Next Article in Journal
An In Silico Molecular Modelling-Based Prediction of Potential Keap1 Inhibitors from Hemidesmus indicus (L.) R.Br. against Oxidative-Stress-Induced Diseases
Next Article in Special Issue
Biocompatible and Biodegradable Surfactants from Orange Peel for Oil Spill Remediation
Previous Article in Journal
A Fast and Efficient Procedure of Iron Species Determination Based on HPLC with a Short Column and Detection in High Resolution ICP OES
Previous Article in Special Issue
Cd(II) and Pd(II) Mixed Ligand Complexes of Dithiocarbamate and Tertiary Phosphine Ligands—Spectroscopic, Anti-Microbial, and Computational Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Imine-Tethering Cationic Surfactants: Synthesis, Surface Activity, and Investigation of the Corrosion Mitigation Impact on Carbon Steel in Acidic Chloride Medium via Various Techniques

by
Hany M. Abd El-Lateef
1,2,*,
Ahmed H. Tantawy
3,*,
Kamal A. Soliman
3,
Salah Eid
3,4 and
Mohamed A. Abo-Riya
3
1
Chemistry Department, College of Science, King Faisal University, Al-Ahsa 31982, Saudi Arabia
2
Chemistry Department, Faculty of Science, Sohag University, Sohag 82534, Egypt
3
Chemistry Department, Faculty of Science, Benha University, Benha 13518, Egypt
4
Chemistry Department, College of Science and Arts, Jouf University, Alqurayat 77455, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(11), 4540; https://doi.org/10.3390/molecules28114540
Submission received: 6 May 2023 / Revised: 25 May 2023 / Accepted: 31 May 2023 / Published: 3 June 2023
(This article belongs to the Special Issue Synthesis, Characterization and Application of Surfactants II)

Abstract

:
Novel imine-tethering cationic surfactants, namely (E)-3-((2-chlorobenzylidene)amino)-N-(2-(decyloxy)-2-oxoethyl)-N,N-dimethylpropan-1-aminium chloride (ICS-10) and (E)-3-((2-chlorobenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(tetradecyloxy)ethyl)propan-1-aminium chloride (ICS-14), were synthesized, and the chemical structures were elucidated by various spectroscopic approaches. The surface properties of the target-prepared imine-tethering cationic surfactants were investigated. The effects of both synthesized imine surfactants on carbon steel corrosion in a 1.0 M HCl solution were investigated by weight loss (WL), potentiodynamic polarization (PDP), and scanning electron microscopy (SEM) methods. The outcomes show that the inhibition effectiveness rises with raising the concentration and diminishes with raising the temperature. The inhibition efficiency of 91.53 and 94.58 % were attained in the presence of the optimum concentration of 0.5 mM of ICS-10 and ICS-14, respectively. The activation energy (Ea) and heat of adsorption (Qads) were calculated and explained. Additionally, the synthesized compounds were investigated using density functional theory (DFT). Monte Carlo (MC) simulation was utilized to understand the mechanism of adsorption of inhibitors on the Fe (110) surface.

