Nothing Special   »   [go: up one dir, main page]

Academia.eduAcademia.edu

Effective Carbon/TiO2 Gel for Enhanced Adsorption and Demonstrable Visible Light Driven Photocatalytic Performance

Gels

A new strategy to synthesise carbon/TiO2 gel by a sol–gel method is proposed. Textural, morphological, and chemical properties were characterised in detail and the synthesised material was proven to be an active adsorbent, as well as a visible light photocatalyst. Homogenously distributed TiO2 is mesoporous with high surface area and, hence, exhibited a high adsorption capacity. The adsorption equilibrium experimental data were well explained by the Sips isotherm model. Kinetic experiments demonstrated that experimental data fitted a pseudo second order model. The modification in electronic structure of TiO2 resulted in a reduced bandgap compared to commercial P25. The absorption edge studied through UV-Vis shifted to the visible region, hence, daylight photocatalytic activity was efficient against degradation of MB dye, as an example pollutant molecule. The material was easily removed post treatment, demonstrating potential for employment in industrial water treatment processes.

gels Article Effective Carbon/TiO2 Gel for Enhanced Adsorption and Demonstrable Visible Light Driven Photocatalytic Performance Anam Safri and Ashleigh Jane Fletcher * Department of Chemical and Process Engineering, University of Strathclyde, Glasgow G1 1XJ, UK; anam.safri@strath.ac.uk * Correspondence: ashleigh.fletcher@strath.ac.uk; Tel.: +44-141-5482-431 Abstract: A new strategy to synthesise carbon/TiO2 gel by a sol–gel method is proposed. Textural, morphological, and chemical properties were characterised in detail and the synthesised material was proven to be an active adsorbent, as well as a visible light photocatalyst. Homogenously distributed TiO2 is mesoporous with high surface area and, hence, exhibited a high adsorption capacity. The adsorption equilibrium experimental data were well explained by the Sips isotherm model. Kinetic experiments demonstrated that experimental data fitted a pseudo second order model. The modification in electronic structure of TiO2 resulted in a reduced bandgap compared to commercial P25. The absorption edge studied through UV-Vis shifted to the visible region, hence, daylight photocatalytic activity was efficient against degradation of MB dye, as an example pollutant molecule. The material was easily removed post treatment, demonstrating potential for employment in industrial water treatment processes.   Keywords: adsorption; carbon/TiO2 gels; resorcinol formaldehyde RF/TiO2 gels; photocatalysis; adsorption kinetics; methylene blue dye degradation Citation: Safri, A.; Fletcher, A.J. Effective Carbon/TiO2 Gel for Enhanced Adsorption and Demonstrable Visible Light Driven Photocatalytic Performance. Gels 2022, 8, 215. https://doi.org/ 10.3390/gels8040215 Academic Editors: Hiroyuki Takeno and Avinash J. Patil Received: 11 February 2022 Accepted: 29 March 2022 Published: 1 April 2022 Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. Copyright: © 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https:// creativecommons.org/licenses/by/ 4.0/). 1. Introduction Adsorption of carbon is perhaps the most widely used water treatment technique. However, there is an ongoing effort to develop efficient adsorbents with reduced regeneration costs. Currently, the combination of carbon and titanium dioxide (TiO2 ) appears to offer a promising route to obtain an adsorbent with self-regeneration properties. Additionally, the synergistic effect of both carbon and TiO2 enhances the degradation process due to respective adsorptive and photocatalytic properties. Literature reports several studies to address the synergy of adsorption and photodegradation by experimental demonstration of various carbon/TiO2 composite materials [1,2]. However, there is still a need to better understand the phenomenon of pollutant-adsorbent interactions in order to have a good knowledge to design an efficient water treatment process. Additionally, the improvement in design involves the type of materials and synthesis process employed to attain maximum efficiency of the system. Previously, carbon has been combined with TiO2 through various approaches, in different forms, such as carbon nanotubes [3–5], graphene [6–8], and activated carbon [9,10]. Lately, focus has been shifted to highly porous carbon materials as support matrix for industrial applications, due to the high surface area and tuneable porosity. Ideally, welldeveloped mesoporous structures with large pore volumes and uniform pore size distributions are preferred, due to enhanced accessible surface sites contributing to superior adsorption capacity of pollutants from the aqueous phase. However, the preparation process of these mesoporous carbons is costly and complicated, usually resulting in materials with moderate or low surface area. The efficiency of the material is also limited, since most TiO2 nanoparticles incorporated in the pores of the carbon are unavailable for photocatalysis [11]. Gels 2022, 8, 215. https://doi.org/10.3390/gels8040215 https://www.mdpi.com/journal/gels Gels 2022, 8, 215 2 of 19 Amongst mesoporous carbon materials, carbon gels are a new type of nanocarbon with potential applications in photocatalysis [2,12,13]. Carbon gels produced by polycondensation of resorcinol (R) with formaldehyde (F) are highly porous and have flexible properties. A comprehensive review of sol–gel synthesis of RF gel reveals that the material can be easily tailored to attain desired properties, mainly tuneable porosity, and acts as a support for metals [14]. Hence, RF gels can be promising materials for water treatment applications, mainly due to their stability, owing to aromatic resorcinol rings and their overall interconnected mesoporous carbon structure. For industrial applications where a continuous process system is often required, carbon derived from RF gels can be more efficient and cost-effective than commercial adsorbents, which are in the form of granules or powders and are unsuitable for use in continuous systems. The aim of this study is to synthesize an adsorbent with visible light driven photocatalytic activity by incorporating TiO2 nanoparticles into RF gels. A typical synthesis route of an RF gel [15] is modulated in this study to integrate TiO2 nanoparticles by formulating a twostep synthesis scheme. In addition to enhancement in textural properties of this newly synthesised adsorbent, improvement in photocatalytic properties is expected by (i) modification in electronic structure of TiO2 , due to the presence of RF gel as a carbon source, shifting the absorption edge to the visible light region, hence, enabling TiO2 to activate under visible light irradiation; (ii) the carbon phase can entrap the photogenerated electron and hole pairs, which would otherwise recombine and dissipate heat energy; and (iii) the porous RF gel helps facilitate dispersion of TiO2 and easy post treatment removal of the adsorbent/photocatalyst. Here, we report a study of the textural and optical characteristics of the adsorbent/photocatalyst. Detailed adsorption experiments were carried out to study the effect of several parameters on adsorption capacity. Additionally, the interaction behaviour between potential pollutants and the material were investigated, using methylene blue (MB) as a model adsorptive. Equilibrium sorption data were modelled using Langmuir, Freundlich, Sips, and Toth isotherm models. Kinetic analyses were carried out by comparing the experimental data with pseudo first order and pseudo second order expressions, as well as a diffusion model to better understand the transfer behaviour of the adsorbate species. Further, photocatalytic application tests were performed under visible light irradiation and the data were modelled to study the kinetics of photocatalysis. 2. Results and Discussion 2.1. Morphology The morphology of sample, studied using FESEM, is shown in Figure 1. Figure 1a shows a heterogenous nature of synthesised RF/TiO2 with homogenously distributed TiO2 , as represented in Figure 1b. The overall structure shows the nanospheres connected to form a three-dimensional porous network, as represented in Figure 1c [14]. The heterogenous surface is more evident in Figure 1d where organic and inorganic phases can be differentiated. The diameter of microspheres ranged around 0.76–1.66 µm, indicating that the size of the primary particles was slightly larger than pristine RF, which generally is in the nanometre range [16]. Energy dispersive X-ray (EDX) spectra of the microspheres is shown in Figure 1e, (EDX zone shown in supplementary information, Figure S1) which evidently corresponds to the recorded spectra. 