Next Article in Journal
SIV: Raise the Correlation of Second-Order Correlation Power Analysis to 1.00
Previous Article in Journal
User–Topic Modeling for Online Community Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review on Corrosion Inhibitors for Oil and Gas Corrosion Issues

Department of Petroleum Engineering, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(10), 3389; https://doi.org/10.3390/app10103389
Submission received: 3 January 2020 / Revised: 29 January 2020 / Accepted: 10 February 2020 / Published: 14 May 2020

Abstract

:
The pipeline system in the oil and gas industry is the heart for transportation of crude and refined petroleum. Nevertheless, continuous exposure of the pipeline surfaces to impurities and sources of corrosion such as sulfur and chromate is totally unavoidable. Vast employment of commercial corrosion inhibitors to minimize the corrosion is being restrained due to toxicity towards the environment. The emergence of “green” chemistry has led to the use of plant extracts and fruit wastes which have proven to be good corrosion inhibitors. This paper aims to provide insight into carrying out further investigation under this research theme for accurate inhibition efficiency measurement.

1. Introduction

In oil and gas industries, the corrosion issue has always been of great importance, with consequences similar result to those of natural disaster. Corrosion normally occurs in oil and gas pipelines. Since the pipelines play the role of transporting oil and gas from the wellheads to the processing facilities, they are exposed to the continuous threat of corrosion, from the date of commissioning up to decommissioning or abandonment. According to [1], the rough estimation of the aggregate yearly cost of corrosion is $1.372 billion, which is the total of surface pipeline and facility costs ($589 million), down-hole tubing costs ($463 million), as well as capital expenses ($320 million).
Corrosion inhibitors are one of the mediums applied to minimize corrosion in petroleum industries. For an optimum inhibition to be achieved, the inhibitors must be added above a certain minimum concentration. There are plenty of techniques, e.g., cathodic protection [2,3], organic coatings [4,5,6], and application of first-rate corrosion-resistant alloys [7], that can be implemented to fight against corrosion, yet film-forming inhibitors are still known to be the unrivalled method of defense for mild steel in an acidic environment [8,9]. The film-forming inhibitors are used in industries to create a molecular layer right on the surface of the steel and aliphatic tail as a second layer in hydrocarbon to prevent the water from contacting the steel surface and causing corrosion [10].
Recently, the rise of the “green” chemistry concept in the fields of science, technology and engineering [11,12] is restraining the application of commercial corrosion inhibitors by implementing certain theories or ideas to reduce the contamination [13] from being discharged into the environment as well as coming up with the eco-friendly chemicals [14,15,16]. As a movement to support this concept, the use of green-based corrosion inhibitors like plant extracts [17], chemical drugs [18], and ionic liquids [19,20] are being practiced. These green inhibitors are organic compounds that function through the adsorption on the surface of the metal to prevent the occurrence of corrosion. Moreover, fruit-based corrosion inhibitors are also one of the natural elements utilized due to their richness in vitamins, minerals, and phenolic compounds. Nevertheless, the corrosion inhibitors adhere to certain factors like concentration, rate of dispersion, velocity, temperature, film persistency, pH, flow regime, and fluid composition, as well as the presence of instabilities able to perturbate the flow in minimizing corrosion.
This review paper assesses the trilateral view of corrosion: mechanism of corrosion, sources of corrosion, commercial and green-based corrosion inhibitors. Hence, this paper is anticipated to serve as the foundation for future research on green corrosion inhibitors. Note that comprehensive discussions or analysis regarding the current state and the future of green-based inhibitors will not be covered.

2. Corrosion

Corrosion is observed as one of the main reasons for the failures of oil and gas infrastructure. The existence of corrosion is the consequence of chemicals such as naphthenic acid (NA) reacting with iron particles or developing a surface film; this occurs with sulfur particles (S) in the hydrocarbon industries. As the foremost drivers of corrosion, sulfur and naphthenic acid exist as organic acids in various crude oils. However, the rate of corrosion is also dependent on the quality of the crude oil, its acidic constituents, and the environment of the transport [21]. It is crucial to study the nature of these acids and the amount of sulfur and naphthenic acid components present in the crude oil to understand the performance and the root of corrosion. Despite the defects in oil and gas infrastructure (e.g., pipelines), the nature of crude oil itself promotes corrosion due to its harmful impurities like naphthenic acid and sulfur [22].
Pipelines, as one of the common tools of the oil and gas industries, have seen an increased demand in infrastructure due to the augmentation of the Canadian oil and gas industry to create improved operational and management conditions. It is vital to maintain the integrity of this pipeline infrastructure from being affected by the environment in ways that will have consequences of economic loss [23,24]. Moreover, internal corrosion of the pipelines turns out to be a key threat to the initial stage of production [25]. In accordance with this, more than 9000 failures due to internal corrosion were reported from 1990 to 2012 [26], which accounted for 54.8% of all spills. The United States’ (US) oil and gas companies disburse 1.052 billion dollars annually to prevent internal corrosion [25]. Considering these issues, there is an urge to come up with an effective corrosion prevention approach within the given budget of the companies.

2.1. Mechanism of Corrosion

Anodes and cathodes are the two cells of corrosion generally used to demonstrate the transfer of charges between the iron and electrolyte as well as within the iron itself. The presence of charged ions in the electrolyte causes the electrolyte to transform into an electrically conductive solution. During the corrosion process, the metal ions shift from the active site (anode) into the solution and pass the electrons from the metal at the lower active site (cathode) to an acceptor. Electron acceptors like oxygen, oxidizing agents, or hydrogen ions are required for the cathodic process to take place. Equation (1) represents the general chemical corrosion reaction in the presence of oxygen in moist air. Figure 1 signifies the basic electrochemical cell built using the anode and cathode cells partially immersed into an electrolyte.
The anode cell experiences rusting during the oxidation process, whereas the cathode cell reduces but does not rust. In the production of oil and gas, the Fe2+ ions are produced at the anode when the iron from steel is driven into the solution. These ions act in response with oxygen, hydrogen sulfide, or carbon dioxide to form decay yields as presented below. The additional electrons change from the anode to the cathode where hydroxyl ions are produced by reducing water. The hydrogen ions are broken into hydrogen gas by the electrons if the oxygen does not exist at the cathode. The anodic and cathodic positions are regions on the outer surface of the metal that vary in electric potential. Due to the occurrence of salts, the electrolyte is normally transformed from water into being conductive.
The chemical reaction of corrosion is given as follows:
4 F e + 3 O 2 = 2 F e 2 O 3

