Startseite Corrosion inhibition by imidazoline and imidazoline derivatives: a review
Artikel Öffentlich zugänglich

Corrosion inhibition by imidazoline and imidazoline derivatives: a review

  • Nipaporn Sriplai

    Nipaporn Sriplai received a PhD degree in material science and nanotechnology from Khon Kaen University, Thailand. Presently, she works as a postdoctoral researcher at National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA), Thailand. Her research interests lie in the development of corrosion inhibitor formulations to prevent corrosion in gas pipelines and synthesis of corrosion inhibition compounds.

    und Korakot Sombatmankhong

    Korakot Sombatmankhong received a PhD in chemical engineering from of University of Cambridge, UK. Currently, she is a senior researcher at National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA), Thailand. She has more than 10 years of research experience on chemical deformulation/formulation, polymer gel electrolyte, fuel cells, organic synthesis and arsenic/mercury removal. She has been working in several industry-sponsored research projects, such as the development of corrosion inhibition formulation, imidazoline synthesis and pipeline decommissioning processes.

    ORCID logo EMAIL logo
Veröffentlicht/Copyright: 8. März 2023

Abstract

Imidazoline and imidazoline derivatives are extensively employed as effective corrosion inhibitors due to their low toxicity, low cost and environmental friendliness. Their chemical structure consists of a 5-membered heterocyclic ring (C3N2H4) with two nitrogen atoms that are readily adsorbed onto metal surfaces. Also, a pendant side chain or alkyl amine substituent acts as an anchor that helps to maintain its adsorption on steel surfaces. The tail portion is a long hydrocarbon chain that can form a hydrophobic film on a surface. These molecular structures make it very attractive as a starting point for several enhancements in corrosion inhibition research. Moreover, modification of an imidazoline structure can be more effective in enhancing its effectiveness in corrosion inhibition. This review compiled all information regarding imidazoline and imidazoline derivatives used as effective corrosion inhibitors in the petroleum industry. It includes their chemical structures and properties, synthesis processes, characterisation and performance evaluations. The review also gives an overview of various types of imidazoline inhibitors with their preparation processes, metal types, corrosive media and concentration range for measurements.

1 Introduction

CO2 corrosion is a serious cause of operational failure in the oil and gas industry where carbon steel is normally used as a pipeline material (Zhang et al. 2017). In a pipeline transport system, dissolved CO2 reacts with H2O to form H2CO3 which then dissociates into H+, HCO3 and CO32−, lowering the solution pH (Geng et al. 2022; Zhang et al. 2007). This process may lead to the collapse of a carbon steel pipeline, resulting in the shutdown of production processes, increased maintenance, loss of production efficiency, product contamination and hence unnecessary expenses with environmental and ecology issues (Guo et al. 2021; Shamsa et al. 2021). In order to reduce the adverse effects of metal corrosion in the petroleum industry, surface coatings and fluid-mediated corrosion inhibitors are widely employed (Shamsa et al. 2022). The use of corrosion inhibitors has proven to be one of the most economically favourable methods to combat corrosion in oil and gas pipelines as it is the most practical and cost effective method (He et al. 2014; Zhang et al. 2015d). It can significantly reduce the corrosion rate even at a low dosage of 10 ppm or sometimes at higher dosage of up to 5000 ppm depending on the operating conditions and interferences (Zhang et al. 2015d). Corrosion mitigation using effective inhibitors is mainly based on modification of the metal surface through adsorption of inhibitor molecules and subsequent formation of a protective film blocking layer that helps slow corrosion in harsh media (Zhang et al. 2007).

Generally, the commonly used corrosion inhibitors for steel pipelines in CO2-rich environments are organic compounds containing heteroatoms (e.g., O, P, S and N atoms), heterocyclic rings, functional groups (–OH, –COOH and NH2) and multiple bonds (Li et al. 2017; Munis et al. 2020; Yoo et al. 2012). Imidazoline and imidazoline derivatives have been reported as an excellent effective organic corrosion inhibitors with several advantages such as low cost, low toxicity, ease of synthesis and high efficacy for corrosion inhibition (Ouakki et al. 2022; Pan et al. 2020), containing nitrogen atoms are well-known as the most effective compounds in this regard since they readily adsorb onto steel surfaces (Solomon et al. 2019b). Imidazoline can be classified as a cationic surfactant. It is also an interesting organic compound for corrosion inhibition as it consists of three main sections, an imidazoline head group, a pendant group and a hydrocarbon tail (Bajpai and Tyagi 2006; Yoo et al. 2012). An imidazoline ring with the presence of a lone pair of electrons in a nitrogen atom facilitates strong bonding of the molecule to steel surfaces (Bajpai and Tyagi 2006; Obike et al. 2018). The pendant group or alkyl amine substituent of the imidazoline ring acts as an anchor that helps maintain its adsorption on steel surfaces (Obike et al. 2018). The tail is a long hydrocarbon chain that can form a hydrophobic film on surfaces (Bajpai and Tyagi 2006; Obike et al. 2018).

The adsorption mechanisms of the imidazoline inhibitors on metal surfaces have been reported. Zheng et al. (2022) simulated adsorption of imidazoline molecules on a steel surface in CO2-saturated water employing covalent bonding between donated electrons in the imidazoline ring and Fe atoms. He et al. (2014) reported that imidazoline molecules could be coordinated with vacant orbitals of metal surfaces. This is done through monolayer chemical adsorption by two N atoms in the imidazoline ring and physisorption through N-onium ions raised from the reaction between an imidazoline molecule and H+ in an acidic aqueous solution. Moreover, Qian and Cheng (2019) showed that a non-polar long hydrocarbon chain formed as a hydrophobic layer on a metal surface to prevent it from exposure to a corrosive environment and thus hindered interaction with water/oxygen. However, the performance of an imidazoline corrosion inhibitor is limited by the number of adsorption sites (Geng et al. 2022) and adsorption capability (Zheng et al. 2022). Chemical modification of an imidazoline molecule is a good way to enhance its inhibition efficacy. Mercaptopropionic acid can be used as a chemical modifying agent that significantly improves inhibition capability. Accordingly, mercapto oleic imidazoline (MOI) accepts electrons from Fe atoms through its sulfhydryl group to form feedback bonds and hence shows outstanding adsorption stability, greater than 95% within 144 h (Zheng et al. 2022). Zhang and co-workers showed that imidazoline derivatives had higher inhibition efficacy than that of imidazoline. They did this by introducing thiourea to affect the kinetic processes of cathodic and anodic reactions and increased reaction activation (Zhang et al. 2007). Guo et al. (2021) introduced a phosphonate (–PH2O3) group into imidazoline derivatives and suggested that this group could improve the adsorption capacity on steel surfaces and also inhibit interference of hydrogen bonds formed by –NH2 groups and H2O molecules. Geng et al. (2022) used resin as a raw material to synthesize an imidazoline derivative leading to an increase in adsorption sites since the resin has excellent surface activity. This imidazoline derivative was able to adsorb spontaneously onto steel surfaces. Similarly, He et al. (2014) compared the corrosion inhibition mechanisms of imidazole (IM) and 2-phenyl-2-imidazoline (2-PI) for an AA5052 surface. It was concluded that the adsorption force of 2-PI is stronger than that of IM because there are two bonds between a 2-PI molecule and a steel surface leading to better inhibition by 2-PI. However, it is also believed that the adsorption performance of corrosion inhibitors on metal surfaces strongly depends on their molecular structures and characteristics (He et al. 2014; Zheng et al. 2022), their concentrations (Solomon et al. 2019a,b), nature and surface charge of metal (Zhang et al. 2017) and electrolyte chemistry (Haladu et al. 2019; Zhang et al. 2017).

In recent years, a considerable amount of research has been done in the development of imidazoline and its derivatives for corrosion inhibition (Munis et al. 2020; Qian and Cheng 2019; Yoo et al. 2012; Zhang et al. 2017; Zheng et al. 2022) in terms of improvement of functionality together with inhibition efficacy. The few review papers that are available focus on their preparation and properties in various applications (Tyagi et al. 2007). An application of imidazoline and imidazoline derivatives as an effective corrosion inhibitor for different metals and alloys has also been reviewed (Alaoui et al. 2020; Hooshmand Zaferani et al. 2013; Mishra et al. 2020; Ouakki et al. 2022; Vinutha and Venkatesha 2016). Vinutha and Venkatesha (2016) published a review that discussed possible inhibition mechanisms for different classes of inhibitors. However, they superficially described the details of imidazoline compounds. Hooshmand Zaferani et al. (2013) focused only on eco-friendly inhibitors covering their properties, mechanisms of inhibition and their efficiencies, but this review lacked deep information on the imidazolines and imidazoline derivatives. In addition, there are several reviews such as Bashir and Shaikh (Usman and Ali 2018), Fazal et al. (Fazal et al. 2022) and Rafida et al. (Jaal et al. 2014) that summarised corrosion inhibition of pipeline steels in CO2-riched aqueous solution but just mentioned briefly about the use of imidazoline as an example of various CO2 corrosion inhibitors. A recent review of Xhanari et al. (2021) discussed the use of organic compounds and natural products as corrosion inhibitors, not only imidazolines. This 2021 review provides insights into the corrosion mechanisms, the influence of steel composition and the influence of various factors on inhibition performance. Aisha and Ime (Al-Moubaraki and Obot 2021) presented various corrosion topics such as physicochemical basics of corrosion, corrosion sources and mitigation methods. However, they focused only on the corrosion-related issues in petroleum refinery operations and were unrelated to pipeline corrosion and imidazoline inhibitor. Although details of imidazoline compounds and their corrosion mechanisms have been reviewed previously (Al-Moubaraki and Obot 2021; Fazal et al. 2022; Jaal et al. 2014; Usman and Ali 2018; Xhanari et al. 2021), all pieces of information have not been gathered and reviewed that have covered from the background knowledge (e.g., basic structures, properties, synthesis processes and basic characterisation) until corrosion inhibition performance of imidazoline in various factors.

In the current review, the most recent studies done on the development of imidazoline and its derivatives for corrosion inhibitors are reviewed as there is a lack of information on this topic. A brief background of this scientific field is presented, followed by chemical structure and properties of imidazoline and its derivatives that extend their applications into corrosion inhibition. Then, this work highlights the chemical reactions and synthesis processes for production of imidazoline and imidazoline derivatives. Molecular level characterisation studies using Fourier transform infrared (FT-IR) and nuclear magnetic resonance (NMR) techniques are discussed. After that, the factors used for performance evaluations of these compounds are thoroughly covered such as metal type (carbon steel), corrosive medium (CO2-saturated solution, HCl, H2SO4, HCl or NH4 solution) and concentration range (0–500 ppm), the evaluation techniques both electrochemical measurements (e.g., OCP, LPR, PDP and EIS) and weigh loss measurements and surface characterisations (adsorption mechanisms, SEM, EDS and contact angle techniques). Finally, a summary and outlook for future directions of this research field are discussed. This article can serve as a reference for researchers to develop more promising corrosion inhibitors with highly cost-effective methods for use in oil and gas applications.

2 Imidazoline and imidazoline derivatives

2.1 Background

Imidazoline or dihydroimidazole (Ji Ram et al. 2019) was first used as a cationic surfactant for dye levelling agents in textile industry in the 1930s (Levinson 1999; Tyagi et al. 2007). In the 1950s, it was first used in fabric softeners as a household product for aqueous dispersion (Tyagi et al. 2007). In the 1960s, it was launched as a household product in Europe and Japan (Tyagi et al. 2007). It is unclear when it was first exploited as a corrosion inhibitor. However, Ferm and Riebsomer (1954) mentioned that imidazoline derivatives were reported to inhibit corrosion against brines, weak inorganic acids, organic acids, carbon dioxide and hydrogen sulfide in 1950–1952. In 1985, an article entitled “The Existence of Imidazoline Corrosion Inhibitors” by Martin and Valone mentioned that the first successful use of hydrocarbon solutions of imidazolines and their salts as corrosion inhibitors was carried out in 1946 at the Richard King lease, Crane County, Texas (Martin and Valone 1985). They indicated that imidazolines were easy-to-use and inexpensive substances that were readily applicable for commercialisation (Martin and Valone 1985). Additionally, they also highlighted that imidazoline salts were more resistant to hydrolysis than the conjugate base (deprotonated) form, enabling an improved steric barrier to corrodent penetration. Thus, it is highly probable that imidazolines were first utilised in the field of corrosion inhibition in the 1940s. Since then, the use of imidazoline inhibitors has been considered a new trend in electromigration inhibitor research. They have been reported as effective organic corrosion inhibitors with several advantages such as low toxicity and high efficacy (Pan et al. 2020).

Recently, several research studies focused on the role of various substituents, namely mercapto, benzene and methyl groups, on promoting the corrosion inhibition of imidazoline molecules. It was reported that these groups have remarkable influence on adsorption of organic molecules on metal surfaces (Solomon et al. 2019a). Guo et al. (2021) suggested that the –PH2O3 functional group could be introduced into the structure of imidazoline derivatives and form coordination compounds that exhibit great adsorption on metal surfaces. Wang et al. (2022) prepared complex inhibitors called SMIF: (E)-1-(2-(1-butyl-2-(hexadec-7-en-1-yl)-1-iodo-2,5-dihydro-1H-1λ5-imidazol-1-yl)ethyl)-3-phenylthiourea and benzimidazole (BMZ) that helped reduce the cost of inhibitors. Migahed et al. (2017) synthesised 1-(2-aminoethyl)-1-dodecyl-2-hexadecyl-4,5-dihydro-1H-imidazol-1-ium (PQI) and 1-(2-aminoethyl)-1-dodecyl-2-heptadecyl-4,5-dihydro-1H-imidazol-1-ium (SQI) with improved inhibition efficacy for carbon steel in oil well formation water. Zhang et al. (2007) prepared an imidazoline derivative by addition of thiourea substituents in its pendant side chain. Obike et al. (2018) reported synthesis and inhibition behaviours of imidazoline derivatives in various structures of their pendant side chains.

2.2 Structure and properties

Generally, imidazoline compounds consist of three parts, an imidazoline head group, a pendant group and a hydrocarbon tail (Bajpai and Tyagi 2006; Yoo et al. 2012) as shown in Figure 1. The head group is the main structure of the imidazoline molecule. It consists of a 5-membered heterocyclic ring (C3N2H4) with two nitrogen atoms as active sites used to adsorb on metal surfaces (Obike et al. 2018). The imidazoline ring can be divided into three different isomeric forms, 2 imidazoline (4,5-dihydroimidazole), 3-imidazoline (2,5-dihydroimidazole) and 4 imidazoline (2,3-dihydroimidazole) (Ferm and Riebsomer 1954; Ji Ram et al. 2019; Sharma and Ghuge 2018; Solomon et al. 2019a), as shown in Figure 2. The 2 imidazoline and 3-imidazoline structures contain an imine centre, whereas 4-imidazoline contains an alkene substructure (Ji Ram et al. 2019). The 2-imidazoline isomer was most commonly found in some natural products and drugs (Liu and Du 2009; Solomon et al. 2019a). The pendant side chain acts as an anchor that can help maintain adsorption on metal surfaces (Obike et al. 2018). The characteristics of pendent group are dependent on the amine types used in imidazoline preparation. The imidazolines were synthesised through a reaction between natural fatty acids and various amines, such as diethylenetriamine (DETA), triethylenetetramine (TETA), tetraethylenepentamine (TEPA) (Muktiarti et al. 2020), ethylenediamine (EDA) and aminoethylethanolamine (AEEA) (Obike et al. 2018). The inhibition efficacy relies strongly on the solubility of imidazoline (Muktiarti et al. 2020) and the presence of polycentric adsorption sites (Obike et al. 2018).

Figure 1: 
						Molecular structure of imidazoline using diethylenetriamine and fatty acid as starting materials.
Figure 1:

Molecular structure of imidazoline using diethylenetriamine and fatty acid as starting materials.

Figure 2: 
						Chemical structure of (a) 2-imidazoline, (b) 3-imidazoline and (c) 4-imidazoline.
Figure 2:

Chemical structure of (a) 2-imidazoline, (b) 3-imidazoline and (c) 4-imidazoline.