2.2. FTIR Analysis The IR absorption bands of RF/TiO2 overall resembled those of the pristine RF gel, as also observed through FESEM images with clear uniform spheres illustrating a porous network and the retention of the gel structure even after addition of TiO2 . Typical characteristic peaks, such as the previously reported C=C stretching, CH2 , and C-O-C of aromatic rings, methylene bridges, and methylene ether bridges [17,18], were observed. The broad peak at 3300 cm−1 is characteristic of stretching vibrations associated with phenolic OH groups. Weak vibrations in the range of 2000–1700 cm−1 are attributed to CH bending of Gels 2022, 8, 215 3 of 19 aromatic compounds. The absorption bands at 1605 and 1473 cm−1 correspond to aromatic ether bridges, attributed to condensation of resorcinol to form the RF gel network. A strong IR peak, expected in the range 1740–1700 cm−1 , associated with C=O stretching of aldehyde, was not observed, which confirms that the sol–gel reaction was complete. In comparison to a spectrum of pristine RF, a few additional peaks were observed that verify the chemical linkages between RF and TiO2 , as marked in Figure 2. It has been established that the oxygenated surface groups of carbon materials support the attachment of TiO2 . [19]. Here, crosslinking of TiO2 with RF, via the hydroxyl groups, can be observed through the peaks in the vicinity of 1400 cm−1 , attributed to OH groups of RF, which appeared weak in the spectrum of RF/TiO2 , signifying the reaction of OH and TiO2 . Meanwhile, new signals observed at 1200 and 1084 cm−1 suggest formation of Ti-O-C functionalities. Similar crosslinking has previously been reported in TiO2 /phenol resol hybrid structures, where chemical interactions between TiO2 and phenol resol form Ti-O-C complexes. This heterojunction is responsive to visible light due to formation of a charge complex between the interface of TiO2 and mesoporous phenol resol producing new electronic interactions [20]. Hence, it can be concluded that the interactions between RF and TiO2 are chemical in nature. Additional signals below 1000 cm−1 , such as bands at 963 and 880 cm−1 , are associated with titanium ethoxide functional groups. Additionally, the broad band observed in the range of 600 cm−1 corresponds to the vibration of Ti-O-Ti bonds [21]. 2.3. Surface Area Analysis A nitrogen sorption isotherm was measured to determine the specific surface area and pore volume of RF/TiO2 . Figure 3 shows N2 sorption isotherm and pore size distribution (inset Figure 3). As can be seen, the isotherm of RF/TiO2 is of Type IV classification [22] with a sharp capillary condensation at P/Po = 0.4–0.9 and a well-defined hysteresis loop of Type H1, associated with open ended pores whilst suggesting a mesoporous structure [16]. Pore filling occurs at low relative pressure and the calculated mesoporosity in the structure was ~94%. The SBET , corresponding pore size and total pore volume of as prepared RF/TiO2 is 439 m2 g−1 , 9.4 nm and 0.71 cm3 g−1 , respectively. The SBET value of pristine RF gel obtained in this study is 588 m2 g−1 . The reason in reduced SBET value for RF/TiO2 is attributed to blockage of pores of RF gel matrix with inclusion of TiO2 nanoparticles. Meanwhile, in comparison with pristine TiO2 , the SBET value is significantly higher for the synthesised RF/TiO2 . Additionally, noteworthy SBET value for pristine TiO2 (i.e., 111 m2 g−1 ) is obtained in this study, contrary to commercial P25 with SBET value of 57 m2 g−1 . 2.4. Effect of pH The influence of MB sorption was studied by varying the solution pH from 2–12 (25 mL, 100 mg L−1 , 0.01 g of adsorbent). The adsorption capacities at different pH values are shown in Figure 4. The efficiency of uptake increases from 47.24 to 65.96 mg g−1 when the pH increases from 2–5. Thereafter, a sharp increase in adsorption capacity is observed at pH ≥ 6. The variation in adsorption behaviour of MB on RF/TiO2 can be explained by considering the structure of MB and evaluated point of zero charge (pzc). The pHpzc value for RF/TiO2 is determined to be 7.2 (Figure S2). RF/TiO2 can be amphoteric having both positively and negatively charged surface sites in aqueous solution due to the varying amount and nature of surface oxygen [23]. At pH lower than the pHpzc, the surface of RF/TiO2 is positively charged, which repels the cationic dye (MB), and resultant interactions are hindered in acidic media due to electrostatic repulsion between the competing H+ ions on the surface of adsorbent and MB dye molecules. As the pH increases, the surface of RF/TiO2 becomes deprotonated and the adsorption sites available for interaction with cationic species increase, therefore, increased adsorption capacity is observed. This suggests that the electrostatic forces of attraction between MB and the surface of RF/TiO2 increases due to increased ion density and positive Gels 2022, 8, 215 4 of 19 Gels 2022, 8, x FORcharges PEER REVIEW on the surface. Further, the OH groups on the surface of RF/TiO2 can also attract MB dye molecules under higher pH conditions. Figureof1.RF/TiO Morphology of RF/TiO 2 sample of RF/TiO 2, (b) TiO 2 distribution Figure 1. Morphology (a) FESEM image(a)ofFESEM RF/TiOimage deter2 sample 2 , (b) TiO 2 distribution by EDX the sample, (c) distinct appearance of (d) micro/nanospheres, (d) isolated mined by EDX onmined the sample, (c) on distinct appearance of micro/nanospheres, isolated microsphere sphere with differentiation between phase,EDX andspectra. (e) corresponding EDX sp with differentiation between organic–inorganic phase,organic–inorganic and (e) corresponding 2.2. FTIR Analysis The IR absorption bands of RF/TiO2 overall resembled those of the pristine RF also observed through FESEM images with clear uniform spheres illustrating a p Gels 2022, 8, 215 junction is responsive to visible light due to formation of a charge complex between the interface of TiO2 and mesoporous phenol resol producing new electronic interactions [20]. Hence, it can be concluded that the interactions between RF and TiO2 are chemical in nature. Additional signals below 1000 cm−1, such as bands at 963 and 880 cm−1, are associated with titanium ethoxide functional groups. Additionally, the broad band observed in5the of 19 range of 600 cm−1 corresponds to the vibration of Ti-O-Ti bonds [21]. Pristine RF RF/TiO2 C=O C=C C-H Gels 2022, 8, x FOR PEER REVIEW 5 of 19 OH 2 is 439 m2 g−1, 9.4 nm and 0.71 cm3 g−1 Ti-O-Ti prepared RF/TiO , respectively. The SBET value of pristine RF gel obtained in this study is 588 m2 OH g−1. The reason in reduced SBET value for RF/TiO2 is attributed to blockage of pores of RF gel matrix with inclusion of TiO2 nanoTi-O-C particles. Meanwhile, in comparison with pristine TiO2, the SBET value is significantly higher the synthesised RF/TiO 2. Additionally, S500 BET value for pristine TiO2 4000 for3500 3000 2500 2000 1500 noteworthy 1000 -1 (cmcontrary ) (i.e., 111 m2 g−1) is obtainedWavenumber in this study, to commercial P25 with SBET value of 57 2 g−1. m Figure 2 gel compared to to pristine RFRF gel. Figure2.2.FTIR FTIRspectra spectraofofsynthesised synthesisedRF/TiO RF/TiO gel compared pristine gel. 2 600 2.3. Surface Area Analysis Adsorption Desorption 2.0 dV/dlog(w) Pore Volume (cm³g -1 ) -1 Quantity Adsorbed (cm³g ) A nitrogen sorption isotherm was measured to determine the specific surface area 500 and pore volume of RF/TiO2. Figure 3 shows N2 sorption isotherm and pore size distribution (inset Figure 3). As can be seen, the isotherm of RF/TiO2 is of Type IV classification [22] with a sharp capillary condensation at P/Po = 0.4–0.9 and a well-defined hysteresis 400 loop of Type H1, associated with open ended pores whilst suggesting a mesoporous structure [16]. Pore filling occurs at low relative pressure and the calculated mesoporosity in 300 the structure was ~94%. The SBET, corresponding pore size and total pore volume of as 1.5 1.0 0.5 0.0 10 20 30 40 50 60 70 80 90 100 Pore width (nm) 200 100 0 0.0 0.2 0.4 0.6 0.8 1.0 Relative Pressure (p/p0) Figure Figure3.3.Nitrogen Nitrogensorption sorptionisotherms isothermsand andpore poresize sizedistribution distributionof ofsynthesised synthesisedRF/TiO RF/TiO2 2gel. gel. Overall, 2.4. Effect of pHa good adsorption capacity for MB is observed at pH higher than the pHpzc due to an increased number of negative sites in the higher pH range. This is in good The influence of MB sorption was studied by varying the solution pH from 2–12 (25 agreement with the fact that, due to the presence of COO− and OH- functional groups, −1 mL, 100 mg L , 0.01 g of adsorbent). The adsorption capacities at different pH values are MB dye adsorption is favoured at pH > pHpzc [24]. The same trend has been observed in shown in Figure 4. The efficiency of uptake increases from 47.24 to 65.96 mg g−1 when the previous studies with activated carbon and TiO2 composites where reduced activity was pH increases from 2–5. Thereafter, a sharp increase