Next Article in Journal
Factors Influencing the Roles of Environmental Non-Governmental Organizations (ENGOs) on Environmental Bargaining in Yunnan, China
Next Article in Special Issue
Water Shortage Simulation Using a System Dynamics Approach: A Case Study of the Rafsanjan City
Previous Article in Journal
Transboundary Waters and Their Status in Today’s Water-Scarce World
Previous Article in Special Issue
Treatment of High Nutrient-Loaded Wastewater in a Constructed Floating Wetland with Different Configurations: Role of Lantana Biochar Addition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Challenges and Emerging Trends in Advanced Oxidation Technologies and Integration of Advanced Oxidation Processes with Biological Processes for Wastewater Treatment

by
Ginni Gopalakrishnan
1,
Rajesh Banu Jeyakumar
2 and
Adishkumar Somanathan
3,*
1
Department of Civil Engineering, Francis Xavier Engineering College, Tirunelveli 627003, Tamil Nadu, India
2
Department of Biotechnology, Central University of Tamil Nadu, Neelakudi, Thiruvarur 610005, Tamil Nadu, India
3
Department of Civil Engineering, Anna University Regional Campus, Tirunelveli 627007, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(5), 4235; https://doi.org/10.3390/su15054235
Submission received: 17 January 2023 / Revised: 14 February 2023 / Accepted: 22 February 2023 / Published: 27 February 2023

Abstract

:
One of the biggest problems the world is currently experiencing is wastewater treatment. Numerous pollutants are released into water bodies by enormous amounts of effluents from varied sources. This paper provides a concise summary of the state of the art in AOPs, biological treatment, and their efficient application for the degradation of the numerous recalcitrant organic contaminants found in wastewater. The employment of a photoreactor is necessary for the efficient execution of the majority of photocatalytic processes. This review examines the effectiveness of several reactor configurations with varying geometries. Integrating different AOPs and AOPs with biological approaches for degrading pollutants in wastewater was also discussed. It is important to emphasize that an integrated AOP/biological system produces by-products that are not hazardous, uses little energy, and completely degrades pollutants. The review also outlines the challenges and issues of wastewater treatment for an environmentally and economically feasible process.

1. Introduction

One of the essential needs of all living things, including humans, is safe and uncontaminated drinking water. However, these days, finding it has become a big issue. The water used by humans for most activities produces wastewater. Wastewater accounts for over 80% of the water provided for domestic use [1]. About 5–10 billion tons of industrial waste are generated each year and discharged untreated into the environment [2]. If released directly into the environment without treatment, the wastewater generated from various industries can cause significant environmental problems [3]. Therefore, removing emerging pollutants is essential to making the water safe for drinking, and the release of wastewater should not cause any adverse effect on the environment. Moreover, recycling and reusing treated wastewater is necessary to reduce freshwater consumption. There are physical, chemical, and biological treatment methods; these are the most commonly employed wastewater treatment technologies. In the case of physical treatment, the removal of organics from the wastewater is challenging. The existing biological and chemical treatment methods include ozonation, electrochemical treatment, photocatalysis, aerobic treatment, activated sludge, anaerobic treatment, coagulation–flocculation treatment, etc. [4]. Aerobic biological activities require a lot of energy and generate a significant amount of biomass. Anaerobic biological processes are susceptible to shock loading, which requires further treatment of the wastewater produced before the final release. Although biological processes are efficient and cost-effective, they primarily need a vast area and have a significant energy demand (for aeration) and a substantial quantity of produced sludge [5]. Chemical treatments can quickly oxidize and totally breakdown the organic contaminants, making them an effective wastewater treatment approach [6]. As a result, the development of advanced oxidation processes (AOPs), which have an excellent potential for ultimately destroying a variety of resistant pollutants, has shown to be a promising alternative [7].
It involves the generation of powerful oxidizing radical groups, such as hydroxyl radicals, which function as oxidizing agents and mineralize organic chemical substances into CO2 and H2O [7]. This is possible because hydroxyl radicals (OH), after fluorine radicals, have the highest oxidation potential (E0: 2.8 eV vs. normal hydrogen electrode (NHE)). The most potent oxidant, fluorine, which has an oxidation potential of 3.06 V, cannot be utilized to treat wastewater due to its toxicity [8]. Ozone is an extremely potent oxidant with a redox potential of 2.08 eV that may directly oxidize microorganisms and a variety of organic substances. It can also cause secondary reactions by generating hydroxyl radicals that can subsequently react with the microcontaminants. However, this process has some drawbacks, including the poor solubility of ozone in water, the high energy requirement, and the generation of reaction products that could be even more hazardous than the parent compounds [9]. Other frequently found oxidants, such as hypochlorite and hydrogen peroxide, are less reactive because they have paired electrons in their chemical structures. Hydrogen peroxide (H2O2) is an environmentally friendly oxidant with high oxidation potentials across the entire pH range. These properties give it exceptional bleaching and antiseptic capabilities that have led to many applications in wastewater treatment industries [10] (Tong et al., 2022). Thus the oxidation potential of an oxidant is used to measure the capacity of an oxidant to initiate chemical processes [11]. Table 1 shows the oxidation potential of various oxidative species [12].
AOPs such as photocatalysis and photo-Fenton have been widely considered to be very effective in removing persistent organic pollutants. It has been proved that AOPs can be utilized as a pretreatment to convert contaminants into shorter-chain compounds [13]. Mineralization by AOPs becomes exceedingly expensive because the oxidation intermediates generated during chemical oxidation become more difficult to completely degrade. The hydroxyl radicals, however, transform the resistant substances into compounds that are more readily biodegradable, which can then be further mineralized by microorganisms [14]. A single advanced oxidation processes such as the solar photo-Fenton and solar photocatalytic process have inherent challenges and limitations with respect to efficiency and economics because the solar photo-Fenton process is pH dependent and the quantum yield is less in the solar photocatalytic process. Considering the above drawbacks, there is a need for the combination of the two processes rather than the individual process and for the optimization of the combined treatment system to achieve the required target within a short reaction time. However, in general, AOPs are expensive, and one of the shortcomings that prevent the wide application of this treatment process in developing countries lies in its high operational cost due to its high energy consumption and chemical reagents. In order to reduce the energy cost, renewable solar energy is used for the treatment process. Combining several AOPs can greatly increase the oxidation efficiency of the pollutants compared to an individual treatment method because of the synergistic effect of different compounds [15]. To improve the overall treatment efficiency of the process and to reduce the chemical cost, the advanced oxidation processes are coupled with the biological treatment process [16].
This review paper briefly summarizes the different types of AOPs, the combination of two or more AOPs, and the integration of AOPs with the biological processes for the removal of toxic contaminants from wastewater. Moreover, this review outlines the challenges and issues of wastewater treatment technologies and the economic aspects of the treatment processes.

2. Mechanism of Advanced Oxidation Processes

There are two types of AOPs: homogeneous processes and heterogeneous processes. The homogeneous processes in the AOP are defined by the chemical alterations resulting solely from the interactions between the chemical reagents and the target substances [17]. In heterogeneous processes, the reactants and products adsorb and desorb on the catalyst’s active sites. As the reaction occurs, the desorption of products and the absorption of new species takes place on the active sites; thereby, the efficiency is affected by the surface characteristics and the pore structure of the catalyst [18]. The different methods for the production of hydroxyl radicals take place via several combinations, such as the Fenton process (Fe2+/H2O2), the solar photo-Fenton process (solar/Fe2+/H2O2), solar photocatalysis (solar/TiO2/H2O2), peroxone (Ozone/H2O2), the combination of peroxone with ultraviolet (Ozone/H2O2), Ozone/UV, and photolysis (H2O2/UV). The oxidation by-product formation mechanisms in the various AOPs are discussed.

2.1. Solar Photo-Fenton Process

Among the AOPs, the solar photo-Fenton process is particularly interesting for treating a large variety of hazardous pollutants and is one of the most environmentally benign treatment systems [19]. Due to the potential utilization of sunlight, this technique has drawn more attention. As a result, it may be considered to be an affordable AOP as sunlight is always available in a tropical country such as India [20,21]. Figure 1 shows the schematic diagram of the SPF.
As shown in Equation (1), catalysis by ferrous iron breaks down hydrogen peroxide in an acidic medium to produce highly reactive species such as OH radicals without requiring high pressure or temperature. When solar irradiation is used, Fe3+ hydroxy complexes (Fe(OH)2+) and Fe3+ organic complexes (FeOOCR2+) can perform the photo-reduction in Equation (2) and the photodecarboxylation in Equation (3), enabling the regeneration of iron and generating extra OH radicals [22]. The hydroxyl radicals break down the complex organic structures into simpler compounds, and the partial oxidation of the non-biodegradable organics contributes to the biodegradability of the wastewater.
Fe2+ + H2O2 → Fe3+ + OH + HO
[Fe(OH)]2+ + hν → Fe2+ + OH + H+
Fe(OOCR)2+ + hν → Fe2+ + CO2 + R
In the solar photo-Fenton process, the significant characteristic of iron is that it undergoes cyclic oxidation and reduction. Table 2 shows the treatment of wastewater by various solar photo-Fenton processes.
Xu et al. (2007) [23] used the solar photo-Fenton technique to treat a pulp and paper mill’s bleaching wastewater. According to the reports, the TOC removal was usually faster for 15 min using a cost-effective source of solar light irradiation. After that, there was a slow increase for 3 h. Sheik et al. (2008) [24] explored the photocatalytic oxidation of organic contaminants such as p-nitroaniline, p-aminophenols, and acetanilide. It was found that the organic compounds were entirely oxidized and degraded into CO2 and H2O. Karimi et al. (2011) [25] focused on the discoloration of wastewater from a paper mill by the photo-Fenton process. The research revealed that after 15 min of photo-Fenton treatment, the soda effluent had dropped nearly 65% of its original color.
Lucas et al. (2012) [26] reported experimentally that about 90% of DOC degradation was attained with 5 mg/L of Fe2+ and 50 mM of H2O2 for the treatment of pulp and paper mill wastewater. The same DOC degradation can be achieved by solar photo-Fenton using less H2O2 and a shorter time. Bernabeu et al. (2012) [27] employed the photo-Fenton process to degrade emerging pollutants. According to the findings, emerging contaminants at high concentrations are removed at an acidic pH rather than a neutral pH. Trovo et al. (2012) [28] evaluated the photodegradation of paracetamol using ferrous sulphate and potassium ferrioxalate under artificial solar light. In contrast to FeOx, FeSO4 decreased Fe3+ to Fe2+, enhancing the decomposition by generating higher concentrations of hydroxylated intermediates. Garcia and Buitron (2013) [22] proposed the photo-Fenton process and obtained 97% decolorization of each azo dye and the development of chemometric tools, a strategy that reduced the peroxide from 33% to 65%. Turbay et al. (2013) [29] examined the photo-Fenton reaction-based tetracycline antibiotic degradation under varying hydrogen peroxide doses. The outcome demonstrated the improved effectiveness of the treatment and a more effective use of the hydroxyl radicals generated in the reaction medium.
Rocha et al. (2014) [30] treated extracted petroleum wastewater utilizing sunlight as the irradiation source. After 7 h of exposure to sunlight, the experimental results showed a reduction of about 92.7% and 96.2% of the polycyclic aromatic hydrocarbons and the aromaticity, respectively. Rodriguez et al. (2014) [31] explored the potential of the photo-Fenton process to decolorize and mineralize simulated wastewaters from wool dyeing tanks. The results indicated a high level of efficiency in decreasing TOC, COD, and BOD5. Velegraki and Mantzavinos (2015) [32] studied the solar photo-Fenton technique for the treatment of winery effluent and achieved mineralization at a low catalyst dose (Fe2+ = 5 mg/L) and relied on an oxidant consumption of 500 mg/L. Chueca et al. (2015) [33] assessed the efficacy of a mild solar photo-Fenton system for disinfecting the actual effluents containing fecal bacteria. Despite the complexity and variety of the pollutants in the effluents, the tests conducted on the real effluents yielded highly encouraging results. Soares et al. (2015) [34] treated the textile dye wastewater using ferric organic ligands with the solar photo-Fenton process. It was reported that 87% of the mineralization was achieved at neutral pH values with biodegradability enhancement. Guzman et al. (2016) [35] assessed solar photo-Fenton treatment and various ozone-based processes for the pretreatment of synthetic samples of citrus wastewater. According to the findings, the photo-Fenton method removes COD and DOC at rates of 76.9% and 53.3%, respectively, which are higher than those of ozone-based systems.
Benitez and Penuela (2018) [36] examined the complete removal of benzophenone and the biodegradability increase using the solar photo-Fenton process under simulated solar irradiation. Costa et al. (2018) [37] treated biodiesel effluent, employing the solar photo-Fenton approach to decrease COD by adding oxalate at 50 kJ m−2 of accumulated UVA radiation. About 72% of COD and 76% of BOD removal were achieved using 1 mmol L−1 of ferrioxalate. Garcia et al. (2021) [38] investigated an advanced method for post-treating laundry effluent using the solar photo-Fenton process. It was found that a high removal efficiency was achieved, reaching 89% and 96%, respectively, in just 8 min. Arka et al. (2021) [39] investigated the solar photo-Fenton method for the treatment of institutional wastewater. According to the study, the solar UV/Fe2+/H2O2 process was very efficient in treating the wastewater from institutions, achieving a greater pollutant removal rate. The removal of six representative pharmaceuticals (PHCs) in raw hospital wastewater was investigated by Lumbaque et al. (2021) [40]. It was observed how the molar ratios (1:1 and 1:2) affected the breakdown of PHCs and the production of their transformation products during the treatments. The primary findings showed that PHC degradation is greatly favored in Fe3+:EDDS (1:2) due to the overall H2O2 consumption and the enhanced iron stability, with degradation efficiencies of above 77% in most PHCs.
Lin and Lin (2021) [41] used a solar photo-Fenton technique to remove cytostatic medicines. In comparison to three other Fe(III)-ligand complexes, the addition of EDDS caused the degradation of cyclophosphamide and 5-fluorouracil to occur more quickly. Gualda-Alonso et al. (2022) [42] carried out a continuous flow solar photo-Fenton process, employing a 100 m2 raceway pond reactor (RPR), for the elimination of emerging contaminants. The RPR was run in continuous flow mode with 0.1 mM FeSO4 and 1.47 mM H2O2 at an acidic pH and 1 h of hydraulic residence time. The liquid depth was fixed at 10 cm for winter and 18 cm for summer, removing >85% of the CEC in both cases. Pandey et al. (2023) [43] treated contaminated water by solar photo-Fenton using Fe-rich catalysts and obtained 82% COD removal.
The discussions mentioned above provide a general overview of the refractory compounds degraded by the photo-Fenton process and the effectiveness of the mineralization of these chemical pollutants under observed optimal conditions.
Table 2. Treatment of wastewaters by various solar photo-Fenton processes.
Table 2. Treatment of wastewaters by various solar photo-Fenton processes.
S. No.WastewaterProcessOperational ConditionsDegradation EfficiencyReference
1Pulp and paper millSolar photo-FentonFe(II) concentration from 31 to 310 mg L−1 (initial pH 3.0, 30 °C), initial H2O2 concentration from 0.5 to 3 DthTOC—82%
Time—120 min
Xu et al., 2007 [23]
2p-nitroanilineSolar photo-Fenton Sheik et al. (2008) [24]
3Soda pulping effluentFenton and photo-FentonFeSO4—1 mM
H2O2—570 μL
Color—65%Karimi et al. (2011) [25]
4Pulp and paper millSolar photo-Fenton5 mg/L of Fe2+ and H2O2 of 50 mM90% of DOC mineralizationLucas et al. (2012) [26]
5Azo dye, Acid Blue 161Solar photo-FentonH2O2/Fe2+ = 12, pH = 2.5 to 4.0Degradation 40 %Trovo et al., 2016 [28]
6Sulfonated azo dyesSolar photo-FentonFe2+ = 5 to 38 mg/L, H2O2 = 98 to 828 mg/LColor—97%Garcia and Buitron (2013) [22]
7Petroleum extractionPhoto-Fenton485.3 mmol/L of H2O2Hydrocarbons and aromaticity of about 92.7% and 96.2%Rocha et al. (2014) [30]
8Winery effluentPhoto-Fenton(Fe2+ = 5 mg/L)-Velegraki and Mantzavinos (2015) [32]
9Citrus wastewaterPhoto-FentonH2O2 = 1017 mg/L, pH = 7COD—77%Guzman et al., 2016 [35]
10Biodiesel effluentSolar photo-Fenton1 mmol L−1 of ferrioxalate72% of COD and 76% of BODCosta et al. (2018) [37]
11Laundry effluentSolar photo-FentonH2O2 (50–400 mg/L), Fe2+ (2.75–10 mg/L)SDS removal—96%Garcia et al. (2021) [38]
12Industrial effluentSolar photo-Fenton(H2O2)—0.25 to 1.25 g/L, Fe2+—0.005 to 0.12 g/L, pH (2 to 10), time (30 to 180 min)color (91%), turbidity (90%), and COD—86%)Arka et al. (2021) [39]
13Hospital wastewaterSolar photo-FentonFe3+:EDDS = 1:2,
H2O2 (230 mg L−1)
Degradation = 77%Lumbaque et al., 2021 [40]
14Cytostatic drugsSolar photo-FentonpH 3.0–8.5, 0.1 mM Fe(III)-EDDS, 1 mM H2O2 Lin and Lin, 2021 [41]
15Emerging contaminantsSolar photo-Fenton0.1 mM FeSO4 and 1.47 mM H2O2 at an acidic pH and 1 h of hydraulic residence time>85%Gualda-Alonso et al., 2022 [42]
16Paraquat (PQ)-contaminated waterPhoto-FentonFe-containing industrial wasteCOD—82%Pandey et al., 2023 [43]

