Next Article in Journal
Microwave-Catalyzed Conversion of Phenolic Resin Waste to Activated Carbon and Its Applications for Removing Ammonium from Water
Next Article in Special Issue
Photocatalytic and Sonocatalytic Degradation of EDTA and Rhodamine B over Ti0 and Ti@TiO2 Nanoparticles
Previous Article in Journal
Altering the Stereoselectivity of Whole-Cell Biotransformations via the Physicochemical Parameters Impacting the Processes
Previous Article in Special Issue
Enhancement of Iron-Based Photo-Driven Processes by the Presence of Catechol Moieties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Treatment of Petroleum Refinery Wastewater Using Advanced Oxidation Processes

by
Wamda Faisal Elmobarak
1,
Bassim H. Hameed
1,
Fares Almomani
1,* and
Ahmad Zuhairi Abdullah
2
1
Department of Chemical Engineering, Qatar University, Doha P.O. Box 2713, Qatar
2
School of Chemical Engineering, Engineering Campus, University Sains Malaysia, Nibong Tebal 14300, Pulau Pinang, Malaysia
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(7), 782; https://doi.org/10.3390/catal11070782
Submission received: 2 June 2021 / Revised: 22 June 2021 / Accepted: 23 June 2021 / Published: 27 June 2021
(This article belongs to the Special Issue Trends in Environmental Applications of Advanced Oxidation Processes)

Abstract

:
The petroleum industry is one of the most rapidly developing industries and is projected to grow faster in the coming years. The recent environmental activities and global requirements for cleaner methods are pushing the petroleum refining industries for the use of green techniques and industrial wastewater treatment. Petroleum industry wastewater contains a broad diversity of contaminants such as petroleum hydrocarbons, oil and grease, phenol, ammonia, sulfides, and other organic composites, etc. All of these compounds within discharged water from the petroleum industry exist in an extremely complicated form, which is unsafe for the environment. Conventional treatment systems treating refinery wastewater have shown major drawbacks including low efficiency, high capital and operating cost, and sensitivity to low biodegradability and toxicity. The advanced oxidation process (AOP) method is one of the methods applied for petroleum refinery wastewater treatment. The objective of this work is to review the current application of AOP technologies in the treatment of petroleum industry wastewater. The petroleum wastewater treatment using AOP methods includes Fenton and photo-Fenton, H2O2/UV, photocatalysis, ozonation, and biological processes. This review reports that the treatment efficiencies strongly depend on the chosen AOP type, the physical and chemical properties of target contaminants, and the operating conditions. It is reported that other mechanisms, as well as hydroxyl radical oxidation, might occur throughout the AOP treatment and donate to the decrease in target contaminants. Mainly, the recent advances in the AOP treatment of petroleum wastewater are discussed. Moreover, the review identifies scientific literature on knowledge gaps, and future research ways are provided to assess the effects of these technologies in the treatment of petroleum wastewater.

Graphical Abstract

1. Introduction

Water is crucial for life since it is a fundamental necessity of all organisms. Fast industrial and economic growth have guided a rapid increase in population and development [1]. The world is observing growth and development in the industry, owing to the several industrial methods that are applied [2]. The industrial regions incessantly produce significant quantities of wastewater at highs rates and generally dispose of the wastewater with no appropriate management and treatment [3,4]. Petroleum refineries and industries are crucial from an economic development perspective [5]. Petroleum refineries are meeting the difficulty of how to dispose this wastewater.
Wastewater produced by petroleum industries includes various kinds of organic and inorganic contaminants, for example sulfides, phenol, BTEX, hydrocarbons, heavy metals, etc. [6]. Huge amounts of harmful materials are produced through petroleum industry activities, for example through the oil production process, oil refinery, transportation, storage, etc., which are all unsafe for the environment and human health [6]. The treatment of wastewater produced from the petroleum industry includes different processes that comprise physical, chemical, and biological methods [7]; however, the majority of these techniques are ideal to match certain treatment requirements for every application and generally do not recommend the treatment to separate the various groups of compounds [8].
The chemical oxidation method is one of the techniques used for wastewater treatment, and it can be classified into two types: conventional chemical treatments and advanced oxidation processes [9]. Advanced oxidation methods are highly efficient techniques needed for the treatment of several wastewater types, including the oil and gas industry, pharmaceutical industries wastewaters, etc. In previous years, several works have been carried out to examine the efficiency of advanced oxidation processes in the treatment of different wastewater types that contain recalcitrant and toxic pollutants [10]. AOPs are defined as the methods that depend on the production of hydroxyl free radicals, which have great electrochemical oxidant power and strong oxidizing potential [11,12]. The great oxidizing potential allows them to easily offense and degrade nearly all organic compounds to H2O, carbon dioxide (CO2), and inorganic ions, through dehydrogenation or hydroxylation. In general, AOPs are extensively employed in petroleum wastewater treatment for decreasing organic materials, the removal of certain contaminants, and sludge treatment [13]. The possibility of AOPs being used for wastewater treatment is improved by their capability to generate hydroxyl radical (OH•) via various methods.
The very common AOP method types studied for wastewater treatment involve heterogeneous photocatalytic oxidation [14], ozonation processes (O3) [15], Fenton and photo-Fenton reaction [16,17], and electrochemical oxidation [18]. These methods are favorable in the treatment of wastewater containing different organic and inorganic pollutants. According to previous reviews, several combinations of various AOPs are more effective in comparison to a single oxidation process, owing to the production of additional radicals and excessive energy efficiency. Moreover, the utilization of heterogeneous and homogenous catalysts and energy-dispersing elements considerably improves the degradation of the wastewater. Wang et al. [19] summarized hydrogen peroxide (H2O2), O3, ultraviolet (UV)/H2O2 or O3, O3/H2O2, and photo-Fenton processes and their use in wastewater treatment. Furthermore, Chen et al. [20] and Ani et al. [21] reported a review on the basics of the heterogeneous photocatalytic degradation of organic pollutants and stated the benefits of utilizing titanium as an effective material for the removal of organic contaminants by applying advanced oxidation processes. Azizah and Widiasa [22] reviewed the application of H2O2/UV and H2O2/UV/O3 advanced oxidation process methods in synthetic residual fluid catalytic cracking (RFCC) wastewater to verify the suitable condition of AOPs to greet the phenol separation ranges. It is found that a high phenol degradation of 93.75% was achieved by using the H2O2/UV/O3 AOP process with a H2O2 amount of 1000 ppm; this clarified that the H2O2/UV method requires extended time and greater H2O2 amounts for the removal of phenols from these wastewater types to levels below the target. Lofrano et al. [23] propose a generally updated review of AOPs in antibiotics removal considering physio-chemical and toxicity implications. They discussed the technical and economic problems, and the environmental compatibility of effluents.
An environmentally friendly preparation of TiO2 nanoparticles supported by microwaves from titanium tetrachloride and water was carried out by de Oliveira et al. for the treatment of petroleum refinery effluent [24]. The highest rates of TOC separation were achieved below pH conditions of 10 and a catalyst dosage of 100 mg/Lin 90 min of reaction time. The catalyst demonstrated steadiness for four cycles. The attained data revealed that the produced catalyst is a favorable material for the separation of remaining recalcitrant organic material from petroleum refinery effluent.
As far as the authors are aware, there are different studies on the treatment of petroleum refinery wastewater using the different types of an advanced oxidation process method as a tertiary treatment to disinfect and degrade the wastewater compounds and reuse the treated wastewater for different purposes. Therefore, the objective of the review is to present the different methods of the advanced oxidation process and its use in petroleum refinery wastewater treatment. The review reports that the treatment efficiencies depended strongly on the selected AOP method type, whether the properties of target contaminants were physical or chemical, and the operating conditions. Moreover, the knowledge gaps and future research methods are also provided to assess the effects of these technologies in the treatment of petroleum wastewater.

2. Methodology

The review provides an overview of the findings from recent studies from 2017 to 2021 which have been applied to different types of AOPs to degrade and remove different compounds from petroleum refinery wastewater. Few articles published from 2012–2015 were also considered as they presented crucial data. The major source of these articles was google-scholar and Scopus databases; the keywords applied for the search were “advanced oxidation”, “petroleum”, “refinery”, “wastewater pollutants”, “Fenton”, “catalyst”, and “review”. The assembled data were introduced in Tables 1–5 and studied in four sections. Information on the characteristics of petroleum refinery wastewater and the environmental impact of petroleum refinery wastewater was presented in Tables 1 and 2, and discussed in Section 3 and Section 4, respectively. The fundamentals of AOPs were discussed in Section 5, an overview of the process in petroleum refinery wastewater treatment was represented by Figure 1, and the classification of the AOPs was presented in Table 3. In this section (Section 5) four types of AOPs including H2O2/UV, Fenton, photo-Fenton, and ozonation AOPs and their applications in petroleum refinery wastewater treatment were discussed. The data relating to each process was presented in Tables 4 and 5, and Figures 2–6 were included to present crucial concepts related to each process (the reaction mechanism, the drawbacks of Fenton, and photo-Fenton processes, solar photoreactor setup, experimental setup of photoreactor, and detailed steps of the mechanism of organic degradation of wastewater contaminants through O3/UV) [25,26]. In Section 6, the integrated AOPs were reviewed, and a diagram of the integrated system was presented in Figure 7. Finally, from the analysis of the assembled data, conclusions are presented and recommendations for additional research are provided.

3. Characteristics of Petroleum Refinery Wastewater

Petroleum refinery effluent has numerous amounts of complex substances present in it. They were present in the form of oil, gas, wax, grease, metals, minerals, hydrocarbons, etc. Few components present in very low amounts; such components are N, S, O, and a few other metals which have minor importance. Aromatic organic compounds containing solely carbon and hydrogen atoms are colorless, white, or pale-yellow solids. It has been proven that, when it enters the human body, our potential immune system will suppress it. This can cause carcinogenic and mutagenic damage [27].
The constituents present in the produced include oil and gases and different organic and inorganic matters that are complex in nature. The particles implicated are hydrocarbons, dissolved oil, dissolved salts [28], naturally formed radioactive elements, and heavy metals [29]. Furthermore, with gases such as hydrogen sulfides, carbon dioxide composition will be encompassed as dissolved gases. The generation of produced water is strictly controlled by different countries because the components present in it will increase the pollution rate [30]. The concentration of salt present in seawater is lower than compared to the concentration of salt present in the produced water. The percentage of salinity is enhanced due to the concentration of the additional component present in it. Those components are Ca, K, Mg, and dissolved Cl, etc. [31].
The previous survey [31] reveals that the electrical conductivity of the produced water lies in a range from 200 µs/cm to 300 µs/cm, while the TDS concentration range of produced water lies in a range from 120 g/L to 140 g/L. Some of the major compositions are sodium with a range of (42,720 ± 2093 mg/L), calcium ions (4247 ± 752 mg/L), magnesium ions (727 ± 54 mg/L), potassium ion (805 ± 230 mg/L), strontium ions (257 ± 20 mg/L), chlorine ions (65,800 ±1600 mg/L), sulphate (1010 ± 9 mg/L), bromide ion (591 ± 16 mg/L) and silicon dioxide (32 ± 2 mg/L) [27]. The concentration of ammonia present in the PW has a range of (655 ± 77 mg/L) [31]. The value of pH, as well as alkalinity, are 7.30 ± 0.21 and 2345 ± 329 mg/L, respectively, such as CaCO3 [32]. Table 1 shows the composition present in the petroleum refinery wastewater [33]. The composition of nitrogen in the produced water coming from the oil platform is 50 v/v%. The summation of all nitrogen compounds in the water is termed total nitrogen; it contains ammonia nitrogen (NH3-N), nitrite nitrogen (NO2-N), nitrate nitrogen (NO3-N), and organically bonded nitrogen [34].
The configuration of these substances and their tenacious character affects the rate of chemical oxygen demand and the level of toxicity. The wastewater encompasses a high level of COD. The presence of these toxic oils considerably disperses the chemical oxidation demand [32]. If the produced water contains a high content of BOD value, it is due to the large content of organic materials present in the drilling fluids. Before discharging a high value of BOD content water into the water bodies, oxidation of the water must be maintained so that only it can resist the wastewater with a larger BOD content [29]. The PW containing the range of the TOC is (0–1500 mg/L). The naturally formed water has a TOC concentration that lies in the range of <0.01–11,000 mg/L [35]. The World Health Organization proposes that a fringe of the permittable volume of benzene and toluene in drinking water are 10 and 700 mg/L, respectively, while the permissible proportion of ethylbenzene and xylene is 300 mg/L and 500 mg/L, respectively [27].
The characteristic of the petroleum refinery wastewater fluctuates according to their base. The constitution of the wastewater depends on the final product induced, the undertaking process, and the type of crude oil that is utilized. The produced water is assorted with aromatic and aliphatic hydrocarbons, MTBE, PAHs, phenols, naphthenic acid, sulfides, metal products as well as benzene, toluene, ethylbenzene and xylene (BTEX) that usually possess a high toxicity. In addition, xylene has three more topical isomers named o-xylene, m-xylene, and p-xylene [36].

