Then the mixture was transferred into an autoclave to obtain the final nanocomposite.

After the successful synthesis of double-shelled periodic mesoporous organosilica nanospheres/MIL-88A-Fe composite, its structure was assessed using different spectroscopic methods, including Fourier Transform infrared spectroscopic analysis (FTIR), X-ray powder diffraction (XRD), Brunauer, Emmett and Teller (BET) surface area analysis, transmission electron microscopy (TEM), field emission scanning electron microscopy (FE-SEM), energy-dispersive X-ray spectroscopy, and inductively coupled plasma optical emission spectroscopy (ICP-OES).

The FTIR spectra of (a) DSS nanospheres, (b) MOF MIL-88(A)-Fe, and (c) DSS/MIL-88(A)-Fe composite, have been shown in Fig. 1. As it is evident from Fig. 1a, the typical absorption bands located at 1084, 819, and 465 cm−1 are assigned to the asymmetric, symmetric, and bending vibrations of the Si–O–Si bond, respectively. The FT-IR spectra of the DSS nanospheres showed absorbance bands of about 2930 cm−1, which can be assigned to the vibration of C–H bond in –CH2–CH2– group, clearly indicating the ethane-bridged frameworks.

FTIR spectra of (a) DSS nanospheres, (b) MIL-88(A)-Fe MOF, and (c) DSS/MIL-88(A)-Fe composite.

The broad absorption band at 3440 cm−1 and the distinctive band at 1623 cm−1 can be related to the stretching and bending modes of the surface-attached hydroxyl groups (ν O–H) and adsorbed water molecules in DSS nanospheres, respectively140. In the FTIR spectrum of MOF MIL-88(A)-Fe (Fig. 1b), the band corresponding to the ν(C=C) for the fumarate ligand emerges as a sharp band in the region of 1690 cm−1142. Also, two influential bands at 1607, and 1398 cm−1 can be attributed to the asymmetric and symmetric vibration modes of the carboxyl group from fumaric acid, respectively. The characteristic peak at 790 cm−1 can be associated to the C–H bending vibration of the organic linker143. Besides, the absorption band at 640 cm−1 is allocated to carbonyl group144,145. In the case of DSS/MIL-88(A)-Fe composite, as can be observed in Fig. 1c, the characteristic stretching vibration of O–H (attributed to the double-shelled SiO2) is found at 3420 cm−1, which is covered by O–H vibrational mode of water content (related to the adsorption of moisture in the air to MOF. Furthermore, the definite structure of the nanocomposite was corroborated by the advent of the absorption bands at 1608 and 1398 cm−1 (attributed to coordination between the carboxyl group and Fe3+), and absorption bands at 1054 and 800 cm−1 (associated with the vibrations of the Si–O–Si bond). The obtained results approved that the spectral information agrees with the XRD analysis results described in the following.

In order to confirm the crystalline structure of (a) DSS nanospheres, (b) MIL-88(A)-Fe MOF, and (c) DSS/MIL-88(A)-Fe composite, the powder X-ray diffraction (PXRD) technique was performed, and the results are shown in Fig. 2. It can be observed (Fig. 2a), that the amorphous structure of double-shelled SiO2 (DSS) nanospheres (I) exhibits a broad diffraction peak at 2θ = 22.5°146. For the MIL-88(A)-Fe MOF, the major distinct peaks are located at 2θ = 8.4°, 9.9°, 12.8°, 15.3°,15.9°, 21.1°, and 22.5° related to the (010), (101), (110), (002), (012), (022) and (103) crystal planes (Fig. 2b)141,147. In the XRD pattern of the DSS/MIL-88(A)-Fe composite, all peaks were well-matched with those of their corresponding pure MIL-88(A)-Fe MOF. Notably, the diffraction peak at 2θ = 8.5° attributed to the (100) crystallographic facet developed more than that of pure MIL-88(A)-Fe (Fig. 2c). The shift might be due to the double-shelled periodic mesoporous organosilica nanospheres that controlled the MIL-88(A)-Fe MOF crystal orientation in the modified composite. The XRD results suggested that the pure MIL-88(A)-Fe MOF and desired composite were successfully synthesized.

XRD patterns of DSS nanospheres (a), MIL-88(A)-Fe MOF (b), and DSS/MIL-88(A)-Fe composite (c).

It is known that the BET surface area and pore structure of the as-synthesized nanocomposite are substantial factors, which influence the catalytic activity. To gain further insights into the textural properties (the pore-size distributions and BET surface areas), N2 adsorption-desorption analysis of DSS nanospheres (a), MIL-88(A)-Fe MOF (b), DSS/MIL-88(A)-Fe nanocomposite (c), and DSS/MIL-88(A)-Fe nanocomposite after Pb2+ adsorption (d) were presented in Fig. 3. As seen in Fig. 3a, DSS nanospheres exhibit typical type IV isotherm with a large hysteresis loop indicating the presence of mesoporous structure with a high surface area of about 244.87 m2 g−1 and pore volume of 0.864 cm3 g−1. It was found in Fig. 3b that the hysteresis loops in the isotherm curve of MIL-88(A)-Fe MOF can be attributed to typical type IV isotherm with H3 hysteresis loop148. Furthermore, the derived BET (obtained from Brunauer–Emmett–Teller (BET) theory) surface area and pore volume were estimated to be 236 m2 g−1 and 0.180 cm3 g−1, respectively. As it is evident from the data summarized in Table 1, the BET surface area of DSS/MIL-88(A)-Fe nanocomposite is larger than the pure DSS and MOF (Fig. 3c). The larger pore volume of DSS/MIL-88(A)-Fe nanocomposite compared to the pure DSS and MOF offered a mesoporous architecture for composite samples, providing a suitable pathway for mass transport (Table 1). Moreover, the mean pore volume of DSS/MIL-88(A)-Fe nanocomposite (2.919 nm) is slightly lower than those of the pure DSS (14.115 nm) and MOF (3.056 nm), which is probably due to the synergistic effect and implies the successful combination of MOF and PMOs. Furthermore, investigating on the Barrett–Joyner–Halenda (BJH) pore size distribution which was calculated using the adsorption branch (presented in Fig. 3e) clearly shows three peaks centered at 8, 19 and 61 nm, corresponding to the mesoporous of the shell, hollow void (the shell-in-shell distance) and also macro porous of MOF, respectively. By comparing the BJH before and after Pb2+ adsorption, the conclusion could be derived that no significant changes were observed in the pore size distribution (Fig. 3e,f).

N2 adsorption–desorption analysis of DSS nanospheres (a), MIL-88(A)-Fe MOF (b), DSS/MIL-88(A)-Fe nanocomposite (c), and DSS/MIL-88(A)-Fe nanocomposite after Pb2+ adsorption (d). BJH plots of DSS/MIL-88(A)-Fe nanocomposite (e), and DSS/MIL-88(A)-Fe nanocomposite after Pb2+ adsorption (f).

The morphological studies of DSS/MIL-88(A)-Fe nanocomposite was performed by investigating the field emission scanning electron microscopy (FE-SEM) and transmission electron microscopy (TEM). The obtained results are presented in Fig. 4. In order to testify the well synthesized DSS/MIL-88(A)-Fe nanocomposite, initially we examined the pure DDS nanospheres and MIL-88(A)-Fe MOF by FE-SEM analysis. As can be observed from Fig. 4a, DDS exhibits a uniformly spherical shape, monodisperse size distribution, and high surface area with particle sizes of about 280 nm. Also, the original MIL-88(A)-Fe MOF shows well-crystallized rods with a hexagonal face and an average size of 1.2 μm (Fig. 4b). The formation of DSS/MIL-88(A)-Fe nanocomposite is obviously confirmed from the FE-SEM images (Fig. 4c,d). As revealed by Fig. 4c,d, the obtained DSS/MIL-88(A)-Fe nanocomposite preserved the original shape from the parent DSS and MOF with an average size of 280 nm and 1.1 μm, respectively. Interestingly, the nanocomposite structure shows the formation of a regular arrangement of DDS nanospheres, as indicated in Fig. 4a, along with the formation of a hexagonal-rods structure with dimensions in the range of the former MOF morphologies. Moreover, to investigate the effect of Pb2+ adsorption on the morphology, the FE-SEM image of DSS/MIL-88(A)-Fe nanocomposite after adsorption was recorded and illustrated in Fig. 4e. By comparing the FE-SEM images of DSS/MIL-88(A)-Fe nanocomposite before (Fig. 4c,d) and after (Fig. 4e) adsorption of Pb2+, it could be concluded that Pb2+ ions adsorbed into the surface of DSS/MIL-88(A)-Fe nanocomposite.