2.2. Solar Photocatalytic Process

Solar photocatalysis has been widely used to degrade organic compounds. According to numerous studies, many organic contaminants have been totally oxidized in irradiated semiconductor solutions [44]. Photocatalysts remove pollutants from wastewater primarily through hydroxyl radical (OH) attacks by transforming them into harmless compounds such as water and CO2 [45]. The photocatalyst that has been researched the most is TiO2. However, its usage is restricted to ultraviolet light because of the wide energy band gap (3.2 eV) [46]. When exposed to ultraviolet light, a semiconductor catalyst such as TiO2 or another transition metal oxide is initiated by the photon absorption with enough energy to be equal to or greater than the catalyst’s bandgap energy, stimulating an electron to move from the valence band to the conduction band, creating a hole in the valence band while using just 4% of the solar radiation [47,48,49].
TiO2 photocatalysis uses UV-A radiation (λ < 387 nm) in the presence of oxygen and water to produce hydroxyl radicals for the excitation of the photocatalyst [50]. The overall rates of pollutant degradation are significantly influenced by the surface area of the TiO2 catalysts that contains active sites. Thus, increasing the TiO2-specific surface area effectively increases the pollutant adsorption capacity on the catalyst. As a result, a larger surface area with a more significant number of active sites will lead to a faster and more extensive reaction [51].
The activation Equation (4) can be written as
TiO2 + hν → h+ + e
It is possible to express the oxidation and reduction reactions as given in Equations (5) and (6)
h+ + OH → -OH
e + O2ads → O2ads
Another critical factor in the process of photocatalytic oxidation is the pH level. In alkaline media, the reaction of OH with holes on the TiO2 surface causes the formation of hydroxyl radicals. Thus, hydroxyl radicals accelerate photocatalytic oxidation at higher pH solutions and vice versa [52]. Figure 2 shows the schematic diagram of the SPC.
Adish and Kanmani (2010) [53] used a single baffle reactor to photo-catalyze H2O2 to treat phenolic wastewater. According to reports, the solar/Fe2+/H2O2 process degrades phenol two to three times more quickly than the solar/TiO2 process. Kumar et al. (2011) [54] studied the degradation of pulp and paper mill wastewater by a photocatalytic process. The findings showed that the increase in the BOD/COD ratio after photocatalytic oxidation was 0.09 for the primary wastewater with the removal of the COD (57.9%), BOD (42.9%), and color (89.2%). Dimitrakopoulou et al. (2012) [55] exploited eight different TiO2 catalysts for the degradation of amoxicillin, and it was reported that Degussa P25 was the most active of the catalysts, with 93% mineralization after 25 and 95 min of reaction time at the amoxicillin concentrations of 10 mg/L and 250 mg/L of titania. Singh et al. (2013) [52] examined the photocatalytic oxidation of the organic content in synthetic wastewater at the pH values 2, 4, 6, 8, and 10 and at normal pH. The outcomes demonstrated clearly that an 86% reduction in COD was achieved at a normal pH level (pH = 6.8), and it was reported that an acidic pH was unfavorable for the reduction of organic content. Nagpure et al. (2013) [56] studied the photocatalytic decomposition of paper mill wastewater with ultraviolet and catalysts such as TiO2. The degradation efficiency was observed as 91.34% using TiO2 as a photocatalyst, with an optimum catalyst dose of 80 mg/50 mL for an irradiation time of 8 h under continuous stirring with maximum intensity. Ruiz et al. (2013) [57] studied the photodegradation of trimethoprim by various nanoparticles deposited on TiO2. It was reported that the degradation of trimethoprim was only 50% of the organic matter mineralization with pure TiO2-P25.
Shao et al. (2013) [58] produced mesoporous TiO2 nanotubes with a large diameter for the photodecomposition of effluent from paper manufacturing. The result indicated that after 12 h of photodegradation, the chemical oxygen demand and chroma percent degradations of the paper manufacturing wastewater were approximately 73% and 99.5%, respectively. Metribuzin, a common herbicide, was transformed and mineralized in extensive detail by TiO2-driven photocatalysis under simulated sunlight, as described by Antonopoulou et al. (2014) [59]. The results pointed out that the complete transformation of metribuzin was attained within 40 min at a metribuzin concentration of 10 mg/L, a TiO2 concentration of 100 mg/L, and I = 750 W/m2, whereas 80% mineralization was achieved within 300 min of irradiation.
Thomas et al. (2014) [60] compared the efficiency of TiO2 and few-layer graphene nanocomposites for degrading rhodamine B under solar radiation. The titania incorporation on the few-layer graphene improved titania’s visible light photocatalytic activity, decreased electron–hole recombination, and increased electron–hole mobility. Fernandez et al. (2015) [50] investigated the photocatalytic treatment of polluted wastewater. The effects of many factors, including the type of microorganism, the water temperature, the dissolved oxygen concentration, and the water matrix composition were studied, and the results showed that this process has great potential for the chemical reduction and disinfection of pathogens. Murgolo et al. (2015) [61] investigated a novel photocatalyst based on nanosized TiO2 upon radiation by both UV and solar-simulated light, resulting in degradation efficiencies ranging from 9 to 87% and 9 to 96%, respectively.
Solano et al. (2018) [62] evaluated the removal efficiency of hexavalent chromium (Cr6+) and divalent zinc (Zn2+) using heterogeneous photocatalysis with TiO2. The result revealed that this process was efficient for the total removal of a Cr6+ concentration of 5 mg/L and was less effective for higher concentrations (15, 25 ppm); however, this technology was ineffective for Zn2+ removal. Yang and Yang (2018) [63] addressed the decomposition of rhodamine B using photocatalyst TiO2. The result indicated that the catalyst reuse was found to be efficient up to 8 times with the same degradation efficiency. Aljouboury et al., 2021 [64] studied the ideal conditions for natural petroleum wastewater, utilizing TiO2/ZnO/Fenton/solar and TiO2/ZnO/air/solar. According to the reports, the TiO2/ZnO/air/solar process achieved the highest treatment effectiveness with the COD removal of 74% and the TOC removal of 99% under ideal circumstances.
Kader et al. (2022) [65] investigated the synthesis of TiO2-based photocatalysts, the production of TiO2-based immobilized borosilicate glass reactors, and the use of the reactors to the treatment of methyl orange dye under UV light. The results revealed that the photocatalyst effectiveness dropped with the increase in pH and the initial dye concentrations. Total organic halogen was utilized as an analytical tool by Abusallout and Hua (2022) [66] to assess the effectiveness of solar TiO2 photocatalysis in dehalogenating disinfection by-products in the water. Compared to pH 5, the TOX dehalogenation was improved at pH 9, and the addition of hydrogen peroxide slightly improved the TOX elimination. The findings demonstrated that compared to the anatase and rutile TiO2 particles, the mixed-phase TiO2 (Aeroxide P25) was significantly more efficient at removing TOX.
To immobilize TiO2 for the removal of paraben from wastewater, Martins et al., 2022 [67] tested the suitability of utilizing polymeric supports. According to the results, polydimethylsiloxane is a suitable material to support TiO2 for wastewater treatment by solar photocatalytic oxidation. Using two nanocomposites (TiO2 and TiO2 (Ag) doped), Behera et al. (2022) [68] compared the photocatalytic activity on the breakdown of phenol from water. The nanocomposites were created using the UV photo-reduction method with silver loadings of 0.25, 0.5, 0.75, and 1% (w/w). At pH 7, about 98% of the phenol was degraded in 180 min with 0.5 g L−1 photo-catalyst TiO2 (Ag-1.0) EY. Rueda-Marquez et al., 2020 [69] examined the solar photocatalytic process using immobilized TiO2 for treating urban wastewater and obtained a maximum removal of greater than 40% of the pharmaceutically active compounds. Gai et al., 2021 [70] synthesized potassium and iodide TiO2 and observed the enhancement of photocatalytic activity under simulated sunlight. Giang et al., 2023 [71] synthesized sulfur-doped TiO2 for the degradation of crystal violet. The findings revealed the potential uses of the material in environmental treatments. Rapti et al., 2022 [72] treated hospital wastewater using TiO2 with the catalyst loading between 150 mg/L and 200 mg/L and attained the degradation efficiency of 73%.
Although numerous research studies on the photocatalytic degradation of wastewater pollutants using TiO2 as a semiconductor have been reported, its industrial application has some constraints, including the recombination of the photogenerated electron–hole pair, the lack of and the inability to produce a catalyst with high photon efficiency and low cost that can absorb a broader range of the solar spectrum, and the surface structure [47]. Moreover, the aggregation of the TiO2 catalyst reduces the surface area and hence the effectiveness of the catalyst [73]. Because of the large amounts of chemicals required for proper performance, photocatalysis is only sometimes considered a viable treatment to be used alone; however, their combination with biological processes may improve the overall process efficiency, increasing their viability. Table 3 shows the treatment of wastewater by various solar photocatalytic processes.

3. Biological Treatment Processes

Biological treatment is an essential component of any wastewater treatment plant that handles wastewater with soluble organic pollutants from municipalities, industries, or a mix of the two. Aerobic microorganisms require oxygen and organic substances to function and develop. The organic matter in the wastewater provides nutrients, and the air is typically pumped into the treatment tank to supply oxygen. Aerobic digestion produces carbon dioxide, energy, and metabolized solids, which precipitate as its by-products. Bacteria and other microorganisms that survive without oxygen are known as anaerobic microbes. Compared to aerobic bacteria, these microbes degrade organic pollutants more slowly. Anaerobic microbes produce methane, carbon dioxide, and more anaerobic bacteria.
Vohra et al. (2005) [74] examined the ability of white rot fungi to degrade lignin. According to the studies, the wastewater treatment processes eliminated 71% of the lignin and 48% of the COD. Chelliapan et al. (2012) [75] studied the degradation of the paper mill wastewater by an anaerobic process. The results showed that 98% COD removal was attained with an organic loading rate of 1.560 kg/m3.d. Pleurotus sajor caju and P. ostreatus were used by Belem et al. (2008) [76] to remediate the pulp and paper mill waste. After 14 days, no changes were seen for any of the effluent compositions or inoculation species, and the biological treatment reduced COD by 65–67%. For the treatment of tannery effluent, Durai et al. (2011) [77] investigated the effectiveness of a bench-size aerobic sequencing batch reactor. The optimized factors were the initial substrate concentration, the hydraulic retention period, and the organic loading rate. Based on the findings, it was determined that the highest removals of COD and color were 79% and 51%, respectively.
Ventura et al. (2011) [78] treated real domestic wastewater with a pilot-scale anaerobic-anoxic-oxic process, with a submerged membrane in the oxic tank and a thermophilic aerobic digestion reactor. The removal of MLSS, TCOD, BOD, TN, TP, and E. coli by the A2O-MBR-TAD process was reported to be nearly 99%, 96%, 70%, 83%, and 99%, respectively. Sludge generation was reported to be reduced to almost zero by adjusting the flow rate of the MBR’s waste stream. In a sequencing batch reactor with bacterial consortia (Klebsiella sp., Alcaligens sp. and Cronobacter sp.), Kumar et al. (2014) [79] explored the treatment of wastewater from pulp and paper mills. According to the experimental findings, the demands for chemical and biological oxygen were reduced by 72.3% and 91.1%, respectively. Chen et al. (2018) [80] investigated the wastewater treatment of the pulping industry using fly ash in SBR. According to the results, the addition of fly ash in the bioreactors enhanced the settling and flocculation affinity of the activated sludge without affecting the efficiency and stability of the biological process.
Rhodococcus opacus was studied by Goswami et al. (2018) [81] in a ternary substrate system to determine whether it could simultaneously biodegrade the polycyclic aromatic hydrocarbons (PAHs) and accumulate lipid. It was found that the individual effect of the PAH content was more important, and the initial concentration and aromatic complexity of the PAHs were the key factors affecting the PAH biodegradation effectiveness in the combination. Su et al., 2021 [82] explored the efficiency of the combined Fenton process with an aerated biological filter for the treatment of organic compounds collected from the secondary treatment of an industrial park wastewater treatment plant. It was shown that the combined process resulted in a synergistic effect with more than 70% mineralization at reduced dosages of Fenton’s reagent.
Although these methods are economically viable, the disadvantage of aerobic digestion is that it produces significant amounts of biosolids or sludge that must be disposed of. When the nutrient-rich sludge is improperly released into ponds or rivers, it can lead to eutrophication, which destroys fish and other aquatic species. Methanogenic archea are bacteria with a slow growth rate, which is a drawback of anaerobic processes. If the influent is strong in sulfur, this process emits volatile organic acids that cause odors. Removing the ammonia-nitrogen is ineffective, and the pH must be regulated together with the monitoring of the volatile fatty acids. Additionally, biological systems seldom achieve their maximum efficiency due to the intermittent load and the lack of an efficient monitoring parameter for the live biomass. Refractory substances are not effectively removed, even if the traditional biological processes are often effective at degrading the contaminants found in wastewater [83]. Therefore, a low-cost alternative treatment is required to eliminate the persistent organic compounds in wastewater.