2.2. Sources of Corrosion

Mostly, corrosion is thought to occur in the water phase, except when the water is restricted in the middle of the stream or surrounded by oil, in which there will be no occurrence of corrosion. The major sources of corrosion in the oil industries are hydrochloric acid and its aqueous solutions, hydrogen sulfide, corrosion of steel at hydrocarbon–electrolyte interfaces and in emulsified two-phase environments, oxygen, naphthenic acids, carbon dioxide, as well as water cut. The sources and respective roles are described below:
a.
Hydrogen Sulfide
Hydrogen sulfide is known to be very harmful in the corrosion of metals or alloys, regardless of its application in oil and gas. In a matter of fact, hydrogen sulfide can be the root of sulfide stress corrosion cracking (SSCC) failure in pipelines. However, hydrogen sulfide is only corrosive when it is dissolved in water, where its solubility is relatively higher than that of carbon dioxide and oxygen. Moreover, hydrogen sulfide can cause danger in sour oil and gas fields due to its abundance in oil and production processes. The issues associated with the corrosion by the hydrogen sulfide acid are becoming more significant, as the sulfur availability in crude oil is growing proportionally with the reduction of existing sweet oil [28].
Iron sulfide and hydrogen are the products of the corrosion by sulfide. The hydrogen crack is created by the internal stresses that are caused by the molecular hydrogen. The outcomes of hydrogen cracking are unexpected and disastrous most of the time, since there are not any visible signs shown during the early phase [29]. Therefore, it is vital to select the optimal material for well completions, particularly in fields containing sour oil and gas, as the presence of hydrogen sulfide and tensile strength can possibly be the root of sulfide stress cracking (SSC) that might result in equipment loss.
b.
Chloride
Chloride, which can be found in the mineralized water at the well bottom is known as another vital substance that causes severe corrosion at high temperatures. This can lead to failure that results from intergranular corrosion and chloride stress corrosion cracking (CSCC). Both the existence of sulfide and chloride have been detected in deep sour gas production that consists mainly of methane without liquid hydrocarbons under high pressure [30].
c.
Carbon dioxide
Normally, the corrosion that results from carbon dioxide is labelled as “sweet corrosion”. The carbon dioxide that is produced along with oil and gas dissolves in water to procedure carbonic acid, ensuing a decrease in pH. Moreover, using an injection method for enhanced oil recovery can lead to further corrosion that is caused by CO2. This will eventually lead to downhole and surface equipment corrosion. One of the characteristics of CO2-based corrosion is pitting, which is imaged as deep, sharp-edged pits.
In the oil and gas industries, one of the most acidic surroundings found is the aqueous system that has higher concentrations of carbon dioxide. The partial pressure, temperature, speed, and pH are the examples of factors that contribute to corrosion in the existence of carbon dioxide [31,32].
d.
Oxygen
Oxygen plays a vital role as the corrosive agent in the secondary recovery by water flooding. The corrosion caused by the dissolved oxygen causes pits in the drill pipe, where propagated fatigue cracks occur due to the stress. Additionally, oxygen can be detrimental to water injection equipment like pumps, piping, and water storage tanks, and the byproduct of corrosion might plug the formation.
Figure 2 shows a comparison of the rusting rates of oxygen, hydrogen sulfide, and carbon dioxide on carbon steel in a water solution containing 2 to 5 g/L sodium chloride at the temperature of 25 °C. From this, it can be derived that oxygen ought to be the source for higher rusting rates at much lesser concentrations than carbon dioxide and hydrogen sulfide. It should be considered that the rusting rates in pits can be a few times larger and that the carbon dioxide rusting rate is inconsistent at low temperatures. Mixtures of oxygen with carbon dioxide or hydrogen sulfide prompt immediate-corrosion environments with smaller oxygen concentrations such as 0.1 ppm [33].
a.
Bacteria
Microorganism activity is the root of bacterial corrosion, which can be damaging, especially in enhanced recovery processes. Sulphate-reducing and iron bacteria are known to be the most common bacteria that promote corrosion. Microbiologically-influenced corrosion (MIC) was thoroughly examined from the industrial practice perspective, in order to attain profitable solutions in observing large water injection systems [34].
b.
Water Cut
The presence of oil is considered advantageous as it employs a kind of inhibition effect by forming a film on a steel surface to prevent water wetting. On the other hand, gas and condensates do not employ any useful effect due to its non-inhibition characteristic. As for the vertical tubing, an oil film produced on the surface of the steel is steady up to about 20%–40% water cuts. Using the de Waard and Milliams method, the corrosion rate for higher water quantities can be predicted, as the steel can be considered water-wet. Besides, the volume of water is not a significant factor for the horizontal pipes.
Commonly, water is heavier compared to oil, gas and condensed products, thus it may segregate on the lowest surfaces at the 6 o’clock position for stratified flow. As for this case, the corrosion is anticipated to occur only on the water-wet surfaces. For the stratified flow, corrosion is probable to happen in the highest point of the line because of the condensation of water droplets from the wet gas. In this case, the inhibition effect is said to be poor and the corrosion rate in the highest point of the line can be expected to be 10% of the projected rate in a completely immersed condition, regardless of the carbon dioxide content.
c.
Strong Acids
In the oil and gas industries, the acid stimulation method is widely applied to enhance production by improving the formation permeability. The commonly used fluids for this method are hydrofluoric acid and hydrochloric or acetic acids for sandstone and carbonates, respectively, which can cause corrosion of production tubing, downhole tools and casing in the absence of corrosion inhibitors [27].
d.
Brines
Usually, brines are utilized during the completion stage of oil and gas wells to aid the final operations before the fluid production begins. This is because brines consist of zinc chlorides or calcium bromides and have the capability to regulate the well without damaging the formation. However, they can be harmful to the downhole equipment of the well, since there is the presence of dissolved oxygen.

3. Commercial Corrosion Inhibitors

As problems related to the corrosion issue have been skyrocketing, especially in the oil and gas industries, corrosion scientists have developed various techniques to mitigate this issue, such as corrosion inhibitors [35], cathodic protection [36], and paint-based corrosion inhibitors [37]. Most of the corrosion inhibitors utilized are mainly organic and inorganic composites, where organic inhibitors reduce corrosion through adsorption techniques, while the inorganic ones prevent corrosion by reacting with the anodic or cathodic parts of the process [38,39].
Organic compounds that consist of nitrogen, oxygen, and/or sulfur are considered as competent industrial corrosion inhibitors [40]. These inhibitors inherit the ability to form a protective layer between the metal surface and corrosive environment [41] through the adsorption process to delay the metal disintegration [42,43,44]. For instance, azole and pyrimidine byproducts are the most-used inhibitors to minimize the corrosion of metals and alloys by injecting the inhibitors into the system at low treatment concentrations [45,46]. In addition, organic inhibitors have the tendency to work productively in all acid concentrations and do not poison the refinery catalysts.
Nevertheless, the disposal of contaminated industrial corrosion inhibitors results in damaging effects on the ecosystem; thus, the use of green and environmental-friendly inhibitors is on the rise [47,48]. To illustrate, the benzotriazoles that are usually utilized to minimize the corrosion of heat exchangers were proven to poison marine creatures, even at concentrations as low as 3 ppm [49]. However, organic inhibitors chemically reduce as time passes in the presence of acid yet cannot withstand at temperatures above 95 °C; furthermore, their use is expensive.
Next, inorganic corrosion inhibitors contain the salts of zinc, copper, nickel, arsenic, and additional metals, with the arsenic compounds being the ones that are most commonly used. When these arsenic compounds are mixed with the corrosive solution, they scrape at the cathode cell of the unprotected metal surfaces. The plating reduces the percentage of hydrogen ion interchange due to the formation of iron sulfide amid the steel and acids that act as an obstacle. The reaction of acid with iron sulfide is known as a dynamic process. There are advantages as well as disadvantages when using the inorganic inhibitors. The advantages are they work excellently at high temperatures for longer periods and are less expensive than organic inhibitors. As for the shortcomings, inorganic inhibitors are more likely to lose their grip in acid solutions that are stronger than 17% hydrochloric acid, tougher to combine, and might release toxic arsine gas as the product of corrosion.
Although these inhibitors are potent at high concentrations, they are unarguably damaging to the environment. In conjunction with this problem, new inhibitors concentrating on polymers [50,51,52] and plant-based [53,54] substances were studied. However, polymers were defined as moderate inhibitors, needing secondary refinement that will only result in higher cost [55,56]. Thus, extensive research is being undertaken on plant parts, as they are easily accessible in an inexpensive way.

