Next Article in Journal
Production Optimization in a Grain Facility through Mixed-Integer Linear Programming
Next Article in Special Issue
Extraction Study of Lignite Coalbed Methane as a Potential Supplement to Natural Gas for Enhancing Energy Security of Western Macedonia Region in Greece
Previous Article in Journal
Experimental and Modeling of Residual Deformation of Soil–Rock Mixture under Freeze–Thaw Cycles
Previous Article in Special Issue
Toxic Elements Behavior during Plasma Treatment for Waste Collected from Power Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

UiO-66/Palygorskite/TiO2 Ternary Composites as Adsorbents and Photocatalysts for Methyl Orange Removal

by
Thaleia Ioannidou
1,
Maria Anagnostopoulou
2,
Dimitrios Papoulis
3,
Konstantinos C. Christoforidis
1,2,* and
Ioanna A. Vasiliadou
1,*
1
Department of Environmental Engineering, Democritus University of Thrace, 67100 Xanthi, Greece
2
Institute of Chemistry and Processes for Energy, Environment and Health (ICPEES) UMR7515 CNRS, ECPM, University of Strasbourg, 25 rue Becquerel, CEDEX 2, 67087 Strasbourg, France
3
Department of Geology, University of Patras, 26504 Patras, Greece
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2022, 12(16), 8223; https://doi.org/10.3390/app12168223
Submission received: 19 July 2022 / Revised: 11 August 2022 / Accepted: 15 August 2022 / Published: 17 August 2022
(This article belongs to the Special Issue Treatment of Wastes and Energy Recovery)

Abstract

:
Metal–organic frameworks are recognized as a new generation of emerging porous materials in a variety of applications including adsorption and photocatalysis. The present study presents the development of ternary composite materials made through the coupling of UiO-66 with palygorskite (Pal) clay mineral and titanium dioxide (TiO2) applied as adsorbent and photocatalyst for the removal of methyl orange (MO) from aqueous solutions as a typical anionic dye. The prepared materials were characterized using XRD, ATR, DR UV/Vis, and TGA analysis. Detailed kinetic experiments revealed that the presence of the clay at low amounts in the composite outperformed the adsorption efficiency of pure UiO-66, increasing MO adsorption by ca. 8%. In addition, coupling Pal/UiO-66 with TiO2 for the production of ternary composites provided photocatalytic properties that resulted in complete removal of MO. This was not observed in the pure UiO-66, the Pal/UiO-66 composite, or the pure TiO2 material. This study presents the first example of clay mineral/MOF/TiO2 composites with improved performance in removing dyes from aqueous solutions and highlights the importance of coupling MOFs with low-cost clay minerals and photocatalysts for the development of multifunctional advanced composites.

1. Introduction

The decrease in environmental quality made clear the need to fill the void between development and sustainability. High amounts of pollutants end up in the aquatic environment [1]. Among the different xenobiotics that enter into the environment, synthetic dyes are a common class of organic pollutants with high concentrations in wastewater since they are used in many different industries (paper, printing, textiles, food, and pharmaceutical industry) [2,3,4,5]. Different technologies have been suggested for dye removal from effluents, among which adsorption [6,7,8,9] and photocatalysis [10,11,12] are considered suitable mostly due to the environmentally friendly character of the process. Adsorption offers the benefit of simplicity, high efficiency, and cost-effectiveness. On the other hand, photocatalysis is highly favorable when solid materials are applied as photocatalysts and when solar light is used for the decomposition of organic pollutants.
In both processes, efficiency is driven by the properties of the material used. Regarding adsorption, many different types of materials have been developed and applied in dye removal from the aquatic phase [13,14,15]. Metal–organic frameworks (MOFs) have been effectively applied as solid absorbing media including dye removal [2,16,17,18,19,20,21]. MOFs are highly porous synthetic materials that offer the benefit of chemical, textural, and compositional tunability [2]. In addition, they have also been applied as photocatalysts for the degradation of organic dyes [11,22] and have been coupled efficiently with traditional photocatalysts to improve performance [16,17,23,24] or the development of multifunctional composites [17,25]. Clay minerals have been also extensively studied as adsorbents for dyes removal [19,26,27]. They are low-cost, nontoxic materials with significant cation exchange capacity. However, their negative charge in neutral conditions limits their application as adsorbents of anionic chemical species [28,29]; therefore, chemical modification is required [30,31].
TiO2 is the most used photocatalyst owing to its inherent properties such as high photocatalytic activity and nontoxicity [32,33,34]. Among the different polymorphs of TiO2, anatase is preferred mostly due to the better charge handling properties [33]. Numerous studies have applied TiO2-based materials for the degradation of dyes [35]. Formation of composites through the coupling of TiO2 with MOF structures has also been exploited in photocatalytic reactions [11,16,17,22,24,25,36]. Clay mineral/TiO2 composites have also proven to be more active than the individual counter parts in a variety of photocatalytic reactions [37,38,39]. Since adsorption is a prerequisite for photocatalytic degradation reactions, adsorption of dyes on TiO2 [40,41] or MOF/TiO2 composites [42] has also been investigated. As expected, pure TiO2 materials present much lower adsorption capacity than MOF structures. Recently, Wu et al. showed that coupling MIL-101 (Cr) with TiO2 increased the adsorption performance against methyl orange (MO) (Scheme 1) [42]. In addition, the adsorption of cationic dyes onto UiO-66/TiO2 composites was not affected by the presence of TiO2, while the composites presented superior photocatalytic activity [25]. Clay/MOF composites have also been developed and studied as adsorbents. However, the effect of the presence of both clay minerals and TiO2 has never been studied.
Herein, novel ternary composites bearing UiO-66, palygorskite, and TiO2 were developed using wet chemical processes. The materials were characterized and tested as adsorbents and photocatalysts for the removal of MO from aqueous solutions. The presence of palygorskite in the composite increased the adsorption efficiency while TiO2 provided photocatalytic properties able to decompose MO. To the best of our knowledge, such composites made of these particular phases have never been reported before.

2. Materials and Methods

All chemicals were used as received without further purification. Palygorskite (Pal, from Ventzia continental basin, Western Macedonia, Greece) was obtained as follows: particles <2 μm were collected by gravity sedimentation, and the clay fraction was separated using centrifugation methods [38,39]. Only the clay fraction of the sample was used for the preparation of the composites in order to avoid other mineral impurities.