4. Environmental Impact of Petroleum Refinery Wastewater

Enormous amounts of contaminants were released by petroleum products in the environment [37]. As known, petroleum wastewater is the main of three waste types (wastewater, waste gas, and industrial residue) whose harmful impacts cannot be misjudged [38]. Wastewater discharged by petroleum industries include significant amounts of heavy metals, petroleum hydrocarbons [39], polyaromatic hydrocarbons [40], benzene, toluene, ethylbenzene, xylene (BTEX), phenolic compounds, and other contaminated materials [41]. Some of the major petroleum wastewater contaminants and their environmental impacts are presented in Table 2. Owing to the incompetence of treatment methods, petroleum refinery wastewater is dangerous to the environment and other life systems which could become mutagenic and toxic to human beings [42]. This wastewater can influence various elements of the environment, for example clean water, human health, the different sources of groundwater, air, marine life, crop manufacturing, etc. [43]. An increase in contaminants in the water bodies raises significant concerns, either long-term or short-term, for the ecology and living organisms, which can be severe and chronic [44]. Different studies revealed that petroleum refinery wastewater composites have the ability to biologically build up, which is the main reason to create a method for the separation of pollutants from water before the release [39]. The following problems were created when a huge quantity of petroleum oily wastewater was arbitrarily released into water bodies [7].
Regarding aquatic life where the wastewater was released into the river and oceans, oil and other contaminants create a film on the water surface that lead to a block in the air and sunlight, thus causing a shortage of oxygen for the aquatic organisms and affecting the plants and animals’ growth [45]. Furthermore, owing to the existence of heavy oil on the surface of seawater, serious problems were observed, including slow growth of the head of aquatic organisms, unusual neural growth in the embryo of the aquatic organisms, and other issues [46]. Benzene, toluene, ethylbenzene, and xylene (BTEX) are some of the main pollutants that harm aquatic organisms, the nervous system, genotoxicity, and the respiratory system [46], due to their highly carcinogenic nature as distinguished by The U.S. Environmental Protection Agency (USEPA) [42]. Moreover, petroleum refinery wastewater includes a solvable aromatic hydrocarbon, many radioactive substances, and heavy metals which can generate negative impacts on some organisms accountable for water purification and biological treatment [36]. In the case of direct release, pollutants in the wastewater are transported by the food chain through the infected foods, and, hence, human health will be critically affected [47]. Aquatic contamination will also occur, owing to the spread of toxic components between the food chain via the marine ecosystem from the influenced marine environment [48]. Some studies reported that the consumption of the contaminated oil with marine food could produce many toxic consequences on human health, such as skin tumors due to the existence of the contaminated hydrocarbons in the polluted wastewater [49]. Another study presented the effect of wastewater pollutants during the separation processes of oil throughout the marine water [40]. It was found that the co-workers that engaged in certain processes developed a urinary infection due to the presence of a high concentration of volatile organic composites and polycyclic aromatic hydrocarbons which can be absorbed by their body. Furthermore, it was found that contaminated oil in the wastewater represents a toxic material to certain microorganisms that is accountable for the biological treatment of wastewater [50].
Phenols are other toxic compounds that generally appear in the wastewater produced from petroleum refinery and can impact the enzymatic processes of marine microorganisms even at small phenol concentrations. As the degradation of the phenolic composite takes a long time, they will gather in the aquatic organisms’ tissues and consequently produce biomagnification. The phenolic compounds have been identified to be genetically, physiologically, and haematologically toxic [51]. Thus, the phenolic compounds have been classified by USEPA as a significant contaminant [52]. The phytotoxicity of phenols was experimentally verified on EU cyclops agilis by Kottuparambil et al. [53]. The results showed the high toxicity of phenol compounds even in the presence of a small dose of phenols and how this negatively impacts human life [54].
Polycyclic aromatic hydrocarbon (PAHs) compounds are other toxic contaminants that existed in the petroleum industry wastewater and have toxic effects on aquatic organisms. The study reported by Badibostan et al. [55] showed that the presence of specific polycyclic aromatic hydrocarbons, for example chrysene (CHR) compounds and benzo[b]fluoranthene (BbF) with high concentrations, causes serious problems including drinking, inhalation, etc. When its overall concentration exceeded the concentrations of the standards defined by USEPA for toxic chemicals, it negatively impacted human life and health.
Another concern is that, when petroleum oily wastewater used for irrigation, these compounds cab have harmful impacts on the soil and make it stronger [56]. This change on soil texture inhibits the breathing of plants and organisms as the soil microorganisms have a high sensitivity to MTBE compared to the microorganisms present in the water. In addition, oily wastewater could have a harmful effect on environmental processes such as nutrient cycles and biodegradation [56]. Consequently, the microorganism’s declination rate will decrease, and then the plants and animal growth will be influenced [57].

5. Advanced Oxidation Process (AOPs) for Petroleum Refinery Wastewater

5.1. Fundamentals/Chemistry of the Advanced Oxidation Process

Advanced oxidation processes (AOPs) are the methods defined through the creation of hydroxyl radicals; the core purpose of the advanced oxidation process is to produce highly reactive free radicals which are extremely sensitive, and to apply non-selective materials to reduce contaminated organic composites that exist in a medium for example wastewater, soil, marine mediums, etc. [58].
Hydroxyl radicals are efficient in eliminating organic chemicals since they are reactive electrophiles [59]. They are powerful oxidizing agents to demolish compounds that cannot be oxidized by the standard oxidant. Hydroxyl radicals have a more rapid oxidation rate compared to KMnO4 or H2O2 [60]. The produced hydroxyl radicals can remove the organic chemicals through different ways such as radical addition, hydrogen extraction, or electron transmission [61]. The creation of hydroxyl radicals is generally enhanced by merging O3, H2O2, TiO2, UV radiation, ultrasound, and electron-beam irradiation [62].
The hydroxyl radical has great oxidation potential (E0 = 2.8 V) [63] and is competent to react with almost all types of organic composites, leading to full mineralization of these composites which form from carbon dioxide (CO2), inorganic salts, and water, or their transformation into fewer destructive substances [64]. The removal of contaminants and the inhibition of the production of toxic compounds are some of the major advantages of AOPs since traditional techniques of water treatment. For example filtration, adsorption, and flotation are not effective in completely separating the organic pollutants (non-destructive physical separation procedures), which only eliminate the pollutants, transporting them to other stages, and thus producing intense residues [65]. An overview of the AOP in petroleum wastewater treatment is presented in Figure 1.
The main reaction produce hydroxyl radicals (HO•) that attack organic composites in electrophilic position leading to degradation. In addition HO• reaction with organic composites (aromatic or unsaturated) that include a π bond result in the formation of the organic radicals (Equation (1)) [66]. This includes hydrogen removal through the reaction of the HO• radicals with a saturated aliphatic compound (Equation (2)) and electron transmission by decreasing the HO• radical into a hydroxyl anion through an organic substrate (Equation (3)) [67]. The attack of the HO• radical on organic substrates could be affected by the existence of different chemical types in water (which are created in the mineralization method), for example, bicarbonate and carbonate ions [68]. These ions can react with the HO• radicals (Equations (4) and (5)), and, by consequently vying with the organic substrates via the HO• radicals. HO•, radicals can be created from numerous AOPs, allowing the use of a superior method for every treatment condition.
Catalysts 11 00782 i001
HO + R H R + H 2 O
HO + R X   [ R X ] + + HO
HO +   HCO 3 CO 3   + H 2 O
HO +   CO 3 2 CO 3 HO
Following Flouret and his colleagues (2018), the AOP can be classed as homogeneous and heterogeneous [69]. They introduced this categorization in terms of the light type that is utilized in the method. Table 3 shows the categorization of AOPs consistent with these studies. The AOP are costly methods due to the high cost of chemicals applied (e.g., H2O2 and O3), also due of electric energy for UV radiation. Hence, they are studied in other treatments of wastewater that cannot be organically handled. An additional substantial part is the weight of contaminants that appear in the wastewater, commonly identified according to the existing amounts of COD. The application of these methods is revealed solely for wastewater that contains COD under 5 g/L, since a greater index of COD inhibiting the treatment would need an extremely higher utilization of chemicals [70]. For high organic wastewaters, pretreatment processes, such as dilution, coagulation, and flocculation, are needed to lessen the initial concentrations [71]. Additionally, AOPs can be utilized for treating the polluted waters containing very small organic contents (in ppb) which include dispersed organic composites that are hard to eliminate. One method of decreasing the AOPs’ costs is to utilize them as initial treatment processes to decrease harmfulness, supported by biological treatment. This alternative has demonstrated to be very exciting from an economic perspective and has been examined by numerous studies [72,73].

5.2. H2O2/UV Advanced Oxidation Process and Their Application in Petroleum Refinery Wastewater

Hydrogen peroxide (H2O2) is a powerful oxidant among other oxidant types with a standard reduction potential = 1.77 V that is used to decrease the small levels of contaminants that exist in wastewaters [74]. However, the specific utilization of H2O2 is not effective in treating more compounds and mutinous substances with a low reaction rate. The application of H2O2 is more successful when it performs in combination with other components or sources of energy that are competent of separating it to produce HO• radicals, which can behave as oxidizing factors [75]. Through UV radiation in wavelengths >300 nm, hydrogen peroxide (H2O2) can decompose and create HO•, as demonstrated in Equation (6).
H 2 O 2 hv 2 HO
H2O2 can also react with HO• and created transitional products in line with the reaction, the method is explained easily by Equations (7)–(11) [76].
H 2 O 2 + HO HO 2 + H 2 O
H 2 O 2 + HO 2   HO + H 2 O + O 2
2 HO H 2 O 2
2 HO 2 H 2 O 2 + O 2
HO + HO 2   H 2 O 2 + O 2
The damage on organic composites occurs when HO• and hydroperoxyl (HO2•) radicals are created. Nevertheless, the hydroperoxyl radicals have a smaller reduction potential with a value of 1.7 V compared to hydroxyl radicals with a value of 2.8 V; as a result, their production is not attractive to the method. Numerous researchers have indicated that the increase in the first amount of H2O2 enhances the pollutants degradation rate up to the highest value, after which, when they achieve very high H2O2 levels, they start to decrease. Khan and coworkers (2020) ascribe this reduction in the H2O2/UV process outputs to hydroxyl radicals reacting with additional H2O2 (Equation (7)), rather than reacting by the organic substrates, resulting in the creation of the HO2•. The H2O2 photolysis is generally achieved using different mercury vapor lights (low or medium pressure) with great intensity, which decreases the required quantity of H2O2. Almost 50% of the energy expended is wasted in heat form or in emissions of wavelengths >185 nm, which are noticed by the jacket made of quartz [77].
A cheaper and widely used alternative is the germicide lamp; though, the efficiency is smaller as it radiates in a range from 210 to 240 nm, and H2O2 absorption achieves a peak of 220 nm [78]. The application of the H2O2/UV process has different advantages that can be ascribed to different facts, for example: the H2O2 reagent is completely solvable in water, there is no restriction of mass transfer, it is an active supply of HO•, and there is no necessity for a removal method after the treatment [79]. The operating pH should be small (pH < 4) to void the impact of separating radical types, mainly ionic types such as bicarbonate and carbonate ions, which results in improving the degradation rate.
The applications of H2O2 was evaluated in the existence and the nonappearance of UV radiation, and in the degradation of petroleum refinery wastewater which is pretreated using different separation processes such as coagulation and flotation, studying total petroleum hydrocarbons (TPH), 1,2-dichloroethane (DCE), dichloromethane (DCM), and MTBE [80]. It was remarked that UV emission did not substantially impact the compound degradation, except for DCM which was highly stable compared to others. Overall, 83% was eliminated with 11.76 nM of H2O2 in the existence of UV radiation. The studies showed sluggish degradation of the TPH, which separated by 69% in the first 8 days of testing. They found that the degradation of MTBE was identical to that of DCM, and that total degradation occurred after 24 h. The application of the H2O2/UV process for the purification of oily wastewater from a lubricant production unit was analyzed by Gilpavas et al. (2019). Chromatographic characterizations that were carried out in the study indicated that most of the composites detected in the wastewater were separated with the ending of the treatment, utilizing great amounts of H2O2. However, a decrease in just 40% of preliminary wastewater COD was achieved, which is equivalent to around 9000 mg/L. This was assigned to the creation of organic acids, generated because of the decomposition of the organic compounds that exist in the wastewater, which are more renitent to photo-degradation with H2O2. The influence of pH on the process was also evaluated and it was found that, in an acidic condition of pH 3.5, there was a greater decrease in the wastewater COD, in comparison to basic and neutral pH conditions [81]. The degradation of MTBE was investigated by Neisi and coworkers (2018), and the gasoline was evaluated as one of the highly popular pollutants in polluted underground waters. The H2O2/UV process was evaluated at certain conditions, and the removal percentage of MTBE was found to be more than 98%. A rise in the removal percentage was obtained with an increase in the concentration of H2O2, after which the percentage began to drop [82]. This result was assigned to the competitive reactions of H2O2 and HO•, presented in Equations (7)–(11). A comparison between some studies applying H2O2 advanced oxidation-based process in the treatment of different types of wastewater was illustrated in Table 4.
Table 4 shows a comparison between recent studies for wastewater treatment using H2O2 advanced oxidation-based processes at different conditions. The studies showed the high efficiency of the H2O2-based advanced oxidation process in the degradation and removal of different wastewater compounds, such as COD, TOC, MTBE, phenol, etc., from different wastewater types. The efficiency was found to be more than 90% for each wastewater that either utilized the hydrogen peroxide H2O2 alone or added in the existence of ultraviolet (UV) light.

