Next Article in Journal
A Novel Hybrid Membrane Process Coupled with Freeze Concentration for Phosphorus Recovery from Cheese Whey
Previous Article in Journal
Characterization of New Polymer Material of Amino-β-Cyclodextrin and Sodium Alginate for Environmental Purposes
Previous Article in Special Issue
Polyhexanide-Releasing Membranes for Antimicrobial Wound Dressings: A Critical Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Performance of TiO2-Based Tubular Membranes in the Photocatalytic Degradation of Organic Compounds

1
Departamento de Ingenierías Química y Biomolecular, Universidad de Cantabria, Avda. Los Castros s/n, 39005 Santander, Spain
2
Graduate School of Science and Technology for Innovation, Graduate School Science and Engineering, Yamaguchi University, Ube 755-8611, Japan
*
Author to whom correspondence should be addressed.
Membranes 2023, 13(4), 448; https://doi.org/10.3390/membranes13040448
Submission received: 23 March 2023 / Revised: 17 April 2023 / Accepted: 19 April 2023 / Published: 20 April 2023

Abstract

:
This work presents the photocatalytic degradation of organic pollutants in water with TiO2 and TiO2/Ag membranes prepared by immobilising photocatalysts on ceramic porous tubular supports. The permeation capacity of TiO2 and TiO2/Ag membranes was checked before the photocatalytic application, showing high water fluxes (≈758 and 690 L m−2 h−1 bar−1, respectively) and <2% rejection against the model pollutants sodium dodecylbenzene sulfonate (DBS) and dichloroacetic acid (DCA). When the membranes were submerged in the aqueous solutions and irradiated with UV-A LEDs, the photocatalytic performance factors for the degradation of DCA were similar to those obtained with suspended TiO2 particles (1.1-fold and 1.2-fold increase, respectively). However, when the aqueous solution permeated through the pores of the photocatalytic membrane, the performance factors and kinetics were two-fold higher than for the submerged membranes, mostly due to the enhanced contact between the pollutants and the membranes photocatalytic sites where reactive species were generated. These results confirm the advantages of working in a flow-through mode with submerged photocatalytic membranes for the treatment of water polluted with persistent organic molecules, thanks to the reduction in the mass transfer limitations.

1. Introduction

Clean water is an essential resource for sustainable development. However, the water supply is struggling to meet the increasing demand in many places [1]. Furthermore, climate change is expected to intensify water scarcity [2]. Accordingly, the importance of water reclamation and reuse is increasing. In addition, industrial developments have increased the variety of chemicals in water bodies. Some organic pollutants are toxic even at low concentrations and have low or no biodegradability, which limits the implementation of conventional biological treatments.
Photocatalysis is a promising alternative technology for various applications, including air and water purification, renewable energy production and materials’ synthesis [3]. With its potential to address global environmental challenges, photocatalysis has attracted significant research interest and investment in recent years. Special emphasis is placed on the removal of persistent organic pollutants from water, as this method is attractive because it can both degrade and mineralize dissolved organics compounds under ambient conditions without the addition of extra chemicals [4]. Photocatalysis is based on the activation of a semiconductor material when it is irradiated with light photons. These photons cause an electron (e) to move from the valence band to the conduction band if the energy of the photons is greater than the band-gap of the semiconductor. Once the electron has been absorbed in the conduction band, a positive hole (h+) is formed on the valence band. When the electron/hole pair is produced, several reactive oxygen species are generated in the reaction medium, such as hydroxyl radicals or superoxide radicals, which are responsible for the degradation of the pollutants [5,6]. Titanium dioxide (TiO2) is a widely studied photocatalyst that has various advantages such as low toxicity, high stability and cost-effectiveness [7]. One of the major drawbacks of TiO2 is its wide band gap, ≈3.1–3.2 eV, which requires UV light for activation. Therefore, efforts have been made to develop visible light-active photocatalysts by modifying TiO2 or using other types of semiconductor materials [8,9,10,11,12,13,14,15,16]. Noble metals, such as Ag and Au, are commonly used to enhance the photocatalytic performance of TiO2, as they act as an electron sink. The formation of Schottky junctions between semiconductors and noble metal nanoparticles increases the efficiency of charge carrier separation [17,18].
In laboratory research, photocatalytic activity is often evaluated by mixing the powdery photocatalysts into the solution. The particles must therefore be removed from the treated water at the end of the experiment, by filtration, centrifugation or an external magnetic field in the presence of magnetic photocatalysts [19,20,21,22]. Alternatively, photocatalysts can be immobilised on a support, and then an extra facility to separate particles from water is not required [23]. One of the challenges of supported photocatalyst is the increase in mass transfer resistance, which can lead to slower kinetics. Another issue is the reactor design, as the total catalytic surface area per reactor volume tends to be smaller compared to a slurry-type reactor. Proper illumination plays a critical role in the removal of contaminants of emerging concern and the energy efficiency of photoreactors [24]. One possibility to assess whether the photocatalytic activity is improved is to work with a porous support, such as a porous membrane [25,26,27]. The use of ceramic membranes can be advantageous compared to organic polymeric membranes, overcoming the degradation of polymers under UV exposure [28,29]. Ceramic membranes typically have asymmetric structures with a thin separation layer and a macroporous support that provides the required mechanical strength. Geltmeyer et al. (2017) immobilised TiO2 nanoparticles on ceramic nanofibrous membranes by dip-coating for the photodegradation of the herbicide isoproturon [30]. Ahmad et al. (2020) coated TiO2 layer on a porous Al2O3 membrane substrate, developed using low-cost poly (vinyl chloride) (PVC) to generate interstitial voids for application in a photocatalytic membrane reactor for the treatment of organic dye contaminant. The photocatalytic activity of the membranes was evaluated using a batch photocatalytic reactor for the degradation of two dyes: acid orange (AO) and congo red (CR) [31]. Deepracha et al. (2021) deposited an active layer of commercial TiO2 powders (P25) on a microporous alumina membrane with pore size of 0.2 µm by the dip-coating method. This membrane was tested to examine the decomposition of phenol in water. The reaction was carried out using an easily up-scalable photocatalytic membrane reactor with four single-channel tubular membranes at transmembrane pressures of 250 mbar [32]. From another perspective, Singhapong et al. (2019) used porous mullites as ceramic membranes and coated them with TiO2 powders for disinfection applications. Uncoated mullites were used as control samples. The coating of TiO2 was able to inactivate E. coli under light exposure [33].
A combination of photocatalysis and membranes may realize synergetic performances, improving the contact between pollutants and photocatalyst. Very few studies have been carried out related to this topic. Albu et al. (2007) synthesised a dense and free-standing TiO2 nanotube membrane that was vertically oriented. Working in a flow-through operation, they reported the complete degradation of 2 × 10−3 mmol L−1 of the dye methylene blue after 4 h of UV irradiation [34]. Berger et al. (2020) evaluated the photocatalytic degradation of methylene blue. They operated under flow-through conditions with an aluminum oxide membrane coated with TiO2 via atomic layer deposition [35]. Regarding the degradation of volatile organic compounds, Moulis et al. (2015) described the oxidative degradation of acetone and methanol in gaseous phase in a flow-through photoreactor with immobilized TiO2. They focused their attention on the influence of the wavelength of UV light [36]. Presumido et al. (2022) presented a ceramic tubular membrane coated with a continuous graphene-TiO2 nanocomposite thin-film for the removal of contaminants of emerging concern in a single-pass flow-through operation [37]. They observed an important reduction in the fouling of the membranes during the filtration of the contaminated fluid when UV light was applied. Lofti et al. (2022) used a polyethersulfone-TiO2 membrane for the photocatalytic degradation of steroid hormone micropollutants in a continuous flow-through process [38]. The results indicate that the efficiency of the photocatalytic degradation of the micropollutants varied from 33% to 94%, depending on the type of hormone, concentration, and operating conditions.
In the present study, commercial TiO2 powder (P25, Evonik Industries, Essen, Germany) was immobilised on porous ceramic mullite tubes producing membranes with porosity in the range of microfiltration, in contrast to previous studies. Additionally, silver was photochemically deposited on some of the membranes. The permeation properties and photocatalytic performance of TiO2 and TiO2/Ag membranes were tested both with and without flow-through the membranes, all while under the influence of UV light irradiation. Two organic pollutants were chosen as examples of contaminants of emerging concern (CECs) in water: (i) sodium dodecylbenzene sulfonate (DBS), C18H29NaO3S, an important anionic surfactant used in the formulation of shampoos, detergents or cleaning products and, (ii) dichloroacetic acid (DCA), C2H2Cl2O2, a haloacetic acid used as fungicide and as a chemical intermediate in pharmaceuticals manufacture.