4. Green-Based Corrosion Inhibitors

Green-based inhibitors that are nontoxic in nature, such as plant extracts, have higher demand compared to commercial inhibitors [57,58,59]. This is because plant extracts are observed to be green and sustainable materials due to their natural and biological properties and will be able to inhibit the metals and alloys from corroding [60]. The leaf, out of all parts of the plant, has the utmost preference for its abundance of phytochemicals (active components) produced through synthesis, that act similarly to commercial inhibitors. It is also vital to acknowledge that the extract of other parts of a plant such as root, bark, flower, fruit, wood, seed and peel have contributed to the inhibition efficiency [61,62,63]. Furthermore, the phytochemical synthesis uses carbon dioxide, which is known as the highly poisonous greenhouse gas, to undergo the photosynthesis, contributing to the green chemistry theory as well.
The green corrosion inhibitors function when used in a very low concentration to inhibit the metal surface from a corrosive medium. The rate of corrosion by adsorption process on the metal surface is affected by plant extracts via influencing either the anodic or cathodic reaction kinetics and then affecting the rate of diffusion of aggressive ions from interacting with the metal surface. Consecutively, a layer of film can be established by increasing the electrical resistance of the metal surface [64]. Besides, green corrosion inhibitors are well known for their adsorptive properties (site-blocking elements), enabling the active molecules from the plant extract to adsorb on exposed metal surfaces [65]. Equation (2) shows the working mechanism of the inhibitor molecules in the form of neutral molecules adsorbing on the metal surface instead of the hydrogen ions [66]:
Inhibitor + nHads → Inhibitorads + H2
where nHads is adsorbed hydrogen ions sourced from water and Inhibitorads is adsorbed neutral molecules sourced from plant extract.
In conjunction with this, there are a few limitations that must be examined during the preparation of plant extracts. Normally, the solvents used for extraction will diffuse into plant tissue, solubilize, and extract the available phytochemicals [67,68]. Hence, it is essential to select the right solvent for better results. One of the solvents that is readily available, cheap, and safe is water [69,70], yet ethanol and methanol are still in demand for selective plant extracts [71,72]. Next, the temperature creates a noticeable result when it comes to extract preparation. The solubility of the phytochemicals will be hindered at relatively low temperature, whereas at high temperatures, the phytochemicals result in decomposition. The recommended temperature for an ideal extraction falls between the range of 60–80 °C [73,74]. As for the drying temperature, oven drying is advisable, since drying at room temperature can take up to months to accomplish.
Green-based corrosion inhibitors can be divided into two classes: organic and inorganic [75,76,77]. The organic class of green-based corrosion inhibitors consist of synthetic substances that are nontoxic for the environment. Flavonoids, alkaloids, and byproducts of plants are example of organic inhibitors [78]. The inorganic class of inhibitors are vastly utilized in aqueous systems due to their high productivity [79]. The chromates exhibit a toxic nature in which the employment of this inorganic inhibitor for industrial use is limited. Concerning this issue, lanthanide salts were studied as an eco-friendly inhibitor substitute [80].
Correspondingly, the fruit wastes, e.g., seeds and peels, have gained noticeable attention for their natural antioxidant properties. For instance, mango, orange, passionfruit, and cashew peels are known to contain ample amounts of antioxidants like polyphenols, carotenoids, and vitamins C and E. Emphasizing on the phenolic compounds, specifically flavonoids, the efficiency of the antioxidant activity depends on their structural characteristics like figure and location of phenolic hydroxyls. On the other hand, mango, orange and passion fruits are said to have ample sources of pectin, which belongs to the polysaccharide group [81].
Recently, a study on banana peel extracts was conducted for mild steel in an acidic environment, and another study was performed simultaneously, utilizing mango and orange peel to combat corrosion of mild steel, aluminum, zinc, and copper in acid solutions (HCl and H2SO4) [82]. In conjunction with this, it can be derived that the industrial waste and aqueous fruit peel extracts can be used as green-based corrosion inhibitors, as they fulfil the criteria of green chemistry. Table 1 and Table 2 show the gap analysis from the previous case studies of inhibitors from plant- and fruit-based origin from 2010 to 2019. Remarks listed under the gap column give insight towards carrying out further investigation under this research theme for accurate inhibition efficiency measurement.

5. Conclusions

The issue is unavoidable for the oil and gas industry, creating a similar impact to those of natural disasters. Hence, completely stopping this issue is not possible, but taking preventative measure to inhibit the metal surface from corroding is more economical. The use of conventional corrosion inhibitors has long been in practice. Nonetheless, employment of commercial inhibitors has been restrained over time. This is due to their toxic properties that contribute to the destruction of the environment. Hence, with the aim of combating this, knowledge concerning the green (plant- and fruit-based) corrosion inhibitors has been consolidated in this paper to aid in mitigating the corrosion of pipelines.