5.3. Fenton and Photo-Fenton Advanced Oxidation Processes and Their Application in Petroleum Refinery Wastewater

Fenton’s chemical is a mixture of H2O2 and ferric ions found by Henry John Horstman Fenton (1854–1929). In 1894, he stated that this mixture had great oxidizing potency in an acid medium [83]. However, the use of Fenton’s chemical in oxidation procedures to remove contaminated organic composites was reported by different researchers [84,85]. There is significant disagreement about the mechanism of reaction, including Fenton’s reaction. The standard reaction of the Fenton process is taken by Haber and Weiss et al. (1934), as mentioned by Zouanti et al. [86], which contains an aqueous integration of hydrogen peroxide (H2O2) and ferrous ions (Fe2+) in an acidic medium, leading to H2O2 breaking into a hydroxyl ion and a hydroxyl radical, and the oxidation from Fe2+ to Fe3+, as characterized by Equation (12) [87].
Fe 2 + + H 2 O 2   Fe 3 + + HO + HO
The Fe2+ and Fe3+ types exhibited in an easy manner coincide with the [Fe(OH)(H2O)5]+ aqueous compounds, which, combined with H2O2, create the [Fe(OH)(H2O2)(H2O)4]+ and [Fe(OH)(H2O)5]2+ composites, respectively, as stated by [88]. The Fe3+ created in Equation (12) are able to react with the H2O2 that exists in the medium, and can be decreased to Fe2+ once more by producing the hydroperoxyl radical, according to Equation (13). This reaction happens more sluggishly than reaction 12, as described by Kim et al. [89]. The Fe3+ ions also react with the HO2 and are decreased to Fe+2, as indicated with Equation (14).
Fe 3 + + H 2 O 2 [ Fe . . OOH ] 2 + + H +   Fe 2 + + HO 2
Fe 3 + + HO 2   Fe 2 + + O 2 + H +
The preliminary rate of degradation employing Fe+3 is significantly smaller than that utilizing Fe+2, as presented by Yuan et al. [90]. The standard pH in Fenton’s reaction observed in several reports is 3 [91,92] and, consequently, can be considered the suitable operational pH. At an extremely low pH (<2.5), the creation of compounds, for example [Fe (H2O)6]2+, occurs. These react more sluggishly with H2O2 than [Fe (OH) (H2O)5]+, generating a smaller volume of hydroxyl radicals and consequently reducing the efficiency of the system [93]. In the case of basic pH conditions, the reaction between the iron and hydroxide ions (HO) produced the iron hydroxide (Fe (OH)2 or Fe (OH)3 which does not react with H2O2, thus prohibiting the Fenton reaction. Hence, a change in pH is necessary for wastewater treatment before the addition of Fenton reagents. Previously, it is crucial to determine the ideal correlation between H2O2 ions and ferrous ions. Once the ferrous ions were applied with higher concentrations more than hydrogen peroxide ions, the HO• created by Equation (12) can react with additional ferrous ions following Equation (15), and then reduce the raid of HO• on organic reacted material [94].
Fe 2 + + HO   Fe 3 + + HO
Wang et al. (2019) noticed that, while the ratio of [Fe2+]/[H2O2] is 1:2 and has a greater degradation rate, it is generally suggested to utilize a ratio of 1:5, which generates identical findings and needs fewer reagents. In general, only the reaction of the Fenton process is not able to depredate organic composites and mineralize them effectively. The reaction of the Fenton method is efficient up to the instant where each Fe2+ that exists in the reaction is reacted to Fe3+, hence disrupting the production of hydroxyl radicals and subsequently degrading the organic compound [95]. Wang et al. [89] mentioned that one of the parameters that have a great effect in disturbing the Fenton’s reaction under normal conditions with the degradation products Fe3+ ions can produce steady organic composites, specifically the organic acids existing in the reaction [91], as revealed by Equation (16).
Fe 3 + + RCO 2 2 + Fe 3 +   ( RCO 2 ) 2 +
While Fenton’s reaction has been extensively reviewed and has presented great results in the wastewater treatment process, the reaction of Fenton was identified as a strong means to reduce organic composites when UV radiation is applied to the system. This was accomplished in the initial 1990s when used for treating wastewaters that include polluted carbon-based contaminants. It has acted as an initial step for different researchers associated with the use of the photo-Fenton method [96,97]. It has observed that UV irradiation greatly increases the organic contaminants degradation rate from Fenton’s chemical, which has the feature of being susceptible to UV-Vis emission for wavelengths over 300 nm. Below these situations, the Fe3+ compounds photolysis allows the existence of Fenton’s reaction and the Fe2+ to be reused many times if H2O2 is presented. Clarizia et al. [98] stated that, in a pH range from 2.5 to 5, the ferric compound Fe(OH)2+ is the main type in the photo-Fenton method, a shortened form of demonstrating the aqueous composite Fe(OH) (H2O5)2+. The photolysis of this compound at wavelengths less than 410 nm is found to be the major HO• radical source (Equation (17)) [98]. Additional photoreactive classes also exist in the medium of the reaction, for example Fe2(OH2)4+ and Fe (OH)2+, which could lead to the creation of hydroxyl radicals [99].
Fe   ( OH ) 2 +   hv     Fe 2 + + HO
The regeneration of Fe2+ can also happen with the produced compounds photolysis (Equation (16)) from Fe3+ and the organic compounds created in the degradation method, as demonstrated by Equation (18) [89].
Fe 3 + + RCO 2 2 + hv     Fe 2 + + CO 2 + R
Hence, the ferrous ions reformed by Equations (17) and (18) can react once again with the H2O2 in solution (Equation (12)) to produce more hydroxyl radicals, forming a photocatalytic cycle in the Fe2+/Fe3+ system. The utilization of radiation considerably lowers the essential number of ferrous ions, in comparison to the dim Fenton reaction (lack of light). The number of ferrous ions applied in the Fenton’s reaction must be reduced to attain good efficiency, because adding high concentrations of ferrous ions rises the opacity of the solution, prohibiting the radiation diffusion and the Fe2+ production when oxidized to Fe3+. As a result, the rate of degradation is reduced [100]. It is also found that the elevated intensity of light dispersion and the strong connection between the contaminant and the oxidizing factor act as benefits of the photo-Fenton method because it is a consistent method [101,102]. Figure 2 shows the reaction mechanism for the photo-Fenton process.
The drawbacks of the Fenton and photo-Fenton processes are related to need for pH values which are too small (usually less than 2) and the need to eliminate the iron after reaction [103]. The solution pH in Fenton and photo-Fenton processes should be shifted to protect the catalyst stability, as at pH 6 iron hydroxide is normally generated. For various materials, the proper pH for the Fenton reaction is between 2.8 and 3, and the ideal ratio between catalyst and peroxide is generally 1: 5 wt/wt. At an operational pH of >3 (acidic pH), the decomposition rate declines as a result of the reduction in free iron types in the solution, possibly owing to the creation of Fe(II) complexes with the barrier hindering the production of free radicals. At pH greater than 3, Fe3+ begins precipitating as ferric oxyhydroxides and breaks down the H2O2 into O2 and H2O [104], consequently decreasing the production of ferrous ions. Moreover, the OH radical oxidation potential is known to reduce with a rise in pH [105]. Nevertheless, at low pH (pH = 2.5), the creation of Fe(II) (H2O)2+ happens, which reacts more slowly with hydrogen peroxide, generating a reduced volume of reactive hydroxyl radicals and consequently decreasing the degradation efficiency [75]. The main drawbacks for Fenton and photo-Fenton processes are summarized in Figure 3.
As stated by Rajala et al., [106] iron separation could not be required if it is applied with amounts under the removal range determined by regulation. It is crucial to have the previous familiarity of the original wastewater’s physical and chemical features (untreated wastewater). Using this method, since some materials or inorganic ions for example Cl, (SO4)2−, [(H2PO4)/(HPO4)2−] exist in the wastewater or are added as reagents such as (FeSO4, FeCl3, HCl, H2SO4), may negatively affect the Fenton and photo-Fenton system reaction mechanism, thus hindering the degradation method [107]. The main reason for this is the scavenging effect of these ion (Fe2+ or Fe3+ or any other inorganic ions) to OH radicals. It was highlighted that these reactions with OH could create new radicals such as (Cl•−, Cl2•− and SO4•−), which are of lo oxidation potential [108].
Owing to the existence of Cl• − ions in the reaction, the production of Cl2• − radical anions can be evaded by monitoring the pH throughout the reaction time at a value of 3, which indicates that the degradation process of organic substrate thru the photo-Fenton method aid in the formation of the acid consequently reduces the pH. This happens due to pH reduction (smaller than 2.5) in the existence of Cl• − ions, resulting in more severe production of ferric chloride compounds (FeCl2+ and FeCl2+). These components also endure photolysis, thus reducing the volume of Fe (OH)+2, which is the major supply of OH in the photo-Fenton method. Furthermore the creation of the Cl2• − radical anion can react with Fe2+ (reacting to Fe3+ with no production of OH) and the organic substrate. One of the major benefits of utilizing the photo-Fenton method in comparison to the other oxidation methods is the application of solar radiation in its reaction process [109]. The existence of oxalate ions (C2O4)2− and the Fe3+ present in the reaction medium can produce the [Fe(C2O4)]+ compound (Equation (19)) with light absorption at wavelengths lower than 570 nm [110,111]. The solar photoreactor setup is shown in Figure 4 by Pourehie and Saien [111].
2 [ Fe ( C 2 O 4 ) ] +   hv   2 Fe 2 + + C 2 O 4 2 + 2 CO 2
The ability to utilize solar radiation represents massive conservation from the energetic perspective, given that there is no necessity to utilize lamps. The progress of pilot plants for treating the wastewater that utilizes solar reactors used to the AOPs has been one of the major aims of study in the field [13,103].
The application of Fenton and photo-Fenton processes in the wastewater treatment process was studied by Giwa et al. [102] who use ferrioxalate as an iron source to analyze and reduce various wastewaters components, including BTEX and pentachlorophenol. Applying the photo-Fenton method has contrasted with additional photodegradation methods. Depending on the utilization of energy, the outcomes revealed that the method was above 30 times more effective than the UV/H2O2 and UV/Fe2+/H2O2 methods. Therefore, the authors determined that the method utilizing ferrioxalate needs less electric energy compared to the UV/H2O2 process, leading to a decrease in the treatment costs.
The photo-Fenton method was applied to reduce the existing hydrocarbons in the produced wastewater synthesized using gasoline oil to mimic the water generated from oil industries in oil production fields that include dissolved salts with great amounts. It was noticed that the existence of NaCl (salts) prohibit the total degradation of contaminants. This was attained in the assessments without adding any amount of salt to the prepared wastewater [112]. The researchers assigned this impact to Fe3+ ions that reacted with Cl ions producing FeCl2+ composites, which, in the existence of UV radiation, can provide an increase to Cl2• − radicals with smaller oxidation potential compared to HO• radicals, thus decreasing the productivity of the method. Fenton and photo-Fenton methods oxidation potential were examined to reduce different contaminants in the solution, including BTX and water polluted with gasoline including ethanol with a percentage of 25%. The degradation of prompt BTX composites was noted, along with the creation of transitional phenolic complexes in the primary reaction times that were degraded with 30 min of reaction. About 75% of total hydrocarbons introduced in gasoline-polluted water were removed and degraded, and the BTX composites were eliminated in the initial reaction mins (when phenolic composites were produced), after which they degraded sluggishly [113].
The impact of the photo-Fenton process in degrading phenol composites was examined by Brillas et al. 2020, which is an extremely popular contaminant in petrochemical production utilizing black light fluorescent lights as a source of UV-A emission. These are not expensive and are more efficient from the energy use perspective in comparison to medium-pressure mercury vapor lights. The effectiveness of black light fluorescent lamps was verified when compared to the degradation of phenol applying dark Fenton and photo-Fenton methods. The rate of degradation was twice as high when applying the photo-Fenton method. The degradation rate of phenol improved when the intensity of the radiation emission source increased, due to the increase in Fe2+ ion production (Equation (17)) [114]. Thus, this then increased the number of HO• radicals that were produced. These studies showed that the analysis of this factor is crucial as it facilitates the construction of the reactor that utilizes lamps as a source of UV radiation, which is essential for continuous wastewater treatment. This cannot be achieved by solar reactors only, which indicates that they are just used in cases of sunlight availability.
The application of the photo-Fenton method for the treatment of wastewater including diesel oil was examined by Hassan and his colleagues (2020). They performed research to enhance the number of chemicals utilized and to attain high degradation rates using a lesser number of chemicals. It was found that, by utilizing a small concentration of 0.1 mM Fe2+ ion, the rate of degradation of the organic composites existing in the wastewater was 90% classified as total organic carbon (TOC) [115]. The photo-Fenton method needs to be efficient from the commercial approach below these operating conditions, as it does not need successive treatment to eliminate the remaining iron. Besides, the degraded acid that exists in the wastewater from oil processing plants includes emulsified oil and a sequence of small-biodegradability micro contaminants, for example sulfides, phenols, ammonia, and mercaptans using different advanced oxidation process (AOP) methods.
Among the methods studied, Fenton and photo-Fenton achieved the most significant findings. The reaction of the dark Fenton process was applied until reaction stagnation with a separation efficiency for initial dissolved organic carbon (DOC) of only 27%, supported with the photo-Fenton method using UV radiation. The use of both methods together utilized constantly decreased the preliminary amount of DOC with a high percentage (more than 90%). The BTEX contaminants existing in the initial wastewater were eliminated to non-measurable amounts. The studies confirmed that the two methods were effective in eliminating contaminants from wastewater, but by utilizing high amounts of chemicals. They confirm that there is a need for optimizing the processes to decrease the number of chemicals used and thus increasing the efficiency of the process [116].
The productivity of Fenton chemicals in mineralizing organic complexes existing in water polluted with crude oil was also examined. The process was evaluated by observing the H2O2 using spectrophotometric analysis and via the CO2 titrimetric quantification which allowed researchers to determine the level of mineralized TOC. These assessments indicated a 75% decrease in the initial TOC volume of water polluted by petroleum oil [117]. The efficiencies of the Fenton and photo-Fenton methods in the treatment of petroleum wastewater are summarized in Table 5.
Table 5 shows the efficiencies of the Fenton and photo-Fenton process in the treatment of petroleum refinery wastewater and the separation of different contaminates including TOC and COD compounds. The studies showed that, at different conditions including the pH, reaction time, and H2O2 concentration, Fe+2 concentrations under both processes were effective in eliminating contaminants from the wastewater but utilizing high amounts of reagents.