2.1. Synthesis of UiO-66

Synthesis of UiO-66 (U) was based on a previous reported process [17,43]. Briefly, 1.25 g of zirconium (IV) chloride (ZrCl4, 99.5%, Aldrich, Germany) was dissolved in 50 mL of N,N-dimethyl formamide (DMF, 99.9%, Aldrich) followed by the addition of 10 mL of acetic acid, forming solution A. On the other hand, 1.23 g of terephthalic acid (Aldrich) was mixed in 100 mL of DMF, forming solution B. The two solutions were mixed and heated under stirring to 150 °C for 24 h in an oil bath. The material formed was collected by centrifugation and washed three times with ethanol and with acetone. Solvent exchange with acetone was performed over a period of 4 days. The solvent was changed every second day. Finally, the material was dried at 100 °C over night and stored in a desiccator.

2.2. Synthesis of TiO2

Anatase TiO2 (T) was synthesized using the microemulsion process [44]. Titanium(IV) isopropoxide (Aldrich, Germany) was added to an inverse emulsion of 50 mL of water and n-heptane (85/10 v/v vs. H2O). The water/Ti molar ratio was equal to 100. Triton X-100 was used as a surfactant in 1-hexanol (105/100 v/v vs. surfactant). The solution was stirred for 24 h at room temperature. The material was collected and washed thoroughly with methanol by centrifugation and dried at 110 °C overnight. Anatase TiO2 was obtained by calcination for 2 h at static air and 400 °C using a ramp of 1 °C·min−1.

2.3. Synthesis of Binary Palygorskite/UiO-66 Composite

For the development of the palygorskite/UiO-66 composite (PU), a similar synthesis process was used with a slide modification. Firstly, 0.1 g of palygorskite was added to the ZrCl4 solution, and the amounts of ZrCl4 and terephthalic acid were adjusted, keeping the same ratio, such that the final composite contained a theoretical nominal amount of 10% by weight of palygorskite. The same washing and activation process was used as in the U sample.

2.4. Synthesis of Palygorskite/UiO-66/TiO2 Composite

The Pal/UiO-66/TiO2 ternary composite (PUT) was developed using a facile wet impregnation method. Preformed PU and TiO2 were mixed at a mass ratio equal to 6/4 in 100 mL of double-distilled water. The mixture was first sonicated for 30 min and then stirred overnight. Mixing was followed by removing the solvent using rotary evaporation, drying overnight at 100 °C, and a thermal treatment step at 200 °C for 4 h (ramp 1 °C·min−1).

2.5. Materials Characterization

X-ray diffraction (XRD) measurements were performed using a Bruker D8-Advance diffractometer equipped with a Lynx Eye detector operating at 40 kV and 40 mA, using Kα radiation of Cu at 1.5418°. A Thermo Fisher Nicolet iS10 spectrometer was used to obtain the attenuated total reflectance (ATR) IR spectrum in the 400–4000 cm−1 range. Thermogravimetric analysis (TGA) was performed using a TA Instrument Q5000IR. Each sample was heated under air flow (25 mL·min−1) up to 650 °C with a dynamic ramping rate depending on the weight loss. Diffuse reflectance UV/Vis absorption spectra were recorded in the range of 200–800 nm using a PerkinElmer Lambda 950 Scan spectrophotometer equipped with integrating sphere. For this analysis, BaSO4 was used as the reference. An X-band Bruker ER200D electron paramagnetic resonance (EPR) spectrometer operating at the X-band frequencies was used to perform spin trap experiments. The 5,5-dimethyl-1-pyrroline N-oxide (DMPO) was used as spit trap for the detection of hydroxyl radicals ( OH ) at room temperature. A suspension of 1 mg·mL−1 of the catalyst in water was sonicated for 10 min. Then, 100 μL of the above suspension was mixed with 100 μL of 0.005 M DMPO directly in the EPR flat cell. A quartz EPR flat cell was used.

2.6. Adsorption Experiments

The material (20 mg) was placed in an Erlenmeyer flask containing 50 mL of 90 mg·L−1 MO aqueous solution. The pH of the suspension was adjusted to 5. Adsorption experiments were performed in a temperature-controlled incubator. The temperature was maintained at 25 °C, and the adsorption process was performed under dark conditions throughout the whole experiment. For the preparation of all solutions, Milli-Q water was used. Adsorption of MO was quantified through the following process; aliquots of 1.5 mL were periodically taken, filtered using a 0.45 μm PTFE Millipore disc, and measured using a Varian Cary 100 Scan spectrophotometer. Quantification of MO was based on comparison with standards. The adsorption capacity evaluation was performed using the following equation:
q e = ( C 0 C e ) × V m ,
where qe (mg·g−1) is the adsorbed amount at equilibrium, C0 and Ce (mg·L−1) are the initial dye concentrations and at equilibrium, respectively, V is the volume of the mixture (mL), and m is the mass of the adsorbent (mg). The removal efficiency was calculated using the following equation:
RE = ( C 0 C e ) × 100 C 0 .

2.7. Photocatalytic Reactions

Photocatalytic tests were carried out after the adsorption process reached equilibrium. In brief, after incubating the material with the MO solution under dark conditions for 1.5 h, the mixture was irradiated using six UV lamps (Philips 24W/10/4P lamps, λmax = 365 nm) symmetrically positioned in the periphery of the photoreactor. The temperature of the mixture was maintained at 25 °C with water circulation by means of a cryostat. Reactions were performed at ambient pressure. Samples were periodically taken and filtered using a 0.45 μm PTFE Millipore disc. The UV/Vis spectra were then collected using a Varian Cary 100 Scan spectrophotometer.

2.8. Recycling of the Materials

Recycling was performed by redispersing the materials in water at pH 10 to accelerate the desorption process followed by sonication for 10 min and washing thoroughly via centrifugation until pH was neutral. Then, the material was washed ethanol three times and finally dried at 100 °C under vacuum for 5 h.