SEM images of DSS nanospheres (a), MIL-88(A)-Fe MOF (b), DSS/MIL-88(A)-Fe nanocomposite (c,d), after Pb2+ adsorption (e), and reusability of DSS/MIL-88(A)-Fe nanocomposite after 4 adsorption–desorption cycles (f).

For further insight on the morphology of the synthesized DSS/MIL-88(A)-Fe nanocomposite, the TEM images were surveyed and presented in Fig. 5a–c. The TEM images of the nanocomposite confirm the existence of bright and dark areas in the double-shelled structures. Furthermore, as it is evident from Fig. 5a,b, the cavities observed from TEM images suggested the presence of mesoporous channels on the surface of the DSS nanospheres. Of note that the formation of the hexagonal-rods structure of MIL-88(A)-Fe MOF was clearly observed in Fig. 5c. Also, the thickness of the external layer, internal layer and the hollow void (the shell-to-shell distance) was estimated to be ∼13, ∼ 20 and ∼33 nm, respectively. Accordingly, the overall outer diameter is about ∼280 nm for DSS nanospheres which is very close to the FE-SEM data.

TEM images of DSS (a), and DSS/MIL-88(A)-Fe nanocomposite (b,c).

Surveying the EDX spectrum confirmed the existence of C, O, and Fe elements in the nanocomposite structure (Fig. 6a). After adsorption, the Pb2+ signals have been detected in the used adsorbent surface (Fig. 6b).

EDX of DSS/MIL-88(A)-Fe nanocomposite (a), and after adsorption of Pb2+ (b).

To further evaluate the DSS/MIL-88(A)-Fe nanocomposite, EDX-mapping analysis was performed and the presence of C, O, Si, and Fe with uniform distributions was approved (Fig. 7a). In addition, after Pb2+ adsorption, the elemental distributions were almost identical. These findings evidenced that Pb2+ had successfully attached to the adsorbent surface (Fig. 7b).

EDX-mapping of DSS/MIL-88(A)-Fe nanocomposite (a), and after adsorption of Pb2+ (b).

In this study, the removal of Pb2+ from aqueous solution has been investigated using double-shelled periodic mesoporous organosilica nanospheres/MIL-88A-Fe composite. In order to check the adsorption of lead ions from an aqueous solution, in the first, to the 50 mL solutions of Pb2+ with varying concentrations (10, 20, 40, 60, 80, 100 mg/L), 15 mg DSS/MIL-88A-Fe composite was separately added at pH 6. Then, initial concentrations of Pb2+ were measured at an interval of 5, 15, 30, 45, 60, 75, 90, and 110 min at room temperature. The effects of contact time and initial concentration of Pb2+ on the adsorption of Pb2+ by DSS/MIL-88A-Fe were studied. It was found from Fig. 8 that the adsorption capacity of Pb2+ with an initial mass concentration of 10–100 mg/L on DSS/MIL-88A-Fe composite reached an adsorption equilibrium within 30 min.

Effect of contact initial concentration of Pb2+ and time on its adsorption.

From Fig. 9, DSS/MIL-88A-Fe composite has the highest adsorption capacity compared with MIL-88A-Fe and DDS, which is attributed to the higher surface area and more numerous active sites for adsorption of Pb2+ ions (Pb2+ solution with a concentration of 100 mg/L). It became apparent that the adsorption capacity increased with an increase in the initial Pb2+ concentration. This is due to the increase in the driving force of the concentration gradient with the increase in the initial Pb2+ concentration.

Pb2+ adsorption isotherms for DSS, MIL-88A-Fe and DSS/MIL-88A-Fe composite at room temperature with a concentration of 100 mg/L of Pb2+ solution.

The effect of adsorbent dosage on the removal efficiency of Pb2+ by DSS/ MIL-88A-Fe composite was investigated. For this purpose, 50 mL of Pb2+ solution with a concentration of 40 mg/L, and 5–35 mg of DSS/MIL-88A-Fe composite were added at room temperature and stirred for 30 min. As can be seen from Fig. 10, with the increase of adsorbent dosage, the percentage removal initially increases, which could be due to the increase of the total surface area of the adsorbent, and the adsorption site number of the adsorbent in a certain amount of solution. But beyond a value of 20 mg, the percentage removal reaches an almost constant value. This may be due to an overlapping of adsorption sites and, consequently, of the adsorbent particles overcrowding. Maximum removal of 86.6% was observed at an adsorbent dosage of 20 mg/L (for economic purposes) at pH 6.

Effect of adsorbent dosage on adsorption capacity and removal efficiency of Pb2+.

It is well-known that the pH value of the solution is an essential factor affecting the adsorption performance of Pb2+. For this purpose, different pH values (3, 4, 5, 6, and 7) affecting the adsorption performance of Pb2+ by DSS/MIL-88A-Fe were investigated (Fig. 11). The adsorption capacity of Pb2+ on composite depends on the pH value, which increased slowly with an increase of pH until pH 6.

Effect of pH value on adsorption capacity of Pb2+.

The surface of the DSS/MIL-88A-Fe composite is protonated in the acidic solutions (pH 3, 4, and 5), which contains a positive charge causing electrostatic repulsion between DSS/MIL-88A-Fe and Pb2+. This phenomenon demonstrated that a large amount of H+ ions competed with Pb2+ ions for the adsorption sites in the composite in an acidic environment149. Thus, the adsorption capacity of Pb2+ by the adsorbent DSS/MIL-88A-Fe composite was reduced. When the pH exceeds 6, Pb2+ ions quickly form Pb2+ hydroxide (precipitation) with OH−150. Hence the optimum pH of 6 was selected for the subsequent adsorption experiments.

To obtain mechanism information on the adsorption process, adsorption isotherm models are important for designing the adsorption system. The adsorption capacity of Pb2+ by DSS/MIL-88A-Fe composite was evaluated using Langmuir (3), Freundlich (4), and Temkin (5) isothermal adsorption model to find the interaction between DSS/MIL-88A-Fe composite and Pb2+. The corresponding equations of isothermal adsorption models are as follows;

where \({q}_{e}\)(mg/g) is the equilibrium adsorption capacity of Pb2+ in solution, \({q}_{L}\)(mg g−1) represents the maximum adsorption capacity, KL (L mg−1) is the Langmuir constant ascribed to the affinity of the binding sites between the adsorbent and the target substance. KF ((mg g−1)/(mg g−1)1/n) and n ((mg(1−(1/n)) L(1/n) g−1) are Freundlich constants representing adsorption capacity and adsorption intensity, respectively. R is the gas constant (8.314 J/mol⋅K), T is the temperature (K).

Three models of experimental data from adsorption processes were applied to explain the Pb2+ adsorption mechanism between the liquid and adsorbent phases. The fitting results of the adsorption of Pb2+ are shown in Fig. 12. The parameters of the models are listed in Table 2.

Linear plot for (a) Langmuir, (b) Freundlich, and (c) Temkin adsorption isotherm models.

The Langmuir model is normally associated with monolayer adsorption characteristics and the energy level of a homogeneous system, which has no following interaction between adsorbed species151. The Freundlich model is generally an empirical one related to heterogeneous systems and applied to multi-layer adsorption of the adsorbent152.

The Temkin model is commonly described by a uniform distribution of binding energies to explain adsorbate–adsorbent interactions on adsorption sites153.

The most favourable isothermal adsorption model was provided by the Freundlich model, as it yielded a higher R2 value in contrast to the value of the correlation coefficient between the Langmuir and Temkin models. The adsorption behaviour of Pb2+ by DSS/MIL-88A-Fe composite mainly occurred on heterogeneous pores or surfaces as the main adsorption sites, and multi-layer adsorption could exist.

\(\frac{1}{\mathrm{n}}\) is a constant reflecting the Pb2+ adsorption intensity by the adsorbent; if the value of \(\frac{1}{\mathrm{n}}\) sits between 0.1 and 1.0, the process of adsorption is favourable154.