in adsorption capacity is observed at observed at acidic pH and maximum activity was observed in the pH range 6–10 [25–27]. pH ≥ 6. The variation in adsorption behaviour of MB on RF/TiO2 can be explained by considering of MB and evaluated point of zero charge (pzc). The pHpzc value 2.5. Effectthe of structure Contact Time for RF/TiO 2 is determined to be 7.2 (Figure S2). Figure 5 shows the effect of contact time on the amount of MB molecules adsorbed -1 qe (mg g ) by RF/TiO2 gel under different initial MB concentrations. As shown, the adsorption 220 capacity increases with increase in initial concentration. The equilibrium adsorption ca200 pacity increases from 102 mg g−1 to 207 mg g−1 by increasing the initial concentration 180 of MB from 50 mg L−1 to 200 mg L−1 . Initially, the adsorption capacity overall is rapid 160 for timeframes up to 30 min. This trend is expected, due to the greater driving force of MB140dye molecules and immediate availability of vacant adsorption sites, hence resulting 120 100 80 60 600 Adsorption Desorption -1 Quantity Adsorbed (cm³g ) 500 400 1.5 6 of 19 1.0 in increased in frequency of collisions between MB dye molecules and the RF/TiO2 gel. 300 Additionally, mesoporosity throughout the RF/TiO2 gel structure provides a high surface area for greater adsorption of MB molecules. It is noteworthy that at higher MB concenPore width (nm) 200 the adsorption rate is greater and adsorption capacity attains equilibrium faster tration than at low concentration. The reason is attributed to immediate occupancy of available active sites by a large amount of adsorbate molecules. This rapid occurrence of sorption 100 is due to the presence of mesoporosity within the RF/TiO2 gel, which corresponds to a large portion of the adsorption sites. In this case, the mesoporous structure provides a large 0surface area to solution volume within the porous network of the adsorbent gel. 0.0 0.2 0.4 0.6 0.8 1.0 Additionally, within the mesopores, MB dye molecules are confined to be in close proximity Relative Pressure (p/p ) to the surface. Such observations have 0been reported in previous research, particularly for activated carbons [28]. Over time, saturation of active occurs, and adsorption Figure 3. Nitrogen sorption isotherms and pore size distribution of sites synthesised RF/TiO 2 gel. becomes difficult on the fewer available active sites due to repulsive forces between the MBEffect molecules 2.4. of pH and the RF/TiO2 gel surface. Additionally, the blockage of pores and charge repulsion of MB dye may decelerate theby adsorption progress. phenomena The influence of species MB sorption was studied varying the solutionSimilar pH from 2–12 (25 have been explained for porous TiO and other carbon/TiO porous composite materials, 2 2 −1 mL, 100 mg L , 0.01 g of adsorbent). The adsorption capacities at different pH values are where it may have taken longer for the adsorbate to diffuse deeper in the fine pores [29]. shown in Figure 4. The efficiency of uptake increases from 47.24 to 65.96 mg g−1 when the Thereafter, the adsorption capacity increases gradually until 90 min, and equilibrium is pH increases from 2–5. Thereafter, a sharp increase in adsorption capacity is observed at attained for the entire concentration range. Thus, equilibrium time was considered as pH ≥ 6. The variation in adsorption behaviour of MB on RF/TiO2 can be explained by con90 min which was considered sufficient for removal of MB ions by RF/TiO2 gel. Hence, sidering the structure of MB and evaluated point of zero charge (pzc). The pHpzc value the contact time was set to 90 min in the remaining experiments to ensure equilibrium for RF/TiO2 is determined to be 7.2 (Figure S2). was achieved. 0.5 0.0 10 20 30 40 50 60 70 80 90 100 220 200 180 160 140 -1 qe (mg g ) Gels 2022, 8, 215 dV/dlog(w) Pore Volume (cm³g -1 ) 2.0 120 100 80 60 40 20 0 0 2 4 6 8 10 12 pH Figure4. 4. Effect Effectof ofpH pHon onthe theadsorption adsorption of of MB MB dye dye by by RF/TiO RF/TiO2 2gel. gel. Figure 2.6. Effect of Sorbent Dose The percentage removal of MB dye increased with increase in the adsorbent dose from 0.005 to 0.01 g but remained almost constant with further increase in the dose range 0.01 to 0.1 g, as represented in Figure 6. Percentage removal was calculated using Equation (2), and showed an increase with increase in adsorbent dose, due to greater availability of vacant active sites, a large surface area, and a greater number of adsorptive sites present on the surface of RF/TiO2 . With further increase in adsorbent dose (>0.01 g), the rate of MB removal becomes low, as the concentrations at the surface and solution reach equilibrium. The resultant reduction in adsorption rate is attributed to unoccupied adsorbent sites, as well as overcrowding or aggregation of adsorbent particles [30]. Hence, the surface area available for MB adsorption per unit mass of the adsorbent reduces, whereby percentage removal was not significantly enhanced with further increase adsorbent dose. -1 qe (mg g Gels FOR PEER REVIEW Gels2022, 2022,8,8,x215 100 7 of 7 of1919 50 qt (50 mg/L) qt (100 mg/L) qt (150 mg/L) qt (200 mg/L) 0 0 50 100 200 150 200 250 time (min) Figure 5. Effect of adsorption on contact time and initial concentration of MB dye by RF/TiO2 gel. 150 -1 qe (mg g ) 2.6. Effect of Sorbent Dose The percentage removal of MB dye increased with increase in the adsorbent dose from 0.005 100to 0.01 g but remained almost constant with further increase in the dose range 0.01 to 0.1 g, as represented in Figure 6. Percentage removal was calculated using Equation (2), and showed an increase with increase in adsorbent dose, due to greater availability of vacant active 50 sites, a large surface area, and a greater number of adsorptive sites present on the qt (50 mg/L) surface of RF/TiO2. With further increase in adsorbent dose (>0.01 g), the rate of MB reqt (100 mg/L) qt (150 mg/L) moval becomes low, as the concentrations at the surface and solution reach equilibrium. qt (200 mg/L) The resultant reduction in adsorption rate is attributed to unoccupied adsorbent sites, as 0 50 100 150 200 particles 250 [30]. Hence, the surface area well as0 overcrowding or aggregation of adsorbent time (min) available for MB adsorption per unit mass of the adsorbent reduces, whereby percentage removal not significantly increase adsorbent dose. Figure 5. was Effect of adsorption on enhanced contact timewith and further initial concentration of MB dye by RF/TiO2 gel. Figure 5. Effect of adsorption on contact time and initial concentration of MB dye by RF/TiO2 gel. 300 2.6. Effect of Sorbent Dose qe % Removal 75 The percentage removal of MB dye increased with increase in the adsorbent dose from 0.005 to 0.01 g but remained almost constant with further increase in the dose range 0.01 to 0.1 g, as represented in Figure 6. Percentage removal was calculated using Equation (2), and 70 showed an increase with increase in adsorbent dose, due to greater availability of vacant 200 active sites, a large surface area, and a greater number of adsorptive sites present on the surface of RF/TiO2. With further increase in adsorbent dose65(>0.01 g), the rate of MB re150 moval becomes low, as the concentrations at the surface and solution reach equilibrium. The resultant reduction in adsorption rate is attributed to unoccupied adsorbent sites, as 60 well100 as overcrowding or aggregation of adsorbent particles [30]. Hence, the surface area available for MB adsorption per unit mass of the adsorbent reduces, whereby percentage removal was not significantly enhanced with further increase 55 adsorbent dose. 50 -1 qe (mg g ) % Removal 250 300 0 0.00 qe 0.02 250 0.04 0.06 0.08% Removal0.10 50 75 Dose (g) Figure6.6.Effect Effectof ofadsorbent adsorbentdose doseon onthe theremoval removaland andadsorption adsorptionof ofMB MBdye dyeby byRF/TiO RF/TiO2 2gel. gel. 70 Figure 200 2.7. Adsorption Kinetics % Removal -1 qe (mg g ) The adsorption kinetics were studied using a contact time 65 of 240 min in the concentration150range 50–200 mg L−1 . The experimental data obtained for MB dye adsorption capacity vs. time (t) were fitted with PFO and PSO, as presented in Figure 7a–d. The parameters 60 determined, including measured equilibrium adsorption capacity qe (experimental), theo100 retical equilibrium adsorption capacity qe (calculated), first order rate constant K1 , second order rate constant K2 , and regression coefficient R2 , are presented in Table 1. 