3. Results and Discussion

3.1. Materials Characterization

Structural information of the prepared materials was first obtained by means of X-ray diffraction (XRD) analysis. Figure 1 presents the diffraction patters of all materials. A typical diffraction pattern for the pure UiO-66 was observed [25,45]. The PU composite presented the same diffraction pattern as the U sample. This observation verified that the presence of Pal did not affect the synthesis of UiO-66. No diffraction peak attributed to Pal was detected in the PU sample. Two reasons are probably responsible for this. Firstly, the amount of Pal in the composite was small. Secondly, there was an intense diffraction peak from UiO-66 in the ca. 8.5° region that overlapped with the main diffraction peak of Pal. TiO2 presented the typical diffraction pattern of anatase. The PUT composite had a diffraction pattern that was a combination of the pure UiO-66 and TiO2 sample, without any obvious differences compared with the patterns of their pure counterparts.
The materials were further characterized using attenuated total reflection Fourier-transform infrared (ATR-FTIR) spectroscopy, and the data are given in Figure 2a. The characteristic peaks of UiO-66 were observed in the materials containing this specific MOF (U, PU, and PUT) [46,47,48]. The symmetric and asymmetric stretch vibrations in the COO group of the organic linker were observed at 1392 cm−1 and 1582 cm−1, while the vibrational mode of the C=C bond in the benzene ring was centered at 1510 cm−1. Vibrational modes of the C–H, C=C, O–H, and OCO bonds were detected in the region 810–710 cm−1. The broad peak at the beginning of the spectra (400–800 cm−1) was attributed to the presence of TiO2 [44,49]. Regarding the Pal sample, an intense peak was observed in the region 97–1060 cm−1. This was not observed in the pure UiO-66 sample but was present in all composites bearing Pal (i.e., PU and PUT). This peak was attributed to the Si–O–Mg asymmetric stretching vibration [39,46,50] and verified the presence of Pal in the composite materials. It should be noted that the peak centered at 1582 cm−1 in the UiO-66 sample was shifted to higher wavenumbers in the composite samples bearing Pal (Figure 2b). This is probably linked with the in-situ development of UiO-66 in the presence of Pal and suggests an interaction of the clay mineral with the carboxylate groups of the organic linker in the MOF structure. Overall, ATR verified the presence of UiO-66, Pal, and TiO2 in the composite materials.
In order to get information regarding the amount of the different parts in the composites, the prepared materials were subjected to thermogravimetric (TGA) analysis. The TGA profiles of all materials are illustrated in Figure 3. A smooth mass decrease was observed in the pure TiO2 that ended at ca. 7% total mass loss at 550 °C. The mass loss was greater in the Pal sample (ca. 18% at 550 °C). In this temperature range, the mass loss in the Pal sample was attributed to both superficially adsorbed and loosely bound zeolitic water, as well as structural OH2 [51,52]. In the materials containing UiO-66, the small weight loss in the temperature range 50–400 °C was due to the loss of guest molecules (DMF). The fast drop in mass loss at temperatures higher than 450 °C was attributed to the decomposition of the organic framework of UiO-66. This second drop in mass was observed at lower temperatures in the composite bearing TiO2. Taking into account the mass loss in the composites at temperatures higher than 400 °C (indicated by Δmass in Figure 3 for sample U) and the corresponding weight loss in the pure Pal and TiO2, the Pal content was calculated equal to 7 wt.% in the PU and PUT samples, and the TiO2 content was calculated equal to 39 wt.% in the PUT composite.
Diffuse reflectance UV/Vis spectroscopy was applied to study the electronic properties related with light absorption. Figure 4a presents the UV/Vis absorption spectra of the materials. The corresponding Tauc plots [38,39,53] used to estimate the bandgap energies (Eg) are given in Figure 4b. Materials containing TiO2 were treated as direct, while the remaining materials were treated as indirect semiconductors [17]. Pure UiO-66 absorbed light at wavelengths lower than 310 nm in accordance with the literature [24]. The same absorption onset was observed for the PU sample. This corresponded to Eg values of ca. 4 eV. The absorption edge of TiO2 was observed at ca. 390 nm, while Pal absorbed light in the whole visible region [38,39]. Light absorption of the composites was a combination of the absorption of the individual counter parts. The absorption onset of the PUT material was observed at ca. 390 nm, as for the pure TiO2. This corresponded to an Eg value of 3.18 eV, typical for pure anatase TiO2 [16]. A broad absorption was observed in visible region (400–800 nm) for both PU and PUT samples. This was strictly attributed to the presence of Pal. Overall, it was verified through DR UV/Vis spectroscopy that all parts in the composite contributed to the light absorption properties of the material.

3.2. Dye Adsorption Evaluation

The effect of contact time on MO adsorption over the prepared materials is given in Figure 5a. Adsorption equilibrium was achieved within less than 5 min for all materials containing UiO-66. On the contrary, equilibrium was reached at longer reaction times for pure Pal and TiO2. Significant differences were also detected in the experimentally observed adsorption capacity at equilibrium (qe,exp). The qe,exp together with the removal efficiency (RE %) are given in Table 1. Both TiO2 and Pal presented low qe and removal efficiency. On the contrary, the pure U and PU composite presented high adsorption capacity, with the PU sample being the most efficient (qe = 213.3 mg·g−1). This suggests that the presence of Pal at low amounts in the composite was beneficial for MO adsorption. An increase in the adsorption capacity of dyes was previously shown in oxide/MOFs or clay mineral/MOF composites including UiO-66 [25,42,54,55], as well as in other MOF-based composites [20,21]. In the specific case of Pal/UiO-66 composites, formation of microporosity at the interface of the two parts was shown to increase the specific surface area that may affect adsorption efficiency [56]. A decrease in the qe and the removal efficiency was observed in the composite bearing TiO2 (i.e., PUT). This was probably related to the relatively high amount of TiO2 in the PUT composite that presented low qe. However, the removal efficiency in these composites was not a linear combination of the individual counter parts considering their content in the composites. Taking into account the fast kinetic profile of MO adsorption, the high qe, and the removal efficiency, the adsorption behavior of the PUT composite was dominated by UiO-66. Lastly, in order to verify the stability of the composites, recycling adsorption experiments using the most efficient composite were performed (Figure S1). The stability was evidenced by the small loss of adsorption efficiency after three consecutive cycles [56].
The pseudo-second-order kinetic model [54,55,57,58,59] was used to fit the experimental data as represented by Equation (3).
t q t = 1 k 2 q e 2 + t q e ,
where qe is the amount of MO adsorbed at equilibrium, qt the MO adsorbed at time t, and k2 is the rate constant (g·mg−1·min−1) of the second-order adsorption kinetics. The fitted adsorption kinetic curves for the materials containing UiO-66 as a function of time are given in Figure 5b. The correlation coefficient R2 together with the calculated adsorption at equilibrium (qe,cal) is given in Table 1. High correlation coefficients were obtained for all materials. In addition, qe,cal matched the experimental observation. These findings indicate that MO adsorption onto the pure U sample and all composites containing UiO-66 followed the pseudo-second-order kinetic model [54,55,57,58]. Overall, the adsorption experiments showed that Pal induced an increase in MO adsorption, while the presence of ca. 40 wt.% TiO2 did not significantly affect the adsorption given its relatively high content in the composite and the low adsorption capacity.