1. Introduction

Carbon steel (C-Steel) is the most popular and common construction material across a lot of industries. Carbon steel suffers from serious corrosion problems [1,2,3]. Inhibitors are a common and effective way to safeguard metals in acidic media. [4]. Acidic solutions, particularly HCl acid solutions, are frequently utilized in chemical and industrial operations [5]. The majority of organic molecule inhibitors function by adhering to the metal’s surface and binding to heteroatoms, such as sulfur, oxygen, and nitrogen, as well as many linkages. [6,7]. The surface charge and nature of metals, the chemical construction of organic inhibitors, and the type of aggressive solution all influence this process [8]. Numerous studies have been conducted on the corrosion and inhibition of C-steel in acidic solutions [9,10,11,12].
Cationic surfactants, mainly quaternary amines, are considered one of the most important surfactants having numerous uses in different fields. They are frequently used in a variety of industrial applications, including medicines, pharmaceutical formulations, food processing, and oil recovery [13,14,15]. In recent decades, cationic surfactants have shown promising results as corrosion inhibitors [16,17,18]. This is accredited to the absorption of cationic surfactants on the steel interface, forming an adsorption film layer, which separates the metal surface from the surrounding environment [19]. The efficiency of this material as an inhibitor depends on its adsorption capacity [20]. Most cationic surfactants can be used as corrosion inhibitors containing O, N, S, P, heteroatoms, and/or π-bonds that permit the adsorption of the surfactant molecule on the electrode interface [21,22]. The inhibition process is achieved by the buildup of a coordinate covalent bond between the free electrons of these atoms or multiple bonds and lower unoccupied orbitals on the metal [23,24].
Therefore, compounds with these atoms or bonds have a strong ability to resist corrosion. Hence, began an investigation of Schiff base compounds, which contain a heteroatom and π-bond as corrosion inhibitors. Organic substances known as Schiff bases have numerous uses in biology and materials research [25,26]. Schiff-based cationic surfactants can adsorb onto metal surfaces by charge-transfer complex bonds between the polar groups of Schiff base and metal surfaces [27,28]. Cationic surfactants bearing Schiff base moiety proved high efficiency in an acidic solution, where a nitrogen atom is accomplished to form coordinate covalent bonds with the steel interface by its unshared, lone pairs of electrons, whereas the π-bond interacts physically, increasing their adsorption attraction to the metal interface. Imine surfactant molecules are categorized by their high capability to adsorb at the interfaces, owing to their unique amphipathic construction, which contains two opposing parts, a tail, and a head. The head is considered by its high electronic richness, which is a set of some electronic-rich functional groups, such as double, carbonyl, nitrogen, the imine group, oxygen, etc. A successive film on the metal surface could be shaped by the surfactant’s hydrocarbon, which has the capability to separate the corroded metal surface from contact with corrosive mediums [29,30,31,32,33,34]. In addition, the imine cationic surfactant is considered by its biocidal activity, which improves its capacity in the petroleum sector. Furthermore, in acidic conditions, these compounds exhibit effective anticorrosive properties.
The objective of our study was to prepare and study the inhibition effect of (E)-3-((2-chlorobenzylidene)amino)-N-(2-(decyloxy)-2-oxoethyl)-N,N-dimethylpropan-1-aminium chloride (ICS-10) and (E)-3-((2-chlorobenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(tetradecyloxy)ethyl)propan-1-aminium chloride (ICS-14) on the corrosion of carbon steel in an acidic chloride medium, utilizing the PDP, WL, and SEM methods. In addition, activation energy (Ea) and the heat of adsorption (Qads) were computed and explained. In addition, the two inhibitors, ICS-10 and ICS-14, were studied by DFT and MC simulation to understand the corrosion inhibition mechanism.

2. Results and Discussion

2.1. Synthesis

As shown in Scheme 1, the overall synthesis procedure of the target imine-tethering cationic surfactants consists of two steps. Initially, 2-chlorobenzaldehyde was reserved to react with 3-(N,N Dimethylamino)-1-propylamine to form 3-((2-chlorobenzylidene)amino)-N,N-dimethylpropan-1-amine (1). The prepared Schiff base (1) was refluxed with decyl 2-chloroacetate and tetradecyl 2-chloroacetate in ethyl acetate as a solvent at 70 °C to produce novel cationic imine surfactants [ICS-10] and [ICS-14] with yields equal to 77 and 81%, respectively. The investigations were conducted on the structure of the synthetic surfactants using various spectroscopic tools, such as IR and NMR spectra. The recorded data in the spectra elucidated the functional groups in the obtained compounds, as recorded in the experimental section.