2. Materials and Methods

2.1. Materials

TiO2 (P25) photocatalyst (≈80% anatase and 20% rutile) with a specific surface area of approximately 55 m2 g−1 [39,40] was provided from Evonik Industries, Essen, Germany. Porous mullite tubes (outer diameter 12mm, inner diameter 9 mm, length 100 mm, mean pore size 1.3 μm, porosity 2%) were purchased from NIKKATO Corporation, Osaka, Japan. Silver acetate (CH3COOAg, purity 97%) was purchased from FUJIFILM Wako Pure Chemical Corporation, Japan. Sodium dodecylbenzene sulfonate (DBS) and dichloroacetic acid (DCA) were purchased from Sigma-Aldrich Chemie GmbH (Buchs, Switzerland) and Acros Organics (ThermoFisher Scientific, Geel, Belgium), respectively.

2.2. Methods

2.2.1. Membrane Preparation and Characterisation

TiO2 particles were mechanically scrubbed onto the outer surface of the mullite tubes and heated in air at 673 K for three hours. After cooling down to room temperature, the membranes were washed with water to remove any loosely attached particles and dried at 353 K. Approximately 1 × 10−2 mmol cm−2 of TiO2 was immobilised on the membrane, calculated from the mass uptake after the deposition.
TiO2 membranes were immersed in 1 × 10−2 mmol L−1 silver acetate solution and exposed to UV light for 1.5 h. Black lamps (FL8BLB, Toshiba Corporation, Tokyo, Japan, λmax = 352 nm, 3.3 mW cm−2) were used as the light source. Details can be found elsewhere [41]. During light irradiation, silver ions were photochemically reduced on the TiO2 surface [42]. The amount of silver deposited was calculated from the decrease in silver concentration in the solution, which was measured by inductively coupled plasma (ICP, SPS3500, S2 nanotechnology corporation, Yokohama, Japan). Approximately 3 × 10−5 mmol cm−2 of silver was deposited on the membrane. The membrane area was estimated to be about 29 cm2 for TiO2 membrane and 30 cm2 for TiO2/Ag membrane.
Characterisation of the membranes included scanning electron microscopy (FE-SEM, JSM-7600F, JEOL Ltd., Tokyo, Japan) and X-ray diffraction (XRD, SmartLab. Rigaku corporation, Tokyo, Japan).

2.2.2. Permeation Tests

The permeation properties of the membranes were evaluated at room temperature (approximately 295 K) using either ultrapure water (MilliQ, Millipore, Burlington, VT, USA) or DBS model aqueous solutions. The experimental system used in the permeation tests is shown in Figure 1. The membrane was immersed in a liquid reservoir. One end of the membrane was blocked with a sealed plug and the other end was connected to a vacuum pump (Millivac Maxi SD1P014M04, Millipore, Darmstadt, Germany), which sucked the liquid from the reservoir through the membrane. A glass trap was connected between the membrane and the vacuum pump to collect the permeate, which was continuously monitored with a weight balance (PS 6000.R2, RADWAG, Spain) and recorded using Pomiar Win software. The transmembrane pressure ΔP was estimated as the difference between the atmospheric pressure (Pr) and the vacuum pressure (Pv) adjustable in the pump (Equation (1)).
P = P r P v
The membrane flux (L m−2 h−1), J, was calculated using Equation (2), where Q p e r m e a t e is the permeate flux (L h−1) and A e (m2) is the effective membrane area:
J = Q p e r m e a t e A e
The hydraulic permeability (L m−2 h−1 bar−1), K w , was estimated as indicated in Equation (3), considering the ultrapure water flux, J w :
K w = J w P
For the analysis of the permeability of organic model solutions, DBS (348.48 g mol−1) was chosen because of its large molecular size compared to DCA (128.94 g mol−1). A feed solution containing 50 mg L−1 of DBS was prepared. In addition to the flux of the model solution through the membrane samples, the compound rejection was estimated (Equation (4)):
R % = 1 C 0 C × 100
After each use of the membrane, a cleaning protocol was followed, which consisted of washing the membrane several times with ultrapure (UP) water and allowing it to dry at room temperature.