Temperature is one of the essential factors that influence the adsorption capacity of composite on absorbing Pb2+. The effect of temperature on the adsorption of Pb2+on DSS/MIL-88A-Fe composite was investigated at three temperatures at 298, 308, and 318 K. Obviously, on increasing the temperature, the adsorption capacity of the adsorbent on absorbing Pb2+ increased. This showed that the adsorption process was endothermic.

To further investigate the thermodynamic features, the thermodynamic parameters such as Gibb's free energy change (ΔG°, kJ/mol), enthalpy change (ΔH°, kJ/mol), and entropy change (ΔS°, J/(mol·K)) were calculated using the following equations:

where R is the universal gas constant (8.314 J/mol K); T is the Kelvin temperature (K); Kc represents the thermodynamic equilibrium constant.

The ΔH° and ΔS° values were acquired from the slopes, and intercepts of the Van't Hoff curve. Results are shown in Fig. 13 and the thermodynamic parameters are provided in Table 3.

Van Der Hoff equation. Experimental conditions: [Pb2+] = 80 mg L−1, pH 6.0, adsorbent dose = 15 mg, solution volume = 50 mL; time = 30 min.

The negative values of ΔG° indicated the adsorption of Pb2+ on DSS/MIL-88A-Fe composite was spontaneous. Moreover, with the elevated temperature, the absolute value of ΔG° increased, revealing that high temperature can promote the adsorption process.

The positive values of ΔH° suggest that the adsorption process was endothermic, in nature, whereas the positive value of entropy change (ΔS°) reveals the increase in randomness at the solid/solution interface during the adsorption of Pb2+. Therefore, the adsorption was an endothermic, and spontaneous process.

The adsorption kinetics of Pb2+ on DSS/MIL-88A-Fe composite using 80 mg L−1 Pb2+ solution of pH 6 at room temperature were also studied. As shown in Fig. 8, the fast Pb2+ adsorption process on DSS/MIL-88A-Fe composite in the first 5 min may be due to the presence of sufficient active adsorption sites available on the surface of the adsorbent. The adsorption was almost attained equilibrium within 30 min.

To study the adsorption kinetics and accurately interpret the adsorption behavior of Pb2+ adsorption on DSS/MIL-88A-Fe composite, four types of kinetic models, including the pseudo-first-order (9), pseudo-second-order (10), Elovich model (11), and particle diffusion (12) are expressed as follows155,156;

where \({q}_{e}\) and \({q}_{t}\) are the adsorption capacity at equilibrium and time (mg g−1), K1 (min−1), and K2 [g (mg min)−1] are adsorption rate constant of pseudo-first-order and pseudo-second-order kinetics, respectively. Ki (mg/g \({\mathrm{min}}^{0.5})\) and a are adsorption rate constant of intra-particle diffusion, and intra-particle diffusion constants that reflecting boundary layer effect, respectively. \(\alpha\)[mg (g min)−1] and \(\beta\)(g mg−1) illustrate the initial constant adsorption and desorption constants, respectively157,158,159.

The adsorption process usually has various steps and out of which the slowest step controls the rate of the adsorption process.

The experimental data were fitted by using the above four adsorption kinetic models. According to the fitting results in Fig. 14a–d, and Table 4, it is evident that the R2 value of the pseudo-second-order dynamics model is greater than that of the other three models (the correlation coefficient R2 of the pseudo-second-order dynamic model is equal 0.998). This result indicates that the adsorption rate on the surface of the adsorbent is the rate-determining step and the adsorbent surface corresponds to a heterogeneous system.

Plots of pseudo-first-order kinetic model for the adsorption of Pb2+ (a), pseudo-second-order kinetic model for the adsorption of Pb2+ (b), Elovich kinetic model for the adsorption of Pb2+ (c), and intra-particle-diffusion kinetic model for the adsorption of Pb2+ (d). Conditions: T = 398 K, adsorbent = 15 mg/L, metal = 80 mg/ L, pH 6, and contact time = 90 min.

According to the above analysis and characterization, a possible mechanism for Pb2+ removal is suggested. To further understand the Pb2+ adsorption process and the composite-heavy metal interaction, the zeta potential of the DSS and MIL-88(A)-Fe was measured at pH 6. On the basis of the obtained results, the surface charge of DDS was negative (− 40 mV) indicating the electrostatic attraction enhancing for more Pb2+ adsorption. So, the mechanism of adsorption between DSS-heavy metals is due to the electrostatic attraction of unlike charges at pH 6.

On the other hand, DSS has provided a high specific surface area and mesoporous channel microstructure in which many hydroxyl-functional groups were exposed on the surface of DSS materials as active sites for the adsorption of heavy metal ions. The adsorption has occurred mainly through electrostatic interactions between the surface hydroxyl group of DSS and the heavy metal ions.

Furthermore, the Pb2+ adsorption mechanism in the solution via MIL-88(A)-Fe could readily happen by ion exchange protons on the surface of the adsorbent with Pb2+ (This observation has been confirmed by comparison before and after Pb2+ adsorption through FTIR and SEM analysis).

It could be understood that the Pb2+ adsorption mechanism involved competitive ion exchange with MIL-88(A)-Fe and electrostatic interactions with the DSS of the composite. Also, the Pb2+ adsorption mechanism could be occurred from binding to open metal sites on the MOF of composite (a pore-filling mechanism) or interacting with active sites on the surface of DSS containing hydroxyl-functional groups.

The MIL-88(A)-Fe showed a positive surface charge (+ 30 mV). It is thus concluded, from this standpoint, the interaction between the DDS nanoparticles and MIL-88(A)-Fe is through electrostatic attraction to form a composite material denoted as DSS/MIL-88(A)-Fe.

The recyclability of the adsorbent is an important index, both economically and environmentally. In order to evaluate the recycling ability of DSS/MIL-88(A)-Fe composite, the adsorbent was placed in an HCl solution with a concentration of 0.05 M and stirred for 2 h, followed by filtering and washing with nitric acid for 3 times. Afterward, the composite was cleaned with deionized water for several times and activated at 85 °C for 10 h to adsorb Pb2+ again in 40 mg L−1 solution. The effect of DSS/MIL-88(A)-Fe composite cycling times on adsorption capacity is revealed in Fig. 15. The adsorption capacity of Pb2+ by DSS/MIL-88(A)-Fe composite decreases slightly after 4 periods of regeneration to reuse the adsorbent (from 77.8% at the first cycle to 70.3% at the fourth cycle). The decrease in adsorption capacity during repeated use could be caused by the mass loss of the composite adsorbent in acid treatments. The results demonstrate good recycling capability in the cyclic adsorption process for Pb2+ adsorption.

Cycle performance of DSS/MIL-88(A)-Fe composite for Pb2+ adsorption.

In order to further investigate the adsorption mechanism and the recycling capability of DSS/MIL-88(A)-Fe composite in the cyclic adsorption process for Pb2+ adsorption, the FT-IR spectra of DSS/MIL-88(A)-Fe composite before (Fig. 16a), after adsorption of Pb2+ (Fig. 16b), and also reused adsorbent after four catalytic runs (Fig. 16c) were analyzed.

FTIR spectra of fresh DSS/MIL-88(A)-Fe composite (a), after pb2+ adsorption (b), and reused DSS/MIL-88(A)-Fe composite (c).

As it is beheld from Fig. 16b, the characteristic peaks at 3420 and 1054 cm−1 attributed to vibrations of the DSS, shifted to 3407 and 1043 cm−1 and the intensity of the peaks weakened after the adsorption of Pb2+. Furthermore, the significant peaks at 1608 and 1398 cm−1 represented coordination between the carboxyl group and Fe3+, were red-shifted to 1568 and 1390 cm−1, and significantly weakened (Supplementary Information).

Also, from the FT-IR spectrum of reused adsorbent after four catalytic runs can be concluded that were matched in all the characteristic absorption bands such as shapes, positions, and frequencies with the FT-IR spectra of the fresh catalyst (Fig. 16c).

According to the data shown in Fig. 4f, the reused FE-SEM image of the DSS/MIL-88(A)-Fe composite asserts that not much change in the particle size or shape and morphology was observed after four catalytic runs.