2.2. Surface Activity Measurements

The surface tension (γ) of the imine surfactant solutions vs. their bulk concentrations, in mol/L at 25 °C, were plotted, and the CMC (Critical Micelle Concentration) values were graphically acquired and recorded at the breakpoint of a definite concentration of the prepared surfactants, as shown in Figure 1A. Table 1 lists the CMC values and the surface tension at the CMC (γCMC). It could be proven that the generated cationic surfactant’s CMC value decreases as the hydrophobic chain lengthens; this could be related to the growing hydrophobicity and a reduction of the dissolving of the imine surfactants, so the system’s freedom energy increases, which results in a surfactant molecule aggregated into the micelles. In order to avoid contact with the aqueous medium, the hydrophilic group directs to the solution, while the hydrophobic series is towards the interior body of the system, lowering the freedom of the system. So, by the length of the hydrophobic tail, the inclination of the surfactant compound to produce a micelle is, therefore, reduced by the CMC. As depicted in Table 1, the values of the CMC, determined in 1.0 M of HCl (Figure 1B), were partially reduced compared with the values in pure aqueous solutions (Table 1).
Based on the electrical conductivity measurements (K) of the as-prepared surfactants, the CMC values and the dissociation degree of the counter ion (α) were evaluated and determined at a certain temperature (298 K), as shown in Figure 1C, where the α values attained from the ratio of the slopes above and lower the break, revealing the CMC, were determined. Usually, the degrees of counterion binding (β) values are determined according to the following recorded relationship: β = 1 − α [35,36]. The values of β and α are recorded in Table 1, which shows that the β values reduce and the α values rise with a lengthening chain of hydrophobic alkyl at 25 °C [37]. It was noted that the prepared cationic surfactants with higher hydrophobic chain tails had a low CMC value. Furthermore, the CMC values were determined via electrical conductivity, which is the same as those obtained using surface tension.
From the calculated values of γ, the effectiveness values of the imine compounds were carried out via the next equation [38]:
ΠCMC = γ0γCMC
where γCMC and γ0 are the γ values at the CMC and pure H2O, respectively. From the results shown in Table 1, it was detected that the most effective imine surfactant was ICS-14 which revealed the highest decrease in the γ at the CMC, whereas the highest lessening of the γ at the CMC reached 37.54 [39].
In addition, the surface area of Amin refers to the packaging density of the as-prepared cationic surfactants at the interface of air/water; this is very significant for explaining the surface features of the imine surfactants. The Γmax value, showing the efficacy of the interfacial adsorption of the surfactant adsorption values, was investigated by the Gibbs adsorption equation [38]:
Γ max = 1 2.303 n R T d γ d log C T
where T = 298 K, R denotes the gas constant, /dlogC characterizes the slop, and n symbolizes the number of ions that generate inside the solution via the surfactant molecule dissociation (equal to two in our calculations). The Γmax values are listed in Table 1, where it is exhibited that the lengthening of the hydrophobic chain of the imine surfactants shifts Γmax to lower values of concentrations. Additionally, the Amin calculations, with the minimum surface area in nm2, were carried out via the below equation [40]:
A min = 10 14 N A × Γ max
where NA equals 6.022 × 1023. As shown in Table 1, the Γmax and Amin values are listed. The data showed that the Amin values were raised by lengthening the hydrophobic chain, owing to increasing the area of the hydrophobic chain employed by each inhibitor molecule upsurge; consequently, the Amin was augmented, and the Γmax decreased, whereby in decreasing the Γmax, the distances among the surfactant compound upsurged, leading to upsurged Amin values.
According to Gibb’s adsorption equations, the calculation of the thermodynamic parameters of the adsorption and micellization of the imine surfactants (ICS-10 and ICS-14) was conducted, as shown below [41]:
Δ G m i c o = 2.303 ( 2 α ) R T log C M C
Δ G a d s o = Δ G m i c o 0.0602 π C M C A min
where Δ G m i c o and Δ G a d s o are the micellization and adsorption-free energies, respectively. As seen in Table 1, the values of Δ G m i c o and Δ G a d s o are continuously negative, indicating the spontaneousness of these two routes, but there are higher increases in the negativity values of Δ G a d s o (via the lengthening of the hydrophobic moiety) than in Δ G m i c o , referring to the inclination of the surfactants to be adsorbed at the interface. By inspection of the data in Table 1, it was found that the values of the Δ G m i c o and Δ G a d s o increase by the lengthening of a hydrophobic chain, indicating that the tendency of the molecules is absorbed on the interface.