The pendant side chain is the main part for structural modification to enhance inhibition. For example, the pendant side of an imidazoline structure was modified using thiourea (Zhang et al. 2007) and sulfhydryl substituents (Zheng et al. 2022). Solomon et al. (2019a) prepared an imidazoline derivative, namely N-(2-(2-tridecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)tetradecanamide (NTETD), which consists of 13 carbon atoms linked to an imidazoline moiety on one of the nitrogen atoms. In addition, the sulfhydryl group could serve as a powerful active site for adsorption and exhibited a great ability to bond with metal surfaces through coordination bonds, leading to an enhanced inhibition (Zhang et al. 2007; Zheng et al. 2022). Moreover, an effective imidazoline-based corrosion inhibitor should have a pendant chain length longer than 12 carbon atoms that is directly attached on the ring N atoms (Solomon et al. 2019a). Alternatively, it could have a relatively small organic radical attached to one of the N atoms of the imidazoline ring. Solomon et al. (Solomon et al. 2019a) found that the partitioning cum inhibiting properties of imidazoline could be improved in corrosive media by introduction of suitable pendant groups with appropriate lengths and anchoring positions.

The part of the long hydrophobic hydrocarbon chain was designed as an initial structure of a tall oil fatty acid chain (Martínez-Palou et al. 2004; Wahyuningrum et al. 2008; Yoo et al. 2012). The influence of length and the number of double bonds in the hydrocarbon tail was studied by Yoo et al. (2012). Biodiesel-based 2-(2-alkyl-4,5-dihydro-1H-imidazol-1-yl) ethanol derivatives (A-DIEs) were prepared from various sources such as Palm BD, Soy-BD and castor oil with a controlled number of double bonds, i.e., C12(0)-DIE, C18(1)-DIE, C18(2)-DIE. Their inhibition efficiency depended strongly on both the degree of unsaturation and the alkyl chain length.

Most commonly used corrosion inhibitors are either based on a long-chain aliphatic diamine or long carbon chain imidazolines. Imidazoline products are solid forms with high thermal stability in organic nitrogenous bases and easily form quaternary salts. Accordingly, the final products of imidazolines with substituents at position 2 are generally present in solid form or a heavy viscous liquid probably caused by hydrogen bonding. N1 substituents are normally liquid with lower solubility in polar solvents (Ji Ram et al. 2019). 2-Alkyl-2-imidazoline is highly soluble in water, alcohol, acetone and chloroform compared to its 1 alkyl-2-imidazoline analogue. In contrast, its solubility is relatively low in benzene, petroleum ether and carbon tetrachloride (Ji Ram et al. 2019).

To improve solubility and other related properties, the imidazoline structure needs to be modified. Several approaches are possible (Guo et al. 2021; Migahed et al. 2017; Wang et al. 2019, 2022; Zhang et al. 2015c; Zhao and Zou 2013). Guo et al. (2021) found that imidazoline derivatives functionalised by a –PH2O3 group rendered the effect of hydrogen bonding between –NH2 and water molecules, which helped to increase the adsorption capacity by complexation of –PH2O3 with iron ions. Zhang et al. (2015c) prepared a bromide-substituted imidazoline (3-IM) with a good inhibition efficacy in a hydrochloric acid environment. Composite inhibitor molecules could be stable on the surfaces of carbon steel with a complex inhibitor, e.g., 0.1 g/L SMIF: (E)-1-(2-(1-butyl-2-(hexadec-7-en-1-yl)-1-iodo-2,5-dihydro-1H-1λ5-imidazol-1-yl)ethyl)-3-phenylthiourea and 0.9 g/L BMZ (Wang et al. 2022). Migahed et al. (2017) modified the structure of 2-(2-trifluoromethyl-4,5-dihydro-imidazol-1-yl)-ethylamine(1-IM) and 2-(2-trichloromethyl 4,5-dihydro-imidazol-1-yl)-ethylamine (2-IM) through a quaternization process using dodecyl chloride. This resulted in formation of the following final products: 1-(2-aminoethyl)-1-dodecyl-2-hexadecyl-4,5-dihydro-1H-[imidazol-1-ium (PQI) and 1-(2-aminoethyl)-1-dodecyl-2-heptadecyl-4,5-dihydro-1H-imidazol-1-ium (SQI). A long-chain fatty acid amide (LFA) could be employed as a corrosion inhibitor that provides a stable adsorption film on steel surfaces via chemical bonds between the heteroatoms (in the acyl, amine and phosphate groups of an LFA molecule) and Fe atoms, yielding multiple adsorption centres (Wang et al. 2019).

2.3 Synthesis processes

In 1888, Hofmann first synthesised 2-methyl-imidazoline by heating N,N′-diacetylethylenediamine in dry HCl (Mehedi and Tepe 2020). Since then, a variety of synthesis methods have been developed to produce imidazolines from various starting materials. The methods can be categorised according to functional groups such as isocyanide, amidine, imine, aziridine, cyanide, amide and others (Mehedi and Tepe 2020) for various prospective applications. However, this review focuses only on the synthesis processes of imidazoline corrosion inhibitors. In general, the optimum conditions used in imidazoline synthesis are relatively simple and involve addition of fatty acids and polyamine compounds under solvent-free conditions (Mehedi and Tepe 2020). The fatty acids or tall oil fatty acids commonly used are stearic acid, palmitic acid, lauric acid, oleic acid and some oils. Examples of the polyamines are ethylenediamine (EDA), diethylenetriamine (DETA), aminoethyl ethanolamine, aminoethylethanolamine (AEEA), triethylenetetramine (TETA) and tetraethylenepentamine (TEPA) (Bajpai and Tyagi 2008). Tall oil fatty acids are considered as low-cost raw materials for the synthesis of imidazoline inhibitors as they are common by-products of coniferous wood in the pulping industry (Demirbas 2008). Thus, the estimated price of imidazoline compounds is approximately 0.02–30 USD/m3 (Zhong et al. 2020).

The synthesis process is a facile method that is controllable by means of modulating reaction temperature and pressure (Bajpai and Tyagi 2008; Martínez-Palou et al. 2003; Tyagi et al. 2007). Formation of imidazoline compounds can be carried out in two steps (Figure 3). First, the carboxylic acid group(s) in the fatty acids undergo an amidation reaction with the amine group(s) in polyamine compounds. Second, an intramolecular cyclisation reaction yields the corresponding imidazoline ring (Tyagi et al. 2007; Zhao et al. 2021). Ditama et al. (2020) prepared imidazolines based on a reaction between methyl oleic fatty acid and N-(2-aminoetil)-3-aminopropiltrimetoksisilan (DAMO) under an inert atmosphere with a reaction time that was over 17 h. Similarly, Wahyuningrum et al. (2008) used a total time of approximately 13 h to prepare imidazoline derivative compounds.

Figure 3: 
						The synthesis pathway for imidazoline, starting from diethylenetriamine and fatty acid.
Figure 3:

The synthesis pathway for imidazoline, starting from diethylenetriamine and fatty acid.

Geng et al. (2022) prepared rosin imidazoline by slowly adding TETA in a mixture of rosin and xylene at 120 °C for 4 h. Then, the temperature was increased to 220 °C for 6 h. Guo et al. (2021) synthesised 2-[2-(7-isopropyl-1,4-dimethyl-9,10-octahydro-phenanthren-1-yl)-4,5-dihydro-1-yl]-ethylamino}-methyl)-phosphonicimidazole (GSIM) via an amidation reaction at 140 °C for 2 h, followed by a cyclization reaction at 220 °C for 2 h. Moreover, Zheng et al. (2022) reported that an imidazoline derivative, mercapto-oleic imidazoline (MOI), was produced from oleic imidazoline (OI) and mercaptopropionic acid under agitation and reflux for 4 h.

All of the above-mentioned processes are conventional synthesis or traditional methods with limitations that include long reaction times, low yields and tedious work-up (Martínez-Palou et al. 2003). Several literature reports proposed a one-step preparation of imidazolines and their derivatives by microwave irradiation (Bajpai and Tyagi 2008; Martínez-Palou et al. 2003; Usyatinsky and Khmelnitsky 2000; Wahyuningrum et al. 2008; Wolkenberg et al. 2004). Martínez-Palou et al. (2003) showed that long chain 2-alkyl-1-(2-hydroxyethyl)-2-imidazolines and their amide precursors can be prepared by both domestic microwave (DMW) and monomode microwave (MMW) ovens. The reactions were fully completed within a short period of time, yielding high purity products (>95%) (Martínez-Palou et al. 2003). Similarly, Bajpai and Tyagi (2008) presented the production of long chain dialkyldiamido imidazolines using DETA and several fatty acids under non-solvent microwave irradiation. It was suggested that this method could provide higher yields and faster reaction times compared to conventional methods. Solomon et al. employed a microwave oven to promote an acid–base reaction for 7 min. This resulted in formation of a novel palmitic imidazoline compound, N-(2-(2-pentadecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)palmitamide (NIMP) (Solomon et al. 2019a), and an imidazoline derivative, N-(2-(2-tridecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)tetradecanamide (NTETD) (Solomon et al. 2019b). In the same way, microwave irradiation is an efficient and quick method for preparation of long chain 2-(2-heptadecyl-4, 5-dihydro-1H-imidazole-1-yl) ethanol (HDIE) (Munis et al. 2020). In some cases, high chemical yields of 91–94% can be obtained within 7–10 min of reaction time (Wahyuningrum et al. 2008). Accordingly, this novel synthesis route is an alternative to the conventional thermal condensation method, especially when using a solvent-free reaction.

It is possible to do an additional step in the chemical modification after imidazoline synthesis. For example, Wang et al. (2022) mixed 1-iodobutane with n-phenylthiourea at 140 °C for 1.5 h after the end process of oleic imidazoline synthesis for production of novel complex inhibitors (SMIF + BMZ). Moreover, two imidazoline derivatives, including SMIF and SMID, were attained from two precursors, n-phenylthiourea and n-methylthiourea (Xiong et al. 2021). Substituents of thiourea on the pendant side of imidazoline ring were also studied by Zhang et al. (2007).

2.4 Characterisation

FT-IR and NMR (1H NMR and 13C NMR) techniques are commonly used to assess chemical structure of imidazoline and imidazoline derivatives. Table 1 summarises the FT-IR analysis of imidazoline and imidazoline derivatives including wavenumbers and their corresponding assignments. The characteristic peaks of imidazoline compounds can be found at approx. 1633–1655 cm−1 corresponding to C=N stretching in the imidazoline ring (Bajpai and Tyagi 2008; Geng et al. 2022; Migahed et al. 2017; Obike et al. 2018; Rahayu et al. 2021; Xiong et al. 2021; Zheng et al. 2022). Also, imidazoline ring C–N stretching is located at around 1550 cm−1 (Kong et al. 2016; Obike et al. 2018; Zheng et al. 2022), and the characteristic peak of N–H stretching is noted in the range between 3275 and 3380 cm−1 (Kousar et al. 2020; Migahed et al. 2017; Munis et al. 2020; Rahayu et al. 2021; Solomon et al. 2019a; Tripathy and Mishra 2017; Zhang et al. 2015d; Zheng et al. 2022). Additionally, other absorption peaks can be found depending on substituents and modified groups.

Table 1:

Major IR absorption bands and possible assignments of imidazoline and imidazoline derivatives along with corresponding references.

Compound Wave number (cm−1) Assignment References
((Z)-2-(2-(Heptadec-8-en-1-yl)-4,5-dihydro-1H-imidazol-1-yl)ethan-1-amine (OMID) 3354, 3280 N–H stretch Kousar et al. (2020)
3005 C–H stretch
1657 C=C stretch
1608 C=N stretch in imidazoline ring
2921, 2851 C–H stretch
1461 C–H scissor
1359, 722 C–H rock
N-(2-(2-Pentadecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)palmitamide (NIMP) 3435 N–H stretch (secondary amine) Solomon et al. (2019a)
2951 C–H stretch
1734 C=O (amide)
1655 C–N in imidazoline ring
719 CH2 skeletal band
N-(2-(2-Tridecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)tetradecanamide (NTETD) 3290 N–H stretch Solomon et al. (2019b)
2956 C–H stretch
1637 C=O stretch
1569 N–H stretch in imidazoline ring
2-(2-Heptadecyl-4,5-dihydro-1H-imidazole-1-yl) ethanol (HDIE) 2914, 2855 C–H stretch Munis et al. (2020)
718 CH2 skeleton of long alkyl chain
1614 C=N stretch in imidazoline ring
1737 C=O stretch
3310 N–H stretch
1460 O–H bending
1604 Imidazoline ring
695 Quaternization of ring
726
758
3381 N–H stretch of amide linkage
2-Methyl-4-phenyl-1-tosyl-4, 5-dihydro-1H-imidazole (IMI) 3275 N–H stretch Zhang et al. (2015d)
2941, 2876 –CH2
1656 C–N stretch in imidazoline ring
1328 C=N stretch in imidazoline ring
3063, 814 C–H stretch in benzene ring
1157 C–SO2–N
Oleic imidazoline (OI) and mercapto-oleic imidazoline (MOI) 1655, 1652 C=N stretch in imidazoline ring Zheng et al. (2022)
1553 C–N bond
2723 –SH group
1367 C=O bond
1135 C=C
2857 CH2
2931 CH3
3298 –NH
Mannich-modified imidazoline (MMI) 1708 HO–C=O Kong et al. (2016)
1555 O=C–NH
1609 C=N
1708 Carbonyl groups
2-Henicos-10-enyl-4,5-dihydro-1H-imidazole (HDI) 720 (CH2)n skeletal Obike et al. (2018)
1550 N=C–N
1607 Imidazoline ring
1650 C=N
2924 –C–H CH2, CH3
3291 N–H stretch
1-(2-Aminoethyl)-1-dodecyl-2-hexadecyl-4,5-dihydro-1H-imidazol-1-ium (PQI) 2851 –C–H Migahed et al. (2017)
2922 CH3
723 (CH2)n skeletal
1744 Ester
3301 Primary amine
1642 C=N
1603 Imidazoline ring
Resin imidazoline (RI) 3288 –N–H Geng et al. (2022)
2929 Asymmetric stretching vibration of –CH2
2866 Symmetric stretching vibration of –CH2
1633 C=N stretch in imidazoline ring
Bio-diesel-based 2-(2-alkyl-4,5-dihydro-1H-imidazol-1-yl)ethanol derivatives (A-DIEs) 3296 Primary amine Yoo et al. (2012)
1640 Amide
1605 C=N stretch in imidazoline ring
1640 Amide group
Imidazoline derivative (SMIF and SMID) 3275 N–H stretch Xiong et al. (2021)
1650 C=N stretch
1650–1640 C=C stretch
1170 C=S stretch
2925, 2842 CH2 stretch
1460 C–C stretch
1250–950 Phenyl stretch
2900 CH3 stretch
2-Heptadecyl-1-[2-(octadecanoylamino)ethyl]-2-imidazoline 2951 CH3 C–H stretch Bajpai and Tyagi (2008)
719 (CH2)16 skeletal
3435 N–H stretch
1734 C=O stretch (amide)
1655 C=N stretch in imidazoline ring
2-Heptadecyl-1-[2-(octadecanoylamino)ethyl]-2-imidazolinium methyl sulphate 2918 CH3 C–H stretch Bajpai and Tyagi (2008)
713 (CH2)16 skeletal
3441 N–H stretch
1734 C=O stretch (amide)
1655 C=N stretch in imidazoline ring
1460 S=O stretch
Oleic-imidazoline 1656 C=N stretch in imidazoline ring Rahayu et al. (2021)
1271 C–N stretch in imidazoline ring
3008 C–H sp2
1643 C=C
723 Cis C–H
2924–2853 C–H stretch sp3
1465–1375 C–H bend sp3
3291 N–H stretch
1547 N–H bend

Structural characterisation of imidazolines and their derivatives should also be carried out using an NMR technique. NMR peaks can be used to identify the chemical structure of imidazolines and their derivatives through 1H NMR and 13C NMR spectra. For example, the 1H NMR spectra of CH2 attached to the nitrogen of an imidazoline ring could be observed at δ values of 2.66, 2.75 and 2.80 ppm (Bajpai and Tyagi 2008). CH2 attached to the carbon of an imidazoline ring is at δ values of 2.36, 2.37 and 2.39 ppm (Bajpai and Tyagi 2008). An equivalent methylene group in the imidazoline ring appears at δ values of 3.591, 3.613 ppm (Migahed et al. 2017). C sp2 from C=C alkene in an unsaturated hydrocarbon chain that can be found at 13C NMR spectra at 130.1, 129.9 ppm. C=N is at 164.6 ppm. C sp3 is bonded to the nitrogen atom in both the imidazoline ring and pendant group at 56.7–52.9 ppm as well as C sp3 in the hydrocarbon chain at 32.0–14.3 ppm (Budiana et al. 2020). Additionally, the NMR spectra of imidazoline derivatives can be varied depending on the preparation process, the structure of the substituent group and the structure of final product. A complete list of both the 1H NMR and 13C NMR spectra found in imidazolines and imidazoline derivatives is presented in Table 2.

Table 2:

The 1H NMR and 13C NMR spectra data of imidazolines and imidazoline derivatives.