5.4. Ozone Based AOPs and Their Application in Petroleum Wastewater Treatment

The use of ozone (O3) has been extensively applied to eliminate undesirable organic pollutants in both purification of drinking water and wastewater treatment [118]. Ozone (O3) is a selective oxidant that reacts with electron-rich organic compounds at a high reaction rate constant (kO3, pH7 > 105 M−1 s−1) [119,120]. Hydroxyl radical (OH•) produced from the consumption of O3 is a nonselective oxidant, which quickly reacts with several organic composites at closely controlled rates of diffusion [121].
The creation of radicals, for example the superoxide radical (O2−•) and hydroxyl radical (HO•), happens when ozone is combined with water, via a complicated series of reactions. The ozone degeneration rate in water is improved at a greater pH [122]. The reaction integration with HO• and molecular ozone produces the oxidation of the organic compound.
3 O 3 + OH + H +   2 HO + 4 O 2
The ozone decomposition cycle can be improved by adding hydrogen peroxide (H2O2) to cause the creation of HO• [123]. H2O2, to some extent, separates in water to generate the hydroperoxide ion. These HO2- ions react with the O3, promptly generating the HO•. In brief, the reactions integration results in the next equation:
2 O 3 + H 2 O 2   2 HO + 3 O 2
Different parameters determine the performance of the process including the contact time, water alkalinity, and ozone concentrations. By presenting hydrogen peroxide (H2O2) when the highly reactive materials oxidized with O3, it improves the degradation rate. The ozonation method using H2O2/O3 systems was found to be the greatest proper AOP method in the purification of water in comparison to other technologies of AOPs. Therefore, there is a discipline-verified operation history in the application of H2O2/O3 techniques. Another approach to accelerate ozonation is by utilizing catalysts (homogenous or heterogeneous). Research has been conducted with several metal ions, and metal oxides have demonstrated considerable variations in decomposition, but just in certain situations. The amount of ozone in the inlet gas and the ozone process time have been deemed as major parameters to develop the degradation rate. The research of oxidation has been conducted on numerous organic pollutants utilizing TiO2, Fe2O3, Mn2+, MnO2, Fe2+, and Fe3+. The system of ozone/catalyst was found to be more efficient for the COD and TOC reduction, and ozone oxidation only at a high pH [124]. In the tests conducted with [125,126], it was noticed that the chemical decomposition is successfully enhanced with the synergistic impact of photocatalytic reactions happening on TiO2-photocatalyst. The experimental setup of the photoreactor is shown in Figure 5 [126].
The ozone process can also be improved through the decomposition of ozone with direct absorption via UV radiation (λ = 254 nm) which produces H2O2 as a transitional and then decomposes to HO• radicals. The reactions shown in Equations (22) and (23) demonstrate the photolysis of ozone to produce H2O2, which leads to the creation of a very reactive HO• radical [127]. Though numerous oxidation methods can damage and eliminate organic pollutants, HO• is the main separation method.
O 3 + H 2 O + h υ   O 2 + H 2 O 2   at   ( h υ :   λ < 300   nm )
2 O 3 + H 2 O 2   2 HO + 3 O 2
This technique includes all the mechanisms of organic degradation via O3/UV, H2O2/O3 and H2O2/UV, O3/Fenton, catalytic ozonation (O3/Metal oxide), and photocatalytic ozonation, etc. [125]. The detailed steps of the mechanism of organic degradation of wastewater contaminants through O3/UV are described in Figure 6. It is more efficient to separate organic materials than applying UV or O3 separately, and it is an additional effective means of producing HO• in comparison to the UV/H2O2 technique at the same oxidant dosages [128,129]. However, this procedure is costly and energy-intensive [130]. Moreover, ozone techniques have the possibility to create bromate [131,132]. Inhibiting composites can influence the absorption of UV radiation, therefore decreasing the effectiveness of the method. Nevertheless, this difficulty will be lessened by increasing the ratio between the H2O2 and O3, decreasing the pH, or increasing the hydroxyl radical amount [133,134]. In addition, it is found that the absorption of UV radiation by O3 was affected by the turbidity. The ozone production energy costs are the major operating expenses for this procedure. The existence of carbonate or bicarbonate may behave as hunters of the hydroxyl radical in normal structures which is an additional problem of these methods [135]. The usage of AOPs with ozone has more industrial uses in water treatment in comparison to all other AOPs. The method of ozone has been utilized to handle the petroleum refineries wastewater pollution.
A study conducted by [136] on the biochemical wastewater sewage treatment plant uses a combination of O3/UV advanced oxidation process. The authors showed that the major parameters affecting the efficiency of the UV/O3 process were the efficient O3 distribution means, monitoring of the comparative pressure in the reactor, the ratio of ozone added, and UV radiation intensity. Cruz and his workers named the ozone concentration that consumes reactions with the initiator’s composites as the ozone dose line, beyond which the addition of H2O2 rises the rate of HO• production. The authors noticed that, in the case of H2O2 addition, ozonation does not have much effect on the exposure of HO• until the concentrations become more than the threshold ozone dosages used [137,138]. It was noted that the usage of a homogeneous reactor allows the quick and high formation of the radical species directly in the bulk liquid without any movement and mass transfer limits in comparison to a classical gas–liquid reactor, whereby the oxidation of the contaminants through hydroxyl radicals could be restricted via their transportation from the liquid phase to the interface as these radicals are extensively reactive and do not disperse in the solution [139]. The study carried out by Meshref et al. [140] on the identification and monitoring of the volatile organic compounds (VOCs) presents that the effluent’s wastewater showed 43% and 34% of COD and BOD reduction, respectively, resulting in an effective AOP process for VOCs degradation [141]. It also showed that, at certain ozone concentrations between 0.5–1.5 mg O3/mg dissolved organic carbon, the efficiencies of removing the contaminants increased up to ~14–18% in groundwater, above 6–10% in surface water, and insignificantly in secondary effluent through the O3/H2O2 treatment in comparison to the conventional ozonation method.

6. Integrated AOP Processes

In recent years, AOP methods have been extensively used for the treatment of petroleum wastewater applying different types of AOP methods. A combination of AOPs methods and additional conventional methods for wastewater treatment have been confirmed to be more efficient for the treatment of contaminated industrial wastewater sources efficiently. The selection means of these methods rely on the waste stream properties, environmental policies, and the cost.
AOPs were integrated with biological methods and other different processes to increase the efficiency of the AOP process in the degradation and separation of contaminants. Biological treatment methods include the decomposition of every residual oil, the degradation of organic composites, the breaking of organic contaminants, the removal of trace metals and nutrients, etc. [4,142]. Various methods of biological treatment have been effective in treating petroleum refinery wastewater. Between them, the activated sludge process (ASP) is the most biological of treatment methods used [143]. Ebrahimi et al. and Jain et al. [31,144] noticed a COD separation effectiveness ranging from 70–80% through the treatment of petroleum refinery wastewater utilizing ASP integrated with the AOP methods. A comparatively greater COD separation efficiency of 96% was achieved by Mirbagheri et al. [145], which was reported for the great synthetic aeration generated in the method. The integration of binary or extra-biological methods or the combination of the biological method with membrane-based AOP techniques were also found to be an efficient method for wastewater treatment. The study carried out by [146] combined an expanded granular sludge bed bioreactor (EGSB-BR) with an ASP-based AOP for petroleum wastewater treatment, with a COD of 4600–5300 mg/L, which achieved a decrease in COD of 85% after a 62.8 h retention time. An application of biological-AOP was carried out by Razavi et al. [147], utilizing hollow fiber membrane bioreactors (HF-MBR) for the treatment of refinery effluent and attaining a COD separation of 82% after a 36 h retention time. The research reported by El-Naas et al. [148] utilized the anaerobic baffled reactor (ABR) for the treatment of produced water including heavy oil by applying a biological-based advanced oxidation process. The removal efficiency of 65% and 88% were then attained for heavy oil and oil and grease, respectively. The study reported by [149] examined the feasibility of combined electrocoagulation and membrane process. The combination of electrocoagulation and membrane filtration methods was used with integrated intelligent automation and process optimization in the structure of a mobile treatment platform (Figure 7). Up to 95% of oil, TOC, COD, TSS, and turbidity in produced water could be removed. The system could also effectively remove about 90% of ions, including chloride, sulfate, sodium, calcium, manganese, etc.
However, most of these procedures generate a huge quantity of sludge, trained staff, and frequent maintenance. Certain methods also require extremely high hydraulic retention time (HRT), which is an essential drawback to the Bio-AOP techniques.
Another integrated advanced oxidation process was reported by [150]. They joined photo-bioreactors (PBR) with conventional oxic/anoxic methods for the treatment of petrochemical wastewater with an insignificant COD concentration of 312.8 mg/L and attained a separation rate of 71% using an electrocoagulation-based AOP method. A high concentration of petrochemical water with COD values in the range 3600–5300 mg/L was reduced and a separation rate of 97% was achieved in a spouted bed bioreactor (SBBR) with a packed activated carbon treatment (PACT) and an electrocoagulation cell (EC). There are several studies in integrating other AOPs, for example integration of photocatalysis and ultrasound [151], ozonation and ultrasound integration [152], the integration of Fenton, and photo-Fenton processes with ultrasound [153]. Fernandes et al. [151] showed that the integration of photocatalytic and electrochemical processes achieved a high degradation rate compared to the photocatalytic AOP process alone. The major factor in the integration methods to assess the efficiency of the system is the synergetic impact. The synergetic impact is a factor that indicates the development in the degradation of organic compounds below the integrated process. In general, the studies found that the integration of advanced oxidation technology with other methods is simple, effective, and economic for the treatment of wastewater, except when employing biological methods. The biological process is uncommon due to the creation of hydroxyl radicals throughout the AOPs which can negatively affect the biomass. Furthermore, the existence of H2O2 is also toxic to microorganisms. Consequently, it is better to utilize the integrated procedure in the AOP part to improve biodegradability and oxidation in a short time.
AOP-bioremediation is a new type of integrated method applied for industrial wastewater to separate micropollutants. The AOP-bioremediation integrated process concentrates on the treatment of petroleum industrial wastewater including bio-resistant pollutants. The combined two-step process has been studied, partially owing to the capability of AOPs for the degradation of these toxic pollutants to biodegradable compounds [103]. However, the recent literature from the last two years has indicated that the feasible application of the integrated process might be difficult, owing to the variety of potential effluent properties, operating conditions, and design structures. Consequently, the integrated process of AOPs with bioremediation has been recently studied for wastewater treatment [154].
AOP methods were also combined with the membrane (membrane-based AOP) as an innovative technique for the degradation of pollutants in the wastewater. This method is strongly related to the photocatalytic process which utilizes semiconductor catalysts [155,156]. A membrane gives an extra advantage of filtration. [157] This explains the mechanism of the process where the nanoparticle catalysts are inserted and dispersed in the reactor suspension in the form of soluble particles, and where the membrane performs as a barrier for the photocatalyst particles, i.e., the particles are settled inside or on membrane surface. The combination of photocatalyst nanoparticles is predictable to develop the quality of the permeate via the reduction oxidation process which generated active radicals to degrade the pollutants.

7. Knowledge Gaps and Future Perspectives

Hydrocarbon existing in petroleum are categorized as important contaminants. The contaminants which are introduced in wastewater of the petroleum industry can be efficiently removed by utilizing various techniques. Studies on the review of new techniques with the least environmental and economic impact showed that it is a research focus field. This review concentrates on reviewing AOP methods applied for petroleum wastewater treatment produced with the activities of the petroleum industry. The wastewater produced by the petroleum industry includes several harmful materials, for example xylene, toluene, ethylbenzene, benzene, phenols, polycyclic aromatic hydrocarbons (PAHs), etc. It is extremely challenging to immediately separate the contaminants through applying a single technique, especially biological treatment, which is deemed as a green method. Therefore, advanced oxidation process treatments in integration with other treatment techniques are needed. It has been described that the combination of different methods could provide greater outcomes more than a single process utilized for oily wastewater treatment. However, understanding the concept of technology combination is still at its initial stages which require to be investigated by scientists. The latest improvements in the process are mainly more costly, and need maintenance and long time. Future techniques require simple operation techniques which are appropriate for the petroleum industry and other different industries. A massive volume of solid and liquid waste is generated, owing to the activities of the petroleum industry. Controlling and treating the produced waste is currently a significant problem in any country. Owing to the increase in wastewater production, appropriate removal, treatment, and recycling are presenting extra problems, as treatment and disposal are extremely expensive. In addition, waste recovery is developing as a focus research field since it presents sustainability of the environment and social possibilities. Several studies are concentrating on the recovery of several supplies produced by human activities, for example energy, bio-products, metals, and oil from wastewater. Significant pollutants are introduced in petroleum industrial wastewaters, which can be considered as resources after recovery. Effective resource recovery and recycle can generate green life, while, on the other hand, it can help in developing economy through decreasing waste and enhance environmental health and recovery cost. Consequently, there is a necessity to recover and recycle the produced waste from petroleum industry activities in an effective way. Viable treatment methods to generate less contaminated products generate a new approach for the sustainability of the environment and economy. To enhance the development of the petroleum wastewater industry, there is a necessity to create sustainable techniques such as AOP processes. The development of AOP methods can open up new chances for the market and attain effective application of resources.
Moreover, these technologies are favorable and valuable methods where toxic compounds are efficiently degraded and separated from the wastewater. Therefore, these methods were assessed at a pilot scale. Different combination with other types of studied AOP were studied at highly recyclable catalysts to evaluate the treatment efficiency of the wastewater before discharged with no hazardous impacts to the environment. In addition to the future studies on this method at pilot scale treatment unit, the studies must also contain the application of accessible AOPs based on modified catalysts, for example titanium dioxide, with an improvement in the process efficiency in visible light, following that a remarkable decrease in the process costs related to the energy consumption. However, the difference between the catalyst costs (modified catalysts) and energy costs for other types of AOP methods stays a wide-open case. AOP methods give a green, economic, and efficient methodology for the deep degradation of contaminants from petroleum refinery wastewater. Other future research viewpoints would be targeted near the use of some combined methods with AOP methods, such as fluidized-bed Fenton, fixed bed Fenton processes, etc., which study the competence of electro-Fenton processes for treating petroleum wastewater and identify the reaction kinetics and mechanisms that utilize newly developed catalysts as an effort to enhance the total oxidation productivity.

8. Conclusions

Petroleum wastewaters can be handled via the oxidation of chemical and biological treatment methods. There are various harmful components introduced in wastewater produced through the petroleum industry activities. An effective treatment approach should be designed to consider the environment and human health. AOP is one of the techniques which is getting a lot of worldwide concern for cleaning petroleum contaminants such as hydrocarbons. Nevertheless, the combination of different methods provide greater findings more than a single procedure utilized for the treatment of wastewater. The present wastewater treatment methods were concentrated to eliminate the contaminants from wastewaters to match the wastewater release criteria. To settle the risks related to the components of petroleum, an appropriate technique that handles the waste and removes those contaminants from wastewater would be a favorable choice. Recent studies related to the application of advanced oxidation processes (AOPs) in petroleum wastewater treatment that use the latest advanced oxidation types either separately or by combination with revere to wastewater characteristics for generating non-toxic by-products were presented. A highly comprehensive literature review was presented for comparing the different AOP technologies described in this review, and a discussion about their advantages and disadvantages was also highlighted.