2.3. Weight Loss (WL) Studies

The C-steel corrosion in molar hydrochloric acid in the absence and presence of diverse concentrations of imine surfactants was investigated by WL studies. The effects of the addition of various concentrations of ICS-10 and ICS-14 on the weight loss (A) and inhibition capacity (B) of carbon steel in a 1.0 M HCl solution are presented in Figure 2A,B, and the values are recorded in Table S1. From Figure 2, it is observed that the rates of corrosion (weight loss) declined with the presence of the imine surfactants, as compared to the blank HCl (the free inhibitor). The IE rises with the increasing surfactant concentration, displaying a maximum rise in the IE of 91.13 and 94.16% at surfactant doses of 0.5 mmol of ICS-10 and ICS-14, respectively. The protection of C-steel corrosion with the presence of imine surfactant molecules can be ascribed to the adsorption of the surfactant onto metallic substrates, which prevents steel contact with the corrosive medium and, consequently, does not allow the corrosion process to occur. The rise in the IE with the increasing surfactant concentration designates that more inhibitor species are adsorbed on the metal interface, leading to the development of a defensive layer at the steel–electrolyte interface [42]. As shown in Table S2, the inhibition powers of our synthesized imine surfactants (~95) are close to or slightly higher than that of the previous synthetic Schiff base compounds (~94) [34,43,44,45]. Thus, in spite of the yield products’ accounts, the use of synthesized compounds, such as our synthesized cationic surfactants, for the protection of the metal surfaces in the corrosive solutions containing chloride produces more yield than the previous synthetic compounds containing the hydroxyl group [44].

2.4. Potentiodynamic Polarization (PDP)

PDP plots for the C-steel corrosion both in the blank and inhibited solutions containing various concentrations of ICS-10 and ICS-14 compounds in a 1.0 M HCl medium are depicted in Figure 3.
It was detected that, from Figure 3, all the diagrams display comparable profiles, which verifies that neither the anodic (viz. metal corrosion) nor cathodic (viz. the reaction of hydrogen evolution, RHE) mechanisms were altered by the addition of the inhibitor to the aggressive medium [46]. From the extrapolation of Tafel branches, numerous electrochemical parameters were calculated; cathodic (βc) and anodic (βa) Tafel slopes, the corrosion current density (icor), and the potential corrosion (Ecor), which is mentioned in Table 2. The inhibition effectiveness (IE/%) and surface coverage (θ) were estimated utilizing the next equation [46].
I E / % = 1 i s u r f i f r e e × 100 = θ × 100
where ifree and isurf are the icor without and with the ICS-10 and ICS-14 surfactants, respectively.
Examining Figure 3 exposes that there is no substantial modification in the values of either the βc and βa upon adding different ICS-10 and ICS-14 amounts, demonstrating that the mechanism of inhibition by the investigated imine surfactants continues via the hindrance of the efficient cathodic and anodic locations on the metal interface [47]. The inhibitory influence of ICS-10 and ICS-14 is correspondingly shown as the current density decline in both the cathodic and anodic lines, which agrees well with previously predicted thermodynamic and surface-active characteristics.
Table 2 outlines the protection capacities’ rise with the increasing ICS-10 and ICS-14 concentration, accomplishing ca. 91.53 and 94.58% in the presence of 0.5 mM [ICS-10 and ICS-14] and agreeing well with the gravimetric investigation findings. Such performance could be reasonable by the rising adsorption of ICS-10 and ICS-14 molecules onto the steel–HCl solution interface [47,48], which is reinforced by increasing a part of the surface coverage (cf. with Table 2). The inhibitor is classified as an anodic or cathodic type if the change of the Ecor in the presence of the surfactant inhibitor is ±85 mV from that in the absence of the inhibitor [48]. The presence of the ICS-10 and ICS-14 surfactants shifts the Ecor to values less than 85 mV compared to the blank system (the free inhibitor), indicating that ICS-10 and ICS-14 surfactants can be characterized as mixed-type inhibitors, i.e., supporting the retardation of both the cathodic hydrogen evolution and anodic dissolution of the C-steel reactions [48].