Compound 13C NMR (δ values) 1H NMR (δ values) References
2-Heptadecyl-1-[2-(octadecanoylamino)ethyl]-2-imidazoline C2 of imidazoline ring 118.8 ppm CH3 (0.86, 0.87, 0.89 ppm) Bajpai and Tyagi (2008)
Equivalent ring of methylene carbon 49.4 ppm (CH2)n (1.25 ppm)
Equivalent methylene carbon 33.1 ppm CH2 attached to nitrogen of imidazoline ring (2.66, 2.75, 2.80 ppm)
NHCOR 127.32 ppm, (CH2) 29.5 ppm CH2 attached to carbon of imidazoline ring (2.36, 2.37, 2.39 ppm)
CH2 attached to carbon of amide group 1.5, 1.6, 1.7 ppm)
CH2 attached to nitrogen of amide group 2.1 ppm
–CONH– 6.1 ppm
Equivalent ring methylene group (3.34, 3.64 ppm)
2-Heptadecyl-1-[2-(octadecanoylamino)ethyl]-2-imidazolinium methyl sulphate C2 of imidazoline ring 125 ppm CH3 (0.86, 0.87, 0.89 ppm) Bajpai and Tyagi (2008)
Equivalent ring of methylene carbon 45.13 (CH2)n (1.25 ppm)
Equivalent methylene carbon 32.6 ppm CH2 attached to nitrogen of imidazoline ring (2.66, 2.75, 2.80 ppm)
NHCOR 131.2 ppm, (CH2) 29.48 CH2 attached to carbon of imidazoline ring (2.36, 2.37, 2.39 ppm)
CH3SO4 − methylene carbon 48.2 CH2 attached to carbon of amide group 1.5, 1.6, 1.7 ppm)
CH2 attached to nitrogen of amide group 2.1 ppm
–CONH– 6.1 ppm
Equivalent ring methylene group (3.34, 3.64 ppm)
CH3SO4 (7.25 ppm)
1-(2-Aminoethyl)-1-dodecyl-2-heptadecyl-4,5-dihydro-1H-imidazol-1-ium (SQI) CH3 (0.837, 0.834, 0.857 ppm) Migahed et al. (2017)
(CH2)n (1.242 ppm)
CH2 attached to nitrogen of imidazoline ring (2.479, 2.501, 2.503 ppm)
CH2 attached to carbon of imidazoline (2.207, 2.222, 2.234 ppm)
Equivalent methylene group in imidazoline ring (3.591, 3.613 ppm)
1-(2-Aminoethyl)-1-dodecyl-2-hexadecyl-4,5-dihydro-1H-imidazol-1-ium (PQI) CH3 (0.857, 0.844, 0.837 ppm
(CH2)n (1.243 ppm)
CH2 attached to nitrogen of imidazoline ring (2.494, 2.500, 2.506 ppm)
CH2 attached to carbon of imidazoline (2.210, 2.234, 2.258 ppm)
Equivalent methylene group in imidazoline ring (3.586, 3.608 ppm)
H of NH2 group (7.104 ppm)
Oleic imidazoline (OI) S, 2 H, –CH=CH– (5.32 ppm) Zheng et al. (2022)
S, 4 H, =N–CH2–CH2–N– (3.49 ppm)
d, 2 H, –N–CH2– (3.08 ppm)
M, 12 H –CH2–CH2–NH– (2.69–2.47 ppm)
M, 2 H –N=C–CH2 (2.33 ppm)
d, 4 H, –CH2–CH= (2.01 ppm)
S, 2 H –NH2 (1.45 ppm)
M, 24 H –CH2–CH2–CH2– (1.27–1.23 ppm)
t, 3 H, –CH3 (0.86 ppm)
Mercapto-oleic imidazoline (MOI) S, 2 H, –CH=CH– (5.32 ppm)
S, 4 H, =N–CH2–CH2–N– (3.65 ppm)
S, 2 H, –N–CH2– (3.32 ppm)
M, 16 H –CH2–CH2–NH– (2.83–2.63 ppm)
M, 2 H –N=C–CH2– (2.46 ppm)
S, 1 H –NH–C=O (2.16 ppm)
M, 4 H –CH2–CH= (1.99 ppm)
S, 1 H –SH (1.60 ppm)
M, 22 H –CH2–CH2–CH2– (1.35–1.18 ppm)
M, 3 H –CH3 (0.86 ppm)
2-Methyl-4-phenyl-1-tosyl-4, 5-dihydro-1H-imidazole (IMI) d, 2H, benzene-H (7.83 ppm) Zhang et al. (2015d)
d, 2H, benzene-H (7.71 ppm)
M, 3H, benzene-H (7.32 ppm)
d, 2H, benzene-H (7.21 ppm)
dd, 1H, imidazoline-H (5.04 ppm)
ddm, 2H, imidazoline-H (3.29 ppm)
S, 3H, benzene-CH3 (2.04 ppm)
S, 3H, benzene-CH3 (1.61 ppm)
N-(2-(2-Tridecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)tetradecanamide (NTETD) Methylene protons (2.46 ppm) Solomon et al. (2019b)
–CONH groups (2.10 ppm)
N-(2-(2-Pentadecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)palmitamide (NIMP) C2 of imidazoline ring (118.8 ppm) CH3 (0.86, 0.87, 0.89 ppm) Solomon et al. (2019a)
Equivalent ring methyl carbon (49.4 ppm) (CH2)n (1.25 ppm)
Equivalent methylene carbon (33.1 ppm) CH2 attached to imidazoline ring (2.66, 2.75, 2.80 ppm)
CH2 (29.5 ppm) –CONH– (6.1 ppm)
Equivalent ring methylene group (3.34, 3.64 ppm)
((Z)-2-(2-(Heptadec-8-en-1-yl)-4,5-dihydro-1H-imidazol-1-yl)ethan-1-amine (OMID) Aliphatic tail (129.2 ppm) Aliphatic tail (5.16 ppm) Kousar et al. (2020)
Equivalent in imidazoline ring (167.4 ppm) –CH2–CH2–(C(3)–C(4)) fragment of ring (3.08, 3.5 ppm)
Oleic imidazoline C=N (164.6 ppm) 4H, m imidazoline ring (3.62–3.26 ppm) Budiana et al. (2020)
C sp2 from C=C alkene in unsaturated hydrocarbon chain (130.1, 129.9 ppm) 2H, t, 7 Hz hydrocarbon chain, H–C–C=N (2.16 ppm)
C sp2 region (130.4–128.0 ppm) 2H, q, 7 Hz unsaturated hydrocarbon chain for proton attached on C=C alkene (5.33 ppm)
C sp3 bonded to the nitrogen atom in both imidazoline ring and pendant group (56.7–52.9 ppm) 4H, q, 7 Hz C=C alkene (2.01 ppm)
C sp3 in the hydrocarbon chain (32.0–14.3 ppm) 2H, qi, 9.5 Hz (1.59 ppm)
20H, m (1.37–1.24 ppm)
3H, t, 7 Hz (0.87 ppm)
8H, m –CH2 group (2.92–2.42)
3H, br s (4.23 ppm)
Resin imidazoline (RI) S, 1H, C=CH–C (5.73 ppm) Geng et al. (2022)
S, 1H, C=CH–C (5.31 ppm)
dt, J = 30.9, 9.8 Hz, 1H, –N–CH2–C (4.90 ppm)
S, 1H, –CH2–N (3.76 ppm)
M, 5H, –CH2–N= (3.73–3.59 ppm)
M, 7H, –CH2–N (3.43–3.22 ppm)
M, 10H, –CH2– (2.92–2.65 ppm)
M, 7H, –CH2– (2.54–2.38 ppm)
p, J = 7.0 Hz, 2H, –CH (2.20 ppm)
S, 2H, –CH2– (2.06 ppm)
M, 3H, –CH2– (2.02–1.93 ppm)
S, 2H, –CH (1.91 ppm)
d, J = 12.5 Hz 3H, –CH (1.86 ppm)
M, Hz, 8H, –CH2– (1.53–1.35 ppm)
M, 17H, –CH2−/−NH1– (1.24–1.09 ppm)
M, 8H, –CH3/–CH2– (1.05–0.95 ppm)
M, 9H, –CH3 (0.95–0.73 ppm)
Bio-diesel-based 2-(2-alkyl-4,5-dihydro-1H-imidazol-1-yl)ethanol derivatives (A-DIEs) t, 2H, –C=N–C–CH2 –N– (3.23 ppm) Yoo et al. (2012)
t, 2H, –CH2–C–OH (3.32 ppm)
t, 2H, –CH2–OH (3.67 ppm)
t, 2H, –C=N–CH2– (3.71 ppm)
Br, 1H, –OH (2.91 ppm)
t, 2H, O=CCH2– methylene right next to carbonyl group (2.30 ppm)
Imidazoline derivative (SMIF) –N– C=S (129 ppm) –C=N in inidazoline (3.25 ppm) Xiong et al. (2021)
1-Benzene (128.4 ppm) 1-Benzene (7.70 ppm)
1 beta –C(=S)–N (42.1 ppm) 1–N C=S (7.43 ppm)
–N=C (32.9 ppm) –C=C (2.16 ppm)
Imidazoline derivative (SMID) 1-Ethylene (130.6 ppm) Aldimine (7.5 ppm)
–C=N (28.0 ppm) –NC(=S) (3.55 ppm)
1 beta –C(=S)–N (15.2 ppm) –C=N (3.25 ppm)
–C=C (2.16 ppm)
Methylene (1.25 ppm)
Oleic-imidazoline C=N (164.6 ppm) 4H, m imidazoline ring (3.62–3.26 ppm) Rahayu et al. (2021)
C sp2 from C=C alkene in unsaturated hydrocarbon chain (130.1, 129.9 ppm) 2H, t 7 Hz hydrocarbon chain H–C–C=N (2.16 ppm)
C sp2 region (130.4–128.0 ppm) 2H, q, 7 Hz unsaturated hydrocarbon chain on C=C alkene (5.33 ppm)
C sp3 bonded to the nitrogen atom in both imidazoline ring and pendant group (56.7–52.9 ppm) 4H, q, 7 Hz C=C alkene (2.01 ppm)
C sp3 in the hydrocarbon chain (32.0–14.3 ppm) 2H, qi, 9.5 Hz (1.59 ppm)
20H, m (1.37–1.24 ppm)
3H, t, 7 Hz hydrocarbon chain (0.87 ppm)
8H, m –CH2 group (2.92–2.42)
3H, br s (4.23 ppm)

3 Evaluation of corrosion inhibition efficiency

Suitable test procedures and conditions should be clearly defined to evaluate the corrosion inhibition efficacy of imidazolines and their derivatives. These parameters may include metal type, corrosive medium, inhibitor concentration, evaluation value and other factors (e.g., test duration and temperature).

Conventional characterisation methods usually rely on electrochemical parameters (e.g., open circuit potential (OCP), linear polarization resistance (LPR), potentiodynamic polarization (PDP), electrochemical impedance spectroscopy (EIS)), weight loss measurements, adsorption studies and corrosion inhibition mechanisms. Surface characterisation through SEM, EDS and contact angle techniques is also employed to investigate corrosion processes and to explain the possible inhibition mechanism(s) of imidazolines and imidazoline derivatives.

3.1 Conditions of corrosion inhibition evaluation

3.1.1 Metal type electrode materials in electrochemical measurements

Carbon steel is one of the principal materials used in oil and gas production. It is also used in evaluation of corrosion inhibition efficacy via both electrochemical and weight loss measurements (Geng et al. 2022; Qian and Cheng 2019; Umoren 2016; Vinutha and Venkatesha 2016; Wahyuningrum et al. 2008; Zhang et al. 2007, 2015b,d). Apart from carbon steel, there are also other metal types that can be tested, as given in Table 3. These metals have different chemical compositions and morphologies and are sometimes studied when subjected to specific working conditions and applications (Xhanari et al. 2021).

Table 3:

The designation and chemical composition (in wt%) of steel samples used in an evaluation of corrosion inhibition efficiencies examining imidazolines and imidazoline derivatives.

Metal Chemical composition (wt%) References
AA5052 0.25 Si, 0.10 Cu, 2.2–2.8 Mg, 0.10 Zn, 0.10 Mn, 0.15–0.35 Cr, 0.40 Fe and aluminium balanced He et al. (2014)
X52 steel 0.24 C, 1.4 Mn, 0.45 Si, 0.025 P, 0.015 S, 0.10 V, 0.05 Nb, 0.04 Ti and Fe balance Qian and Cheng (2019)
Mild steel 0.17 C, 0.37 Mn, 0.20 Si, 0.03 S, 0.01 P and balance Fe Zhang et al. (2015b)
Q345 steel 0.19 C, 0.025 P, 0.015 S, 0.35 Si, 1.45 Mn, 0.07 Nb, 0.15 V, 0.20 Ti, 0.30 Cr, 0.012 Ni, 0.10 Mo and balance Fe Guo et al. (2021)
Q235 steel 0.18 C, 0.02 Si, 0.45 Mn, 0.02 S, 0.01 P and balance Fe Zhang et al. (2012)
P110 steel 0.26 C, 0.19 Si, 1.37 Mn, 0.004 P, 0.004 S, 0.148 Cr, 0.028 Ni, 0.019 Cu, 0.013 Mo, 0.006 V, 0.062 Al and balanced Fe Zhang et al. (2015d)
N80 steel 0.31 C, 0.19 Si, 0.92 Mn, 0.010 P, 0.008 S, 0.2 Cr and balance Fe Solomon et al. (2019a)
Carbon steel 0.65 Mn, 0.45 C, 0.20 Si, 0.17 Ni, 0.03 S, 0.15 Cr, 0.02 P and balance Fe Zheng et al. (2022)
L360 steel 0.06 V, 0.05 Cr, 1.09 Mn, 0.00 Ni, 0.10 Cu, 0.09 Mo, 0.00 W and 96.55 Fe Obike et al. (2018)
API X65 0.04 C, 0.2 Si, 1.5 Mn, 0.011 P, 0.003 S, 0.02 Mo and balance Fe Zhang et al. (2007)
1018 steel 0.18 C, 0.35 Mn, 0.17 Si, 0.025 S, 0.03 P and balance Fe Martínez-Palou et al. (2004)

Zhang et al. (2018) studied the effect of microstructure on the inhibition performance of an imidazoline-based inhibitor. Two different types of microstructure were selected for this study including H steel (coarse laminar pearlite) and T steel (globular and shot rod shaped pearlite) prepared under different heat-treatment conditions. The results indicated that less localised corrosion took place on T steel than that on H steel. This is because T steel has a homogeneous distribution of pearlite with a low density of dislocations. The imidazoline molecules were able to adsorb more uniformly onto T steel than H steel. A more compact inhibitor film was formed in a shorter time. Furthermore, the corrosion inhibition of other metal types such as X70 and Q235 steels was also done (Zhang et al. 2017). It was reported that Q235 steel exhibited higher inhibition and better adsorption capability due to larger amounts of uniformly distributed carbides than with X70 steel. Consequently, the type of carbon steel used in a corrosion test should be clearly specified. In addition, imidazoline and its derivatives have been used extensively as a corrosion inhibitor with other metallic materials such as iron (%IE = 96%) (Bereket et al. 2002), copper (%IE = 95%) (Hou et al. 2022), zinc (%IE = 87%) (Stupnisek et al. 1995) and their alloys (%IE = >90%) (El-Katori et al. 2022).

3.1.2 Corrosive medium

The corrosive medium is one of significant parameters that strongly affects the interaction between inhibitor molecules and metal surfaces. It helps to define possible inhibition mechanism(s) of corrosion inhibitors. Most related studies on corrosion testing using imidazoline inhibitors employed either (i) a CO2-saturated solution with various concentrations of NaCl or (ii) other corrosion systems containing HCl, H2SO4, HCl or NH4 at various concentrations (see Table 4). Basically, the imidazoline compounds containing hydrophobic alkyl chain are oil soluble and have low solubility in water phase (Abd El-Lateef et al. 2012; McMahon 1991). To promote their solubility in 3 wt% NaCl solution saturated with CO2, Zhang et al. (2015a) dissolved imidazoline in alcohol with the ratio of 1:9 (wt%, the former was inhibitor) before testing their corrosion efficiencies. Moreover, imidazoline derivatives consisting of polar substituents such as –OH, –NH2, –COOH, –SO3H, –CN and –NO2 increase their solubility in corrosive media. Shamsa et al. (2022) suggested that the fully protonated inhibitor increased inhibitor solubility/dispersibility and also enhanced molecular adsorption of the positively charged imidazoline on the metal surface, thereby increasing inhibition performance (Shamsa et al. 2022).