4. Combination of Different Advanced Oxidation Processes

As wastewater is reused, the research has recently concentrated on integrated treatments for the breakdown of complex organic contaminants. Depending on the chemical composition of the wastewater, different AOPs can be used alone or in combination because various reagent systems are used to produce the radical species. Combining the two other AOP processes is an attractive option, and the different combinations of AOPs result in synergistic effects that enhance the removal efficiency. Additionally, when several advanced oxidation processes are utilized in combination systems, the amount of reagent consumed is reduced, which results in a decrease in the production of inhibitory by-products. Table 4 shows the treatment of wastewater by integrated AOPs.
The following is some of the literature based on the combined advanced oxidation processes.
Berberidou et al. (2007) [84] used ultrasound irradiation in combination with heterogeneous (TiO2) and homogeneous photocatalysis to investigate the decomposition of Malachite green in the water. Due to increased reactive radical formation and a potential increase in the catalyst’s active surface area caused by ultrasound, it was concluded that TiO2 sono-photocatalysis was more rapid than the individual process. Zhang and Pagilla (2010) [85] addressed the treatment of pesticide wastewater by combining nanofiltration (NF) and photo-Fenton’s reagent and evaluated the economic analysis. It was reported that enhancing NF could minimize the treatment cost. Na et al. (2012) [86] examined the degradation of diethylpthalate by combining sono-photolysis and sono-photocatalysis to confirm the synergistic degradation of diethylpthalate. With sono-photolysis and sono-photocatalysis, diethylpthalate was shown to degrade and mineralize significantly with synergistic effects of 1.29 and 1.95, respectively. Jadhav et al. (2013) [87] utilized hydrodynamic cavitation (HC) and a combination of HC and H2O2 to investigate the degradation of imidacloprid in aqueous solution. The rate of degradation of imidacloprid was significantly accelerated by HC + H2O2, resulting in complete degradation in 45 min with a synergistic coefficient of 22.79. Basturk and Karatas (2014) [88] studied the decolorization of CI Reactive blue 181 by the sono-Fenton and Fenton processes. It was reported that the color removal of 88% was attained for the Fenton process under optimum conditions, whereas 93.5% color removal was attained for the sono-Fenton process in optimum conditions. Thus, it was noticed that high decolorization was obtained by the sono-Fenton process.
Gogate and Patil (2014) [89] investigated the degradation of triazophos by hydrodynamic cavitation and advanced oxidation processes. Under optimum operational conditions, hydrodynamic cavitation (HC) alone was reported to degrade triazophos by about 50% and 80% by integrating HC and Fenton’s reagent. In contrast, the combination of HC and ozonation resulted in complete degradation. Qiu et al. (2014) [90] investigated the combined electrolysis and ozonation methods for treating wastewater contaminated with nitrophenol contaminants. The outcomes showed that the combined procedure had a significant synergy for deterioration. Giannakis et al. (2015) [91] examined the efficiency of the coupling of sonication and photo-Fenton for the inactivation of bacteria in the secondary treated effluent. The combined approach increased the effectiveness of removal and suggested feasible, affordable solutions for implementing this method. Zhan et al. (2019) [92] explored 3D electrolysis, ozonation, and the combined process (3D/O3) for pharmaceutical wastewater treatment. The findings revealed that ozonation and a 3D electrochemical method synergized the breakdown of hazardous organics in pharmaceutical effluent. Chanikya et al., 2020 [93] compared the efficacy of the combined sulphate radical-based electrochemical advanced oxidation process (EAOP) + electrocoagulation (EC) and EC+EAOP methods for the treatment of textile wastewater. The iron plate served as the cathode and the Pt/Ti plate as the anode. The results showed that a higher COD reduction of 93.5% was observed with EAOP, followed by EC with less specific energy consumption and sludge production.
Bui and Minh (2021) [94] investigated the treatment effectiveness of red wastewater by examining COD reduction with several types of advanced oxidation technologies and their combinations. The findings showed that Fenton/TiO2/O3/UV was the most efficient approach for treating red wastewater among the investigated procedures. Under the ideal operating parameters, it allowed a reduction of more than 99% of COD during 30 h of treatment. Titchou et al. (2021) [95] examined the combinations of AOPs. The findings showed a synergistic effect on the degradation of organic compounds by combining UV or ultrasonic radiation with Fenton sulfate-based oxidation and photocatalysis, resulting in complete removal efficiency. Dogan et al. (2021) [96] evaluated the combination of Fenton and enhanced photo-Fenton UF system performances using a membrane oxidation reactor (MOR). This study shows the MOR performance synergy allows effluents to be discharged directly into the sewage system after pH correction with complete treatment.
Dindas et al., 2020 [97] treated pharmaceutical wastewater with the combination of electrocoagulation, electro-Fenton, and photocatalytic oxidation. The result reported that the sequential process removed 64% of TOC and 70.2% of COD. Maifadi et al., 2022 [98] treated salon wastewater with the integrated AOP/membrane process and obtained TOC removal of 89.74%. An et al., 2022 [99] investigated the integration of AOP with microfiltration for treating semiconductor wastewater. The result showed the highest permeability and low energy consumption of the membrane used in the microfiltration. Doltade et al., 2022 [100] studied the integrated oxidation process for treating dye wastewater. The findings revealed that about 91% of COD removal was obtained when ozone and H2O2 were used as oxidizing agents.
The above kinds of literature provide the efficiencies of various combinations of AOPs. When used individually, the solar photo-Fenton and photocatalytic processes produce difficulties with restricted applicability and increased treatment costs. The two advanced oxidation processes are thus integrated to increase the degradation efficiency and generate various synergistic reactions by generating hydroxyl radicals under mild experimental conditions to widen the practical applications of industrial and environmental interest [89]. Therefore, integrating several modern oxidation systems is conceptually advantageous and enables the reduction in the scope of the individual processes without affecting the overall degradation rates, while also lowering investment and operating costs [101]. Figure 3 depicts a schematic illustration of the combined solar photo-Fenton and photocatalytic processes.
Table 4. Treatment of wastewater by integrated AOPs.
Table 4. Treatment of wastewater by integrated AOPs.
S. No.WastewaterProcessOperational ConditionsDegradation EfficiencyReference
1Malachite greenUltrasound irradiation+ TiO2+ solar photo-FentonUltrasound power—75–135 W
TiO2—0.1–0.5 g L−1, Fe3+—7–20 mg L−1
Complete malachite green degradationBerberidou et al., 2007 [84]
2Pesticide wastewaterNanofiltration and solar photo-FentonpH 3, H2O2 = 1:100 and H2O2:Fe(II) = 40:1Maximum removal efficiencyZhang and Padilla (2010) [85]
3Diethyl phthalateSonolytic, photolytic, and sonophotolytic processesUltrasound (283 kHz)Synergistic effects of 1.68 and 1.23Na et al. (2012) [86]
4ImidaclopridHydrodynamic cavitation (HC) + Fenton, HC + photo-Fenton, Hc + photolytic, HC + photocatalytic Imidacloprid:H2O2 as 1:40 and the molar ratio of FeSO4·7H2O:H2O2 as 1:40Synergetic index of 3.636 (HC) + Fenton and 2.912 HC + photo-Fenton Jadhav et al. (2013) [87]
5CI Reactive Blue 181Fenton and sono-Fenton processFenton process— [Fe2+] = 30 mg/L, [H2O2] = 50 mg/L and pH = 3
sono-Fenton [Fe2+] = 10 mg/L, [H2O2] = 40 mg/L and pH = 3
Color removal
Fenton—88%
Sono-Fenton—93.5%
Basturk and Karatas (2014) [88]
6TriazophosHydrodynamic cavitation and advanced oxidation processesInlet pressure (1–8 bar) and initial pH (2.5–8)80% degradationGogate and Patil (2014) [89]
7NitrophenolOzonation and electrolysisReaction time—60 minTOC—91% Qiu et al. (2014) [90]
8Synthetic wastewaterSonication, mild photo-Fenton treatmentUltrasound 20 W recirculating flow rate: 4.39 L/h
Iron 1 ppm treated volume: 500 mL
H2O2 10 ppm
Maximum E.coli inactivationGiannakis et al. (2015) [91]
9 Pharmaceutical3D electrolysis, ozonation-TOC abatement to ~71%Zhan et al. (2019) [92]
10PharmaceuticalElectrocoagulation + electro-Fenton, electrocoagulation + photocatalytic oxidation1 h, 5 mA/cm2 pulsed current density, Fe:H2O2—1:10 4 h PcO using 1.5 g/L TiO2 and 10 mM H2O264.0 % TOC, 70.2 % COD, 97.8 % BODDindas et al., 2020 [97]
11Red wastewateradvanced oxidation processes-COD > 99% Bui and Minh, 2021 [94]
12Organic pollutantsOzonation, Fenton, sulfate-based processes, photocatalysis, and sonolysis-Removal efficiency > 90%Titchou et al., 2021 [95]
13Pulp and paper mill wastewaterHybrid Fenton and photo-Fenton-Fenton—64.0–74.9%Dogan et al., 2021 [96]
14Salon wastewaterPeroxydisulfate-based AOPspH 2–6, reaction time—7 h TOC—89.74 %Maifadi et al., 2022 [98]
15Semiconductor wastewaterOzone + microfiltrationpH—6.8–8.0Maximum permeability, low energy consumptionAn et al., 2022 [99]
16Industrial wastewaterOzone + H2O2Treatment time—60 min, stirring at 240 RPMCOD—91%Doltade et al., 2022 [100]