55 50 As observed from the data, the correlation factor R2 deviates significantly from 1 for PFO and, therefore, pseudo first order model does not exhibit good compliance with the 0 50 implies that the adsorption experimental data for the entire concentration range. This 0.00 0.02 0.04 0.06 0.08 0.10 reaction is not inclined towards physisorption, and the MB dye molecules adsorb to specific Dose (g) sites on the surface of RF/TiO2 gel. The argument regarding the failure of the pseudo first order model suggests that otherand interactions responsible for the sorption Figure 6. Effect of adsorbent dose onseveral the removal adsorptionare of MB dye by RF/TiO 2 gel. mechanism. Hence, the correlation coefficients R2 of the pseudo second order model were compared with pseudo first order parameters. R2 values for pseudo second order behaviour are approximately 0.99 for the entire concentration range, indicating that the system is more appropriately described by the pseudo second order equation. The dependence on qe experimental (mg g−1) 112.75 175.98 201.46 212.56 Pseudo first order 167.45 183.10 206.99 0.11886 0.1261 0.17078 8 of 19 0.9632 0.948 0.985 Pseudo second order −1 qe, mg g 116.97 178.54 203.58 217.59 initial concentration of MB dye is verified by good compliance of the experimental data K2 (×10−3 g mg−1 min−1) 1.01 1.10 1.06 1.48 with the pseudo second order equation, where the adsorption capacity is affected by the 2 R 0.998 0.989 0.987 0.996 initial MB dye concentration, subsequent surface-active sites, and adsorption rate (Other error analyses are represented in Table S1). qe, mg g−1 K1 (min−1) R2 Gels 2022, 8, 215 107.65 0.08087 0.9758 200 120 (b) 180 160 100 140 120 -1 qe (mg g ) -1 qe (mg g ) 80 100 60 40 80 60 40 20 -1 qe (50 mg L ) PFO PSO 0 -1 20 qe (100 mg L ) PFO PSO 0 0 50 100 150 200 250 0 50 100 time (min) 150 200 250 time (min) (a) (b) 250 200 200 -1 qe (mg g ) -1 qe (mg g ) 150 100 150 100 50 50 -1 150 (mg L ) PFO PSO -1 0 0 50 100 150 200 200 (mg L ) PFO PSO 0 250 0 50 100 time (min) 150 200 250 time (min) (c) (d) −1 mg L−1, (c) 150 mg −1 L−1, (d) 200 mg − Figure7.7. MB onon RF/TiO 2 gel at (a) 50 mg L−1, (b) 100 L1−1, Figure MBuptakes uptakes RF/TiO 2 gel at (a) 50 mg L , (b) 100 mg L , (c) 150 mg L , and fitted data for pseudo firstfor order and first pseudo second order kinetic −1 , and (d) 200 mg L fitted data pseudo order and pseudo secondmodels. order kinetic models. Table 1. Kinetic parameters obtained by fitting kinetic data for MB adsorption to RF/TiO2 . Model qe experimental (mg g−1 ) 50 mg L−1 100 mg L−1 150 mg L−1 200 mg L−1 112.75 175.98 201.46 212.56 183.10 0.1261 0.948 206.99 0.17078 0.985 203.58 1.06 0.987 217.59 1.48 0.996 Pseudo first order g−1 qe , mg K1 (min−1 ) R2 107.65 0.08087 0.9758 167.45 0.11886 0.9632 Pseudo second order g−1 qe , mg K2 (×10−3 g mg−1 min−1 ) R2 116.97 1.01 0.998 178.54 1.10 0.989 The equilibrium sorption capacity increased from 116.97 to 217.59 mg g−1 when initial dye concentration was increased from 50 to 200 mg g−1 confirming that MB dye removal is dependent on initial concentration, where the rate limiting step is determined by both adsorbate (MB) and adsorbent (RF/TiO2 ) concentration. This signifies that the sorption mechanism is chemisorption. Previous studies have explained theoretically that if 200 150 -1 qe (mg g ) Gels 2022, 8, 215 initial MB dye concentration, subsequent surface-active sites, and adsorption rate. (Other error analyses are represented in Table S1) The equilibrium sorption capacity increased from 116.97 to 217.59 mg g−1 when initial dye concentration was increased from 50 to 200 mg g−1 confirming that MB dye removal is dependent on initial concentration, where the rate limiting step is determined by 9both of 19 adsorbate (MB) and adsorbent (RF/TiO2) concentration. This signifies that the sorption mechanism is chemisorption. Previous studies have explained theoretically that if diffusion is not the rate limiting factor, then higher adsorbate concentrations would give a good diffusion is not the rate limiting factor, then higher adsorbate concentrations would give a pseudo first order fit whereas, for low concentrations, pseudo second order better repregood pseudo first order fit whereas, for low concentrations, pseudo second order better sents the kinetics of sorption, analogous to the observations made here [31]. Previously, represents the kinetics of sorption, analogous to the observations made here [31]. Previously, the adsorption processes of MB on TiO2/carbon composites have also exhibited strong dethe adsorption processes of MB on TiO2 /carbon composites have also exhibited strong pendencies of pseudo second order fitting parameters on initial concentrations [27]. dependencies of pseudo second order fitting parameters on initial concentrations [27]. 0.5 Figure 8 shows a plot of MB dye uptake (qe) on synthesised RF/TiO2 against (time)0.5. Figure 8 shows a plot of MB dye uptake (qe) on synthesised RF/TiO2 against (time) . The plots exhibit multi-linearity, rather than two straight lines, indicating that the adsorpThe plots exhibit multi-linearity, rather than two straight lines, indicating that the adsorption influenced by by several several steps. steps. The tion process process is is influenced The initial initial segment segment of of the the plots plots shows shows that that diffusion comparison diffusion across across the the boundary boundary of of the the adsorbent adsorbent only only lasts lasts for for aa short short time time in in comparison to whole adsorption adsorption process. to the the whole process. This This second second section section is is attributed attributed to to diffusion diffusion into into the the mesopores of the adsorbent, i.e., the MB dye molecules enter less accessible pore mesopores of the adsorbent, i.e., the MB dye molecules enter less accessible pore sites. sites. Resultantly, the diffusion diffusion resistance resistance increases, increases, and and the the diffusion diffusion rate rate decreases. Resultantly, the decreases. This This stage stage is a slow and gradual stage of the adsorption process. The third segment represents is a slow and gradual stage of the adsorption process. The third segment represents the the final equilibrium stage where intra-particle diffusion slows down to an extremely low rate final equilibrium stage where intra-particle diffusion slows down to an extremely low rate due remaining concentration concentrationof ofthe theMB MBdye dyemolecules moleculesininthe thesolution. solution.This Thisimplies impliesa due to to the the remaining aslow slowtransport transportrate rateofofMB MBdye dyemolecules moleculesfrom fromthe the solution solution (through (through the the gel–dye solution gel–dye solution interface) sites. Here, the surface surface of of the the RF/TiO RF/TiO22 gel, interface) to to available available sites. Here, the gel, and and micropores, micropores, may may be be responsible MB dye dye molecules. molecules. responsible for for the the uptake uptake of of MB 100 50 -1 50 (mg L ) -1 100 (mg L ) -1 150 (mg L ) -1 200 (mg L ) 0 0 2 4 6 8 0.5 time 10 12 14 16 0.5 (min ) Intra-particle diffusion kinetics of MB dye adsorption on RF/TiO RF/TiO2.2 . Figure 8. Intra-particle 2.8. Adsorption Isotherms The equilibrium data were analysed using Langmuir, Freundlich, Sips, and Toth isotherm equations to obtain the best fit. The isotherm data plots, and fitting model parameters are shown in Figure 9 and Table 2, respectively. Comparison of the correlation factor R2 indicates that qe,exp fitted well to the Sips model with the lowest χ2 value. The qe,cal value, calculated using the Sips model, is closest to qe,exp with R2 closest to 1. The Sips model is a combination of the Langmuir and Freundlich adsorption isotherms, hence, the model suggests both monolayer and multilayer adsorption. At low MB dye concentrations, the model predicts Freundlich adsorption isotherms as a heterogenous adsorption system and localised adsorption without adsorbate–adsorbate interactions, whereas at high concentrations the model predicts monolayer adsorption as in Langmuir isotherm [32,33]. In the present study, the value of constant ns from Equation (11), the heterogeneity factor, is greater than 1 (i.e., ns = 1.91), hence, the adsorption system is predicted to be heterogenous [33]. Further, the Toth isotherm model validates multilayer and heterogeneous adsorption, where the factor nT determines heterogeneity. Here, again the value of nT is greater than 1, and, therefore, the system confirms heterogeneity. It is evident that the equilibrium uptakes follow the Sips model according to the correlation factor R2 (other error analyses are represented in Table S2) and the isotherm models fit the data in the order Sips > Toth > Langmuir > Freundlich. erogenous [33]. Further, the Toth isotherm model validates multilayer and heterogeneous adsorption, where the factor nT determines heterogeneity. Here, again the value of nT is greater than 1, and, therefore, the system confirms heterogeneity. It is evident that the equilibrium uptakes follow the Sips model according to the correlation factor R2 (other error analyses are represented in Table S2) and the isotherm models fit the data in the 10 of 19 order Sips > Toth > Langmuir > Freundlich. Gels 