Table 4:

The corrosion inhibition effectiveness of imidazoline and imidazoline derivatives.

Inhibitor Preparation process Metal type Corrosive medium Concentration range Efficiency References
Oleic imidazoline (OI) and mercapto-oleic imidazoline (MOI) Conventional Carbon steel CO2-saturated formation water 0–100 ppm Weight loss method; 86.26% (100 ppm of OI) 96.79% (100 ppm of MOI) Zheng et al. (2022)
Resin imidazoline (RI) + thiourea (TU) Conventional Carbon steel CO2-saturated solution 0–400 mg/L RI + 0–80 mg/L TU Weight loss method; 89.8% (200 mg/L RI + 10 mg/L TU) Geng et al. (2022)
Imidazoline derivative thiourea Conventional API X65 steel CO2-saturated 5% NaCl solution 0–200 ppm PDP method; 98.95% (200 ppm) Zhang et al. (2007)
Imidazole (IM) and 2-phenyl-2-imidazoline (2-PI) Commercial (Sigma-Aldrich) AA5052 1.0 M HCl solution 0–14.0 mM (IM), 0–6.5 mM (2-PI) Weight loss method; 80.1% (14.0 mM of IM) 86.5% (6.5 mM of 2-PI) He et al. (2014)
Imidazoline (IM) and sodium dodecylbenzenesulphonate (SDBS) Commercial (Fisher Scientific) X52 steel 1 wt% NaCl solution saturated with CO2 gas 0–150 mg/L IM + 0–150 mg/L SDBS Weight loss method; 90.1% (150 mg/L IM + 50 mg/L SDBS) Qian and Cheng (2019)
Bio-diesel–based 2-(2-alkyl-4,5-dihydro-1H-imidazol-1-yl)ethanol derivatives (A-DIEs): C12(0)-DIE, C18(1)-DIE, C18(2)-DIE, soy-DIE, palm-DIE and castor-DIE Conventional Mild steel 1 M HCl 0–500 ppm PDP method; 99.3% (500 ppm C18(2)-DIE) Yoo et al. (2012)
2-(2-Heptadecyl-4,5-dihydro-1H-imidazole-1-yl) ethanol (HDIE) Microwave Q-235 7.5% NH4 solution 0–1.0 mM PDP method; 90.9% (0.5 mM) Munis et al. (2020)
2-[2-(7-Isopropyl-1,4a-dimethyl-1,2,3,4,4a,9, 10,10a-octahydro-phenanthren-1-yl)-4,5-dihydro-imidazol-1-yl]-ethylamineimidazole (SIM) Conventional Q345 steel 1 M HCl 0–500 ppm PDP method; 90.22% (500 ppm of SIM) Guo et al. (2021)
({2-[2-(7-Isopropyl-1,4-dimethyl-9,10-octahydro-phenanthren-1-yl)-4,5-dihydro-1-yl]-ethylamino}-methyl)-phosphonicimidazole (GSIM) Conventional Q345 steel 1 M HCl 0–500 ppm PDP method; 94.02% (500 ppm of GSIM) Guo et al. (2021)
((Z)-2-(2-(Heptadec-8-en-1-yl)-4,5-dihydro-1H-imidazol-1-yl)ethan-1-amine (OMID) Conventional Carbon steel 1 M HCl or 1 M H2SO4 0.18 mM in 1 M HCl or 0.22 mM in 1 M H2SO4 LPR method; 97% (1 M HCl), 99% (1 M H2SO4) Kousar et al. (2020)
[2-(2-Henicos-10-enyl-4,5-dihydro-imidazol-1-yl)-ethyl]-methylamine (HDM), 2-(2-henicos-10-enyl-4,5-dihydro-imidazol-1-yl)-ethanol (HDE) and 2-henicos-10-enyl-4,5-dihydro-1H-imidazole (HDI) Microwave L360 mild steel CO2-saturated 3.5% NaCl solution 0–300 ppm Weight loss method; 71.8% (100 ppm of HDM), 60.2% (100 ppm of HDE), 65.5% (100 ppm of HDI) Obike et al. (2018)
1-(2-Aminoethyl)-2-heptadecyl imidazoline (IM, CAS: 3010-23-9) Commercial (Wuhan Chuboss Technology Corporation) N80 carbon steel CO2-saturated 1.65 wt% NaCl solution

Containing 1 g/L or 3 g/L acetic acid (HAc)
0.1 g/L PDP method; 99.3% (1 g/L HAc), 99.4% (3 g/L HAc) Li et al. (2017)
N-(2-(2-Tridecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)tetradecanamide (NTETD) Microwave St37-2 steel 15% HCl 0–400 ppm Weight loss method; 96.51% (300 ppm) Solomon et al. (2019b)
2-(2-Trifluoromethyl-4,5-dihydro-imidazol-1-yl)-ethylamine (1-IM) and 2-(2-trichloromethyl-4,5-dihydro-imidazol-1-yl)-ethylamine (2-IM) Conventional Mild steel 0.5 M HCl 0–2 mM Weight loss method; 96% (2 mM of 1-IM), 98% (1.5 or 2 mM of 2-IM) Zhang et al. (2015b)
2-Methyl-4-phenyl-1-tosyl-4, 5-dihydro-1H-imidazole (IMI) Conventional P110 carbon steel 1 M HCl 0–300 mg/L Weight loss method; 94.6% (300 mg/L) Zhang et al. (2015d)
1-(2-Aminoethyl)-1-dodecyl-2-hexadecyl-4,5-dihydro-1H-imidazol-1-ium (PQI) and 1-(2-aminoethyl)-1-dodecyl-2-heptadecyl-4,5-dihydro-1H-imidazol-1-ium (SQI) Conventional X-65 carbon steel Oil wells formation water 0–500 ppm PDP method; 93.3% (500 ppm of PQI), 94.68% (500 ppm of SQI) Migahed et al. (2017)
Thioureido imidazoline inhibitor (TAI) Conventional Q235 steel CO2-saturated 2% NaCl solution 0–0.31 mmol/dm3 Weight loss method; 92.2% (0.15 mmol/dm3) Wang et al. (2011)
Mixture of 2-undecyl-N-carboxymethyl-N-hydroxyethyl imidazoline (UHCI) with potassium iodide (KI) Commercial (Shanghai Fakai Chemical Industry Co. Ltd) Carbon steel 8% amidosulfuric acid solution N/A Weight loss method; 97.48% UHKI:KI = 9:1 Mu et al. (2017)
N-(2-(2-Pentadecyl-4,5-dihydro-1H-imidazol-1-yl)ethyl)palmitamide (NIMP) Microwave N80 steel 15% HCl 0–500 ppm Weight loss method; 97.92% (300 ppm) Solomon et al. (2019a)
2-(2-Tribromomethyl-4,5-dihydro-imidazol-1-yl)-ethylamine (3-IM) Conventional Mild steel 0.5 M HCl 0–2 mM Weight loss method; 85% (2 mM) Zhang et al. (2015c)
Oleic-imidazoline Conventional Carbon steel 1% NaCl solution 0–500 ppm Weight loss method; 83.33% (500 ppm) Rahayu et al. (2021)
Imidazoline Commercial X70 and Q235 steel CO2-saturated 3 wt% NaCl 0–4.878 × 10−5 mol/L PDP method; 97.8% (X70), 98.6% (Q235) Zhang et al. (2017)
2-Mercaptobenzimidazole (MBI) Commercial Carbon steel 0.01, 0.1 and 1 M HCl 0–2 mM PDP method; >90% Morales-Gil et al. (2015)
Silane functional group imidazoline compound (DAMO imidazoline) Conventional Mild steel 5% NaCl 0–0.44 wt% PDP method; 96% (0.28 wt%) Ditama et al. (2020)
Compound 1a and 1b: (Z)-2-(2-(heptadec-8-enyl)-4,5-dihydroimidazol-1-yl)ethanamine

Compound 2a and 2b: (Z)-2-(2-(heptadec-8-enyl)-4,5-dihydroimidazol-1-yl)ethanol

Compound 3a and 3b: 2-(2-heptadecyl-4,5-dihydroimidazol-1-yl)ethanamine
(a) Conventional and (b) microwave Carbon steel 1% NaCl solution 0–120 ppm LPR method; 32.18% (8 ppm of 1b), 39.59%

(8 ppm of 2b) and 12.73% (8 ppm of 3b)
Wahyuningrum et al. (2008)
Imidazoline phosphate quaternary ammonium salt Conventional Q235 steel CO2-saturated 2 wt% NaCl 0–1000 mg/L Weight loss method; 90.9% (1000 mg/L) Wang et al. (2019)
Imidazoline amides (IM) Conventional Armco iron 3.0 mol/L NaCl solution saturated with 0.1 MPa CO2 0–20 ppm PDP method; 97% (20 ppm) Zhang et al. (2001)
Thioureido imidazoline Conventional X70 steel 0.2 mol/L NaCl solution with 0.05 mol/L NaNO2, adding HCl solution ∼ pH 5.5 0–240 ppm N/A Zuo et al. (2017)
Imidazoline Commercial Carbon steel (H steel: coarse laminar pearlite; T steel: globular and shot rod shaped pearlite) CO2-saturated 3 wt% NaCl solution 0–4.88 × 10−5 mol/L PDP method; 98.0% (4.88 × 10−5 mol/L, H), 98.6%

(4.88 × 10−5 mol/L, T)
Zhang et al. (2018)
Imidazoline derivative (SMIF) + benzimidazole (BMZ) Conventional Q235 steel 3.5 wt% simulated concrete solution SMIF 0–1.0 g/L + BMZ 0–1.0 g/L PDP method; 90.22% (0.1 g/L SMIF and 0.9 g/L BMZ) Wang et al. (2022)
Imidazoline Commercial X65 carbon steel CO2-saturated 3 wt% NaCl brine 0–30 ppm LPR method; 73% (30 ppm) Shamsa et al. (2022)
Imidazoline Commercial (M-I SWACO) X65 carbon steel CO2-saturated 3 wt% NaCl brine 30 ppm LPR method; 96% (30 ppm) Shamsa et al. (2021)
((Z)-2-(2-(Heptadec-8-en-1-yl)-4,5-dihydro-1H-imidazol-1-yl)ethan-1-amine (OMID) Conventional Carbon steel 1 M HCl 0–0.1 mM PDP method; 95% (0.01 mM) Kousar et al. (2021)
Imidazoline (OIM) and sodium benzoate (SB) Conventional Mild steel CO2-saturated brine solution OIM 10 mg/L + SB 10 mg/L PDP method; 91.45% Zhao and Chen (2012)
Imidazoline derivative (SMIF and SMID) Conventional Q235 carbon steel Synthetic concrete pore solution (SCP solution): 0.6 mol/L potassium hydroxide + 0.2 mol/L sodium hydroxide + saturated calcium hydroxide solution with 3.5 wt% NaCl 0–4.0 g/L PDP method; 95.92% (4.0 g/L of SMIF), 89.80% (4.0 g/L of SMID) Xiong et al. (2021)
Imidazoline Conventional X65 carbon steel CO2-saturated 3 wt% NaCl solution 0–50 ppm LPR method; >90% (30 ppm) Shamsa et al. (2020)
Imidazoline derivative Conventional Mild steels Natrium chloride (5 wt% NaCl) N/A PDP method; 90.43% Muktiarti et al. (2020)

Kousar et al. (2020) measured critical micelle concentration (CMC) when using an OMID inhibitor in two different corrosive media, 1 M HCl and 1 M H2SO4. The different media systems had negligible impact on the CMC concentration observed, for example, 0.18 ± 0.03 mM and 0.22 ± 0.05 mM in the respective cases of 1M HCl and 1M H2SO4. It is established that a minimal corrosion rate can be achieved at the CMC. An increase in inhibitor concentration could only result in the formation of more micelles in the bulk solution rather than having them increasingly adsorb on the metal surface (Fuchs-Godec 2006; Moradighadi et al. 2021). So, these CMC values can be used to estimate concentration of inhibitors for evaluating corrosion inhibition performance in both acids. Haladu et al. reported the performance of a tetrapolymer in 15 wt% HCl and 15 wt% H2SO4 acid media (Haladu et al. 2019). The as-prepared inhibitor functioned as a moderate corrosion inhibitor in both acid media, but the medium containing HCl offered superior performance. This is because in HCl, the specific adsorption of free chloride ions is on the anodic site or on the metal surface while sulphate ions are adsorbed in the H2SO4 medium. The chloride ions have a higher shielding power than the sulphate ions so more protonated inhibitors are expected to be adsorbed on steel surfaces recharged by chloride ions. He et al. (2014) proposed an adsorption mechanism of imidazoline molecules on AA5052 surfaces in a 1.0 M HCl solution. The imidazoline molecules can react with H+ in an acid medium to form N-onium ions. Thus, imidazoline molecules can form as neutral molecules and N-onium ions in an aqueous acid solution. In neutral and acidic media, imidazoline molecules consist of two N atoms in the imidazoline ring, which has free electron pairs that can coordinate with the vacant orbitals of a steel surface, causing the formation of a chemical adsorption monolayer. The N-onium ions in the imidazoline ring are prone to undergo physisorption on steel surfaces, whereas the free N atom, which is not protonated, can generally form a chemical adsorption monolayer. Additionally, acid concentration in the corrosive media has an influence on the behaviour of inhibitors. Li et al. (2017) reported inhibition efficacy using a commercial imidazoline (1-(2-aminoethyl)-2-heptadecyl imidazoline) in preventing crevice corrosion of N80 carbon steel in a CO2-saturated NaCl solution containing various concentrations of acetic acid (HAc). The results showed that imidazoline could act as a mixed-type inhibitor with a predominantly anodic effectiveness at 1 g/L HAc and predominantly cathodic effectiveness at 3 g/L HAc. In the same way, the corrosion inhibition efficacy of 2-mercaptobenzimidazole (MBI) was evaluated at various HCl concentrations, 1 M, 0.1 M and 0.01 M (Morales-Gil et al. 2015). In the higher acid concentrations, 0.1 M and 1 M HCl, the MBI became a mixed-type inhibitor while evolving as a predominantly anodic form inhibitor at the lowest HCl concentration, 0.01 M.

3.1.3 Concentration of inhibitors (dosage)

One important factor in corrosion evaluation is the concentration of inhibitors or dosage. In general, the inhibition efficacy increases with the dosage (He et al. 2014; Qian and Cheng 2019; Yoo et al. 2012; Zhang et al. 2015b,c), since at higher concentrations, greater numbers of adsorption sites on steel surfaces are occupied by inhibitor molecules. Hence, there are fewer active sites available for corrosion reactions (Wang et al. 2019). The effect of blocking the active sites on the metal surface by adsorbed imidazoline molecules is the most effective way to decrease corrosion rates. Yoo et al. (2012) found that the partial dielectric constant and the double-layer capacitance of the system decreased with increasing inhibitor concentrations. Moreover, formation of a very dense inhibitor film on metal surfaces reduces the number of the kinetic active sites where anodic and cathodic reactions take place, thus lowering corrosion rates (Monticelli 2018; Ouakki et al. 2019).

In contrast, other researchers showed that inhibition efficacy was not improved by increasing inhibitor concentration. Geng et al. (2022) suggested that the number of adsorption sites increased with the inhibitor concentration to some extent and was limited when the concentration of inhibitors reached their critical CMC. Solomon et al. (2019a) investigated the inhibition efficacy of NIMP molecules and observed significant improvement with NIMP concentration, but not over 300 ppm of dosage. The saturation point was reached at a concentration greater than 300 ppm. Hence, the adsorbed inhibitor species began to interact with the unabsorbed species resulting in the desorption of the inhibitor film. Similarly, 300 ppm of NTETD inhibitor showed maximum inhibition of >93% as a steel surface was fully covered by inhibitor molecules. However, the desorption of pre-adsorbed species occurred at higher concentrations and thus adversely affected inhibition (Solomon et al. 2019b). It was possible that the NTETD concentration was optimised at 300 ppm or even lower in other cases. For instance, Zheng et al. (2022) reported excellent inhibition efficiency of 95% at a very low concentration of 20 ppm when the molecular structure of imidazoline was modified with mercaptopropionic acid. Munis et al. (2020) indicated that the optimum concentration of HDIE inhibitor was 0.5 mM/l. When the inhibitor dosage exceeds its optimum concentration, interaction between the pre-absorbed molecules and free inhibitor molecules reaches a point that gives rise to desorption of the pre-adsorbed inhibitor molecules. Moreover, the use of an inhibitor at a minimal level is beneficial as it reduces operational costs. All available data on inhibitor concentration with their corresponding performance are given in Table 4.