5. Coupling of the Advanced Oxidation Process and the Biological Process

Recently, the combination of advanced oxidation and biological processes has been effectively used as a viable approach to treating recalcitrant wastewater. The implementation of biological treatment is both economical and environmentally friendly. However, it cannot sufficiently remove the organic materials in high concentration levels, and it usually requires a longer time to degrade the pollutant, and the efficiency of the treatment is affected by the formation of toxic by-products [102]. Consequently, chemical pretreatment is necessary to transform the toxic compounds into less harmful and more biodegradable molecules when direct biological treatment is not viable [103]. Thus, the AOP is used as a pretreatment to improve biodegradability and make the wastewater compatible with further biological treatment. Table 5 shows the treatment of wastewater by the combined AOPs and biological processes.
The following is some of the literature based on the coupled AOP–biological process.
By combining activated sludge and ozonation processes, Assalin et al. (2009) [4] investigated wastewater treatment from pulp and paper mills. The wastewater was acclimatized for 39 days for the growth of microorganisms using biomass from domestic sewage. The wastewater was subjected to ozonation at pH 8.3 and 10 for 60 min. From kraft E1 effluent at pH 10, the combined process eliminated 80% of COD, TOC, and color. The feasibility of completely removing the phenol was 75%. It was determined that it was only the use of integrated processes that resulted in less consumption of ozone and that they were, hence, economical. To decontaminate wastewater contaminated with commercial pesticides, Zapata et al. (2010) [104] suggested a design plan for a solar photo-Fenton/biological system that performed in tandem. The combined system had an efficiency of 84%, with 49% attributable to the biological stage and 35% to the photo-Fenton treatment.
Dhir et al. (2011) [105] examined the coupled biological and photochemical processes for treating pulp and paper mill wastewater. The results revealed that about 81% and 93% of degradation efficiency was obtained for C and E1 wastewater, respectively, for the combined AOP and biological process under operating conditions. Ruas et al. (2012) [106] explored the removal of refractory organic compounds from kraft bleaching wastewater using advanced oxidation techniques that combined hydrogen peroxide with UV radiation. A horizontal flow anaerobic immobilized biomass reactor was used for anaerobic pretreatment. The wastewater was treated in the reactor for 19 h of hydraulic retention time. The predicted removal efficiency for COD was 61%, TOC 69%, BOD5 90%, and AOX 55%. The H2O2/UV process as a post-treatment resulted in the removal efficiency of COD in the range from 0 to 11%, UV254 in the range from 16% to 35%, lignin from 0 to 29%, and AOX from 23 to 54%.
Botia et al. (2012) [107] investigated using a biological–photocatalytic coupled system to treat effluent from pulp bleaching. The biological pretreatment system was injected with the fungus Trametus pubescens immobilized on polyurethane foam. According to the findings, it was found that about 96% of the initial total organic carbon was degraded completely. To treat wastewater containing pesticides from the washing of phytopharmaceutical plastic containers, Moreira et al. (2012) [108] treated the wastewater containing pesticides using the combined biological treatment and AOPs. The biological treatment was conducted before the AOPs because of the high biodegradability of the material. This resulted in reductions in COD, DOC, and BOD5, and this process was followed by the photo-Fenton reaction, which resulted in 86% mineralization. The COD concentration was brought down below the detection threshold by this integrated treatment method.
A multistage treatment system was developed by Silva et al. (2013) [109] to treat landfill leachate using biological oxidation, the solar photo-Fenton process, and, again, the activated sludge biological oxidation. The first biological oxidation process eliminated 95% of the nitrogen and 39% of the DOC. The experiment without sludge removal resulted in a 30% increased H2O2 and UV energy consumption during the photo-Fenton reaction. The overall treatment process results in high efficiency in removing identified organic trace contaminants. Merayo et al. (2013) [110] treated the pulp and paper mill wastewater by ozonation and solar photocatalysis with a combination of biological processes. As a pre- or post-treatment for the biological process, it was revealed that the photocatalytic process did not significantly reduce COD; however, the use of ozonation significantly increased the COD removal to 90% for the combined treatment. Souza et al. (2013) [111] studied the treatment of winery wastewater utilizing biological oxidation and the solar photo-Fenton process combined with the biological approach. The results showed that it takes roughly 10 and 6 days, respectively, to attain the COD value utilizing simply the biological system or the combined system.
Marsolek et al. (2014) [112] investigated the AOP–biological process that was sequentially coupled for wastewater treatment. By adjusting the wavelength of light, the catalyst, and the treatment time in a photocatalytic reactor, four different effluents were produced from 2,4,5-trichlorophenol, revealing that photocatalytic pretreatment conditions can be controlled to attain a range of treatment objectives. Blanco et al. (2014) [113] studied textile wastewater degradation by integrating an aerobic sequencing batch reactor and photo-Fenton oxidation. It was found that 97% COD and 95% TOC reductions were achieved by the combined SBR and photo-Fenton process under the optimum conditions. After the combined treatment, it was determined that the resulting water was subjected to reverse osmosis to achieve the required water characteristics for 100% internal reuse.
Mendez et al. (2015) [114] explored heterogeneous photodegradation, Fenton, biological techniques, and their combinations for the degradation of phenol, formaldehyde, and phenol–formaldehyde mixtures from water. The results revealed that the combined process was found to be optimal for wastewater treatment. Amoxicillin and cloxacillin-containing antibiotic wastewater were subjected to combined photo-Fenton–SBR treatment in a study by Elmolla and Chaudhuri (2012) [115]. The photo-treated wastewater was subjected to the sequencing batch reactor, which operated under different hydraulic retention times. It was observed that about 89% of the COD was removed under the operating conditions at 90 min of reaction time and 12 h of HRT.
Sathya et al. (2019) [116] treated textile wastewater with the integration of a membrane bioreactor with ozone and achieved the maximum removal efficiency of 94% color and 93% COD removal. Liu et al. (2022) [117] treated Congo red wastewater by integrated photocatalysis and biodegradation. The result clearly indicated that about 76% of the COD removal was obtained for the degradation of Congo red. Faggiano et al. (2023) [118] treated the olive mill wastewater with a moving bed biofilm reactor and a photo-Fenton-like process. It was observed that about 91% of the COD removal was achieved with MBBR, but the phenol removal was about 57%. However, with the photo-Fenton-like process, the phenol removal was found to be 99%. Liu et al. (2023) [119] explored the integration of photocatalysis with MBBR for treating emerging organic pollutants and attained a COD removal of 80.7%. Ali et al. (2023) [120] studied the efficiency of treating pharmaceutical wastewater containing acetylsalicylic acid, and it was observed that by using this integrated system, 99.7% of the acetylsalicylic acid was removed with a high quality of treated water.
Table 5. Treatment of wastewater by combined AOP and biological processes.
Table 5. Treatment of wastewater by combined AOP and biological processes.
S. NoWastewaterProcessOperational ConditionsDegradation EfficiencyReference
1Industrial park wastewaterFenton oxidation + biological aerated filter2.0 mM FeII, 10 mM H2O2, pH 3.0 and 3.0 h of the reaction timeEfOM (>70%)Su et al., 2021 [82]
2Dyeing wastewaterElectrochemical advanced oxidation process (EAOP) and electrocoagulation (EC) -COD reduction of 93.5%Chanikya et al., 2021 [93]
3Mixed industrial wastewaterElectrochemical advanced oxidation processes (EAOPs) and biological treatmentpH-3, effective area-25 cm2, electrode spacing-1 cm while voltage10 V, persulphate dosage-200 mg L−1, and catalyst dosage-10 mg L−194% COD removalPopat et al., 2019 [121]
4Organic pollutantsAOP + membrane process-Removal efficiency > 90%Titchou et al., 2021 [95]
5Hospital AOP + adsorptionH2O2 of 100 mg L−1 and multiple additions of Fe2+ of 5 mg L−1 at times of t = 0, 5, and 10 min98–100% removal rateDella-Flora et al., 2021 [122]
6Citrus wastewaterozone-based processes150 min. at 1.9gO3/L, 1017 mg/L of H2O2, UV radiation and pH ~ 777% of chemical oxygen demand and 53% of dissolved organic
carbon
Guzman et al., 2016 [35]
7CI Reactive Blue 181Fenton and
sono-Fenton
[Fe2+] = 30 mg/L, [H2O2] = 50 mg/L and pH = 3
[Fe2+] = 10 mg/L, [H2O2] = 40 mg/L and pH = 3
Colour removals
Fenton—88%
sono-Fenton 93.5%
Basturk and Karatas (2014) [88]
8Synthetic wastewaterSonication, mild photo-Fenton treatmentUltrasound 20 W Recirculating Flow rate: 4.39 L/h
Iron 1 ppm Treated Volume: 500 mL
H2O2 10 ppm
Maximum E. coli inactivationGiannakis et al. (2015) [91]
9 Pharmaceutical3D electrolysis ozonation-TOC abatement to ∼71%Zhan et al. (2019) [92]
10Paper millActivated sludge-ozonationpH 8.3COD—75.5%, TOC- 59.1%, colour—77%, Total phenols—52.3%Assalin et al. (2009) [4]
11Pulp and paper millBiological and photochemical3 g L−1 TiO2, pH 6.0 and 0.01 mol L−1 NaOCl81% and 93% in C
and E1 effluent
Dhir et al. (2011) [105]
12Kraft bleaching effluentAnaerobic immobilised biomass reactor and hydrogen peroxide with UV radiation19 h of hydraulic retention time (HRT)COD (61 ± 3%), TOC (69 ± 9%), BOD5 (90 ± 5%) and AOX (55 ± 14%)Ruas et al. (2012) [106]
13Textile wastewaterMembrane bioreactor+ ozonation and photocatalysisPhotocatalyst-500 mg/L, ozone-5 g/hColor—94%, COD—93%Sathya et al., 2019 [116]
14Congo red wastewaterPhotocatalysis + biodegradation COD—94.3%Liu et al., 2022 [117]
15Olive mill wastewaterMBBR + Photo-FentonMBBR—24h, Photo-Fenton—3 hBOD—97%, COD—91%, Total phenols—57%
Photo-Fenton, total phenols—99%
Faggiano et al., 2023 [118]
16Synthetic wastewaterPhotocatalysis and moving bed biofilm reactorHydraulic retention time of 24 hCOD—67.7–80.7 %Liu et al., 2023 [119]
17Pharmaceutical wastewaterAnoxic/activated sludge and ultrafiltration membranepH 6–7 and F/M of 0.85 at HRT of 10 hRemoval efficiency—90%Ali et al., 2023 [120]
As a result of the reports in the literature that biological wastewater treatment is ineffective at successfully removing recalcitrant and hazardous contaminants, a new hybrid technique that combines advanced oxidation processes and biological processes has been developed. AOPs are the only effective treatment options for wastewater containing recalcitrant organic compounds, although they frequently involve substantial capital and operating costs. By combining the economic and environmental benefits of biological and AOP treatments, it would be possible to remove hazardous substances from wastewater practically. This can be achieved by applying AOP as a pretreatment to the wastewater containing recalcitrant organic compounds and optimizing the reagent consumption within the short reaction time, resulting in the generation of fully biodegradable intermediates. Thus, the pretreated wastewater can be subjected to biological treatment for the complete removal of organic matter.

6. Solar Photocatalytic Reactors

Solar energy is a practical and efficient solution to address the energy crisis and promote environmental protection. The sun is our planet’s most plentiful source of energy, and during clear weather, when the sun is near its zenith, the amount of solar radiation that reaches the earth’s surface equals about 1 kW/m2. Direct radiation, also known as beam radiation, originates from the sun’s disc directly, while diffuse radiation travels to the earth after being dispersed by the atmosphere in all directions. Thus, the total amount of solar radiation equals the sum of the two elements (direct and diffuse radiation). Solar energy is focused on collectors that gather and concentrate the solar radiation onto a smaller receiving surface to reach high temperatures.
Therefore, a photoreactor that can effectively interact with the photocatalyst and solar photons is required for the industrial use of a photocatalytic process. The photons from solar irradiation are used to degrade wastewater in solar photocatalytic reactors. More and more research is focusing on solar photoreactors using sunlight as an economical alternative to expensive UV radiation [123]. To degrade organic compounds, many solar photocatalytic reactor designs have been developed. These reactor systems are categorized as concentrating and non-concentrating reactor systems based on how well they absorb sunlight [124]. Curved mirrors are used in concentrating reactors to focus solar radiation onto a receiver, which is then heated by the radiation. A fluid that travels through the receiver receives the heat that was absorbed. The usage of these promotes the enhancement of their efficiency, which is dependent on direct-beam irradiation. The term “parabolic trough” refers to a solar thermal collector with a parabola form that is straight in one dimension and curved in the other two. It is lined with a polished metal mirror. The focal line, where things are placed that is meant to be heated, is where the sunlight’s energy is focused after entering the mirror parallel to its plane of symmetry.
The direct ultraviolet photons of the solar spectrum are focused using a parabolic trough reactor [125]. Singh et al. (2013) [52] examined the photocatalytic oxidation of synthetic wastewater containing high COD using a parabolic trough reactor. TiO2 has been used as a photocatalyst. It was reported that a maximum COD reduction of 86% was obtained. The drawbacks of parabolic reactors include the fact that they can only receive direct radiation and have a significant cost associated with the tracking system and an inadequate quantum yield when using TiO2. Non-concentrating reactors are simple in design and static without any solar tracking device. They are usually a flat plate with a determined tilt. The main advantage of a non-concentrating reactor is the reduced cost. The following is some of the literature on different types of reactors. Non-concentrating reactors absorb direct and diffuse solar radiation with negligible optical losses, whereas in a concentrating reactor system, the sunlight is concentrated, thus decreasing the size of the reactor compared to that of the non-concentrating reactor. The literature studies on various types of reactors used for wastewater treatment are listed in Table 6.

6.1. Thin Film Fixed-Bed Reactor

The wastewater runs over the sloped P25 TiO2 DEGUSSA-coated plate that makes up the thin film fixed-bed reactor (TFFBR). The pump that is found within the reactor can regulate the flow rate. The key benefits of this TFFBR are (i) its good optical efficiency, (ii) its uncomplicated design, and (iii) the inexpensive capital cost. Kernani et al. (2014) [126] studied the treatment of landfill containing recalcitrant organic compounds using a thin film fixed-bed reactor. It was reported that about 92% COD removal was achieved at a pH of 5.

6.2. Compound Parabolic Reactor

A large-scale solar thermal energy collection is made possible by the compound parabolic concentrator (CPC). Without the need for a sophisticated sun-tracking mechanism, a CPC is made up of two parabolas with a compound rotation [102]. The compound parabolic concentrator has been extensively interpreted as a good option for solar photochemical applications. With the use of a pilot CPC, Velegraki and Mantzavinos (2015) [32] successfully mineralized and detoxified winery wastewater by using the sunlight-driven Fenton process.

6.3. Double-Skin Sheet Reactor

A flat box with two parallel UV translucent Plexiglas panes spaced a few centimeters apart makes up the double-skin sheet reactor. Channels are created inside the reactor by Plexiglas strips wrapped around the double skin. During insolation, this channel system is used to pump the suspension of the photocatalyst [127]. Dillert et al., 1999 [125] treated various types of biologically pretreated industrial wastewater in a double-skin sheet reactor. The disadvantage of this reactor is that it can function in slurry mode and has low optical efficiency.

6.4. Shallow Pond Reactor

Using direct and diffuse radiation, the shallow pond reactor may operate on sunny and gloomy days, and it is a non-concentrating reactor [128]. Toor et al., 2006 [129] investigated the decomposition of diazo Direct Yellow 12 using the semiconductor TiO2 as a catalyst in a non-concentrated shallow pond slurry-type reactor under UV light. It was reported that about 94% COD reduction was achieved after 2.5 h, and complete decolorization was achieved after 1.5 h by UV-vis analysis.

6.5. Pebble Bed Reactor

The innovative pebble bed reactor utilizes white pebbles coated with TiO2 and a horizontal or slope solar trough collector to degrade wastewater. The pebbles are dispersed in a triangular arrangement to aid in the interaction between the liquid and the photocatalyst. The outcome revealed that under sunlight, 72% colour and 3–35% TOC removal was accomplished [130].

6.6. Fountain Reactor

The fountain photocatalytic reactor is made up of a flattened water bell exposed to sunlight or UV lamps. Water is pumped through a specially made nozzle to create a thin film, which emerges from the nozzle in the form of a smooth, roughly horizontal, and radially widening water fountain [131].
In the above discussions, a reactor using immobilized TiO2 with a laminar flow, such as a thin film fixed-bed, pebble bed, or fountain reactor, only has a limited ability to degrade substances. Similarly, a shallow tank reactor with immobilized TiO2 had the highest throughput while only achieving partial degrading efficiency [128]. Despite several types of research on the efficient use of various reactors, large-scale reactor applications are missing due to opacity, light scattering, and depth of radiation penetration.
Table 6. Various types of solar photocatalytic reactors used for the degradation of wastewater.
Table 6. Various types of solar photocatalytic reactors used for the degradation of wastewater.
S. NoTypes of ReactorPollutant/
Wastewaters
Reference
1Raceway pond reactorMicropollutantsArzate et al., 2017 [132]
2Raceway pond reactorChlorpyrifosAmiri et al., 2018 [133]
3Parabolic trough collectorPhenolAbid et al., 2017 [134]
4Water bell reactorIndustrial wastewaterMaksoud et al., 2018 [128]
5Compound parabolic reactorPharmaceuticalsAlmomani et al., 2018 [135]
6Solar pond reactorsReactive orange dyeChavaco et al., 2017 [136]
7Batch stirred and fluidized bed reactorPhenolShet and Shetty, 2016 [137]
8Compound parabolic collectorWinery effluentsVelegraki and Mantzavinos, 2015 [32]
9Thin film fixed-bed reactorAmaranthSudrajat and Babel, 2016 [138]
10Photocatalytic solar tower reactor Volatile organic compoundsNegishi and Sano, 2014 [139]
11Water falling DBD reactor MesotrioneJovic et al., 2014 [140]
12Thin film cascade reactorBenzoic acid Chan et al., 2003 [141]
13Thin film fixed-bed reactor PesticidesShankar et al., 2004 [142]
14Water fountain photocatalytic reactorsTextile dyeing wastewatersKanmani et al., 2003 [124]
15Thin film multi-tubular photoreactor Methyl orangeAdams et al., 2013 [143]
16Double-skin sheet reactorIndustrial wastewaterDillert et al., 1999 [125]
17Shallow pond reactorDirect Yellow 12 dyeToor et al., 2006 [129]
18Falling film reactorGelatin industryVelmurugan et al., 2016 [144]
19Fixed-bed reactorMethyl orangeKhalilian et al., 2015 [145]
20Continuous stirred tank reactor2,4-dinitrophenolMiyawaki et al., 2016 [146]
21Thin film fixed-bed reactorLandfill waterKernani et al., 2014 [126]
22Pebble bed reactorTextile wastewaterRao et al., 2012 [130]
23Fountain reactorWater purificationPuma and Yue, 2001 [131]
24Compound parabolic reactorIndustrial wastewaterNair et al., 2016 [147]
25Annular photoreactorPharmaceuticalSarkar et al., 2017 [148]
26Raceway pondLandfill leachateSpineillo et al., 2023 [149]