3.3. Photocatalytic Activity Evaluation

The materials were further tested as photocatalysts for the degradation of MO. Before starting the catalytic process, the mixture containing 90 mg·L−1 MO and 20 mg of the material was stirred in dark conditions for 1.5 h, as indicated by the adsorption experiments. The shaded area in Figure 6a corresponds to the adsorption process. After adsorption/desorption equilibrium was achieved, the sample was irradiated with UV light. The kinetics of MO photocatalytic degradation was monitored for 2 h. The data are presented in Figure 6a. No obvious degrease in MO concentration was detected by the Pal, pure UiO-66, and PU composite. Clay minerals do not possess any photocatalytic activity. On the contrary, UiO-66 is photoactive [37,53]. However, on the basis of the U and PU light absorption properties (Figure 4), light irradiation at wavelengths lower than 310 nm was required for their activation. Therefore, it is not surprising that U and PU were not active under the conditions of our experiment. MO concentration was reduced when materials containing TiO2 were used. This implies that PUT is photoactive and able to degrade MO. UiO-66/TiO2 composites were previously proven to be active in photocatalytic degradation of dyes [25,60]. Pure TiO2 removed approximately 50% of the initial MO. Comparing the materials presenting photoactivity, complete decolorization was achieved only in the PUT composite within less than 2 h irradiation. The composite bearing UiO-66 and TiO2 (UT) also presented photoactivity, but the removal efficiency was lower than that of the PUT composite. The high removal efficiency of PUT against MO was linked to both functions of the PUT composite (i.e., adsorbent and photocatalyst). Most of the MO (77.5%) was adsorbed, while the remainder was decomposed via photocatalysis.
In order to further evaluate the photocatalytic efficiency of the prepared materials, the photocatalytic degradation rates were calculated. The data were well fitted with a fist-order kinetic model. Rate constants (k) were estimated by fitting the experimental data with the following equation:
ln C 0 C = kt ,
where k is the apparent first-order rate constant (min−1), t is the irradiation time (min), C0 is the initial concentration (i.e., after the adsorption/desorption process reached equilibrium), and C is the concentration of MO at reaction time t. The data are presented in Figure 6b, and the photocatalytic degradation rates (k) are summarized in Table 1. The PUT composite presented higher degradation rates than the pristine TiO2. This has previously been seen in UiO-66/TiO2 [24,60] composites, as well as in clay/TiO2 or TiO2-based composites containing other natural porous minerals [37,38,39,61].
The improved photoactivity in TiO2 composites has been linked with synergistic effects related to the increase in charge availability [11,25,45], as well as the deposition of TiO2 nanoparticles on the support, which decreases agglomeration phenomena [37,39]. Both factors are important for photocatalytic reactions in the liquid phase [12]. An additional contributor to the improved photoactivity was the increase in the contact between TiO2 and the substrate [17,53]. This could be achieved through the increase in MO concentration close to TiO2 due to the high adsorption efficiency of UiO-66. The large size of MO molecular structure and the fast adsorption kinetics suggest that MO was most likely adsorbed onto the external surface of UiO-66 where TiO2 particles were present. This potentially increased the contact between TiO2 and MO, improving photoactivity.
To further verify the photocatalytic activity of the composites, in situ spin-trap electron paramagnetic resonance (EPR) spectroscopy was performed using DMPO to study the reactive oxygen species formed under irradiation. Figure 7 presents the EPR spectrum obtained at room temperature after 15 min irradiation of a PUT aqueous suspension under light irradiation conditions identical to the photocatalytic experiments. An EPR signal with a 1:2:2:1 intensity ratio corresponding to the DMPO–OH spin adduct was observed. These data clearly indicate that the PUT composite was able to produce OH under irradiation. A similar analysis was performed for the PU composite. However, no DMPO–OH was detected (Figure 7). This suggests that TiO2 was required in the composite to produce reactive oxygen species. These data may explain the inability of PU and pure U to decompose MO under the conditions of our experiment.

4. Conclusions

A facile two-step synthesis process was applied for the synthesis of clay mineral/MOF/TiO2 ternary composites as adsorbents and photocatalysts for the removal of MO from aqueous solutions. Pure Pal and TiO2 presented low adsorption capacities. Low amounts of Pal in the composite increased adsorption by ca. 8%. The presence of TiO2 at 40 wt.% reduced the adsorption efficiency but not significantly. In addition, it induced photocatalytic properties in the composite. Materials made of Pal, UiO-66, and TiO2 completely removed MO, while pure TiO2 removed ca. 50% of the initial dye concentration. This study highlights the importance of composites for the development of efficient materials presenting multifunctionality and presents the first example of coupling UiO-66 with low-cost clay minerals and TiO2 for the development of dual-function advanced materials.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/app12168223/s1: Figure S1. Percentage removal efficiency of MO by the PU composite over three consecutive cycles.