Author Contributions

Conceptualization, K.T. and H.H.; Resources, K.T. and H.H.; Writing—original draft preparation, K.T.; Writing—review and editing, K.T. and H.H.; Funding acquisition, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by YUTP-FRG grant (015LC0-064).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simon, M.R. Report of Offshore Technology Conference (OTC) Presentation; NACE International Oil and Gas production: Houston, TX, USA, 2008. [Google Scholar]
  2. Chen, R.; Li, X.; Du, C.; Cheng, Y. Effect of cathodic protection on corrosion of pipeline steel under disbonded coating. Corros. Sci. 2009, 51, 2242–2245. [Google Scholar] [CrossRef]
  3. Yabuki, A.; Tanabe, S.; Fathona, I. Comparative studies of two benzaldehyde thiosemicarbazone derivatives as corrosion inhibitors for mild steel in 1.0 M HCl. Surf. Coat. Technol. 2018, 341, 71–77. [Google Scholar] [CrossRef]
  4. Lyon, S.; Bingham, R.; Mills, D. Corrosion Protection of Carbon Steel by Pongamia glabra Oil- Based Polyetheramide Coatings. Prog. Org. Coat. 2017, 102, 2–7. [Google Scholar] [CrossRef] [Green Version]
  5. Chengduan, W.; Chuan, L.; Bin, X.; Xiaogang, G.; Dong, F.; Bin, L. Corrosion inhibition of mild steel in HCl medium by S-benzyl-O,O′-bis(2-naphthyl)dithiophosphate with ultra-long lifespan. Results Phys. 2018, 10, 558–567. [Google Scholar]
  6. Salman, T.; Al-Azawi, K.; Mohammed, I.; Al-Baghdadi, S.; Al-Amiery, A.; Gaaz, T. Experimental and quantum chemical simulations on the corrosion inhibition of mild steel by 3-((5-(3,5-dinitrophenyl)-1,3,4-thiadiazol-2-yl)imino)indolin-2-one. Results Phys. 2018, 10, 291–296. [Google Scholar] [CrossRef]
  7. Odewunmi, N.; Umoren, S.; Gasem, Z.; Ganiyu, S.; Muhammad, Q. Electrochemical and quantum chemical studies on carbon steel corrosion protection in 1 M H2SO4 using new eco-friendly Schiff base metal complexes. J. Taiwan Inst. Chem. Eng. 2015, 51, 177–185. [Google Scholar] [CrossRef]
  8. Zeino, A.; Abdulazeez, I.; Khaled, M.; Jawich, M.; Obot, I. Electrochemical Corrosion Performance of Aromatic Functionalized Imidazole Inhibitor Under Hydrodynamic Conditions on API X65 Carbon Steel in 1 M HCl Solution. J. Mol. Liq. 2018, 250, 50–62. [Google Scholar] [CrossRef]
  9. Umoren, S.; Eduok, U. Application of carbohydrate polymers as corrosion inhibitors for metal substrates in different media: A review. Carbohydr. Polym. 2016, 140, 314–341. [Google Scholar] [CrossRef]
  10. Yadav, D.; Maiti, B.; Quraishi, M. Electrochemical and quantum chemical studies of 3,4-dihydropyrimidin-2(1H)-ones as corrosion inhibitors for mild steel in hydrochloric acid solution. Corros. Sci. 2010, 52, 3586–3598. [Google Scholar] [CrossRef]
  11. Hu, K.; Zhuang, J.; Zheng, C.; Ma, Z.; Yan, L.; Gu, H.; Zeng, X.; Ding, J. Effect of novel cytosine-l-alanine derivative based corrosion inhibitor on steel surface in acidic solution. J. Mol. Liq. 2016, 222, 109–117. [Google Scholar] [CrossRef]
  12. Ramezanzadeh, B.; Vakili, H.; Amini, R. The effects of addition of poly (vinyl) alcohol (PVA) as a green corrosion inhibitor to the phosphate conversion coating on the anticorrosion and adhesion properties of the epoxy coating on the steel substrate. Appl. Surf. Sci. 2015, 327, 174–181. [Google Scholar] [CrossRef]
  13. Mohammadinejad, R.; Karimi, S.; Iravani, S.; Varma, R. Plant-derived nanostructures: Types and applications. Green Chem. 2016, 18, 20–52. [Google Scholar] [CrossRef]
  14. Varma, R. Journey on greener pathways: From the use of alternate energy inputs and benign reaction media to sustainable applications of nano-catalysts in synthesis and environmental remediation. Green Chem. 2014, 16, 2027–2041. [Google Scholar] [CrossRef]
  15. Jeon, H.; Lim, C.; Lee, M.; Kim, S. Chemical assay-guided natural product isolation via solid-supported chemodosimetric fluorescent probe. Chem. Sci. 2015, 6, 2806–2811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Srivastava, M.; Tiwari, P.; Srivastava, S.; Prakash, R.; Ji, G. Electrochemical investigation of Irbesartan drug molecules as an inhibitor of mild steel corrosion in 1 M HCl and 0.5 M H2SO4 solutions. J. Mol. Liq. 2017, 236, 184–197. [Google Scholar] [CrossRef]
  17. Mo, S.; Li, L.; Luo, H.; Li, N. An example of green copper corrosion inhibitors derived from flavor and medicine: Vanillin and isoniazid. J. Mol. Liq. 2017, 242, 822–830. [Google Scholar] [CrossRef]
  18. Diamanti, M.; Velardi, U.; Brenna, A.; Mele, A.; Pedeferri, M.; Ormellese, M. Compatibility of imidazolium-based ionic liquids for CO2 capture with steel alloys: A corrosion perspective. Electrochim. Acta 2016, 192, 414–421. [Google Scholar] [CrossRef] [Green Version]
  19. Lozano, I.; Mazario, E.; Olivares-Xometl, C.; Likhanova, N.; Herrasti, P. Corrosion behaviour of API 5LX52 steel in HCl and H2SO4 media in the presence of 1, 3-dibencilimidazolio acetate and 1,3-dibencilimidazoliododecanoate ionic liquids as inhibitors. Mater. Chem. Phys. 2014, 147, 191–197. [Google Scholar] [CrossRef]
  20. El-Hajjaji, F.; Messali, M.; Aljuhani, A.; Aouad, M.; Hammouti, B.; Belghiti, M.; Chauhan, D.; Quraishi, M. Pyridazinium-based ionic liquids as novel and green corrosion inhibitors of carbon steel in acid medium: Electrochemical and molecular dynamics simulation studies. J. Mol. Liq. 2018, 249, 997–1008. [Google Scholar] [CrossRef]
  21. Mahmoodian, M.; Qingi, C. Failure assessment and safe life prediction of corroded oil and gas pipelines. J. Pet. Sci. Eng. 2017, 151, 434–438. [Google Scholar] [CrossRef]
  22. Wang, W.; Natelson, R.; Stikeleather, L.; Roberts, W. CFD simulation of transient stage of continuous countercurrent hydrolysis of canola oil. Comput. Chem. Eng. 2012, 43, 108–119. [Google Scholar] [CrossRef]
  23. Nesic, S. Key issues related to modelling of internal corrosion of oil and gaspipelines—A review. Corros. Sci. 2007, 49, 4308–4338. [Google Scholar] [CrossRef]
  24. Revie, R. Oil, Gas Pipelines. In Integrity and Safety Handbook; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  25. Papavinasm, S. Corrosion Control in the Oil and Gas Industry; Elsevier: Houston, TX, USA, 2013. [Google Scholar]
  26. Regulator, A.E. Report 2013-B: Pipeline Performance in Albarta, 1990–2012; Alberta Energy Regulator: Calgary, AB, Canada, 2013. [Google Scholar]
  27. Mahmood, A.; Dawood, H. A Comprehensive Review of Corrosion and its Inhibition in the Oil and Gas Industry. In Proceedings of the SPE Kuwait Oil and Gas Show and Conference, Mishref, Kuwait, 11–14 October 2015; Society of Petroleum Engineers. [Google Scholar] [CrossRef]
  28. Szyprowski, A. Methods of Investigation on Hydrogen Sulfide Corrosion of Steel and Its Inhibitors. Corrosion. Corrosion 2003, 59, 68–81. [Google Scholar] [CrossRef]
  29. Carter, D.R.; Adams, N.J. Hydrogen Sulfide in the Drilling Industry. In Proceedings of the 4th Deep Drilling and Production Symposium, Amarillo, TX, USA, 1–3 April 1979; Society of Petroleum Engineers of AIME: Dallas, TX, USA, 1979; pp. 123–135. [Google Scholar]
  30. Carrell, M.A. Reclaiming Produced Water for Steam Generation in the Kern River Field. In Proceedings of the 54th Ann. Conf. Soc. Pet Eng. A.I.M.E., SPE 8411, Las Vegas, NV, USA, 23–26 September 1979. [Google Scholar]
  31. Fang, H.; Nesic, S.; Brown, B.; Wang, S. General CO2 Corrosion in High Salinity Brines. Corrosion 2006. [Google Scholar]
  32. Ramachandran, S.; Bartrip, K.; Menendez, C.; Coscio, S. Preventing Erosion and Erosion Corrosion Using Specialty Chemicals. In International Symposium on Oilfield Chemistry; Society of Petroleum Engineers: Houston, TX, USA, 2003. [Google Scholar]
  33. Yesudass, S.; Olasunkanmi, L.O.; Bahadur, I.; Kabanda, M.M.; Obot, I.B.; Ebenso, E.E. Experimental and theoretical studies on some selected ionic liquids with different cations/anions as corrosion inhibitors for mild steel in acidic medium. J. Taiwan Inst. Chem. Eng. 2016, 64, 252–268. [Google Scholar] [CrossRef]
  34. Farquhar, G. A review of trends in Microbiologically influenced corrosion. Mater. Perform. 1993, 32, 53. [Google Scholar]
  35. Umoren, S.; Solomon, M. Recent developments on the use of polymers as corrosion Inhibitors—A review. Open Mater. Sci. J. 2014, 8, 39–54. [Google Scholar] [CrossRef]
  36. Abreu, C.; Izquierdo, M.; Merino, P.; Nóvoa, X.; Pérez, C. A new approach to the determination of the cathodic protection period in zinc-rich paints. Corrosion 1999, 55, 1173–1181. [Google Scholar] [CrossRef]
  37. Mansfeld, F.; Tsai, C. Determination of coating deterioration with EIS: I. Basic relationships. Corrosion 1991, 47, 958–963. [Google Scholar] [CrossRef]
  38. Umoren, S.; Obot, I.; Madhankumar, A.; Gasem, Z. Performance evaluation of pectin as ecofriendly corrosion inhibitor for X60 pipeline steel in acid medium. Carbohydr. Polym. 2015, 124, 280–291. [Google Scholar] [CrossRef]
  39. Tiu, B.D.B.; Advincula, R.C. Polymeric corrosion inhibitors for the oil and gas industry: Design graphic and mechanism. React. Funct. Polym. 2015, 95, 25–45. [Google Scholar] [CrossRef]
  40. Goyal, M.; Kumar, S.; Bahadur, I.; Verma, C.; Ebenso, E. Organic corrosion inhibitors for industrial cleaning of ferrous and non-ferrous metals in acidic solutions: A review. J. Mol. Liq. 2018, 256, 565–573. [Google Scholar] [CrossRef]
  41. Verma, C.; Quraishi, M.; Singh, A. 5-Substituted 1H-tetrazoles as effective corrosion inhibitors for mild steel in 1M hydrochloric acid. J. Taibah Univ. Sci. 2016, 10, 718–733. [Google Scholar] [CrossRef] [Green Version]
  42. Azhar, M.; Mernari, M.; Traisnel, M.; Bentiss, F.; Lagrenee, M. Corrosion inhibition of mild steel by the new class of inhibitors [2, 5-bis(n-pyridyl)-1,3,4-thiadiazoles] in acidic media. Corros. Sci. 2011, 43, 2229–2243. [Google Scholar] [CrossRef]
  43. Mobin, M.; Zehra, S.; Parveen, M. l-Cysteine as corrosion inhibitor for mild steel in 1M HCl and synergistic effect of anionic, cationic and non-ionic surfactants. J. Mol. Liq. 2016, 216, 598–607. [Google Scholar] [CrossRef]
  44. Da Silva, A.; D’Elia, E.; Gomes, J. Carbon steel corrosion inhibition in hydrochloric acid solution using a reduced Schiff base of ethylenediamine. Corros. Sci. 2010, 52, 788–793. [Google Scholar] [CrossRef]
  45. Gong, Y.; Wang, Z.; Gao, F.; Zhang, S.; Li, H. Synthesis of new benzotriazole derivatives containing carbon chains as the corrosion inhibitors for copper in sodium chloride solution. Ind. Eng. Chem. Res. 2015, 54, 12245–12253. [Google Scholar] [CrossRef]
  46. Ghazoui, A.; Saddik, R.; Benchat, N.; Guenbour, M.; Hammouti, B.; Al-Deyab, S.S.; Zarrouk, A. Comparative study of pyridine and pyrimidine derivaitves as corrosion inhibitors of C38 steel in molar HCl. Int. J. Electrochem. Sci. 2012, 7, 7080–7097. [Google Scholar]
  47. Raja, P.; Sethuraman, M. Natural products as corrosion inhibitor for metals in corrosive media—A review. Mater. Lett. 2008, 62, 113–116. [Google Scholar] [CrossRef]
  48. Verma, C.; Ebenso, E.; Bahadur, I.; Quraishi, M. An overview on plant extracts as environmental sustainable and green corrosion inhibitors for metals and alloys in aggressive corrosive media. J. Mol. Liq. 2018, 266, 577–590. [Google Scholar] [CrossRef]
  49. Pillard, D.; Cornell, J.; Dufresne, D.; Hernandez, M. Toxicity of benzotriazole and benzotriazole derivatives to three aquatic species. Water Res. 2001, 35, 557–560. [Google Scholar] [CrossRef]
  50. Abdallah, M.; Megahed, H.; Radwan, M.; Abdfattah, E. Polyethylene glycol compounds as corrosion inhibitors for aluminum in 0.5 M hydrochloric acid solution. J. Am. Sci. 2012, 8, 49–55. [Google Scholar]
  51. Awad, M.; Metwally, M.; Soliman, S.; El-Zomrawy, A.; Bedair, M. Experimental and quantum chemical studies of the effect of polyethylene glycol as corrosion inhibitors of aluminum surface. J. Ind. Eng. Chem. 2013, 20, 796–808. [Google Scholar] [CrossRef]
  52. Umoren, S.; Solomon, M.; Israel, A.; Eduok, U.; Jonah, A. Comparative study of the corrosion inhibition efficacy of polypropylene glycol and poly(methacrylic acid) for mild steel in acid solution. J. Dispers. Sci. Technol. 2015, 36, 1721–1735. [Google Scholar] [CrossRef]
  53. Al-Otaibi, M.; Al-Mayouf, A.; Khan, M.; Mousa, A.; Al-Mazroa, S.; Alkhathlan, H. Corrosion inhibitory action of some plant extracts on the corrosion of mild steel in acidic media. Arab J. Chem. 2014, 7, 340–346. [Google Scholar] [CrossRef] [Green Version]
  54. Gerengi, H.; Jazdzewska, A.; Kurtay, M. A comprehensive evaluation of mimosa extract as a corrosion inhibitor on AA6060 alloy in acid rain solution: Part I. Electrochemical AC methods. J. Adhes. Sci. Technol. 2015, 29, 36–48. [Google Scholar] [CrossRef]
  55. Rajeswari, V.; Kesavan, D.; Gopiraman, M.; Viswanathamurthi, P. Physicochemical studies of glucose, gellan gum, and hydroxypropyl cellulose—Inhibition of cast iron corrosion. Carbohydr. Polym. 2013, 95, 288–294. [Google Scholar] [CrossRef]
  56. Solomon, M.; Umoren, S.; Udousoro, I.; Udoh, A. Inhibitive and adsorption behavior of carboxymethyl cellulose on mild steel corrosion in sulphuric acid solution. Corros. Sci. 2010, 52, 1317–1325. [Google Scholar] [CrossRef]
  57. Hussin, M.; Rahim, A.; Ibrahim, M.; Brosse, N. The capability of ultrafiltrated alkaline and organosolv oil palm (Elaeis guineensis) fronds lignin as green corrosion inhibitor formild steel in 0.5 M HCl solution. Measurement 2016, 78, 90–103. [Google Scholar] [CrossRef]
  58. Mohammed, M.; Khan, Z.; Siddiquee, A. Surface modifications of titanium materials for developing corrosion behavior in human body environment: A review. Proc. Mater. Sci. 2014, 6, 1610–1618. [Google Scholar] [CrossRef] [Green Version]
  59. Negm, N.; Kandile, N.; Aiad, I.; Mohammad, M. New eco-friendly cationic surfactants synthesis, characterization and applicability as corrosion inhibitors for carbon steel in 1 M HCl. Colloids Surf. A Phys. Eng. Asp. 2011, 391, 224–233. [Google Scholar] [CrossRef]
  60. Chemat, F.; Vian, M.; Cravotto, G. Green extraction of natural products: Concept and principles. Int. J. Mol. Sci. 2012, 13, 8615–8627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. El-Etre, A. Inhibition of C-steel corrosion in acidic solution using the aqueous extract of zallouh root. Mater. Chem. Phys. 2008, 108, 278–282. [Google Scholar] [CrossRef]
  62. Ji, G.; Dwivedi, P.; Sundaram, S.; Prakash, R. Inhibitive effect of Chlorophytum borivilianum root extract on mild steel corrosion in HCl and H2SO4 solutions. Ind. Eng. Chem. Res. 2013, 52, 10673–10681. [Google Scholar] [CrossRef]
  63. Ji, G.; Dwivedi, P.; Sundaram, S.; Prakash, R. Aqueous extract of Argemone Mexicana roots for effective protection of mild steel in an HCl environment. Res. Chem. Intermed. 2016, 42, 439–459. [Google Scholar] [CrossRef]
  64. Krishnaveni, K.; Ravichandran, J. Effect of aqueous extract of leaves of Morinda tinctoria on corrosion inhibition of aluminium surface in HCl medium. Trans. Nonferrous Met. Soc. China 2014, 24, 2704–2712. [Google Scholar] [CrossRef]
  65. Sanatkumar, B.; Nayak, J.; Shetty, A. Influence of 2-(4-chlorophenyl)-2-oxoethyl benzoate on the hydrogen evolution and corrosion inhibition of 18 Ni 250 grade weld aged maraging steel in 1.0 M sulfuric acid medium. Int. J. Hydrog. Energy 2012, 37, 9431–9442. [Google Scholar] [CrossRef]
  66. Manamela, K.; Murulana, L.; Kabanda, M.; Ebenso, E. Adsorptive and DFT studies of some imidazolium based ionic liquids as corrosion inhibitors for zinc in acidic medium.of some imidazolium based ionic liquids as corrosion inhibitors for zinc in acidic medium. Int. J. Electrochem. Sci. 2014, 9, 3029–3046. [Google Scholar]
  67. Capello, C.; Fischer, U.; Hungerbühler, K. What is a green solvent? A comprehensive framework for the environmental assessment of solvents. Green Chem. 2007, 9, 927–934. [Google Scholar] [CrossRef]
  68. Nasrollahzadeh, M.; Sajadi, S.; Khalaj, M. Green synthesis of copper nanoparticles using aqueous extract of the leaves of Euphorbia esula L and their catalytic activity for ligand-free Ullmann-coupling reaction and reduction of 4-nitrophenol. RSC Adv. 2014, 4, 47313–47318. [Google Scholar] [CrossRef]
  69. Sharghi, H.; Khalifeh, R.; Doroodmand, M. Copper nanoparticles on charcoal for multicomponent catalytic synthesis of 1, 2, 3-triazole derivatives from benzyl halides or alkyl halides, terminal alkynes and sodium Azide in water as a “green” solvent. Adv. Synth. Catal. 2009, 351, 207–218. [Google Scholar] [CrossRef]
  70. Bose, D.; Fatima, L.; Mereyala, H.B. Green chemistry approaches to the synthesis of 5-alkoxycarbonyl-4-aryl-3, 4-dihydropyrimidin-2 (1 H)-ones by a three component coupling of one-pot condensation reaction: Comparison of ethanol, water, and solvent-free connditions. J. Organomet. Chem. 2003, 68, 587–590. [Google Scholar] [CrossRef] [PubMed]
  71. Duan, H.; Wang, D.; Li, Y. Green chemistry for nanoparticle synthesis. Chem. Soc. Rev. 2015, 44, 5778–5792. [Google Scholar] [CrossRef]
  72. Varma, R. Greener and sustainable trends in synthesis of organics and nanomaterials. Chem. Eng. 2016, 4, 5866–5878. [Google Scholar] [CrossRef]
  73. Mohamad, N.; Arham, N.; Jai, J.; Hadi, A. Plant extract as reducing agent in synthesis of metallic nanoparticles: A review. Adv. Mater. Res. Trans. Tech. Publ. 2014, 832, 350–355. [Google Scholar] [CrossRef]
  74. Seo, J.; Lee, S.; Elam, M.; Johnson, S.; Kang, J.; Arjmandi, B. Study to find the best extraction solvent for use with guava leaves (Psidium guajava L.) for high antioxidantefficacy. Food Sci. Nutr. 2014, 2, 174–180. [Google Scholar] [CrossRef]
  75. Aida, Z.; Razika, A.; Laid, M.; Kamel, B.; Boualem, S. Inhibition of acid corrosion of mild steel by aqueous nettle extracts. Pigment. Resin Technol. 2014, 43, 127–138. [Google Scholar]
  76. Devarayan, K.; Mayakrishnan, G.; Sulochana, N. Green inhibitors for corrosion of metals: A review. Chem. Sci. Rev. Lett. 2012, 1, 1–8. [Google Scholar]
  77. Gobara, M.; Zaghloul, B.; Baraka, A.; Elsayed, M.; Zorainy, M.; Kotb, M.; Elnabarawy, H. Green corrosion inhibition of mild steel to aqueous sulfuric acid by the extract of Corchorus ollitorius stems. Mater. Res. Express 2017, 4, 391–401. [Google Scholar] [CrossRef]
  78. McCafferty, E. Thermodynamic aspects of the crevice corrosion of iron in chromate/chloride solutions. J. Electrochem. Soc. 1979, 126. [Google Scholar] [CrossRef]
  79. Bethencourt, M.; Botana, F.; Calvino, J.; Marcos, M.; Rodriguez-Chacon, M. Lanthanide compounds as environmentally-friendly corrosion inhibitors of aluminium alloys: A review. Corros. Sci. 1998, 40, 1803–1819. [Google Scholar] [CrossRef]
  80. Michodjehoun-Mestres, L.; Jean-Marc, S.; Fulcrand, H.; Bouchut, C.; Reynes, M.; Jean-Marc, B. Characterisation of highly polymerised prodelphinidins from skin and flesh of four cashew apple (Anacardium occidentale L.) genotypes. Food Chem. 2009, 112, 851. [Google Scholar] [CrossRef]
  81. Saleh, R.; Ismail, A.; Hosary, A.E.B. Inhibitory mechanism of low-carbon steel corrosion by mimosa tannin in sulphuric acid solutions. Corros. J. 1982, 3, 17. [Google Scholar]
  82. Dehghani, A.; Bahlakeha, G.; Ramezanzadeh, B.; Ramezanzadeh, M. Potential of Borage flower aqueous extract as an environmentally sustainable corrosion inhibitor for acid corrosion of mild steel: Electrochemical and theoretical studies. J. Mol. Liq. 2019, 277, 895–911. [Google Scholar] [CrossRef]
  83. Othman, N.K.; Yahya, S.; Ismail, M.C. Corrosion inhibition of steel in 3.5% NaCl by rice straw extract. J. Ind. Eng. Chem. 2019, 70, 299–310. [Google Scholar] [CrossRef]
  84. Alibakhshi, E.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B.; Mahdavian, M.; Motamedi, M. Glycyrrhiza glabra leaves extract as a green corrosion inhibitor for mild steel in 1 M hydrochloric acid solution: Experimental, molecular dynamics, Monte Carlo and quantum mechnics study. J. Mol. Liq. 2018, 255, 185–198. [Google Scholar] [CrossRef]
  85. Asadi, N.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Utilizing Lemon Balm extract as an effective green corrosion inhibitor for mild steel in 1 M HCl solution: A detailed experimental, molecular dynamics, Monte Carlo and quantum mechanics study. J. Taiwan Inst. Chem. Eng. 2019, 95, 252–272. [Google Scholar] [CrossRef]
  86. Haldhara, R.; Prasada, D.; Saxenaa, A.; Kumarb, R. Experimental and theoretical studies of Ficus religiosa as green corrosion inhibitor for mild steel in 0.5 M H2SO4 solution. Sustain. Chem. Pharm. 2018, 9, 95–105. [Google Scholar] [CrossRef]
  87. Haldhar, R.; Prasad, D.; Saxena, A. Myristica fragrans extract as an eco-friendly corrosion inhibitor for mild steel in 0.5 M H2SO4 solution. J. Environ. Chem. Eng. 2018, 6, 2290–2301. [Google Scholar] [CrossRef]
  88. Hassannejad, H.; Nouri, A. Sunflower seed hull extract as a novel green corrosion inhibitor for mild steel in HCl solution. J. Mol. Liq. 2018, 254, 377–382. [Google Scholar] [CrossRef]
  89. Ikeuba, A.; Okafor, P. Green corrosion protection for mild steel in acidic media: Saponins and crude extracts of Gongronema latifolium. Pigment Resin Technol. 2018, 48, 57–64. [Google Scholar] [CrossRef]
  90. Jmiai, A.; El Ibrahimi, B.; Tara, A.; Chadili, M.; El Issami, S.; Jbara, O.; Khallaayoun, A.; Bazzi, L. Application of Zizyphus Lotuse-pulp of Jujube extract as green and promising corrosion inhibitor for copper in acidic medium. J. Mol. Liq. 2018, 268, 102–113. [Google Scholar] [CrossRef]
  91. Khadom, A.; Abd, A.; Ahmed, N. Xanthium strumarium leaves extracts as a friendly corrosion inhibitor of low carbon steel in hydrochloric acid: Kinetics and mathematical studies. S. Afr. J. Chem. Eng. 2018, 25, 13–21. [Google Scholar] [CrossRef]
  92. Saxena, A.; Prasad, D.; Haldhar, R. Investigation of corrosion inhibition effect and adsorption activities of Cuscuta reflexa extract for mild steel in 0.5 M H2SO4. Bioelectrochemistry 2018, 124, 156–164. [Google Scholar] [CrossRef]
  93. Ugi, B.U.; Obeten, M.E.; Magu, T.O. Phytochemical constituents of Taraxacum officinale leaves as eco-friendly and nontoxic organic inhibitors for stainless steel corrosion in 0.2 M HCl acid medium. Int. J. Chem. Sci. 2018, 2, 35–43. [Google Scholar]
  94. Aribo, S.; Olusegun, S.J.; Ibhadiyi, L.J.; Oyetunji, A.; Folorunso, D. Green inhibitors for corrosion protection in acidizing oilfield environment. J. Assoc. Arab Univ. Basic Appl. Sci. 2017, 24, 34–38. [Google Scholar] [CrossRef] [Green Version]
  95. Gerengia, H.; Uygura, I.; Solomona, M.; Yildiza, M.; Goksub, H. Evaluation of the inhibitive effect of Diospyros kaki (Persimmon) leaves extract on St37 steel corrosion in acid medium. Sustain. Chem. Pharm. 2016, 4, 57–66. [Google Scholar] [CrossRef]
  96. Umoren, S.A.; Eduok, U.M.; Solomon, M.M.; Udoh, A.P. Corrosion inhibition by leaves and stem extracts of Sida acuta for mild steel in 1 M H2SO4 solutions investigated by chemical and spectroscopic techniques. Arab. J. Chem. 2016, 9, S209–S224. [Google Scholar] [CrossRef] [Green Version]
  97. Li, X.; Deng, S.; Fu, H. Inhibition of the corrosion of steel in HCl, H2SO4 solutions by bamboo leaf extract. Corros. Sci. 2012, 62, 163–175. [Google Scholar] [CrossRef]
  98. Quraishi, M.; Singh, A.; Singh, K.V.; Yadav, D.; Singh, A. Green approach to corrosion inhibition of mild steel in hydrochloric acid and sulphuric acid solutions by the extract of Murraya koenigii leaves. Mater. Chem. Phys. 2010, 122, 114–122. [Google Scholar] [CrossRef]
  99. Sanaei, Z.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Use of Rosa canina fruit extract as a green corrosion inhibitor for mild steel in 1 M HCl solution: A complementary experimental, molecular dynamics and quantum mechanics investigation. J. Ind. Eng. Chem. 2019, 69, 18–31. [Google Scholar] [CrossRef]
  100. Liao, L.; Moa, S.; Luo, H.; Li, N. Corrosion protection for mild steel by extract from the waste of lychee fruit in HCl solution: Experimental and theoretical studies. J. Colloid Interface Sci. 2018, 520, 41–49. [Google Scholar] [CrossRef]
  101. Tiwari, P.; Srivastava, M.; Mishra, G.R.; Prakash, R. Economic use of waste Musa paradisica peels for effective control of mild steel loss in aggressive acid solutions. J. Environ. Chem. Eng. 2018, 6, 4773–4783. [Google Scholar] [CrossRef]
  102. Liao, L.; Mo, S.; Luo, H.; Li, N. Longan seed and peel as environmentally friendly corrosion inhibitor for mild steel in acid solution: Experimental and theoretical studies. J. Colloid Interface Sci. 2017, 499, 110–119. [Google Scholar] [CrossRef]
  103. Paul, S.; Koley, I. Corrosion Inhibition of Carbon Steel in Acidic Environment by Papaya Seed as Green Inhibitor. J. Bio Tribo Corros. 2016, 2, 6. [Google Scholar] [CrossRef] [Green Version]
  104. Reddy, C.; Sanketi, B.; Kumar, S. Corrosion inhibition of mild steel by Capsicum annuum fruit paste. Perspect. Sci. 2016, 8, 603–605. [Google Scholar] [CrossRef] [Green Version]
  105. Singh, A.; Lin, Y.; Ebenso, E.; Liu, W.; Pan, J.; Huang, B. Gingko biloba fruit extract as an eco-friendly corrosion inhibitor for J55 steel in CO2 saturated 3.5% NaCl solution. J. Ind. Eng. Chem. 2015, 24, 219–228. [Google Scholar] [CrossRef]
  106. Singh, M.; Gupta, P.; Gupta, K. The litchi (Litchi chinensis) peels extract as a potential green inhibitor in prevention of corrosion of mild steel in 0.5 M H2SO4 solution. Arab. J. Chem. 2015. [Google Scholar] [CrossRef] [Green Version]
  107. Odewunmi, N.; Umoren, S.; Gasem, Z. Watermelonwaste products as green corrosion inhibitors for mild steel in HCl solution. J. Environ. Chem. Eng. 2015, 3, 286–296. [Google Scholar] [CrossRef]
  108. Yaro, A.; Khadom, A.; Wael, R. Apricot juice as green corrosion inhibitor of mild steel in phosphoric acid. Alexandra Eng. J. 2013, 52, 129–135. [Google Scholar] [CrossRef] [Green Version]
  109. Rocha, J.; Gomes, J.P.; D’Elia, E. Corrosion inhibition of carbon steel in hydrochloric acid solution by fruit peel aqueous extracts. Corros. Sci. 2010, 52, 2341–2348. [Google Scholar] [CrossRef]
  110. Ossai, C. Advances in Asset Management Techniques: An Overview of Corrosion Mechanisms and Mitigation Strategies for Oil and Gas Pipelines. Int. Sch. Res. Not. 2012, 2012, 570143. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Electrochemical cell [27].
Figure 1. Electrochemical cell [27].
Applsci 10 03389 g001
Figure 2. Comparison of rusting rates of three gases [27].
Figure 2. Comparison of rusting rates of three gases [27].
Applsci 10 03389 g002
Table 1. Summary of research gaps for plant-based green corrosion inhibitors from 2010 to 2019. -Note: EIS-Electrochemical impedance spectroscopy; IE- inhibition efficiency; FTIR- fourier transformed infrared spectroscopy
Table 1. Summary of research gaps for plant-based green corrosion inhibitors from 2010 to 2019. -Note: EIS-Electrochemical impedance spectroscopy; IE- inhibition efficiency; FTIR- fourier transformed infrared spectroscopy
DetailsGapReference/Year
(a) Borage flower
(b) Experiments: Weight loss, EIS, surface analysis
(c) Parameters: Concentration (200, 400, 600, 800 ppm) and immersion time (0.5, 2.5, 5.0 h)
(d) Results: 800 ppm; 2.5 h; 91% IE
Limitation: Constant temperature (25 °C)
Remark: Vary the temperature from 25 up to 90 °
[83]/2019
(a) Rice straw extract
(b) Experiments: Weight loss, surface and morphology analysis, and electrochemical test
(c) Parameters: Immersion time (7, 14, 21, 28, 35, 42 days) at room temperature (25 °C)
(d) Results: Immersion time of 7 to 14 days and 85% IE
Limitation: Constant temperature (25 °C) used throughout 42 days
Remark: Vary the temperature from 25 up to 90 °C
[84]/2019
(a) Glycyrrhiza glabra (Pea and bean family) leaves
(b) Experiments: EIS, surface characterization
(c) Parameters: Concentration of inhibitor (200, 400, 600, 800 ppm)
(d) Results: 800 ppm gave 88% IE
Limitation:Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[85]/2018
(a) Lemon balm extracts
(b) Experiments: Characterization technique (LBE, EIS, surface analysis)
(c) Parameters: Inhibitor concentration (200, 400, 600, 800 ppm) and immersion time (0.5, 2, 4, 6, 12, 24 h)
(d) Result: 800 ppm with immersion time of 24 h and 94.6% IE
Limitation: Constant temperature (25 °C)
Remark: Vary the temperature from 25 up to 90 °C
[86]/2018
(a) Ficus religiose (leaf, bodhi tree)
(b) Experiments: EIS, gravimetric measurements, quantum chemical study, SEM
(c) Parameters: Temperature (25, 35, 45 °C), inhibitor concentration (100–500 ppm)
(d) Results: 50 ppm gave 88.29% IE at 25 °C
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[87]/2018
(a) Myristica fragrans (nutmeg fruit)
(b) Experiments: Weight loss, UV-vis spectroscopy, FT-IR spectroscopy, NMR analysis, quantum chemical studies, SEM
(c) Parameter: Inhibitor concentration (100, 200, 300, 400, 500 ppm)
D) Results: 500 ppm gave 87.81% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[88]/2018
(a) Sunflower seed hull (flower)
(b) Experiments: FT-IR, UV-vis
(c) Parameters: Inhibitor concentration (50, 100, 200, 300, 400 ppm) and temperatures (25, 40, 50, 60 °C)
(d) Result: 400 ppm gave 98.46% IE at 60 °C
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[89]/2018
(a) Gongronema latifolium (utazi, herb)
(b) Experiments: Gasometric method
(c) Parameters: Inhibitor concentration (50, 100, 250, 500, 1000ppm) and temperature (30, 40, 50, 60 °C)
(d) Results
i. EEGL: 1000 ppm gave 93.7% IE at 30 °C
ii. SEGL: 1000 ppm gave 96.5% IE at 50 °C
Limitation: Immersion time was not stated in this article, and only one major experiment was carried out
Remark:
Immersion time should be tested from 3 up to 30 days
[90]/2018
(a) Zizyphus lotuse (lotus)
(b) Active ingredients: Vitamin C (ascorbic acid), linoleic acid, oleanolic acid, flavonoid compound, triterpenoic acid, jujuboside
(c) Experiments: Electrochemical methods, potentiodynamic polarization, SEM and EDS analysis
(d) Parameters: Concentration of inhibitors (0.05–2 g L−1) and Temperatures (25, 35, 45, 55 °C)
(e) Results:
i. Concentration effect: 1000 ppm gave 93% IE
ii. Temperature effect at 1000 ppm: 25 °C; 93.16% IE
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[91]/2018
(a) Xanthium strumarium (cocklebur) leaf extract
(b) Experiments: SEM, FTIR, weight loss
(c) Parameters: Inhibitor concentration (200, 400, 600, 800, 1000 ppm) and temperature (30, 40, 50, 60 °C)
(d) Results: 1000 ppm gave 94.82% IE at 60 °C
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[92]/2018
(a) Cuscuta reflexa (morning glory family, fruit extract)
(b) Experiments: Weight loss, electrochemical measurement, UV-visible spectroscopy, FT-IR spectroscopy, surface analyses, quantum chemical studies
(c) Parameter: Inhibitor concentration (100, 200, 300, 400, 500 ppm)
(d) Results: 500 ppm gave 95.47% IE
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[93]/2018
(a) Taraxacum officinale (dandelion, flower)
(b) Experiments: Weight loss, Thermometric measurements, Electrochemical measurements, Gravimetric
(c) Parameters: Type of crude (saponins—SETOL; flavonoids—FETOL; alkaloids—AETOL) and inhibitor concentration (10, 30, 70, 150, 300 ppm) and temperature (25, 40, 60 °C)
(d) Results
i. Gravimetric
AETOL
300 ppm; 99.3% IE
ii. Thermometric
AETOL
300 ppm; 25 °C; 98.2% IE
iii. EIS
AETOL; 79.0% IE
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[94]/2018
(a) Tridax procumbens (daisy flower) and Chromolaena odorata (Christmas bush-leaf)
(b) Experiments: EIS and adsorption isotherm
(c) Parameters: Inhibitor concentration (100, 200, 300, 400 ppm)
(d) Results: 100 ppm gave 95.06% IE
Limitation: Constant temperature (40 °C) and immersion time
Remark: Vary the temperature from 25 up to 90 °C and the immersion time from 3 up to 30 days
[95]/2018
(a) Diospyros kaki (persimmon)
(b) Experiments: EIS and polarization, weight loss, surface analysis
(c) Results: 225 ppm gave 83.45% IE at immersion time of 6 h
Limitation: Constant temperature (25 °C) and short immersion time (3–6 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[96]/2016
(a) Sida acuta leaves and stem (wireweed)
(b) Experiments: Weight loss, hydrogen evaluation measurement, spectrophotometric analysis
(c) Parameters: Temperature (30–60 °C)
(d) Results: 500 ppm gave 85% (leaves) and 52% (stem) IE at 30 °C
Limitation: Constant immersion time (24 h) used throughout experiment
Remark: Vary the immersion time from 3 up to 30 days
[97]/2016
(a) Bamboo leaf extract
(b) Experiments: Weight loss measurements, electrochemical measurements and atomic force microscope
(c) Parameters: Acid concentration (1 M HCl and 0.5 M H2SO4), temperatures (20, 30, 40, 50 °C), immersion time (6–160 h)
(d) Results:
i. 1 M HCl
40 °C temperature
91.2% IE
ii. 0.5 M H2SO4
50 °C temperature
86.5% IE
iii. Immersion time
36 to 160 h
95% IE; 1 M HCl
86% IE; 0.5 M H2SO4
Limitation: Constant inhibitor concentration was used
Remark: Vary the concentration from 50 to 250 ppm
[98]/2012
(a) Murraya koenigii (curry leaves)
(b) Experiments: Weight loss method, EIS
(c) Parameters: Inhibitor concentration, acid concentration
(d) Results
i. Concentration: 600 ppm
96.66% and 94.66% IE in HCl and H2SO4
ii. Acid concentration 1 M HCl; 97.54% IE
Limitation: Small temperature range (35–65 °C) and short immersion time (2 to 8 h) used
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[99]/2010
Table 2. Summary of research gap for fruit-based green corrosion inhibitors from 2010 to 2019.
Table 2. Summary of research gap for fruit-based green corrosion inhibitors from 2010 to 2019.
DetailsGapReferences/Year
(a) Rosa canina fruit
(b) Experiments: Characterization, quantum chemical and EIS
(c) Parameters: Inhibitor concentration (200, 400, 600, 800 ppm) and immersion time (2, 4, 6, 24, 48 h)
(d) Result: 600 ppm gave 85.7% IE at immersion time of 6 h
Limitation: Constant temperature (25 °C)
Remark: Vary the temperature from 25 up to 90 °C
[100]/2019
(a) Lychee waste
(b) Experiments: Weight loss, EIS, FTIR and SEM, and computational studies
(c) Parameters: Extraction process (blank, etoh-U, etoh-R, H2O-U), immersion time (1.5, 3.0, 4.5 h) and inhibitor concentration
(300, 400, 500, 600, 700 ppm)
(d) Results
Etoh-U: 97.95% IE
1.5 h: 97.95% IE
600 ppm: 97.95% IE
Limitation: Constant temperature (25 °C)
Remark: Vary the temperature from 25 up to 90 °C
[101]/2018
(a) Musa paradisica peels (banana)
(b) Experiments: EIS, polarization, surface analysis
(c) Parameters: Acid solution (1 M HCl and 0.5 M H2SO4) and inhibitor concentration (200, 300, 400 ppm)
(d) Results: 1 M HCl, 400 ppm gave 90% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[102]/2018
(a) Longan seed and peel
(b) Experiment: EIS, Weight loss, FTIR, SEM and computational studies
(c) Parameters: Inhibitor concentration
(300, 400, 500, 600 ppm) and temperature (25, 35, 45, 55 °C)
(d) Results:
600ppm: 92.93% IE
55 °C: 89.29% IE
Limitation: Constant immersion time (24 h)
Remark: Vary the immersion time from 3 up to 30 days
[103]/2017
(a) Papaya Seed
(b) Experiments: Electrochemical studies, adsorption studies
(c) Parameter: Different H2SO4 solutions (0.5 M, 1 M, 3 M)
(d) Results: 3 M H2SO4 gave 90% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[104]/2016
(a) Capsicum annuum fruit paste
(b) Experiments: Weight loss, contact angle measurements, analysis of protective film
(c) Parameter: Immersion time (24, 96, 168 h)
(d) Results: 96.48% IE at immersion time of 24 h
Limitation: Constant temperature (25 °C) and concentration
Remark: Vary the temperature from 25 up to 90 °C and concentrations from 50 to 250 ppm
[105]/2016
(a) Gingko biloba fruit
(b) Experiments: MS, FTIR, EIS, contact angle measurement and SEM
(c) Parameters: Inhibitor concentration (250, 500, 1000 ppm)
(d) Results: 1000 ppm gave 97% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[106]/2015
(a) Litchi fruit
(b) Experiments: Weight loss, EIS, surface analysis
(c) Parameter: Inhibitor concentration (25, 75, 100, 150, 200, 300 ppm)
(d) Results: 300 ppm gave 97.8% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[107]/2015
(a) Watermelon waste
(b) Experiments: EIS, SEM, UV-vis and FTIR
(c) Parameters: Watermelon waste (rind, seed, peel) and inhibitor concentration (10, 50, 100, 200 ppm)
(d) Results:
Rind: 200 ppm, 79.86% IE
Seed: 200 ppm, 83.67% IE
Peel: 200 ppm, 72.42% IE
Limitation: Constant temperature (25 °C) and immersion time (24 h)
Remark: Vary the temperature from 25 up to 90 °C and extend the immersion time from 3 up to 30 days
[108]/2015
(a) Apricot juice
(b) Experiments: Adsorption study and inhibition mechanism
(c) Parameters: Inhibitor concentration (100, 200, 300, 400 ppm) and temperature (30, 40, 50, 60 °C)
(d) Results: 400 ppm gave 75% IE at 30 °C
Limitation: Constant immersion time (24 h) used throughout experiment
Remark:
Vary the immersion time from 3 up to 30 days
[109]/2013
(a) Fruit peels
(b) Experiments: EIS, polarization, weight loss, SEM
(c) Parameters: Type of fruit peels (mango, orange, passion, and cashew), inhibitor concentration (100–800 ppm), immersion time (1, 4, 24 h) and temperature (25, 40, 60 °C)
(d) Results:
i. Mango
600 mg/L: 91% IE
Orange
400 mg/L: 95% IE
Passion
500 mg/L: 90% IE
Cashew
800 mg/L: 80% IE
ii. Immersion time
24 h: 96% IE
iii. Temperature
25 °C: 92% IE
No limitation or remarks[110]/2010

Share and Cite

MDPI and ACS Style

Tamalmani, K.; Husin, H. Review on Corrosion Inhibitors for Oil and Gas Corrosion Issues. Appl. Sci. 2020, 10, 3389. https://doi.org/10.3390/app10103389

AMA Style

Tamalmani K, Husin H. Review on Corrosion Inhibitors for Oil and Gas Corrosion Issues. Applied Sciences. 2020; 10(10):3389. https://doi.org/10.3390/app10103389

Chicago/Turabian Style

Tamalmani, Kausalya, and Hazlina Husin. 2020. "Review on Corrosion Inhibitors for Oil and Gas Corrosion Issues" Applied Sciences 10, no. 10: 3389. https://doi.org/10.3390/app10103389

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