Author Contributions

Data curation, W.F.E.; Formal analysis, F.A.; Investigation, B.H.H. and A.Z.A.; Methodology, W.F.E. and B.H.H.; Project administration, F.A. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was made possible by the grant provided by Qatar University, International Research Collaboration Co-Fund, Grant number IRCC-2020-011. The statements made herein are solely the responsibility of the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviation

AOPAdvanced oxidation process
BTEXBenzene, Toluene, Ethylbenzene Xylene
RFCCResidual Fluid Catalytic Cracking
UVUltraviolet
BODBiological oxygen demand
CODChemical oxygen demand
PAHsPolycyclic aromatic hydrocarbons
SBBRSpouted bed bioreactor
PACTPacked activated carbon
ECElectrocoagulation cell
ABRAnaerobic baffled reactor
ASPActivated Sludge Process
HF-MBRHollow fiber membrane bioreactors
PBRPhoto-bioreactors
EGSB-BRExpanded granular sludge bed bioreactor
VOCsVolatile Organic Compounds
TOCTotal organic carbon
O3Ozone
OH•Hydroxyl radical
O2−Superoxide radical
H2O2Hydrogen peroxide
MnO2Manganese dioxide
Fe2O3Iron III oxide
TiO2Titanium dioxide
Fe2+Ferrous iron
Fe3+Ferric iron
Mn2+Manganese
MinMinute
ppmPart per million
RefReference
DOCDissolved Organic Compounds
C2O42−Oxalate ions
MTBEMethyl tert-butyl ether
mg/Lmilligram/litter
g/lgram/litter
TPHTotal petroleum hydrocarbon
PpbParts per billion
E0Oxidation potential
CHRChrysene
BbFBenzo[b]fluoranthene
USEPAUnited States Environmental Protection Agency
VFAVolatile fatty acids
TDSTotal dissolved solids
TSSTotal suspended solids
DODissolved oxygen
HEMHexane Extractable Material
MBASMethylene Blue Active Substance
APIAmerican Petroleum Institute
O & Goil and grease
hrhour