3.1.4 Evaluation value

Most literature reported that the corrosion inhibition performance of imidazolines and imidazoline derivatives are in terms of inhibition efficiency (%IE or %η). The inhibition efficiency value is compared to a system with no inhibitor under the same test conditions.

3.1.5 Other factors

Test duration also affects the corrosion inhibition. Wang et al. (2019) published inhibition efficiency values of LFA molecules measured via EIS measurements. They reached a maximum value of 98.6% at 168 h of immersion time. After 600 h, inhibition efficiencies could be maintained at above 90%, indicating that LFA was an effective corrosion inhibitor for field applications with long action duration.

A few relevant studies also reported the influence of temperature on corrosion inhibition. The corrosion rates increased from chemical and electrochemical reactions with temperature due to acceleration of transfer process of reactive (and corrosive) ions from the bulk solution to steel surfaces (Desimone et al. 2011; Wang et al. 2019; Zeng et al. 2018) and also an increased reaction rate coefficient (Zhang et al. 2007). Increased temperature alters the adsorption–desorption equilibrium of pre-adsorbed inhibitor molecules. This causes a decrease in surface coverage and, therefore, inhibitor action (Obike et al. 2018; Popova et al. 2007; Wang et al. 2019). Zhang et al. (2015c) found that anodic dissolution of iron was accelerated with increasing temperature while the adsorption of 2-(2-tribromomethyl-4,5-dihydro-imidazol-1-yl)-ethylamine molecules was hindered and competed with the cathodic reaction to yield a mass of hydrogen.

The evaluation of anticorrosive properties of imidazolines and imidazoline derivatives can be done using both chemical (weight loss) and electrochemical (e.g., OCP, LPR, PDP, and EIS) measurements. Each analytical technique needs unique test procedures and conditions as detailed below.

3.2 Electrochemical measurements: OCP studies

OCP monitoring during corrosion testing is a useful tool to study corrosion resistance behaviour. The evolution of OCP curves as a function of immersion time can be used to determine the effect of a corrosive medium and inhibitor addition on steel surfaces. Zhang et al. (2015c) and Umoren (2016) suggested that an immersion time of 1 h was suitable for the electrochemical experiments since the Ecorr curves reached a stable state in this timeframe.

In an inhibitor free system, the corrosion potential (Ecorr) moves toward a more negative state and potentially has the lowest values due to the breakdown of the oxide film on carbon steel surfaces (Li et al. 2021; Migahed et al. 2017; Xiong et al. 2021). Munis et al. (Munis et al. 2020; Wang et al. 2022) proposed that the negative value of Ecorr curves was the corrosive nature of the blank solution because of the absorption of negative chloride ions on the working electrode (WE) in a NH4Cl medium. In an efficiently inhibited system, Ecorr should move to more positive values (Li et al. 2021; Migahed et al. 2017; Umoren 2016; Wang et al. 2022) as shown in Figure 4. The difference in OCP values in the absence and presence of a particular inhibitor indicates that the carbon surface was covered and protected by the inhibitor film (Umoren 2016). Additionally, the Ecorr shift in the inhibited system could also be used to predict inhibition type, e.g., anodic or cathodic, if the difference in displacement of Ecorr with and without inhibitor is greater than 85 mV (Solomon et al. 2019b; Umoren 2016).

Figure 4: 
						OCP curves of corrosion inhibitors at different concentrations of the imidazoline derivative and benzimidazole (polypropylene glycol, PPG) (Umoren 2016). Reprinted with permission from Elsevier.
Figure 4:

OCP curves of corrosion inhibitors at different concentrations of the imidazoline derivative and benzimidazole (polypropylene glycol, PPG) (Umoren 2016). Reprinted with permission from Elsevier.

3.3 PDP and LPR studies

PDP and LPR measurements can be used to gain insights into the kinetics of anodic (metal dissolution) and cathodic reactions (hydrogen evolution) (Alhaffar et al. 2018). PDP studies are some of the most commonly used DC electrochemical methods in corrosion measurements (Telegdi et al. 2018). In PDP, a wide range of voltage is applied to a test electrode and the current response is monitored and recorded according to the oxidation or reduction reactions on electrode surface. As shown in Figure 5, the data presentation is called a polarisation curve, which is a plot of the applied potential and current density (I or log I) (Telegdi et al. 2018, Zhang et al. 2015d). This curve can be used to determine the values of various parameters such as Ecorr, anodic and cathodic Tafel slopes (ba, bc), corrosion current density (Icorr) and inhibition efficiency (η%) (Pavithra et al. 2010; Wang et al. 2019). The Icorr value is usually reduced when imidazolines or imidazoline derivatives are added. Alteration of the peak position and Tafel slope should also be determined. With an inhibitor in the system, the polarisation curve is remarkably shifted to a more negative potential in comparison to the blank, which verifies its inhibition efficacy in rendering anodic and cathodic reactions (Zhang et al. 2012, 2007). In some cases, the shift in potential is more pronounced on the cathodic branch than the anodic branch. Here, the polarisation curve indicates that a particular inhibitor is predominantly under cathodic control (Solomon et al. 2019b).

Figure 5: 
						Potentiodynamic polarisation curves of pipeline steel in 15% HCl solution with the absence and presence of an inhibitor with various concentrations of polypropylene glycol (PPG) additives (Umoren 2016). Reprinted with permission from Elsevier.
Figure 5:

Potentiodynamic polarisation curves of pipeline steel in 15% HCl solution with the absence and presence of an inhibitor with various concentrations of polypropylene glycol (PPG) additives (Umoren 2016). Reprinted with permission from Elsevier.

The shape of Tafel curve is associated with ba and bc values. In the inhibited system, the shape of PDP curve significantly changes, implying that corrosion processes on steel surfaces are greatly affected by a given inhibitor (Zhang et al. 2015d). The change of ba and bc values can provide information on the corrosion inhibition mechanism(s) of imidazolines or imidazoline derivatives in protecting carbon steel surfaces (Munis et al. 2020; Zhang et al. 2015b). For example, the difference of ba values obtained from the inhibited and uninhibited systems confirmed that 1-IM and 2-IM corrosion inhibitors underwent the kinetics processes of an anodic reaction (Zhang et al. 2015b). Solomon et al. (2019b) showed the change of bc and ba values in inhibited and uninhibited systems based on the utilisation of NTETD inhibitor. They found that the difference in bc values was greater than that of ba values as NTETD acted as a mixed type corrosion inhibitor with a predominant effect on cathodic-corrosion half reactions. When adding Gemini surfactant, the ba value increased approximately two times due to an alteration of the anodic reaction progress (Zhang et al. 2012). Alternatively, Zhang et al. (2015b) reported that the bc value could be unchanged if an inhibitor does not involve hydrogen production and the cathodic process in an electrochemical system.

The Ecorr value in an inhibited system can be shifted to either a more cathodic side (toward negative value) or a more anodic side (less negative value) with respect to that of the blank solution. Similar to OCP studies, PDP curves can also be used to determine a cathodic or anodic inhibition when the shift in corrosion potential is over 85 mV (Munis et al. 2020; Zhang et al. 2015c). For instance, Munis et al. (2020) found that the difference between Ecorr of blank and inhibited systems was 97 mV using an HDIE inhibitor. This implies that the HDIE functioned as a mixed-type inhibitor predominantly of the anodic reaction. Solomon et al. (2019a) reported that Ecorr shifted toward more positive values using NIMP inhibitor. Consequently, NIMP had greater inhibitive influence on the cathodic corrosion half reactions than anodic corrosion half reactions. However, most imidazolines and imidazoline derivatives act as mixed-type corrosion inhibitors (Alhaffar et al. 2018; Migahed et al. 2017; Obike et al. 2018; Pavithra et al. 2010; Zhang et al. 2007, 2015d, Zheng et al. 2022).

The LPR technique provides a rapid corrosion analysis in which the WE is polarised by a very small potential perturbation of less than ±30 mV versus OCP. The change in the applied potential induces a change in current response between the counter and working electrodes. The polarisation resistance (Rp) is defined by the slope of the linear portion of a voltage versus current curve near the Ecorr (Arya and Joseph 2021). This Rp value can be used to analyse the icorr value of the metal through the Stern and Geary method (Angst and Büchler 2016; Law et al. 2000) Eq. (1):

(1)icorr=babc2.303Rp(ba+bc)

The value of η(%) is obtained by the following equation Eq. (2):

(2)η(%)=icorr0icorricorr0×100

The ba and bc coefficients can be obtained from a Tafel plot or estimated from experience with the testing system. In commercial instruments, the b values are an assumed value of about 18 mV (Agrawal 2001). The Rp value should increase in an inhibited solution compared to the blank test, implying formation of protective layer on carbon steel surfaces (Migahed et al. 2017; Solomon et al. 2019a). The trend of the Rp value as a function of the inhibition efficacy is the same as for PDP analysis (Alhaffar et al. 2018; Umoren 2016).

3.4 EIS studies

EIS is a useful technique to study the formation and destruction of inhibitor films and their corrosion protection mechanism(s) (Tan et al. 1996). Prior to measurements, the WE is immersed in a solution with and without inhibitors to insure a steady OCP. EIS information can be analysed using Nyquist and Bode plots. Figure 6 shows EIS plots including (a) Nyquist, (b) Bode modulus and (c) phase angle plots.

Figure 6: 
						Electrochemical impedance spectra including (a) Nyquist, (b) Bode modulus and (c) phase angle plots (Umoren 2016). Reprinted with permission from Elsevier.
Figure 6:

Electrochemical impedance spectra including (a) Nyquist, (b) Bode modulus and (c) phase angle plots (Umoren 2016). Reprinted with permission from Elsevier.

As shown in Figure 6(a), a Nyquist plot displays a correlation between the real and imaginary parts of the impedance over a wide frequency range (Guo et al. 2021; He et al. 2014; Huang et al. 2016; Zhang et al. 2015d). A semi-circular plot is obtained in the high-frequency side of the impedance spectra, followed by a steep curve at the low-frequency side, corresponding to electrically conductive properties and electrode polarisation effect, respectively (Hadjichristov et al. 2021). An ideal Nyquist plot according to the principles of EIS should present a perfect semicircle. A real Nyquist plot of an inhibited system is commonly a distorted semicircle (Yoo et al. 2012). This distortion is a characteristic of solid electrodes due to physical phenomena such as surface roughness and inhomogeneities of solid electrodes as a result of corrosion processes (Alhaffar et al. 2018; Umoren 2016). The shape of Nyquist plots in the blank and the inhibited system are relatively similar because addition of inhibitors to a corrosive environment does not influence the corrosion mechanism (Alhaffar et al. 2018; Umoren 2016; Zhang et al. 2015d). However, the size of Nyquist impedance for the inhibited system is usually larger than that of the blank system as the charge transfer process can be retarded in the presence of an inhibitor film on carbon steel surfaces (Guo et al. 2021; Solomon et al. 2019b; Zhang et al. 2015d; Zheng et al. 2022).

The lack of frequency information in a Nyquist plot leads to a Bode plot (Poursaee 2014), which is an effective technique for EIS data presentation. A Bode plot is a graph of log frequency on the abscissa (Figure 6(b)), and a combination of the magnitude (log of the absolute value of the impedance (|Z|)) and the phase shift of the frequency response (Figure 6(c)). The performance of a corrosion inhibitor can be analysed from the Bode magnitude plot at low frequencies (He et al. 2014; Zhang et al. 2015d). A large impedance modulus obtained in an inhibited system directly relates to high corrosion protection (He et al. 2014; Zhang et al. 2015b). A Bode phase plot presents the phase shift (θ) on the ordinate versus log of frequency. This is a quick way to assess the formation of an inhibitor film on metal surfaces.

The presence of a time-constant (or frequency) distribution is usually modelled with a constant phase element (CPE). The impedance response for electrochemical systems typically reflects a distribution of reactivity that is commonly represented as a CPE in equivalent electrical circuits (Nkomo and Masia 2021). The number of CPEs can be determined from the number of phase peaks over a wide range of frequencies (Zhang et al. 2015b). For instance, Zhang et al. (2015c) showed phase angle plots obtained from a system containing imidazoline inhibitor in 0.5 M HCl. All the phase angle plots exhibited only one phase peak at a middle frequency.

Moreover, the change in phase angle indicates elimination of diffusion control (Alhaffar et al. 2018). Saviour et al. (Umoren 2016) found a wide phase angle peak showing the presence of more than two time constants in the absence and presence of inhibitor. Zhang et al. (2007) also found two time constants at two frequencies, about 10 and 1000 Hz, corresponding to the influence of the double-layer capacitance/charge transfer resistance and the relaxation process of an adsorbed inhibitor film, respectively. However, Solomon et al. (2019b) reported that one time constant is present in the blank system while double peaks or two time constants appear in inhibited systems due to the influence of film resistance as an additional element. Also, the magnitude of phase angles increases with the addition of imidazolines or imidazoline derivatives because of the surface coverage by the inhibitors molecules (He et al. 2014). This indicates the non-ideal behaviour of a capacitor with an increase in capacitance by the presence of interference, i.e., the adsorbed inhibitor molecules at the interface (Porcayo-Calderon et al. 2015). The imperfections within the capacitor’s material with the existing protective film to prevent corrosion of metal surface create resistance causing the capacitor to dissipate energy. As a result, the polarisation resistance (Rp) values should be increased while the double-layer capacitance (Cdl) values should be decreased with increasing the inhibitor dosage. This also leads to a decrease in the corrosion rate (Ouakki et al. 2019).

Interpretation of the electrochemical behaviour from EIS spectra requires an appropriate physical model that represents the electrochemical process and system. Solomon et al. (2019b) proposed equivalent circuit diagrams that can be perfectly fitted with impedance data in both the blank and inhibited systems. The blank system is fitted with a single time constant model consisting of the solution resistance (Rs), charge transfer resistance (Rct) and constant phase element (CPE). The inhibited system is usually described by the two time constant model having a film resistance (Rf) as additional element. A CPE, a non-integer power that is dependent on frequency, is employed to describe the impedance behaviour of the electric double layer and is also used to calculate the value of double-layer capacitance (Cdl). The Rt and Cdl values can be analysed through data fitting of the impedance spectra (Zhang et al. 2007). The Rt values increase with addition of inhibitors, which leads to the film formation of imidazolines or imidazoline derivatives. This can be used to assess corrosion inhibition efficacy (He et al. 2014; Umoren 2016; Zhang et al. 2007). Alternatively, Cdl values are reduced as the surface coverage of inhibitor molecules replaces that of water molecules. Hence, the local dielectric constant and the thickness of adsorbed inhibitor film at the electrode interface becomes smaller (Zhang et al. 2007; Umoren 2016).

3.5 Weight loss testing

A weight loss technique is a simple and inexpensive method for determining the corrosion rate (CR) of carbon steel and corrosion inhibition efficacy. Weight loss measurements take much time (usually more than a week) to provide accurate results. However, multiple samples can be tested simultaneously (Pearson and Cousins 2016). Rahayu et al. (2021) suggested that long-term measurements weaken the interaction between metal surface and inhibitor molecules leading to an increased corrosion rate. Also, the behaviour and reaction type of inhibitor molecules cannot be obtained using this method. Normally, the weight loss of carbon steel sheet and the corresponding corrosion rate decrease with the addition of effective inhibitors. Furthermore, inhibition efficacy is improved with increased inhibitor dosage (Munis et al. 2020). Additionally, the corrosion inhibition efficacy depends strongly on film formation and surface coverage by inhibitor molecules (He et al. 2014).

The procedure involves an immersion test followed by weight loss measurement done to calculate the average CR value of a specimen in a well-defined environment over an extended period of time as expressed by the following equation (Vedalakshmi et al. 2009) Eq. (3):

(3)CR(mmy)=Weightloss×Kρ×A×t

where weight loss is in g, K = a conversion factor, i.e., 8.76 × 104 (in mm/y), ρ = density of sample (g/cm3) (density of steel = 7.87 g/cm3), A = area of specimen (cm2) and t = time of exposure (h).