A comparison of adsorption capacity and adsorption equilibrium time for Pb2+ by DSS/MIL-88(A)-Fe nanocomposite with some different adsorbents is listed in Table 5. Obviously, in previously reported works, despite the short adsorption equilibrium time of Pb2+, adsorbents demonstrated poor adsorption capacity compared to this work (Table 5, entries 1–6). Also, modified biochar showed a long adsorption time and relatively low adsorption capacity (Table 5, entry 7). In this study, the DSS/MIL-88(A)-Fe nanocomposite as an effective adsorbent of Pb2+ from an aqueous solution has high adsorption performance as well as a short adsorption time.

In summary, we synthesized a double-shelled periodic mesoporous organosilica nanospheres/MIL-88(A)-Fe nanocomposite in a conventional manner and comprehensively characterized through various techniques, including FTIR, XRD, BET, TEM, FE-SEM, EDX, and EDX-mapping analysis. Thanks to unique structure and remarkable properties of DSS/MIL-88A-Fe composite (such as an average size of 280 nm and 1.1 μm long attributed to the DSS and MOF, respectively, microporous structure and relatively large specific surface area (312.87 m2/g), which resulted from the coexistence of PMO and MOF, the composite described above exhibited excellent performance in the separation of Pb2+ from water with a maximum adsorption capacity of 230 mg/g with an effective adsorption rate of around 90 min. More importantly, one of the main advantages of this unprecedented composite is its cycling stability. To this end, the reusability of the DSS/MIL-88(A)-Fe was investigate, and the obtained results demonstrated that this adsorbent was preserved after 4 times regeneration, which illustrated its favorable performance in the removal of lead metal pollutants.

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Zhao, J., Wang, C., Wang, S. & Zhou, Y. Experimental and DFT study of selective adsorption mechanisms of Pb (II) by UiO-66-NH2 modified with 1, 8 dihydroxyanthraquinone. J. Ind. Eng. Chem. 83, 111–122 (2020).

Article CAS Google Scholar

Briffa, J., Sinagra, E. & Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 6, e04691 (2020).

Article CAS PubMed PubMed Central Google Scholar

Zaynab, M. et al. Health and environmental effects of heavy metals. J. King Saud Univ. Sci. 34, 101653 (2022).

Article Google Scholar

Griffiths, C. et al. US Environmental Protection Agency valuation of surface water quality improvements. Rev. Environ. Econ. Policy 6, 130 (2012).

Article Google Scholar

World Health Organization. Guidelines for Drinking-Water Quality Vol. 1 (World Health Organization, 2004).

Google Scholar

Inoue, K. H. & Hagerman, A. E. Determination of gallotannin with rhodanine. Anal. Biochem. 169, 363–369 (1988).

Article CAS PubMed Google Scholar

Yamaguchi, H., Higasida, R., Higuchi, M. & Sakata, I. Adsorption mechanism of heavy-metal ion by microspherical tannin resin. J. Appl. Polym. Sci. 45, 1463–1472 (1992).

Article CAS Google Scholar

Özacar, M. & Şengil, İA. Adsorption of metal complex dyes from aqueous solutions by pine sawdust. Bioresour. Technol. 96, 791–795 (2005).

Article PubMed Google Scholar

Shukla, S. & Pai, R. S. Removal of Pb (II) from solution using cellulose-containing materials. J. Chem. Technol. Biotechnol. 80, 176–183 (2005).

Article CAS Google Scholar

Karnitz, O. et al. Adsorption of heavy metal ion from aqueous single metal solution by chemically modified sugarcane bagasse. Bioresour. Technol. 98, 1291–1297 (2007).

Article CAS PubMed Google Scholar

Mitra, S. et al. Impact of heavy metals on the environment and human health: Novel therapeutic insights to counter the toxicity. J. King Saud Univ. Sci. 1, 101865 (2022).

Article Google Scholar

Jan, A. T. et al. Heavy metals and human health: Mechanistic insight into toxicity and counter defense system of antioxidants. Int. J. Mol. Sci. 16, 29592–29630 (2015).

Article CAS PubMed PubMed Central Google Scholar

Çelebi, H., Bahadır, T., Şimşek, İ & Tulun, Ş. An environmental and green process for Pb2+ pollution: An experimental research from the perspective of adsorption. Eng. Proc. 19, 20 (2022).

Google Scholar

Ethaib, S., Al-Qutaifia, S., Al-Ansari, N. & Zubaidi, S. L. Function of nanomaterials in removing heavy metals for water and wastewater remediation: A review. Environment 9, 123 (2022).

Google Scholar

Shakarami, M. & Maroufi, S. Wastewater and sewage sludge effects on the absorption of some heavy metals in soil and mint. J. Health Sci. Surveill. Syst. 45, 1–15 (2019).

Google Scholar

Bansal, R. C. & Goyal, M. Activated Carbon Adsorption (CRC Press, 2005).

Book Google Scholar

Gunawan, P. et al. Hollow fiber membrane decorated with Ag/MWNTs: Toward effective water disinfection and biofouling control. ACS Nano 5, 10033–10040 (2011).

Article CAS PubMed Google Scholar

Khare, P., Yadav, A., Ramkumar, J. & Verma, N. Microchannel-embedded metal–carbon–polymer nanocomposite as a novel support for chitosan for efficient removal of hexavalent chromium from water under dynamic conditions. J. Chem. Eng. 293, 44–54 (2016).

Article CAS Google Scholar

Di Mauro, A. et al. Novel synthesis of ZnO/PMMA nanocomposites for photocatalytic applications. Sci. Rep. 7, 40895 (2017).

Article PubMed PubMed Central Google Scholar

Paz, Y. Application of TiO2 photocatalysis for air treatment: Patents’ overview. Appl. Catal. B Environ. 99, 448–460 (2010).

Article CAS Google Scholar

Zaimee, M. Z. A., Sarjadi, M. S. & Rahman, M. L. Heavy metals removal from water by efficient adsorbents. Water 13, 2659 (2021).

Article CAS Google Scholar

Hu, J. S., Zhong, L. S., Song, W. G. & Wan, L. Synthesis of hierarchically structured metal oxides and their application in heavy metal ion removal. Adv. Mater. 20, 2977–2982 (2008).

Article CAS Google Scholar

Ravindranath, R. et al. Fe2O3/Al2O3 microboxes for efficient removal of heavy metal ions. J. Chem. 41, 7751–7757 (2017).

CAS Google Scholar

Zhang, Y., Bian, T., Gu, J., Zheng, X. & Li, Z. Controllable ZnO architectures with the assistance of ethanolamine and their application for removing divalent heavy metals (Cu, Pb, Ni) from water. New J. Chem. 42, 3356–3362 (2018).

Article CAS Google Scholar

Hua, M. et al. Heavy metal removal from water/wastewater by nanosized metal oxides: A review. J. Hazard. Mater. 211, 317–331 (2012).

Article PubMed Google Scholar

Li, M., Liu, J., Xu, Y. & Qian, G. Phosphate adsorption on metal oxides and metal hydroxides: A comparative review. Environ. Rev. 24, 319–332 (2016).

Article CAS Google Scholar

Wang, L., Shi, C., Pan, L., Zhang, X. & Zou, J.-J. Rational design, synthesis, adsorption principles and applications of metal oxide adsorbents: A review. Nanoscale 12, 4790–4815 (2020).

Article CAS PubMed Google Scholar

Alimohammady, M., Jahangiri, M., Kiani, F. & Tahermansouri, H. Highly efficient simultaneous adsorption of Cd (II), Hg (II) and As (III) ions from aqueous solutions by modification of graphene oxide with 3-aminopyrazole: Central composite design optimization. New J. Chem. 41, 8905–8919 (2017).

Article CAS Google Scholar

Kołodyńska, D., Krukowska, J. & Thomas, P. Comparison of sorption and desorption studies of heavy metal ions from biochar and commercial active carbon. J. Chem. Eng. 307, 353–363 (2017).

Article Google Scholar

Sawant, S. Y., Pawar, R. R., Lee, S.-M. & Cho, M. H. Binder-free production of 3D N-doped porous carbon cubes for efficient Pb2+ removal through batch and fixed bed adsorption. J. Clean. Prod. 168, 290–301 (2017).

Article CAS Google Scholar

Wei, M.-P., Chai, H., Cao, Y.-L. & Jia, D.-Z. Sulfonated graphene oxide as an adsorbent for removal of Pb2+ and methylene blue. J. Colloid Interface Sci. 524, 297–305 (2018).