Author Contributions

Conceptualization, I.A.V.; methodology, K.C.C. and I.A.V.; validation, I.A.V., K.C.C., and D.P.; formal analysis, I.A.V.; investigation, T.I., M.A., and I.A.V.; resources, I.A.V., K.C.C., and D.P.; data curation, I.A.V. and K.C.C.; writing—original draft preparation, I.A.V.; writing—review and editing, I.A.V. and K.C.C.; visualization, I.A.V. and M.A.; supervision, I.A.V. and K.C.C.; project administration, I.A.V. and K.C.C.; funding acquisition, K.C.C. and I.A.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work benefited from state aid managed by the French National Research Agency (ANR) under the Investments for the Future (program “Make Our Planet Great Again”) with reference ANR-18-MOPGA-0014. The research was partially funded by Democritus University of Thrace (project 82268).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saravanan, A.; Kumar, P.S.; Yaashikaa, P.R.; Karishma, S.; Jeevanantham, S.; Swetha, S. Mixed Biosorbent of Agro Waste and Bacterial Biomass for the Separation of Pb(II) Ions from Water System. Chemosphere 2021, 277, 130236. [Google Scholar] [CrossRef] [PubMed]
  2. Uddin, M.J.; Ampiaw, R.E.; Lee, W. Adsorptive Removal of Dyes from Wastewater Using a Metal-Organic Framework: A Review. Chemosphere 2021, 284, 131314. [Google Scholar] [CrossRef] [PubMed]
  3. Murugesan, A.; Loganathan, M.; Senthil Kumar, P.; Vo, D.-V.N. Cobalt and Nickel Oxides Supported Activated Carbon as an Effective Photocatalysts for the Degradation Methylene Blue Dye from Aquatic Environment. Sustain. Chem. Pharm. 2021, 21, 100406. [Google Scholar] [CrossRef]
  4. Renita, A.A.; Vardhan, K.H.; Kumar, P.S.; Ngueagni, P.T.; Abilarasu, A.; Nath, S.; Kumari, P.; Saravanan, R. Effective Removal of Malachite Green Dye from Aqueous Solution in Hybrid System Utilizing Agricultural Waste as Particle Electrodes. Chemosphere 2021, 273, 129634. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, D.; Chen, M.; Wang, H.; Zeng, G.; Huang, D.; Cheng, M.; Liu, Y.; Xue, W.; Wang, Z. The Application of Different Typological and Structural MOFs-Based Materials for the Dyes Adsorption. Coord. Chem. Rev. 2019, 380, 471–483. [Google Scholar] [CrossRef]
  6. Abhinaya, M.; Parthiban, R.; Kumar, P.S.; Vo, D.-V.N. A Review on Cleaner Strategies for Extraction of Chitosan and Its Application in Toxic Pollutant Removal. Environ. Res. 2021, 196, 110996. [Google Scholar] [CrossRef] [PubMed]
  7. Sharma, G.; AlGarni, T.S.; Kumar, P.S.; Bhogal, S.; Kumar, A.; Sharma, S.; Naushad, M.; ALOthman, Z.A.; Stadler, F.J. Utilization of Ag2O–Al2O3–ZrO2 Decorated onto RGO as Adsorbent for the Removal of Congo Red from Aqueous Solution. Environ. Res. 2021, 197, 111179. [Google Scholar] [CrossRef]
  8. Ullah, S.; Al-Sehemi, A.G.; Mubashir, M.; Mukhtar, A.; Saqib, S.; Bustam, M.A.; Cheng, C.K.; Ibrahim, M.; Show, P.L. Adsorption Behavior of Mercury over Hydrated Lime: Experimental Investigation and Adsorption Process Characteristic Study. Chemosphere 2021, 271, 129504. [Google Scholar] [CrossRef]
  9. Yang, R.; Li, D.; Li, A.; Yang, H. Adsorption Properties and Mechanisms of Palygorskite for Removal of Various Ionic Dyes from Water. Appl. Clay Sci. 2018, 151, 20–28. [Google Scholar] [CrossRef]
  10. Christoforidis, K.C.; Montini, T.; Bontempi, E.; Zafeiratos, S.; Jaén, J.J.D.; Fornasiero, P. Synthesis and Photocatalytic Application of Visible-Light Active β-Fe2O3/g-C3N4 Hybrid Nanocomposites. Appl. Catal. B Environ. 2016, 187, 171–180. [Google Scholar] [CrossRef]
  11. Dias, E.M.; Petit, C. Towards the Use of Metal–Organic Frameworks for Water Reuse: A Review of the Recent Advances in the Field of Organic Pollutants Removal and Degradation and the next Steps in the Field. J. Mater. Chem. A 2015, 3, 22484–22506. [Google Scholar] [CrossRef]
  12. Akpan, U.G.; Hameed, B.H. Parameters Affecting the Photocatalytic Degradation of Dyes Using TiO2-Based Photocatalysts: A Review. J. Hazard. Mater. 2009, 170, 520–529. [Google Scholar] [CrossRef]
  13. Salleh, M.A.M.; Mahmoud, D.K.; Karim, W.A.W.A.; Idris, A. Cationic and Anionic Dye Adsorption by Agricultural Solid Wastes: A Comprehensive Review. Desalination 2011, 280, 1–13. [Google Scholar] [CrossRef]
  14. Cheung, W.H.; Szeto, Y.S.; McKay, G. Enhancing the Adsorption Capacities of Acid Dyes by Chitosan Nano Particles. Bioresour. Technol. 2009, 100, 1143–1148. [Google Scholar] [CrossRef] [PubMed]
  15. Belaid, K.D.; Kacha, S.; Kameche, M.; Derriche, Z. Adsorption Kinetics of Some Textile Dyes onto Granular Activated Carbon. J. Environ. Chem. Eng. 2013, 1, 496–503. [Google Scholar] [CrossRef]
  16. Crake, A.; Christoforidis, K.C.; Gregg, A.; Moss, B.; Kafizas, A.; Petit, C. The Effect of Materials Architecture in TiO2 /MOF Composites on CO2 Photoreduction and Charge Transfer. Small 2019, 15, 1805473. [Google Scholar] [CrossRef]
  17. Crake, A.; Christoforidis, K.C.; Kafizas, A.; Zafeiratos, S.; Petit, C. CO2 Capture and Photocatalytic Reduction Using Bifunctional TiO2/MOF Nanocomposites under UV–Vis Irradiation. Appl. Catal. B Environ. 2017, 210, 131–140. [Google Scholar] [CrossRef]