The results obtained from weight loss studies can be used to determine the relative corrosiveness of different operating parameters such as inhibitor type and dosage, steel type, vapour versus liquid phase and type of corrosive media (Pearson and Cousins 2016). Not only those operating variables but the CR value is also strongly influenced by the operating temperature. In principle, the CR value increases with temperature (Jevremović et al. 2013; Solomon et al. 2020; Zhang et al. 2013) due to a higher rate of corrosion reaction by corrosive species, e.g., CO2. The operating temperature also causes an alteration of adsorption and desorption behaviours of inhibitor molecules to reach a dynamic equilibrium (Zhang et al. 2013). Solomon et al. (2020) published the CR results obtained under various temperatures together with the calculated corrosion kinetic parameters. For weight loss measurements, the duration of immersion was ranged between 72 and 168 h at various temperatures of 25–60 °C (Ansari et al. 2016; Geng et al. 2022; He et al. 2014; Munis et al. 2020; Solomon et al. 2019a,b; Xiong et al. 2021; Zhang et al. 2012, 2015d). A variety of weight loss data using imidazoline inhibitors was reported depending on the inhibitor dosage and temperature, for example, 0.76 mmpy (400 ppm, at 25 °C) (Solomon et al. 2019b), 0.016 mmpy (1.0 mM/L, at 25 °C) (Munis et al. 2020), 0.51 mmpy (500 ppm, at 25 °C) and 14.35 mmpy (500 ppm, at 60 °C) (Ansari et al. 2016).

3.6 Adsorption studies and corrosion inhibition mechanisms

The analysis of different adsorption isotherms (Guo et al. 2021; Migahed et al. 2017; Solomon et al. 2019a,b; Xiong et al. 2021; Zhang et al. 2015c,d) and computational simulations (Abd El-Lateef et al. 2021; Munis et al. 2020; Rodríguez-Valdez et al. 2006; Xiong et al. 2021; Zhang et al. 2015b) were studied in order to understand the adsorption behaviour of imidazolines and their derivatives on metal surface. Adsorption isotherms can be used to evaluate the strength of interaction between inhibitors and metal surface (Solomon et al. 2019b). The surface coverage (θ) value can be fitted into a Langmuir adsorption isotherm model through the general equation (Guo et al. 2021; Xiong et al. 2021; Zhang et al. 2015d) Eq. (4):

(4)CinhƟ=1Kads+Cinh

where θ is surface coverage inhibition efficiency (η) obtained from weight loss measurement (=η100), Kads is an adsorption–desorption equilibrium constant and Cinh is the concentration of inhibitors. The Langmuir adsorption isotherm model is exploited to explain the adsorption characteristics of imidazolines and their derivatives (Guo et al. 2021; Xiong et al. 2021; Zhang et al. 2015c,d). The Kads value indicates the bond strength between inhibitors and metal surface. High Kads value means strong interaction (Solomon et al. 2019a) and high rate of inhibitor adsorption on the metal surface (Munis et al. 2020). The Kads value also relates to the free energy of adsorption (ΔGads0) as shown by the following expression Eq. (5):

(5)ΔGads0=RTln(55.5)Kads

where R is an ideal gas constant (J mol−1 K−1), T is the thermodynamic temperature (K) and 55.5 is the molar concentration of water in solution (mol dm−3) (Xiong et al. 2021; Zhang et al. 2015b,c). The ΔGads0 value can be taken as a benchmark for comprehensive understanding of the adsorption mechanism(s) of imidazolines and their derivatives on metal surface. Normally, the ΔGads0 value is less negative than −20 kJ mol−1 indicating that the physisorption mechanism or electrostatic interaction occurs between the inhibitor molecules and the metal surface. When the ΔGads0 value is more negative than −40 kJ mol−1, the inhibitors are adsorbed on the metal surface by chemisorption mechanism. The adsorption mechanisms of imidazolines and their derivatives can be both via chemisorption and physisorption (Guo et al. 2021; Munis et al. 2020; Xiong et al. 2021; Zheng et al. 2022) and this infers spontaneous adsorption (He et al. 2014; Solomon et al. 2019a,b, Zhang et al. 2015c).

To understand the adsorption mechanisms of the imidazolines and their derivatives on the metal surface, quantum chemical was employed to describe fractional transfer of electrons from the inhibitor molecules to metal surface and the molecular orbital energies. The frontier orbital model (FOM) could be used to explain the electron transfer behaviour of inhibitor molecules through the lowest unoccupied molecular orbital (LUMO) energy level and the highest occupied molecular orbital (HOMO), indicating the capacity of a particular inhibitor molecule to accept and provide electrons, respectively. The higher HOMO energy (EHUMO) value means that the imidazoline molecule donates electrons to the empty d-orbital of the metal more easily and this links to a strong susceptibility for nucleophilic attack. On the other hand, the lower LUMO energy (ELUMO) value relates to the easier transfer of electrons from the metal surface to the imidazoline molecule or the electron-accepting properties of the inhibitor molecule (Zhang et al. 2015b). Zheng et al. (2022) discussed computational simulations of the optimised molecular structures of various inhibitors and the electronic distributions of HOMO and LUMO. It was suggested that the HOMO and the LUMO of imidazoline molecule distributed around the imidazoline ring indicating that the imidazoline molecule can accept electrons from the Fe atoms and can also transfer electrons to the Fe atoms. The low energy band gap (ΔE) is related to the stability of transition complex that determines the interaction between the inhibitor molecules and the metal surface. The low ΔE value is synonymous with high chemical reactivity and also high inhibition efficiency (Xiong et al. 2021). The imidazoline molecules have low ΔE value of around 2–4 eV (Munis et al. 2020; Zhang et al. 2015b,c) with additional studies on electrostatic potential (ESP) regions (Zheng et al. 2022) to fully explain the existing electrostatic interaction. Figure 7 illustrates ESP distributions of various imidazoline molecules in negative and positive regions as blue and red areas, respectively. Basically, the more positive ESP and more negative ESP regions refer to attraction of nucleophilic and electrophilic species, respectively. The results in Figure 7 show that several extreme points undergoing electrophilic (or nucleophilic) reactions with the Fe atoms are located around the imidazoline ring and the sulfhydryl (for imidazoline derivatives) regions (Zheng et al. 2022). Similarly, Geng et al. (2022) confirmed that two N atoms of the imidazoline ring were highly reactive towards electrophilic reactions (i.e., accept electrons from the Fe atoms).

Figure 7: 
						Electrostatic potential of (a) imidazoline and (b) imidazoline derivatives (Zheng et al. 2022). Reprinted with permission from Elsevier.
Figure 7:

Electrostatic potential of (a) imidazoline and (b) imidazoline derivatives (Zheng et al. 2022). Reprinted with permission from Elsevier.

There are also various authors presenting the mechanisms of corrosion inhibition by imidazoline inhibitors. The corrosion inhibition mechanisms of imidazolines and their derivatives on the metal surface can be explained by physisorption and chemisorption. This is also related to the steel surface, which can appear as (i) normal metal surface and (ii) steel surface hydrated with chloride ions and excess negative charges. The imidazolines and their derivatives have three adsorption centres including N, O and the imidazoline ring (Solomon et al. 2019a). The O and N heteroatoms are protonated and then interact with the hydrated steel surface through physisorption mechanism (Solomon et al. 2019a). On the contrary, the electron pair from the imidazoline ring can be donated to the empty d-orbital of Fe resulting in the formation of covalent kind of bonding via chemisorption mechanism (Solomon et al. 2019a,b; Zheng et al. 2022). Zhang et al. (2015b) reported that the hydrated chloride ions were adsorbed on the electropositive carbon steel surface leading to excess negative charges raised from chloride-containing environment and hence enhanced the adsorption of cations. For example, imidazoline molecules could react with H+ to form N-onium ions (He et al. 2014) and then were electrostatically attracted to the negatively charged steel surface hydrated with chloride ions (Solomon et al. 2019b). This enabled the formation of a protective film of the protonated imidazoline molecules on the solid/liquid interface. In addition, the free N atoms existing on the imidazoline ring can directly bond on the Fe surface yielding a chemical adsorption monolayer (He et al. 2014). The long-chain alkyl group of imidazoline molecule forms a dense hydrophobic layer on the steel surface, which hinders the passage of corrosive species to the metal surface (Qian and Cheng 2019; Solomon et al. 2019b). Also, Munis et al. (2020) reported that the hydrophobic alkyl chain shaded the movement of Cl ions and electrons from reaching the metal surface. Consequently, the more compact and intact the imidazoline film is formed on the steel surface, the better inhibition efficacy it can be achieved through a combination of chemisorption and physisorption.

3.7 Surface characterisation

After corrosion testing, surface examination and surface chemical composition analysis are done using scanning electron microscopy (SEM), energy dispersive X-ray spectroscopy (EDS) and contact angle techniques. SEM is employed to confirm the formation of an inhibitor film on carbon steel surfaces by examining surface morphology at a microscopic level. Rough surface morphology is usually attained on carbon steel surfaces in a blank test because the surface is damaged in an acid solution, yielding deposition of corrosion products on the surface (Solomon et al. 2019b; Umoren 2016; Wang et al. 2019; Zhang et al. 2012). In an inhibited system, carbon steel exhibited fewer corroded surfaces (Alhaffar et al. 2018; Solomon et al. 2019b; Zhang et al. 2012) as shown in Figure 8. The EDX spectrum of a polished sample showed high peak intensity of Fe, which is the main constituent of carbon steel (Migahed et al. 2017). When soaking in a chloride media with no inhibitor, Munis et al. (2020) reported that the surface was rich in chloride ions because they reacted with Fe and produced soluble FeCl3 as a corrosion product. Umoren (2016) found an increase in O atoms caused by the formation of an oxide layer on carbon steel surfaces. In contrast, the presence of inhibitors in the system causes a reduction of O atoms on a weight percent basis, assessed by EDX (Guo et al. 2021; Solomon et al. 2019a,b). Solomon et al. (2019b) believed that the inhibitor molecules were probably adsorbed on the surfaces and effectively protected them from further dissolution. They also observed a higher content of carbon atoms (in wt%) caused by adsorption of NTETD molecules on carbon steel surfaces. However, the N atom in the imidazoline ring could only be detected at some points over the entire metal surface as a specific deposit film spot (Solomon et al. 2019b). The inhibited system exhibits an increase in water contact angle that reveals the hydrophobicity of the surface. Moreover, the surface roughness decreases with the adsorption of inhibitor molecules (He et al. 2014; Zheng et al. 2022).

Figure 8: 
						SEM images of carbon steel samples tested in 1 M HCl solution with an absence or a presence of particular inhibitor: (a) the metal surface before corrosion test, (b) the metal surface after immersion with no inhibitor and (c) the metal surface after immersion containing 500 mg/L inhibitor (Zhang et al. 2012). Reprinted with permission from Wiley Online Library.
Figure 8:

SEM images of carbon steel samples tested in 1 M HCl solution with an absence or a presence of particular inhibitor: (a) the metal surface before corrosion test, (b) the metal surface after immersion with no inhibitor and (c) the metal surface after immersion containing 500 mg/L inhibitor (Zhang et al. 2012). Reprinted with permission from Wiley Online Library.

Many models that have been proposed to explain the corrosion inhibition mechanism(s) of imidazolines and imidazoline derivatives based on electrochemical results and surface characterisation. It is often not possible to define a single general mechanism of inhibition to an inhibitor because the mechanism may change with test conditions. Thus, the predominant mechanism(s) of imidazolines or imidazoline derivatives may vary with these parameters. Table 4 summarises all studies related to imidazoline inhibitors that cover preparation processes, corrosion inhibition measurements, metal types, corrosive media, inhibitor dosage and inhibition efficacy.

4 Conclusions and outlook

In this review, the present knowledge of imidazoline and imidazoline derivatives being used as corrosion inhibitors in petroleum industries was summarised thoroughly. In general, the molecular structure of imidazoline consists of three main parts including (i) an imidazoline head group, (ii) a pendant group that facilitates the molecular bonding on steel surfaces and (iii) a hydrocarbon tail that can function as a hydrophobic film on steel surfaces. The basic characterisation techniques such as FT-IR and NMR are usually employed to identify the resulting molecular structure in the synthesis of imidazolines and their derivatives (via an amidation reaction, followed by a cyclisation reaction to yield an imidazoline ring). As we focused on their prospective application in corrosion inhibition, the evaluation of inhibition efficacy was also given with the details of testing conditions such as metal type (carbon steel), corrosive medium (CO2-saturated solution, HCl, H2SO4, HCl or NH4 solution) and concentration range (0–500 ppm). Electrochemical measurements (e.g., OCP, LPR, PDP and EIS) are commonly used to estimate corrosion resistance behaviour over time. It has been reported that the imidazoline and imidazoline derivatives had the inhibition efficiency of >90%. Adsorption mechanisms of the imidazoline and their derivatives on the metal surface can be explained by quantum chemical. The adsorption of the imidazoline and their derivatives obey the Langmuir adsorption isotherm through three adsorption centres: N, O and the imidazoline ring. Also, SEM, EDS and contact angle techniques are commonly used to examine surface morphology and its chemical composition before and after testing. The inhibited system exhibits a decrease in surface roughness and oxide layer on carbon steel surfaces, whereas the water contact angle is usually found to increase implying an enhanced hydrophobicity of the surface.

Despite the scant research work related to corrosion inhibitors, there are still plenty of opportunities for further studies. For instance, novel and effective technologies should be applied to synthesise imidazoline and imidazoline derivatives. Apart from their inhibition efficacy, their physical and chemical properties should also be considered to ensure their suitability for real applications. Accordingly, near future research trends should be on molecular modification of imidazoline and imidazoline derivatives using versatile and inexpensive preparation processes. Several imidazoline structures with novel functional properties will be developed and proposed but still adhere to the main goal of improving corrosion inhibition.


Corresponding author: Korakot Sombatmankhong, National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA), 111 Thailand Science Park, Phahonyothin Road, Khlong Nueng, Khlong Luang, Pathum Thani, 12120, Thailand, E-mail:

Funding source: National Energy Technology Center (ENTEC), National Science and Technology Development Agency

About the authors

Nipaporn Sriplai

Nipaporn Sriplai received a PhD degree in material science and nanotechnology from Khon Kaen University, Thailand. Presently, she works as a postdoctoral researcher at National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA), Thailand. Her research interests lie in the development of corrosion inhibitor formulations to prevent corrosion in gas pipelines and synthesis of corrosion inhibition compounds.

Korakot Sombatmankhong

Korakot Sombatmankhong received a PhD in chemical engineering from of University of Cambridge, UK. Currently, she is a senior researcher at National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA), Thailand. She has more than 10 years of research experience on chemical deformulation/formulation, polymer gel electrolyte, fuel cells, organic synthesis and arsenic/mercury removal. She has been working in several industry-sponsored research projects, such as the development of corrosion inhibition formulation, imidazoline synthesis and pipeline decommissioning processes.

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was supported by National Energy Technology Center (ENTEC), National Science and Technology Development Agency (NSTDA).

  3. Conflicts of interest: The authors declare no conflicts of interest regarding this article.

References

Abd El-Lateef, H.M., Shalabi, K., and Abdelhamid, A.A. (2021). One-pot synthesis of novel triphenyl hexyl imidazole derivatives catalyzed by ionic liquid for acid corrosion inhibition of C1018 steel: experimental and computational perspectives. J. Mol. Liq. 334: 116081, https://doi.org/10.1016/j.molliq.2021.116081.Suche in Google Scholar

Abd El-Lateef, H.M., Maharram, A.A., Leylufer, I., and Ismayilov, T.A. (2012). Corrosion protection of steel pipelines against CO2 corrosion-a review. Chem. J. 2: 52–63.Suche in Google Scholar

Agrawal, A.K. (2001). Corrosion monitoring. In: Buschow, K.H.J., Cahn, R.W., Flemings, M.C., Ilschner, B., Kramer, E.J., Mahajan, S., and Veyssière, P. (Eds.), Encyclopedia of materials: science and technology. Elsevier, Oxford.Suche in Google Scholar

Al-Moubaraki, A.H. and Obot, I.B. (2021). Corrosion challenges in petroleum refinery operations: sources, mechanisms, mitigation, and future outlook. J. Saudi Chem. Soc. 25: 101370, https://doi.org/10.1016/j.jscs.2021.101370.Suche in Google Scholar

Alaoui, K., Dkhireche, N., Touhami, M.E., and El Kacimi, Y. (2020). Review of application of imidazole and imidazole derivatives as corrosion inhibitors of metals. In: New Challenges and Industrial Applications for Corrosion Prevention and Control. IGI Global, United States.10.4018/978-1-7998-2775-7.ch005Suche in Google Scholar

Alhaffar, M., Umoren, S., Obot, I., and Ali, S. (2018). Isoxazolidine derivatives as corrosion inhibitors for low carbon steel in HCl solution: experimental, theoretical and effect of KI studies. RSC Adv. 8: 1764–1777, https://doi.org/10.1039/c7ra11549k.Suche in Google Scholar PubMed PubMed Central

Angst, U.M. and Büchler, M. (2016). Corrosion rate measurements in concrete-a closer look at the linear polarization resistance method. In: Dehn, F., Beushausen, H.-D., Alexander, M.G., Moyo, P. (Eds.), Concrete repair, rehabilitation and retrofitting IV. Taylor & Francis Group, London, pp. 198–199.10.1201/b18972-125Suche in Google Scholar

Ansari, K.R., Quraishi, M.A., Singh, A., Ramkumar, S., and Obote, I.B. (2016). Corrosion inhibition of N80 steel in 15% HCl by pyrazolone derivatives: electrochemical, surface and quantum chemical studies. RSC Adv. 6: 24130–24141, https://doi.org/10.1039/c5ra25441h.Suche in Google Scholar

Arya, S.B. and Joseph, F.J. (2021). Electrochemical methods in tribocorrosion. In: Siddaiah, A., Ramachandran, R., and Menezes, P.L. (Eds.), Tribocorrosion, Chapter 3. Academic Press, United States.Suche in Google Scholar

Bajpai, D. and Tyagi, V.K. (2006). Fatty imidazolines: chemistry, synthesis, properties and their industrial applications. J. Oleo Sci. 55: 319–329, https://doi.org/10.5650/jos.55.319.Suche in Google Scholar

Bajpai, D. and Tyagi, V.K. (2008). Microwave synthesis of cationic fatty imidazolines and their characterization. J. Surfactants Deterg. 11: 79–87, https://doi.org/10.1007/s11743-007-1057-z.Suche in Google Scholar

Bereket, G., Hür, E., and Öğretir, C. (2002). Quantum chemical studies on some imidazole derivatives as corrosion inhibitors for iron in acidic medium. J. Mol. Struct. Theochem 578: 79–88, https://doi.org/10.1016/S0166-1280(01)00684-4.Suche in Google Scholar

Budiana, S., Andreana, B., Rahayu, D.U.C., Nurani, D.A., Krisnandi, Y.K., and Budianto, E. (2020). Synthesis of imidazoline-palmitic derivative using MAOS (microwave assisted organic synthesis) method. IOP Conf. Ser. Mater. Sci. Eng. 763: 012006, https://doi.org/10.1088/1757-899x/763/1/012006.Suche in Google Scholar

Demirbas, A. (2008). Production of biodiesel from tall oil. Energy Sources, Part A Recovery, Util. Environ. Eff. 30: 1896–1902, https://doi.org/10.1080/15567030701468050.Suche in Google Scholar

Desimone, M.P., Gordillo, G., and Simison, S.N. (2011). The effect of temperature and concentration on the corrosion inhibition mechanism of an amphiphilic amido-amine in CO2 saturated solution. Corrosion Sci. 53: 4033–4043, https://doi.org/10.1016/j.corsci.2011.08.009.Suche in Google Scholar

Ditama, I., Muktiarti, N., and Soegijono, B. (2020). Imidazoline derivatives based on silane functional group as corrosion inhibitor on mild steel. AIP Conf. Proc. 2242: 020019.10.1038/s41529-021-00201-5Suche in Google Scholar

El-Katori, E.E., Ahmed, M., and Nady, H. (2022). Imidazole derivatives based on glycourils as efficient anti-corrosion inhibitors for copper in HNO3 solution: synthesis, electrochemical, surface, and theoretical approaches. Colloids Surf. A Physicochem. Eng. Asp. 649: 129391, https://doi.org/10.1016/j.colsurfa.2022.129391.Suche in Google Scholar

Fazal, B.R., Becker, T., Kinsella, B. and Lepkova, K. (2022). A review of plant extracts as green corrosion inhibitors for CO2 corrosion of carbon steel. NPJ Mater. Degrad. 6: 5, https://doi.org/10.1038/s41529-021-00201-5.Suche in Google Scholar

Ferm, R.J. and Riebsomer, J.L. (1954). The chemistry of the 2-imidazolines and imidazolidines. Chem. Rev. 54: 593–613, https://doi.org/10.1021/cr60170a002.Suche in Google Scholar

Fuchs-Godec, R. (2006). The adsorption, CMC determination and corrosion inhibition of some N-alkyl quaternary ammonium salts on carbon steel surface in 2M H2SO4. Colloids Surf. A Physicochem. Eng. Asp. 280: 130–139, https://doi.org/10.1016/j.colsurfa.2006.01.046.Suche in Google Scholar

Geng, S., Hu, J., Yu, J., Zhang, C., Wang, H., and Zhong, X. (2022). Rosin imidazoline as an eco-friendly corrosion inhibitor for the carbon steel in CO2-containing solution and its synergistic effect with thiourea. J. Mol. Struct. 1250: 131778, https://doi.org/10.1016/j.molstruc.2021.131778.Suche in Google Scholar

Guo, W., Talha, M., Lin, Y., Ma, Y., and Kong, X. (2021). Effect of phosphonate functional group on corrosion inhibition of imidazoline derivatives in acidic environment. J. Colloid Interface Sci. 597: 242–259, https://doi.org/10.1016/j.jcis.2021.02.131.Suche in Google Scholar PubMed

Hadjichristov, G.B., Marinov, Y.G., Vlakhov, T.E., and Scaramuzza, N. (2021). Phospholipid Langmuir-Blodgett nano-thin monolayers: electrical response to cadmium ions and harmful volatile organic compounds. In: Iglič, A., Rappolt, M., and García-Sáez, A.J. (Eds.), Advances in biomembranes and lipid self-assembly, Chapter 5. Academic Press, United States.10.1016/j.corsci.2014.02.004Suche in Google Scholar

Haladu, S.A., Umoren, S.A., Ali, S.A., Solomon, M.M. and Mohammed, A.-R.I. (2019). Synthesis, characterization and electrochemical evaluation of anticorrosion property of a tetrapolymer for carbon steel in strong acid media. Chin. J. Chem. Eng. 27: 965–978, https://doi.org/10.1016/j.cjche.2018.07.015.Suche in Google Scholar

He, X., Jiang, Y., Li, C., Wang, W., Hou, B., and Wu, L. (2014). Inhibition properties and adsorption behavior of imidazole and 2-phenyl-2-imidazoline on AA5052 in 1.0 M HCl solution. Corrosion Sci. 83: 124–136, https://doi.org/10.1016/j.corsci.2014.02.004.Suche in Google Scholar

Hooshmand Zaferani, S., Sharifi, M., Zaarei, D., and Shishesaz, M.R. (2013). Application of eco-friendly products as corrosion inhibitors for metals in acid pickling processes – a review. J. Environ. Chem. Eng. 1: 652–657, https://doi.org/10.1016/j.jece.2013.09.019.Suche in Google Scholar

Hou, Y., Zhu, L., He, K., Yang, Z., Ma, S., and Lei, J. (2022). Synthesis of three imidazole derivatives and corrosion inhibition performance for copper. J. Mol. Liq. 348: 118432, https://doi.org/10.1016/j.molliq.2021.118432.Suche in Google Scholar

Huang, J., Li, Z., Liaw, B.Y., and Zhang, J. (2016). Graphical analysis of electrochemical impedance spectroscopy data in Bode and Nyquist representations. J. Power Sources 309: 82–98, https://doi.org/10.1016/j.jpowsour.2016.01.073.Suche in Google Scholar

Jaal, R.A., Ismail, M.C., and Ariwahjoedi, B. (2014). A review of CO2 corrosion inhibition by imidazoline-based inhibitor. In: MATEC Web of conferences. EDP Sciences, France, 05012.10.15255/KUI.2016.014Suche in Google Scholar

Jevremović, I., Singer, M., Nešić, S., and Mišković-Stanković, V. (2013). Inhibition properties of self-assembled corrosion inhibitor talloil diethylenetriamine imidazoline for mild steel corrosion in chloride solution saturated with carbon dioxide. Corrosion Sci. 77: 265–272, https://doi.org/10.1016/j.corsci.2013.08.012.Suche in Google Scholar

Ji Ram, V., Sethi, A., Nath, M., and Pratap, R. (2019). Five-membered heterocycles. In: The chemistry of heterocycles, Chapter 5. Elsevier, United States.10.1002/jsde.12363Suche in Google Scholar

Kong, X., FanWeiyu, L., Ming, Q., ChengduoLuo, H., and Yao, Y. (2016). Study on the synthesis and corrosion inhibition performance of Mannich-modified imidazoline. Kemija u Industriji 65: 353–358, https://doi.org/10.15255/kui.2016.014.Suche in Google Scholar

Kousar, K., Walczak, M.S., Ljungdahl, T., Wetzel, A., Oskarsson, H., Restuccia, P., Ahmad, E.A., Harrison, N.M., and Lindsay, R. (2021). Corrosion inhibition of carbon steel in hydrochloric acid: elucidating the performance of an imidazoline-based surfactant. Corrosion Sci. 180: 109195, https://doi.org/10.1016/j.corsci.2020.109195.Suche in Google Scholar

Kousar, K., Ljungdahl, T., Wetzel, A., Dowhyj, M., Oskarsson, H., Walton, A.S., Walczak, M.S., and Lindsay, R. (2020). An exemplar imidazoline surfactant for corrosion inhibitor studies: synthesis, characterization, and physicochemical properties. J. Surfactants Deterg. 23: 225–234, https://doi.org/10.1002/jsde.12363.Suche in Google Scholar

Law, D.W., Millard, S.G., and Bungey, J.H. (2000). Linear polarisation resistance measurements using a potentiostatically controlled guard ring. NDT E Int. 33: 15–21, https://doi.org/10.1016/s0963-8695(99)00015-8.Suche in Google Scholar

Levinson, M.I. (1999). Rinse-added fabric softener technology at the close of the twentieth century. J. Surfactants Deterg. 2: 223–235, https://doi.org/10.1007/s11743-999-0077-4.Suche in Google Scholar

Li, Y.Z., Xu, N., Guo, X.P., and Zhang, G.A. (2017). Inhibition effect of imidazoline inhibitor on the crevice corrosion of N80 carbon steel in the CO2-saturated NaCl solution containing acetic acid. Corrosion Sci. 126: 127–141, https://doi.org/10.1016/j.corsci.2017.06.021.Suche in Google Scholar

Li, Z., Xu, X., Li, Y. and Liu, Z. (2021). Effect of imidazoline inhibitor on stress corrosion cracking of P110 steel in simulated annulus environments of CO2 injection well. J. Electroanal. Chem. 886: 115105, https://doi.org/10.1016/j.jelechem.2021.115105.Suche in Google Scholar

Liu, H. and Du, D.M. (2009). Recent advances in the synthesis of 2-imidazolines and their applications in homogeneous catalysis. Adv. Synth. Catal. 351: 489–519, https://doi.org/10.1002/adsc.200800797.Suche in Google Scholar

Martin, J.A. and Valone, F.W. (1985). The existence of imidazoline corrosion inhibitors. Corrosion 41: 5, https://doi.org/10.5006/1.3582003.Suche in Google Scholar

Martínez-Palou, R., Rivera, J., Zepeda, L.G., Rodríguez, A.N., Hernández, M.A., Marín-Cruz, J., and Estrada, A. (2004). Evaluation of corrosion inhibitors synthesized from fatty acids and fatty alcohols isolated from sugar cane wax. Corrosion 60: 465–470, https://doi.org/10.5006/1.3299242.Suche in Google Scholar

Martínez-Palou, R., de Paz, G., Marín-Cruz, J. and Zepeda, L.G. (2003). Synthesis of long chain 2-alkyl-1-(2-hydroxyethyl)-2-imidazolines under microwave in solvent-free conditions. Synlett 2003: 1847–1849, https://doi.org/10.1055/s-2003-41410.Suche in Google Scholar

McMahon, A.J. (1991). The mechanism of action of an oleic imidazoline based corrosion inhibitor for oilfield use. Colloid. Surface. 59: 187–208, https://doi.org/10.1016/0166-6622(91)80247-l.Suche in Google Scholar

Mehedi, M.S.A. and Tepe, J.J. (2020). Recent advances in the synthesis of imidazolines (2009–2020). Adv. Synth. Catal. 362: 4189–4225, https://doi.org/10.1002/adsc.202000709.Suche in Google Scholar

Migahed, M.A., El-Rabiei, M.M., Nady, H., Gomaa, H.M., and Zaki, E.G. (2017). Corrosion inhibition behavior of synthesized imidazolium ionic liquids for carbon steel in deep oil wells formation water. J. Bio- and Tribo-Corrosion 3: 22, https://doi.org/10.1007/s40735-017-0080-5.Suche in Google Scholar

Mishra, A., Aslam, J., Verma, C., Quraishi, M.A., and Ebenso, E.E. (2020). Imidazoles as highly effective heterocyclic corrosion inhibitors for metals and alloys in aqueous electrolytes: a review. J. Taiwan Inst. Chem. Eng. 114: 341–358, https://doi.org/10.1016/j.jtice.2020.08.034.Suche in Google Scholar

Monticelli, C. (2018). Corrosion inhibitors. In: Wandelt, K. (Ed.), Encyclopedia of interfacial chemistry. Oxford: Elsevier.10.1063/5.0007980Suche in Google Scholar

Moradighadi, N., Lewis, S., Olivo, J.D., Young, D., Brown, B., and Nešić, S. (2021). Determining critical micelle concentration of organic corrosion inhibitors and its effectiveness in corrosion mitigation. Corrosion 77: 266–275, https://doi.org/10.5006/3679.Suche in Google Scholar

Morales-Gil, P., Walczak, M.S., Camargo, C., Cottis, R., Romero, J.M., and Lindsay, R. (2015). Corrosion inhibition of carbon-steel with 2-mercaptobenzimidazole in hydrochloric acid. Corrosion Sci. 101: 47–55, https://doi.org/10.1016/j.corsci.2015.08.032.Suche in Google Scholar

Mu, M., Chen, Y., Dong, Y., Aijun, Y., and Liao, Q. (2017). Synergistic inhibition effect of imidazoline derivative and potassium iodide on carbon steel corrosion in amidosulfuric acid solution. Corrosion Rev. 35: 463–471, https://doi.org/10.1515/corrrev-2017-0035.Suche in Google Scholar

Muktiarti, N., Ditama, I., and Soegijono, B. (2020). Characterization of imidazoline derivates synthesized from soybean oil fatty acids as corrosion inhibitors on mild steel. AIP Conf. Proc. 2242: 020023.10.1016/j.molliq.2021.117815Suche in Google Scholar

Munis, A., Zhao, T., Zheng, M., Rehman, A.U. and Wang, F. (2020). A newly synthesized green corrosion inhibitor imidazoline derivative for carbon steel in 7.5% NH4Cl solution. Sustain. Chem. Pharm. 16: 100258, https://doi.org/10.1016/j.scp.2020.100258.Suche in Google Scholar

Nkomo, D. and Masia, N. (2021). An insight on corrosion resistance ability of biocompatible dental implants through electrochemical impedance spectroscopy. IntechOpen, United Kingdom.10.3390/ma13020398Suche in Google Scholar PubMed PubMed Central

Obike, A., Okafor, P., Uwakwe, K., Jiang, X., and Qu, D. (2018). The inhibition of CO2 corrosion of L360 mild steel in 3.5% NaCl solution by imidazoline derivatives. Int. J. Corrosion Scale Inhib. 7: 318–330.10.1016/j.corsci.2010.07.034Suche in Google Scholar

Ouakki, M., Galai, M. and Cherkaoui, M. (2022). Imidazole derivatives as efficient and potential class of corrosion inhibitors for metals and alloys in aqueous electrolytes: a review. J. Mol. Liq. 345: 117815, https://doi.org/10.1016/j.molliq.2021.117815.Suche in Google Scholar

Ouakki, M., Galai, M., Rbaa, M., Abousalem, A.S., Lakhrissi, B., Rifi, E.H., and Cherkaoui, M. (2019). Quantum chemical and experimental evaluation of the inhibitory action of two imidazole derivatives on mild steel corrosion in sulphuric acid medium. Heliyon 5: e02759, https://doi.org/10.1016/j.heliyon.2019.e02759.Suche in Google Scholar PubMed PubMed Central

Pan, C., Mao, J., and Jin, W. (2020). Effect of imidazoline inhibitor on the rehabilitation of reinforced concrete with electromigration method. Materials 13: 398, https://doi.org/10.3390/ma13020398.Suche in Google Scholar PubMed PubMed Central

Pavithra, M.K., Venkatesha, T.V., Vathsala, K., and Nayana, K.O. (2010). Synergistic effect of halide ions on improving corrosion inhibition behaviour of benzisothiozole-3-piperizine hydrochloride on mild steel in 0.5M H2SO4 medium. Corrosion Sci. 52: 3811–3819, https://doi.org/10.1016/j.corsci.2010.07.034.Suche in Google Scholar

Pearson, P. and Cousins, A. (2016). Assessment of corrosion in amine-based post-combustion capture of carbon dioxide systems. In: Feron, P.H.M. (Ed.). Absorption-based post-combustion capture of carbon dioxide. Woodhead Publishing, United States, p. 18.10.1016/j.molliq.2019.111674Suche in Google Scholar