2022, 8, 215 250 200 -1 qe (mg g ) 150 100 50 qe Langmuir isotherm Freundlich isotherm Sips isotherm Toth isotherm 0 0 20 40 60 80 100 120 -1 Ce (mg L ) Figure 9. Adsorption Adsorption data data for for RF/TiO RF/TiO2 2onto ontoMB MBdye dyecorresponding correspondingfits fitsto toLangmuir, Langmuir, Freundlich, Freundlich, Figure Sips, Sips, and and Toth Toth equation. Table 2. Isotherm Isotherm parameters parameters obtained obtained by by fitting fitting MB MB adsorption adsorption data data for for RF/TiO RF/TiO22to tothe theLangmuir, Langmuir, Table 2. Freundlich, equations. Freundlich, Sips, Sips, and and Toth Toth equations. Langmuir Freundlich Sips Toth qm (mg g−1) qm (mg g−1 ) KL (L mg−1) Langmuir KL (L mg−1 ) R2 R2 KF mg g−1 (L mgK-1)1/n −1 (L mg-1 )1/n mg g F nF nF Freundlich R2 R2 qs (mg g−1) qs (mg g−1 ) KS KS Sips ns ns 2 2 R R qm (mg g−1 ) qm (mg g−1) KT KT Toth nT nT R2 254.65 254.65 0.0732 0.0732 0.960 0.960 54.85 54.85 3.19933.1993 0.865 0.865 218.71218.71 0.010 0.010 1.913 1.913 0.994 0.994 558.47558.47 0.02950.0295 1.403 1.403 0.991 2.9. Thermodynamic Study Thermodynamic parameters for the adsorption system are recorded in Table 3. Negative values of free energy changes are evident from the data, which signifies the spontaneous adsorption of MB dye molecules on the sample for the studied temperature range. Adsorption capacity increases with an increase in temperature and a positive ∆H0 (Table 3) suggests that the adsorption is endothermic in nature. Positive ∆S0 indicates some structural changes in the MB dye and RF/TiO2 gel causing an increase in the degree of freedom of the MB dye species and consequently increased randomness at the adsorbent–adsorbate interface. At high temperature, the release of high-energy desolvated water molecules from the MB dye molecules and/or aggregates arise after adsorption on RF/TiO2 gel, which relates to a positive ∆S0 [34]. Before sorption begins, the MB ions are surrounded by highly ordered water clusters strongly bound via hydrogen bonding. Once MB ions come in close contact with the surface of RF/TiO2 , the interaction results in agitation of the ordered water molecules, subsequently increasing the randomness of the system. Although, the adsorption of MB dye onto RF/TiO2 gel may reduce the freedom of the system, the entropy increase in water molecules is much higher than the entropy decrease in MB ions. Therefore, the driving force for the adsorption of MB on RF/TiO2 is controlled by an entropic effect rather than an enthalpic change. Similar phenomena have previously been reported Gels 2022, 8, 215 11 of 19 in order to explain the fact that thermodynamic parameters are not only related to the properties of the adsorbate but also to the properties of other solid particles [35,36]. Table 3. Thermodynamic data for MB adsorption onto RF/TiO2 at various temperatures. T (K) lnk ∆G0 (KJ/mol) ∆S0 (J/mol) 112 281 296 305 313 1.29 2.20 2.40 2.50 −3.01 −5.41 −6.09 −6.51 ∆H0 (KJ/mol) 28.2 3. Photocatalytic Tests Absorbance (a.u.) 0 min 30 min 60 min 90 min 500 600 700 800 Wavelength (nm) Figure10. 10.UV-Vis UV-Visspectra spectraofofMB MBdye dyedegradation degradationusing usingRF/TiO RF/TiO Figure 2 gel. 2 gel.  (h) (eVcm) 1/2 Within studied systems, no photodegradation (reduction in concentration The dye the degradation data obtained after treatmentactivity with RF/TiO 2 showed ~73% MB andremoval decolourisation of MB dye) was observed insynergy the absence adsorbent/catalyst, dye after 90 min. This is attributed to the of RFofand TiO2, enabling anas well as in the presence of pristine RF, indicating that the properties of MB are11). more stable. absorption shift to a higher wavelength, as λmax is detected at 410 nm (Figure Further Additionally, RF solely may not be recommended for photocatalysis due to slow charge analysis indicates modification in the electronic structure and a subsequent reduction in transfer occurs properties, has also been proven by the study carried out by Zang, Ni, and bandgap due which to doping of TiO 2 similar to when combined with carbon [39]. The Liu, where the researchers employed pristine RFin resins for11 visible light photocatalysis [37]. calculated band gap energy is 2.97 eV, as shown Figure (inset). The value achieved Slight photodegradation is observed in the presence of pristine TiO2 , which may be atis significantly lower than pristine TiO2 (i.e., 3.2 eV [21]), indicating photodegradation of tributed to the potential absorbance of UV-Vis light from the surroundings confirming MB dye under visible light irradiation. The RF matrix enables entrapment of a photogenthat the process of MB degradation is light driven. Although the TiO obtained for use erated electron and hole pairs and, therefore, rapid generation of ROS is 2possible for effiin this study has a high surface area, which may possess good adsorption properties to cient degradation of the MB dye. These findings are comparable to other carbon/TiO2 sysexhibit efficient adsorption of MB dye, since TiO2 only activates upon UV light irradiation tems where synergistic effects have substantially enhanced the performance of the system (~280 nm), it does not produce enough reactive oxide species (ROS) to be an effective due to improved optical properties of the material [1,40,41]. photodegradation system [38]. The dye degradation data obtained after treatment with RF/TiO2 showed ~73% MB dye removal after 90 min. This is attributed to the synergy of RF and TiO2 , enabling an max=410shift nm to a higher wavelength, as λmax is detected at 410 nm (Figure 11). Further absorption analysis indicates modification in the electronic structure and a subsequent reduction in bandgap occurs due to doping of TiO2 similar to when combined with carbon [39]. The calculated band gap energy is 2.97 eV, as shown in Figure 11 (inset). The value achieved is bsorbance (a.u.) Gels 2022, 8, x FOR PEER REVIEW Photocatalytic activity was determined by testing the efficiency of RF/TiO2 against degradation of methylene blue (MB) under visible light irradiation. The maximum ab12 of 19 sorbance vs. wavelength spectra (in the range of 550–700 nm) were collected and subsequent activity, after 30 min, intervals was recorded, as shown in Figure 10. Eg=2.97 0 1 2 3 4 5 6 7 8 9 10 (h) (eVcm) 1/2 max=410 nm  Absorbance (a.u.) Gels 2022, 8, 215 The dye degradation data obtained after treatment with RF/TiO2 showed ~73% MB dye removal after 90 min. This is attributed to the synergy of RF and TiO2, enabling an absorption shift to a higher wavelength, as λmax is detected at 410 nm (Figure 11). Further analysis indicates modification in the electronic structure and a subsequent reduction in 12 of 19 bandgap occurs due to doping of TiO2 similar to when combined with carbon [39]. The calculated band gap energy is 2.97 eV, as shown in Figure 11 (inset). The value achieved is significantly lower than pristine TiO2 (i.e., 3.2 eV [21]), indicating photodegradation of significantly TiO2 (i.e., 3.2 matrix eV [21]), indicating photodegradation of MB MB dye under lower visiblethan lightpristine irradiation. The RF enables entrapment of a photogendye under visible light irradiation. The RF matrix enables entrapment of a photogenerated erated electron and hole pairs and, therefore, rapid generation of ROS is possible for effielectron and hole and, therefore, rapid generation of ROS is possible for efficient cient degradation of pairs the MB dye. These findings are comparable to other carbon/TiO 2 sysdegradation of the MB dye. These findings are comparable to other carbon/TiO systems tems where synergistic effects have substantially enhanced the performance of the2system where synergistic effects have substantially enhanced the performance of the system due to due to improved optical properties of the material [1,40,41]. improved optical properties of the material [1,40,41]. Eg=2.97 0 1 2 3 4 5 6 7 8 9 10 h 400 450 500 550 600 650 700 750 800 Wavelength (nm) Figure Absorption wavelength spectrum RF/TiO ethanol measured through Figure 11.11. Absorption vs.vs. wavelength spectrum ofof RF/TiO 2 dispersed in in ethanol measured through 2 dispersed UV-Vis spectrophotometer, 2. . UV-Vis spectrophotometer,inset insetshows showscalculated calculatedband bandgap gapofofsynthesised synthesisedRF/TiO RF/TiO 2 The photodegradation MB dye, that the reduction concentration with time The photodegradation ofof MB dye, that is,is, the reduction inin concentration with time is is recorded Figure and the recorded data modelled using pseudo first order kinetics, recorded inin Figure 1212 and the recorded data is is modelled using pseudo first order kinetics, shownininFigure Figure12a,b. 12a,b.The Thedata data are are fitted equation (ln(ln (C(C = =kt) shown fitted to tothe thefirst firstorder orderkinetic kinetic equation o/Ce) o /Ce) to evaluate the value of the rate constant by slope of plot