  18. Au, V.K.-M. Recent Advances in the Use of Metal-Organic Frameworks for Dye Adsorption. Front. Chem. 2020, 8, 708. [Google Scholar] [CrossRef]
  19. Ramírez, D.J.; Alfonso Herrera, L.A.; Colorado-Peralta, R.; Rodríguez, R.P.; Camarillo Reyes, P.K.; Chiñas, L.E.; Sánchez, M.; Rivera, J.M. Highly Efficient Methyl Orange Adsorption by UV-012, a New Crystalline Coii MOF. CrystEngComm 2021, 23, 3537–3548. [Google Scholar] [CrossRef]
  20. Cheng, J.; Liu, K.; Li, X.; Huang, L.; Liang, J.; Zheng, G.; Shan, G. Nickel-Metal-Organic Framework Nanobelt Based Composite Membranes for Efficient Sr2+ Removal from Aqueous Solution. Environ. Sci. Ecotechnology 2020, 3, 100035. [Google Scholar] [CrossRef]
  21. Cheng, J.; Liang, J.; Dong, L.; Chai, J.; Zhao, N.; Ullah, S.; Wang, H.; Zhang, D.; Imtiaz, S.; Shan, G.; et al. Self-Assembly of 2D-Metal–Organic Framework/Graphene Oxide Membranes as Highly Efficient Adsorbents for the Removal of Cs+ from Aqueous Solutions. RSC Adv. 2018, 8, 40813–40822. [Google Scholar] [CrossRef] [PubMed]
  22. Mansouri, M.; Sadeghian, S.; Mansouri, G.; Setareshenas, N. Enhanced Photocatalytic Performance of UiO-66-NH2/TiO2 Composite for Dye Degradation. Environ. Sci. Pollut. Res. 2021, 28, 25552–25565. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, Z.; Zhuang, Y.; Dong, L.; Mu, H.; Tian, S.; Wang, L.; Huang, A. Preparation of CeO2/UiO-66-NH2 Heterojunction and Study on a Photocatalytic Degradation Mechanism. Materials 2022, 15, 2564. [Google Scholar] [CrossRef] [PubMed]
  24. Man, Z.; Meng, Y.; Lin, X.; Dai, X.; Wang, L.; Liu, D. Assembling UiO-66@TiO2 Nanocomposites for Efficient Photocatalytic Degradation of Dimethyl Sulfide. Chem. Eng. J. 2022, 431, 133952. [Google Scholar] [CrossRef]
  25. Wang, Y.; Liu, H.; Zhang, M.; Duan, W.; Liu, B. A Dual-Functional UiO-66/TiO2 Composite for Water Treatment and CO 2 Capture. RSC Adv. 2017, 7, 16232–16237. [Google Scholar] [CrossRef]
  26. Abidi, N.; Errais, E.; Duplay, J.; Berez, A.; Jrad, A.; Schäfer, G.; Ghazi, M.; Semhi, K.; Trabelsi-Ayadi, M. Treatment of Dye-Containing Effluent by Natural Clay. J. Clean. Prod. 2015, 86, 432–440. [Google Scholar] [CrossRef]
  27. Lagaly, G. Colloid Clay Science. In Developments in Clay Science; Bergaya, F., Theng, B.K.G., Lagaly, G., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2006; Volume 1, pp. 141–245. ISBN 978-0-08-044183-2. [Google Scholar]
  28. Moreira, M.A.; Ciuffi, K.J.; Rives, V.; Vicente, M.A.; Trujillano, R.; Gil, A.; Korili, S.A.; de Faria, E.H. Effect of Chemical Modification of Palygorskite and Sepiolite by 3-Aminopropyltriethoxisilane on Adsorption of Cationic and Anionic Dyes. Appl. Clay Sci. 2017, 135, 394–404. [Google Scholar] [CrossRef]
  29. Middea, A.; Spinelli, L.S.; Souza Jr, F.G.; Neumann, R.; Fernandes, T.L.A.P.; Gomes, O. da F.M. Preparation and Characterization of an Organo-Palygorskite-Fe3O4 Nanomaterial for Removal of Anionic Dyes from Wastewater. Appl. Clay Sci. 2017, 139, 45–53. [Google Scholar] [CrossRef]
  30. Chen, H.; Zhong, A.; Wu, J.; Zhao, J.; Yan, H. Adsorption Behaviors and Mechanisms of Methyl Orange on Heat-Treated Palygorskite Clays. Ind. Eng. Chem. Res. 2012, 51, 14026–14036. [Google Scholar] [CrossRef]
  31. Zhao, G.; Shi, L.; Feng, X.; Yu, W.; Zhang, D.; Fu, J. Palygorskite-Cerium Oxide Filled Rubber Nanocomposites. Appl. Clay Sci. 2012, 67–68, 44–49. [Google Scholar] [CrossRef]
  32. Christoforidis, K.C.; Montini, T.; Fittipaldi, M.; Jaén, J.J.D.; Fornasiero, P. Photocatalytic Hydrogen Production by Boron Modified TiO2/Carbon Nitride Heterojunctions. ChemCatChem 2019, 11, 6408–6416. [Google Scholar] [CrossRef]
  33. Christoforidis, K.C.; Fornasiero, P. Photocatalytic Hydrogen Production: A Rift into the Future Energy Supply. ChemCatChem 2017, 9, 1523–1544. [Google Scholar] [CrossRef]
  34. Gao, Y.; Zheng, Y.; Chai, J.; Tian, J.; Jing, T.; Zhang, D.; Cheng, J.; Peng, H.; Liu, B.; Zheng, G. Highly Effective Photocatalytic Performance of {001}-TiO2/MoS2/RGO Hybrid Heterostructures for the Reduction of Rh B. RSC Adv. 2019, 9, 15033–15041. [Google Scholar] [CrossRef] [PubMed]
  35. Reza, K.M.; Kurny, A.; Gulshan, F. Parameters Affecting the Photocatalytic Degradation of Dyes Using TiO2: A Review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef]
  36. Ling, L.; Wang, Y.; Zhang, W.; Ge, Z.; Duan, W.; Liu, B. Preparation of a Novel Ternary Composite of TiO2/UiO-66-NH2/Graphene Oxide with Enhanced Photocatalytic Activities. Catal. Lett. 2018, 148, 1978–1984. [Google Scholar] [CrossRef]
  37. Papoulis, D.; Panagiotaras, D.; Tsigrou, P.; Christoforidis, K.C.; Petit, C.; Apostolopoulou, A.; Stathatos, E.; Komarneni, S.; Koukouvelas, I. Halloysite and Sepiolite–TiO2 nanocomposites: Synthesis Characterization and Photocatalytic Activity in Three Aquatic Wastes. Mater. Sci. Semicond. Processing 2018, 85, 1–8. [Google Scholar] [CrossRef]
  38. Papoulis, D.; Komarneni, S.; Panagiotaras, D.; Nikolopoulou, A.; Christoforidis, K.C.; Fernández-Garcia, M.; Li, H.; Shu, Y.; Sato, T. Palygorskite-TiO2 Nanocomposites: Part 2. Photocatalytic Activities in Decomposing Air and Organic Pollutants. Appl. Clay Sci. 2013, 83–84, 198–202. [Google Scholar] [CrossRef]