2.2.3. Photocatatlytic Activity Tests

Photocatalytic experiments were performed in a 1 L Pyrex glass photoreactor purchased to APRIA Systems SL (Cantabria, Spain). Figure 2 depicts the system. The equipment was provided with UV-A LED technology emitting at a fixed wavelength of 365 nm. The reactor housing had 30 LEDs (ENGIN LZ1-00UV00) distributed in 10 strips, 3 LEDs per strip, so as to homogeneously illuminate the entire height of the reactor. Strips were uniformly placed at a distance of 1.50 cm from the photoreactor. The solution was irradiated with an average irradiance of 200 W m−2. The reactor was placed on a magnetic stirrer to constantly mix the solution. The membrane was completely immersed in the aqueous feed solution. A lid, which ensured sufficient oxygenation of the aqueous medium, was placed to fix the membrane in the center of the vessel (see detail of the cross-section of the reactor in Figure 2).
First, dark adsorption experiments were carried out for 2 h. A total of 50 mg L−1 of DCA and DBS solutions were used with constant stirring. In order to evaluate the photocatalytic activity, the membranes were submerged inside the aqueous solutions. In this first approach, photocatalyst was immobilized on the membrane and operated without any permeation. The photocatalytic performance of the submerged membranes was then tested for 6 h at temperature and pH conditions similar to those used in the adsorption experiments.
In a second approach, coupled photocatalysis and permeation (flow-through mode) experiments were performed. In this case, the vacuum system used in the permeation tests was connected to the photocatalytic reactor. As a result, the aqueous solution was forced to permeate through the membrane, thus improving the solid–liquid contact between the photocatalyst and the pollutant solutions. The permeated volume, which had a total volume of 200 mL, was collected in the vacuum trap and subsequently returned to the feed tank at regular intervals of approximately 4–5 min. The total duration of the experiment was 6 h. The total flux through the membranes, the concentration of organic compounds in the feed tank and the accumulated permeate volume (4–5 min) were monitored.
All the experiments were carried out at room temperature (approximately at 295–298 K), natural pH (3.4 for DCA and 5.6 for DBS model solutions) and under constant stirring. All the experiments were performed with the same membrane samples.
The concentration of DCA was determined using an AS9-HC column in an ICS-5000 (Dionex, ThermoFisher Scientific, Waltham, MA, USA) ion chromatograph with a 9 mM Na2CO3 solution as eluent, at a flow rate of 1 mL min−1 and a pressure of approximately 2000 psi. The concentration of DBS was determined by UV spectrophotometry (UV-1800, Shimadzu Europe, Duisburg, Germany) at a wavelength of 223 nm. For UV spectrophotometry, a calibration curve was established with standard solutions up to 60 mg L−1.

3. Results

3.1. Characterisation

Figure 3 shows the membrane surface and a cross-sectional view. The outer surface of the ceramic tube was completely covered with TiO2 particles. The thickness of the TiO2 layer was approximately 2–3 μm. Some of the TiO2 membranes were immersed in a silver acetate solution and exposed to UV light. The concentration of silver in the solution decreased due to photoreduction, because silver ions were photochemically reduced on TiO2. The colour of the membrane changed from white to light yellow after UV irradiation. Under these conditions, silver can be deposited as metallic silver and silver oxide [43].
Figure 4 shows the XRD patterns of TiO2 particles (P25), mullite support, TiO2 membrane by immobilisation of P25 particles on mullite support and TiO2/Ag membrane. The P25 powder consists mainly of anatase phase with some rutile phase (≈80/20) [44]. The standard XRD pattern of TiO2 was checked through the Crystallographic Open Database (COD), card no. 7206075. The intensified diffraction peaks in the 2θ range at 25.4, 37.1, 48.14, 54.0, 55.2, 62.9, 68.7, 69.7 and 75.1° correspond to the (101), (004), (200), (105), (211), (204), (116), (220) and (215) crystallographic planes belonging to the anatase phase with tetragonal structure, respectively [45,46,47]. The peak at 27.4° is the characteristic reflection for rutile, whose crystallographic plane is (110) [48]. The uncoated mullites could be classified as microfiltration membranes [33]. The XRD pattern of mullite carrier at diffraction angles 2θ of 16.4, 26.1, 33.2, 35.3, 40.8° are assigned to (110), (210), (220), (111) and (121) crystallographic planes respectively [28]. Alumina phase is also detected in the sample, which could be explained by the transformation of mullite [49,50,51]. Peaks corresponding to mullite support and TiO2 were found in the TiO2 and TiO2/Ag membranes. The immobilisation of TiO2 on ceramic tubes did not significantly affect the TiO2 phases, since the annealing temperature used in this study, 673 K, is lower than the temperature that causes anatase to transition to rutile. The presence of silver was confirmed in earlier research through X-ray photoelectron spectroscopy (XPS) analysis [43]. Further characterisations of the membranes can be found in previous publications [41,43].

3.2. Permeation Performance

TiO2 and TiO2/Ag membranes showed an average hydraulic permeability of 758 ± 109 and 690 ± 96 L m−2 h−1 bar−1, respectively. The transmembrane pressure of the hydraulic permeability experiments varied from 0.95 to 1.07 bar. A similar hydraulic permeance (845 L m−2 h−1 bar−1) was reported by Ma et al. (2010) for porous α-Al2O3 ceramic disc supports coated with a hydroxyapatite (HPA) and TiO2/Ag layer [52]. The permeances of the DBS solutions were 611 ± 46 and 426 ± 63 L m−2 h−1 bar−1, for TiO2 and TiO2/Ag membranes, respectively. In both cases, the permeances for the DBS solutions are in the range of the hydraulic permeances of the TiO2 and TiO2/Ag membranes for water. Although the membrane permeance of the DBS solutions was generally slightly lower than the hydraulic permeance, none of the membranes presented permanent fouling and the membrane flux of all feed solutions was within the range of water permeation values. This is coherent with the low rejection of DBS (<2%) provided by the two types of membranes. Therefore, DBS was not retained by either TiO2 or TiO2/Ag membranes. Workneh and Shukla (2008) showed a rejection of sodium dodecyl sulfate in the range of 10–45% and a hydraulic permeability of 270 L m−2 h−1 bar−1 in a sodalite octahydrate-zeolite clay composite membrane [53]. Zhang et al. (2006) reported a rejection of 60–70% of DBS with a silica/titania nanorods/nanotubes composite membrane and a DBS solution permeation flux of 47 L m−2 h−1 at 0.5 bar ΔP [54]. When comparing results from the literature, it was observed that the membranes prepared in this work offer higher hydraulic permeabilities at the expense of a low rejection of organic molecules in the size range of DBS, confirming that the pore size of the tubular ceramic membranes remained in the microfiltration range after photocatalyst deposition.
In order to assess the stability and activity of the two membranes, TiO2 and TiO2/Ag, several interspersed cycles with water and DBS were carried out. These cycles involved the permeation of both water and DBS solutions through the membranes. After five consecutive cycles, the results showed no significant change in flux for either membrane when water or DBS solutions were permeated. This indicates that the activity of the membranes remained stable, and no loss of activity was detected during the cycles. These results provide valuable insights into the durability and performance of the TiO2 and TiO2/Ag membranes, indicating their potential usefulness for various applications in this field.

3.3. Photocatalytic Performance

TiO2 and TiO2/Ag membranes did not adsorb either DCA or DBS after 4 h of contact in the absence of light. Figure 5 shows the change in the dimensionless concentration of DCA and DBS during the photocatalytic tests using TiO2 and TiO2/Ag membranes. It could be observed that the TiO2/Ag membrane provided a slightly higher degradation percentage than TiO2 for both pollutants. The TiO2 membrane showed a degradation percentage of 16.6% for DBS and 37% for DCA after 6 h, whereas TiO2/Ag presented a degradation percentage of 21% for DBS and 44% for DCA. Experiments were carried out in duplicate with less than 5% error.
As observed, there is a slight increase in the removal of pollutants with the TiO2/Ag membrane, which could be attributed to the decrease in the electron/hole recombination rate due to the presence of small amounts of Ag, as Ag can act as an electron sink [9,55,56,57,58,59]. The electrons in the conduction band can be easily transferred to Ag, suppressing the recombination of the electron/hole pair, leading to more electrons becoming involved in the photocatalytic reaction. Therefore, the effective spatial separation of charge carriers was achieved, and led to the enhancement of the photocatalytic reaction [18]. The data fitted well to a pseudo-first-order kinetic model, typical for the degradation of organic pollutants by TiO2 photocatalysts, with 1.30 × 10−3 min−1 and 1.60 × 10−3 min−1 representing the kinetic constants corresponding to DCA degradation, and 5.02 × 10−4 min−1 and 6.27 × 10−4 min−1 representing the kinetic constants for DBS degradation, respectively, for the membranes without and with Ag.
Figure 6 presents the total flux and the dimensionless DCA concentration decay during the simultaneous photocatalysis and flow-through permeation experiments with TiO2 (Figure 6a) and TiO2/Ag membranes (Figure 6b). It also compares the DCA concentration decay in the photocatalytic experiments without permeation with the photocatalytic material supported on the membrane. Experimental results are shown with error bars; the average standard deviation was lower than 5% for all samples.
A negligible loss of activity with these operation cycles was previously confirmed. Furthermore, the DCA concentration in the tank and in the collected permeate were the same each time, also confirming the null DCA rejection of TiO2 and TiO2/Ag membranes. Interestingly, when the feed solution permeated the membrane pores, the DCA degradation rate was increased. For the TiO2 membrane, a kinetic constant k of 2.50 × 10−3 min−1 (R2 = 0.99) was obtained, which is almost two-fold higher than that obtained when working with the submerged membrane. The TiO2/Ag membrane operating in a flow-through mode provided a value of k of 2.80 × 10−3 min−1 (R2 = 0.99), which is 1.8-fold higher than the kinetic constant corresponding to the submerged membrane without permeation. These results confirm that as the aqueous solution permeated the pores of the membrane, the contact between the pollutants and the oxidant species generated on the surface of the TiO2 and TiO2/Ag membranes was facilitated, which otherwise could not react with the pollutant molecules due to the mass transport limitations [38]. The flow-through mode, therefore, enhanced the degradation rate of the photocatalytic process and is emerging as a promising option [34,35,36,37,38,60]. Again, the results of the TiO2/Ag membrane are slightly better than those of the TiO2 membrane, due to the positive effect of Ag as an electron sink [61,62]. This confirms that the presence of Ag, even in small amounts, has the ability to improve the photocatalytic results. Bian et al. (2013) reported a strategy for loading uniformly Ag nanoparticles in flow-through TiO2 nanotube arrays [61]. They tested the nanotubes arrays in the degradation of methyl orange under UV light, reaching complete degradation in 30 min. They attributed the good results to the good distribution of silver in the nanotube arrays. Working in a flow-through operation mode significantly attenuates the difference in the photocatalytic performance between TiO2 and TiO2/Ag membranes. Furthermore, in economic terms, it does not compensate for the presence of silver in the preparation of the materials.
For comparison, DCA degradation experiments with commercial TiO2 in suspension were carried out in the same photocatalytic reactor. To facilitate the comparison between suspended and immobilised TiO2, a performance factor was calculated as the ratio of initial degradation rate and catalyst concentration (Table 1).
An analysis of the results in Table 1 showed that the performance factor was similar (1.1-fold higher) when TiO2 was used in suspension and for the submerged membrane, indicating that, in this case, the immobilised photocatalyst preserved the photocatalytic activity compared to slurry systems. Interestingly, the performance factor of the immobilised photocatalysts operating with the submerged membrane in flow-through mode was 2.2-fold higher than with TiO2 in suspension. The membranes did not retain the pollutants, so the increase in the kinetic constants could be attributed to the improved contact between the pollutants and the oxidative species favored by the transport through the pores of the membrane. Therefore, these results demonstrate that the performance of photocatalytic membranes for the degradation of aqueous pollutants is enhanced when working in a flow-through mode because the contact with reactive species is favored and the resistance to mass transport is reduced, leading to an increase in the overall degradation rate.

4. Conclusions

Water remediation is crucial to environmental preservation, removing organic compounds from water sources. The use of combined technologies is an interesting approach for this purpose. In this work, mullite tubular membranes were employed as supports for immobilised TiO2 and TiO2/Ag. These membranes were used to evaluate the hydraulic permeability, with values of ≈758 and 690 L m2 h1 bar1 obtained for TiO2 and TiO2/Ag membranes, respectively. The permeabilities for DBS solutions were in the range of the hydraulic permeabilities of the membranes. In addition, neither TiO2 nor TiO2/Ag membranes retained DBS. Afterwards, the same membranes were tested for the photocatalytic degradation of DCA and DBS synthetic solutions. Although silver improves the photocatalytic properties, the improvement achieved under the conditions of this work did not support the large-scale manufacture of silver-containing membranes. Finally, the combination of both technologies, filtration and photocatalysis, operating in a single flow-through mode for the treatment of synthetic solutions activated by UV-A light improves the process’ effectiveness. Working in flow-through mode with the TiO2 membrane improves the performance factor by 2.2-fold compared to suspended TiO2. This improvement is directly related to the enhanced contact between the organic molecules and the oxidative species generated on the surface of the membrane under UV-A light, and reinforces one of the main advantages of immobilised membranes, facilitating the catalyst recovery after the photocatalytic treatment. The results of this study indicate that working with the submerged membrane in flow-through mode is a promising approach for practical environmental applications.

Author Contributions

Conceptualization, I.K., N.D. and M.J.R.; methodology, C.B., A.V.-G., N.D. and M.J.R.; investigation, C.B. and A.V.-G..; writing—original draft preparation, C.B., I.K., N.D. and M.J.R.; writing—review and editing, C.B., I.K., N.D., M.J.R., A.U. and I.O.; supervision, N.D. and M.J.R.; project administration, I.K., N.D. and M.J.R.; funding acquisition, I.K. and N.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by JST SICORP Grant Number JPMJSC18C5 (Japan) and the grant number PCI2018-092929 funded by MCIN/ AEI/10.13039/501100011033/ (Spain) as part of the project X-MEM within the framework of the EIG CONCERT-Japan 5th Joint Call “Functional Porous Materials”.

Institutional Review Board Statement

Not applicable.

Acknowledgments

Carmen Barquín is grateful for the FPI contract PRE2019-089096.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hoekstra, A.Y. Water scarcity challenges to business. Nat. Clim. Chang. 2014, 4, 318–320. [Google Scholar] [CrossRef]
  2. Schewe, J.; Heinke, J.; Gerten, D.; Haddeland, I.; Arnell, N.W.; Clark, D.B.; Dankers, R.; Eisner, S.; Fekete, B.M.; Colón-González, F.J. Multimodel assessment of water scarcity under climate change. Proc. Natl. Acad. Sci. USA 2014, 111, 3245–3250. [Google Scholar] [CrossRef]
  3. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  4. Qu, X.; Alvarez, P.J.; Li, Q. Applications of nanotechnology in water and wastewater treatment. Water Res. 2013, 47, 3931–3946. [Google Scholar] [CrossRef] [PubMed]
  5. Nosaka, Y.; Nosaka, A.Y. Generation and detection of reactive oxygen species in photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  6. Byrne, C.; Subramanian, G.; Pillai, S.C. Recent advances in photocatalysis for environmental applications. J. Environ. Chem. Eng. 2018, 6, 3531–3555. [Google Scholar] [CrossRef]
  7. Wang, Y.; Sun, C.; Zhao, X.; Cui, B.; Zeng, Z.; Wang, A.; Liu, G.; Cui, H. The application of nano-TiO2 photo semiconductors in agriculture. Nanoscale Res. Lett. 2016, 11, 1–7. [Google Scholar] [CrossRef]
  8. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. [Google Scholar] [CrossRef] [PubMed]
  9. Ribao, P.; Rivero, M.J.; Ortiz, I. TiO2 structures doped with noble metals and/or graphene oxide to improve the photocatalytic degradation of dichloroacetic acid. Environ. Sci. Pollut. Res. 2017, 24, 12628–12637. [Google Scholar] [CrossRef] [PubMed]
  10. Zarrin, S.; Heshmatpour, F. Photocatalytic activity of TiO2/Nb2O5/PANI and TiO2/Nb2O5/RGO as new nanocomposites for degradation of organic pollutants. J. Hazard. Mater. 2018, 351, 147–159. [Google Scholar] [CrossRef]
  11. Xu, L.; Yang, L.; Johansson, E.M.J.; Wang, Y.; Jin, P. Photocatalytic activity and mechanism of bisphenol a removal over TiO2−x/rGO nanocomposite driven by visible light. Chem. Eng. J. 2018, 350, 1043–1055. [Google Scholar] [CrossRef]
  12. Zhu, D.; Zhou, Q. Action and mechanism of semiconductor photocatalysis on degradation of organic pollutants in water treatment: A review. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100255–100265. [Google Scholar] [CrossRef]
  13. Rivero, M.J.; Ribao, P.; Gomez-Ruiz, B.; Urtiaga, A.; Ortiz, I. Comparative performance of TiO2-rGO photocatalyst in the degradation of dichloroacetic and perfluorooctanoic acids. Sep. Purif. Technol. 2020, 240, 116637–116643. [Google Scholar] [CrossRef]
  14. Dorosheva, I.B.; Valeeva, A.A.; Rempel, A.A.; Trestsova, M.A.; Utepova, I.A.; Chupahkin, O.N. Synthesis and physicochemical properties of nanostructured TiO2 with enhanced photocatalytic activity. Inorg. Mater. 2021, 57, 503–510. [Google Scholar] [CrossRef]
  15. Tsebriienko, T.; Popov, A.I. Effect of poly(titanium oxide) on the viscoelastic and thermophysical properties of interpenetrating polymer networks. Crystals 2021, 11, 794. [Google Scholar] [CrossRef]
  16. Trestsova, M.A.; Utepova, I.A.; Chupakhin, O.N.; Semenov, M.V.; Pevtsov, D.N.; Nikolenko, L.M.; Tovstun, S.A.; Gadomska, A.V.; Shchepochkin, A.V.; Kim, G.A.; et al. Oxidative C-H/C-H coupling of dipyrromethanes with azines by TiO2-based photocatalytic system. Synthesis of new BODIPY dyes and their photophysical and electrochemical properties. Molecules 2021, 26, 5549. [Google Scholar] [CrossRef]
  17. de Almeida, G.C.; Mohallem, N.D.S.; Viana, M.M. Ag/GO/TiO2 nanocomposites: The role of the interfacial charge transfer for application in photocatalysis. Nanotechnology 2022, 33, 35710. [Google Scholar] [CrossRef]
  18. Lee, Y.; Kim, T.; Kim, B.; Choi, S.; Kim, K. Synthesis of TiO2/MoSx/Ag nanocomposites via photodeposition for enhanced photocatalysis and membrane fouling mitigation. J. Environ. Chem. Eng. 2023, 11, 109266. [Google Scholar] [CrossRef]
  19. Molinari, R.; Borgese, M.; Drioli, E.; Palmisano, L.; Schiavello, M. Hybrid processes coupling photocatalysis and membranes for degradation of organic pollutants in water. Catal. Today 2002, 75, 77–85. [Google Scholar] [CrossRef]
  20. Kochkodan, V.M.; Rolya, E.A.; Goncharuk, V.V. Photocatalytic membrane reactors for water treatment from organic pollutants. J. Water. Chem. Technol. 2009, 31, 227–237. [Google Scholar] [CrossRef]
  21. Mozia, S. Photocatalytic membrane reactors (PMRs) in water and wastewater treatment. A review. Sep. Purif. Technol. 2010, 73, 71–91. [Google Scholar] [CrossRef]
  22. Yang, X.; Sun, H.; Li, G.; An, T.; Choi, W. Fouling of TiO2 induced by natural organic matters during photocatalytic water treatment: Mechanisms and regeneration strategy. Appl. Catal. B Environ. 2021, 294, 120252–120262. [Google Scholar] [CrossRef]
  23. Balasubramanian, G.; Dionysiou, D.D.; Suida, M.T.; Baudin, I.; Laîne, J.M. Evaluating the activities of immobilized TiO2 powder films for the photocatalytic degradation of organic contaminants in water. Appl. Catal. B Environ. 2004, 47, 73–84. [Google Scholar] [CrossRef]
  24. Espíndola, J.C.; Vilar, V.J.P. Innovative light-driven chemical/catalytic reactors towards contaminants of emerging concern mitigation: A review. Chem. Eng. J. 2022, 394, 124865. [Google Scholar] [CrossRef]
  25. Iglesias, O.; Rivero, M.J.; Urtiaga, A.M.; Ortiz, I. Membrane-based photocatalytic systems for process intensification. Chem. Eng. J. 2016, 305, 138–148. [Google Scholar] [CrossRef]
  26. Tetteh, E.K.; Rathilal, S.; Asante-Sackey, D.; Chollom, M.N. Prospects of synthesized magnetic TiO2-based membranes for wastewater treatment: A review. Materials 2021, 14, 3524. [Google Scholar] [CrossRef]
  27. Romay, M.; Diban, N.; Rivero, M.J.; Urtiaga, A.; Ortiz, I. Critical issues and guidelines to improve the performance of photocatalytic polymeric membranes. Catalysts 2020, 10, 570. [Google Scholar] [CrossRef]
  28. Kim, S.H.; Kwak, S.Y.; Sohn, B.H.; Park, T.H. Design of TiO2 nanoparticle self-assembled aromatic polyamide thin-film-composite (TFC) membrane as an approach to solve biofouling problem. J. Membr. Sci. 2003, 211, 157–165. [Google Scholar] [CrossRef]
  29. Song, H.; Shao, J.; He, Y.; Liu, B.; Zhong, X. Natural organic matter removal and flux decline with PEG–TiO2-doped PVDF membranes by integration of ultrafiltration with photocatalysis. J. Membr. Sci. 2012, 405, 48–56. [Google Scholar] [CrossRef]
  30. Geltmeyer, J.; Teixido, H.; Meire, M.; Van Acker, T.; Deventer, K.; Vanhaecke, F.; Van Hulle, S.; De Buysser, K.; De Clerck, K. TiO2 functionalized nanofibrous membranes for removal of organic (micro)pollutants from water. Sep. Purif. Technol. 2017, 179, 533–541. [Google Scholar] [CrossRef]
  31. Ahmad, R.; Lee, C.S.; Kim, J.H.; Kim, J. Partially coated TiO2 on Al2O3 membrane for high water flux and photodegradation by novel filtration strategy in photocatalytic membrane reactors. Chem. Eng. Res. Des. 2020, 163, 138–148. [Google Scholar] [CrossRef]
  32. Deeoracha, S.; Atfane, L.; Ayral, A.; Ogawa, M. Simple and efficient method for functionalizing photocatalytic ceramic membranes and assessment of its applicability for wastewater treatment in up-scalable membrane reactors. Sep. Purif. Technol. 2021, 262, 118307. [Google Scholar] [CrossRef]
  33. Singhapong, W.; Jaroenworaluck, A.; Pansri, R.; Chokevivat, W.; Manpetch, P.; Miyauchi, M.; Srinophakun, P. Mullite membrane coatings: Antibacterial activities of nanosized TiO2 and Cu-grafted TiO2 in the presence of visible light illumination. Appl. Phys. A 2019, 125, 244. [Google Scholar] [CrossRef]
  34. Albu, S.P.; Ghicov, A.; Macak, J.M.; Hahn, R.; Schmuki, P. Self-organized, free-standing TiO2 nanotube membrane for flow-through photocatalytic applications. Nano Lett. 2007, 7, 1286–1289. [Google Scholar] [CrossRef]
  35. Berger, T.E.; Regmi, C.; Schäfer, A.I.; Richards, B.S. Photocatalytic degradation of organic dye via atomic layer deposited TiO2 on ceramic membranes in single-pass flow-through operation. J. Membr. Sci. 2020, 604, 118015. [Google Scholar] [CrossRef]
  36. Moulis, F.; Krýsa, J. Photocatalytic degradation of acetone and methanol in a flow-through photoreactor with immobilized TiO2. Res. Chem. Intermed. 2015, 41, 9233–9242. [Google Scholar] [CrossRef]
  37. Presumido, P.H.; dos Santos, L.F.; Neuparth, T.; Santos, M.M.; Feliciano, M.; Primo, A.; Garcia, H.; Ðolić, M.B.; Vilar, V.J.P. A novel ceramic tubular membrane coated with a continuous graphene-TiO2 nanocomposite thin-film for CECs mitigation. Chem. Eng. J. 2022, 430, 132639–132652. [Google Scholar] [CrossRef]
  38. Lofti, S.; Fischer, K.; Schulze, A.; Schäfer, A.I. Photocatalytic degradation of steroid hormone micropollutants by TiO2-coated polyethersulfone membranes in a continuous flow-through process. Nat. Nanotechnol. 2022, 17, 417–423. [Google Scholar] [CrossRef]
  39. Fernandez-Machado, N.R.C.; Santana, V.S. Influence of thermal treatment on the structure and photocatalytic activity of TiO2 P25. Catal. Today 2005, 107, 595–601. [Google Scholar] [CrossRef]
  40. Deiana, C.; Fois, E.; Coluccia, S.; Martra, G. Surface structure of TiO2 P25 nanoparticles: Infrared study of hydroxy groups on coordinative defect sites. J. Phys. Chem. C 2010, 114, 21531–21538. [Google Scholar] [CrossRef]
  41. Kumakiri, I.; Murasaki, K.; Yamada, S.; Abdul Rahim, A.N.B.C.; Ishii, H. A greener procedure to prepare TiO2 membranes for photocatalytic water treatment applications. J. Membr. Sci. Res. 2022, 8, 549416–549422. [Google Scholar] [CrossRef]
  42. Wodka, D.; Bielanska, E.; Socha, R.P.; Elzbieciak-Wodka, M.; Gurgul, J.; Nowak, P.; Warszyński, P.; Kumakiri, I. Photocatalytic activity of titanium dioxide modified by silver nanoparticles. ACS App. Mater. Interfaces 2010, 2, 1945–1953. [Google Scholar] [CrossRef] [PubMed]
  43. Che Abdul Rahim, A.N.; Yamada, S.; Bonkohara, H.; Mestre, S.; Imai, T.; Hung, Y.-T.; Kumakiri, I. Influence of salts on the photocatalytic degradation of formic acid in wastewater. Int. J. Environ. Res. Public Health 2022, 19, 15736. [Google Scholar] [CrossRef]
  44. Jiang, X.; Manawan, M.; Feng, T.; Qian, R.; Zhao, T.; Zhou, G.; Kong, F.; Wang, Q.; Dai, S.; Pan, J.H. Anatase and rutile in evonik aeroxide P25: Heterojunctioned or individual nanoparticles? Catal. Today 2018, 300, 12–17. [Google Scholar] [CrossRef]
  45. Senthilnathan, J.; Philip, L. Photocatalytic degradation of lindane under UV and visible light using N-doped TiO2. Chem. Eng. J. 2010, 161, 83–92. [Google Scholar] [CrossRef]
  46. Zhang, J.; Tao, H.; Wu, S.; Yang, J.; Zhu, M. Enhanced durability of nitric oxide removal on TiO2 (P25) under visible light: Enabled by the direct Z-scheme mechanism and enhanced structure defects through coupling with C3N5. Appl. Catal. B Environ. 2021, 296, 120372–120381. [Google Scholar] [CrossRef]
  47. Barquín, C.; Rivero, M.J.; Ortiz, I. Shedding light on the performance of magnetically recoverable TiO2/Fe3O4/rGO-5 photocatalyst. Degradation of S-metolachlor as case study. Chemosphere 2022, 307, 135991–135999. [Google Scholar] [CrossRef]
  48. Doustkhah, E.; Assadi, M.H.N.; Komaguchi, K.; Tsunoji, N.; Esmat, M.; Fukata, N.; Tomita, O.; Abe, R.; Ohtani, B.; Ide, Y. In situ blue titania via band shapes engineering for exceptional solar H2 production in rutile TiO2. Appl. Catal. B-Environ. 2021, 297, 120380–120390. [Google Scholar] [CrossRef]
  49. Sainz, M.A.; Serrano, F.J.; Amigo, J.M.; Bastida, J.; Caballero, A. XRD microstructural analysis of mullites obtained from kaolite-alumina mixtures. J. Eur. Ceram. Soc. 2000, 20, 403–412. [Google Scholar] [CrossRef]
  50. Orooji, Y.; Ghasali, E.; Moradi, M.; Derakhshandeh, M.R.; Alizadeh, M.; Asl, M.S.; Ebadzadeh, T. Preparation of mullite-TiB2-CNTs hybrid composite through spark plasma sintering. Ceram. Int. 2019, 45, 16288–16296. [Google Scholar] [CrossRef]
  51. Arroyo, R.; Cordoba, G.; Padilla, J.; Lara, V. Influence of manganese ions on the anatase–rutile phase transition of TiO2 prepared by the sol–gel process. Mater. Lett. 2002, 54, 397–402. [Google Scholar] [CrossRef]
  52. Ma, N.; Zhang, Y.; Quan, X.; Fan, X.; Zhao, H. Performing a microfiltration integrated with photocatalysis using an Ag-TiO2/HAP/Al2O3 composite membrane for water treatment: Evaluating effectiveness for humic acid removal and anti-fouling properties. Water Res. 2010, 44, 6104–6114. [Google Scholar] [CrossRef] [PubMed]
  53. Workneh, S.; Shukla, A. Synthesis of sodalite octahydrate zeolite-clay composite membrane and its use in separation of SDS. J. Membr. Sci. 2008, 309, 189–195. [Google Scholar] [CrossRef]
  54. Zhang, H.; Quan, X.; Chen, S.; Zhao, H.; Zhao, Y. The removal of sodium dodecylbenzene sulfonate surfactant from water using silica/titania nanorods/nanotubes composite membrane with photocatalytic capability. Appl. Surf. Sci. 2006, 252, 8598–8604. [Google Scholar] [CrossRef]
  55. Shi, Y.; Yang, D.; Li, Y.; Qu, J.; Yu, Z.Z. Fabrication of PAN@TiO2/Ag nanofibrous membrane with high visible light response and satisfactory recyclability for dye photocatalytic degradation. Appl. Surf. Sci. 2017, 426, 622–629. [Google Scholar] [CrossRef]
  56. Diban, N.; Pacuła, A.; Kumakiri, I.; Barquín, C.; Rivero, M.J.; Urtiaga, A.; Ortiz, I. TiO2–Zeolite metal composites for photocatalytic degradation of organic pollutants in water. Catalysts 2021, 11, 1367. [Google Scholar] [CrossRef]
  57. Liu, L.; Liu, Z.; Bai, H.; Sun, D.D. Concurrent filtration and solar photocatalytic disinfection/degradation using high-performance Ag/TiO2 nanofiber membrane. Water Res. 2012, 49, 1101–1112. [Google Scholar] [CrossRef]
  58. Goei, R.; Lim, T.T. Ag-decorated TiO2 photocatalytic membrane with hierarchical architecture: Photocatalytic and anti-bacterial activities. Water Res. 2014, 59, 207–218. [Google Scholar] [CrossRef]
  59. Li, W.; Li, B.; Meng, M.; Cui, Y.; Wu, Y.; Zhang, Y.; Dong, H.; Feng, Y. Bimetallic Au/Ag decorated TiO2 nanocomposite membrane for enhanced photocatalytic degradation of tetracycline and bactericidal efficiency. App. Surf. Sci. 2019, 487, 1008–1017. [Google Scholar] [CrossRef]
  60. Giusu, D.; Ampelli, C.; Genovese, C.; Perathoner, S.; Centi, G. A novel gas flow-through photocatalytic reactor based on copper-functionalized nanomembranes for the photoreduction of CO2 to C1-C2 carboxylic acids and C1-C3 alcohols. Chem. Eng. J. 2021, 408, 127250. [Google Scholar] [CrossRef]
  61. Bian, H.; Wang, Y.; Yuan, B.; Cui, J.; Shu, X.; Wu, Y.; Zhang, X.; Adeloju, S. Flow-through TiO2 nanotube arrays: A modified support with homogeneous distribution of Ag nanoparticles and their photocatalytic activities. New J. Chem. 2013, 37, 752–760. [Google Scholar] [CrossRef]
  62. Chen, H.S.; Chen, P.H.; Huang, S.H.; Perng, T.P. Toward highly efficient photocatalysis: A flow-through Pt@TiO2@AAO membrane nanoreactor prepared by atomic layer deposition. Chem. Commun. 2014, 50, 4379–4382. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Permeability system consisting on: (a) membrane, (b) solution reservoir, (c) weight balance, (d) vacuum trap and (e) vacuum pump.
Figure 1. Permeability system consisting on: (a) membrane, (b) solution reservoir, (c) weight balance, (d) vacuum trap and (e) vacuum pump.
Membranes 13 00448 g001
Figure 2. Schematics of the photocatalytic system.
Figure 2. Schematics of the photocatalytic system.
Membranes 13 00448 g002
Figure 3. TiO2 membrane morphology: (a) surface view, (b) cross-sectional view.
Figure 3. TiO2 membrane morphology: (a) surface view, (b) cross-sectional view.
Membranes 13 00448 g003
Figure 4. XRD patterns: (a) P25 powder, (b) mullite support, (c) TiO2 membrane, (d) TiO2/Ag membrane.
Figure 4. XRD patterns: (a) P25 powder, (b) mullite support, (c) TiO2 membrane, (d) TiO2/Ag membrane.
Membranes 13 00448 g004
Figure 5. Photocatalytic degradation data of DBS and DCA with TiO2 and TiO2/Ag submerged membranes without permeation. Dotted lines represent the fitted curves to pseudo-first-order kinetic model. P: photocatalytic experiments.
Figure 5. Photocatalytic degradation data of DBS and DCA with TiO2 and TiO2/Ag submerged membranes without permeation. Dotted lines represent the fitted curves to pseudo-first-order kinetic model. P: photocatalytic experiments.
Membranes 13 00448 g005
Figure 6. Comparison of DCA degradation in photocatalytic submerged membranes (P: photocatalytic experiments) and coupled photocatalysis and flow-through permeation experiments (PF: photocatalytic and flow-through). Total flux, in grey lines, of flow-through permeation experiments at a transmembrane pressure of 1 bar. (a) TiO2 membrane and (b) TiO2/Ag membrane.
Figure 6. Comparison of DCA degradation in photocatalytic submerged membranes (P: photocatalytic experiments) and coupled photocatalysis and flow-through permeation experiments (PF: photocatalytic and flow-through). Total flux, in grey lines, of flow-through permeation experiments at a transmembrane pressure of 1 bar. (a) TiO2 membrane and (b) TiO2/Ag membrane.
Membranes 13 00448 g006
Table 1. Comparison of the photocatalytic performance of TiO2-based catalysts: suspended, P: submerged photocatalytic membrane, PF: photocatalytic membrane and flow-through permeation experiments for DCA degradation.
Table 1. Comparison of the photocatalytic performance of TiO2-based catalysts: suspended, P: submerged photocatalytic membrane, PF: photocatalytic membrane and flow-through permeation experiments for DCA degradation.
Photocatalytic
Experiment
[TiO2]
(g L−1)
r0
(mg L−1 min−1)
Performance Factor
(mg L−1 min−1 gcat−1)
TiO2-P25 in suspension0.300.79 ± (2.79 × 10−2)2.63 ± 0.11
P-TiO20.020.06 ± (1.16 × 10−3)3.01 ± 0.12
PF-TiO20.020.12 ± (4.13 × 10−3)5.80 ± 0.24
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Barquín, C.; Vital-Grappin, A.; Kumakiri, I.; Diban, N.; Rivero, M.J.; Urtiaga, A.; Ortiz, I. Performance of TiO2-Based Tubular Membranes in the Photocatalytic Degradation of Organic Compounds. Membranes 2023, 13, 448. https://doi.org/10.3390/membranes13040448

AMA Style

Barquín C, Vital-Grappin A, Kumakiri I, Diban N, Rivero MJ, Urtiaga A, Ortiz I. Performance of TiO2-Based Tubular Membranes in the Photocatalytic Degradation of Organic Compounds. Membranes. 2023; 13(4):448. https://doi.org/10.3390/membranes13040448

Chicago/Turabian Style

Barquín, Carmen, Aranza Vital-Grappin, Izumi Kumakiri, Nazely Diban, Maria J. Rivero, Ane Urtiaga, and Inmaculada Ortiz. 2023. "Performance of TiO2-Based Tubular Membranes in the Photocatalytic Degradation of Organic Compounds" Membranes 13, no. 4: 448. https://doi.org/10.3390/membranes13040448

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