Article ADS CAS PubMed Google Scholar

Anfar, Z. et al. Recent trends on numerical investigations of response surface methodology for pollutants adsorption onto activated carbon materials: A review. Crit. Rev. Environ. Sci. Technol. 50, 1043–1084 (2020).

Article CAS Google Scholar

Shafeeyan, M. S., Daud, W. M. A. W., Houshmand, A. & Shamiri, A. A review on surface modification of activated carbon for carbon dioxide adsorption. J. Anal. Appl. Pyrol. 89, 143–151 (2010).

Article CAS Google Scholar

Wang, X. et al. Key factors and primary modification methods of activated carbon and their application in adsorption of carbon-based gases: A review. Chemosphere 287, 131995 (2022).

Article ADS CAS PubMed Google Scholar

Gupta, V., Agarwal, S. & Saleh, T. A. Chromium removal by combining the magnetic properties of iron oxide with adsorption properties of carbon nanotubes. Water Res. 45, 2207–2212 (2011).

Article CAS PubMed Google Scholar

Gupta, V. K., Agarwal, S. & Saleh, T. Synthesis and characterization of alumina-coated carbon nanotubes and their application for lead removal. J. Hazard. Mater. 185, 17–23 (2011).

Article CAS PubMed Google Scholar

Darkrim, F. L., Malbrunot, P. & Tartaglia, G. Review of hydrogen storage by adsorption in carbon nanotubes. Int. J. Hydrog. Energy 27, 193–202 (2002).

Article CAS Google Scholar

Fiyadh, S. S. et al. Review on heavy metal adsorption processes by carbon nanotubes. J. Clean. Prod. 230, 783–793 (2019).

Article CAS Google Scholar

Stafiej, A. & Pyrzynska, K. J. S. Adsorption of heavy metal ions with carbon nanotubes. Sep. Purif. Technol. 58, 49–52 (2007).

Article CAS Google Scholar

Xu, J. et al. A review of functionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation, application, and mechanism. Chemosphere 195, 351–364 (2018).

Article ADS CAS PubMed Google Scholar

Min, X. et al. Ultra-high capacity of lanthanum-doped UiO-66 for phosphate capture: Unusual doping of lanthanum by the reduction of coordination number. J. Chem. Eng. 358, 321–330 (2019).

Article CAS Google Scholar

Ouyang, H. et al. Selective capture of toxic selenite anions by bismuth-based metal–organic frameworks. Angew. Chem. Int. Ed Engl. 57, 13197–13201 (2018).

Article CAS PubMed Google Scholar

Zhou, H.-C., Long, J. R. & Yaghi, O. M. Introduction to metal–organic frameworks. Chem. Rev. 112, 673–674 (2012).

Article CAS PubMed Google Scholar

Ahmed, I. & Jhung, S. H. Composites of metal–organic frameworks: Preparation and application in adsorption. Mater. Today 17, 136–146 (2014).

Article CAS Google Scholar

Burtch, N. C., Jasuja, H. & Walton, K. S. Water stability and adsorption in metal–organic frameworks. Chem. Rev. 114, 10575–10612 (2014).

Article CAS PubMed Google Scholar

Küsgens, P. et al. Characterization of metal-organic frameworks by water adsorption. Microporous Mesoporous Mater. 120, 325–330 (2009).

Article Google Scholar

Bordonhos, M., Lourenço, M., Gomes, J. R., Ferreira, P. & Pinto, M. L. Exploring periodic mesoporous organosilicas for ethane–ethylene adsorption–separation. Microporous Mesoporous Mater. 317, 110975 (2021).

Article CAS Google Scholar

Gao, J. et al. Clickable periodic mesoporous organosilicas: Synthesis, click reactions, and adsorption of antibiotics. Chem. Eur. J. 20, 1957–1963 (2014).

Article CAS PubMed Google Scholar

Grüning, W. R., Siddiqi, G., Safonova, O. V. & Copéret, C. Bipyridine periodic mesoporous organosilica: A solid ligand for the iridium-catalyzed borylation of CH bonds. Adv. Synth. Catal. 356, 673–679 (2014).

Article Google Scholar

Ide, M., De Canck, E., Van Driessche, I., Lynen, F. & Van Der Voort, P. Developing a new and versatile ordered mesoporous organosilica as a pH and temperature stable chromatographic packing material. RSC Adv. 5, 5546–5552 (2015).

Article ADS CAS Google Scholar

Martens, S. et al. Periodic mesoporous organosilicas as adsorbents of toxic trace gases out of the ambient air. Anorg. Allg. Chem. 640, 632–640 (2014).

Article CAS Google Scholar

Otero, R. et al. Adsorption of the herbicide S-metolachlor on periodic mesoporous organosilicas. Chem. Eng. J. 228, 205–213 (2013).

Article CAS Google Scholar

Munaweera, I., Hong, J., D’Souza, A. & Balkus, K. J. Novel wrinkled periodic mesoporous organosilica nanoparticles for hydrophobic anticancer drug delivery. J. Porous Mater. 22, 1–10 (2015).

Article CAS Google Scholar

Park, S. S., Santha Moorthy, M. & Ha, C.-S. Periodic mesoporous organosilicas for advanced applications. NPG Asia Mater. 6, e96 (2014).

Article CAS Google Scholar

Teng, Z. et al. Mesoporous organosilica hollow nanoparticles: Synthesis and applications. Adv. Mater. 31, 1707612 (2019).

Article Google Scholar

Van Der Voort, P. et al. Periodic mesoporous organosilicas: From simple to complex bridges; a comprehensive overview of functions, morphologies and applications. Chem. Soc. Rev. 42, 3913–3955 (2013).

Article PubMed Google Scholar

Ahmad, N., Chughtai, A. H., Younus, H. A. & Verpoort, F. Discrete metal-carboxylate self-assembled cages: Design, synthesis and applications. Coord. Chem. Rev. 280, 1–27 (2014).

Article CAS Google Scholar

Batten, S. R., Hoskins, B. F. & Robson, R. Two interpenetrating 3D networks which generate spacious sealed-off compartments enclosing of the order of 20 solvent molecules in the structures of Zn (CN)(NO3)(tpt) 2/3. cntdot. solv (tpt = 2, 4, 6-tri (4-pyridyl)-1, 3, 5-triazine, solv = . apprx. 3/4C2H2Cl4. cntdot. 3/4CH3OH or. apprx. 3/2CHCl3. cntdot. 1/3CH3OH). J. Am. Chem. Soc. 117, 5385–5386 (1995).

Article CAS Google Scholar

Kitagawa, S., Kawata, S., Nozaka, Y. & Munakata, M. Synthesis and crystal structures of novel copper (I) co-ordination polymers and a hexacopper (I) cluster of quinoline-2-thione. J. Chem. Soc. Dalton Trans. 9, 1399–1404 (1993).

Article Google Scholar

Chen, Q. et al. A controllable gate effect in cobalt (II) organic frameworks by reversible structure transformations. Angew. Chem. 125, 11764–11767 (2013).

Article ADS Google Scholar

Furukawa, H., Cordova, K. E., O’Keeffe, M. & Yaghi, O. M. The chemistry and applications of metal-organic frameworks. Science 341, 1230444 (2013).

Article PubMed Google Scholar

Li, J. R. et al. Selective gas adsorption and unique structural topology of a highly stable guest-free zeolite-type MOF material with N-rich chiral open channels. Chem. Eur. J. 14, 2771–2776 (2008).

Article CAS PubMed Google Scholar

Li, N., Xu, J., Feng, R., Hu, T.-L. & Bu, X.-H. Governing metal–organic frameworks towards high stability. Chem. Commun. 52, 8501–8513 (2016).

Article CAS Google Scholar

Wang, B., Côté, A. P., Furukawa, H., O’Keeffe, M. & Yaghi, O. M. Colossal cages in zeolitic imidazolate frameworks as selective carbon dioxide reservoirs. Nature 453, 207–211 (2008).

Article ADS CAS PubMed Google Scholar

Bergstrom, J. C. & Cordell, H. K. An analysis of the demand for and value of outdoor recreation in the United States. J. Leis. Res. 23, 67–86 (1991).

Article Google Scholar

Yaghi, O. M. et al. Reticular synthesis and the design of new materials. Nature 423, 705–714 (2003).