  39. Papoulis, D.; Komarneni, S.; Panagiotaras, D.; Stathatos, E.; Christoforidis, K.C.; Fernández-García, M.; Li, H.; Shu, Y.; Sato, T.; Katsuki, H. Three-Phase Nanocomposites of Two Nanoclays and TiO2: Synthesis, Characterization and Photacatalytic Activities. Appl. Catal. B Environ. 2014, 147, 526–533. [Google Scholar] [CrossRef]
  40. Gao, L.; Zhang, Q.; Li, J.; Feng, R.; Xu, H.; Xue, C. Adsorption of Methyl Orange on Magnetically Separable Mesoporous Titania Nanocomposite. Chin. J. Chem. Eng. 2014, 22, 1168–1173. [Google Scholar] [CrossRef]
  41. Jafari, S.; Sillanpää, M. Adsorption of Dyes onto Modified Titanium Dioxide. In Advanced Water Treatment; Elsevier: Amsterdam, The Netherlands, 2020; pp. 85–160. [Google Scholar]
  42. Wu, W.; Yao, T.; Xiang, Y.; Zou, H.; Zhou, Y. Efficient Removal of Methyl Orange by a Flower-like TiO2 /MIL-101(Cr) Composite Nanomaterial. Dalton Trans. 2020, 49, 5722–5729. [Google Scholar] [CrossRef]
  43. Zhao, Y.; Zhang, Q.; Li, Y.; Zhang, R.; Lu, G. Large-Scale Synthesis of Monodisperse UiO-66 Crystals with Tunable Sizes and Missing Linker Defects via Acid/Base Co-Modulation. ACS Appl. Mater. Interfaces 2017, 9, 15079–15085. [Google Scholar] [CrossRef] [PubMed]
  44. Christoforidis, K.C.; Figueroa, S.J.A.; Fernández-García, M. Iron-Sulfur Codoped TiO2 Anatase Nano-Materials: UV and Sunlight Activity for Toluene Degradation. Appl. Catal. B Environ. 2012, 117–118, 310–316. [Google Scholar] [CrossRef]
  45. Zhang, J.; Guo, Z.; Yang, Z.; Wang, J.; Xie, J.; Fu, M.; Hu, Y. TiO 2 @UiO-66 Composites with Efficient Adsorption and Photocatalytic Oxidation of VOCs: Investigation of Synergistic Effects and Reaction Mechanism. ChemCatChem 2021, 13, 581–591. [Google Scholar] [CrossRef]
  46. Han, Y.; Liu, M.; Li, K.; Zuo, Y.; Wei, Y.; Xu, S.; Zhang, G.; Song, C.; Zhang, Z.; Guo, X. Facile Synthesis of Morphology and Size-Controlled Zirconium Metal–Organic Framework UiO-66: The Role of Hydrofluoric Acid in Crystallization. CrystEngComm 2015, 17, 6434–6440. [Google Scholar] [CrossRef]
  47. Tang, J.; Dong, W.; Wang, G.; Yao, Y.; Cai, L.; Liu, Y.; Zhao, X.; Xu, J.; Tan, L. Efficient Molybdenumvi Modified Zr-MOF Catalysts for Epoxidation of Olefins. RSC Adv. 2014, 4, 42977–42982. [Google Scholar] [CrossRef]
  48. Zhang, Q.; Yang, T.; Liu, X.; Yue, C.; Ao, L.; Deng, T.; Zhang, Y. Heteropoly Acid-Encapsulated Metal–Organic Framework as a Stable and Highly Efficient Nanocatalyst for Esterification Reaction. RSC Adv. 2019, 9, 16357–16365. [Google Scholar] [CrossRef]
  49. Christoforidis, K.C.; Iglesias-Juez, A.; Figueroa, S.J.A.; Newton, M.A.; Michiel, M.D.; Fernández-García, M. A Structural and Surface Approach to Size and Shape Control of Sulfur-Modified Undoped and Fe-Doped TiO2 Anatase Nano-Materials. Phys. Chem. Chem. Phys. 2012, 14, 5628–5634. [Google Scholar] [CrossRef]
  50. Zhu, J.; Zhang, P.; Wang, Y.; Wen, K.; Su, X.; Zhu, R.; He, H.; Xi, Y. Effect of Acid Activation of Palygorskite on Their Toluene Adsorption Behaviors. Appl. Clay Sci. 2018, 159, 60–67. [Google Scholar] [CrossRef]
  51. Giustetto, R.; Chiari, G. Crystal Structure Refinement of Palygorskite from Neutron Powder Diffraction. Eur. J. Mineral. 2004, 16, 521–532. [Google Scholar] [CrossRef]
  52. Giustetto, R.; Seenivasan, K.; Pellerej, D.; Ricchiardi, G.; Bordiga, S. Spectroscopic Characterization and Photo/Thermal Resistance of a Hybrid Palygorskite/Methyl Red Mayan Pigment. Microporous Mesoporous Mater. 2012, 155, 167–176. [Google Scholar] [CrossRef]
  53. Christoforidis, K.C.; Melchionna, M.; Montini, T.; Papoulis, D.; Stathatos, E.; Zafeiratos, S.; Kordouli, E.; Fornasiero, P. Solar and Visible Light Photocatalytic Enhancement of Halloysite Nanotubes/g-C3N4 Heteroarchitectures. RSC Adv. 2016, 6, 86617–86626. [Google Scholar] [CrossRef]
  54. Shahriyari Far, H.; Hasanzadeh, M.; Najafi, M.; Rahimi, R. In-Situ Self-Assembly of Mono- and Bi-Metal Organic Frameworks onto Clay Mineral for Highly Efficient Adsorption of Pollutants from Wastewater. Chem. Phys. Lett. 2022, 799, 139626. [Google Scholar] [CrossRef]
  55. Hidayat, A.R.P.; Sulistiono, D.O.; Murwani, I.K.; Endrawati, B.F.; Fansuri, H.; Zulfa, L.L.; Ediati, R. Linear and Nonlinear Isotherm, Kinetic and Thermodynamic Behavior of Methyl Orange Adsorption Using Modulated Al2O3@UiO-66 via Acetic Acid. J. Environ. Chem. Eng. 2021, 9, 106675. [Google Scholar] [CrossRef]
  56. Vasiliadou, I.A.; Ioannidou, T.; Anagnostopoulou, M.; Polyzotou, A.; Papoulis, D.; Christoforidis, K.C. Adsorption of Methyl Orange on a Novel Palygorskite/UiO-66 Nanocomposite. Appl. Sci. 2022, 12, 7468. [Google Scholar] [CrossRef]
  57. Yang, J.-M. A Facile Approach to Fabricate an Immobilized-Phosphate Zirconium-Based Metal-Organic Framework Composite (UiO-66-P) and Its Activity in the Adsorption and Separation of Organic Dyes. J. Colloid Interface Sci. 2017, 505, 178–185. [Google Scholar] [CrossRef]