Popova, A., Christov, M., and Vasilev, A. (2007). Inhibitive properties of quaternary ammonium bromides of N-containing heterocycles on acid mild steel corrosion. Part II: EIS results. Corrosion Sci. 49: 3290–3302, https://doi.org/10.1016/j.corsci.2007.03.012.Suche in Google Scholar

Porcayo-Calderon, J., De La Escalera, L.M.M., Canto, J., and Casales-Diaz, M. (2015). Imidazoline derivatives based on coffee oil as CO2 corrosion inhibitor. Int. J. Electrochem. Sci. 10: 3160–3176.10.1016/j.corsci.2006.05.036Suche in Google Scholar

Poursaee, A. (2014). Corrosion sensing for assessing and monitoring civil infrastructures. In: Wang, M.L., Lynch, J.P., and Sohn, H. (Eds.), Sensor technologies for civil infrastructures. Woodhead Publishing, United States, p. 13.10.1016/j.corsci.2020.108916Suche in Google Scholar

Qian, S. and Cheng, Y.F. (2019). Synergism of imidazoline and sodium dodecylbenzenesulphonate inhibitors on corrosion inhibition of X52 carbon steel in CO2-saturated chloride solutions. J. Mol. Liq. 294: 111674, https://doi.org/10.1016/j.molliq.2019.111674.Suche in Google Scholar

Rahayu, D.U.C., Krisnandi, Y.K., Susanto, B.H., Abdullah, I., Nurani, D.A., Saragi, I.R., Yuniastuti, A., Mujahiduzzakka, M., Cahyani, S., Fitriani, S., et al.. (2021). Synthesis of oleic-imidazoline and investigation on its inhibition efficiency for the corrosion of low carbon steel in chloride environments. Int. J. Corrosion Scale Inhib. 10: 1282–1293.10.1016/j.petrol.2021.109235Suche in Google Scholar

Rodríguez-Valdez, L.M., Villamisar, W., Casales, M., González-Rodriguez, J.G., Martínez-Villafañe, A., Martinez, L., and Glossman-Mitnik, D. (2006). Computational simulations of the molecular structure and corrosion properties of amidoethyl, aminoethyl and hydroxyethyl imidazolines inhibitors. Corrosion Sci. 48: 4053–4064, https://doi.org/10.1016/j.corsci.2006.05.036.Suche in Google Scholar

Shamsa, A., Barker, R., Hua, Y., Barmatov, E., Hughes, T.L., and Neville, A. (2020). Performance evaluation of an imidazoline corrosion inhibitor in a CO2-saturated environment with emphasis on localised corrosion. Corrosion Sci. 176: 108916, https://doi.org/10.1016/j.corsci.2020.108916.Suche in Google Scholar

Shamsa, A., Barker, R., Hua, Y., Barmatov, E., Hughes, T.L., and Neville, A. (2021). Impact of corrosion products on performance of imidazoline corrosion inhibitor on X65 carbon steel in CO2 environments. Corrosion Sci. 185: 109423, https://doi.org/10.1016/j.corsci.2021.109423.Suche in Google Scholar

Shamsa, A., Barmatov, E., Hughes, T.L., Hua, Y., Neville, A., and Barker, R. (2022). Hydrolysis of imidazoline based corrosion inhibitor and effects on inhibition performance of X65 steel in CO2 saturated brine. J. Petrol. Sci. Eng. 208: 109235, https://doi.org/10.1016/j.petrol.2021.109235.Suche in Google Scholar

Sharma, M.K. and Ghuge, R.B. (2018). Chemical and biological profiles of vicinal diaryl-substituted thiophenes, imidazolines, selendiazoles, and isoselenazoles. In: Vicinal diaryl substituted heterocycles, Chapter 10. Elsevier, United States, pp. 305–326.10.1016/B978-0-12-409547-2.13591-7Suche in Google Scholar

Solomon, M.M., Umoren, S.A., Quraishi, M.A. and Mazumder, M.A.J. (2019a). Corrosion inhibition of N80 steel in simulated acidizing environment by N-(2-(2-pentadecyl-4, 5-dihydro-1H-imidazol-1-YL) ethyl) palmitamide. J. Mol. Liq. 273: 476–487, https://doi.org/10.1016/j.molliq.2018.10.032.Suche in Google Scholar

Solomon, M.M., Umoren, S.A., Quraishi, M.A., and Salman, M. (2019b). Myristic acid based imidazoline derivative as effective corrosion inhibitor for steel in 15% HCl medium. J. Colloid Interface Sci. 551: 47–60, https://doi.org/10.1016/j.jcis.2019.05.004.Suche in Google Scholar PubMed

Solomon, M.M., Umoren, S.A., Quraishi, M.A., Tripathy, D.B. and Abai, E.J. (2020). Effect of akyl chain length, flow, and temperature on the corrosion inhibition of carbon steel in a simulated acidizing environment by an imidazoline-based inhibitor. J. Petrol. Sci. Eng. 187: 106801, https://doi.org/10.1016/j.petrol.2019.106801.Suche in Google Scholar

Stupnisek, L.E., Kasunic, D., and Vorkapic, F.J. (1995). Imidazole derivatives as corrosion inhibitors for zinc in hydrochloric acid. Corrosion 51: 767–772, https://doi.org/10.5006/1.3293554.Suche in Google Scholar

Tan, Y.J., Bailey, S., and Kinsella, B. (1996). An investigation of the formation and destruction of corrosion inhibitor films using electrochemical impedance spectroscopy (EIS). Corrosion Sci. 38: 1545–1561, https://doi.org/10.1016/0010-938x(96)00047-9.Suche in Google Scholar

Telegdi, J., Shaban, A., and Vastag, G. (2018). Biocorrosion—steel. In: Wandelt, K. (Ed.), Encyclopedia of interfacial chemistry. Elsevier, Oxford.10.1016/j.corsci.2009.07.014Suche in Google Scholar

Tripathy, D.B. and Mishra, A. (2017). Convenient synthesis, characterization and surface active properties of novel cationic gemini surfactants with carbonate linkage based on C12C18 sat./unsat. fatty acids. J. Appl. Res. Technol. 15: 93–101, https://doi.org/10.1016/j.jart.2016.12.004.Suche in Google Scholar

Tyagi, R., Tyagi, V.K., and Pandey, S.K. (2007). Imidazoline and its derivatives: an overview. J. Oleo Sci. 56: 211–222, https://doi.org/10.5650/jos.56.211.Suche in Google Scholar PubMed

Umoren, S.A. (2016). Polypropylene glycol: a novel corrosion inhibitor for ×60 pipeline steel in 15% HCl solution. J. Mol. Liq. 219: 946–958, https://doi.org/10.1016/j.molliq.2016.03.077.Suche in Google Scholar

Usman, B.J. and Ali, S.A. (2018). Carbon dioxide corrosion inhibitors: a review. Arabian J. Sci. Eng. 43: 1–22, https://doi.org/10.1007/s13369-017-2949-5.Suche in Google Scholar

Usyatinsky, A.Y. and Khmelnitsky, Y.L. (2000). Microwave-assisted synthesis of substituted imidazoles on a solid support under solvent-free conditions. Tetrahedron Lett. 41: 5031–5034, https://doi.org/10.1016/s0040-4039(00)00771-1.Suche in Google Scholar

Vedalakshmi, R., Manoharan, S.P., Song, H.-W., and Palaniswamy, N. (2009). Application of harmonic analysis in measuring the corrosion rate of rebar in concrete. Corrosion Sci. 51: 2777–2789, https://doi.org/10.1016/j.corsci.2009.07.014.Suche in Google Scholar

Vinutha, M.R. and Venkatesha, T.V. (2016). Review on mechanistic action of inhibitors on steel corrosion in acidic media. Port. Electrochim. Acta 34: 157–184, https://doi.org/10.4152/pea.201603157.Suche in Google Scholar

Wahyuningrum, D., Achmad, S., Syah, Y., Buchari, B., and Ariwahjoedi, B. (2008). The synthesis of imidazoline derivative compounds as corrosion inhibitor towards carbon steel in 1% NaCl Solution. ITB J. Sci. 40: 33–48, https://doi.org/10.5614/itbj.sci.2008.40.1.4.Suche in Google Scholar

Wang, B., Du, M., Zhang, J., and Gao, C.J. (2011). Electrochemical and surface analysis studies on corrosion inhibition of Q235 steel by imidazoline derivative against CO2 corrosion. Corrosion Sci. 53: 353–361, https://doi.org/10.1016/j.corsci.2010.09.042.Suche in Google Scholar

Wang, B., Du, M., Zhang, J., Li, C., Liu, J., Liu, H., Li, R., and Li, Z. (2019). Corrosion inhibition of mild steel by the hydrolysate of an imidazoline-based inhibitor in CO2-saturated solution. RSC Adv. 9: 36546–36557, https://doi.org/10.1039/c9ra05322k.Suche in Google Scholar PubMed PubMed Central

Wang, P., Xiong, L., He, Z., Xu, X., Hu, J., Chen, Q., Zhang, R., Pu, J., and Guo, L. (2022). Synergistic effect of imidazoline derivative and benzimidazole as corrosion inhibitors for Q235 steel: an electrochemical, XPS, FT-IR and MD study. Arabian J. Sci. Eng. 47: 7123–7134, https://doi.org/10.1007/s13369-021-06540-4.Suche in Google Scholar

Wolkenberg, S., Wisnoski, D., Leister, W., Wang, Y., Zhao, Z., and Lindsley, C. (2004). Efficient synthesis of imidazoles from aldehydes and 1,2-diketones using microwave irradiation. Org. Lett. 6: 1453–1456, https://doi.org/10.1021/ol049682b.Suche in Google Scholar PubMed

Xhanari, K., Wang, Y., Yang, Z., and Finšgar, M. (2021). A review of recent advances in the inhibition of sweet corrosion. Chem. Rec. 21: 1845–1875.10.1016/j.apsusc.2018.02.115Suche in Google Scholar

Xiong, L., Wang, P., He, Z., Chen, Q., Pu, J., and Zhang, R. (2021). Corrosion behaviors of Q235 carbon steel under imidazoline derivatives as corrosion inhibitors: experimental and computational investigations. Arab. J. Chem. 14: 102952, https://doi.org/10.1016/j.arabjc.2020.102952.Suche in Google Scholar

Yoo, S.-H., Kim, Y.-W., Chung, K., Baik, S.-Y., and Kim, J.-S. (2012). Synthesis and corrosion inhibition behavior of imidazoline derivatives based on vegetable oil. Corrosion Sci. 59: 42–54, https://doi.org/10.1016/j.corsci.2012.02.011.Suche in Google Scholar

Zeng, D., Dong, B., Shi, S., Pan, J., Yu, H., Li, J., Ding, Y., and Cai, L. (2018). Effects of temperature on corrosion of N80 and 3Cr steels in the simulated CO2 auxiliary steam drive environment. Arabian J. Sci. Eng. 43: 3845–3854, https://doi.org/10.1007/s13369-018-3081-x.Suche in Google Scholar

Zhang, G., Chen, C., Lu, M., Chai, C., and Wu, Y. (2007). Evaluation of inhibition efficiency of an imidazoline derivative in CO2-containing aqueous solution. Mater. Chem. Phys. 105: 331–340, https://doi.org/10.1016/j.matchemphys.2007.04.076.Suche in Google Scholar

Zhang, H., Gao, K., Yan, L., and Pang, X. (2017). Inhibition of the corrosion of X70 and Q235 steel in CO2-saturated brine by imidazoline-based inhibitor. J. Electroanal. Chem. 791: 83–94, https://doi.org/10.1016/j.jelechem.2017.02.046.Suche in Google Scholar

Zhang, H., Pang, X. and Gao, K. (2018). Localized CO2 corrosion of carbon steel with different microstructures in brine solutions with an imidazoline-based inhibitor. Appl. Surf. Sci. 442: 446–460, https://doi.org/10.1016/j.apsusc.2018.02.115.Suche in Google Scholar

Zhang, H., Pang, X., Zhou, M., Liu, C., Wei, L., and Gao, K. (2015a). The behavior of pre-corrosion effect on the performance of imidazoline-based inhibitor in 3 wt.% NaCl solution saturated with CO2. Appl. Surf. Sci. 356: 63–72, https://doi.org/10.1016/j.apsusc.2015.08.003.Suche in Google Scholar

Zhang, J., Gong, X.‐L., Song, W.‐W., Jiang, B., and Du, M. (2012). Synthesis of imidazoline‐based dissymmetric bis‐quaternary ammonium gemini surfactant and its inhibition mechanism on Q235 steel in hydrochloric acid medium. Mater. Corros. 63: 636–645.10.1016/j.electacta.2012.02.101Suche in Google Scholar

Zhang, J., Niu, L., Zhu, F., Li, C., and Du, M. (2013). Theoretical and experimental studies for corrosion inhibition performance of Q235 steel by imidazoline inhibitors against CO2 corrosion. J. Surfactants Deterg. 16: 947–956, https://doi.org/10.1007/s11743-013-1515-8.Suche in Google Scholar

Zhang, K., Xu, B., Yang, W., Yin, X., Liu, Y., and Chen, Y. (2015b). Halogen-substituted imidazoline derivatives as corrosion inhibitors for mild steel in hydrochloric acid solution. Corrosion Sci. 90: 284–295, https://doi.org/10.1016/j.corsci.2014.10.032.Suche in Google Scholar

Zhang, K., Yang, W., Xu, B., Liu, Y., Yin, X., and Chen, Y. (2015c). Corrosion inhibition of mild steel by bromide-substituted imidazoline in hydrochloric acid. J. Taiwan Inst. Chem. Eng. 57: 167–174, https://doi.org/10.1016/j.jtice.2015.05.021.Suche in Google Scholar

Zhang, L., He, Y., Zhou, Y., Yang, R., Yang, Q., Qing, D. and Niu, Q. (2015d). A novel imidazoline derivative as corrosion inhibitor for P110 carbon steel in hydrochloric acid environment. Petroleum 1: 237–243, https://doi.org/10.1016/j.petlm.2015.10.007.Suche in Google Scholar

Zhang, X., Wang, F., He, Y., and Du, Y. (2001). Study of the inhibition mechanism of imidazoline amide on CO2 corrosion of Armco iron. Corrosion Sci. 43: 1417–1431, https://doi.org/10.1016/s0010-938x(00)00160-8.Suche in Google Scholar

Zhao, J. and Chen, G. (2012). The synergistic inhibition effect of oleic-based imidazoline and sodium benzoate on mild steel corrosion in a CO2-saturated brine solution. Electrochim. Acta 69: 247–255, https://doi.org/10.1016/j.electacta.2012.02.101.Suche in Google Scholar

Zhao, P., Ren, R., Zhou, H., Ouyang, J., Li, Z., and Sun, D. (2021). Preparation and properties of imidazoline surfactant as additive for warm mix asphalt. Construct. Build. Mater. 273: 121692, https://doi.org/10.1016/j.conbuildmat.2020.121692.Suche in Google Scholar

Zhao, B. and Zou, L.K. (2013). Corrosion inhibition of chloroacetic-acid modified imidazoline for Q235 steel in H2SO4 solution. Adv. Mater. Res. 710: 41–44, https://doi.org/10.4028/www.scientific.net/AMR.710.41.Suche in Google Scholar

Zheng, Z., Hu, J., Eliaz, N., Zhou, L., Yuan, X., and Zhong, X. (2022). Mercaptopropionic acid-modified oleic imidazoline as a highly efficient corrosion inhibitor for carbon steel in CO2-saturated formation water. Corrosion Sci. 194: 109930, https://doi.org/10.1016/j.corsci.2021.109930.Suche in Google Scholar

Zhong, H., Shi, Z., Jiang, G., and Yuan, Z. (2020). Synergistic inhibitory effects of free nitrous acid and imidazoline derivative on metal corrosion in a simulated water injection system. Water Res. 184: 116122, https://doi.org/10.1016/j.watres.2020.116122.Suche in Google Scholar PubMed

Zuo, Y., Yang, L., Tan, Y., Wang, Y. and Zhao, J. (2017). The effects of thioureido imidazoline and NaNO2 on passivation and pitting corrosion of X70 steel in acidic NaCl solution. Corrosion Sci. 120: 99–106, https://doi.org/10.1016/j.corsci.2016.12.015.Suche in Google Scholar

Received: 2022-10-09
Accepted: 2023-02-02
Published Online: 2023-03-08
Published in Print: 2023-06-27

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 22.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2022-0092/html
Button zum nach oben scrollen