7. Economic Analysis of Advanced Oxidation Processes

Cost is a major consideration whenever new technologies are being developed. It shows whether any new, generally accepted commercial techniques or technologies are significantly lowering processing costs and whether they are commercially feasible or not [150]. The primary variables affecting solar detoxification systems are the installed costs, operation costs, and treatment costs. The installed cost comprises all expenses related to the design, installation, and startup of the facility. The operational cost is the second deciding aspect in the cost of treatment after the total installed cost. The total costs of operating and maintaining the facility, including all labor and material expenditures (such as those associated with procurement, maintenance, and supervision), chemical supplies, and electricity costs, must be included in the operating cost. The cost comparison is an important component in comparing AOPs, depending on their intended application, such as a pre- or partial treatment to attain total mineralization (>99%) [151].
By using an integrated photo-Fenton and electrocoagulation process, [152] Modenes et al. (2012) conducted a cost study for tannery effluent while only taking into account the direct cost elements. According to [21], who handled the wastewater from polyester dyeing, it costs too much to use artificial radiation in the photo-Fenton process (17.4, 9.6, and 2.9 EUR/m3 for acrylic, cotton, and polyester effluents, respectively). Using (simulated) solar light makes it possible to achieve high levels of color removal and a considerable decrease in COD and DOC removal at a reduced total cost of a factor of approximately three. According to the published research, the solar photo-Fenton process will cost roughly 2.11 EUR/m3 to run for five years with a treatment capacity of 50 m3/d (USD 2.2155) [153]. Alalm et al. (2014) [154] explored the solar photo-Fenton process using a compound parabolic collector reactor for extracting phenol from an aqueous solution. The costs connected with a large-scale reactor were examined. It was observed that the overall cost of the best economic scenario with the greatest phenol degradation was 2.54 EUR/m3. Torres et al., 2015 [155] studied the economic assessment of landfill leachate by the solar photo–Fenton process with the design flow rate of 40 m3/d, and the estimated treatment cost was 43 EUR/m3 (USD 45.15). Selvabharathi et al. (2016) [156] carried out a cost analysis by analyzing the degradation of tannery wastewater by the combined homogeneous and heterogeneous process and estimated the annual treatment cost as USD 19.35.
According to research by Aljouboury et al. [64] from 2021, the best economic condition with the greatest phenol degradation costs was found to be 2.54 EUR/m3. Furthermore, the TiO2/ZnO/air process was estimated to cost 9 USD/m3, while the TiO2/ZnO/Fenton method was estimated to be about 8 USD/m3. (Mousset et al., 2021) [151] evaluated the operating costs for each AOP by normalizing them based on the treated volume and mass of TOC removed and taking into account sludge treatment and chemical and power consumption. It was clear that due to its electrocatalytic behavior, electro-Fenton was the most cost-effective AOP (108–125 EUR/m3), regardless of the mineralization target (50%, 75%, or 99%). Benis et al. (2021) [157] treated laundry wastewater by the combined biological and physicochemical process and reported that the operating cost analysis for the combined process was 0.65 EUR/m3. It has been estimated that the operating costs and investment involved in the photocatalytic treatment of wastewater are lower than other wastewater treatment technologies such as ozone treatment, the chlorination of water, and the wet oxidation process of water, incineration, and membrane separation [150]. One of the main drawbacks of the use of AOPs for the treatment of new compounds is their high cost. Hybrid techniques involving biological treatment and advanced oxidation processes were created to make them more energetically effective and applicable. AOPs have been recommended as a supplement to biological treatment for the removal of evolving compounds because they offer considerable advantages over using biological approaches alone. One of the ways used to reduce costs, increase permeate quality, and improve treatment efficiency is the combination of AOPs and biological treatments.
Thus, the literature above discussed the cost of treatment for various advanced oxidation processes. It was observed that the treatment cost of the individual process is more compared to the combined process, and therefore, the integration of AOPs with a biological treatment may be a suitable solution to reduce the operating cost and make the process more profitable.

8. Wastewater Treatment: Challenges and Opportunities

The discharge of harmful chemicals into water sources, which disrupts the ecosystem’s ability to operate, is one of the most important problems people now confront. Currently, the treatment of wastewater is a difficult endeavor that depends not only on regional and socioeconomic factors but also on laws for controlling the waste. As each treatment has unique advantages and difficulties in terms of operational complexity, ecological consequences, sludge generation, viability, applicability, and cost-efficiency, the optimal technique has not yet been determined [158]. Solar photocatalysis is a powerful and environmentally friendly energy source for degrading vulnerable organic pollutants in water sources. Even though these single AOPs are quite effective at treating wastewater, they have certain limitations such as the possibility of the formation of secondary pollutants, the necessity of expensive catalysts and electrode materials, and high costs.
The integration of AOPs and the coupling of AOPs with traditional technologies have advanced in several areas, including economic efficiency, environmental friendliness, and operability [159]. A key challenge in the homogeneous Fenton method is that it has a significant issue with iron-containing sludge, which needs additional treatment. The fact that the reaction occurs at low pH levels is another disadvantage of Fenton’s method. Although maintaining a low pH is not a concerning issue, a scaled-up process will require low pH values to be taken into consideration when designing the materials for reactors and pipelines. Though much of the literature reported the application of the solar photo-Fenton process for the degradation of wastewater containing different complex organic compounds with high efficiency, the Fenton process is still not considered to be the ultimate treatment technique because it works at very low pH. Therefore, this process is pH-dependent and strict pH control is required, thus avoiding any precipitation of inactive iron oxyhydroxides and the inevitability of its post-treatment expense when neutralizing the treated water before disposal. However, an important drawback to industrial applications of the photo-Fenton process is the consumption of oxidizing reagents and the chemicals required for increasing the alkalinity of the wastewater during pH adjustment, which inhibits its wide application, potentially resulting in higher costs.
Finding an appropriate catalyst (second phase) to treat wastewater is the most difficult aspect of heterogenous AOPs. However, the photocatalysis process has certain noticeable drawbacks, such as the high costs associated with removing the catalyst after use and the challenge of achieving radiation uniformly throughout the entire catalyst surface at a greater scale [158]. The removal of recalcitrant organic compounds is accomplished through the integration of two or more treatment approaches. In terms of energy-saving and treatment effectiveness, this method is reliable and stable. Numerous studies have shown that these hybrid approaches improve water contaminant separation efficiency while consuming less energy [160]. Wastewater treatment methods that are integrated have demonstrated tremendous promise for the mitigation of environmental pollution issues and the improvement of the availability of potable water. As a result, in addition to enhancing individual processes, future studies will focus on developing possible integrated treatment processes.
To increase the use of AOPs and AOP–biological processes in wastewater treatment, a techno-economic comparison with other technologies is being conducted [161]. When compared to traditional wastewater and water treatment systems, combination processes are more efficient. However, for each water and wastewater sample, a different type of pretreatment and combination may be required. However, several restrictions have prevented integrated wastewater/water systems from being widely used on an industrial scale. Large-scale integrated wastewater and water treatment processes have not been used due to the lack of technological maturity and feasible data, the difficulty of managing integrated systems, and the activity reduction in treatment materials such as photocatalysts and adsorbents over time [160].

9. Conclusions

This review focuses on individual AOPs, integrated AOPs, and AOPs + biological processes for treating various industrial wastewaters. In this study, a methodical way to evaluate different AOPs was provided. Advanced oxidation processes provide effective methods for removing organic pollutants from water. Compared to conventional wastewater treatment systems, hybrid and integrated procedures are more efficient. Combining AOPs with a biological treatment may be an ideal approach for reducing operating costs and increasing the profitability of the treatment process. AOPs are often more expensive and require more reagents and energy sources, which limits their industrial applicability. AOP efficiency can be increased by adopting specific reactor configuration. A solar photoreactor is required for the photochemical treatment of wastewater using a light source. Due to the high operational costs of UV-based AOPs, solar irradiation may provide a long-term solution to these constraints. Therefore, it is necessary to assure that the new technology will be cost-competitive by including all the expenditures in the economic strategy.