ln(Co /Ce) vs. time (t) in minutes, where Co and Ce is the initial at t = 0 and final concentration at given time of MB concentration, respectively. The value of k here is the measure of photocatalytic performance, as it defines the concentration of reacting substances, that is, photogenerated reactive oxide species, therefore, a higher value of k signifies higher photocatalytic efficiency. As compared to no catalyst (k = 2.43 × 10−6 min−1 ) pristine TiO2 (k = 1.74 × 10−3 min−1 ) and pristine RF (k = 6.89 × 10−4 min−1 ), the rate of RF/TiO2 was the highest (k = 1.27 × 10−2 min−1 ). Clearly, the rate constant obtained for photodegradation of MB using RF/TiO2 was the highest. Mainly, improved optical property was the most important advancement in forming RF/TiO2 gel which is photocatalytically active under visible light (410 nm) irradiation. The RF/TiO2 material created in this study exhibits excellent photoactivity under visible light, which can further be explained by the mechanism of MB photodegradation represented in Equations (1)–(8). The system activates when RF/TiO2 absorbs light with photon energy (hν) and generates conduction band (CB) electron (e− ) and valence band (VB) hole (h+ ) pairs upon under visible light irradiation. The holes interact with moisture on the surface of the adsorbent gel yielding hydroxy free radicals or reactive oxide species (H+ or OH• ), which are oxidation agents that can mineralise a wide range of organic pollutants, ultimately producing CO2 and H2 O as end products. The reaction sequence below represents the photodegradation of MB, showing a simplified mechanism of photoactivation by a photocatalyst (Equations (1)–(4)) [19]. For the mechanism of photoinactivation of MB in the presence of RF/TiO2 , hydroxy free radicals or reactive oxide species (H+ or OH• ) attack the aromatic ring of the MB structure, degrading it into a single ring structure product, which then finally degrades to CO2 and H2 O (Equations (5)–(8)) [42,43]. active oxide species, therefore, a higher value of k signifies higher photocatalytic efficiency. As compared to no catalyst (k = 2.43 × 10−6 min−1) pristine TiO2 (k = 1.74 × 10−3 min−1) and pristine RF (k = 6.89 × 10−4 min−1), the rate of RF/TiO2 was the highest (k = 1.27 × 10−2 min−1). Clearly, the rate constant obtained for photodegradation of MB using RF/TiO2 was the highest. Mainly, improved optical property was the most important advancement in 13 of 19 forming RF/TiO2 gel which is photocatalytically active under visible light (410 nm) irradiation. Gels 2022, 8, 215 1.2 (a) 1.0 (b) k=1.27x10-2 RF/TiO2 no catalyst pristine TiO2 1.0 pristine RF 0.8 ln(C0/Ce) Ce/C0 0.8 0.6 0.6 0.4 0.4 RF/TiO2 no catalyst pristine RF pristine TiO2 0.2 k=1.74x10-3 0.2 k=6.89x10-4 k=2.43x10-6 0.0 0 20 40 60 time (min) 80 100 0 20 40 60 80 100 time (min) Figure Figure 12. 12. (a) (a)Photocatalytic Photocatalytic performance performance regarding regarding MB MB dye dye degradation degradation without without catalyst, catalyst, and and with with pristine RF, pristine TiO2 and RF/TiO2 gel (b) First-order kinetics of photoactivity without catalyst, pristine RF, pristine TiO2 and RF/TiO2 gel (b) First-order kinetics of photoactivity without catalyst, and with pristine RF, pristine TiO2, and RF/TiO2 gel. and with pristine RF, pristine TiO2 , and RF/TiO2 gel. The RF/TiO2 material created in this study exhibits excellent photoactivity under vis+ − RF/TiO2 + hν = eCB + hvB (1) ible light, which can further be explained by the mechanism of MB photodegradation rep− + . resented in Equations a–h. The system RF/TiO 2 absorbs light with photon hCB + activates H2 O = Hwhen + OH (2) energy (hν) and generates conduction−band (CB) electron (e−) and valence band (VB) hole .− eCB + O2 = O2 (3) (h+) pairs upon under visible light irradiation. The holes interact with moisture on the .− + . = HO surface of the adsorbent gel yieldingOhydroxy radicals or reactive oxide species (H+ (4) or 2 2 + H free OH●), which are oxidation agents that can mineralise a wide range of organic pollutants, MB + RF/TiO2 = MB.+ + e− (RF/TiO2 ) (5) ultimately producing CO2 and H2O as end products.CB The reaction sequence below repreO2 + e−a=simplified O.2− (6) sents the photodegradation of MB, showing mechanism of photoactivation by a photocatalyst (Equations (1)–(4)) of photoinactivation of MB .+ [19]. For − the mechanism . MB + OH = MB+OH (7) in the presence of RF/TiO2, hydroxy free radicals or reactive oxide species (H+ or OH●) .+ . OHstructure, = H2 O +degrading CO2 + other products (8) attack the aromatic ringMB of the+MB it into a single ring structure product, which then finally degrades to CO2 and H2O (Equations (5)–(8)) [42,43]. 4. Conclusions (1) RF/TiO + hν = e + h An RF/TiO2 gel was successfully synthesised using sol–gel techniques via a straightforH O = H excellent + OH˙ adsorption–photodegradation (2) ward route. The synergy of RF andh TiO+2 exhibited activity due to the corresponding characteristics, mesoporosity and photocatalysis. (3) e + O =mainly O. The synergy of contributing materials allowed modification in the electronic structure of . . (4) O linkages, + H = HO TiO2 by formation of Ti-O-C chemical responsible for a reduction in the band gap of TiO2 for photodegradation upon visible (5) MB + RF/TiO = MB . light + e irradiation. (RF/TiO ) Kinetic studies revealed a pseudo second order reaction, signifying chemisorption phenomenon is involved in (6) O + eisotherm = O. study showed that the system was the adsorption mechanism. The adsorption . heterogeneous following the Sips MBmodel + OHequation. = MB +The OH˙spontaneity of the process was (7) validated via thermodynamic studies, which signified an entropically driven adsorption mechanism. Effective photodegradation results were observed due to the high adsorption capacity and improved optical properties of the material, enabling significant MB dye degradation within 90 min. Overall, the material possesses properties that have potential to effectively reduce/eliminate a wide range of pollutants and, therefore, can be employed as a low-cost photocatalytic adsorbent for water treatment Especially in the industrial applications where post treatment separation and recovery of the adsorbent is difficult, employing this material can reduce the costs since in this case the adsorbent precipitates and easy separation is possible just by filtration or even decantation. Gels 2022, 8, 215 14 of 19 5. Material and Methods 5.1. Synthesis Synthesis of RF and TiO2 precursors was carried out in two separate systems, which were integrated and processed further in order to obtain the final gel. System 1: Preparation of Titania Sol For preparation of the titania sol, 1.78 g of titanium precursor: titanium isopropoxide (TTIP) (98+%, ACROS Organics™, Geel, Belgium) was dissolved in ethanol and stirred for 30 min. A mixture of water and HCl was added dropwise to the titania/EtOH solution under constant stirring, at room temperature, to begin hydrolysis. After 2 h, a homogenous solution was obtained. The molar ratios for these parameters were 1 TTIP:10 EtOH:0.3 HCl:0.1 H2 O. System 2: Preparation of RF sol In total, 7.74 g of resorcinol (SigmaAldrich, ReagentPlus, 99%, Poole, UK) was added to 50 mL of deionised water until completely dissolved. 0.0249 g of sodium carbonate (Na2 CO3 , Sigma-Aldrich, anhydrous, ≥99.5%), as a catalyst, and 4.23 g of formaldehyde (37wt%) were added to the dissolved resorcinol under continuous stirring, at room temperature. Finally, the prepared titania sol (system 1) was gradually transferred to the RF sol (from system 2) under constant stirring, at room temperature. The resulting sol was stirred at room temperature for 2 h after which the sol mixture was aged at 85 ◦ C for 72 h. After aging, the process of solvent exchange and drying the RF/TiO2 gel, first involved cutting the gel into smaller pieces. These pieces were then immersed in acetone for 72 h to facilitate solvent exchange prior to drying, followed by vacuum drying at 110 ◦ C for 48 h to obtain the final RF/TiO2 adsorbent gel. In this way, the final gel corresponded to 10 wt% TiO2 (theoretical percentage) incorporated in the RF gel matrix. 5.2. Adsorbent Characterisation Morphology of the synthesised sample was studied by field emission electron scanning microscope (FESEM) TESCAN-MIRA. The functional groups on the surface of synthesised RF/TiO2 , and the chemical linkages between the constituent RF and TiO2 components, were verified using Fourier Transform Infrared Spectroscopy (FTIR) (MB3000 series, scan range 4000–400 nm). BET surface area measurements were carried out using a Micromeritics ASAP 2420 to obtain N2 adsorption isotherm at 77 K and pore size was determined via BJH theory [22]. A UV-Vis Spectrophotometer (Varian Cary 5000 UV-Vis NIR Spectrophotometer Hellma Analytics) was used to collect absorption spectra and the data used to interpret the change in electronic structure of RF/TiO2 [44]. The data were manipulated to calculate the band gap energy values through the Tauc method described in previous studies [44]. 