References

  1. Chen, Y.-C. Evaluating greenhouse gas emissions and energy recovery from municipal and industrial solid waste using waste-to-energy technology. J. Clean. Prod. 2018, 192, 262–269. [Google Scholar] [CrossRef]
  2. Zhang, X.; He, W.; Ren, L.; Stager, J.; Evans, P.J.; Logan, B.E. COD removal characteristics in air-cathode microbial fuel cells. Bioresour. Technol. 2015, 176, 23–31. [Google Scholar] [CrossRef] [PubMed]
  3. Varjani, S.; Kumar, G.; Rene, E.R. Developments in biochar application for pesticide remediation: Current knowledge and future research directions. J. Environ. Manag. 2019, 232, 505–513. [Google Scholar] [CrossRef]
  4. Jafarinejad, S.; Jiang, S.C. Current technologies and future directions for treating petroleum refineries and petrochemical plants (PRPP) wastewaters. J. Environ. Chem. Eng. 2019, 7, 103326. [Google Scholar] [CrossRef]
  5. Abdulredha, M.M.; Aslina, H.S.; Luqman, C.A. Overview on petroleum emulsions, formation, influence and demulsification treatment techniques. Arab. J. Chem. 2020, 13, 3403–3428. [Google Scholar] [CrossRef]
  6. Raza, W.; Lee, J.; Raza, N.; Luo, Y.; Kim, K.-H.; Yang, J. Removal of phenolic compounds from industrial waste water based on membrane-based technologies. J. Ind. Eng. Chem. 2019, 71, 1–18. [Google Scholar] [CrossRef]
  7. Varjani, S.J. Microbial degradation of petroleum hydrocarbons. Bioresour. Technol. 2017, 223, 277–286. [Google Scholar] [CrossRef]
  8. Jiménez, S.; Andreozzi, M.; Micó, M.M.; Álvarez, M.G.; Contreras, S. Produced water treatment by advanced oxidation processes. Sci. Total Environ. 2019, 666, 12–21. [Google Scholar] [CrossRef]
  9. Almomani, F.A.; Shawaqfah, M.; Bhosale, R.R.; Kumar, A. Removal of emerging pharmaceuticals from wastewater by ozone-based advanced oxidation processes. Environ. Prog. Sustain. Energy 2016, 35, 982–995. [Google Scholar] [CrossRef]
  10. Almomani, F.; Bhosale, R.; Kumar, A.; Khraisheh, M. Potential use of solar photocatalytic oxidation in removing emerging pharmaceuticals from wastewater: A pilot plant study. Solar Energy 2018, 172, 128–140. [Google Scholar] [CrossRef]
  11. Al Mayyahi, A.; Al-Asadi, H.A.A. Advanced oxidation processes (AOPs) for wastewater treatment and reuse: A brief review. Asian J. Appl. Sci. Technol. 2018, 2, 18–30. [Google Scholar]
  12. Kilic, M.Y.; Abdelraheem, W.H.; He, X.; Kestioglu, K.; Dionysiou, D.D. Photochemical treatment of tyrosol, a model phenolic compound present in olive mill wastewater, by hydroxyl and sulfate radical-based advanced oxidation processes (AOPs). J. Hazard. Mater. 2019, 367, 734–742. [Google Scholar] [CrossRef]
  13. Garrido-Cardenas, J.A.; Esteban-García, B.; Agüera, A.; Sánchez-Pérez, J.A.; Manzano-Agugliaro, F. Wastewater treatment by advanced oxidation process and their worldwide research trends. Int. J. Environ. Res. Public Health 2020, 17, 170. [Google Scholar] [CrossRef] [Green Version]
  14. Lei, J.; Alvarado, J.; Li, H.; Zhu, X.; Tian, Y.; Liang, L. The Advanced Oxidation Processes of Oilfield Wastewater: A Review. In Proceedings of the 4th International Conference on Water Resource and Environment (WRE 2018), Kaohsiung City, Taiwan, 17–21 July 2018. [Google Scholar]
  15. Ikehata, K.; Li, Y. Ozone-based processes. In Advanced Oxidation Processes for Waste Water Treatment; Elsevier: Amsterdam, The Netherlands, 2018; pp. 115–134. [Google Scholar]
  16. Silva, P.C.; Ferraz, N.P.; Perpetuo, E.A.; Asencios, Y.J.O. Oil produced water treatment using advanced oxidative processes: Heterogeneous-photocatalysis and photo-Fenton. J. Sediment. Environ. 2019, 4, 99–107. [Google Scholar] [CrossRef]
  17. Gaurav, G.K.; Yadav, D. Probing the excellence of wastewater PAHs removal approaches: A critical review. J. Contam. Hydrol. 2020, 236, 103715. [Google Scholar] [CrossRef]
  18. 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. 2019, 376, 120597. [Google Scholar] [CrossRef]
  19. Wang, B.; Zhang, H.; Wang, F.; Xiong, X.; Tian, K.; Sun, Y.; Yu, T. Application of heterogeneous catalytic ozonation for refractory organics in wastewater. Catalysts 2019, 9, 241. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  21. Ani, I.; Akpan, U.; Olutoye, M.; Hameed, B. Photocatalytic degradation of pollutants in petroleum refinery wastewater by TiO2-and ZnO-based photocatalysts: Recent development. J. Clean. Prod. 2018, 205, 930–954. [Google Scholar] [CrossRef]
  22. Azizah, A.N.; Widiasa, I.N. Advanced Oxidation Processes (AOPs) for Refinery Wastewater Treatment Contains High Phenol Concentration; MATEC Web of Conferences, EDP Sciences: Les Ulis, France, 2018; p. 03012. [Google Scholar]
  23. Lofrano, G.; Pedrazzani, R.; Libralato, G.; Carotenuto, M. Advanced oxidation processes for antibiotics removal: A review. Curr. Org. Chem. 2017, 21, 1054–1067. [Google Scholar] [CrossRef]
  24. de Oliveira, C.P.M.; Viana, M.M.; Amaral, M.C.S. Coupling photocatalytic degradation using a green TiO2 catalyst to membrane bioreactor for petroleum refinery wastewater reclamation. J. Water Process. Eng. 2020, 34, 101093. [Google Scholar] [CrossRef]
  25. Al Momani, F.A.; Shawaqfeh, A.T.; Mo‘ayyed, S.S. Solar wastewater treatment plant for aqueous solution of pesticide. Solar Energy 2007, 81, 1213–1218. [Google Scholar] [CrossRef]
  26. Al Momani, F.; Gonzalez, O.; Sans, C.; Esplugas, S. Combining photo-Fenton process with biological sequencing batch reactor for 2, 4-dichlorophenol degradation. Water Sci. Technol. 2004, 49, 293–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Mohammadi, L.; Rahdar, A.; Bazrafshan, E.; Dahmardeh, H.; Susan, M.; Hasan, A.B.; Kyzas, G.Z. Petroleum Hydrocarbon Removal from Wastewaters: A Review. Processes 2020, 8, 447. [Google Scholar] [CrossRef]
  28. Jiménez, S.; Micó, M.; Arnaldos, M.; Medina, F.; Contreras, S. State of the art of produced water treatment. Chemosphere 2018, 192, 186–208. [Google Scholar] [CrossRef]
  29. Lusinier, N.; Seyssiecq, I.; Sambusiti, C.; Jacob, M.; Lesage, N.; Roche, N. A comparative study of conventional activated sludge and fixed bed hybrid biological reactor for oilfield produced water treatment: Influence of hydraulic retention time. Chem. Eng. J. 2020, 420, 127611. [Google Scholar] [CrossRef]
  30. Ye, H.; Liu, B.; Wang, Q.; How, Z.T.; Zhan, Y.; Chelme-Ayala, P.; Guo, S.; El-Din, M.G.; Chen, C. Comprehensive chemical analysis and characterization of heavy oil electric desalting wastewaters in petroleum refineries. Sci. Total Environ. 2020, 724, 138117. [Google Scholar] [CrossRef]
  31. Jain, M.; Majumder, A.; Ghosal, P.S.; Gupta, A.K. A review on treatment of petroleum refinery and petrochemical plant wastewater: A special emphasis on constructed wetlands. J. Environ. Manag. 2020, 272, 111057. [Google Scholar] [CrossRef]
  32. Rodriguez, A.Z.; Wang, H.; Hu, L.; Zhang, Y.; Xu, P. Treatment of Produced Water in the Permian Basin for Hydraulic Fracturing: Comparison of Different Coagulation Processes and Innovative Filter Media. Water 2020, 12, 770. [Google Scholar] [CrossRef] [Green Version]
  33. Olajire, A.A. Recent advances on the treatment technology of oil and gas produced water for sustainable energy industry-Mechanistic aspects and process chemistry perspectives. Chem. Eng. J. Adv. 2020, 4, 100049. [Google Scholar] [CrossRef]
  34. Sun, Y.; Liu, Y.; Chen, J.; Huang, Y.; Lu, H.; Yuan, W.; Yang, Q.; Hu, J.; Yu, B.; Wang, D. Physical pretreatment of petroleum refinery wastewater instead of chemicals addition for collaborative removal of oil and suspended solids. J. Clean. Prod. 2021, 278, 123821. [Google Scholar] [CrossRef]
  35. Nanda, S.; Berruti, F. A technical review of bioenergy and resource recovery from municipal solid waste. J. Hazard. Mater. 2020, 403, 123970. [Google Scholar] [CrossRef] [PubMed]
  36. Bustillo-Lecompte, C.F.; Kakar, D.; Mehrvar, M. Photochemical treatment of benzene, toluene, ethylbenzene, and xylenes (BTEX) in aqueous solutions using advanced oxidation processes: Towards a cleaner production in the petroleum refining and petrochemical industries. J. Clean. Prod. 2018, 186, 609–617. [Google Scholar] [CrossRef]
  37. Kuppusamy, S.; Maddela, N.R.; Megharaj, M.; Venkateswarlu, K. An Overview of Total Petroleum Hydrocarbons. In Total Petroleum Hydrocarbons; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–27. [Google Scholar]
  38. Thakur, C.; Srivastava, V.C.; Mall, I.D.; Hiwarkar, A.D. Mechanistic Study and Multi-Response Optimization of the Electrochemical Treatment of Petroleum Refinery Wastewater. CLEAN–Soil Air Water 2018, 46, 1700624. [Google Scholar] [CrossRef]
  39. Al-Hawash, A.B.; Dragh, M.A.; Li, S.; Alhujaily, A.; Abbood, H.A.; Zhang, X.; Ma, F. Principles of microbial degradation of petroleum hydrocarbons in the environment. Egypt. J. Aquat. Res. 2018, 44, 71–76. [Google Scholar] [CrossRef]
  40. Putatunda, S.; Bhattacharya, S.; Sen, D.; Bhattacharjee, C. A review on the application of different treatment processes for emulsified oily wastewater. Int. J. Environ. Sci. Technol. 2019, 16, 2525–2536. [Google Scholar] [CrossRef]
  41. Varjani, S.J.; Joshi, R.R.; Kumar, P.S.; Srivastava, V.K.; Kumar, V.; Banerjee, C.; Kumar, R.P. Polycyclic aromatic hydrocarbons from petroleum oil industry activities: Effect on human health and their biodegradation. In Waste Bioremediation; Springer: Berlin/Heidelberg, Germany, 2018; pp. 185–199. [Google Scholar]
  42. Varjani, S.J.; Gnansounou, E.; Pandey, A. Comprehensive review on toxicity of persistent organic pollutants from petroleum refinery waste and their degradation by microorganisms. Chemosphere 2017, 188, 280–291. [Google Scholar] [CrossRef] [PubMed]
  43. Zainab, S.M.; Junaid, M.; Xu, N.; Malik, R.N. Antibiotics and antibiotic resistant genes (ARGs) in groundwater: A global review on dissemination, sources, interactions, environmental and human health risks. Water Res. 2020, 187, 116455. [Google Scholar] [CrossRef]
  44. Chowdhary, P.; Bharagava, R.N.; Mishra, S.; Khan, N. Role of industries in water scarcity and its adverse effects on environment and human health. In Environmental Concerns and Sustainable Development; Springer: Berlin/Heidelberg, Germany, 2020; pp. 235–256. [Google Scholar]
  45. Schweitzer, L.; Noblet, J. Water contamination and pollution. In Green Chemistry; Elsevier: Amsterdam, The Netherlands, 2018; pp. 261–290. [Google Scholar]
  46. Armstrong, T.; Khursigara, A.J.; Killen, S.S.; Fearnley, H.; Parsons, K.J.; Esbaugh, A.J. Oil exposure alters social group cohesion in fish. Sci. Rep. 2019, 9, 1–9. [Google Scholar]
  47. Sharma, S.; Chatterjee, S. Microplastic pollution, a threat to marine ecosystem and human health: A short review. Environ. Sci. Pollut. Res. 2017, 24, 21530–21547. [Google Scholar] [CrossRef] [PubMed]
  48. Mojiri, A.; Zhou, J.L.; Ohashi, A.; Ozaki, N.; Kindaichi, T. Comprehensive review of polycyclic aromatic hydrocarbons in water sources, their effects and treatments. Sci. Total Environ. 2019, 696, 133971. [Google Scholar] [CrossRef] [PubMed]
  49. Li, J.; Luo, C.; Song, M.; Dai, Q.; Jiang, L.; Zhang, D.; Zhang, G. Biodegradation of phenanthrene in polycyclic aromatic hydrocarbon-contaminated wastewater revealed by coupling cultivation-dependent and-independent approaches. Environ. Sci. Technol. 2017, 51, 3391–3401. [Google Scholar] [CrossRef] [PubMed]
  50. Zango, Z.U.; Sambudi, N.S.; Jumbri, K.; Ramli, A.; Abu Bakar, N.H.H.; Saad, B.; Rozaini, M.N.H.; Isiyaka, H.A.; Osman, A.M.; Sulieman, A. An Overview and Evaluation of Highly Porous Adsorbent Materials for Polycyclic Aromatic Hydrocarbons and Phenols Removal from Wastewater. Water 2020, 12, 2921. [Google Scholar] [CrossRef]
  51. Mustapha, H.I.; van Bruggen, H.J.J.A.; Lens, P.N. Vertical subsurface flow constructed wetlands for the removal of petroleum contaminants from secondary refinery effluent at the Kaduna refining plant (Kaduna, Nigeria). Environ. Sci. Pollut. Res. 2018, 25, 30451–30462. [Google Scholar] [CrossRef]
  52. Saleem, H.; Rehman, K.; Arslan, M.; Afzal, M. Enhanced degradation of phenol in floating treatment wetlands by plant-bacterial synergism. Int. J. Phytoremediation 2018, 20, 692–698. [Google Scholar] [CrossRef]
  53. Kottuparambil, S.; Kim, Y.-J.; Choi, H.; Kim, M.-S.; Park, A.; Park, J.; Shin, W.; Han, T. A rapid phenol toxicity test based on photosynthesis and movement of the freshwater flagellate, Euglena agilis Carter. Aquat. Toxicol. 2014, 155, 9–14. [Google Scholar] [CrossRef] [PubMed]
  54. Li, J.; Li, F. Polycyclic aromatic hydrocarbons in the Yellow River estuary: Levels, sources and toxic potency assessment. Mar. Pollut. Bull. 2017, 116, 479–487. [Google Scholar] [CrossRef]
  55. Badibostan, H.; Feizy, J.; Daraei, B.; Shoeibi, S.; Rajabnejad, S.H.; Asili, J.; Taghizadeh, S.F.; Giesy, J.P.; Karimi, G. Polycyclic aromatic hydrocarbons in infant formulae, follow-on formulae, and baby foods in Iran: An assessment of risk. Food Chem. Toxicol. 2019, 131, 110640. [Google Scholar] [CrossRef]
  56. Shaikh, S.S.; Abu-Dieyeh, M.H.; Al Naemi, F.A.; Ahmed, T.; Al-Ghouti, M.A. environmental impact of utilization of “produced water” from oil and gas operations in turfgrass systems. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef]
  57. Njoku, L.; Okporuanefe, F.; Ude, E. Responses of Accessions of Zea Mays to Crude Oil Pollution Using Growth Indices and Enzyme Activities As Markers; University of Lagos: Lagos, Nigeria, 2018. [Google Scholar]
  58. Sathasivam, M.; Shanmugapriya, S.; Yogeshwaran, V.; Priya, A. Industrial waste water treatment using advanced oxidation process–A review. Int. J. Eng. Adv. Technol. 2019, 8, 485–488. [Google Scholar]
  59. Cao, J.; Yang, J.; Yue, K.; Wang, Z. Preparation of modified citrus pectin (MCP) using an advanced oxidation process with hydroxyl radicals generated by UV-H2O2. Food Hydrocoll. 2020, 102, 105587. [Google Scholar] [CrossRef]
  60. Li, D.; Sun, T.; Wang, L.; Wang, N. Enhanced electro-catalytic generation of hydrogen peroxide and hydroxyl radical for degradation of phenol wastewater using MnO2/Nano-G|Foam-Ni/Pd composite cathode. Electrochim. Acta 2018, 282, 416–426. [Google Scholar] [CrossRef]
  61. Yang, Y.; Hoffmann, M.R. Synthesis and stabilization of blue-black TiO2 nanotube arrays for electrochemical oxidant generation and wastewater treatment. Environ. Sci. Technol. 2016, 50, 11888–11894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Rostam, A.B.; Taghizadeh, M. Advanced oxidation processes integrated by membrane reactors and bioreactors for various wastewater treatments: A critical review. J. Environ. Chem. Eng. 2020, 8, 104566. [Google Scholar] [CrossRef]
  63. Qu, R.; Li, C.; Liu, J.; Xiao, R.; Pan, X.; Zeng, X.; Wang, Z.; Wu, J. Hydroxyl radical based photocatalytic degradation of halogenated organic contaminants and paraffin on silica gel. Environ. Sci. Technol. 2018, 52, 7220–7229. [Google Scholar] [CrossRef] [PubMed]
  64. Ghanbari, F.; Khatebasreh, M.; Mahdavianpour, M.; Lin, K.-Y.A. Oxidative removal of benzotriazole using peroxymonosulfate/ozone/ultrasound: Synergy, optimization, degradation intermediates and utilizing for real wastewater. Chemosphere 2020, 244, 125326. [Google Scholar] [CrossRef]
  65. Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Gupta, V.K.; Singh, P. Review on augmentation in photocatalytic activity of CoFe2O4 via heterojunction formation for photocatalysis of organic pollutants in water. J. Saudi Chem. Soc. 2019, 23, 1119–1136. [Google Scholar]
  66. Abdelraheem, W.H.; Nadagouda, M.N.; Dionysiou, D.D. Solar light-assisted remediation of domestic wastewater by NB-TiO2 nanoparticles for potable reuse. Appl. Catal. B Environ. 2020, 269, 118807. [Google Scholar] [CrossRef]
  67. Mishra, N.S.; Reddy, R.; Kuila, A.; Rani, A.; Mukherjee, P.; Nawaz, A.; Pichiah, S. A review on advanced oxidation processes for effective water treatment. Curr. World Environ. 2017, 12, 470. [Google Scholar] [CrossRef]
  68. Tufail, A.; Price, W.E.; Mohseni, M.; Pramanik, B.K.; Hai, F.I. A critical review of advanced oxidation processes for emerging trace organic contaminant degradation: Mechanisms, factors, degradation products, and effluent toxicity. J. Water Process Eng. 2020, 40, 101778. [Google Scholar] [CrossRef]
  69. Flouret, A.; de Almeida, M.C.; de Oliveira, T.F.; de Sá, F.P. Advanced treatment of phenol by H2O2/UV/activated carbon coupling: Influence of homogeneous and heterogeneous phase. Can. J. Chem. Eng. 2018, 96, 1979–1985. [Google Scholar] [CrossRef]
  70. da Silva, J.R.P.; Monteiro, M.A.; de Mendonça Ochs, S.; da Silva Moura, C.; da Fonseca, F.V.; Borges, C.P. Study of effects of pharmaceuticals on the activated sludge process combining advanced oxidation using ultraviolet/hydrogen peroxide to increase their removal and mineralization of wastewater. J. Environ. Chem. Eng. 2020, 9, 104576. [Google Scholar] [CrossRef]
  71. Enaime, G.; Baçaoui, A.; Yaacoubi, A.; Belaqziz, M.; Wichern, M.; Lübken, M. Phytotoxicity assessment of olive mill wastewater treated by different technologies: Effect on seed germination of maize and tomato. Environ. Sci. Pollut. Res. 2020, 27, 8034–8045. [Google Scholar] [CrossRef]
  72. Ferreira, L.; Salmerón, I.; Peres, J.; Tavares, P.; Lucas, M.; Malato, S. Advanced Oxidation Processes as sustainable technologies for the reduction of elderberry agro-industrial water impact. Water Resour. Ind. 2020, 24, 100137. [Google Scholar] [CrossRef]
  73. Rueda-Márquez, J.J.; Levchuk, I.; Manzano, M.; Sillanpää, M. Toxicity Reduction of Industrial and Municipal Wastewater by Advanced Oxidation Processes (Photo-Fenton, UVC/H2O2, Electro-Fenton and Galvanic Fenton): A Review. Catalysts 2020, 10, 612. [Google Scholar] [CrossRef]
  74. Rott, E.; Kuch, B.; Lange, C.; Richter, P.; Kugele, A.; Minke, R. Removal of emerging contaminants and estrogenic activity from wastewater treatment plant effluent with UV/chlorine and UV/H2O2 advanced oxidation treatment at pilot scale. Int. J. Environ. research Public Health 2018, 15, 935. [Google Scholar] [CrossRef] [Green Version]
  75. Liu, Y.; Zhao, Y.; Wang, J. Fenton/Fenton-like processes with in-situ production of hydrogen peroxide/hydroxyl radical for degradation of emerging contaminants: Advances and prospects. J. Hazard. Mater. 2020, 404, 124191. [Google Scholar] [CrossRef]
  76. Hassanshahi, N.; Karimi-Jashni, A. Comparison of photo-Fenton, O3/H2O2/UV and photocatalytic processes for the treatment of gray water. Ecotoxicol. Environ. Saf. 2018, 161, 683–690. [Google Scholar] [CrossRef]
  77. García, C.A.; Hodaifa, G. Real olive oil mill wastewater treatment by photo-Fenton system using artificial ultraviolet light lamps. J. Clean. Prod. 2017, 162, 743–753. [Google Scholar] [CrossRef]
  78. Nguyen, T.M.H.; Suwan, P.; Koottatep, T.; Beck, S.E. Application of a novel, continuous-feeding ultraviolet light emitting diode (UV-LED) system to disinfect domestic wastewater for discharge or agricultural reuse. Water Res. 2019, 153, 53–62. [Google Scholar] [CrossRef] [PubMed]
  79. Dhivakar, D.T. BTEX Compounds Removal from Waste Water by using UV&UV/H2O2 Process. Int. J. Recent Eng. Sci. 2018, 5, 22–25. [Google Scholar]
  80. Aljuboury, D.; Palaniandy, P.; Abdul Aziz, H.; Feroz, S. Treatment of petroleum wastewater by conventional and new technologies-A review. Glob. Nest J. 2017, 19, 439–452. [Google Scholar]
  81. GilPavas, E.; Dobrosz-Gómez, I.; Gómez-García, M.-Á. Optimization and toxicity assessment of a combined electrocoagulation, H2O2/Fe2+/UV and activated carbon adsorption for textile wastewater treatment. Sci. Total Environ. 2019, 651, 551–560. [Google Scholar] [CrossRef] [PubMed]
  82. Neisi, A.; Afshin, S.; Rashtbari, Y.; Babaei, A.A.; Khaniabadi, Y.O.; Asadi, A.; Shirmardi, M.; Vosoughi, M. Efficiency of sequencing batch reactor for removal of organic matter in the effluent of petroleum wastewater. Data Brief 2018, 19, 2041–2046. [Google Scholar] [CrossRef]
  83. Raji, M.; Mirbagheri, S.A. A global trend of Fenton-based AOPs focused on wastewater treatment: A bibliometric and visualization analysis. Water Pract. Technol. 2020, 16, 19–34. [Google Scholar] [CrossRef]
  84. Ghime, D.; Ghosh, P. Removal of organic compounds found in the wastewater through electrochemical advanced oxidation processes: A review. Russ. J. Electrochem. 2019, 55, 591–620. [Google Scholar] [CrossRef]
  85. Ghasemi, H.; Aghabarari, B.; Alizadeh, M.; Khanlarkhani, A.; Abu-Zahra, N. High efficiency decolorization of wastewater by Fenton catalyst: Magnetic iron-copper hybrid oxides. J. Water Process. Eng. 2020, 37, 101540. [Google Scholar] [CrossRef]
  86. Zouanti, M.; Bezzina, M.; Dhib, R. Experimental study of degradation and biodegradability of oxytetracycline antibiotic in aqueous solution using Fenton process. Environ. Eng. Res. 2020, 25, 316–323. [Google Scholar] [CrossRef] [Green Version]
  87. Litter, M.I.; Slodowicz, M. An overview on heterogeneous Fenton and photoFenton reactions using zerovalent iron materials. J. Adv. Oxid. Technol. 2017, 20. [Google Scholar] [CrossRef]
  88. Ushie, E. Simultaneous Removal of Alkylphenols and Oils in Simulated Produced Water by UV/MW/Fenton-Like Process Using a Novel Surface Functionalised Heterogeneous Pan Catalyst; De Montfort University: Leicester, UK, 2018. [Google Scholar]
  89. Kim, D.-H.; Lee, D.; Monllor-Satoca, D.; Kim, K.; Lee, W.; Choi, W. Homogeneous photocatalytic Fe3+/Fe2+ redox cycle for simultaneous Cr (VI) reduction and organic pollutant oxidation: Roles of hydroxyl radical and degradation intermediates. J. Hazard. Mater. 2019, 372, 121–128. [Google Scholar] [CrossRef] [PubMed]
  90. Yuan, D.; Zhang, C.; Tang, S.; Li, X.; Tang, J.; Rao, Y.; Wang, Z.; Zhang, Q. Enhancing CaO2 fenton-like process by Fe (II)-oxalic acid complexation for organic wastewater treatment. Water Res. 2019, 163, 114861. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, Y.; Lin, X.; Shao, Z.; Shan, D.; Li, G.; Irini, A. Comparison of Fenton, UV-Fenton and nano-Fe3O4 catalyzed UV-Fenton in degradation of phloroglucinol under neutral and alkaline conditions: Role of complexation of Fe3+ with hydroxyl group in phloroglucinol. Chem. Eng. J. 2017, 313, 938–945. [Google Scholar] [CrossRef]
  92. Jiang, Z.; Wang, L.; Lei, J.; Liu, Y.; Zhang, J. Photo-Fenton degradation of phenol by CdS/rGO/Fe2+ at natural pH with in situ-generated H2O2. Appl. Catal. B Environ. 2019, 241, 367–374. [Google Scholar] [CrossRef]
  93. Thakare, Y.D.; Wani, K.S. Treatment of Industrial Waste Water by Fenton Process. J. Archit. Technol. 2019, XI, 43–49. [Google Scholar]
  94. Villegas-Guzman, P.; Giannakis, S.; Rtimi, S.; Grandjean, D.; Bensimon, M.; De Alencastro, L.F.; Torres-Palma, R.; Pulgarin, C. A green solar photo-Fenton process for the elimination of bacteria and micropollutants in municipal wastewater treatment using mineral iron and natural organic acids. Appl. Catal. B Environ. 2017, 219, 538–549. [Google Scholar] [CrossRef]
  95. Wang, B.; Liu, Y.; Zhang, Y.; Shen, F.; Yang, G.; Zhang, X.; Wang, L.; Luo, L.; He, Y.; Deng, S. Degradation process and kinetics study of actual urotropine wastewater by Fenton method. Desalination Water Treat. 2019, 160, 219–228. [Google Scholar] [CrossRef] [Green Version]
  96. Perini, J.A.L.; Tonetti, A.L.; Vidal, C.; Montagner, C.C.; Nogueira, R.F.P. Simultaneous degradation of ciprofloxacin, amoxicillin, sulfathiazole and sulfamethazine, and disinfection of hospital effluent after biological treatment via photo-Fenton process under ultraviolet germicidal irradiation. Appl. Catal. B Environ. 2018, 224, 761–771. [Google Scholar] [CrossRef] [Green Version]
  97. Shokri, A. Application of Sono–photo-Fenton process for degradation of phenol derivatives in petrochemical wastewater using full factorial design of experiment. Int. J. Ind. Chem. 2018, 9, 295–303. [Google Scholar] [CrossRef] [Green Version]
  98. Clarizia, L.; Russo, D.; Di Somma, I.; Marotta, R.; Andreozzi, R. Homogeneous photo-Fenton processes at near neutral pH: A review. Appl. Catal. B Environ. 2017, 209, 358–371. [Google Scholar] [CrossRef]
  99. Danforth, R.A. Ultrafast Photochemistry of Aqueous Iron (III) Complexes. Ph.D. Thesis, College of Letters & Science, Montana State University-Bozeman, Bozeman, MT, USA, 2017. [Google Scholar]
  100. Muramatsu, K.; Tokumura, M.; Wang, Q.; Miyake, Y.; Amagai, T.; Makino, M. Mitigation of the inhibitory effects of co-existing substances on the Fenton process by UV light irradiation. J. Environ. Sci. Health Part A 2020, 55, 730–738. [Google Scholar] [CrossRef] [PubMed]
  101. Zhang, M.-H.; Dong, H.; Zhao, L.; Wang, D.-x.; Meng, D. A review on Fenton process for organic wastewater treatment based on optimization perspective. Sci. Total Environ. 2019, 670, 110–121. [Google Scholar] [CrossRef] [PubMed]
  102. Ye, Z.; Sirés, I.; Zhang, H.; Huang, Y.-H. Mineralization of pentachlorophenol by ferrioxalate-assisted solar photo-Fenton process at mild pH. Chemosphere 2019, 217, 475–482. [Google Scholar] [CrossRef]
  103. Giwa, A.; Yusuf, A.; Balogun, H.A.; Sambudi, N.S.; Bilad, M.R.; Adeyemi, I.; Chakraborty, S.; Curcio, S. Recent advances in advanced oxidation processes for removal of contaminants from water: A comprehensive review. Process Saf. Environ. Prot. 2021, 146, 220–256. [Google Scholar] [CrossRef]
  104. Matavos-Aramyan, S.; 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, 1–18. [Google Scholar]
  105. Fernandes, A.; Makoś, P.; Boczkaj, G. Treatment of bitumen post oxidative effluents by sulfate radicals based advanced oxidation processes (S-AOPs) under alkaline pH conditions. J. Clean. Prod. 2018, 195, 374–384. [Google Scholar] [CrossRef]
  106. Rajala, K.; Grönfors, O.; Hesampour, M.; Mikola, A. Removal of microplastics from secondary wastewater treatment plant effluent by coagulation/flocculation with iron, aluminum and polyamine-based chemicals. Water Res. 2020, 183, 116045. [Google Scholar] [CrossRef] [PubMed]
  107. Dalari, B.L.S.K.; Giroletti, C.L.; Dalri-Cecato, L.; Domingos, D.G.; Hassemer, M.E.N. Application of heterogeneous photo-fenton process using chitosan beads for textile wastewater treatment. J. Environ. Chem. Eng. 2020, 8, 103893. [Google Scholar] [CrossRef]
  108. Lumbaque, E.C. Degradation of Pharmaceuticals in Hospital Wastewater by Solar Photo-Fenton Processes; UFRGS: Porto Alegre, Brazil, 2020. [Google Scholar]
  109. Nikravesh, B.; Shomalnasab, A.; Nayyer, A.; Aghababaei, N.; Zarebi, R.; Ghanbari, F. UV/Chlorine process for dye degradation in aqueous solution: Mechanism, affecting factors and toxicity evaluation for textile wastewater. J. Environ. Chem. Eng. 2020, 8, 104244. [Google Scholar] [CrossRef]
  110. Amildon, I.; Paniagua, C.E.; Paiva, V.A.; Gonçalves, B.R.; Sousa, R.M.; Machado, A.E.; Trovó, A.G. Degradation and initial mechanism pathway of chloramphenicol by photo-Fenton process at circumneutral pH. Chem. Eng. J. 2018, 339, 531–538. [Google Scholar] [CrossRef]
  111. Pourehie, O.; Saien, J. Homogeneous solar Fenton and alternative processes in a pilot-scale rotatable reactor for the treatment of petroleum refinery wastewater. Process Saf. Environ. Prot. 2020, 135, 236–243. [Google Scholar] [CrossRef]
  112. El-Qanni, A. Development of Sustainable Nanosorbcats Based Technology for Hydrocarbons and Organic Pollutants Recovery from Industrial Wastewater; University of Calgary: Calgary, Alberta, 2017. [Google Scholar]
  113. Radwan, M.; Alalm, M.G.; El-Etriby, H.K. Application of electro-Fenton process for treatment of water contaminated with benzene, toluene, and p-xylene (BTX) using affordable electrodes. J. Water Process Eng. 2019, 31, 100837. [Google Scholar] [CrossRef]
  114. Brillas, E.; Garcia-Segura, S. Benchmarking recent advances and innovative technology approaches of Fenton, photo-Fenton, electro-Fenton, and related processes: A review on the relevance of phenol as model molecule. Sep. Purif. Technol. 2020, 237, 116337. [Google Scholar] [CrossRef]
  115. Hassan, A.K.; Hassan, M.M.A.; Hasan, A.F. Treatment of Iraqi Petroleum Refinery Wastewater by Advanced Oxidation Processes; Journal of Physics: Conference Series; IOP Publishing: Bristol, UK, 2020; p. 012071. [Google Scholar]
  116. Pintor, A.M.; Vilar, V.J.; Botelho, C.M.; Boaventura, R.A. Oil and grease removal from wastewaters: Sorption treatment as an alternative to state-of-the-art technologies. A critical review. Chem. Eng. J. 2016, 297, 229–255. [Google Scholar] [CrossRef]
  117. Testolin, R.C.; Mater, L.; Sanches-Simões, E.; Dal Conti-Lampert, A.; Corrêa, A.X.; Groth, M.L.; Oliveira-Carneiro, M.; Radetski, C.M. Comparison of the mineralization and biodegradation efficiency of the Fenton reaction and Ozone in the treatment of crude petroleum-contaminated water. J. Environ. Chem. Eng. 2020, 8, 104265. [Google Scholar] [CrossRef]
  118. Boczkaj, G.; Fernandes, A.; Makoś, P. Study of different advanced oxidation processes for wastewater treatment from petroleum bitumen production at basic pH. Ind. Eng. Chem. Res. 2017, 56, 8806–8814. [Google Scholar] [CrossRef]
  119. Cruz Alcalde, A. Contribution to Performance Characterization and Kinetic Modelling of Micropollutants Abatement in Water and Wastewater by Ozone-based Oxidation Processes. Ph.D. Thesis, Universitat de Barcelona, Barcelona, Spain, 2019. [Google Scholar]
  120. Contreras, S.; Rodrıguez, M.; Al Momani, F.; Sans, C.; Esplugas, S. Contribution of the ozonation pre-treatment to the biodegradation of aqueous solutions of 2, 4-dichlorophenol. Water Res. 2003, 37, 3164–3171. [Google Scholar] [CrossRef]
  121. Kim, M.S.; Cha, D.; Lee, K.-M.; Lee, H.-J.; Kim, T.; Lee, C. Modeling of ozone decomposition, oxidant exposures, and the abatement of micropollutants during ozonation processes. Water Res. 2020, 169, 115230. [Google Scholar] [CrossRef]
  122. Rekhate, C.V.; Srivastava, J. Recent advances in ozone-based advanced oxidation processes for treatment of wastewater-A review. Chem. Eng. J. Adv. 2020, 3, 100031. [Google Scholar] [CrossRef]
  123. Malvestiti, J.A.; Dantas, R.F. Disinfection of secondary effluents by O3, O3/H2O2 and UV/H2O2: Influence of carbonate, nitrate, industrial contaminants and regrowth. J. Environ. Chem. Eng. 2018, 6, 560–567. [Google Scholar] [CrossRef]
  124. Wen, C.; Wang, H.; Wang, L.; Lou, Z.; Sun, Z.; Zhou, Z. The reduction of waste lubricant oil distillate through the enhancement of organics degradation by ozonation with elevated temperature and stable pH for the zero discharge. J. Clean. Prod. 2019, 240, 118194. [Google Scholar] [CrossRef]
  125. Krishnan, S.; Rawindran, H.; Sinnathambi, C.; Lim, J. Comparison of various advanced oxidation processes used in remediation of industrial wastewater laden with recalcitrant pollutants. IOP Conf. Ser. Mater. Sci. Eng. 2017, 206, 012089. [Google Scholar] [CrossRef]
  126. Shahrezaei, F.; Mansouri, Y.; Zinatizadeh, A.A.L.; Akhbari, A. Process modeling and kinetic evaluation of petroleum refinery wastewater treatment in a photocatalytic reactor using TiO2 nanoparticles. Powder Technol. 2012, 221, 203–212. [Google Scholar] [CrossRef]
  127. Tao, P.; Yang, C.; Wang, H.; Zhao, Y.; Zhang, X.; Shao, M.; Sun, T. Synergistic effects of ultrasonic-assisted ozonation on the formation of hydrogen peroxide. J. Environ. Chem. Eng. 2020, 9, 104905. [Google Scholar] [CrossRef]
  128. Liu, Z.; Chys, M.; Yang, Y.; Demeestere, K.; Van Hulle, S. Oxidation of trace organic contaminants (TrOCs) in wastewater effluent with different ozone-based AOPs: Comparison of ozone exposure and •OH formation. Ind. Eng. Chem. Res. 2019, 58, 8896–8902. [Google Scholar] [CrossRef]
  129. Al Momani, F. Impact of photo-oxidation technology on the aqueous solutions of nitrobenzene: Degradation efficiency and biodegradability enhancement. J. Photochem. Photobiol. A Chem. 2006, 179, 184–192. [Google Scholar] [CrossRef]
  130. Miklos, D.B.; Remy, C.; Jekel, M.; Linden, K.G.; Drewes, J.E.; Hübner, U. Evaluation of advanced oxidation processes for water and wastewater treatment—A critical review. Water Res. 2018, 139, 118–131. [Google Scholar] [CrossRef]
  131. Buehlmann, P.H. Balancing Bromate Formation, Organics Oxidation, and Pathogen Inactivation: The Impact of Bromate Suppression Techniques on Ozonation System Performance in Reuse Waters. Ph.D. Thesis, Virginia Tech, Blacksburg, VA, USA, 2019. [Google Scholar]
  132. El-Din, M.G.; Smith, D.W.; Al Momani, F.; Wang, W. Oxidation of resin and fatty acids by ozone: Kinetics and toxicity study. Water Res. 2006, 40, 392–400. [Google Scholar] [CrossRef]
  133. Tufail, A.; Price, W.E.; Hai, F.I. A critical review on advanced oxidation processes for the removal of trace organic contaminants: A voyage from individual to integrated processes. Chemosphere 2020, 260, 127460. [Google Scholar] [CrossRef]
  134. Al Momani, F. Degradation of cyanobacteria anatoxin-a by advanced oxidation processes. Sep. Purif. Technol. 2007, 57, 85–93. [Google Scholar] [CrossRef]
  135. Wei, C.; Zhang, F.; Hu, Y.; Feng, C.; Wu, H. Ozonation in water treatment: The generation, basic properties of ozone and its practical application. Rev. Chem. Eng. 2017, 33, 49–89. [Google Scholar] [CrossRef]
  136. Yang, S.; Song, Y.; Chang, F.; Wang, K. Evaluation of chemistry and key reactor parameters for industrial water treatment applications of the UV/O3 process. Environ. Res. 2020, 188, 109660. [Google Scholar] [CrossRef]
  137. Cruz-Alcalde, A.; Esplugas, S.; Sans, C. Continuous versus single H2O2 addition in peroxone process: Performance improvement and modelling in wastewater effluents. J. Hazard. Mater. 2020, 387, 121993. [Google Scholar] [CrossRef]
  138. Al Momani, F.A.; Jarrah, N. Treatment and kinetic study of cyanobacterial toxin by ozone. J. Environ. Sci. Health Part A 2010, 45, 719–731. [Google Scholar] [CrossRef]
  139. Biard, P.-F.; Dang, T.T.; Bocanegra, J.; Couvert, A. Intensification of the O3/H2O2 advanced oxidation process using a continuous tubular reactor filled with static mixers: Proof of concept. Chem. Eng. J. 2018, 344, 574–582. [Google Scholar] [CrossRef]
  140. Meshref, M.N.; Klamerth, N.; Islam, M.S.; McPhedran, K.N.; El-Din, M.G. Understanding the similarities and differences between ozone and peroxone in the degradation of naphthenic acids: Comparative performance for potential treatment. Chemosphere 2017, 180, 149–159. [Google Scholar] [CrossRef] [PubMed]
  141. Wang, H.; Zhan, J.; Yao, W.; Wang, B.; Deng, S.; Huang, J.; Yu, G.; Wang, Y. Comparison of pharmaceutical abatement in various water matrices by conventional ozonation, peroxone (O3/H2O2), and an electro-peroxone process. Water Res. 2018, 130, 127–138. [Google Scholar] [CrossRef]
  142. Khatri, I.; Singh, S.; Garg, A. Performance of electro-Fenton process for phenol removal using Iron electrodes and activated carbon. J. Environ. Chem. Eng. 2018, 6, 7368–7376. [Google Scholar] [CrossRef]
  143. Ahmadi, M.; Jorfi, S.; Kujlu, R.; Ghafari, S.; Soltani, R.D.C.; Haghighifard, N.J. A novel salt-tolerant bacterial consortium for biodegradation of saline and recalcitrant petrochemical wastewater. J. Environ. Manag. 2017, 191, 198–208. [Google Scholar] [CrossRef]
  144. Ebrahimi, M.; Kazemi, H.; Mirbagheri, S.; Rockaway, T.D. An optimized biological approach for treatment of petroleum refinery wastewater. J. Environ. Chem. Eng. 2016, 4, 3401–3408. [Google Scholar] [CrossRef]
  145. Mirbagheri, S.A.; Ebrahimi, M.; Mohammadi, M. Optimization method for the treatment of Tehran petroleum refinery wastewater using activated sludge contact stabilization process. Desalination Water Treat. 2014, 52, 156–163. [Google Scholar] [CrossRef]
  146. Liang, J.; Mai, W.; Tang, J.; Wei, Y. Highly effective treatment of petrochemical wastewater by a super-sized industrial scale plant with expanded granular sludge bed bioreactor and aerobic activated sludge. Chem. Eng. J. 2019, 360, 15–23. [Google Scholar] [CrossRef]
  147. Razavi, S.M.R.; Miri, T. A real petroleum refinery wastewater treatment using hollow fiber membrane bioreactor (HF-MBR). J. Water Process Eng. 2015, 8, 136–141. [Google Scholar] [CrossRef]
  148. El-Naas, M.H.; Alhaija, M.A.; Al-Zuhair, S. Evaluation of a three-step process for the treatment of petroleum refinery wastewater. J. Environ. Chem. Eng. 2014, 2, 56–62. [Google Scholar] [CrossRef]
  149. An, C.; Huang, G.; Yao, Y.; Zhao, S. Emerging usage of electrocoagulation technology for oil removal from wastewater: A review. Sci. Total Environ. 2017, 579, 537–556. [Google Scholar] [CrossRef]
  150. Huo, S.; Zhu, F.; Zou, B.; Xu, L.; Cui, F.; You, W. A two-stage system coupling hydrolytic acidification with algal microcosms for treatment of wastewater from the manufacture of acrylonitrile butadiene styrene (ABS) resin. Biotechnol. Lett. 2018, 40, 689–696. [Google Scholar] [CrossRef]
  151. Fernandes, A.; Gągol, M.; Makoś, P.; Khan, J.A.; Boczkaj, G. Integrated photocatalytic advanced oxidation system (TiO2/UV/O3/H2O2) for degradation of volatile organic compounds. Sep. Purif. Technol. 2019, 224, 1–14. [Google Scholar] [CrossRef]
  152. Dastpak, H.; Pasalari, H.; Jafari, A.J.; Gholami, M.; Farzadkia, M. Improvement of Co-Composting by a combined pretreatment Ozonation/Ultrasonic process in stabilization of raw activated sludge. Sci. Rep. 2020, 10, 1–7. [Google Scholar] [CrossRef]
  153. Zhai, J.; Ma, H.; Liao, J.; Rahaman, M.; Yang, Z.; Chen, Z. Comparison of Fenton, ultraviolet–Fenton and ultrasonic–Fenton processes on organics and colour removal from pre-treated natural gas produced water. Int. J. Environ. Sci. Technol. 2018, 15, 2411–2422. [Google Scholar] [CrossRef]
  154. Shah, A.; Shah, M. Characterisation and bioremediation of wastewater: A review exploring bioremediation as a sustainable technique for pharmaceutical wastewater. Groundw. Sustain. Dev. 2020, 11, 100383. [Google Scholar] [CrossRef]
  155. Horovitz, I.; Avisar, D.; Luster, E.; Lozzi, L.; Luxbacher, T.; Mamane, H. MS2 bacteriophage inactivation using a N-doped TiO2-coated photocatalytic membrane reactor: Influence of water-quality parameters. Chem. Eng. J. 2018, 354, 995–1006. [Google Scholar] [CrossRef]
  156. Yadav, D.N.; Kishore, K.A.; Bethi, B.; Sonawane, S.H.; Bhagawan, D. ZnO nanophotocatalysts coupled with ceramic membrane method for treatment of Rhodamine-B dye waste water. Environ. Dev. Sustain. 2018, 20, 2065–2078. [Google Scholar] [CrossRef]
  157. Giwa, A.; Yusuf, A.; Dindi, A.; Balogun, H.A. Polygeneration in desalination by photovoltaic thermal systems: A comprehensive review. Renew. Sustain. Energy Rev. 2020, 130, 109946. [Google Scholar] [CrossRef]
Figure 1. An overview for the AOP in petroleum refinery wastewater treatment.
Figure 1. An overview for the AOP in petroleum refinery wastewater treatment.
Catalysts 11 00782 g001
Figure 2. (a) Photo-Fenton, (b) Fenton process reactions mechanism.
Figure 2. (a) Photo-Fenton, (b) Fenton process reactions mechanism.
Catalysts 11 00782 g002
Figure 3. The major drawbacks for Fenton and photo-Fenton processes.
Figure 3. The major drawbacks for Fenton and photo-Fenton processes.
Catalysts 11 00782 g003
Figure 4. (a) The used solar photoreactor setup. (1) the surface of solar reactor, (2) light sensor, (3) recycle pump, (4) electromotor for rotation, (5) flow meter, (6) flow recycle valves, (7) storage tank, (8) and dynamic Jack. (b). The surface of the solar reactor; (1) quartz tubes, (2) parabolic polished aluminum reflector sheets, (3) Wind crossing paths, and (4) light sensor (Reprinted with permission from ref. [111]. Copyright 2020 Elsevier).
Figure 4. (a) The used solar photoreactor setup. (1) the surface of solar reactor, (2) light sensor, (3) recycle pump, (4) electromotor for rotation, (5) flow meter, (6) flow recycle valves, (7) storage tank, (8) and dynamic Jack. (b). The surface of the solar reactor; (1) quartz tubes, (2) parabolic polished aluminum reflector sheets, (3) Wind crossing paths, and (4) light sensor (Reprinted with permission from ref. [111]. Copyright 2020 Elsevier).
Catalysts 11 00782 g004
Figure 5. The experimental set up of photoreactor (Reprinted with permission from ref. [126]. Copyright 2012 Elsevier).
Figure 5. The experimental set up of photoreactor (Reprinted with permission from ref. [126]. Copyright 2012 Elsevier).
Catalysts 11 00782 g005
Figure 6. The detailed steps of the mechanism of organic degradation through the O3/UV method.
Figure 6. The detailed steps of the mechanism of organic degradation through the O3/UV method.
Catalysts 11 00782 g006
Figure 7. The integrated mobile system for oilfield produced water treatment. (1) EC equipment; (2) reaction/sedimentation equipment; (3) UF membrane unit; (4) controller; (5) RO unit; (6) trailer platform; (7) pressure filter (Reprinted with permission from ref. [147]. Copyright 2017 Elsevier).
Figure 7. The integrated mobile system for oilfield produced water treatment. (1) EC equipment; (2) reaction/sedimentation equipment; (3) UF membrane unit; (4) controller; (5) RO unit; (6) trailer platform; (7) pressure filter (Reprinted with permission from ref. [147]. Copyright 2017 Elsevier).
Catalysts 11 00782 g007
Table 1. The compositions present in the petroleum refinery wastewater.
Table 1. The compositions present in the petroleum refinery wastewater.
Composition of Each ElementComposition in Petroleum WastewaterLimitationsMetal CompositionsComponents in Petroleum WastewaterLimitations
pH4.3–106.5–8.5Ca18–132,687100
Density1014–1140-Na316–134,652-
TOC3.4–5960-K8.6–14,649-
COD1200≤50Mg4–18,145100
BOD-≤30Fe<0.1–1000.3
TSS1.2–21,82030Al310–410≤0.2
TDS1 × 103–4 × 1051200B5.00–951
DO8.2<4.0Ba0–224001
TPH2–565-Cd<0.005–0.20.005
BTEX0.39–35-Cr0.02–1.1≤0.1
base and neutrals<140-Cu<0.002–1.51.3
Cl80–310,5611400–190,000Li3–50.00-
Br0–12,000150–1149Mn<0.004–1750.05
I0–500-Ni-0.3
HCO31.9–7355-Pb0.002–8.80.015
CO30–800-Sr0.02–1000-
SO40.5–7851<0.1–47Ti<0.01–0.7-
PO40–0.10-Zn0.01–357.4
SO3~10-As<0.005–0.30.02
NO30–3.5-Hg<0.001–0.0020.005
NO20.05-Ag<0.001–0.150.1
NH3-N10–300-Be<0.001–0.004-
Total Polar9.7–600-NORM (pCi/L)--
Higher acids<1–63-Total Ra0.054–32,4005
Phenolstill 23-U0.008–2.7-
VFA2–4900-Th0.008–0.02715
Oil & Grease6.9–2102.3–60Pb1.35–5130-
m-xylene0.01–54-Po0.005–0.17-
MBAS0.01–54----
HEM0.6–20000.02---
Alkalinity6.1–200----
Table 2. Some of the main wastewater contaminants mentioned in the work and their environmental impacts.
Table 2. Some of the main wastewater contaminants mentioned in the work and their environmental impacts.
ContaminantEnvironmental ImpactRef
PhenolHave harmful effects on the muscles, causing moving problems, pain to the gastrointestinal system, and even death.Yang et al. (2020)
Methyl-butyl etherIn an aquatic environment, it quickly causes the reduction in dissolved oxygen. Regarding human health, it causes kidney and blood cancers and affects the nerve. Gallo et al. (2020)
BenzeneThe existence in the aquatic environment affects the stability and health of the organisms, leading to a law regenerative rate, and affecting the behavior of the organisms. It also reduces the growth of plants and animals, and, with long time exposure, it can lead to their death.Rabani et al. (2020)
TolueneHave harmful impacts on the different types of microorganism (aquatic, soil, etc).Poyraz et al. (2020)
EthylbenzeneExposure for long periods affects the kidney which can lead to pain in the inner ear.Wollin et al. (2020)
XyleneAffects the kidney, nervous system, and the dysfunctions of the liver. In addition, various negative impacts on the neurological system have been stated.Egendorf et al. (2020)
Table 3. Types and categorization of advanced oxidation methods (AOPs).
Table 3. Types and categorization of advanced oxidation methods (AOPs).
Non-photochemical-Homogeneous MethodsPhotochemical-Homogeneous MethodsNon-Photochemical-Heterogeneous ProcessesPhotochemical-Heterogeneous Processes
  • Ozonation in alkaline media (O3/HO)
  • Ozonation with hydrogen peroxide (O3/H2O2)
  • Fenton (Fe2+ or Fe3+/H2O2)
  • Electro-oxidation
  • Electrohydraulic discharge-ultrasound
  • Wet air oxidation (WAO)
  • Supercritical water oxidation (SCWO)
  • Photolysis of water in vacuum ultraviolet (VUV)
  • UV/H2O2tyg,v =
  • UV/O3
  • UV/O3/H2O2
  • Photo-Fenton Fe2+ or Fe3+/H2O2/UV
  • Catalytic wet air oxidation (CWAO)
  • Heterogeneous photocatalysis: /UV, SnO2/UV, TiO2/UV, TiO2/H2O2/UV
Table 4. Comparison between recent studies using H2O2 advanced oxidation-based process in wastewater treatment.
Table 4. Comparison between recent studies using H2O2 advanced oxidation-based process in wastewater treatment.
ProcessType of Wastewater (WW)ConditionsEfficiencyComments and DetailsRef
TiO2/UVModel: TPH Benzene (Toluene Phenol)
Naphthalene
[TiO2] 100 mg/L
pH 6.5 for 1 h
COD reduction from 970 mg/L to 65 mg/L
= 93% removal at 90 min, 30 °C and pH 3 using 100 mg/L
Higher degradations rates for with TiO2/UV = 92%, 98.8%, 91.5%, and 93% for B, T, P and N, respectively.Ihtisham et al. 2020
H2O2/UVModel [Phenol] 100 mg/L(H2O2) 3060 mg/L pH 999% phenol removal in 60 minH2O2/UV<O3/UV
0.081 min−1 < 0.0881 min−1
Alrousan et al. 2020
H2O2/UVOil COD 9000 ± 500 mg/L(H2O2) 3330 mg/L
pH 12
40% COD removal in 150 minNegative impact on COD removal at basic pHGilPavas et al. 2019
H2O2/UVModel: [nonylphenol] 4.41 mg/L(H2O2) 1.7–17 mg/L pH 11To achieve 90% of degradation [H2O2] 1.7 mg/L in 20 min
(H2O2) 3.4 mg/L in 16 min
(H2O2) 8.5 mg/L in 12 min
At pH 7 the time needed was 3 times higher in all concentrations studiedKaur et al. 2020
H2O2
H2O2/UV
Linear Alkyl Benzene COD 300–350 mg/L(H2O2) 340 mg/L pH 9H2O2: 27% COD removal in 180 min H2O2/UV: 36% COD removal in 180 minH2O2 (27%) < O3/UV, H2O2/UV (36%) < O3, (37%) < H2O2/O3 (39%)< H2O2/O3/UV (42%)Fernandes et al. 2019
H2O2/UVModel: MTBE 25 mg/L(H2O2) 34 mg/L pH 1298% MTBE removal in 60 minH2O2/UV
(0.126 min−1)
Neisi et al. 2018
Table 5. The efficiencies of Fenton and Photo-Fenton methods in the treatment of petroleum wastewater.
Table 5. The efficiencies of Fenton and Photo-Fenton methods in the treatment of petroleum wastewater.
NoMethodParameterRemoval Efficiency (%)pHH2O2 (ppm)Fe2+
(ppm)
Ratio of H2O2:Fe2+Reaction Time (min)Ref
1H2O2/Fe2+/solarTOC84210.0812.5300Aljuboury
et al. 2017
2H2O2/Fe2+COD/
TOC
56 COD
54 TOC
4.39.71.18.8120Edison
et al. 2019
3H2O2/Fe2+/UVCOD923110353.1492Majed et al. 2020
4H2O2/Fe2+/solarCOD843.22001.5133180Asaithambi
et al. 2017
5H2O2/Fe3+TOC9035002502120Deng
et al. 2017
6H2O2/Fe2+/UVCOD725.617.861.7610.1470Shokri
et al. 2019
7H2O2/Fe3+/TiO2COD69.631600305360Hassan 2018
8H2O2/Fe2+/UVCOD76.83250406.2530Tufaner
2020
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Elmobarak, W.F.; Hameed, B.H.; Almomani, F.; Abdullah, A.Z. A Review on the Treatment of Petroleum Refinery Wastewater Using Advanced Oxidation Processes. Catalysts 2021, 11, 782. https://doi.org/10.3390/catal11070782

AMA Style

Elmobarak WF, Hameed BH, Almomani F, Abdullah AZ. A Review on the Treatment of Petroleum Refinery Wastewater Using Advanced Oxidation Processes. Catalysts. 2021; 11(7):782. https://doi.org/10.3390/catal11070782

Chicago/Turabian Style

Elmobarak, Wamda Faisal, Bassim H. Hameed, Fares Almomani, and Ahmad Zuhairi Abdullah. 2021. "A Review on the Treatment of Petroleum Refinery Wastewater Using Advanced Oxidation Processes" Catalysts 11, no. 7: 782. https://doi.org/10.3390/catal11070782

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