Author Contributions

G.G.: writing original draft, proofreading, conceptualization; R.B.J.: resources, proofreading; A.S.: supervision, proofreading. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Renuka, R.; Mohan, M.S.; Sowmiya, B.; Raj, A.S. Performance evaluation of panelled anaerobic baffle-cum-filter reactor in treating municipal wastewater. Ecol. Eng. 2016, 97, 1–12. [Google Scholar] [CrossRef]
  2. Nandkumar, P. Studies on the effluent generated during the pulping process in paper industry. Curr. World Environ. 2008, 3, 189–193. [Google Scholar] [CrossRef]
  3. Subramonian, W.; Wu, T.Y.; Chai, S.P. A comprehensive study on coagulant performance and floc characterization of natural Cassia obtusifolio seed gum in treatment of raw pulp and paper mill effluent. Ind. Crops Prod. 2014, 61, 317–324. [Google Scholar] [CrossRef] [Green Version]
  4. Assalin, M.R.; Almeida, E.S.; Duran, N. Combined System of Activated Sludge and Ozonation for the Treatment of Kraft E1 Effluent. Int. J. Environ. Res. Public Health 2009, 6, 1145–1154. [Google Scholar] [CrossRef] [Green Version]
  5. Davarnejad, R.; Nasiri, S. Slaughterhouse wastewater treatment using an advanced oxidation process: Optimization study. Environ. Pollut. 2016, 223, 1–10. [Google Scholar] [CrossRef]
  6. Ma, D.; Yi, H.; Lai, C.; Liu, X.; Huo, X.; An, Z.; Li, L.; Fu, Y.; Li, B.; Zhang, M.; et al. Critical review of advanced oxidation processes in organic wastewater treatment. Chemosphere 2021, 275, 130104. [Google Scholar] [CrossRef]
  7. Barcelo, M.A.; Lopez, M.I.P.; Lucena, F.; Jofre, J.; Ibanez, P.F. Solar Advanced Oxidation Processes as disinfection tertiary treatments for real wastewater: Implications for water reclamation. Appl. Catal. B Environ. 2013, 136–137, 341–350. [Google Scholar] [CrossRef]
  8. Hutagalung, S.S.; Muchlis, I.; Khotimah, K. Textile wastewater treatment using Advanced Oxidation Process. Mater. Sci. Eng. 2020, 722, 012032. [Google Scholar] [CrossRef]
  9. Fernandes, E.; Gomes, J.; Martins, R.C. Semiconductors Application Forms and Doping Benefits to Wastewater Treatment: A Comparison of TiO2, WO3, and g-C3N4. Catalysts 2022, 12, 1218. [Google Scholar] [CrossRef]
  10. Tong, Y.; Wang, L.; Hou, F.; Dou, S.X.; Liang, J. Electrocatalytic Oxygen Reduction to Produce Hydrogen Peroxide: Rational Design from Single-Atom Catalysts to Devices. Electrochem. Energy Rev. 2022, 5, 1–34. [Google Scholar] [CrossRef]
  11. Lopez, M.I.P.; Alferez, M.C.; Oller, I.; Ibanez, P.F. Assessment of solar photo-Fenton, photocatalysis, and H2O2 for removal of phytopathogen fungi spores in synthetic and real effluents of urban wastewater. Chem. Eng. J. 2014, 257, 122–130. [Google Scholar] [CrossRef]
  12. Kurian, M. Advanced oxidation processes and nanomaterials–A review. Clean. Eng. Technol. 2021, 2, 100090. [Google Scholar] [CrossRef]
  13. Garrido-Cradenas, J.A.; Esteban-Garcia, B.; Aguera, A.; Sanchez-Perez, J.A.; Manzano-Agugliaro, F. Wastewater Treatment by Advanced Oxidation Process and Their Worldwide Research Trends. Int. J. Environ. Res. Public Health 2019, 17, 170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Paździor, K.; Bilińska, L.; Ledakowicz, S. A review of the existing and emerging technologies in the combination of AOPs and biological processes in industrial textile wastewater treatment. Chem. Eng. J. 2018, 376, 120597. [Google Scholar] [CrossRef]
  15. Abdelhaleem, A.; Chu, W. Prediction of Carbofuran Degradation Based on the Hydroxyl Radical’s Generation using the FeIII impregnated N doped-TiO2/H2O2/Visible LED Photo-Fenton-Like Process. Chem. Eng. J. 2019, 382, 122930. [Google Scholar] [CrossRef]
  16. Nidheesh, P.V.; Couras, C.; Karim, A.V.; Nadais, H. A review of integrated advanced oxidation processes for organic pollutant removal. Chem. Eng. Commun. 2021, 209, 390–432. [Google Scholar] [CrossRef]
  17. Aramyan, S.M.; Moussavi, M. Advances in Fenton and Fenton based Oxidation processes for industrial effluent contaminants control—A review. Int. J. Environ. Sci. Nat. Resour. 2017, 2, 555594. [Google Scholar] [CrossRef] [Green Version]
  18. Ribeiro, A.R.; Nunes, O.C.; Pereira, M.F.R.; Silva, A.M.T. An overview on the advanced oxidation processes applied for the treatment of water pollutants defined in the recently launched Directive 2013/39/EU. Environ. Int. 2015, 75, 33–51. [Google Scholar] [CrossRef] [Green Version]
  19. Gomez, E.; Martin, M.M.B.; Carratala, A.P.; Ibanez, F.; Pérez, J.A.S.; Pulgarin, C. Principal parameters affecting virus inactivation by the solar photo-Fenton process at neutral pH and concentrations of H2O2 and Fe2+. Appl. Catal. B Environ. 2015, 174–175, 395–402. [Google Scholar] [CrossRef]
  20. Barwal, A.; Chaudhary, R. Effectiveness of solar photo—Fenton process for simultaneous detoxification of heavy metals and disinfection in municipal wastewater by using response surface method. Environ. Prog. Sustain. Energy 2016, 36, 448–459. [Google Scholar] [CrossRef]
  21. Rodriguez, L.P.; Oller, I.; Klamerth, N.; Aguera, A.; Rodriguez, E.M.; Malato, S. Application of AOPs and ozonation for elimination of micropollutants in municipal wastewater treatment plant effluents. Water Res. Nano Energy 2013, 47, 1521–1528. [Google Scholar]
  22. Garcia, P.D.; Buitron, G. Improvement of the robustness of solar photo-Fenton processes using chemometric techniques for the decolorization of azo dye mixtures. J. Environ. Manag. 2013, 131, 66–73. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, M.; Wang, Q.; Hao, Y. Removal of organic carbon from wastewater pulp effluent by lab-scale solar photo-Fenton process. J. Hazard. Mater. 2007, 148, 103–109. [Google Scholar] [CrossRef] [PubMed]
  24. Sheik, M.A.; Kumar, A.; Paliwal, M.; Ameta, R.; Khandelwal, R.C. Degradation of organic effluents containing wastewater by photo-Fenton oxidation process. Indian J. Chem. 2008, 4, 1681–1684. [Google Scholar]
  25. Karimi, S.; Abdolkhani, A.; Karimi, A. Discoloration of Soda Pulping Effluent by Advanced Oxidation Processes. Eng. e-Trans. 2011, 6, 20–25, ISSN 1823-6379. [Google Scholar]
  26. Lucas, M.S.; Peres, J.A.; Amor, C.; Rodriguez, L.P.; Maldonado, M.I.; Malato, S. Tertiary treatment of pulp mill wastewater by solar photo—Fenton. J. Hazard. Mater. 2012, 226, 173–181. [Google Scholar] [CrossRef]
  27. Bernabeu, A.; Palacios, S.; Vicente, R.; Vercher, R.F.; Malato, S.; Arques, A.; Amat, A.M. Solar photo-Fenton at mild conditions to treat a mixture of six emerging pollutants. Chem. Eng. J. 2012, 198–199, 65–72. [Google Scholar] [CrossRef]
  28. Trovo, A.G.; Nogueira, R.F.P.; Aguera, A.; Alba, A.R.F.; Malato, S. Paracetamol degradation intermediates and toxicity during photo-Fenton treatment using different iron species. Water Res. 2012, 46, 5374–5380. [Google Scholar] [CrossRef]
  29. Turbay, E.Y.; Jaen, E.; Graells, M.; Moya, M.P. Enhanced photo-Fenton process for tetracycline degradation using efficient hydrogen peroxide dosage. J. Photochem. Photobiol. A Chem. 2013, 267, 11–16. [Google Scholar] [CrossRef]
  30. Rocha, O.R.S.; Dantas, R.F.; Duarte, M.M.M.B.; Duarte, M.M.L.; da Silva, V.L. Solar photo-Fenton treatment of petroleum extraction wastewater. Desal. Water Treat. 2014, 51, 5785–5791. [Google Scholar] [CrossRef]
  31. Rodrıguez, M.J.H.; Rodrıguez, C.F.; Rodrıguez, J.M.D.; Dıaz, O.M.G.; Zerbani, D.; Pena, J.P. Treatment of effluents from wool dyeing process by photo-Fenton at solar pilot plant. J. Environ. Chem. Eng. 2014, 2, 163–171. [Google Scholar] [CrossRef]
  32. Velegraki, T.; Mantzavinos, D. Solar photo-Fenton treatment of winery effluents in a pilot photocatalytic reactor. Catal. Today 2015, 240, 153–159. [Google Scholar] [CrossRef]
  33. Chueca, J.R.; Lopez, M.I.P.; Mosteo, R.; Ormad, M.R.; Ibanez, P.F. Disinfection of real and simulated urban wastewater effluents using a mild solar photo-Fenton. Appl. Catal. B Environ. 2015, 150–151, 619–629. [Google Scholar]
  34. Soares, P.A.; Batalha, M.; Souza, S.M.A.G.U.; Boaventura, R.A.R.; Vilar, V.J.P. Enhancemnet of a solar photo-Fenton reaction with ferric-organic ligands for the treatment of acrylic-textile dyeing wastewater. J. Environ. Manag. 2015, 152, 120–131. [Google Scholar] [CrossRef]
  35. Guzman, J.; Mosteo, R.; Sarasa, J.; Alba, J.A.; Ovelleiro, J.L. Evaluation of solar photo-Fenton and ozone-based processes as citrus wastewater pre-treatments. Sep. Purif. Technol. 2016, 164, 155–162. [Google Scholar] [CrossRef]
  36. Benitez, H.Z.N.; Penuela, G.A. Application of solar photo-Fenton for benzophenone-type UV filters removal. J. Environ. Manag. 2018, 217, 929–938. [Google Scholar] [CrossRef]
  37. Costa, N.M.; Silva, G.D.; Marson, E.O.; Richter, E.M.; Machado, A.E.H.; Trovo, A.G. Enhanced treatment of a biodiesel effluent using ferrioxalate in a photo-Fenton process based on the use of solar radiation. Fuel 2018, 221, 110–115. [Google Scholar] [CrossRef]
  38. García, A.B.E.; Szymanski, K.; Mozia, S.; Perez, J.A.S. Treatment of laundry wastewater by solar photo-Fenton process at pilot plant scale. Environ. Sci. Pollut. Res. 2021, 28, 8576–8584. [Google Scholar] [CrossRef]
  39. Arka, A.; Asaithambi, P.; Debela, S.K. Development of Solar Photo-Fenton Process for the Removal of Color, COD and Turbidity from Institutional Wastewater. J. Energy Environ. Chem. Eng. 2022, 7, 26–35. [Google Scholar] [CrossRef]
  40. Lumbaque, E.C.; Cardoso, R.M.; de Araújo Gomes, A.; Malato, S.; Sanchez Perez, J.A.; Sirtori, C. Removal of pharmaceuticals in hospital wastewater by solar photo-Fenton with Fe3+-EDDS using a pilot raceway pond reactor: Transformation productsand in silico toxicity assessment. Microchem. J. 2021, 164, 106014. [Google Scholar] [CrossRef]
  41. Lin, H.H.H.; Lin, A.Y.C. Solar photo-Fenton oxiadation of cytostatic drugs via FeIII-EDDS at circumneutral pHin an aqueous environment. J. Water Process Eng. 2021, 41, 102066. [Google Scholar] [CrossRef]
  42. Gualda-Alonso, E.; Soriano-Molina, P.; Lopez, J.L.C.; Sanchez, J.L.G.; Plaza-Bolaños, P.; Agüera, A.; Perez, J.A.S. Large scale raceway pond reactor for CEC removal from municipal WWTP effluents by solar photo-Fenton. Appl. Catal. B Environ. 2022, 319, 121908. [Google Scholar] [CrossRef]
  43. Pandey, Y.; Verma, A.; Toor, A.P. Abatement of paraquat contaminated water using solar assisted heterogeneous photo-Fenton like treatment with iron-containing industrial wastes as catalysts. J. Environ. Manag. 2023, 325, 116550. [Google Scholar] [CrossRef] [PubMed]
  44. Antunes, C.S.A.; Bietti, M.; Salamone, M.; Scione, N. Early stages in the TiO2-photocatalyzed degradation of simple phenolic and non-phenolic lignin model compounds. J. Photochem. Photobiol. A Chem. 2004, 163, 453–462. [Google Scholar] [CrossRef]
  45. Fotiou, T.; Triantis, T.; Kaloudis, T.; Hiskia, A. Photocatalytic degradation of cylindrospermopsin under UV-A, solar and visible light using TiO2. Mineralization and intermediate products. Chemosphere 2015, 119, 589–594. [Google Scholar] [CrossRef]
  46. Rashid, J.; Barakat, M.A. Ag3PO4 enhanced TiO2 for visible light photocatalysis of 2-chlorophenol in wastewater. Int. J. Environ. Eng. 2015, 2, 93–97. [Google Scholar]
  47. Shinde, S.S.; Bhosale, C.H.; Rajpure, K.Y. Photodegradation of organic pollutants using N-titanium oxide catalyst. J. Photochem. Photobiol. B Biol. 2014, 141, 186–191. [Google Scholar] [CrossRef]
  48. Portjanskaja, E.; Stepnova, K.; Klauson, D.; Preis, S. The influence of titanium dioxide modifications on photocatalytic oxidation of lignin and humic acids. Catal. Today 2009, 144, 26–30. [Google Scholar] [CrossRef]
  49. Amat, A.M.; Arques, A.; Lopez, F.; Miranda, M.A. Solar photocatalysis to remove paper mill wastewater pollutants. Sol. Energy 2005, 79, 393–401. [Google Scholar] [CrossRef]
  50. Fernandez, I.G.; Calderero, I.F.; Lopez, M.I.P.; Ibanez, P.F. Disinfection of urban effluents using solar TiO2 photocatalysis: A study of significance of dissolved oxygen, temperature, type of microorganism and water matrix. Catal. Today 2015, 240, 30–38. [Google Scholar] [CrossRef]
  51. Karat, I. Advanced Oxidation Processes for Removal of COD from Pulp and Paper Mill Effluents A Technical, Economical and Environmental Evaluation, Royal Institute of Technology. Stockholm 2013, 6, 1–104. [Google Scholar]
  52. Singh, C.; Chaudhary, R.; Gandhi, K. Preliminary study on optimization of pH, oxidant and catalyst dose for high COD content: Solar parabolic trough collector. Iran. J. Environ. Health Sci. Eng. 2013, 10, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Adishkumar, S.; Kanmani, S. Treatment of phenolics wastewater in single baffle reactor by solar/TiO2/H2O2 process. Desal. Water Treat. 2010, 24, 67–73. [Google Scholar] [CrossRef]
  54. Kumar, P.; Kumar, S.; Bhardwaj, N.K.; Choudhary, A.K. Advanced Oxidation of Pulp and Paper Industry Effluent. Int. Conf. Environ. Agric. Eng. Singap. 2011, 15, 170–178. [Google Scholar]
  55. Dimitrakopoulou, D.; Rethemiotaki, I.; Frontistis, Z.; Xekuokuolotakis, N.P.; Veniere, D.; Mantzavinos, D. Degradation, mineralization and antibiotic inactivation of amoxicillin by UV-A/TiO2 photocatalysis. J. Environ. Manag. 2012, 98, 168–174. [Google Scholar] [CrossRef] [PubMed]
  56. Nagpure, H.; Banakar, V.; Dhanda, R.; Wani, K.S. Degradation of Paper Mill Wastewater using Batch (Photo catalytic) Reactor. Int. J. Green Chem. Bioprocess 2013, 3, 24–29. [Google Scholar]
  57. Ruiz, O.S.; Zanella, R.; Prado, B. Photocatalytic degradation of trimethoprim by pulp metallic nano particles TiO2 P25. J. Hazard. Mater. 2013, 263, 28–35. [Google Scholar] [CrossRef]
  58. Shao, C.; Zhou, G.; Li, Z.; Wu, Y.; Xu, D.; Sun, B. Fabrication of large-diameter tube-like mesoporous TiO2 via homogeneous precipitation and photocatalytic decomposition of papermaking wastewater. Chem. Eng. J. 2013, 230, 227–235. [Google Scholar] [CrossRef]
  59. Antonopoulou, M.; Konstantinou, I. Photocatalytic treatment of metribuzin herbicide over TiO2 aqueous suspensions: Removal efficiency, identification of transformation products, reaction pathways and ecotoxicity evaluation. J. Photochem. Photobiol. A Chem. 2014, 294, 110–120. [Google Scholar] [CrossRef]
  60. Thomas, R.T.; Rasheed, P.A.; Sandhyarani, N. Synthesis of nanotitania decorated few-layer graphene for enhanced visible light driven photocatalysis. J. Colloid Interface Sci. 2014, 428, 214–221. [Google Scholar] [CrossRef]
  61. 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. [Google Scholar] [CrossRef]
  62. Solano, R.; Cerri, G.; Herrera, A.; Vargas, X. Cr6+ and Zn2+ removal for heterogeneous photocatalysis with TiO2 in synthetic wastewater. Int. J. Chem. Tech. Res. 2018, 11, 312–320. [Google Scholar]
  63. Yang, H.; Yang, J. Photocatalytic degradation of rhodamine B catalysed by TiO2 films on a capillary column. RSC Adv. 2018, 8, 11921–11929. [Google Scholar] [CrossRef] [Green Version]
  64. Aljouboury, D.A.D.A.; Shaik, F. Optimization of the petroleum wastewater treatment process using TiO2/Zn photocatalyst. S. Afr. J. Chem. Eng. 2021, 38, 61–69. [Google Scholar]
  65. Kader, S.; Al-Mamun, M.R.; Suhan, M.B.K.; Shuchi, S.B.; Islam, M.S. Enhanced photodegradation of methyl orange dye under UV irradiation using MoO3and Ag doped TiO2 photocatalysts. Environ. Technol. Innov. 2022, 27, 102476. [Google Scholar] [CrossRef]
  66. Abusallout, I.; Hua, G. Solar photocatalytic degradation of total organic halogen in water using TiO2 catalyst. Chemopshere 2022, 308, 136206. [Google Scholar] [CrossRef]
  67. Martins, R.C.; Sacras, A.; Jovanovic, S.; Alves, P.; Ferreira, P.; Gomes, J. Solar energy for liquid wastewater treatment with novel TiO2 supported catalysts. Energy Rep. 2022, 8, 489–494. [Google Scholar] [CrossRef]
  68. Behera, A.K.; Shadangi, K.P.; Sarangi, P.K. Synthesis of dye-sensitized TiO2/Ag doped nano-composites using UV photoreduction process for phenol degradation: A comparative study. Environ. Pollut. 2022, 312, 120019. [Google Scholar] [CrossRef]
  69. Rueda-Marquez, J.J.; Palacios-Villarreal, C.; Manzano, M.; Blanco, E.; del Solar, M.R.; Levchuk, I. Photocatalytic degradation of pharmaceutically active compounds (PhACs) in urban wastewater treatment plants effluents under controlled and natural solar irradiation using immobilized TiO2. Sol. Energy 2020, 208, 480–492. [Google Scholar] [CrossRef]
  70. Gai, H.; Wang, H.; Liu, L.; Feng, B.; Xiao, M.; Tang, Y.; Qu, X.; Song, H.; Huang, T. Potassium and iodide codoped mesoporous titanium dioxide for enhancing photocatalytic degradation of phenolic compounds. Chem. Phys. Lett. 2021, 767, 138367. [Google Scholar] [CrossRef]
  71. Giang, N.T.H.; Tan, N.N.; Huang, L.M.; Hai, N.D.; Thinh, N.T.; Phuc, N.T.; Dat, N.M.; Phong, M.T.; Hieu, N.H. Photocatalytic degradation of crystal violet on titanium dioxide/grapheme aerogel doped sulfur. J. Mol. Struct. 2023, 1271, 134031. [Google Scholar] [CrossRef]
  72. Rapti, I.; Kosma, C.; Albanis, T.; Konstantinou, I. Solar photocatalytic degradation of inherent pharmaceutical residues in real hospital WWTP effluents using titanium dioxide on a CPC pilot scale reactor. Catal. Today, 2022; in press. [Google Scholar] [CrossRef]
  73. Cunha, D.L.; Kuznetzov, A.; Achete, C.A.; Machado, A.E.d.H.; Marques, M. Immobilised TiO2 on glass spheres applied to heterogeneous photocatalysis: Photoactivity, leaching and regeneration. PeerJ 2018, 6, e4464. [Google Scholar] [CrossRef] [Green Version]
  74. Vohra, M.S.; Al-Suwaiyan, M.S.; Essa, M.H.; Chowdhury, M.M.I.; Rahman Wu, J.; Xiao, Y.Z.; Yu, H.Q. Degradation of lignin in pulp mill wastewaters by white-rot fungi on biofilm. Bioresour. Technol. 2005, 96, 1357–1363. [Google Scholar]
  75. Chelliapan, S.; Mahat, S.B.; Din, M.F.M.; Yuzir, A.; Othman, N. Anaerobic digestion of paper mill wastewater. Iran. J. Energy Environ. 2012, 3, 85–90. [Google Scholar] [CrossRef]
  76. Belem, A.; Panteleitchouk, A.V.; Duarte, A.C.; Rocha-Santos, T.A.P.; Freitas, A.C. Treatment of the effluent from a Kraft bleach plant with white rot fungi Pleurotus sajar caju and Pleurotus ostreatus. Glob. Nest 2008, 10, 426–431. [Google Scholar]
  77. Durai, G.; Rajasimman, M.; Rajamohan, N. Kinetic studies on biodegradation of tannery wastewater in a sequential batch bioreactor. J. Biotech Res. 2011, 3, 19–26. [Google Scholar]
  78. Ventura, J.R.; Seo, S.; Chung, I.; Yeom, I.; Kim, H.; Oh, Y.; Jahng, D. Enhanced reduction of excess sludge and nutriet removal in a pilot-scale A2(O)- MBR-TAD system. Water Sci. Technol. 2011, 63, 1547–1556. [Google Scholar] [CrossRef]
  79. Kumar, V.; Dhall, P.; Naithani, S.; Kumar, A.; Kumar, R. Biological approach for the treatment of pulp and paper industry effluent in sequencing batch reactor. J. Bioremediat. Biodegrad. 2014, 5, 218. [Google Scholar]
  80. Chen, X.; Si, C.; Fateh, P. Enhancement in biological treatment of pulping wastewater by flyash. Chemosphere 2018, 210, 1–9. [Google Scholar] [CrossRef]
  81. Goswami, L.; Manikandan, N.A.; Dolman, B.; Pakshirajan, K. Biological treatment of wastewater containing a mixture of polycyclic aromatic hydrocarbons using the oleaginous bacterium Rhodococcus opacus. J. Clean. Prod. 2018, 196, 1282–1291. [Google Scholar] [CrossRef]
  82. Su, T.; Wang, Z.; Zhou, K.; Chen, X.; Cheng, Y.; Zhang, G.; Wu, D.W.; Sun, S.P. Advanced treatment of secondary effluent organic, matters (EfOM) from an industrial park wastewater treatment plant by Fenton oxidation combining with aerated filter. Sci. Total Environ. 2021, 784, 147204. [Google Scholar] [CrossRef] [PubMed]
  83. Cesaro, A.; Naddeo, V.; Belgiorno, V. Wastewater Treatment by Combination of Advanced Oxidation Processes and Conventional Biological Systems. J. Bioremediat. Biodegrad. 2013, 4. [Google Scholar] [CrossRef]
  84. Berberidou, C.; Poulios, I.; Xekoukoulotakis, N.P.; Mantzavinos, D. Sonolytic, photocatalytic and sonophotocatalytic degradation of malachite green in aqueous solution. Appl. Catal. B Environ. 2007, 74, 63–72. [Google Scholar] [CrossRef]
  85. Zhang, Y.; Pagilla, K. Treatment of malathion pesticide wastewater with nanofiltration and photo-Fenton oxidation. Desalination 2010, 263, 36–44. [Google Scholar] [CrossRef]
  86. Na, S.; Ahn, G.; Cui, M.; Khim, J. Significant diethyl phthalate (DEP) degradation by advanced oxidation process in aqueous solution. J. Environ. Manag. 2012, 101, 104–110. [Google Scholar] [CrossRef]
  87. Jadhav, S.R.; Saharan, V.K.; Pinjari, D.; Sonawane, S.; Saini, D.; Pandit, A. Synergetic effect of combination of AOP’s (hydrodynamic cavitation and H2O2) on the degradation of neonicotinoid class of insecticide. J. Hazard. Mater. 2013, 261, 139–147. [Google Scholar] [CrossRef]
  88. Basturk, E.; Karatas, M. Advanced oxidation of Reactive Blue 181 solution: A comparison between Fenton and Sono-Fenton process. Ultrason. Sonochem. 2014, 21, 1881–1885. [Google Scholar] [CrossRef]
  89. Gogate, P.R.; Patil, P.N. Combined treatment technology based on synergism between hydrodynamic cavitation and advanced oxidation processes. Ultrason. Sonochem. 2014, 25, 60–69. [Google Scholar] [CrossRef]
  90. Qiu, C.; Yuan, S.; Li, X.; Wang, H.; Bakheet, B.; Komarneni, S.; Wang, Y. Investigation of the synergistic effects of p-nitrophenol mineralization by a combined process of ozonation and electrolysis using a boron-doped diamond anode. J. Hazard. Mater. 2014, 280, 644–653. [Google Scholar] [CrossRef]
  91. Giannakis, S.; Papoutsakis, S.; Darakas, E.; Canellas, A.E.; Petriere, C.; Pulgarin, C. Ultrasound enhancement of near-neutral photo-Fenton for effective E. coli inactivation in wastewater. Ultrason. Sonochem. 2015, 22, 515–526. [Google Scholar] [CrossRef] [Green Version]
  92. Zhan, J.; Li, Z.; Yu, G.; Pan, X.; Wang, J.; Zhu, W.; Han, X.; Wang, Y. Enhanced treatment of pharmaceutical wastewater by combining three dimensional electrochemical process with ozonation to in situ regenerate granular activated particle electrodes. Sep. Purif. Technol. 2019, 208, 12–18. [Google Scholar] [CrossRef]
  93. Chanikya, P.; Nidheesh, P.V.; Babu, D.S.; Gopinath, A.; Suresh Kumar, M. Treatment of dyeing wastewater by combined sulfate radical based electro-chemical advanced oxidation and electrocoagulation processes. Sep. Purif. Technol. 2020, 254, 117570. [Google Scholar] [CrossRef]
  94. Bui, D.N.; Minh, T.T. Investigation of TNT red wastewater treatment technology using the combination of advanced oxidation processes. Sci. Total Environ. 2020, 756, 143852. [Google Scholar] [CrossRef]
  95. Titchou, F.E.; Zazou, H.; Afanga, H.; Gaayda, J.E.; Akbour, R.A.; Nidheesh, P.V.; Hamdani, M. Removal of organic pollutants from wastewater by advanced oxidation processes and its combination with membrane processes. Chem. Eng. Process. Process Intensif. 2021, 169, 108631. [Google Scholar] [CrossRef]
  96. Dogan, E.C.; Kilicoglu, O.; Narci, A.O.; Mert, B.K.; Durna, E.; Akbacak, U.A.; Aydiner, C. Fenton and photo-Fenton processes integrated with submerged ultrafiltration for the treatment of pulp and paper industry wastewater. J. Environ. Chem. Eng. 2021, 9, 105878. [Google Scholar] [CrossRef]
  97. Dindas, G.B.; Caliskan, Y.; Celebi, E.E.; Tekbas, M.; Bektas, N.; Yatmaz, H.C. Treatment of pharmaceutical wastewater by combination of electrocoagulation, electro-fentonand photocatalytic oxidation processes. J. Environ. Chem. Eng. 2020, 8, 103777. [Google Scholar] [CrossRef]
  98. Maifadi, S.; Mhlanga, S.D.; Nxumalo, E.N.; Motsa, M.M.; Kuvarega, A.T. Treatment of salon wastewater by peroxydisulfate based advanced oxidation process (PDS-AOP) under solar light. Synergy through integrated technologies. J. Water Process Eng. 2022, 49, 103062. [Google Scholar] [CrossRef]
  99. An, S.A.; Lee, J.; Sim, J.; Park, C.G.; Lee, J.S.; Rho, H.; Park, K.D.; Kim, H.S.; Woo, Y.C. Evaluation of the advanced oxidation process integrated with microfiltration for reverse osmosis to treat semiconductor wastewater. Process Saf. Environ. Prot. 2022, 162, 1057–1066. [Google Scholar] [CrossRef]
  100. Doltade, S.B.; Yadav, Y.J.; Jadhav, N.L. Industrial wastewater treatment using oxidative integrated approach. S. Afr. J. Chem. Eng. 2022, 40, 100–106. [Google Scholar] [CrossRef]
  101. Cesaro, A.; Belgiorno, V. Removal of Endocrine Disruptors from Urban Wastewater by Advanced Oxidation Processes (AOPs): A Review. Open Biotechnol. J. 2016, 10, 151–172. [Google Scholar] [CrossRef]
  102. Ahmad, M.; Haque, S.E.; Nisar, N.; Qureshi, K.; Ghaffar, A.; Abbas, M.; Nisar, J.; Iqbal, M. Detoxification of photo-catalytically treated 2-chlorophenol: Optimization through response surface methodology. Water Sci. Technol. 2017, 76, 323–336. [Google Scholar] [CrossRef] [PubMed]
  103. Guieysse, B.; Norvill, Z.N. Sequential chemical–biological processes for the treatment of industrial wastewaters: Review of recent progresses and critical assessment. J. Hazard. Mater. 2014, 267, 142–152. [Google Scholar] [CrossRef] [PubMed]
  104. Zapata, A.; Malato, S.; Perez, J.A.S.; Oller, I.; Maldonado, M.I. Scale-up strategy for a combined solar photo-Fenton/biological system for remediation of pesticide-contaminated water. Catal. Today 2010, 151, 100–106. [Google Scholar] [CrossRef]
  105. Dhir, A.; Prakash, N.T.; Sud, D. Studies on coupled biological and photochemical treatment of soda pulp bleaching effluents from agro residue based pulp and paper mill. Soc. Chem. Ind. 2011, 86, 1508–1513. [Google Scholar] [CrossRef]
  106. Ruas, D.B.; Chaparro, T.R.; Pires, E.C. Advanced oxidation process H2O2/UV combined with anaerobic digestion to remove chlorinated organics from bleached kraft pulp mill wastewater. Rev. Fac. Ing. Univ. De Antioq. 2012, 63, 43–54. [Google Scholar]
  107. Botia, D.C.; Rodriguez, M.S.; Sarria, V.M. Evaluation of UV/TiO2 and UV/ZnO photocatalytic systems coupled to a biological process for the treatment of bleaching pulp mill effluent. Chemosphere 2012, 89, 732–736. [Google Scholar] [CrossRef]
  108. Moreira, F.C.; Vilar, V.J.P.; Ferreira, A.C.C.; dos Santos, F.R.A.; Dezotti, M.; Sousa, M.A.; Goncalves, C.; Boaventura, R.A.R.; Alpendurada, M.F. Treatment of a pesticide containing wastewater using combined biological and solar-driven AOPs at pilot scale. Chem. Eng. J. 2012, 209, 429–441. [Google Scholar] [CrossRef]
  109. Silva, T.F.C.V.; Silva, M.E.F.; Queda, A.C.C.; Fonseca, A.; Saraiva, I.; Sousa, M.A.; Goncalves, C.; Alpendurada, M.F.; Boaventura, R.A.R.; Vilar, V.J.P. Multistage treatment system for raw leachate from sanitary landfill combining biological nitrification-denitrification/solar photo-Fenton/biological processes, at a scale close to industrial-Biodegradability enhancement and evolution profile of trace pollutants. Water Res. 2013, 47, 6167–6186. [Google Scholar]
  110. Merayo, N.; Hermosilla, D.; Blanco, L.; Cortijo, L.; Blanco, A. Assessing the application of advanced oxidation processes, and their combination with biological treatment, to effluents from pulp and paper industry. J. Hazard. Mater. 2013, 262, 420–427. [Google Scholar] [CrossRef] [Green Version]
  111. Souza, B.S.; Moreira, F.C.; Dezotti, M.W.C.; Vilar, V.J.P. Application of biological oxidation and solar driven advanced oxidation processes to remedition of winery wastewater. Catal. Today 2013, 209, 201–208. [Google Scholar] [CrossRef]
  112. Marsolek, M.D.; Kirisits, M.J.; Gray, K.A.; Rittmann, B.E. Coupled photocatalytic-biodegradation of 2,4,5 trichlorophenol: Effects of photolytic and photocatalytic effluent composition on bioreactor process performance, community diversity, and resistance and resilience to perturbation. Water Res. 2014, 50, 59–69. [Google Scholar] [CrossRef] [PubMed]
  113. Blanco, J.; Torrades, F.; Moron, M.; Agnesa, M.B.; Montano, J.G. Photo-Fenton and sequencing batch reactor coupled to photo-Fenton processes for textile wastewater reclamation: Feasibility of reuse in dyeing processes. Chem. Eng. J. 2014, 240, 469–475. [Google Scholar] [CrossRef] [Green Version]
  114. Mendez, J.A.O.; Melián, J.A.H.; Arana, J.; Rodríguez, J.M.D.; Díaz, O.G.; Pena, J.P. Detoxification of waters contaminated with phenol, formaldehyde and phenol–formaldehyde mixtures using a combination of biological treatments and advanced oxidation techniques. Appl. Catal. B Environ. 2015, 163, 63–73. [Google Scholar] [CrossRef]
  115. Elmolla, E.S.; Chaudhuri, M. The feasibility of using combined Fenton-SBR for antibiotic wastewater treatment. Desalination 2012, 285, 14–21. [Google Scholar] [CrossRef]
  116. Sathya, U.; Keerthi, N.M.; Balasubramanian, N. Evaluation of advanced oxidation processes (AOPs) integrated membrane bioreactor (MBR) for the real textile wastewater treatment. J. Environ. Manag. 2019, 246, 768–775. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, K.; Yang, Y.; Sun, F.; Liu, Y.; Tang, M.; Chen, J. Rapid degradation of Congo red wastewater by Rhodopseudomonas palustris intimately coupled carbon nanotube-Silver modified titanium dioxide photocatalytic composite with sodium alginate. Chemosphere 2022, 299, 134417. [Google Scholar] [CrossRef] [PubMed]
  118. Faggiano, A.; De Carluccio, M.; Fiorentino, A.; Ricciardi, M.; Cucciniello, R.; Proto, A.; Rizzo, L. Photo-Fenton like process as polishing step of biologically co-treated olive mill wastewater for phenols removal. Sep. Purif. Technol. 2023, 305, 122525. [Google Scholar] [CrossRef]
  119. Liu, Q.; Hou, J.; Zeng, Y.; Xia, J.; Miao, L.; Wu, J. Integrated photocatalysis and moving bed biofilm reactor (MBBR) for treating conventional and emerging organic pollutants from synthetic wastewater: Performances and microbial community responses. Bioresour. Technol. 2023, 370, 128530. [Google Scholar] [CrossRef]
  120. Ali, S.S.; Sabry, R.; Ali, H.M.; Gadallah, H.; Mansor, E.S.; Abdallah, H.; Shalaby, M.; Shaban, A.M. Integrated system of anoxic/activated sludge and ultrafiltration membrane for zero liquid discharge of pharmaceutical industrial wastewater treatment. J. Environ. Chem. Eng. 2023, 11, 109068. [Google Scholar] [CrossRef]
  121. Popat, A.; Nidheesh, P.V.; Anantha Singh, T.S.A.; Suresh Kumar, M. Mixed industrial wastewater treatment by combined electrochemical advanced oxidation and biological process. Chemosphere 2019, 237, 124419. [Google Scholar] [CrossRef]
  122. Della-Flora, A.; Wilde, M.L.; Lima, D.; Lima, E.C.; Sirtori, C. Combination of tertiary solar photo-Fenton and adsorption processes in the treatment of hospital wastewater: The removal of pharmaceuticals and their transformation products. J. Environ. Chem. Eng. 2021, 9, 105666. [Google Scholar] [CrossRef]
  123. Barndok, H.; Blanco, L.; Hermosilla, D.; Blanco, A. Heterogeneous photo-Fenton processes using zero valent iron microspheres for the treatment of wastewaters contaminated with 1, 4-dioxane. Chem. Eng. J. 2016, 284, 112–121. [Google Scholar] [CrossRef]
  124. Kanmani, S.; Thanasekaran, K.; Beck, D. Performance study of novel photocatalytic reactors for decolourisation of textile dyeing wastewaters. Indian J. Chem. Technol. 2003, 10, 638–643. [Google Scholar]
  125. Dillert, R.; Vollmer, S.; Schober, M.; Theurich, J.; Bahnemann, D.; Arntz, H.J.; Pahlmann, K.; Wienefeld, J.; Schmedding, T.; Sager, G. Photocatalytic treatment of an industrial wastewater in the Double-Skin Sheet Reactor. Chem. Eng. Technol. 1999, 22, 931–934. [Google Scholar] [CrossRef]
  126. Kernani, R.; Mameri, N.; Lounici, H. The use of TiO2 as catalyst in thin film fixed bed reactor for the treatment of landfill water. Mediterr. J. Chem. 2014, 3, 780–788. [Google Scholar] [CrossRef]
  127. Gulyas, H. Solar Heterogeneous Photocatalytic Oxidation for Water and Wastewater Treatment: Problems and Challenges. J. Adv. Chem. Eng. 2014, 4, 1–11. [Google Scholar] [CrossRef]
  128. Maksoud, Y.A.; Imam, E.; Ramadan, A. TiO2 solar photocatalytic reactor systems: Selection of reactor design for scale up and commercialization-Analytical review. Catalysts 2016, 6, 138. [Google Scholar] [CrossRef] [Green Version]
  129. Toor, A.P.; Verma, A.; Jotshi, C.K.; Bajpai, P.K.; Singh, V. Photocatalytic degradation of Direct Yellow 12 dye using UV/TiO2 in a shallow pond slurry reactor. Dye. Pigm. 2006, 68, 53–60. [Google Scholar] [CrossRef]
  130. Rao, N.N.; Chaturvedi, V.; Pum, L.G. Novel pebble bed photocatalytic reactor for solar treatment of textile wastewater. Chem. Eng. J. 2012, 184, 90–97. [Google Scholar] [CrossRef]
  131. Puma, G.L.; Yue, P.L. A novel fountain photocatalytic reactor: Model development and experimental validation. Chem. Eng. Sci. 2001, 56, 2733–2744. [Google Scholar] [CrossRef]
  132. Arzate, S.; Sanchez, A.L.G.; Soriano Molina, P.; Lopez, J.L.C.; Campos-Manos, M.C.; Aguera, A.; Perez, J.A.S. Effect of residence time on micropollutant removal in WWTP secondary effluents by continuous solar photo-Fenton process in raceway pond reactors. Chem. Eng. J. 2017, 316, 1114–1121. [Google Scholar] [CrossRef]
  133. Amiri, H.; Nabizadeh, R.; Martinez, S.S.; Shahtehari, S.J.; Yaghmaeian, K.; Badiei, A.; Nazmara, S.; Nadafi, K. Response surface methodology modeling to improve degradation of Chlorpyrifos in agriculture runoff using TiO2 solar photocatalytic in a raceway pond reactor. Ecotoxicol. Environ. Saf. 2018, 147, 919–925. [Google Scholar] [CrossRef] [PubMed]
  134. Abid, M.F.; Abdulla, O.N.; Kadhim, A.F. Study on removal of phenol from synthetic wastewater using solar photocatalytic reactor. J. King Saud Univ. Eng. Sci. 2017, 31, 131–139. [Google Scholar] [CrossRef]
  135. Almomani, F.; Bhosale, R.; Kumar, A.; Khraisheh, M. Potential use of solar photocatalytic oxidation in removing emerging pharmaceuticals from wastewater: A pilot plant study. Sol. Energy 2018, 172, 128–140. [Google Scholar] [CrossRef]
  136. Chavaco, L.C.; Arcos, C.A.; Prato-Garcia, D. Decolorization of reactive dyes in solar pond reactors: Perspectives and challenges for the textile industry. J. Environ. Manag. 2017, 198, 203–212. [Google Scholar] [CrossRef]
  137. Shet, A.; Shetty, V.K. Solar light mediated photocatalytic degradation of phenol using Ag-core TiO2 shell (Ag@TiO2) nanoparticles in batch and fluidized bed reactor’. Sol. Energy 2016, 127, 67–78. [Google Scholar] [CrossRef]
  138. Sudrajat, H.; Babel, S. A new cost-effective photocatalytic system N-ZnO@polyster fabric for degradation of recalcitrant compound in a continuous flow reactor. Mater. Res. Bull. 2016, 83, 369–378. [Google Scholar] [CrossRef]
  139. Negishi, N.; Sano, T. Photocatalytic Solar Tower Reactor for the Elimination of a Low Concentration of VOCs. Molecules 2014, 19, 16624–16639. [Google Scholar] [CrossRef] [Green Version]
  140. Jovic, M.S.; Dojcinovic, B.P.; Kovacevic, V.V.; Obradovic, B.M.; Kuraica, M.M.; Gasic, U.M.; Roglic, G.M. Effect of different catalysts on mesotrione degradation in water falling film DBD reactor. Chem. Eng. J. 2014, 248, 63–70. [Google Scholar] [CrossRef]
  141. Chan, A.H.C.; Chan, C.K.; Barford, J.P.; Porter, J.F. Solar photocatalytic thin film cascade reactor for treatment of benzoic acid containing wastewater. Water Res. 2003, 37, 1125–1135. [Google Scholar] [CrossRef]
  142. Shankar, M.V.; Anandan, S.; Venkatachalam, N.; Arabindoo, B.; Murugesan, V. Novel thin-film reactor for photocatalytic degradation of pesticides in an aqueous solution. J. Chem. Technol. Biotechnol. 2004, 79, 1279–1285. [Google Scholar] [CrossRef]
  143. Adams, M.; Skillen, N.; McCullagh, C.; Robertson, P.K.J. Development of a doped titania immobilised thin film multi tubular photoreactor. Appl. Catal. B Environ. 2013, 130–131, 99–105. [Google Scholar] [CrossRef]
  144. Velmurugan, R.; Subash, B.; Krishnakumar, B.; Selvam, K.; Swaminathan, M. Solar photocatalytic treatment of gelatine industry effluent: Performance of pilot scale reactor with suspended TiO2 and supported TiO2. Ind. J. Chem. Technol. 2016, 23, 139–145. [Google Scholar]
  145. Khalilian, H.; Behpoura, M.; Atoufa, V.; Hosseini, S.N. Immobilization of S, N-codoped TiO2 nanoparticles on glass beads for photocatalytic degradation of methyl orange by fixed bed photoreactor under visible and sunlight irradiation. Sol. Energy 2015, 112, 239–245. [Google Scholar] [CrossRef]
  146. Miyawaki, A.; Taira, S.; Shiraishi, F. Performance of continuous stirred-tank reactors connected in series as a photocatalytic reactor system. Chem. Eng. J. 2016, 286, 594–601. [Google Scholar] [CrossRef]
  147. Nair, R.G.; Bharadwaj, P.J.; Samdarshi, S.K. Design improvement and performance evaluation of solar photocatalytic reactor for industrial effluent treatment. Ecotoxicol. Environ. Saf. 2016, 134, 301–307. [Google Scholar] [CrossRef]
  148. Sarkar, S.; Bhattacharjee, C.; Sarkar, S. Studies on the performance of annular photoreactor (APR) for pharmaceutical wastewater treatment. J. Water Proc. Eng. 2017, 19, 26–34. [Google Scholar] [CrossRef]
  149. Spiniello, I.; De Carluccio, M.; Castiglione, S.; Amineva, E.; Kostryyukova, N.; Cicatelli, A.; Rizzo, L.; Guarino, F. Landfill leachate treatment by a combination of a multiple plant hybrid constructed wetland system with a solar photo-Fenton process in a raceway pond reactor. J. Environ. Manag. 2023, 331, 117211. [Google Scholar] [CrossRef]
  150. Pandey, A.; Ahmed, S.; Kumar, V.; Singh, P.; Kothari, R. Solar Photocatalytic Treatments of Wastewater and Factors Affecting Mechanism: A Feasible-Low Cost Approach. Emerg. Energy Altern. Sustain. Environ. 2019, 1–28. [Google Scholar] [CrossRef]
  151. Mousset, E.; Loh, W.H.; Lim, W.S.; Jarry, L.; Wang, Z.; Lefebvre, O. Cost comparison of advanced oxidation processesfor wastewater treatment using accumulated oxygen equivalent-criteria. Water Res. 2021, 200, 117234. [Google Scholar] [CrossRef]
  152. Modenes, A.N.; Espinoza-Quiñones, F.R.; Borba, F.H.; Manenti, D.R. Performance evaluation of an integrated photo-Fenon-Electrocoagulation process applied to pollutant removal from tannery effluent in batch system. Chem. Eng. J. 2012, 197, 1–9. [Google Scholar] [CrossRef]
  153. Michael, I.; Panagi, A.; Ioannou, L.A.; Frontistis, Z.; Kassinos, F.D. Utilizing solar energy for the purification of olive mill wastewater using a pilot-scale photocatalytic reactor after coagulation-flocculation. Water Res. 2014, 60, 28–40. [Google Scholar] [CrossRef] [PubMed]
  154. Alalm, M.G.; Tawfik, A.; Ookawara, S. Investigation of optimum conditions and costs estimation for degradation of phenol by solar photo-Fenton process. Appl. Water Sci. 2017, 7, 375–382. [Google Scholar] [CrossRef] [Green Version]
  155. Torres-Socias, E.D.; Rodriguez, L.P.; Zapata, A.; Calderero, I.F.; Oller, I.; Malato, S. Detailed treatement line for a specific landfill leachate treatment remediation: Brief economic assessment. Chem. Eng. J. 2015, 261, 60–66. [Google Scholar] [CrossRef]
  156. Selvabharathi, G.; Kumar, A.S.; Jenefa, S.; Ginni, G.; Banu, R.J.; Yeom, I.T. Combined homogeneous and heterogeneous advanced oxidation process for the treatment of tannery wastewaters. J. Water Reuse Desal. 2016, 6, 59–71. [Google Scholar] [CrossRef] [Green Version]
  157. Benis, K.Z.; Behnami, A.; Aghayani, E.; Farabi, S.; Pourakbar, M. Water recovery and on-site reuse f laundry wastewater by a facile and cost-effective system: Combined biological and advanced oxidation process. Sci. Total Environ. 2021, 789, 148068. [Google Scholar] [CrossRef]
  158. Ahamed, S.F.; Mofijur, M.; Nuzhat, S.; Chowdhury, A.T.; Rafa, N.; Uddin, M.A.; Inayat, A.; Mahlia, T.M.I.; Ong, H.C.; Chia, W.Y.; et al. Recent developments in physical, biological, chemical and hybrid treatments techniques for removing emerging contaminants from wastewater. J. Hazard. Mater. 2021, 416, 125912. [Google Scholar] [CrossRef]
  159. Liu, L.; Chen, Z.; Zhang, J.; Shan, D.; Wu, Y.; Bai, L.; Wang, B. Treatment of Industrial dye wastewater and pharmaceutical residue wastewater by advanced oxidation processes and its combination with nanocatalysts: A. review. J. Water Process Eng. 2021, 42, 102122. [Google Scholar] [CrossRef]
  160. Moravvej, Z.; Soroush, E.; Rahimpour, M.R. Chapter 9-Achievements in hybrid processes for wastewater and water treatment. In Current Trends and Future Developments on (Bio-) Membranes; Elsevier: Amsterdam, The Netherlands, 2020; pp. 239–262. [Google Scholar] [CrossRef]
  161. Mirza, N.R.; Huang, R.; Du, E.; Peng, M.; Pan, Z.; Ding, H.; Shan, G.; Ling, L.; Xie, Z. A review of the textile wastewater treatment techniques with special focus on advanced oxidation processes (AOPs), membrane separation and integrated AOP-membrane processes. Desalin. Water Treat. 2020, 206, 83–107. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a solar photo-Fenton process.
Figure 1. Schematic diagram of a solar photo-Fenton process.
Sustainability 15 04235 g001
Figure 2. Schematic diagram of the solar photocatalytic process.
Figure 2. Schematic diagram of the solar photocatalytic process.
Sustainability 15 04235 g002
Figure 3. Schematic diagram of combined solar photo-Fenton and photocatalytic process.
Figure 3. Schematic diagram of combined solar photo-Fenton and photocatalytic process.
Sustainability 15 04235 g003
Table 1. The oxidation potential of various oxidative species.
Table 1. The oxidation potential of various oxidative species.
S. No.Oxidation Species Oxidation PotentialeV
1Fluorine3.06
2Hypochlorous acid1.49
3Chlorine1.36
4Hydrogen peroxide1.77
5Ozone2.07
6Perhydroxyl radical1.70
7Hydroxyl radical2.80
8Nascent oxygen2.42
Table 3. Treatment of wastewater by various solar photocatalytic processes.
Table 3. Treatment of wastewater by various solar photocatalytic processes.
S. No.WastewaterProcessOperational ConditionsDegradation EfficiencyReference
1Pulp and paper millSolar photocatalytic (TiO2)-COD—57.9%, BOD—42.9%, color—89.2%Kumar et al. (2011) [54]
2AmoxicillinSolar photocatalytic (TiO2)TiO2 catalysts 100 to 750 mg/L93% mineralizationDimitrakopoulou et al. (2012) [55]
3Synthetic wastewaterPhotocatalytic oxidationpH values 2, 4, 6, 8, 10COD—86%Singh et al. (2013) [52]
4TrimethoprimPhotocatalytic oxidation-50% mineralizationRuiz et al. (2013) [57]
5Pulp and paper millPhotocatalytic oxidation-COD—73%, chroma percent—95%Shao et al. (2013) [58]
6MetribuzinPhotocatalytic oxidationTiO2—100 mg/L and
I = 750 W/m2
80% mineralization
Reaction time—300 min
Antonopoulou et al. (2014) [59]
7Organic pollutantsNanosized TiO2 supported on single-wall carbon nanotubes-UV—9 to 87%, simulated solar light—9 to 96%Murgolo et al. (2015) [61]
8Hexavalent chromiumSolar photocatalytic (TiO2)pH = 2, 6 and 10, TiO2—2.0, 2.5, and 3.0 g/LComplete removal for 5 mg/L Cr6+Solano et al. (2018) [62]
9Rhodamine BSolar photocatalytic (TiO2)Film thickness = 361 nm Specific surface area of 47.72 m2 g−198.33% in 30 minYang and Yang (2018) [63]
10Urban wastewaterImmobilized TiO2 High removal (>40%)Rueda-Marquez et al., 2020 [69]
11Petroleum wastewaterTiO2/ZnO/Fenton/Solar and TiO2/ZnO/Air/SolarZnO dosage of 54 g/L and TiO2 dosage 50 g/LCOD—74%, TOC—99%Aljouboury et al., 2021 [64]
12Organic pollutantSolar TiO2 photocatalysis Perfect acid and alkali resistanceGai et al., 2021 [70]
13Methyl orange dyeTiO2-based photocatalystspH—7.0, reaction time—5.5 h of UV irradiation, photocatalyst (0.12 g)Degradation—75.8%Kader et al. (2022) [65]
14Dehalogenation disinfection by-productsSolar TiO2 photocatalysispH 9, TiO2—100 mg/LEffective removal of TOXAbusallout and Hua (2022) [66]
15Pharmaceutical residuesSolar TiO2 photocatalysisTiO2—200 and 300 mg L−1Higher than 56%Rapti et al. (2022) [72]
16Crystal violetTitanium dioxide/graphene aerogel-doped sulfur Enhanced photo-degradationGiang et al., 2023 [71]
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

Gopalakrishnan, G.; Jeyakumar, R.B.; Somanathan, A. Challenges and Emerging Trends in Advanced Oxidation Technologies and Integration of Advanced Oxidation Processes with Biological Processes for Wastewater Treatment. Sustainability 2023, 15, 4235. https://doi.org/10.3390/su15054235

AMA Style

Gopalakrishnan G, Jeyakumar RB, Somanathan A. Challenges and Emerging Trends in Advanced Oxidation Technologies and Integration of Advanced Oxidation Processes with Biological Processes for Wastewater Treatment. Sustainability. 2023; 15(5):4235. https://doi.org/10.3390/su15054235

Chicago/Turabian Style

Gopalakrishnan, Ginni, Rajesh Banu Jeyakumar, and Adishkumar Somanathan. 2023. "Challenges and Emerging Trends in Advanced Oxidation Technologies and Integration of Advanced Oxidation Processes with Biological Processes for Wastewater Treatment" Sustainability 15, no. 5: 4235. https://doi.org/10.3390/su15054235

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