6. Adsorption Experiments 6.1. Effect of pH The effect of pH on the sorption of methylene blue (MB) dye was investigated with 0.01 g of sample by adjusting the pH of solution (25 mL, 100 mg L− 1 MB) between 2 and 12, at 23 ◦ C. The pH was adjusted using 0.01 M HCl and 0.01 M NaOH. After 2 h of agitation, the solution was centrifuged for 15 min and the supernatant solution was collected via syringe. The initial and final concentrations were measured using a UV-Vis spectrophotometer (Varian Cary 5000 UV-Vis NIR Spectrophotometer, Agilent, UK) and onward calculations were performed. 6.2. Effect of Sorbent Dose The amount of sorbent dose was gradually increased from 0.005 to 0.01 g to study the effect of sorbent dose on the adsorption capacity. pH and temperature were maintained at 7.0 and 23 ◦ C, respectively, against 25 mL of 100 mg L− 1 MB concentrated solution. The pH was adjusted using 0.01 M HCl and 0.01 M NaOH. After 2 h of agitation, the solution was centrifuged for 15 min and the supernatant solution was collected via syringe. The initial and final concentrations were measured using a UV-Vis spectrophotometer (Varian Gels 2022, 8, 215 15 of 19 Cary 5000 UV-Vis NIR Spectrophotometer Hellma Analytics) and onward calculations were performed. 6.3. Effect of Initial Concentration All adsorption experiments were performed at 23 ◦ C in 125 mL conical flasks, using a shaker (VWR 3500 Analog orbital shaker) set to 125 rpm. The first experiment was conducted to study the isothermal equilibrium and the effect of initial MB concentration. Standard solutions of MB were prepared using distilled water, with initial concentrations in the range of 20–200 mg L −1 . Then, 25 mL aliquots were distributed into each flask, and 0.01 g of the adsorbent gel was added individually to each flask. The pH values of all solutions were recorded and adjusted to 7.0, if required, using 1 M HCl and 1 M NaOH. After 2 h of agitation, the solution was centrifuged for 15 min and the supernatant solution was collected via syringe. The initial and final concentrations were measured using UV-Vis spectrophotometer (Varian Cary 5000 UV-Vis NIR Spectrophotometer Hellma Analytics). The equilibrium adsorption capacity, qe (mg g−1 ), was calculated using: qe = (Co − Ce ) × V(l) W (9) while the respective percentage removal of MB was calculated as: Removal % = Co − Ce × 100% Co (10) where Co and Ce are the initial MB and final concentration, respectively. W is the weight (g) of the adsorbent and V is the volume (L) of MB solution. 6.4. Effect of Contact Time The effect of contact time was studied by adding MB solution (pH 7.0, 25 mL, 100 mg L−1 ) and 0.01 g adsorbent gel into a flask, which was then agitated for a predetermined contact time between 5 min and 4 h. The samples were prepared and treated as described in Section 2.5 and the amount of adsorption was calculated using Equation (11): qt = (Co − Ct ) × V W (11) where Ct is the equilibrium MB concentration at a given time, and Co , V, and W are as previously defined. Equilibrium concentration was determined by plotting qt versus time of aliquots collected at different time intervals. Adsorption-photodegradation (absorption) changes of MB dye with time were also recorded via UV-Vis spectrophotometry. 6.5. Effect of Temperature The effect of temperature on the removal of MB (pH 7.0, 25 mL, 20–200 mg L−1 ) was investigated by adding a known concentration MB solution and 0.01 g adsorbent gel to a flask. A hot plate with a stirrer (120 rpm) was used to maintain a constant temperature of 8, 23, 32, and 40 ◦ C, under stirring, for 120 min after which the absorbance versus wavelength spectra were recorded to measure the final concentration, and subsequent adsorption was calculated using Equation (9). 6.6. Kinetic Models The kinetic-based models: pseudo first order (PFO) and pseudo second order (PSO) were applied to study the adsorption kinetics and to explain the mode of sorption of MB onto the synthesised RF/TiO2 . The PFO model [33] has been frequently used to describe kinetic processes under non-equilibrium conditions. PFO is based on the assumption that the rate of adsorption is proportional to the driving force, that is, the difference between Gels 2022, 8, 215 16 of 19 the equilibrium concentration and solid phase concentration, presented as a differential Equation (12): dqt = k1 (qe − qt ) (12) dt Integrating Equation (13) with the initial condition of qt = 0 at t = 0, the PFO model can be rewritten, in a linear form, as:   qt= qe 1 − e−k1 t (13) Several studies have also reported the use of PSO [45] to interpret data obtained for the sorption of contaminants from water, including dyes, organic molecules, and metal ions. The PSO model assumes that the overall adsorption rate is proportional to the square of the driving force and can be expressed as Equation (14): dqt = k1 (qe − qt )2 dt (14) Integrating Equation (14) with the initial condition of qt = 0 at t = 0, and qt = t at t = t, the PSO model can be rewritten as: qt = k2 tq2e 1 + k2 tqe (15) In Equations (13)–(15), qt (mg g−1 ) and qe (mg g−1 ) are the adsorption capacities of MB dye molecules at time t and at equilibrium, respectively. k1 (mg g−1 min−1 ) and k2 (mg g−1 min−1 ) are the PFO and PSO rate constants, respectively. 6.7. Sorption Isotherm Models The equilibrium data for the sorption of MB on RF/TiO2 adsorbent gel as a function of equilibrium concentration (Ce mg L−1 ) was analysed in terms of Langmuir, Freundlich, Sips, and Toth isotherm models [2]. The nonlinear form of Langmuir’s isotherm model is represented as: q KL Ce qe = m (16) 1 + Ce KL where qe (mg g−1 ) is the MB uptake at equilibrium, Ce (mg L−1 ) is the equilibrium concentration, qm (mg g−1 ) is the amount of adsorbate at complete monolayer coverage, and KL is the Langmuir constant. The Freundlich equation can be expressed as follows: qe = KF C1/n e (17) where qe and Ce are as defined in the Langmuir equation, adsorption affinity is related to the adsorption constant KF , and n indicates the magnitude of the adsorption driving force and the distribution of energy sites on the adsorbent surface, if n < 1, then adsorption is a chemical process, whereas if n > 1, then adsorption maybe dependent on distribution of the surface sites. Generally, n values within 1–10 represents good adsorption [46]. The Sips isotherm model is a combination of the Langmuir and Freundlich isotherms and is represented as: q Ks Cns e qe = s (18) 1 + Ks Cns e where qe and Ce are as defined for Equation (16), Ks is the Sips isotherm model constant (L g−1 ), and ns; is the Sips isotherm exponent. Gels 2022, 8, 215 17 of 19 The Toth model also describes heterogeneous systems, considering both low- and high-end concentrations. The Toth expression is as follows: qe = qm KT Ce [1 + (KT Ce ) t ]1/t (19) where qe and Ce are as defined for Equation (16), qm is the maximum adsorption capacity, t is the surface heterogeneity, and KT is the surface affinity. 6.8. Photodegradation Procedure Photocatalytic performance of as prepared RF/TiO2 was investigated through MB dye degradation, by recording the dye degradation spectra with time using UV-Vis Spectrophotometry. 0.01 g of the adsorbent dose were used against 25 mL of 100 mg L−1 dye concentration at pH ~7 and a temperature of 23 ◦ C and light intensity of 111 Wm−2 . For comparison, the measurements were also recorded in the absence of catalyst, as well as pristine RF and TiO2 . All suspensions were stirred in the dark for 60 min to establish sorption equilibrium before exposure to visible light. Supplementary Materials: The following supporting information can be downloaded at: https: //www.mdpi.com/article/10.3390/gels8040215/s1, Figure S1: Zone of Energy dispersive x-ray (EDX) spectra; Figure S2: Point of zero charge (pHpzc) on the surface of RF/TiO2 ; Table S1: Kinetic parameters obtained by fitting kinetic data for MB adsorption to RF/TiO2 .; Table S2: Isotherm parameters obtained by fitting MB adsorption data for RF/TiO2 to the Langmuir, Freundlich, SIPS and Toth equations. Author Contributions: Methodology, A.S.; formal analysis, A.S. and A.J.F.; resources, A.J.F.; writing— original draft preparation, A.S.; writing—review and editing, A.J.F.; supervision, A.J.F.; project administration, A.J.F.; funding acquisition, A.J.F. All authors have read and agreed to the published version of the manuscript. Funding: This research received no external funding. Acknowledgments: Anam Safri thanks Ashleigh Fletcher and the Chemical and Process Engineering Department at the University of Strathclyde for funding this work. The authors gratefully acknowledge the Materials Science and Engineering Department at Institute of Space Technology, Islamabad, for providing support and facilities to conduct the morphological analysis. Conflicts of Interest: The authors declare no conflict of interest. References 1. 2. 3. 4. 5. 6. 7. 8. Xue, G.; Liu, H.; Chen, Q.; Hills, C.; Tyrer, M.; Innocent, F. Synergy between surface adsorption and photocatalysis during degradation of humic acid on TiO2 /activated carbon composites. J. Hazard. Mater. 2011, 186, 765–772. [CrossRef] [PubMed] Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. Adsorption and photocatalytic removal of Rhodamine B from wastewater using carbon-based materials. FlatChem 2021, 29, 100277. [CrossRef] Quyen, N.D.V.; Khieu, D.Q.; Tuyen, T.N.; Tin, D.X.; Diem, B.T.H.; Dung, H.T.T. Highly effective photocatalyst of TiO2 nanoparticles dispersed on carbon nanotubes for methylene blue degradation in aqueous solution. Vietnam. J. Chem. 2021, 59, 167–178. Sampaio, M.J.; Silva, C.G.; Marques, R.R.; Silva, A.M.; Faria, J.L. Carbon nanotube—TiO2 thin films for photocatalytic applications. Catal. Today 2011, 161, 91–96. [CrossRef] Murgolo, S.; Petronella, F.; Ciannarella, R.; Comparelli, R.; Agostiano, A.; Curri, M.L.; Mascolo, G. UV and solar-based photocatalytic degradation of organic pollutants by nano-sized TiO2 grown on carbon nanotubes. Catal. Today 2015, 240, 114–124. [CrossRef] Morawski, A.W.; Kusiak-Nejman, E.; Wanag, A.; Narkiewicz, U.; Edelmannová, M.; Reli, M.; Kočí, K. Influence of the calcination of TiO2 -reduced graphite hybrid for the photocatalytic reduction of carbon dioxide. Catal. Today 2021, 380, 32–40. [CrossRef] Minella, M.; Sordello, F.; Minero, C. Photocatalytic process in TiO2 /graphene hybrid materials. Evidence of charge separation by electron transfer from reduced graphene oxide to TiO2 . Catal. Today 2017, 281, 29–37. [CrossRef] Faraldos, M.; Bahamonde, A. Environmental applications of titania-graphene photocatalysts. Catal. Today 2017, 285, 13–28. [CrossRef] Gels 2022, 8, 215 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 18 of 19 Zeng, G.; You, H.; Du, M.; Zhang, Y.; Ding, Y.; Xu, C.; Liu, B.; Chen, B.; Pan, X. Enhancement of photocatalytic activity of TiO2 by immobilization on activated carbon for degradation of aquatic naphthalene under sunlight irradiation. Chem. Eng. J. 2021, 412, 128498. [CrossRef] Paušová, M.; Riva, M.; Baudys, M.; Krýsa, J.; Barbieriková, Z.; Brezová, V. Composite materials based on active carbon/TiO2 for photocatalytic water purification. Catal. Today 2019, 328, 178–182. [CrossRef] Khalid, N.; Majid, A.; Tahir, M.B.; Niaz, N.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [CrossRef] Biener, J.; Stadermann, M.; Suss, M.; Worsley, M.A.; Biener, M.M.; Rose, K.A.; Baumann, T.F. Advanced carbon aerogels for energy applications. Energy Environ. Sci. 2011, 4, 656–667. [CrossRef] Faria, J.L.; Wang, W. 13 Carbon Materials in Photocatalysis; John Wiley & Sons: Hoboken, NJ, USA, 2009; p. 481. Al-Muhtaseb, S.; Ritter, J. Preparation and Properties of Resorcinol-Formaldehyde Organic and Carbon Gels. Adv. Mater. 2003, 15, 101–114. [CrossRef] Prostredný, M.; Abduljalil, M.G.; Mulheran, P.A.; Fletcher, A.J. Process variable optimization in the manufacture of resorcinol– formaldehyde gel materials. Gels 2018, 4, 36. [CrossRef] [PubMed] Awadallah-F, A.; Al-Muhtaseb, S.A. Nanofeatures of resorcinol-formaldehyde carbon microspheres. Mater. Lett. 2012, 87, 31–34. [CrossRef] Awadallah-F, A.; Elkhatat, A.M.; Al-Muhtaseb, S.A. Impact of synthesis conditions on meso- and macropore structures of resorcinol-formaldehyde xerogels. J. Mater. Sci. 2011, 46, 7760–7769. [CrossRef] Principe, I.A.; Fletcher, A.J. Parametric study of factors affecting melamine-resorcinol-formaldehyde xerogels properties. Mater. Today Chem. 2018, 7, 5–14. [CrossRef] Shevlin, S.A.; Woodley, S.M. Electronic and Optical Properties of Doped and Undoped (TiO2 )n Nanoparticles. J. Phys. Chem. C 2010, 114, 17333–17343. [CrossRef] Jiang, Y.; Meng, L.; Mu, X.; Li, X.; Wang, H.; Chen, X.; Wang, X.; Wang, W.; Wu, F.; Wang, X. Effective TiO2 hybrid heterostructure fabricated on nano mesoporous phenolic resol for visible-light photocatalysis. J. Mater. Chem. 2012, 22, 23642–23649. [CrossRef] Zaleska, A. Doped-TiO2 : A review. Recent Pat. Eng. 2008, 2, 157–164. [CrossRef] Aranovich, G.L.; Donohue, M.D. Adsorption isotherms for microporous adsorbents. Carbon 1995, 33, 1369–1375. [CrossRef] Wang, S.; Zhu, Z.H.; Coomes, A.; Haghseresht, F.; Lu, G.Q. The physical and surface chemical characteristics of activated carbons and the adsorption of methylene blue from wastewater. J. Colloid Interface Sci. 2005, 284, 440–446. [CrossRef] [PubMed] Chham, A.; Khouya, E.; Oumam, M.; Abourriche, A.; Gmouh, S.; Mansouri, S.; Elhammoudi, N.; Hanafi, N.; Hannache, H. The use of insoluble mater of Moroccan oil shale for removal of dyes from aqueous solution. Chem. Int. 2018, 4, 67–77. Atout, H.; Bouguettoucha, A.; Chebli, D.; Gatica, J.M.; Vidal, H.; Yeste, M.P.; Amrane, A. Integration of Adsorption and Photocatalytic Degradation of Methylene Blue Using TiO2 Supported on Granular Activated Carbon. Arab. J. Sci. Eng. 2017, 42, 1475–1486. [CrossRef] Matos, J. Hybrid TiO2 -C Composites for the Photodegradation of Methylene Blue Under Visible Light; Bol Grupo Español Carbón: Zaragoza, Spain, 2013. Natarajan, T.S.; Bajaj, H.C.; Tayade, R.J. Preferential adsorption behavior of methylene blue dye onto surface hydroxyl group enriched TiO2 nanotube and its photocatalytic regeneration. J. Colloid Interface Sci. 2014, 433, 104–114. [CrossRef] [PubMed] Baker, F.S.; Miller, C.E.; Repik, A.J.; Tolles, E.D. Activated carbon. In Kirk-Othmer Encyclopedia of Chemical Technology; Interscience Publishers: Geneva, Switzerland, 2000. Zhang, X.; Zhang, F.; Chan, K.-Y. Synthesis of titania-silica mixed oxide mesoporous materials, characterization and photocatalytic properties. Appl. Catal. A Gen. 2005, 284, 193–198. [CrossRef] Ashraf, M.A.; Peng, W.; Zare, Y.; Rhee, K.Y. Effects of Size and Aggregation/Agglomeration of Nanoparticles on the Interfacial/Interphase Properties and Tensile Strength of Polymer Nanocomposites. Nanoscale Res. Lett. 2018, 13, 1–7. [CrossRef] [PubMed] Azizian, S. Kinetic models of sorption: A theoretical analysis. J. Colloid Interface Sci. 2004, 276, 47–52. [CrossRef] Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [CrossRef] Ayawei, N.; Ebelegi, A.N.; Wankasi, D. Modelling and Interpretation of Adsorption Isotherms. J. Chem. 2017, 2017, 3039817. [CrossRef] Tan, Y.; Kilduff, J.E. Factors affecting selectivity during dissolved organic matter removal by anion-exchange resins. Water Res. 2007, 41, 4211–4221. [CrossRef] [PubMed] Li, J.; Zhang, S.; Chen, C.; Zhao, G.; Yang, X.; Li, J.; Wang, X. Removal of Cu(II) and Fulvic Acid by Graphene Oxide Nanosheets Decorated with Fe3 O4 Nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 4991–5000. [CrossRef] [PubMed] Shao, D.D.; Fan, Q.H.; Li, J.X.; Niu, Z.W.; Wu, W.S.; Chen, Y.X.; Wang, X.K. Removal of Eu(III) from aqueous solution using ZSM-5 zeolite. Microporous Mesoporous Mater. 2009, 123, 1–9. [CrossRef] Zhang, G.; Ni, C.; Liu, L.; Zhao, G.; Fina, F.; Irvine, J.T.S. Macro-mesoporous resorcinol–formaldehyde polymer resins as amorphous metal-free visible light photocatalysts. J. Mater. Chem. A 2015, 3, 15413–15419. [CrossRef] Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [CrossRef] Gels 2022, 8, 215 39. 40. 41. 42. 43. 44. 45. 46. 19 of 19 Huang, X.; Yang, W.; Zhang, G.; Yan, L.; Zhang, Y.; Jiang, A.; Xu, H.; Zhou, M.; Liu, Z.; Tang, H.; et al. Alternative synthesis of nitrogen and carbon co-doped TiO2 for removing fluoroquinolone antibiotics in water under visible light. Catal. Today 2019, 361, 11–16. [CrossRef] Simonetti, E.A.N.; de Simone Cividanes, L.; Campos, T.M.B.; de Menezes, B.R.C.; Brito, F.S.; Thim, G.P. Carbon and TiO2 synergistic effect on methylene blue adsorption. Mater. Chem. Phys. 2016, 177, 330–338. [CrossRef] Wu, C.H.; Kuo, C.Y.; Chen, S.T. Synergistic effects between TiO2 and carbon nanotubes (CNTs) in a TiO2 /CNTs system under visible light irradiation. Environ. Technol. 2013, 34, 2513–2519. [CrossRef] Houas, A.; Lachheb, H.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation pathway of methylene blue in water. Appl. Catal. B: Environ. 2001, 31, 145–157. [CrossRef] Lakshmi, S.; Renganathan, R.; Fujita, S. Study on TiO2 -mediated photocatalytic degradation of methylene blue. J. Photochem. Photobiol. A Chem. 1995, 88, 163–167. [CrossRef] Makuła, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [CrossRef] [PubMed] Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [CrossRef] Sahoo, T.R.; Prelot, B. Adsorption processes for the removal of contaminants from wastewater: The perspective role of nanomaterials and nanotechnology. In Nanomaterials for the Detection and Removal of Wastewater Pollutants; Elsevier: Amsterdam, The Netherlands, 2020; pp. 161–222.