Article ADS CAS PubMed Google Scholar

Zhang, D.-J. et al. Solvothermal synthesis a novel hemidirected 2-D (3, 3)-net metal-organic framework [Pb (HIDC)] n based on the linkages of left-and right-hand helical chains. Inorg. Chem. Commun. 11, 192–195 (2008).

Article CAS Google Scholar

Sun, Y.-G. et al. Synthesis, crystal structures, and properties of novel heterometallic La/Pr−Cu−K and Sm/Eu/Tb−Cu coordination polymers. Cryst. Growth Des. 10, 1059–1067 (2010).

Article CAS Google Scholar

Sun, Y.-G. et al. Hydrothermal synthesis, structure, and photoluminescence of four complexes based on 1H-imidazole-4, 5-dicarboxylate or 1H-imidazole-2-carboxylate ligands. J. Coord. Chem. 63, 4188–4200 (2010).

Article CAS Google Scholar

Grünker, R. R. et al. A new metal–organic framework with ultra-high surface area. Chem. Commun. 50, 3450–3452 (2014).

Article Google Scholar

Deng, H. et al. Large-pore apertures in a series of metal-organic frameworks. Science 336, 1018–1023 (2012).

Article ADS CAS PubMed Google Scholar

Lan, Y. Q., Jiang, H. L., Li, S. L. & Xu, Q. J. A. M. Mesoporous metal-organic frameworks with size-tunable cages: Selective CO2 uptake, encapsulation of Ln3+ cations for luminescence, and column-chromatographic dye separation. Adv. Mater. 23, 5015–5020 (2011).

Article CAS PubMed Google Scholar

Chaemchuen, S., Kabir, N. A., Zhou, K. & Verpoort, F. Metal–organic frameworks for upgrading biogas via CO2 adsorption to biogas green energy. Chem. Soc. Rev. 42, 9304–9332 (2013).

Article CAS PubMed Google Scholar

Chaemchuen, S. et al. Tuning metal sites of DABCO MOF for gas purification at ambient conditions. Microporous Mesoporous Mater. 201, 277–285 (2015).

Article CAS Google Scholar

Kreno, L. E. et al. Metal–organic framework materials as chemical sensors. Chem. Rev. 112, 1105–1125 (2012).

Article CAS PubMed Google Scholar

Horcajada, P. et al. Porous metal–organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 9, 172–178 (2010).

Article ADS CAS PubMed Google Scholar

Liang, W. et al. 3D, eco-friendly metal-organic frameworks@ carbon nanotube aerogels composite materials for removal of pesticides in water. J. Hazard. Mater. 401, 123718 (2021).

Article CAS PubMed Google Scholar

Westerhoff, P., Alvarez, P., Li, Q., Gardea-Torresdey, J. & Zimmerman, J. Overcoming implementation barriers for nanotechnology in drinking water treatment. Environ. Sci. Nano 3, 1241–1253 (2016).

Article CAS Google Scholar

Notter, D. A., Mitrano, D. M. & Nowack, B. Are nanosized or dissolved metals more toxic in the environment? A meta-analysis. Environ. Toxicol. Chem. 33, 2733–2739 (2014).

Article CAS PubMed Google Scholar

Saboori, A. nanoparticle sorbent composed of MIL-101 (Fe) and dithiocarbamate-modified magnetite nanoparticles for speciation of Cr (III) and Cr (VI) prior to their determination by electrothermal AAS. Microchim. Acta 184, 1509–1516 (2017).

Article CAS Google Scholar

Bosch, M., Zhang, M. & Zhou, H.-C. Increasing the stability of metal-organic frameworks. Adv. Chem. 2014, 1155 (2014).

Article Google Scholar

Lu, M., Li, L., Shen, S., Chen, D. & Han, W. Highly efficient removal of Pb2+ by a sandwich structure of metal–organic framework/GO composite with enhanced stability. New J. Chem. 43, 1032–1037 (2019).

Article CAS Google Scholar

Joseph, J. et al. Iron-based metal–organic framework: Synthesis, structure and current technologies for water reclamation with deep insight into framework integrity. Chemosphere 284, 131171 (2021).

Article ADS CAS PubMed Google Scholar

Liu, Z., He, W., Zhang, Q., Shapour, H. & Bakhtari, M. F. Preparation of a GO/MIL-101 (Fe) composite for the removal of methyl orange from aqueous solution. ACS Omega 6, 4597–4608 (2021).

Article CAS PubMed PubMed Central Google Scholar

Megha, R. et al. Removal of heavy metals and dyes from its aqueous solution utilizing metal organic frameworks (MOFs). Mater. Today Proc. 77, 188 (2022).

Google Scholar

Manchanda, A. S. & Kruk, M. M. Synthesis of large-pore face-centered-cubic periodic mesoporous organosilicas with unsaturated bridging groups. Microporous Mesoporous Mater. 222, 153–159 (2016).

Article CAS Google Scholar

Hunks, W. J. & Ozin, G. A. Challenges and advances in the chemistry of periodic mesoporous organosilicas (PMOs). J. Mater. Chem. 15, 3716–3724 (2005).

Article CAS Google Scholar

Qian, K. et al. Pore size-optimized periodic mesoporous organosilicas for the enrichment of peptides and polymers. RSC Adv. 3, 14466–14472 (2013).

Article ADS CAS Google Scholar

Chen, Y. et al. Hollow mesoporous organosilica nanoparticles: A generic intelligent framework-hybridization approach for biomedicine. J. Am. Chem. Soc. 136, 16326–16334 (2014).

Article CAS PubMed Google Scholar

Manayil, J. C., Lee, A. F. & Wilson, K. Functionalized periodic mesoporous organosilicas: Tunable hydrophobic solid acids for biomass conversion. Molecules 24, 239 (2019).

Article PubMed PubMed Central Google Scholar

Borah, P., Ma, X., Nguyen, K. T. & Zhao, Y. vanadyl complex grafted to periodic mesoporous organosilica: A green catalyst for selective hydroxylation of benzene to phenol. Angew. Chem. Int. Ed. 124, 7876–7881 (2012).

Article ADS Google Scholar

Karimi, B. & Esfahani, F. K. Unexpected golden Ullmann reaction catalyzed by Au nanoparticles supported on periodic mesoporous organosilica (PMO). Chem. Commun. 47, 10452–10454 (2011).

Article CAS Google Scholar

Karimi, B., Mirzaei, H. M. & Mobaraki, A. Periodic mesoporous organosilica functionalized sulfonic acids as highly efficient and recyclable catalysts in biodiesel production. Catal. Sci. Technol. 2, 828–834 (2012).

Article CAS Google Scholar

Kumar, P. & Guliants, V. M. Periodic mesoporous organic–inorganic hybrid materials: Applications in membrane separations and adsorption. Microporous Mesoporous Mater. 132, 1–14 (2010).

Article CAS Google Scholar

Melero, J. A. et al. Synthesis and catalytic activity of organic–inorganic hybrid Ti-SBA-15 materials. J. Mater. Chem. 17, 377–385 (2007).

Article CAS Google Scholar

Waki, M. et al. Enhanced fluorescence detection of metal ions using light-harvesting mesoporous organosilica. Chem. Eur. J. 18, 1992–1998 (2012).

Article CAS PubMed Google Scholar

Yang, Q., Liu, J., Zhang, L. & Li, C. Functionalized periodic mesoporous organosilicas for catalysis. J. Mater. Chem. 19, 1945–1955 (2009).

Article CAS Google Scholar

Inagaki, S. et al. Light harvesting by a periodic mesoporous organosilica chromophore. Angew. Chem. Int. Ed. 121, 4102–4106 (2009).

Article ADS Google Scholar

Takeda, H. et al. Visible-light-harvesting periodic mesoporous organosilica. Chem. Commun. 40, 6032–6034 (2009).

Article Google Scholar

Ikai, M., Maegawa, Y., Goto, Y., Tani, T. & Inagaki, S. Synthesis of visible-light-absorptive and hole-transporting periodic mesoporous organosilica thin films for organic solar cells. J. Mater. Chem. A 2, 11857–11865 (2014).

Article CAS Google Scholar

Mizoshita, N., Ikai, M., Tani, T. & Inagaki, S. Hole-transporting periodic mesostructured organosilica. J. Am. Chem. Soc. 131, 14225–14227 (2009).