  58. Chen, Q.; He, Q.; Lv, M.; Xu, Y.; Yang, H.; Liu, X.; Wei, F. Selective Adsorption of Cationic Dyes by UiO-66-NH2. Appl. Surf. Sci. 2015, 327, 77–85. [Google Scholar] [CrossRef]
  59. Ho, Y.S.; McKay, G. Pseudo-Second Order Model for Sorption Processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  60. Yao, P.; Liu, H.; Wang, D.; Chen, J.; Li, G.; An, T. Enhanced Visible-Light Photocatalytic Activity to Volatile Organic Compounds Degradation and Deactivation Resistance Mechanism of Titania Confined inside a Metal-Organic Framework. J. Colloid Interface Sci. 2018, 522, 174–182. [Google Scholar] [CrossRef]
  61. Li, F.; Sun, S.; Jiang, Y.; Xia, M.; Sun, M.; Xue, B. Photodegradation of an Azo Dye Using Immobilized Nanoparticles of TiO2 Supported by Natural Porous Mineral. J. Hazard. Mater. 2008, 152, 1037–1044. [Google Scholar] [CrossRef]
Scheme 1. Structural formula of methyl orange (MO).
Scheme 1. Structural formula of methyl orange (MO).
Applsci 12 08223 sch001
Figure 1. XRD patterns of the pure and composite materials.
Figure 1. XRD patterns of the pure and composite materials.
Applsci 12 08223 g001
Figure 2. ATR spectra of all prepared materials (a); zoomed area of the COO−group vibration mode (b).
Figure 2. ATR spectra of all prepared materials (a); zoomed area of the COO−group vibration mode (b).
Applsci 12 08223 g002
Figure 3. TGA profiles of the pure Pal and TiO2 samples and the prepared U, PU, and PUT materials.
Figure 3. TGA profiles of the pure Pal and TiO2 samples and the prepared U, PU, and PUT materials.
Applsci 12 08223 g003
Figure 4. UV/Vis absorption spectra (a) and plots used to estimate the bandgap energies (b).
Figure 4. UV/Vis absorption spectra (a) and plots used to estimate the bandgap energies (b).
Applsci 12 08223 g004
Figure 5. Time dependence of MO adsorption on the pure Pal, TiO2, and UiO-66 and the prepared composites (PU and PUT) (a). Pseudo-second-order rate law for MO adsorption on materials containing UiO-66 (b). Experimental conditions: 90 mg·L−1 initial MO concentration, 0.4 g·L−1 of the adsorbent, pH = 5, volume = 50 mL, and T = 25 °C.
Figure 5. Time dependence of MO adsorption on the pure Pal, TiO2, and UiO-66 and the prepared composites (PU and PUT) (a). Pseudo-second-order rate law for MO adsorption on materials containing UiO-66 (b). Experimental conditions: 90 mg·L−1 initial MO concentration, 0.4 g·L−1 of the adsorbent, pH = 5, volume = 50 mL, and T = 25 °C.
Applsci 12 08223 g005
Figure 6. Photocatalytic removal of MO following the adsorption process (a) and the corresponding first-order kinetic plots (b). Experimental conditions: 90 mg·L−1 initial MO concentration, 0.4 g·L−1 of the material, pH = 5, volume = 50 mL, T = 25 °C, and adsorption period = 1.5 h.
Figure 6. Photocatalytic removal of MO following the adsorption process (a) and the corresponding first-order kinetic plots (b). Experimental conditions: 90 mg·L−1 initial MO concentration, 0.4 g·L−1 of the material, pH = 5, volume = 50 mL, T = 25 °C, and adsorption period = 1.5 h.
Applsci 12 08223 g006
Figure 7. Room temperature DMPO spin-trapping EPR spectra of the PUT and PU composites in aqueous suspensions under 15 min of light irradiation.
Figure 7. Room temperature DMPO spin-trapping EPR spectra of the PUT and PU composites in aqueous suspensions under 15 min of light irradiation.
Applsci 12 08223 g007
Table 1. Experimentally obtained and calculated values for qe, removal efficiency (RE %), and correlation coefficient (R2) for the pseudo-second-order kinetic model used to simulate the adsorption process and photocatalytic rate constant (k).
Table 1. Experimentally obtained and calculated values for qe, removal efficiency (RE %), and correlation coefficient (R2) for the pseudo-second-order kinetic model used to simulate the adsorption process and photocatalytic rate constant (k).
Sampleqe,exp
(mg·g−1)
RE (%) 1qe,cal
(mg·g−1)
k2
(g·mg−1·min−1) 2
R2k
(min−1) 3
U195.787.1196.10.0120.999
PU213.394.9212.80.0740.999
PUT174.777.5175.40.1080.9990.0153
TiO241.318.3 0.0043
Pal20.99.3
1 Removal efficiency; 2 k2: rate constant of the second-order adsorption kinetics; 3 k: first-order rate constant of the photocatalytic removal of MO.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ioannidou, T.; Anagnostopoulou, M.; Papoulis, D.; Christoforidis, K.C.; Vasiliadou, I.A. UiO-66/Palygorskite/TiO2 Ternary Composites as Adsorbents and Photocatalysts for Methyl Orange Removal. Appl. Sci. 2022, 12, 8223. https://doi.org/10.3390/app12168223

AMA Style

Ioannidou T, Anagnostopoulou M, Papoulis D, Christoforidis KC, Vasiliadou IA. UiO-66/Palygorskite/TiO2 Ternary Composites as Adsorbents and Photocatalysts for Methyl Orange Removal. Applied Sciences. 2022; 12(16):8223. https://doi.org/10.3390/app12168223

Chicago/Turabian Style

Ioannidou, Thaleia, Maria Anagnostopoulou, Dimitrios Papoulis, Konstantinos C. Christoforidis, and Ioanna A. Vasiliadou. 2022. "UiO-66/Palygorskite/TiO2 Ternary Composites as Adsorbents and Photocatalysts for Methyl Orange Removal" Applied Sciences 12, no. 16: 8223. https://doi.org/10.3390/app12168223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop