Next Article in Journal
Extrapolation Framework and Characteristic Analysis of Load Spectrum for Agriculture General Power Machinery
Previous Article in Journal
Development Law of Water-Conducting Fracture Zones in Overburden above Fully Mechanized Top-Coal Caving Face: A Comprehensive Study
Previous Article in Special Issue
Catalytic Potential-Guided Design of Multi-Enzymatic System for DHA Production from Glycerol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Novel Approach of Tackling Wax Deposition Problems in Pipeline Using Enzymatic Degradation Process: Challenges and Potential Solutions

by
Shazleen Saadon
1,2,*,
Raja Noor Zaliha Raja Abd Rahman
1,
Nor Hafizah Ahmad Kamarudin
1,3,
Sara Shahruddin
2,
Siti Rohaida Mohd Shafian
2,
Norhidayah Ahmad Wazir
2 and
Mohd Shukuri Mohamad Ali
1,*
1
Enzyme and Microbial Technology Research Centre, Faculty of Biotechnology and Biomolecular Sciences, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
2
PETRONAS Research Sdn Bhd, Lot 3288 & 3299, Off Jalan Ayer Hitam, Kawasan Institusi Bangi, Kajang 43000, Selangor, Malaysia
3
Centre of Foundation Studies in Science for Universiti Putra Malaysia, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
*
Authors to whom correspondence should be addressed.
Processes 2024, 12(10), 2074; https://doi.org/10.3390/pr12102074
Submission received: 22 August 2024 / Revised: 11 September 2024 / Accepted: 21 September 2024 / Published: 25 September 2024
(This article belongs to the Special Issue Application of Enzymes in Sustainable Biocatalysis)

Abstract

:
Anthropogenic activities have led to hydrocarbon spills, and while traditional bioremediation methods are costly and time-consuming, recent research has focused on engineered enzymes for managing pollutant. The potential of enzymes for resolving wax flow problems in the petroleum industry remains unexplored. This paper offers a comprehensive review of the current state of research activities related to the bioremediation of petroleum-polluted sites and the biodegradation of specific petroleum hydrocarbons. The assayed enzymes that took part in the degradation were discussed in detail. Lipase, laccase, alkane hydroxylase, alcohol dehydrogenase, esterase, AlkB homologs and cytochrome P450 monooxygenase are among the enzymes responsible for the degradation of more than 50% of the hydrocarbons in contaminated soil and wastewater and found to be active on carbon C8 to C40. The possible biodegradation mechanism of petroleum hydrocarbons was also elucidated. The enzymes’ primary metabolic pathways include terminal, subterminal, and ω-oxidation. Next, given the successful evidence of the hydrocarbon treatment efficiency, the authors analyzed the opportunity for the enzymatic degradation approach if it were to be applied to a different scenario: managing wax deposition in petroleum-production lines. With properties such as high transformation efficiency and high specificity, enzymes can be utilized for the treatment of viscous heavy oil for transportability, evidenced by the 20 to 99% removal of hydrocarbons. The challenges associated with the new approach are also discussed. The production cost of enzymes, the characteristics of hydrocarbons and the operating conditions of the production line may affect the biocatalysis reaction to some extent. However, the challenges can be overcome by the usage of extremophilic enzymes. The combination of technological advancement and deployment strategies such as the immobilization of a consortium of highly thermophilic and halotolerant enzymes is suggested. Recovering and reusing enzymes offers an excellent strategy to improve the economics of the technology. This paper provides insights into the opportunity for the enzymatic degradation approach to be expanded for wax deposition problems in pipelines.

1. Introduction

Petroleum hydrocarbons are a blend of aromatic and aliphatic carbon compounds, with the composition changing depending on where the reservoir is found [1]. The composition of petroleum hydrocarbons can be characterized into four primary categories: saturates, aromatics, asphaltenes (which include phenols, fatty acids, and ketones), and resins (which include pyridines and amides) [2,3]. Saturates are nonpolar hydrocarbons that are also called waxes. Aromatic hydrocarbons contain polarizable rings, while resins and asphaltenes have polar substituents with oxygen, sulfur, and nitrogen [4]. These organic compounds are highly insoluble in water, which limits the hydrocarbon uptake needed for microorganisms to perform degradation [5]. Anthropogenic activities, such as petroleum exploration, transportation and refining accumulated hydrocarbon spillage at sites, consequently harm beaches, animal habitats, mangrove forests and human settlements. Figure 1 visualizes some of the oil spill disasters that have ensued over time. The situation had a major ecological impact on the environment, which must be managed responsibly.
Historically, research has focuses more on microbial degradation or the bioremediation of toxic pollutants in water and soil using microbes. This technique was costly, time-consuming, and limited by severe conditions of extreme pH, temperature, and chemical shock [6]. Over the past 15 years, research has focused on the use of enzymes for pollution management.
In the petroleum industry, hydrocarbons are not only managed for their spillage but also from the perspective of flow assurance during the transportation of hydrocarbons in pipelines. When a hydrocarbon is high in wax content, the wax crystallizes and is deposited on the pipeline wall, especially in cold environments where the production is farther offshore [7]. The wax accumulation restricts the flow of hydrocarbon and causes blockages [8]. Oil companies are applying chemicals to reduce the wax, combined with pipeline thermal insulation, and pigging to scrap off the wax deposit [9]. However, these strategies are expensive, inefficient, and less robust due to a lack of understanding of their workability in different fields’ operating conditions [7]. This gap creates a greater opportunity for novel discoveries, as conceptualized in Figure 2. To the authors’ knowledge, the enzymatic degradation of hydrocarbons for managing wax deposition in production lines has not been widely studied.

2. Wax Deposition Scenario and the Limitations of Existing Wax Treatment

Wax deposition occurs when wax precipitates out of the petroleum when its temperature is reduced. This is because alkanes that are heavier than carbon number C18 has a lower solubility in the liquid phase [10]. There are many factors affecting wax deposition in production lines, such as molecular diffusion, operating temperature, pressure, asphaltene composition, hydrocarbon flow velocity and pipeline design [11]. One of the adverse effects of wax deposition in the petroleum industry is the blockage of the pipeline, which causes an increase in pressure drop and decrease in flowability. For petroleum-production lines in cold regions or pipelines located below sea level, the wax build-up issue is very alarming [8]. Substantial expenditures are required for wax deposition control and remediation. The prevention and treatment of the wax deposition problem are conventionally performed using thermal, mechanical, chemical, and biological methods.

2.1. Thermal Treatment

The thermal method involves the installation of a heat tracing device and suitable insulation materials in a transportation pipeline to minimize wax deposition on the walls, and hot fluid circulates to keep the crude oil above its wax appearance temperature [12]. However, electrical heaters are prone to burn-out, which prompts an automatic cut-off system when overheating occurs [13]. The integrated production and transportation system of crude oil will be majorly impacted by this interruption. On the other hand, the limitation of the insulation technique is that it requires a material with suitable conductivity and chemical-resistance properties [12].

2.2. Mechanical Treatment

The mechanical method involves frequently scraping the pipeline using a rubber or metal tool before the wax begins to build up, together with the continuous injection of a solvent or wax inhibitor [14]. This technique has an advantage for a non-heating pipeline whereby the significant energy consumption can be reduced. However, the cutting force of the wax layer is dependent on the tool geometry and the friction at the wax–tool interface. A study that investigates the effect of tool geometry on wax scraping demonstrated that a 45° rake angle removed more of the wax layer than a negative 30° angle [15]. The efficiency of wax removal is also affected by the sliding speed of the scraper, which influences the rubbing force of the wax layer [16]. In addition, there is also the risk of a wax plug forming and the scraper becoming wedged inside the pipeline when handling viscous crude oil [14].

2.3. Ultrasonic and Electromagnetic Wave Propagation

The displacement of wax by ultrasonic and electromagnetic wave propagation is a relatively new approach. It was reported that when the crude oil was treated with ultrasonic waves, the resin and asphaltene fractions in the crude oil were increased due to the modification of the wax structure, which became a more polar compound. These increased fractions were regarded as natural surfactants in crude oil and caused the wax appearance temperature to reduce by 12 °C [17]. A morphological study has shown that the size of the wax crystals became smaller and rounded after irradiation treatment due to a decrease in wax molecules [18]. However, research has indicated that this technique requires combination with chemical [19,20,21] or thermal treatment [22] for it to be effective, which risks doubling the treatment cost.

2.4. Chemical Treatment

Chemical injection is by far the most effective and mature method. Chemicals like solvents, surfactants, wax inhibitors and pour-point depressants were injected to alter the surface characteristics and solubilize the wax deposits, leading to reduction in wax appearance temperature and pour point, so the wax would not form in cold environments [23]. It was reported that wax was reduced by 65 to 73% when a polyacrylate polymer-based pour-point depressant was applied to a waxy crude oil [24]. However, chemicals like wax inhibitors and pour-point depressants are typically solids at operating sites, thus requiring dilution with a solvent, which in turn increases the expenses and the number of possible operational hazards [25]. Additionally, the performance of the chemical is not universal. Its effectiveness is chemically specific to the composition of the crude oil [26,27]. Various type of copolymers that functioned as wax crystal modifiers produced different results, and their mechanisms are poorly understood despite many reported studies [28]. Some research has found that asphaltene constituents in crude oil can hinder the effectiveness of the chemicals [29]. Solvents like kerosene, gasoline and benzol are relatively cheap and are used to solubilized wax deposits. A combination of diesel, xylene, toluene, naphthalene and dispersant was formulated at an optimum ratio, which resulted in more than 85% of wax solubilization [23]. However, the usage of solvents posed toxicity and flammability threats to the handlers, and the effectiveness of the solvents also depends on the permeability of the wax deposits [23].

2.5. Biological Treatment

The biological method highlights an efficient and environmentally friendly solution involving hydrocarbon-degrading microbes. Bacteria are readily obtainable and inexpensive, and their handling is convenient. Because of this, bacteria have been utilized successfully for decades in the bioremediation of hydrocarbon-polluted sites [30,31]. The microbial degradation of petroleum was proven successful for ocean clean-up and onshore wastewater treatment. During the oil spill from the Exxon Valdez tanker, the microbial populations sampled from the shoreline of Prince William Sound beach showed the foremost ability to degrade aromatic hydrocarbons; subsequently, alkane degradation became more dominant over time [32]. Another study examined microbial communities from the Gulf of Mexico shoreline contaminated by the Deepwater Horizon blowout, which was also shown to be responsible for the major degradation of polyaromatic hydrocarbons [33]. A bioreactor was used to treat industrial wastewater rich in hydrocarbons through the acclimatization of microbial consortium, resulting in a hydrocarbon degradation of more than 97%, especially on alkanes C10 to C35 [34]. In fact, many advanced bioreactors have been designed to allow for the precise control of biodegradation parameters to increase the mass transfer and reaction rates during the microbial degradation reaction [35]. However, the survivability of bacteria is dependent on nutrients and growth enhancers. This approach may not be suitable for petroleum-production lines due to factors such as the high temperature and pressure of the incoming petroleum, seawater salinity, the availability of microbial substrates, the nutrient solution, dissolved oxygen and nitrogen sources, whereby bacteria are commonly sensitive to the changes of these factors [36]. On top of that, the presence of sulfate-reducing bacteria leads to biofouling, where the microbial growth forms biofilms that can clog and damaged the production lines [37]. The waxy petroleum treatment in production lines requires a less-complex and faster degradation of the high-molecular-weight hydrocarbons [25], but at the same time, it does not impact the low-molecular-weight hydrocarbons that are valuable to the sale value of the petroleum. However, there is no means to control degradation using the microbial method. Table 1 summarizes the limitations of each wax treatment method.
The advancements in technology and knowledge in protein engineering, recombinant DNA, synthetic biology, and metabolic engineering have enabled the design of novel enzymes with desired properties such as high catalytic activity, substrate specificity, and thermostability, tailored for wax treatment conditions. In this regard, the enzymatic degradation of waxy petroleum in production lines represents a novel approach.

3. Microbial Degradation of Petroleum Hydrocarbon and Enzymes Involved in the Reaction

Since enzymes’ potential can be maximized in a lab setting, biocatalysis— a technique that uses enzymes to perform chemical transformation—is a more strategic approach than dealing with the entire microorganism [43]. Within the biocatalytic degradation research, previous studies have described biodegradation pathways, genes, and the characterization of the biodegradative enzymes. The usage of these enzymes as biocatalysts either for making a useful compound or to transform toxic compounds into nontoxic compounds has also been described [44]—for example, the bioremediation of waste materials such as organic and inorganic pollutants and pharmaceutically active compounds in wastewater, industrial chemicals, and pesticides [45,46,47,48,49,50,51,52,53,54,55].
On the other hand, there are also specific studies on the biocatalytic degradation of petroleum, which is the main topic of this paper. For the past 10 years, the trend in research publications related to the specific biocatalytic degradation of hydrocarbon pollutants is more common than for hydrocarbon wax (Figure 3), probably because pollution clean-up is a more pressing matter globally, while the wax issue is specifically managed in the oil and gas sector. Both subjects are elaborated in detail in the next section, focusing on the species of bacteria hosting the enzymes responsible for removing the petroleum hydrocarbons.

3.1. Bioremediation of Petroleum-Contaminated Sites

The effective bioremediation of polluted sites depends on the survivability of the degrading microbes. One study sought to solve the low yield of degradative enzymes and biosurfactants by the intermittent inoculation of microbial consortia S. chilikensis, B. firmus and H. hamiltonii, to maintain a constant level of the degradative enzymes oxidoreductase, lipase, and catalase, and lipoprotein biosurfactants. As a result, a steep decrement of petroleum was observed [56]. Petroleum-polluted sites are commonly associated with high metal and salinity content. The metal-tolerant bacterium Novosphingobium panipatense P5:ABC was able to degrade 90% of petroleum, and alkane hydroxylase was responsible for this [57]. A comparison between free P. stutzeri and immobilized P. stutzeri showed that immobilized bacteria are more efficient in removing hydrocarbons from produced water and were reused repeatedly for 10 cycles, due to the sorbent-degrader synergies [58]. The microbial consortia consisted of salt-tolerant and biosurfactant producers from Dietzia sp. CN-3 and Acinetobacter sp. HC8–3S had achieved 95.8% degradation efficiency of petroleum, whereby alkane hydroxylase genes alkB and alkM degraded the medium-chain alkane C14-C26 [59]. Immobilized indigenous bacteria with high adaptability to petroleum-refining oil sludge showed a great advantage, with an 88.78% degradation of petroleum due to the high catalytic activity of enzymes [60]. Immobilized microorganisms may suffer from the inconsistency of degradation performance because only the external cells exhibit high metabolic activity, while the cells in the support pores are unable to access the hydrocarbon [61]. These issues were solved with the assistance of a bioemulsifier agent, produced by the biocatalyst from Aspergillus brasiliensis, when exposed to the electrical field. The in situ biosurfactant production modified the surface properties and sorption capacity, improving the hydrocarbon attachment [62]. Very recent studies have used transposon mutagenesis technology to increase the biosurfactant production of Enterobacter xiangfangensis [63] and Enterobacter hormaechei [64] mutants to enhance the bioavailability of petroleum. As a result, the degradation of recalcitrant petroleum in oil sludge was achieved to a high degree.
These studies demonstrate the high efficacy of enzymes as bioremediation agents in polluted sites where the hydrocarbon load is low (Table 2). On the contrary, treating waxy hydrocarbons in production lines is more challenging than the bioremediation of hydrocarbon-polluted sites. The treatment requires enzymes to perform in a non-aqueous, hydrocarbon-rich environment that is probably hostile to the enzymes. In this scenario, petroleum has a higher fraction of high-molecular-weight carbon that makes up the wax composition, limiting the susceptibility of the enzymes to the hydrocarbon. The next section describes the biodegradation of specific waxy petroleum and the enzymes taking part in the reaction.

3.2. Biodegradation of Waxy Petroleum Hydrocarbons

Waxy hydrocarbons are associated with long-chain alkane compounds of C17 to C55 that tend to deposit on the cold surface [78]. According to a study, methane-monooxygenase enzymes oxidize short-chain alkanes C1–C4, cytochrome P450 enzymes oxidize medium-chain alkanes C5–C16, and essentially unknown enzyme systems oxidize long-chain alkanes C17+ [79]. However, a study on the biocatalytic degradation of high-paraffinic crude oil has characterized long-chain alkane hydroxylase homologues AlkMa and AlkMb and encoded a putative flavin-binding monooxygenase, AlmA [80]. AlkB-type alkane hydroxylase homologues alkMa and alkMb, from Alcanivorax borkumensis SK2 [81,82] and Acinetobacter DSM 17874 [83], have been demonstrated to be involved in the degradation of n-alkanes with chain lengths of C10 to C20. It was also reported that AlkB genes in Anoxybacillus sp. WJ-4 were responsible for a 58.75% degradation of C8 to C22 [84]. It has been discovered that the cytochrome P450 monooxygenases gene in Alcanivorax dieselolei B-5 can grow on C6 to C12. [85]. The flavin-dependent monooxygenase LadA was discovered in Geobacillus thermodenitrificans NG80-2, which degrades C15 to C36 [86]. AlmA is the first cloned gene proven to be involved in the bacterial degradation of long-chain n-alkanes of C32 and longer [83]. Other than alkane hydrolase and monooxygenase, lipase in Stenotrophomonas maltophilia was found able to degrade 94.13% of diesel oil, which was mainly composed of C12 to C20 [87]. Asphaltene, a fraction of petroleum, has a complex structure; hence, it is difficult to break down. A study reported that catechol dioxygenase and laccase in Pestalotiopsis sp. NG007 played an important role in the 77% degradation of C13-C30 asphalt [88]. Moreover, 82–90% of C32 and 82–88% of long-chain alkane C40 were degraded by thermophile bacteria strains Geobacillus stearothermophilus IR2, Geobacillus thermoparaffinivorans IR4 and Bacillus licheniformis MN6. The enzyme assays showed the presence of alkane hydroxylase, alcohol dehydrogenase and lipase [89]. A study on the degradation rate of motor oil using Alcanivorax borkumensis successfully achieved a 75% degradation due to the production of biosurfactants, which allowed the media to mix well with the hydrocarbon sample through reduction of interfacial tension, thus allowing degradation to happen [90]. These studies highlight a few important hydrocarbon-degrading enzymes that can be potentially sourced for the purpose of managing wax deposition in transportation lines (Table 3).

4. Biodegrading Mechanism of Petroleum Hydrocarbons

The degradation of petroleum hydrocarbons is a complex process involving a variety of enzymes that work synergistically to break down the complex hydrocarbon molecules into simpler compounds. Some of the key enzymes in hydrocarbon degradation are the oxygenases, dehydrogenases, hydrolases and reductases. Oxygenases are enzymes that introduce one or two oxygen atoms derived from molecular oxygen in the alkane substrate [102], which are monooxygenases and dioxygenases, respectively. Alkane hydroxylases and cytochrome P450 monooxygenases are an example of monooxygenases that have been reported widely. Dioxygenases such as catechol dioxygenase and naphthalene dioxygenase incorporated both atoms of molecular oxygen to cause a ring cleavage in an aromatic compound [103]. Dehydrogenases are the enzymes that catalyze the removal of hydrogen atoms from the hydrocarbon molecules, leading to a formation of oxidized products and facilitating the further breakdown of the hydrocarbon. Alcohol dehydrogenase and aldehyde dehydrogenase have been reported in this category [104,105]. Hydrolases catalyze the hydrolysis reaction, which involves the cleavage of a chemical bond using water. For example, esterase breaks down ester linkages present in some of the hydrocarbon derivatives [106]. Finally, reductases catalyze the addition of hydrogen atoms to a certain hydrocarbon intermediate, where the reduction step helps to activate the molecule for further degradation processes. As an example, carbonyl reductase catalyzes the reduction of carbonyl groups (C=O) present in aldehyde and ketones, and the products can be further metabolized by other enzymes [107,108].
Alkanes are also known as paraffins, which translates to the alkanes having low chemical reactivity [109]. To biodegrade a petroleum, the alkane molecules need to be activated to initiate and continue the stepwise metabolisms [110]. The regio- and stereoselective oxidation of non-activated methyl or methylene groups is a challenging, but it is a crucial chemical process [111]. Alkane degradation can be achieved via oxygen-dependent oxygenase, which triggers the initial activation and breaking of C-H bonds in aerobic conditions [112].
Alkane degradation in microorganisms had been described and illustrated by three peripheral metabolic pathways, which include terminal oxidation, subterminal oxidation, and ω-oxidation (Figure 4). The methyl group of alkanes is oxidized by the alkane-activation enzyme alkane monooxygenase (AMO), involved in both the terminal and subterminal pathways. The reactions render a primary alcohol in terminal oxidation and a secondary alcohol in subterminal oxidation. Alcohol dehydrogenase (ADH) converts the alcohol into aldehydes (for terminal) and ketones (for subterminal). Aldehydes are further oxidized to fatty acid by aldehyde dehydrogenase (ADH), while ketones are oxidized by Baeyer–Villiger monooxygenase (BVMO) to render esters, which are subsequently hydrolyzed to alcohol and fatty acid by carboxylesterase (CE). The fatty acids are conjugated to coenzyme A and enter the β-oxidation cycle to generate acetyl-CoA [113,114]. Alkane molecules can also be oxidized at both ends through the ω-hydroxylation of fatty acids at the terminal ω-methyl group, rendering an ω-hydroxy fatty acid that is further converted into a dicarboxylic acid and processed by β-oxidation. The subterminal oxidation of alkanes generates a secondary alcohol, which is converted to the corresponding ketone and then oxidized by a Baeyer–Villiger monooxygenase to render an ester. The ester is hydrolyzed by an esterase, generating an alcohol and a fatty acid, and enters β-oxidation. β-oxidation plays a crucial role in fatty acid metabolism and energy production, but it is not directly involved in the breakdown of alkane molecules during alkane degradation. Both terminal and subterminal oxidations can coexist in the microorganisms [113,114,115].
The research related to the biodegradation of wax has led to the discovery of many bacteria species hosting enzymes involved in the initial step of aerobic degradation pathways. Acinetobacter sp., Alcanivorax sp., Arthrobacter sp., Bacillus sp., Dietzia sp., Geobacillus sp., Marinobacter sp., Mycobacterium sp., Pseudomas sp., and Rhodococcus sp. are among the bacterial strains capable of utilizing C18 and more [116]. A similar discovery is that alkanes are oxidized to the corresponding primary and secondary alcohol by substrate-specific terminal and subterminal monooxygenases or alkane hydroxylases. Alkane hydroxylases are flexible biocatalysts that perform a variety of beneficial oxidation reactions. Pseudomonas putida GPo1 encoded the functional alkane hydroxylase that oxidizes C5 to C16 [117]. AlkB-type enzymes work with two electron-transfer proteins—a dinuclear iron rubredoxin and a mononuclear iron rubredoxin reductase—to transfer electrons from NADH to the active site of alkane hydroxylase [113]. Acinetobacter sp. Strain M-1 was shown to harbor two integral membrane alkane hydroxylases responsible for the degradation of C10 to C30 [118]. Interestingly, AlkB, found in Acinetobacter oleivorans DR1, was able to utilize long-chain alkanes C12 to C36, but not short-chain alkanes C6 to C10. AlkB1 was responsible for long-chain alkanes C24 to C26, whereas AlkB2 was responsible for medium-chain alkanes C12 to C16 [119]. The recombinant strain Rhodococcus erythtopolis AP-expressing AlkB gene cluster was experimentally confirmed to degrade C12 to C24 [120]. The study comprehensively reported that the alkB gene cluster includes an alkB gene encoded for alkane hydroxylase, two rubredoxins (A and B), and rubredoxin reductase. The proteins rubredoxin and rubredoxin reductase are required to shuttle electrons to AlkB, which the enzyme uses for alkane hydroxylation [120]. Five proteins were expressed by Thalassoliitus oleivorans hosting the terminal oxidation of C14, which were characterized as alkane 1-monooxygenase, oxidoreductase ferredoxin, ferredoxin reductase, alcohol dehydrogenase, and aldehyde hydrogenase. The ferredoxin reductase oxidized NAD(P)H to NAD(P)+, generating electrons that were transferred to ferredoxin, which shuttled the electrons to alkane monooxygenase. Alkane monooxygenase introduces oxygen into alkanes at the terminal site, converting them into primary alcohol. The alcohol is further oxidized to aldehyde and fatty acids by alcohol dehydrogenase and aldehyde dehydrogenase. The pathway was then possibly switched to subterminal oxidation when grown on the longer-chain C28, evidenced by the significant upregulation of Baeyer–Villiger monooxygenase and esterase, responsible for catalyzing ketones and ester metabolism [121]. Long-chain thermophilic alkane monooxygenase was discovered when a genome and proteome analysis of Geobacillus thermodenitrificans strain NG80-2, isolated from a deep-subsurface oil reservoir [122], revealed a plasmid-encoded thermophilic enzyme designated as LadA. LadA was determined to be involved in the terminal oxidation of long-chain alkanes able to convert C15 to C36 to primary alcohol. The research highlighted many attributes of LadA: it was a thermophile, able to act on long-chain alkanes, a single-component with no coenzyme requirement, soluble (extracellular), and easily expressed and purified in E. coli, making it a great biocatalyst for industrial applications [86].

5. Biocatalytic Degradation of Petroleum Hydrocarbons—Opportunity for Application and Gaps in Know-How

The petroleum industry is looking into energetically efficient processes with minimum environmental impact. A novel and nonconventional technique like enzyme catalysis for hydrocarbon management has been explored. With properties such as high transformation efficiency and high specificity, enzymes can be utilized not only for cleaning up oil-contaminated water and soil but also for the treatment of viscous heavy oil for transportability, as well as for refining petroleum [102].
The direct application of microbial metabolites such as enzyme extracts or recombinant enzymes, in contrast to whole organisms, is advantageous for the petroleum industry. This is attributed to better process control in terms of enzyme concentration, faster reactions, the ability to remediate complex and persistent compounds, and specificities that enable enzymes to perform at wide range of temperature, pH and salinity levels [123,124]. Enzymes are also biodegradable proteins, so there is no environmental persistence problem. Furthermore, recombinant DNA technology has been introduced to improve the stability and activity of enzymes at a larger scale and lower cost [125]. The conversion or transformation of inert nonpolar hydrocarbons is a challenging biochemical reaction for a single enzyme but is tolerable for multi-enzyme systems [126]. Research has recently reported the usage of an enzyme consortium consisting of oxygen-dependent oxygenase, which is able to catalyze the cleavage of the C-H bond in long-chain wax in hydrocarbons into carbon of a shorter chain, resulting in improved flowability [127]. The concept is illustrated in Figure 5.
Table 4 summarizes the research findings from a reaction system consisting of enzyme consortiums extracted from microbes and their targeted hydrocarbon substrate. It was evidenced from these studies that intracellular and extracellular enzymes are capable of degrading up to 99% of the hydrocarbon.
However, similar conditions were applied in most of the studies: a long treatment time, i.e., more than 24 h; the petroleum substrates used are mostly of short- and medium-chain alkanes, less than C30; and essentially, the partial degradation of wax or asphaltene leads to a reduction in the viscosity of heavy petroleum [128]. However, the analyses mainly discuss the degradation products. Elaborations on the changes in physical or flow properties of the reacted petroleum, i.e., viscosity, morphology, waxiness, are very limited. On contrary, the application of the enzymatic degradation of wax in production lines needs to occur in hydrocarbon-rich environments (non-aqueous system) in a short treatment time, because the petroleum-production process is a continuous operation. In addition, waxy crude oils are composed of a high percentage of chain alkanes more than C30. Therefore, these limitations needed to be studied in detail.
Table 4. Enzymatic degradation of hydrocarbon.
Table 4. Enzymatic degradation of hydrocarbon.
Hydrocarbon-Degrading EnzymeSubstrateTreatment ConditionHydrocarbon Removal Treatment Efficiency Reference
Intracellular crude enzyme consortium from Providencia rettgeri L1 supplemented with formate dehydrogenasePetroleum residual oil 10 mL reaction system containing crude enzymes and residual oil at 30 °C for 8 h23.8% oil degraded at 1 h; 56.7% at 5 h; 49.5% at 8 h[129]
Enzyme consortium of alkane hydroxylase and lipase from Alcanivorax borkumensis SK2Soil contaminated with diesel fuel5000 L enzyme solution injected into 4 injection wells at 14 lpm for 12 weeks36–99% degradation of C10-C50[130]
Protease, catalase, lipase, and amylase from peel wastes of sweet orange Citrus sinensis Osbeck and watermelon Citrullus lanatusUsed motor oil-contaminated soilsEnzyme solutions added to the contaminated soils for 6 weeks62–74% TOC removal by orange peel enzymes; 39–45% TOC removal by watermelon peel enzymes[131]
Intracellular enzymes from Acinetobacter calcoaceticus 21 and formate dehydrogenase CbFDH from Candida boidinii Oil-contaminated sludge 2 g oily sludge with 10% oil by weight reacted with enzyme 21 and CbFDH at 30 °C at 150 rpm35.6% oil degraded in 12 h using enzyme 21/CbFDH with the protein ratio of 1:4[132]
Ligninolytic enzyme Laccase and
Manganese Peroxidase microencapsulated in bilayer cross-linked Ca-alginate beads
Polluted sediment Incubated in the dark at 25–27 °C with vertical mixing at 50 rpm for 70 days27–28% total petroleum hydrocarbon degraded [133]
Intracellular and extracellular enzymes from Aspergillus sp. RFC-1Oilfield crude oil Extracellular and intracellular enzymes inoculated into 20 mg/L crude oil, cultivated at 30 °C at 120 rpm for 7 daysDegradation efficiencies of crude oil by extracellular enzymes were higher (7–25%) than by intracellular enzymes (5–24%)[134]
Alkane hydroxylase, lipase and esterase from Alcanivorax borkumensis crude enzyme preparationHexadecane, motor oil, contaminated soil50 mL Milli-Q water, 10 mg/mL crude enzymes and petroleum sources, incubated at 30 °C for 7 days at 100 rpmRemoval of hexadecane (73.75% to 59.74%); motor oil (74% to 83%); contaminated soil (88.52%)[37]
Fungal enzymes Aspergillus fumigatus (PJ1, PJ2, PJ3 and PJ5), Aspergillus flavus (PJ4), and Aspergillus terreus (PJ6)Oil-contaminated soil 2 g crude oil and 30 mL of crude enzyme solution in 100 mL glass bottles sealed with rubber stoppers, incubated statically at 40 °C for 4 days, with regular shaking every 4 h under oxygen deprived conditions3.70–15.68% removal of alkane, 23.33–40.56% of aromatics, 13.33–35.56% of resins, and 24.69–34.57% of asphaltenes. Oil viscosity reduced by 40.5–59.0%. Total oil removal efficiency of 83.40% to 87.78%[135]
Dioxygenase immobilized onto single-walled carbon nanotube from Arthrobacter chlorophenolicus A6Aromatic hydrocarbon intermediates catechol, 4-chlorocatechol and 3-methylcatechol)Free and immobilized enzymes added to substrate solution and H2O2 as oxygen source at molar ratio 1:2, then mixed at 25 °C for 1 min. The precipitate was separated by centrifugation.Relative activity of immobilized enzyme is the highest at pH 7.5, temperature 40 °C, and 0.6 M salt; and retained 40% of activity after 7 cyles of reusability[103]
Versatile peroxidase (VP) from white-rot fungus Bjerkandera adustaAntracene in silicone oilTwo-phase partitioning bioreactor with 250 mL reaction medium containing 10% silicone oil saturated with anthracene at 30 °C61.88% oxidation of anthracene after 38 h[123]
Intracellular and extracellular enzymes from Pseudomonas sp., Bacillus sp., Ochrobactrum sp Dehydrated crude oil Extracellular enzyme inoculated into 20 mg/L of crude oil, cultivated at 30 °C g at 120 rpm for 18 days6–15% oil degradation in extracellular enzyme; 4–15% in intracellular enzyme[136]

6. Challenges of Applying Enzymatic Degradation to Petroleum Industry

One of the obstacles that hinders the enzymes from replacing chemicals for industrial applications is the high cost of production and purification. There are a few factors contributing to this. Firstly, enzymes undergo denaturation during production and storage. The unfolding of the enzyme tertiary structure to a disordered polypeptide causes the residues to be unaligned, therefore hindering the interactions in its functional groups. Secondly, the presence of other chemicals in the petroleum-production system causes an irreversible loss of activity [137]. During oil and gas operations, enzymes are prone to denaturation and loss of activity because of three factors: temperature and pressure stability at the hydrocarbon production facilities, e.g., reservoirs, production flowlines and transportation pipelines; high salt stability due to the carry-over of seawater; and stability in non-aqueous media, i.e., a hydrocarbon-rich environment. Research has also highlighted that biocatalysts should be kinetically efficient and have little dependency on enzyme cofactors to be cost-competitive [138]. However, with the advancement of proteomics and molecular biology, enzyme production can be optimized to resist and withstand petroleum-production’s operating conditions.
Apart from the cost competitiveness between enzymes and chemicals for wax treatment, the physicochemical characteristics of hydrocarbons are also a challenge. Hydrocarbon wax is nonpolar, while asphaltene and resin are polar constituents that contain both normal and polycyclic aromatic hydrocarbons [139]. A polyaromatic hydrocarbon is chemically stable due to resonance energy and the high number of aromatic rings [75]. Therefore, petroleum hydrocarbons with high asphaltene and resin content may be associated with slower biodegradation. Petroleum hydrocarbons are chemically inert, and high energy barriers must be overcome to initiate the activation and cleavage of carbon–hydrogen bonds. The hydrophobic property of the petroleum is the main obstacle for the enzymatic attack of the carbon backbone [140]. In addition, petroleum hydrocarbons contain crystalline, semi-crystalline and amorphous regions. Only the amorphous and semi-crystalline regions are more susceptible to enzymatic degradability, as illustrated in Figure 6.
For enzymatic degradation to be successful, one must identify the most suitable enzyme for that application. For example, extracellular enzymes with high substrate affinities can produce high-product turnover, and the ability to maintain stable activity under the operational conditions of the intended application is preferable [141]. However, the enzymatic performance depends on the physical conditions of the sites, such as temperature, pH, and salinity, and the enzyme properties, such as catalytic sites and enzyme activity [124]. The enzymes are exposed to the harsh conditions of high temperature, pressure, and fluid turbulence. The movement of petroleum from a reservoir to the surface is a continuous process. The temperature of a freshly produced hydrocarbon can be as high as 60 to 70 °C, while the pH is typically 7 to 9 [1]. Apart from temperature and pH factor, the catalytic activity of the selected enzymes must not be inhibited by metals. Salts, or dissolved metal ions present in seawater, are produced together with the petroleum from the reservoir containing Na+, Mg2+ and Ca2+ as the major constituents, and the composition is varied for different reservoirs [142]. Under an aerobic environment, the C-H bond activation can be catalyzed by the oxygen-dependent oxygenase, which is ubiquitous in the case of petroleum-polluted sites [103]. The dissolved oxygen from the produced water, which is entrained with petroleum during production, is available for the oxygenase during wax degradation [143,144].

7. Current Approaches and Future Outlook of Applying Enzymatic Degradation for Wax Management in Petroleum Operations

Enzyme stability and efficiency are the main challenges for industrial application; therefore, powerful tools of molecular biochemistry should be used for improvement. Enzymes isolated from extremophilic microorganisms that are resistant to pH, high salinity, and organic solvents; thermostable; and thrive in a reduced oxygen environment [145] are the most beneficial for the petroleum industry. Extremophilic enzymes derived from Halobacterium sp., Haloferax sp., Haloarcula sp. and Pyrococcus sp. have been reported to have salt-adapted ability, while Rhodoccus sp. and Pseudomonas sp. are reported to host organic-solvent active enzymes [146]. Enzymes adapted to high-salt environments, known as halophilic enzymes, can retain their functional conformation in the presence of high ionic strength, and they possess stable, multilayered hydration shells that help to preserve their biological functions [146]. The new discovery of thermostable enzymes from microorganisms living in extremely high temperatures of ~100–200 °C, such as deep reservoirs, is pertinent. Enzymes from the genus Bacillus sp. are commonly reported as being thermophilic [89,147]. Other than Bacillus sp., thermophilic enzymes can also be derived from Thermomicrobium sp. and Thermooleophilum sp., and dominantly from Geobacillus sp., and are typically active at 57 to 80 °C [148]. This is an important physicochemical property for enzyme application as petroleum facilities, from production to refining, typically operate at high temperatures and salinity. In addition, the enzyme catalytic treatment of petroleum performed at high temperatures is beneficial for high-viscosity substrates. At high temperatures, the viscosity of the petroleum is reduced, leading to an increase in diffusion coefficients; therefore, the bioavailability of the enzymes is improved [148]. Enzymes should be able to operate in non-aqueous system due to the hydrophobicity of petroleum, thereby limiting the mass transfer and interaction between the enzymes and the hydrophobic substrates. This probably can be solved using reaction mixtures containing organic solvents to increase the substrate accessibility, but the high enzyme activity must be maintained, as supported by [149]. In addition, the enzyme–wax interaction can be facilitated by specific structural features of the enzyme, such as hydrophobic pockets or active sites that can accommodate nonpolar molecules [150]. In some cases, enzymes can be immobilized on hydrophobic supports to enhance the interaction with hydrophobic substrates [151]. On top of that, there is also research that provides a positive indication of using enzymes from the genus Pseudomanas sp. for the biodegradation of plastic polymers such as polyethylene, polyethylene glycol and polyethylene terephthalate [152]. Furthermore, 95% of polyethylene terephthalate was found to be degraded in a heterogenous enzymatic reaction [153]. In the case of wax treatment in a production line, a similar approach can be adopted due to the resemblance in characteristics between plastic polymers and waxy crude oil, such as hydrophobicity, high molecular weight, and lack of favorable functional groups. To find an adequate solution for petroleum hydrocarbons, the enzyme complexes’ behavior, metabolites, and degradation pathways must be explored [154].
The industrial application of biocatalysts is often restricted by the short-term operational stability and difficulty in the recovery of enzymes. However, these limitations are overcome by immobilization, although there is no specific remedy protocol because each enzyme has different properties. The primary objective of enzyme immobilization is to increase the economics of the biocatalytic process. The immobilization of enzymes is a highly valuable technique that was introduced to overcome the limitation of free enzymes. The immobilization technique has evolved to a simple, fast, and efficient process [155]. Covalent binding, physical adsorption, or enzyme entrapment through immobilization lead to high number of enzymes being in contact with the high surface area of support, subsequently discouraging the dissociation of protein into subunits or the formation of inactive intermolecular aggregates [156]. Immobilization supports like silica encapsulation protect microorganisms from lysis and harsh environments of high salinity and organic load, in addition to concentrating their amount, thus accelerating the degradation process [124]. The potential applications of immobilized enzymes are broadened because of their high versatility such as high enzyme/substrate ratio, increased functional efficiency of the enzyme, minimized reaction time, and minimized contamination in products [157]. Immobilized enzymes have an enhanced tolerance to variations in environmental factors, such as temperature, pH, organic solvents, and long storage [158]. Immobilized enzymes are easy to control, whereby the reaction can be stopped by physical removal rather than by heat inactivation when using soluble or free enzymes [159].
The stability of immobilized enzymes concerning temperature and pH is widely reported in the literature to translate the benefit of the immobilization technique for soluble or free enzymes [160]. For example, Mazlan (2017) immobilized laccase on polymer microspheres. They found that the optimum temperature for laccase is extended from 40 °C (in free form) to 50 °C (in immobilized form), while the optimum pH is shifted from 4 (in free form) to 5 (in immobilized form). The reasons is that the covalent bond formation, through the amino groups of the immobilized enzyme, causes the molecule to have a higher activation energy for the better reorganization and conformation of substrate binding. Also, the multipoint attachments of the enzyme support increased enzyme stabilization; thus, an extended optimum temperature and pH were observed [161]. The immobilization of horseradish peroxidase on iron oxide magnetic nanoparticles using physical bonding also produced a similar optimum temperature and pH shifting [162,163]. Oxidoreductases are not stable under various conditions, whereby pressure, temperature and pH may cause a dissociation of the enzyme subunit. Basri (2022) immobilized Mycobacterium phlei carboxylic acid reductase (MpCAR) onto polymeric support Seplite LX120 via an adsorption technique to achieve 3 weeks of storage stability at an ambient temperature. The optimum reaction temperature and pH of the MpCAR were shifted from 42 °C and 7.5 (in free enzyme) to 60 °C and 9 (in immobilized form) [164]. An immobilized aldehyde dehydrogenase from Anoxybacillus geothermalis D9 exhibited an improvement in temperature stability from 30 to 90 °C in contrast to its free form [104]. Immobilized carboxylesterase from A. geothermalis yielded a broad thermal stability and pH tolerance, with an optimal temperature of 80 °C and an optimal pH of 7 (Johan et al., 2023 [165]). Bolivar (2012) elaborated that under soluble or free conditions, enzyme subunits dissociate under an acidic condition, whilst tertiary and quaternary structures are distorted under an alkaline condition [166]. Through enzyme immobilization, the conformational freedom of the enzyme is narrowed due to the multipoint covalent attachment of the enzyme that causes a reduction in the intermolecular reaction; thus, the pH stability is enhanced [167]. pH affects the ionization of functional groups within enzymes, and when immobilized, the enzyme has a better orientation of active sites and thus a higher affinity towards substrates, as reflected in the higher activity [163]. The rigidity in conformation also enables enzymes to withstand higher temperatures for a longer time [168,169,170]. Zhou (2001) cross-linked ß-galactosidase to graphite surfaces and studied the good shifting of optimum temperature and pH. They provide a more straightforward explanation centered around diffusional effects, in which the immobilized enzyme is easily contacted by substrates with the increase in temperature and pH [171].
However, the success rate of enzyme immobilization to produce high catalytic activity and the stability of enzymes depends on immobilization time, procedure, surface coverage, surface curves, active sites orientation, mass transfer and chemical bonding [158]. A study has pointed out that the porosity and particle size of the immobilized support could influence the mass transfer kinetics between the reactant (the hydrocarbon) and the enzymes. A maximum efficacy of the reactant diffusivity may exist, related to the particle size of the immobilized enzymes, and need to be determined experimentally [62].
Enzymes are always associated with high costs; therefore, efforts should be made to recover and reuse them instead of continuously putting them in production lines, which is a more suitable practice for chemicals. Enzyme immobilization provides an opportunity for separation and reuse because immobilized enzymes are stationary throughout the application process instead of mobile with the reactant. The recovery of enzymes makes the application cost-attractive because of the reduction in production costs, labor costs and space requirements [172]. Where the productivity over time and volume of hydrocarbons to be treated is high, fluidized bed reactors are preferred (Figure 7). Immobilized enzyme particles can be sustained in suspension by the flowing stream of hydrocarbons [173].

8. Conclusions

This paper reviews the potential of exploiting enzymes to tackle the wax deposition problem in petroleum-production lines. Enzymes’ catalytic activity can be enhanced in the laboratory to suit specific applications. Research evidence on the microbial degradation of petroleum-contaminated sites successfully highlighted AlkB homologs, alkane hydroxylases, alkane monooxygenases, lipase, laccase and esterase as highly capable hydrocarbon-degrading enzymes. Based on this precursor, this review narrowed down the research and focused on papers related to the microbial degradation of specific alkane compounds and crude oil. A similar group of enzymes was identified and was reported to be able to remove 12–96% of hydrocarbons in the alkane compound in 3–42 days. Based on these studies, it was evidenced that hydrocarbon-degrading enzymes are a good alternative for wax deposition treatment in transportation lines. This paper then discussed the possible challenges for the enzymes when applied in wax deposition. The costly production of enzymes is the main factor, attributed to enzyme denaturation and the presence of unexpected chemicals that reduce catalytic activity. The enzymes’ dependency on cofactors may also reduce their cost-competitiveness in comparison to chemical treatment. Secondly, the hydrophobicity of petroleum hinders enzymatic degradation. Petroleum hydrocarbons are chemically inert, and high energy barriers must be overcome to initiate the activation and cleavage of the carbon–hydrogen bond. Thirdly, enzymes are subjected to the harsh conditions of high temperature, pressure, and fluid turbulence in production lines; therefore, enzyme performance may be deteriorated. However, the challenges can be overcome by the combination of technological advancement and application strategies. The current approaches and prospects for enzyme application in the petroleum industry were discussed next. Recombinant enzymes with thermophilic and halotolerant properties have been discovered in petroleum degradation, although research is scarce. Hence, efforts should be made to enrich this discovery. The direct application of enzyme extracts, in contrast to whole organisms, is advantageous for the petroleum industry in terms of better process control, faster reactions, and the ability to perform in wide range of temperature, pH and salinity. Additionally, using a consortium of enzymes is favorable since the transformation of inert nonpolar hydrocarbons is a challenging biochemical reaction for a single enzyme but is tolerable for multi-enzyme systems. With the usage of powerful tools of molecular biochemistry, enzyme stability and efficiency can also be improved. Enzymes isolated from extremophilic microorganisms that are resistant to pH and organic solvents, thermostable, and thrive in a reduced-oxygen environment are the most beneficial for the petroleum industry. The advances in the enzyme immobilization technique have significantly improved enzymes’ susceptibility to the harsh conditions of the transportation pipeline. This technique also allows for the recovery and reuse of enzymes, making the enzyme cost-competitive with other wax treatment strategies. This paper intended to contribute to sustainability by discussing the potential of the enzymatic degradation approach being expanded to the petroleum-production process.

Author Contributions

Conceptualization, S.S. (Shazleen Saadon) and R.N.Z.R.A.R.; writing—original draft preparation, S.S. (Shazleen Saadon); writing—review and editing, S.S. (Sara Shahruddin), N.H.A.K., N.A.W. and S.R.M.S.; supervision, M.S.M.A.; funding acquisition, S.S. (Sara Shahruddin). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by PETRONAS Research Sdn Bhd under an R&D collaboration agreement between Petroliam Nasional Berhad (PETRONAS) and UPM Consultancy & Services for “Development of robust non-chemical to lower the pour point of crude oil”. The APC was funded by PETRONAS Research Sdn Bhd.

Acknowledgments

The authors express sincere gratitude to Petroliam Nasional Berhad (PETRONAS) for financially support the research.

Conflicts of Interest

Authors Shazleen Saadon, Sara Shahruddin Siti Rohaida Mohd Shafian, Norhidayah Ahmad Wazir were employed by the company PETRONAS Research Sdn Bhd. The authors declare that the research was funded by PETRONAS Research Sdn Bhd. The funders and the remaining authors were involved in the design of study, the collection, analysis, and interpretation of data, and the writing of the manuscript. However, funders have no objections to the publication of the results and collaborated with the remaining authors in the preparation of the manuscript.

References

  1. Havard, D. Oil and Gas Production Handbook—An Introduction to Oil and Gas Production, Transport, Refining and Petrochemical Industry; Lulu: Morrisville, NC, USA, 2013. [Google Scholar]
  2. Tong, K.; Lin, A.; Ji, G.; Wang, D.; Wang, X. The effects of adsorbing organic pollutants from super heavy oil wastewater by lignite activated coke. J. Hazard. Mater. 2016, 308, 113–119. [Google Scholar] [CrossRef] [PubMed]
  3. Das, N.; Chandran, P. Microbial Degradation of Petroleum Hydrocarbon Contaminants: An Overview. Biotechnol. Res. Int. 2011, 2011, 941810. [Google Scholar] [CrossRef] [PubMed]
  4. Bissada, K.K.A.; Tan, J.; Szymczyk, E.; Darnell, M.; Mei, M.; Zhou, J. Group-type characterization of crude oil and bitumen. Part I: Enhanced separation and quantification of saturates, aromatics, resins and asphaltenes (SARA). Org. Geochem. 2016, 95, 21–28. [Google Scholar] [CrossRef]
  5. Sah, D.; Rai, J.P.N.; Ghosh, A.; Chakraborty, M. A Review on Biosurfactant Producing Bacteria for Remediation of Petroleum Contaminated Soils; Springer: Berlin/Heidelberg, Germany, 2022; Volume 12. [Google Scholar] [CrossRef]
  6. Fan, C.Y.; Krishnamurthy, S. Enzymes for enhancing bioremediation of petroleum-contaminated soils: A brief review. J. Air Waste Manag. Assoc. 1995, 45, 453–460. [Google Scholar] [CrossRef] [PubMed]
  7. Chi, Y.; Yang, J.; Sarica, C.; Daraboina, N. A Critical Review of Controlling Paraffin Deposition in Production Lines Using Chemicals. Energy Fuels 2019, 33, 2797–2809. [Google Scholar] [CrossRef]
  8. Wang, Z.; Yu, X.; Li, J.; Wang, J.; Zhang, L. The use of biobased surfactant obtained by enzymatic syntheses forwax deposition inhibition and drag reduction in crude oil pipelines. Catalysts 2016, 6, 61. [Google Scholar] [CrossRef]
  9. Theyab, M.; Theyab, M.A.; Petroleum, S.J.; Muhammad; Theyab, A. Fluid Flow Assurance Issues: Literature Review Fluid Flow Assurance Issues: Literature Review View project SciFed Journal of Petroleum Fluid Flow Assurance Issues: Literature Review. SciFed J. Pet. 2018, 2, 1–11. Available online: https://www.researchgate.net/publication/323733946 (accessed on 3 August 2024).
  10. Mehrotra, A.K.; Bhat, N.V. Deposition from ‘waxy’ Mixtures under turbulent flow in pipelines: Inclusion of a viscoplastic deformation model for deposit aging. Energy Fuels 2010, 24, 2240–2248. [Google Scholar] [CrossRef]
  11. Bai, J.; Jin, X.; Wu, J.T. Multifunctional Anti-Wax Coatings for Paraffin Control in Oil Pipelines; China University of Petroleum Beijing: Beijing, China, 2019. [Google Scholar] [CrossRef]
  12. Sousa, A.L.; Matos, H.A.; Guerreiro, L.P. Preventing and removing wax deposition inside vertical wells: A review. J. Pet. Explor. Prod. Technol. 2019, 9, 2091–2107. [Google Scholar] [CrossRef]
  13. Huang, Q.; Li, H.; Zhuang, Y.; Ding, Y.; Ma, C.; Xie, Y.; Liang, H.; Han, S.; Zhang, J. Reducing viscosity of waxy crude oil with electric field perpendicular to oil’s flow direction. Fuel 2021, 283, 119345. [Google Scholar] [CrossRef]
  14. Wang, W.; Huang, Q. Prediction for wax deposition in oil pipelines validated by field pigging. J. Energy Inst. 2014, 87, 196–207. [Google Scholar] [CrossRef]
  15. Guo, Y.; Liu, S.; Tan, G.; Yang, L.; Wang, D. Removal and tribological behaviors of waxy deposition layer in cleaning process. Ind. Lubr. Tribol. 2017, 69, 566–573. [Google Scholar] [CrossRef]
  16. Tan, G.B.; Liu, S.H.; Wang, D.G.; Zhang, S.W. Spatio-temporal structure in wax-oil gel scraping at a soft tribological contact. Tribol. Int. 2015, 88, 236–251. [Google Scholar] [CrossRef]
  17. Taheri-Shakib, J.; Shekarifard, A.; Kazemzadeh, E.; Naderi, H.; Rajabi-Kochi, M. Characterization of the wax precipitation in Iranian crude oil based on wax appearance temperature (WAT): The influence of ultrasonic waves. J. Mol. Struct. 2020, 1202, 127239. [Google Scholar] [CrossRef]
  18. Taheri-Shakib, J.; Shekarifard, A.; Naderi, H. Characterization of the wax precipitation in Iranian crude oil based on Wax Appearance Temperature (WAT): Part 1. The influence of electromagnetic waves. J. Pet. Sci. Eng. 2018, 161, 530–540. [Google Scholar] [CrossRef]
  19. Tripathy, A.; Nath, G.; Paikaray, R. Effect of ultrasonication in treatment of crude oil. Indian. J. Pure Appl. Phys. 2018, 56, 914–919. [Google Scholar]
  20. Xu, X.; Bao, T. Research on the removal of near-well blockage caused by asphaltene deposition using sonochemical method. Ultrason. Sonochem. 2020, 64, 104918. [Google Scholar] [CrossRef]
  21. Mo, L.; Sun, W.; Jiang, S.; Zhao, X.; Ma, H.; Liu, B.; Feng, L. Removal of colloidal precipitation plugging with high-power ultrasound. Ultrason. Sonochem. 2020, 69, 105259. [Google Scholar] [CrossRef]
  22. Alade, O.; Abdullah, E.; Alfaraj, M.; Al Hamad, J.; Hassan, A.; Mahmoud, M.; Al Shehri, D.; Solling, T.I.; Nakhli, A. Removal of Wax Deposit from Tight Formation Using Ultrasonic Cavitation with Thermochemical Heat Source. In Proceedings of the SPE Annual Technical Conference and Exhibition, Houston, TX, USA, 3–5 October 2022; Society of Petroleum Engineers (SPE): Houston, TX, USA, 2022. [Google Scholar] [CrossRef]
  23. Banerjee, S.; Kumar, S.; Mandal, A.; Naiya, T.K. Design of novel chemical solvent for treatment of waxy crude. Int. J. Oil Gas Coal Technol. 2017, 15, 363. [Google Scholar] [CrossRef]
  24. Theyab, M.A.; Diaz, P. Experimental study of wax deposition in pipeline—Effect of inhibitor and spiral flow. Int. J. Smart Grid Clean Energy 2016, 5, 174–181. [Google Scholar] [CrossRef]
  25. Sakthipriya, N.; Doble, M.; Sangwai, J.S. Fast degradation and viscosity reduction of waxy crude oil and model waxy crude oil using Bacillus subtilis. J. Pet. Sci. Eng. 2015, 134, 158–166. [Google Scholar] [CrossRef]
  26. Borthakur, A.; Chanda, D.; Choudhury, S.R.D.; Rao, K.V.; Subrahmanyam, B. Alkyl fumarate-vinyl acetate copolymer as flow improver for high waxy Indian crude oils. Energy Fuels 1996, 10, 844–848. [Google Scholar] [CrossRef]
  27. El-Ghazawy, R.A.; Atta, A.M.; Kabel, K.I. Modified maleic anhydride-co-octadecene copolymers as flow improver for waxy Egyptian crude oil. J. Pet. Sci. Eng. 2014, 122, 411–419. [Google Scholar] [CrossRef]
  28. Chi, Y.; Daraboina, N.; Sarica, C. Effect of the Flow Field on the Wax Deposition and Performance of Wax Inhibitors: Cold Finger and Flow Loop Testing. Energy Fuels 2017, 31, 4915–4924. [Google Scholar] [CrossRef]
  29. Joonaki, E.; Hassanpouryouzband, A.; Burgass, R.; Hase, A.; Tohidi, B. Effects of Waxes and the Related Chemicals on Asphaltene Aggregation and Deposition Phenomena: Experimental and Modeling Studies. ACS Omega 2020, 5, 7124–7134. [Google Scholar] [CrossRef] [PubMed]
  30. Pritchard, P.H.; Mueller, J.G.; Rogers, J.C.; Kremer, F.V.; Glaser, J.A. Oil spill bioremediation: Experiences, lessons and results from the Exxon Valdez oil spill in Alaska. Biodegradation 1992, 3, 315–335. [Google Scholar] [CrossRef]
  31. Al-Sayegh, A.; Al-Wahaibi, Y.; Joshi, S.; Al-Bahry, S.; Elshafie, A.; Al-Bemani, A. Bioremediation of Heavy Crude Oil Contamination. Open Biotechnol. J. 2016, 10, 301–311. [Google Scholar] [CrossRef]
  32. Environmental Influences on the Microbial Degradation of Exxon Valdez Oil on the Shorelines of Prince William Sound, Alaska. 1997. Available online: https://pubs.acs.org/sharingguidelines (accessed on 23 July 2024).
  33. Kappell, A.D.; Wei, Y.; Newton, R.J.; Van Nostrand, J.D.; Zhou, J.; McLellan, S.L.; Hristova, K.R. The polycyclic aromatic hydrocarbon degradation potential of Gulf of Mexico native coastal microbial communities after the Deepwater Horizon oil spill. Front. Microbiol. 2014, 5, 205. [Google Scholar] [CrossRef]
  34. Gargouri, B.; Karray, F.; Mhiri, N.; Aloui, F.; Sayadi, S. Application of a continuously stirred tank bioreactor (CSTR) for bioremediation of hydrocarbon-rich industrial wastewater effluents. J. Hazard. Mater. 2011, 189, 427–434. [Google Scholar] [CrossRef]
  35. Kuyukina, M.S.; Krivoruchko, A.V.; Ivshina, I.B. Advanced bioreactor treatments of hydrocarbon-containing wastewater. Appl. Sci. 2020, 10, 831. [Google Scholar] [CrossRef]
  36. Purwasena, I.A.; Sugai, Y.; Sasaki, K. Estimation of the potential of an oil-viscosity-reducing bacterium Petrotoga sp. isolated from an oil field for MEOR. Proc. SPE Annu. Tech. Conf. Exhib. 2010, 5, 3640–3650. [Google Scholar] [CrossRef]
  37. Kadri, T.; Magdouli, S.; Rouissi, T.; Brar, S.K. Ex-situ biodegradation of petroleum hydrocarbons using Alcanivorax borkumensis enzymes. Biochem. Eng. J. 2018, 132, 279–287. [Google Scholar] [CrossRef]
  38. Tan, G.B.; Liu, S.H.; Wang, D.G.; Zhang, S.W. Tribological Behaviours of Wax-in-Oil Gel Deposition in Orthogonal Cleaning Process. Tribol. Lett. 2015, 57, 16. [Google Scholar] [CrossRef]
  39. Tan, G.B.; Liu, S.H.; Wang, D.G.; Zhang, S.W. Measurement and analysis of wax-oil gel scraping process at contact area under pure sliding conditions. Measurement 2016, 80, 29–43. [Google Scholar] [CrossRef]
  40. Bell, E.; Lu, Y.; Daraboina, N.; Sarica, C. Experimental Investigation of active heating in removal of wax deposits. J. Pet. Sci. Eng. 2021, 200, 108346. [Google Scholar] [CrossRef]
  41. Zhu, H.; Li, C.; Yang, F.; Liu, H.; Liu, D.; Sun, G.; Yao, B.; Liu, G.; Zhao, Y. Effect of Thermal Treatment Temperature on the Flowability and Wax Deposition Characteristics of Changqing Waxy Crude Oil. Energy Fuels 2018, 32, 10605–10615. [Google Scholar] [CrossRef]
  42. Wang, Z.; Zeng, J.; Song, H.; Li, F. Research on ultrasonic excitation for the removal of drilling fluid plug, paraffin deposition plug, polymer plug and inorganic scale plug for near-well ultrasonic processing technology. Ultrason. Sonochem. 2017, 36, 162–167. [Google Scholar] [CrossRef]
  43. Torrelo, G.; Hanefeld, U.; Hollmann, F. Biocatalysis. Catal. Lett. 2015, 145, 309–345. [Google Scholar] [CrossRef]
  44. Abdelraheem, E.M.M.; Busch, H.; Hanefeld, U.; Tonin, F. Biocatalysis explained: From pharmaceutical to bulk chemical production. React. Chem. Eng. 2019, 4, 1878–1894. [Google Scholar] [CrossRef]
  45. Asif, M.B.; Hai, F.I.; Kang, J.; van de Merwe, J.P.; Leusch, F.D.; Price, W.E.; Nghiem, L.D. Biocatalytic degradation of pharmaceuticals, personal care products, industrial chemicals, steroid hormones and pesticides in a membrane distillation-enzymatic bioreactor. Bioresour. Technol. 2018, 247, 528–536. [Google Scholar] [CrossRef]
  46. Barth, M.; Honak, A.; Oeser, T.; Wei, R.; Belisário-Ferrari, M.R.; Then, J.; Schmidt, J.; Zimmermann, W. A dual enzyme system composed of a polyester hydrolase and a carboxylesterase enhances the biocatalytic degradation of polyethylene terephthalate films. Biotechnol. J. 2016, 11, 1082–1087. [Google Scholar] [CrossRef] [PubMed]
  47. de Cazes, M.; Belleville, M.P.; Mougel, M.; Kellner, H.; Sanchez-Marcano, J. Characterization of laccase-grafted ceramic membranes for pharmaceuticals degradation. J. Memb. Sci. 2015, 476, 384–393. [Google Scholar] [CrossRef]
  48. Gill, I.; Ballesteros, A. Degradation of organophosphorous nerve agents by enzyme-polymer nanocomposites: Efficient biocatalytic materials for personal protection and large-scale detoxification. Biotechnol. Bioeng. 2000, 70, 400–410. [Google Scholar] [CrossRef] [PubMed]
  49. Ji, C.; Hou, J.; Chen, V. Cross-linked carbon nanotubes-based biocatalytic membranes for micro-pollutants degradation: Performance, stability, and regeneration. J. Memb. Sci. 2016, 520, 869–880. [Google Scholar] [CrossRef]
  50. Ji, C.; Hou, J.; Wang, K.; Zhang, Y.; Chen, V. Biocatalytic degradation of carbamazepine with immobilized laccase-mediator membrane hybrid reactor. J. Memb. Sci. 2016, 502, 11–20. [Google Scholar] [CrossRef]
  51. Orozco, J.; Vilela, D.; Valdés-Ramírez, G.; Fedorak, Y.; Escarpa, A.; Vazquez-Duhalt, R.; Wang, J. Efficient biocatalytic degradation of pollutants by enzyme-releasing self-propelled motors. Chem. A Eur. J. 2014, 20, 2866–2871. [Google Scholar] [CrossRef]
  52. Wei, R.; Breite, D.; Song, C.; Gräsing, D.; Ploss, T.; Hille, P.; Schwerdtfeger, R.; Matysik, J.; Schulze, A.; Zimmermann, W. Biocatalytic Degradation Efficiency of Postconsumer Polyethylene Terephthalate Packaging Determined by Their Polymer Microstructures. Adv. Sci. 2019, 6, 1900491. [Google Scholar] [CrossRef]
  53. Salehi, S.; Abdollahi, K.; Panahi, R.; Rahmanian, N.; Shakeri, M.; Mokhtarani, B. Applications of biocatalysts for sustainable oxidation of phenolic pollutants: A review. Sustainability 2021, 13, 8620. [Google Scholar] [CrossRef]
  54. Betancur-Ramírez, K.J.; Meneses-Jácome, A.; Ruiz-Colorado, A.A.; Gallego-Suárez, D. Life cycle assessment of an alternative enzymatic-biological treatment for effluents from industrial processing of potatoes. J. Clean. Prod. 2021, 324, 129151. [Google Scholar] [CrossRef]
  55. Gaur, N.; Narasimhulu, K.; PydiSetty, Y. Recent advances in the bio-remediation of persistent organic pollutants and its effect on environment. J. Clean. Prod. 2018, 198, 1602–1631. [Google Scholar] [CrossRef]
  56. Suganthi, S.H.; Murshid, S.; Sriram, S.; Ramani, K. Enhanced biodegradation of hydrocarbons in petroleum tank bottom oil sludge and characterization of biocatalysts and biosurfactants. J. Environ. Manag. 2018, 220, 87–95. [Google Scholar] [CrossRef] [PubMed]
  57. Chettri, B.; Singh, A.K. Kinetics of hydrocarbon degradation by a newly isolated heavy metal tolerant bacterium Novosphingobium panipatense P5:ABC. Bioresour. Technol. 2019, 294, 122190. [Google Scholar] [CrossRef] [PubMed]
  58. Zapata, K.; Carrasco-Marín, F.; Arias, J.P.; Castelo-Quibén, J.; Franco, C.A.; Rojano, B.; Cortés, F.B. Novel biomaterial design based on Pseudomonas stutzeri–carbon xerogel microspheres for hydrocarbon removal from oil-in-saltwater emulsions: A new proposed treatment of produced water in oilfields. J. Water Process Eng. 2020, 35, 101222. [Google Scholar] [CrossRef]
  59. Chen, W.; Kong, Y.; Li, J.; Sun, Y.; Min, J.; Hu, X. Enhanced biodegradation of crude oil by constructed bacterial consortium comprising salt-tolerant petroleum degraders and biosurfactant producers. Int. Biodeterior. Biodegrad. 2020, 154, 105047. [Google Scholar] [CrossRef]
  60. Swathi, K.V.; Muneeswari, R.; Ramani, K.; Sekaran, G. Biodegradation of petroleum refining industry oil sludge by microbial-assisted biocarrier matrix: Process optimization using response surface methodology. Biodegradation 2020, 31, 385–405. [Google Scholar] [CrossRef]
  61. Dzionek, A.; Wojcieszyńska, D.; Guzik, U. Natural carriers in bioremediation: A review. Electron. J. Biotechnol. 2016, 23, 28–36. [Google Scholar] [CrossRef]
  62. Sánchez-Vázquez, V.; Shirai, K.; González, I.; Gutiérrez-Rojas, M. Fungal biocatalyst activated by an electric field: Improved mass transfer and non-specificity for hydrocarbon degradation in an airlift bioreactor. J. Hazard. Mater. 2017, 337, 62–71. [Google Scholar] [CrossRef]
  63. Muneeswari, R.; Iyappan, S.; Swathi, K.V.; Vinu, R.; Ramani, K.; Sekaran, G. Biocatalytic lipoprotein bioamphiphile induced treatment of recalcitrant hydrocarbons in petroleum refinery oil sludge through transposon technology. J. Hazard. Mater. 2022, 431, 128520. [Google Scholar] [CrossRef]
  64. Muneeswari, R.; Swathi, K.V.; Sekaran, G.; Ramani, K. Microbial-induced biosurfactant-mediated biocatalytic approach for the bioremediation of simulated marine oil spill. Int. J. Environ. Sci. Technol. 2022, 19, 341–354. [Google Scholar] [CrossRef]
  65. Lázaro-Mass, S.; Gómez-Cornelio, S.; Castillo-Vidal, M.; Alvarez-Villagomez, C.S.; Quintana, P.; De la Rosa-García, S. Biodegradation of hydrocarbons from contaminated soils by microbial consortia: A laboratory microcosm study. Electron. J. Biotechnol. 2023, 61, 24–32. [Google Scholar] [CrossRef]
  66. Cai, Y.; Wang, R.; Rao, P.; Wu, B.; Yan, L.; Hu, L.; Park, S.; Ryu, M.; Zhou, X. Bioremediation of petroleum hydrocarbons using Acinetobacter sp. SCYY-5 isolated from contaminated oil sludge: Strategy and effectiveness study. Int. J. Environ. Res. Public Health 2021, 18, 819. [Google Scholar] [CrossRef] [PubMed]
  67. Khazaal, R.M.; Ismail, Z.Z. Bioremediation and detoxification of real refinery oily sludge using mixed bacterial cells. Pet. Res. 2021, 6, 303–308. [Google Scholar] [CrossRef]
  68. El-Sheshtawy, H.S.; Aman, D.; Nassar, H.N. A Novel Bioremediation Technique for Petroleum Hydrocarbons by Bacterial Consortium Immobilized on Goethite-chitosan Nanocomposite. Soil. Sediment. Contam. 2022, 31, 176–199. [Google Scholar] [CrossRef]
  69. Wang, H.; Lu, L.; Chen, H.; McKenna, A.M.; Liu, J.; Jin, S.; Zuo, Y.; Rosario-Ortiz, F.L.; Ren, Z.J. Molecular Transformation of Crude Oil Contaminated Soil after Bioelectrochemical Degradation Revealed by FT-ICR Mass Spectrometry. Environ. Sci. Technol. 2020, 54, 2500–2509. [Google Scholar] [CrossRef] [PubMed]
  70. Sampaio, C.J.S.; de Souza, J.R.B.; de Carvalho, G.C.; Quintella, C.M.; de Abreu, M.R. Analysis of petroleum biodegradation by a bacterial consortium isolated from worms of the polychaeta class (Annelida): Implications for NPK fertilizer supplementation. J. Environ. Manag. 2019, 246, 617–624. [Google Scholar] [CrossRef]
  71. Wang, C.; Liu, X.; Guo, J.; Lv, Y.; Li, Y. Biodegradation of marine oil spill residues using aboriginal bacterial consortium based on Penglai 19-3 oil spill accident, China. Ecotoxicol. Envirn. Saf. 2018, 159, 20–27. [Google Scholar] [CrossRef]
  72. Kachieng’a, L.; Momba, M.N.B. Kinetics of petroleum oil biodegradation by a consortium of three protozoan isolates (Aspidisca sp., Trachelophyllum sp. and Peranema sp.). Biotechnol. Rep. 2017, 15, 125–131. [Google Scholar] [CrossRef]
  73. Varjani, S.J.; Upasani, V.N. Biodegradation of petroleum hydrocarbons by oleophilic strain of Pseudomonas aeruginosa NCIM 5514. Bioresour. Technol. 2016, 222, 195–201. [Google Scholar] [CrossRef]
  74. Shen, T.; Pi, Y.; Bao, M.; Xu, N.; Li, Y.; Lu, J. Biodegradation of different petroleum hydrocarbons by free and immobilized microbial consortia. Environ. Sci. Process. Impacts 2015, 17, 2022–2033. [Google Scholar] [CrossRef]
  75. Sánchez-Vázquez, V.; González, I.; Gutiérrez-Rojas, M. Electric field as pretreatment to enhance the activity of a whole-cell biocatalyst for hydrocarbon degradation in contaminated water. Chem. Eng. J. 2015, 260, 37–42. [Google Scholar] [CrossRef]
  76. Dellagnezze, B.M.; de Sousa, G.V.; Martins, L.L.; Domingos, D.F.; Limache, E.E.G.; de Vasconcellos, S.P.; da Cruz, G.F.; de Oliveira, V.M. Bioremediation potential of microorganisms derived from petroleum reservoirs. Mar. Pollut. Bull. 2014, 89, 191–200. [Google Scholar] [CrossRef] [PubMed]
  77. Singh, A. Assessment of Bioremediation of Oil and Phenol Contents in Refinery Waste Water via Bacterial Consortium. J. Pet. Environ. Biotechnol. 2013, 4, 1–4. [Google Scholar] [CrossRef]
  78. Bai, C.; Zhang, J. Thermal, macroscopic, and microscopic characteristics of wax deposits in field pipelines. Energy Fuels 2013, 27, 752–759. [Google Scholar] [CrossRef]
  79. Van Beilen, J.B.; Funhoff, E.G. Alkane hydroxylases involved in microbial alkane degradation. Appl. Microbiol. Biotechnol. 2007, 74, 13–21. [Google Scholar] [CrossRef]
  80. Kotlar, H.K.; Wentzel, A.; Throne-Hoist, M.; Zotchev, S.; Ellingensen, T. Wax control by biocatalytic degradation in high-paraffmic crude oils. In Proceedings of the SPE International Symposium on Oilfield Chemistry, Houston, TX, USA, 28 February–2 March 2007; pp. 533–539. [Google Scholar] [CrossRef]
  81. Schneiker, S.; Santos, V.A.M.d.; Bartels, D.; Bekel, T.; Brecht, M.; Buhrmester, J.; Chernikova, T.N.; Denaro, R.; Ferrer, M.; Gertler, C.; et al. Genome sequence of the ubiquitous hydrocarbon-degrading marine bacterium Alcanivorax borkumensis. Nat. Biotechnol. 2006, 24, 997–1004. [Google Scholar] [CrossRef]
  82. Nekouei, F.; Nekouei, S. Enhanced enzymatic and: Ex situ biodegradation of petroleum hydrocarbons in solutions using Alcanivorax borkumensis enzymes in the presence of nitrogen and phosphorus co-doped reduced graphene oxide as a bacterial growth enhancer. J. Mater. Chem. A Mater. 2017, 5, 24462–24471. [Google Scholar] [CrossRef]
  83. Throne-Holst, M.; Wentzel, A.; Ellingsen, T.E.; Kotlar, H.K.; Zotchev, S.B. Identification of novel genes involved in long-chain n-alkane degradation by Acinetobacter sp. strain DSM 17874. Appl. Environ. Microbiol. 2007, 73, 3327–3332. [Google Scholar] [CrossRef]
  84. Xia, W.; Dong, H.; Zheng, C.; Cui, Q.; He, P.; Tang, Y. Hydrocarbon degradation by a newly isolated thermophilic Anoxybacillus sp. with bioemulsifier production and new alkB genes. RSC Adv. 2015, 5, 102367–102377. [Google Scholar] [CrossRef]
  85. Liu, C.; Wang, W.; Wu, Y.; Zhou, Z.; Lai, Q.; Shao, Z. Multiple alkane hydroxylase systems in a marine alkane degrader, Alcanivorax dieselolei B-5. Environ. Microbiol. 2011, 13, 1168–1178. [Google Scholar] [CrossRef]
  86. Feng, L.; Wang, W.; Cheng, J.; Ren, Y.; Zhao, G.; Gao, C.; Tang, Y.; Liu, X.; Han, W.; Peng, X.; et al. Genome and proteome of long-chain alkane degrading Geobacillus thermodenitrificans NG80-2 isolated from a deep-subsurface oil reservoir. Proc. Natl. Acad. Sci. USA 2007, 104, 5602–5607. [Google Scholar] [CrossRef]
  87. Larik, I.A.; Qazi, M.A.; Kanhar, A.R.; Mangi, S.; Ahmed, S.; Jamali, M.R.; Kanhar, N.A. Biodegradation of Petrochemical Hydrocarbons Using an Efficient Bacterial Consortium: A2457. Arab. J. Sci. Eng. 2016, 41, 2077–2086. [Google Scholar] [CrossRef]
  88. Yanto, D.H.Y.; Tachibana, S. Biodegradation of petroleum hydrocarbons by a newly isolated Pestalotiopsis sp. NG007. Int. Biodeterior. Biodegrad. 2013, 85, 438–450. [Google Scholar] [CrossRef]
  89. Elumalai, P.; Parthipan, P.; Karthikeyan, O.P.; Rajasekar, A. Enzyme-mediated biodegradation of long-chain n-alkanes (C32 and C40) by thermophilic bacteria. 3 Biotech 2017, 7, 1–10. [Google Scholar] [CrossRef] [PubMed]
  90. Kadri, T.; Rouissi, T.; Magdouli, S.; Brar, S.K.; Hegde, K.; Khiari, Z.; Daghrir, R.; Lauzon, J.-M. Production and characterization of novel hydrocarbon degrading enzymes from Alcanivorax borkumensis. Int. J. Biol. Macromol. 2018, 112, 230–240. [Google Scholar] [CrossRef]
  91. Wu, B.; Xiu, J.; Yu, L.; Huang, L.; Yi, L.; Ma, Y. Degradation of crude oil in a co-culture system of Bacillus subtilis and Pseudomonas aeruginosa. Front. Microbiol. 2023, 14, 1132831. [Google Scholar] [CrossRef]
  92. Chen, L.; Xu, K.; Zhang, Y.; Hasi, Q.; Luo, X.; Xu, J.; Li, A. Selective adsorption and efficient degradation of petroleum hydrocarbons by a hydrophobic/lipophilic biomass porous foam loaded with microbials. ACS Appl. Mater. Interfaces 2021, 13, 53586–53598. [Google Scholar] [CrossRef]
  93. Wang, D.; Lin, J.; Lin, J.; Wang, W.; Li, S. Biodegradation of petroleum hydrocarbons by bacillus subtilis BL-27, a strain with weak hydrophobicity. Molecules 2019, 24, 3021. [Google Scholar] [CrossRef]
  94. Abena, M.T.B.; Sodbaatar, N.; Li, T.; Damdinsuren, N.; Choidash, B.; Zhong, W. Crude Oil Biodegradation by Newly Isolated Bacterial Strains and Their Consortium Under Soil Microcosm Experiment. Appl. Biochem. Biotechnol. 2019, 189, 1223–1244. [Google Scholar] [CrossRef]
  95. Xia, M.; Fu, D.; Chakraborty, R.; Singh, R.P.; Terry, N. Enhanced crude oil depletion by constructed bacterial consortium comprising bioemulsifier producer and petroleum hydrocarbon degraders. Bioresour. Technol. 2019, 282, 456–463. [Google Scholar] [CrossRef]
  96. French, K.E.; Terry, N. A high-throughput fluorescence-based assay for rapid identification of petroleum-degrading bacteria. Front. Microbiol. 2019, 10, 1318. [Google Scholar] [CrossRef]
  97. Shi, K.; Xue, J.; Xiao, X.; Qiao, Y.; Wu, Y.; Gao, Y. Mechanism of Degrading Petroleum Hydrocarbons by Compound Marine Petroleum-Degrading Bacteria: Surface Adsorption, Cell Uptake, and Biodegradation. Energy Fuels 2019, 33, 11373–11379. [Google Scholar] [CrossRef]
  98. Obafemi, Y.D.; Taiwo, O.S.; Omodara, O.J.; Dahunsi, O.S.; Oranusi, S. Biodegradation of crude petroleum by bacterial consortia from oil-contaminated soils in Ota, Ogun State, South-Western, Nigeria. Env. Technol. Innov. 2018, 12, 230–242. [Google Scholar] [CrossRef]
  99. Shetaia, Y.M.H.; El khalik, W.A.A.; Mohamed, T.M.; Farahat, L.A.; ElMekawy, A. Potential biodegradation of crude petroleum oil by newly isolated halotolerant microbial strains from polluted Red Sea area. Mar. Pollut. Bull. 2016, 111, 435–442. [Google Scholar] [CrossRef] [PubMed]
  100. El-Sheekh, M.M.; Hamouda, R.A.; Nizam, A.A. Biodegradation of crude oil by Scenedesmus obliquus and Chlorella vulgaris growing under heterotrophic conditions. Int. Biodeterior. Biodegrad. 2013, 82, 67–72. [Google Scholar] [CrossRef]
  101. Zhang, J.; Xue, Q.; Gao, H.; Lai, H.; Wang, P. Bacterial degradation of crude oil using solid formulations of bacillus strains isolated from oil-contaminated soil towards microbial enhanced oil recovery application. RSC Adv. 2016, 6, 5566–5574. [Google Scholar] [CrossRef]
  102. Ayala, M.; Vazquez-Duhalt, R. Enzymatic catalysis on petroleum products. Stud. Surf. Sci. Catal. 2004, 151, 67–111. [Google Scholar] [CrossRef]
  103. Suma, Y.; Lim, H.; Kwean, O.S.; Cho, S.; Yang, J.; Kim, Y.; Kang, C.S.; Kim, H.S. Enzymatic degradation of aromatic hydrocarbon intermediates using a recombinant dioxygenase immobilized onto surfactant-activated carbon nanotube. Bioresour. Technol. 2016, 210, 117–122. [Google Scholar] [CrossRef]
  104. Latip, W.; Rosli, N.E.; Ali, M.S.M.; Kamarudin, N.H.A.; Rahman, R.N.Z.R.A. Stability Enhancement of Aldehyde Dehydrogenase from Anoxybacillus geothermalis Strain D9 Immobilized onto Seplite LX120. Catalysts 2023, 13, 368. [Google Scholar] [CrossRef]
  105. Rosli, N.E.; Ali, M.S.M.; Kamarudin, N.H.A.; Masomian, M.; Latip, W.; Saadon, S.; Rahman, R.N.Z.R.A. Structure Prediction and Characterization of Thermostable Aldehyde Dehydrogenase from Newly Isolated Anoxybacillus geothermalis Strain D9. Microorganisms 2022, 10, 1444. [Google Scholar] [CrossRef]
  106. Leis, B.; Angelov, A.; Mientus, M.; Li, H.; Pham, V.T.T.; Lauinger, B.; Bongen, P.; Pietruszka, J.; Gonã§Alves, L.G.; Santos, H.; et al. Identification of novel esterase-active enzymes from hot environments by use of the host bacterium Thermus thermophilus. Front. Microbiol. 2015, 6, 275. [Google Scholar] [CrossRef]
  107. Rodriguez, G.M.; Atsumi, S. Isobutyraldehyde production from Escherichia coli by removing aldehyde reductase activity. Microb. Cell Fact. 2012, 11, 90. [Google Scholar] [CrossRef] [PubMed]
  108. Machielsen, R.; Uria, A.R.; Kengen, S.W.M.; Van Der Oost, J. Production and characterization of a thermostable alcohol dehydrogenase that belongs to the aldo-keto reductase superfamily. Appl. Environ. Microbiol. 2006, 72, 233–238. [Google Scholar] [CrossRef] [PubMed]
  109. Mishra, S.; Singh, S.N. Microbial degradation of n-hexadecane in mineral salt medium as mediated by degradative enzymes. Bioresour. Technol. 2012, 111, 148–154. [Google Scholar] [CrossRef] [PubMed]
  110. Ji, Y.; Mao, G.; Wang, Y.; Bartlam, M. Structural insights into diversity and n-alkane biodegradation mechanisms of alkane hydroxylases. Front. Microbiol. 2013, 4, 58. [Google Scholar] [CrossRef] [PubMed]
  111. Ilie, A.; Agudo, R.; Roiban, G.D.; Reetz, M.T. P450-catalyzed regio- and stereoselective oxidative hydroxylation of disubstituted cyclohexanes: Creation of three centers of chirality in a single CH-activation event This paper is dedicated to the memory of Harry H. Wasserman. Tetrahedron 2015, 71, 470–475. [Google Scholar] [CrossRef]
  112. Peixoto, F.B.S.; Da Cunha Peixoto, J.C.; Motta, D.C.L.; Peixoto, A.T.M.; Pereira, J.O.; Filho, S.A. Assessment of petroleum biodegradation for Bacillus toyonensis by the using redox indicator 2,6 dichlorophenol indophenol. Acta Sci. Biol. Sci. 2018, 40. [Google Scholar] [CrossRef]
  113. van Beilen, J.B.; Li, Z.; Duetz, W.A.; Smits, T.H.M.; Witholt, B. Diversité des systèmes alcane hydroxylase dans l’environnement. Oil Gas. Sci. Technol. 2003, 58, 427–440. [Google Scholar] [CrossRef]
  114. Rojo, F. Degradation of alkanes by bacteria: Minireview. Environ. Microbiol. 2009, 11, 2477–2490. [Google Scholar] [CrossRef]
  115. Varjani, S.J. Microbial degradation of petroleum hydrocarbons. Bioresour. Technol. 2017, 223, 277–286. [Google Scholar] [CrossRef]
  116. Wentzel, A.; Ellingsen, T.E.; Kotlar, H.K.; Zotchev, S.B.; Throne-Holst, M. Bacterial metabolism of long-chain n-alkanes. Appl. Microbiol. Biotechnol. 2007, 76, 1209–1221. [Google Scholar] [CrossRef]
  117. Smits, T.H.M.; Balada, S.B.; Witholt, B.; Van Beilen, J.B. Functional analysis of alkane hydroxylases from gram-negative and gram-positive bacteria. J. Bacteriol. 2002, 184, 1733–1742. [Google Scholar] [CrossRef] [PubMed]
  118. Tani, A.; Ishige, T.; Sakai, Y.; Kato, N. Gene structures and regulation of the alkane hydroxylase complex in Acinetobacter sp. strain M-1. J. Bacteriol. 2001, 183, 1819–1823. [Google Scholar] [CrossRef]
  119. Park, C.; Shin, B.; Jung, J.; Lee, Y.; Park, W. Metabolic and stress responses of Acinetobacter oleivorans DR1 during long-chain alkane degradation. Microb. Biotechnol. 2017, 10, 1809–1823. [Google Scholar] [CrossRef] [PubMed]
  120. Zampolli, J.; Collina, E.; Lasagni, M.; Di Gennaro, P. Biodegradation of variable-chain-length n-alkanes in Rhodococcus opacus R7 and the involvement of an alkane hydroxylase system in the metabolism. AMB Express 2014, 4, 73. [Google Scholar] [CrossRef] [PubMed]
  121. Gregson, B.H.; Metodieva, G.; Metodiev, M.V.; Golyshin, P.N.; McKew, B.A. Differential Protein Expression During Growth on Medium Versus Long-Chain Alkanes in the Obligate Marine Hydrocarbon-Degrading Bacterium Thalassolituus oleivorans MIL-1. Front. Microbiol. 2018, 9, 3130. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, L.; Tang, Y.; Wang, S.; Liu, R.-L.; Liu, M.-Z.; Zhang, Y.; Liang, F.-L.; Feng, L. Isolation and characterization of a novel thermophilic Bacillus strain degrading long-chain n-alkanes. Extremophiles 2006, 10, 347–356. [Google Scholar] [CrossRef] [PubMed]
  123. Eibes, G.; McCann, C.; Pedezert, A.; Moreira, M.T.; Feijoo, G.; Lema, J.M. Study of mass transfer and biocatalyst stability for the enzymatic degradation of anthracene in a two-phase partitioning bioreactor. Biochem. Eng. J. 2010, 51, 79–85. [Google Scholar] [CrossRef]
  124. Arora, P.K. Microbial Metabolism of Xenobiotic Compounds; 2019, Volume 10. Available online: http://link.springer.com/10.1007/978-981-13-7462-3 (accessed on 3 August 2024).
  125. Eibes, G.; Arca-Ramos, A.; Feijoo, G.; Lema, J.M.; Moreira, M.T. Enzymatic technologies for remediation of hydrophobic organic pollutants in soil. Appl. Microbiol. Biotechnol. 2015, 99, 8815–8829. [Google Scholar] [CrossRef] [PubMed]
  126. Morikawa, M. Dioxygen activation responsible for oxidation of aliphatic and aromatic hydrocarbon compounds: Current state and variants. Appl. Microbiol. Biotechnol. 2010, 87, 1595–1603. [Google Scholar] [CrossRef]
  127. Saadon, S.; Ali, M.S.M.; Kamarudin, N.H.A.; Latip, W.; Ishak, S.N.H.; Basri, R.S.; Johan, U.U.M.; Shukri, N.S.A.; Rosli, N.E.; Rahman, R.N.Z.R.A. Benefitting multi-enzyme system for the purpose of improving the flow properties of waxy oil. Geoenergy Sci. Eng. 2023, 230, 212221. [Google Scholar] [CrossRef]
  128. Kotlar, H.K.; Brakstad, O.G.; Markussen, S.; Winnberg, A. Use of Petroleum Biotechnology throughout the value chain of an oil company: An integrated approach. Stud. Surf. Sci. Catal. 2004, 151, 1–27. [Google Scholar] [CrossRef]
  129. Wang, M.; Zhang, X.; Huang, Y.; Zhou, F.; Zhao, Q.; Zhang, W.; Chen, G.; Gao, Y. Isolation of Petroleum Degraders and Petroleum-Degradation Characteristics of Crude Enzymes from Providencia rettgeri L1. Pol. J. Environ. Stud. 2022, 31, 3859–3866. [Google Scholar] [CrossRef] [PubMed]
  130. Kadri, T.; Miri, S.; Robert, T.; Brar, S.K.; Rouissi, T.; LaxmanPachapur, V.; Lauzon, J.-M. Pilot-scale production and in-situ application of petroleum-degrading enzyme cocktail from Alcanivorax borkumensis. Chemosphere 2022, 295, 133840. [Google Scholar] [CrossRef] [PubMed]
  131. Bulai, I.S.; Adamu, H.; Umar, Y.A.; Sabo, A. Biocatalytic remediation of used motor oil-contaminated soil by fruit garbage enzymes. J. Environ. Chem. Eng. 2021, 9, 105465. [Google Scholar] [CrossRef]
  132. Ji, L.; Fu, X.; Wang, M.; Xu, C.; Chen, G.; Song, F.; Guo, S.; Zhang, Q. Enzyme cocktail containing NADH regeneration system for efficient bioremediation of oil sludge contamination. Chemosphere 2019, 233, 132–139. [Google Scholar] [CrossRef]
  133. Kucharzyk, K.H.; Benotti, M.; Darlington, R.; Lalgudi, R. Enhanced biodegradation of sediment-bound heavily weathered crude oil with ligninolytic enzymes encapsulated in calcium-alginate beads. J. Hazard. Mater. 2018, 357, 498–505. [Google Scholar] [CrossRef]
  134. Al-Hawash, A.B.; Zhang, X.; Ma, F. Removal and biodegradation of different petroleum hydrocarbons using the filamentous fungus Aspergillus sp. RFC-1. Microbiologyopen 2019, 8, e00619. [Google Scholar] [CrossRef]
  135. Zhang, J.H.; Xue, Q.H.; Gao, H.; Ma, X.; Wang, P. Degradation of crude oil by fungal enzyme preparations from Aspergillus spp. for potential use in enhanced oil recovery. J. Chem. Technol. Biotechnol. 2016, 91, 865–875. [Google Scholar] [CrossRef]
  136. Xu, N.; Bao, M.; Sun, P.; Li, Y. Study on bioadsorption and biodegradation of petroleum hydrocarbons by a microbial consortium. Bioresour. Technol. 2013, 149, 22–30. [Google Scholar] [CrossRef]
  137. Iyer, P.V.; Ananthanarayan, L. Enzyme stability and stabilization-Aqueous and non-aqueous environment. Process Biochem. 2008, 43, 1019–1032. [Google Scholar] [CrossRef]
  138. Sutherland, T.D.; Horne, I.; Weir, K.; Coppin, C.; Williams, M.R.; Selleck, M.; Russell, R.; Oakeshott, J. Enzymatic bioremediation: From enzyme discovery to applications. Clin. Exp. Pharmacol. Physiol. 2004, 31, 817–821. [Google Scholar] [CrossRef]
  139. Khadim, M.A.; Sarbar, M.A. Role of asphaltene and resin in oil field emulsions. J. Pet. Sci. Eng. 1999, 23, 213–221. [Google Scholar] [CrossRef]
  140. Wei, R.; Zimmermann, W. Microbial enzymes for the recycling of recalcitrant petroleum-based plastics: How far are we? Microb. Biotechnol. 2017, 10, 1308–1322. [Google Scholar] [CrossRef]
  141. Scott, C.; Begley, C.; Taylor, M.J.; Pandey, G.; Momiroski, V.; French, N.; Brearley, C.; Kotsonis, S.E.; Selleck, M.J.; Carino, F.A.; et al. Free-enzyme bioremediation of pesticides: A case study for the enzymatic remediation of organophosphorous insecticide residues. ACS Symp. Ser. 2011, 1075, 155–174. [Google Scholar] [CrossRef]
  142. Gaid, K. A Large Review of the Pre Treatment. In Expanding Issues in Desalination; IntechOpen Limited: London, UK, 2011. [Google Scholar] [CrossRef]
  143. Al-Ghouti, M.A.; Al-Kaabi, M.A.; Ashfaq, M.Y.; Da’na, D.A. Produced water characteristics, treatment and reuse: A review. J. Water Process Eng. 2019, 28, 222–239. [Google Scholar] [CrossRef]
  144. Bonifay, V.; Wawrik, B.; Sunner, J.; Snodgrass, E.C.; Aydin, E.; Duncan, K.E.; Callaghan, A.V.; Oldham, A.; Liengen, T.; Beech, I. Metabolomic and metagenomic analysis of two crude oil production pipelines experiencing differential rates of corrosion. Front. Microbiol. 2017, 8, 99. [Google Scholar] [CrossRef]
  145. Mu, B.-Z.; Nazina, T.N. Recent Advances in Petroleum Microbiology. Microorganisms 2022, 10, 1706. [Google Scholar] [CrossRef]
  146. Karan, R.; Capes, M.D.; DasSarma, S. Function and biotechnology of extremophilic enzymes in low water activity. Aquat. Biosyst. 2012, 8, 1–15. [Google Scholar] [CrossRef]
  147. Yang, B.; Lin, Z.; Liu, Y.; Hu, D.; Zhou, X.; Pan, Q.; Pan, X. Spatial dynamics of biochemical fractions degradation, functional enzymatic activity and bacterial community within co-bioevaporation pile. J. Clean. Prod. 2021, 287, 125552. [Google Scholar] [CrossRef]
  148. Nzila, A. Current status of the degradation of aliphatic and aromatic petroleum hydrocarbons by thermophilic microbes and future perspectives. Int. J. Environ. Res. Public Health 2018, 15, 2782. [Google Scholar] [CrossRef]
  149. Gaur, V.K.; Gupta, S.; Pandey, A. Evolution in mitigation approaches for petroleum oil-polluted environment: Recent advances and future directions. Environ. Sci. Pollut. Res. 2021, 29, 61821–61837. [Google Scholar] [CrossRef]
  150. Lai, C.; Tang, Z.; Liu, Z.; Luo, P.; Zhang, W.; Zhang, T.; Zhang, W.; Dong, Z.; Liu, X.; Yang, X.; et al. Probing the functional hotspots inside protein hydrophobic pockets by in situ photochemical trifluoromethylation and mass spectrometry. Chem. Sci. 2024, 15, 2545–2557. [Google Scholar] [CrossRef]
  151. Pauli, O.; Ecker, A.; Cruz-Izquierdo, A.; Basso, A.; Serban, S. Visualizing Hydrophobic and Hydrophilic Enzyme Interactions during Immobilization by Means of Infrared Microscopy. Catalysts 2022, 12, 989. [Google Scholar] [CrossRef]
  152. Wilkes, R.A.; Aristilde, L. Degradation and metabolism of synthetic plastics and associated products by Pseudomonas sp.: Capabilities and challenges. J. Appl. Microbiol. 2017, 123, 582–593. [Google Scholar] [CrossRef]
  153. Lu, D.; Wu, J.; Jin, S.; Wu, Q.; Deng, L.; Wang, F.; Nie, K. The enhancement of waste PET particles enzymatic degradation with a rotating packed bed reactor. J. Clean. Prod. 2024, 434, 140088. [Google Scholar] [CrossRef]
  154. Sharma, A.; Tewari, R.; Rana, S.S.; Soni, R.; Soni, S.K. Cellulases: Classification, Methods of Determination and Industrial Applications. Appl. Biochem. Biotechnol. 2016, 179, 1346–1380. [Google Scholar] [CrossRef]
  155. Sheldon, R.A.; van Pelt, S. Enzyme immobilisation in biocatalysis: Why, what and how. Chem. Soc. Rev. 2013, 42, 6223–6235. [Google Scholar] [CrossRef]
  156. Fernandez-Lafuente, R. Stabilization of multimeric enzymes: Strategies to prevent subunit dissociation. Enzym. Microb. Technol. 2009, 45, 405–418. [Google Scholar] [CrossRef]
  157. Thangaraj, B.; Solomon, P.R. Immobilization of Lipases—A Review. Part I: Enzyme Immobilization. ChemBioEng Rev. 2019, 6, 157–166. [Google Scholar] [CrossRef]
  158. Guzik, U.; Hupert-Kocurek, K.; Wojcieszynska, D. Immobilization as a strategy for improving enzyme properties- Application to oxidoreductases. Molecules 2014, 19, 8995–9018. [Google Scholar] [CrossRef]
  159. Seenuvasan, M.; Kumar, K.S.; Kumar, A.; Parthiban, R. Review on surface modification of nanocarriers to overcome diffusion limitations: An enzyme immobilization aspect. Biochem. Eng. J. 2020, 158, 107574. [Google Scholar] [CrossRef]
  160. Acevedo, F.; Pizzul, L.; Castillo, M.; González, M.; Cea, M.; Gianfreda, L.; Diez, M. Degradation of polycyclic aromatic hydrocarbons by free and nanoclay-immobilized manganese peroxidase from Anthracophyllum discolor. Chemosphere 2010, 80, 271–278. [Google Scholar] [CrossRef]
  161. Mazlan, S.Z.; Hanifah, S.A. Effects of Temperature and pH on Immobilized Laccase Activity in Conjugated Methacrylate-Acrylate Microspheres. Int. J. Polym. Sci. 2017, 2017, 5657271. [Google Scholar] [CrossRef]
  162. Mohamed, S.A.; Al-Harbi, M.H.; Almulaiky, Y.Q.; Ibrahim, I.H.; El-Shishtawy, R.M. Immobilization of horseradish peroxidase on Fe3O4 magnetic nanoparticles. Electron. J. Biotechnol. 2017, 27, 84–90. [Google Scholar] [CrossRef]
  163. Darwesh, O.M.; Matter, I.A.; Eida, M.F. Development of peroxidase enzyme immobilized magnetic nanoparticles for bioremediation of textile wastewater dye. J. Environ. Chem. Eng. 2019, 7, 102805. [Google Scholar] [CrossRef]
  164. Basri, R.S.; Rahman, R.N.Z.R.A.; Kamarudin, N.H.A.; Latip, W.; Ali, M.S.M. Characterization of Carboxylic Acid Reductase from Mycobacterium phlei Immobilized onto Seplite LX120. Polymers 2022, 14, 4375. [Google Scholar] [CrossRef]
  165. Johan, U.U.M.; Rahman, R.N.Z.R.A.; Kamarudin, N.H.A.; Latip, W.; Ali, M.S.M. Immobilization of Hyperthermostable Carboxylesterase EstD9 from Anoxybacillus geothermalis D9 onto Polymer Material and Its Physicochemical Properties. Polymers 2023, 15, 1361. [Google Scholar] [CrossRef]
  166. Bolivar, J.M.; Rocha-Martín, J.; Mateo, C.; Guisan, J.M. Stabilization of a highly active but unstable alcohol dehydrogenase from yeast using immobilization and post-immobilization techniques. Process Biochem. 2012, 47, 679–686. [Google Scholar] [CrossRef]
  167. Shinde, P.; Musameh, M.; Gao, Y.; Robinson, A.J.; Kyratzis, I.L. Immobilization and stabilization of alcohol dehydrogenase on polyvinyl alcohol fibre. Biotechnol. Rep. 2018, 19, e00260. [Google Scholar] [CrossRef]
  168. Nisha, S.; Karthick, S.A.; Gobi, N. Chemical Science Review and Letters A Review on Methods, Application and Properties of Immobilized Enzyme *Correspondence. Chem. Sci. Rev. Latters 2012, 1, 148–155. Available online: https://www.researchgate.net/publication/233882011_A_Review_on_Methods_Application_and_Properties_of_Immobilized_Enzyme (accessed on 5 August 2024).
  169. Das, R.; Mishra, H.; Srivastava, A.; Kayastha, A.M. Covalent immobilization of Β-amylase onto functionalized molybdenum sulfide nanosheets, its kinetics and stability studies: A gateway to boost enzyme application. Chem. Eng. J. 2017, 328, 215–227. [Google Scholar] [CrossRef]
  170. Asmat, S.; Husain, Q.; Azam, A. Lipase immobilization on facile synthesized polyaniline-coated silver-functionalized graphene oxide nanocomposites as novel biocatalysts: Stability and activity insights. RSC Adv. 2017, 7, 5019–5029. [Google Scholar] [CrossRef]
  171. Zhou, Q.Z.K.; Chen, X.D. Immobilization of β-galactosidase on graphite surface by glutaraldehyde. J. Food Eng. 2001, 48, 69–74. [Google Scholar] [CrossRef]
  172. Sheldon, R.A.; Brady, D. The limits to biocatalysis: Pushing the envelope. Chem. Commun. 2018, 54, 6088–6104. [Google Scholar] [CrossRef]
  173. Basso, A.; Serban, S. Industrial applications of immobilized enzymes—A review. Mol. Catal. 2019, 479, 110607. [Google Scholar] [CrossRef]
Figure 1. Locations of significant historical oil-spill incidents [6].
Figure 1. Locations of significant historical oil-spill incidents [6].
Processes 12 02074 g001
Figure 2. Exploiting biocatalytic degradation of hydrocarbon technique for new purpose.
Figure 2. Exploiting biocatalytic degradation of hydrocarbon technique for new purpose.
Processes 12 02074 g002
Figure 3. Research publications on “biocatalytic degradation”, “pollutants” and “wax”, based on data obtained from Google Scholar on 3 August 2024.
Figure 3. Research publications on “biocatalytic degradation”, “pollutants” and “wax”, based on data obtained from Google Scholar on 3 August 2024.
Processes 12 02074 g003
Figure 4. Reported alkane-degrading pathways, adopted and re-illustrated from Beilen et al., 2007 [79].
Figure 4. Reported alkane-degrading pathways, adopted and re-illustrated from Beilen et al., 2007 [79].
Processes 12 02074 g004
Figure 5. Illustration on using enzyme consortium consisting of different types of biocatalyst for paraffin wax degradation.
Figure 5. Illustration on using enzyme consortium consisting of different types of biocatalyst for paraffin wax degradation.
Processes 12 02074 g005
Figure 6. Illustration of how the amorphous region of petroleum is more susceptible to enzyme attack than the crystalline region (re-illustrated from Wei and Zimmerman, 2017 [140]).
Figure 6. Illustration of how the amorphous region of petroleum is more susceptible to enzyme attack than the crystalline region (re-illustrated from Wei and Zimmerman, 2017 [140]).
Processes 12 02074 g006
Figure 7. Proposed fluidized bed reactor concept for recovery and reuse of enzyme.
Figure 7. Proposed fluidized bed reactor concept for recovery and reuse of enzyme.
Processes 12 02074 g007
Table 1. Wax treatment/inhibition strategies and their limitations.
Table 1. Wax treatment/inhibition strategies and their limitations.
Wax Treatment/Inhibition StrategyPositive Findings from Previous Experimental StudiesLimitations
Scraping of pipelines using rubber/metal tool before the wax begin to buildup Suitable for non-heating pipeline, can reduce energy consumption for pipeline transportation [14]
  • The cutting force of the wax layer is dependent on the tool geometry and friction at its surface, e.g., wax removal is better at 45° rake angle than a negative 30° angle [15]
  • The sliding speed of the scraper affects the rubbing force, thus the wax deposit removal [38,39]
  • Formation of wax plug or scrapper became stuck caused by viscous petroleum [14]
Electrical heating, insulation and circulation of hot fluid to keep the oil above wax appearance temperatureWax removal below wax appearance temperature was possible for wax content below 14% [40]
More effective when asphaltene content in oil is high because activated asphaltenes have strong interactions with wax then decrease the wax appearance temperature [41]
4.
Electrical heaters frequently burn out thus require an automatic cut off system when overheating occurs [12]
5.
The insulation technique requires suitable material with low conductivity and chemical-resistance properties [12]
Ultrasonic and electromagnetic waves propagation to create cavitation and displace the waxThe wax appearance temperature of oil was reduced by 12 °C related to an increased in natural surfactant fractions, i.e., resin and asphaltene caused by modified composition after being treated with ultrasonic wave [17]
By morphology, the size of wax crystals was reduced and rounded after irradiation due to decreased in wax molecules [18]
6.
Requires combination with chemical [19,20,21,42] or thermal [22] to increase wax removal efficiency
Injection of chemical, e.g., wax inhibitor or pour-point depressant to reduce wax appearance temperature 65–73% wax reduced using polyacrylate polymer-based pour-point depressant [24]
7.
Wax inhibitors are typically solids at the operating site. Dilution with solvent increased expenses and raised possible hazards [12]
8.
Not universal, effectiveness is chemically specific to oil composition, varying from different well and over time. Various types of copolymers that functioned as wax crystal modifiers produced different results, and its mechanism is poorly understood despite many reported studies [28]
9.
Asphaltene content in oil may hindered chemical performance [29]
Injection of solvents, e.g., kerosene, gasoline or benzol to solubilize wax depositsSolvent formulation consisted of diesel, xylene, toluene, naphthalene and dispersant solubilized >85% of wax without requires heating above 90 °C [23]
10.
The effectiveness depends on the permeability of the wax deposit
11.
Posed toxicity and flammability risks to handlers [23]
Hydrocarbon-degrading bacteria to cleanly degrade the hydrocarbonBacteria is easy to get, inexpensive, easy to handle and biodegradable (environmentally friendly) [12]
12.
Dependent on nutrients and growth enhancers, and in the presence of sulfate-reducing bacteria, pipeline corrosion is a showstopper [37]
Table 2. Microbial treatment of petroleum-contaminated sites and the assayed enzymes.
Table 2. Microbial treatment of petroleum-contaminated sites and the assayed enzymes.
Source of HydrocarbonHydrocarbon-Degrading MicrobeReported EnzymesHydrocarbon Treatment ConditionHydrocarbon Removal EfficiencyReference
Contaminated soil Pseudomonas aeruginosaNot describedStrains inoculated to 500 g of contaminated soil55–84% in consortium; 25–47% in individual cultures [65]
Oil sludge from petroleum refinery Wild-type strains Enterobacter xiangfangensis STP-3; mutant strains Enterobacter Xiangfangensis M257 and Enterobacter Xiangfangensi M916Lipase, laccase, alkane hydroxylase, alcohol dehydrogenase, esterase10 g oil sludge and 20 g bacterial culture in 500 mL reactor82–87.5% in mutant strains; 72.15% in wild-type strains; in 7 days.[63]
Polluted seawater biosurfactant-producing hydrocarbonoclastic bacteria Enterobacter hormaecheiLipase, laccase, alkane hydroxylase, alcohol dehydrogenase, esterase5 liter polluted seawater with 10 g strains treated in 8 liter tank with bottom aeration. 85% in 10 days[64]
Oil sludge Acinetobacter sp. SCYY-5Not describedIncubated at 30 °C at 150 rpm 69.17% in 10 days[66]
Oil sludge from refineryPseudomonas aeruginosa; Staphylococcus sp.Not describedLab scale-anaerobic bioreactors 90% in 14 days[67]
Contaminated soil from oil processing Immobilized consortium of Flavobacterium johnsoniae BS1 and Shewanella baltica BS2Not describedIncubated at 30 °C at 150 rpm, and pH 7.593.32% in 3 days[68]
Oil sludge from refinery Indigenous hydrocarbonoclastic bacteriaLipase, laccase, esterase5 g oil sludge with 5 g immobilized bacteria88.78% in 7 days[60]
Contaminated soil Consortium of halotolerent and biosurfactant producing bacteria, Dietzia sp. CN-3 and Acinetobacter sp. HC8-3SAlkane hydroxylase genes, alkB and alkM100 g of soil with 5 g of crude oil, inoculated with 10 mL bacteria consortium, mixed and kept at 30 °C95.8% in 10 days[59]
Contaminated soil Indigenous bacteria Not describedDual-chamber reactors with membrane, filled with 250 mL saturated soil and artificial groundwater at 22 °C37.5% in 137 days[69]
Oil-in-saltwater emulsionsPseudomonas stutzeri immobilized on xerogel microspheresNot described0.15 g bacteria inoculated into 15 mL of oil-in-water emulsions cultured at 150 rpm and 35 °C10% in immobilized bacteria; 61% in free bacteria in 72 h [58]
Crude oilAcinetobacter sp.; Bacillus sp.; Pantoea sp. and Enterobacter sp.Not describedAddition of nitrogen, phosphorus and potassium (NPK)20–40% in 28 days[70]
Contaminated soil Heavy metal tolerant bacterium Novosphingobium panipatense P5:ABCAlkane hydroxylase, catechol 1,2-dioxygenase and catechol 2,3-dioxygenaseIncubated at 25 °C at 130 rpm90% degradation in 6 days[57]
Oily sludge from refineryBacteria consortium of Shewanalla chilikensis, Bacillus firmus and Halomonas hamiltonii Lipase, catalase, oxidoreductaseIncubated at 37 °C96% in 30 days[56]
Contaminated sediment from oil platform Indigenous bacteria Not described10 g oily sediments with additional surfactants, incubated at 30 °C at 120 rpm51.29% in 28 days[71]
Polluted water Consortium of protozoan species, Aspidisca sp., Trachelophyllum sp. And Peranema sp.,Not describedBioreactor at 30 °C, shaken continuously at 100 rpm61–90% by natural attenuation; 22–55% by nutrient supplementation with glucose; 10–67% by biostimulation with surfactant[72]
Crude oil Pseudomonas aeruginosa NCIM 5514Not describedIncubated at 37 °C at180 rpmOil viscosity reduced from 1883-1002 cp; 61.03% and 60.63% biodegradation of C8–C36+ [73]
Dehydrated crude oil Pseudomonas sp., Bacillus sp., Ochrobactrum sp.Not described10% inoculation, and shaking at 120 rpm47% in immobilized bacteria; 26% in free bacteria in 5 days[74]
HexadecaneAspergillus nigerNot describedElectric field pretreatment53–86% in 48 h[75]
Polluted seawater Dietzia maris CBMAI 705 and Micrococcus sp. CBMAI 636Not describedIncubated at 28 °C at 120 rpm 99% in 21 days[76]
Refinery wastewaterAlcaligenes odorans, Bacillus subtilis, Corynebacterium propinquum and Pseudomonas aeruginosaNot described37 °C in aerobic condition70%[77]
Table 3. Microbial treatment of petroleum hydrocarbons and the assayed enzymes.
Table 3. Microbial treatment of petroleum hydrocarbons and the assayed enzymes.
Source of HydrocarbonHydrocarbon-Degrading MicrobeReported EnzymesHydrocarbon Treatment ConditionHydrocarbon Removal EfficiencyReference
Crude oil Bacillus subtilis SL and Pseudomonas aeruginosa WJ-1Not describedIncubated for 7 days at 37 °C at 180 rpm32.61% in Bacillus subtilis SL; 54.35% in Pseudomonas sp. WJ-1; 58.60% in SL and WJ-1 [91]
n-hexadecaneImmobilized Bacillus thuringiensis (BTS)Not described37 °C at 120 rpm, and pH 781.62–86.65% in 192 h[92]
Alkane C7-C25Bacillus subtilis BL-27Not describedIncubated at 45 °C at 150 rpm 65% in 5 days[93]
Crude oil Raoultella ornithinolytica PS, Bacillus subtilis BJ11, Acinetobacter lwoffii BJ10, Acinetobacter pittii BJ6, Serratia marcescens PLNot described0.2 g crude oil with 5 mL strain, incubated at 30 °C at 180 rpm for 10 days.70–94% in 10 days[94]
Crude oil Rhodococcus erythropolis OSDS1, Serratia proteamaculans S1BD1, Alcaligenes sp. OPKDS2, Rhizobium erythropolis OSDS1, Rhizobium sp. PNS1, Pseudomonas sp. BSS9BS1Not describedIncubated at 30 °C at 130 rpm85.26% 15 days[95]
Crude oil Cupriavidus sp. OPK, Rhodococcus erythropolis OSPS1, Pseudomonas sp. BSSNot describedIncubated at 120 rpm in dark room74–83% in 3 days[96]
Diesel Halomonas sp. and Aneurinibacillus sp.Not describedIncubated at 151 rpm at 30 °C82.65% in 12 days[97]
Bonny light crude oil Bacillus sp. SB4, Pseudomonas sp. SC8, Serratia sp. SC11, and Acinetobacter sp. SC12Not describedIncubated at 180 rpm at 30 °C12–36%[98]
Petroleum hydrocarbons, hexane, hexadecane and motor oilAlcanivorax borkumensisAlkane hydroxylase, lipase and esteraseIncubated at 30 °C at 150 rpm75–81.5% in 72 h[90]
Hexadecane and BTEXAlcanivorax borkumensis SK2Alkane hydroxylase, lipase and esteraseIncubated at 32 °C at 120 rpm79–96.7%[82]
Long-chain n-alkanes C32 and C40Geobacillus thermoparaffinivorans IR2, Geobacillus stearothermophillus IR4 and Bacillus licheniformis MN6Alkane hydroxylase, alcohol dehydrogenase and lipaseIncubated at 50 °C at 120 rpm87–90% in 20 days (for C32); 82–88% in 20 days (for C40)[89]
Belayim Mix crude oil Lipomyces tetrasporus RS-Y1 and Paecilomyces variotii RS-F3Not describedIncubated at 28 °C at 150 rpm58.15–68.3% in 30 days[99]
Alkane mixture C8–C22Anoxybacillus sp. WJ-4alkB homologs genesIncubated at 70 °C at 180 rpm58.75% in 40 days[84]
Crude oil (C12-C25) and asphalt (C13-C30) Pestalotiopsis sp. NG007Catechol 1,2-dioxygenase, catechol 2,3 dioxygenase, laccase, manganese peroxidase, lignin peroxidaseIncubated at 25 °C77–92% in 30 days[88]
Crude oilScenedesmus obliquus and Chlorella vulgarisNot describedIncubated at 25 °C at 80 rpm under dark condition46–88% in 6 weeks[100]
Crude oil Bacillus atrophaeus 5-2a, Bacillus aryabhattai 6-2a and Bacillus amyloliquefaciens 6-2cNot described2 g crude oil in 20 mL bacterial suspension, incubated at 40 °C82.32–94.50%[101]
Diesel oil and used engine oil Stenotrophomonas maltophilia, Bacillus cereus and Bacillus pumilusLipase Aerobic conditions at 30 °C94.13% in diesel oil; 99.77% in used engine oil[87]
Alkane C8–C36Alcanivorax dieselolei B-5AlkB homologs (AlkB1 and AlkB2), CYP153 homolog (P450), alkane hydroxylase (AlmA) Incubated at 28 °C at 200 rpmp450 was upregulated on C8–C16; AlmA on C22–C36; alkB1 and alkB2 C12–C26[85]
Alkane C5–C36Geobacillus thermodenitrificans NG80-2Alkane monooxygenase LadA LadA was upregulated on C15–C36[86]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saadon, S.; Rahman, R.N.Z.R.A.; Kamarudin, N.H.A.; Shahruddin, S.; Shafian, S.R.M.; Wazir, N.A.; Ali, M.S.M. Novel Approach of Tackling Wax Deposition Problems in Pipeline Using Enzymatic Degradation Process: Challenges and Potential Solutions. Processes 2024, 12, 2074. https://doi.org/10.3390/pr12102074

AMA Style

Saadon S, Rahman RNZRA, Kamarudin NHA, Shahruddin S, Shafian SRM, Wazir NA, Ali MSM. Novel Approach of Tackling Wax Deposition Problems in Pipeline Using Enzymatic Degradation Process: Challenges and Potential Solutions. Processes. 2024; 12(10):2074. https://doi.org/10.3390/pr12102074

Chicago/Turabian Style

Saadon, Shazleen, Raja Noor Zaliha Raja Abd Rahman, Nor Hafizah Ahmad Kamarudin, Sara Shahruddin, Siti Rohaida Mohd Shafian, Norhidayah Ahmad Wazir, and Mohd Shukuri Mohamad Ali. 2024. "Novel Approach of Tackling Wax Deposition Problems in Pipeline Using Enzymatic Degradation Process: Challenges and Potential Solutions" Processes 12, no. 10: 2074. https://doi.org/10.3390/pr12102074

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