Article CAS PubMed Google Scholar

Mizoshita, N., Tani, T., Shinokubo, H. & Inagaki, S. Mesoporous organosilica hybrids consisting of silica-wrapped π–π stacking columns. Angew. Chem. Int. Ed. 124, 1182–1186 (2012).

Article ADS Google Scholar

Aggad, D. et al. Gemcitabine delivery and photodynamic therapy in cancer cells via porphyrin-ethylene-based periodic mesoporous organosilica nanoparticles. ChemNanoMater 4, 46–51 (2018).

Article CAS Google Scholar

Giret, S., Man, M. W. C. & Carcel, C. Mesoporous-silica-functionalized nanoparticles for drug delivery. Chem. Eur. J. 21, 13850–13865 (2015).

Article CAS PubMed Google Scholar

Knežević, N. Ž, Djordjević, S., Kojić, V. & Janaćković, D. Functionalized periodic mesoporous organosilica nanoparticles for loading and delivery of suramin. Inorganics 7, 16 (2019).

Article Google Scholar

Lin, C. X. C., Qiao, S. Z., Yu, C. Z., Ismadji, S. & Lu, G. Q. M. Periodic mesoporous silica and organosilica with controlled morphologies as carriers for drug release. Microporous Mesoporous Mater. 117, 213–219 (2009).

Article CAS Google Scholar

Wu, H. Y. et al. Synthesis, bifunctionalization, and remarkable adsorption performance of benzene-bridged periodic mesoporous organosilicas functionalized with high loadings of carboxylic acids. Chem. Eur. J. 19, 6358–6367 (2013).

Article CAS PubMed Google Scholar

Zhang, J. et al. Cisplatin and doxorubicin high-loaded nanodrug based on biocompatible thioether-and ethane-bridged hollow mesoporous organosilica nanoparticles. J. Colloid Interface Sci. 513, 214–221 (2018).

Article ADS CAS PubMed Google Scholar

Rebbin, V., Schmidt, R. & Fröba, M. Spherical particles of phenylene-bridged periodic mesoporous organosilica for high-performance liquid chromatography. Angew. Chem. Int. Ed. 45, 5210–5214 (2006).

Article CAS Google Scholar

Khademy, M., Karimi, B. & Zareian, S. Ionic liquid-based periodic mesoporous organosilica: An innovative matrix for enzyme immobilization. ChemistrySelect 2, 9953–9957 (2017).

Article CAS Google Scholar

Wang, W., Lofgreen, J. E. & Ozin, G. A. Why PMO? Towards functionality and utility of periodic mesoporous organosilicas. Small 6, 2634–2642 (2010).

Article CAS PubMed Google Scholar

Birault, A. et al. Large-pore periodic mesoporous organosilicas as advanced bactericide platforms. ACS Appl. Bio Mater. 1, 1787–1792 (2018).

Article CAS PubMed Google Scholar

Gehring, J. et al. Sunlight-triggered nanoparticle synergy: Teamwork of reactive oxygen species and nitric oxide released from mesoporous organosilica with advanced antibacterial activity. J. Am. Chem. Soc. 138, 3076–3084 (2016).

Article CAS PubMed Google Scholar

Norouzi, M. & Elhamifar, D. Phenylene and isatin based bifunctional mesoporous organosilica supported schiff-base/manganese complex: An efficient and recoverable nanocatalyst. Catal. Lett. 149, 619–628 (2019).

Article CAS Google Scholar

Shaker, M. & Elhamifar, D. Pd-containing IL-based ordered nanostructured organosilica: A powerful and recoverable catalyst for Sonogashira reaction. Tetrahedron Lett. 61, 152481 (2020).

Article CAS Google Scholar

Shaker, M. & Elhamifar, D. Cu-containing magnetic yolk-shell structured ionic liquid-based organosilica nanocomposite: A powerful catalyst with improved activity. Compos. Commun. 24, 100608 (2021).

Article Google Scholar

Norouzi, M., Elhamifar, D., Mirbagheri, R. & Materials, M. Phenylene-based periodic mesoporous organosilica supported melamine: An efficient, durable and reusable organocatalyst. Microporous Mesoporous Mater. 278, 251–256 (2019).

Article CAS Google Scholar

Elhamifar, D., Kazempoor, S. & Karimi, B. Amine-functionalized ionic liquid-based mesoporous organosilica as a highly efficient nanocatalyst for the Knoevenagel condensation. Catal. Sci. Technol. 6, 4318–4326 (2016).

Article CAS Google Scholar

Ardeshirfard, H. & Elhamifar, D. Ionic liquid/Mn complex immobilized on phenylene based periodic mesoporous organosilica: An efficient and reusable nanocatalyst for green oxidation of alcohols. Appl. Surf. Sci. Adv. 5, 100099 (2021).

Article Google Scholar

Elhamifar, D., Yari, O. & Karimi, B. Highly ordered mesoporous organosilica–titania with ionic liquid framework as very efficient nanocatalyst for green oxidation of alcohols. J. Colloid Interface Sci. 500, 212–219 (2017).

Article ADS CAS PubMed Google Scholar

Elhamifar, D. & Ardeshirfard, H. Phenyl and ionic liquid based bifunctional periodic mesoporous organosilica supported copper: An efficient nanocatalyst for clean production of polyhydroquinolines. J. Colloid Interface Sci. 505, 1177–1184 (2017).

Article ADS CAS PubMed Google Scholar

Neysi, M. & Elhamifar, D. Pd-containing magnetic periodic mesoporous organosilica nanocomposite as an efficient and highly recoverable catalyst. Sci. Rep. 12, 1–10 (2022).

Article Google Scholar

Pearson, R. G. Hard and soft acids and bases. J. Am. Chem. Soc. 85, 3533–3539 (1963).

Article CAS Google Scholar

Pearson, R. G. Hard and soft acids and bases, HSAB, part 1: Fundamental principles. J. Chem. Educ. 45, 581 (1968).

Article CAS Google Scholar

Pearson, R. G. Hard and soft acids and bases, HSAB, part II: Underlying theories. J. Chem. Educ. 45, 643 (1968).

Article CAS Google Scholar

Pearson, R. G. Recent advances in the concept of hard and soft acids and bases. J. Chem. Educ. 64, 561 (1987).

Article CAS Google Scholar

Croissant, J. G., Cattoën, X., Man, M. W. C., Durand, J.-O. & Khashab, N. M. Syntheses and applications of periodic mesoporous organosilica nanoparticles. Nanoscale 7, 20318–20334 (2015).

Article ADS CAS PubMed Google Scholar

Mizoshita, N., Tani, T. & Inagaki, S. Syntheses, properties and applications of periodic mesoporous organosilicas prepared from bridged organosilane precursors. Chem. Soc. Rev. 40, 789–800 (2011).

Article CAS PubMed Google Scholar

Diaz, U., Brunel, D. & Corma, A. Catalysis using multifunctional organosiliceous hybrid materials. Chem. Soc. Rev. 42, 4083–4097 (2013).

Article CAS PubMed Google Scholar

Fujita, S. & Inagaki, S. Self-organization of organosilica solids with molecular-scale and mesoscale periodicities. Chem. Mater. 20, 891–908 (2008).

Article CAS Google Scholar

Hatton, B., Landskron, K., Whitnall, W., Perovic, D. & Ozin, G. A. Past, present, and future of periodic mesoporous organosilicas the PMOs. Acc. Chem. Res. 38, 305–312 (2005).

Article CAS PubMed Google Scholar

Hoffmann, F., Cornelius, M., Morell, J. & Fröba, M. Silica-based mesoporous organic–inorganic hybrid materials. Angew. Chem. Int. Ed. 45, 3216–3251 (2006).

Article CAS Google Scholar

Jaroniec, M. Organosilica the conciliator. Nature 442, 638–640 (2006).

Article ADS CAS PubMed Google Scholar

Kaczmarek, A. M. & Van Der Voort, P. Light-emitting lanthanide periodic mesoporous organosilica (PMO) hybrid materials. Materials 13, 566 (2020).

Article ADS CAS PubMed PubMed Central Google Scholar

Xia, H.-S., Zhou, C.-H., Tong, D. S. & Lin, C. X. Synthesis chemistry and application development of periodic mesoporous organosilicas. J. Porous Mater. 17, 225–252 (2010).

Article CAS Google Scholar

Croissant, J. et al. Biodegradable ethylene-bis (propyl) disulfide-based periodic mesoporous organosilica nanorods and nanospheres for efficient in-vitro drug delivery. Adv. Mater. 26, 6174–6180 (2014).

Article CAS PubMed Google Scholar

De Canck, E., Ascoop, I., Sayari, A. & Van Der Voort, P. Periodic mesoporous organosilicas functionalized with a wide variety of amines for CO2 adsorption. Phys. Chem. Chem. Phys. 15, 9792–9799 (2013).

Article PubMed Google Scholar

Noureddine, A., Trens, P., Toquer, G., Cattoën, X. & Man, M. W. C. Tailoring the hydrophilic/lipophilic balance of clickable mesoporous organosilicas by the copper-catalyzed azide–alkyne cycloaddition click-functionalization. Langmuir 30, 12297–12305 (2014).

Article CAS PubMed Google Scholar

Corma, A., Díaz, U., Arrica, M., Fernández, E. & Ortega, Í. Organic–inorganic nanospheres with responsive molecular gates for drug storage and release. Angew. Chem. 121, 6365–6368 (2009).

Article ADS Google Scholar

Teng, Z. et al. A facile multi-interface transformation approach to monodisperse multiple-shelled periodic mesoporous organosilica hollow spheres. J. Am. Chem. Soc. 137, 7935–7944 (2015).

Article CAS PubMed Google Scholar

Benítez, A. et al. MIL-88A metal-organic framework as a stable sulfur-host cathode for long-cycle Li-S batteries. Nanomaterials 10, 424 (2020).

Article PubMed PubMed Central Google Scholar

Amaro-Gahete, J. et al. Fast ultrasound-assisted synthesis of highly crystalline MIL-88A particles and their application as ethylene adsorbents. Ultrason. Sonochem. 50, 59–66 (2019).

Article CAS PubMed Google Scholar

Wang, D. et al. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 519, 273–284 (2018).

Article ADS CAS PubMed Google Scholar

Jeong, E.-Y. & Park, S.-E. Synthesis of porphyrin-bridged periodic mesoporous organosilica and their catalytic applications. Res. Chem. Intermed. 38, 1237–1248 (2012).

Article CAS Google Scholar

Sanchez, C., Julián, B., Belleville, P. & Popall, M. Applications of hybrid organic–inorganic nanocomposites. J. Mater. Chem. 15, 3559–3592 (2005).

Article CAS Google Scholar

Yaghoubi, A. & Dekamin, M. G. Green and facile synthesis of 4H-pyran scaffold catalyzed by pure nano-ordered periodic mesoporous organosilica with isocyanurate framework (PMO-ICS). ChemistrySelect 2, 9236–9243 (2017).

Article CAS Google Scholar

Kim, D., Lee, G., Oh, S. & Oh, M. Unbalanced MOF-on-MOF growth for the production of a lopsided core–shell of MIL-88B@ MIL-88A with mismatched cell parameters. Chem. Commun. 55, 43–46 (2019).

Article CAS Google Scholar

Benítez, A. et al. MIL-88A metal-organic framework as a stable sulfur-host cathode for long-cycle Li-S batteries. J. Nanomater. 10, 424 (2020).

Article Google Scholar

Yang, X. et al. Surface functional groups of carbon-based adsorbents and their roles in the removal of heavy metals from aqueous solutions: A critical review. J. Chem. Eng. 366, 608–621 (2019).

Article CAS Google Scholar

Wang, N. et al. Preparation and application of a xanthate-modified thiourea chitosan sponge for the removal of Pb (II) from aqueous solutions. Ind. Eng. Chem. Res. 55, 4960–4968 (2016).

Article CAS Google Scholar

Xu, J., Wang, L. & Zhu, Y. Decontamination of bisphenol A from aqueous solution by graphene adsorption. Langmuir 28, 8418–8425 (2012).

Article CAS PubMed Google Scholar

Foo, K. Y. & Hameed, B. H. Insights into the modeling of adsorption isotherm systems. J. Chem. Eng. 156, 2–10 (2010).

Article CAS Google Scholar

Mane, V. S., Mall, I. D. & Srivastava, V. C. Use of bagasse fly ash as an adsorbent for the removal of brilliant green dye from aqueous solution. Dyes Pigm. 73, 269–278 (2007).

Article CAS Google Scholar

Bazargan-Lari, R., Zafarani, H. R., Bahrololoom, M. E. & Nemati, A. Removal of Cu (II) ions from aqueous solutions by low-cost natural hydroxyapatite/chitosan composite: Equilibrium, kinetic and thermodynamic studies. J. Taiwan Inst. Chem. Eng. 45, 1642–1648 (2014).

Article CAS Google Scholar

Ghaedi, A. M. et al. Factorial experimental design for the optimization of highly selective adsorption removal of lead and copper ions using metal organic framework MOF-2 (Cd). J. Mol. Liq. 272, 15–26 (2018).

Article CAS Google Scholar

Xu, Y. et al. Magnetization of a Cu (II)-1, 3, 5-benzenetricarboxylate metal-organic framework for efficient solid-phase extraction of Congo Red. Microchim. Acta 182, 2313–2320 (2015).

Article CAS Google Scholar

Ai, J. et al. Porous anionic uranyl–organic networks for highly efficient Cs+ adsorption and investigation of the mechanism. Inorg. Chem. 57, 4419–4426 (2018).

Article CAS PubMed Google Scholar

Kim, S. et al. Aqueous removal of inorganic and organic contaminants by graphene-based nanoadsorbents: A review. Chemosphere 212, 1104–1124 (2018).

Article ADS CAS PubMed Google Scholar

Yap, P. L. et al. All-in-one bioinspired multifunctional graphene biopolymer foam for simultaneous removal of multiple water pollutants. Adv. Mater. Interfaces 7, 2000664 (2020).

Article CAS Google Scholar

Komárek, M., Koretsky, C. M., Stephen, K. J., Alessi, D. S. & Chrastny, V. Competitive adsorption of Cd (II), Cr (VI), and Pb (II) onto nanomaghemite: A spectroscopic and modeling approach. Environ. Sci. Technol. 49, 12851–12859 (2015).

Article ADS PubMed Google Scholar

Ni, Y. et al. Urchin-like Ni–P microstructures: Facile synthesis, properties and application in the fast removal of heavy-metal ions. Chem. Commun. 47, 5891–5893 (2011).

Article CAS Google Scholar

Chen, W. et al. Enhanced removal of lead ions from aqueous Zhao solution by iron oxide nanomaterials with cobalt and nickel doping. J. Clean. Prod. 211, 1250–1258 (2019).

Article CAS Google Scholar

Gong, B. et al. Gram-scale synthesis of monodisperse sulfonated polystyrene nanospheres for rapid and efficient sequestration of heavy metal ions. Chem. Commun. 53, 12766–12769 (2017).

Article CAS Google Scholar

Kim, H. J. et al. Synthesis and adsorption properties of gelatin-conjugated hematite (α-Fe2O3) nanoparticles for lead removal from wastewater. J. Hazard. Mater. 416, 125696 (2021).

Article CAS PubMed Google Scholar

Liu, J. et al. Resource utilization of swine sludge to prepare modified biochar adsorbent for the efficient removal of Pb (II) from water. J. Clean. Prod. 257, 120322 (2020).

Article CAS Google Scholar

Download references

Sara S. E. Ghodsinia was partially supported by a grant from Ferdowsi University of Mashhad (No. FUM-54707).

Department of Chemistry, Faculty of Science, Ferdowsi University of Mashhad, Mashhad, 9177948974, Iran

Sara S. E. Ghodsinia, Hossein Eshghi & Arezou Mohammadinezhad

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

S.S.E.G.: wrote the main manuscript text and prepared all the results. H.E.: the corresponding author and leader in this study. A.M.: did data control. All authors reviewed the manuscript.

Correspondence to Hossein Eshghi.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Ghodsinia, S.S.E., Eshghi, H. & Mohammadinezhad, A. Synthesis of double-shelled periodic mesoporous organosilica nanospheres/MIL-88A-Fe composite and its elevated performance for Pb2+ removal in water. Sci Rep 13, 8092 (2023). https://doi.org/10.1038/s41598-023-35149-w

Download citation

Received: 16 December 2022

Accepted: 13 May 2023

Published: 19 May 2023

DOI: https://doi.org/10.1038/s41598-023-35149-w

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

SHARE