2.5. Morphology Studies

The surface morphologies determined the severity of the corrosion attack. The SEM observation of the C-steel samples after and before immersion in HCl in the absence and presence of 0. 5 mM of ICS-10 and ICS-14 are presented in Figure 4. In the absence of inhibitors, the morphology of the C-steel samples in the molar HCl revealed that the surface was severely degraded. The surface of the C-steel became smooth in the presence of ICS-10 and ICS-14. These findings show that ICS-10 and ICS-14 adsorbed on the C-steel surface and produced a layer that successfully protected the carbon steel surface from the corrosive ions [10].

2.6. Adsorption Isotherm Considerations

Weight loss and PDP measurements have established that imine surfactant molecules could efficiently impede the C-steel corrosion in an acidic chloride solution. Out of data fitting for identifying the best adsorption isotherm model and then selecting the approaching value of R2, the adsorption route mechanism of a surfactant on steel could be determined as well. The surfactant concentration (Cinh) and the surface coverage (θ) were utilized to compute the adsorption equilibrium constant (Kads), and minor alterations in the coverage could affect the protective efficacy.
The values of θ for differing ICS-10 and ICS-14 surfactant concentrations were estimated by using the weight loss results to select the optimal isotherm of adsorption for these compounds. Different models of adsorption isotherms, for example, the Langmuir, Temkin, Freundlich, and Frumkin models, were utilized to fit the weight loss data to comprehend how the surfactant compounds are adsorbed on the steel interface. The closest match was for the Langmuir adsorption isotherm, based on the next equation [49]:
C s u r f θ = 1 K a d s + C s u r f
where Kads represents the adsorption equilibrium constant, and Csurf is the molar dose of the ICS-10 and ICS-14 surfactants. The plots of Csurf vs. Csurf for steel corrosion in 1.0 M of HCl are illustrated in Figure 5. The attained diagrams are linear with higher than 0.999. The intercept allows for the computing of the Kads. The equilibrium constant is equal to 164 × 103 and 238 × 103 for ICS-10 and ICS-14, respectively. The higher values of the Kads signify an interaction between the steel interface and surfactant, showing that the surfactant additives are powerfully adsorbed on the electrode substrate and, in turn, deliver superior protection efficacy to the surfactant.
The following equation describes the relationship between the adsorption’s standard free energy and Kads, the adsorption equilibrium constant [50]:
K a d s = 1 55.5 exp Δ G a d s o R T
where the molar concentration of H2O is represented by the number (55.5), R is the gas constant (8.314 J. mol−1.K−1), and T is the absolute temperature.
The Δ G a d s o values for the ICS-10 and ICS-14 adsorbed on the surface of the C-steel in the molar HCl are equal to −39.0 kJ mol−1 and −39.9 kJ mol−1. The negative value of Δ G a d s o indicates the spontaneous of adsorption the ICS-10 and ICS-14 on the C-steel surface [10].

2.7. Thermodynamic/Adsorption Calculations

Table 3 depicts the temperature effect on the corrosion of C-steel in 1.0 M of HCl in the absence and presence of 0.0005 M of ICS-10 and ICS-14 after 24 h. The data reveal that ICS-10 and ICS-14 compounds are efficient inhibitors. As the temperature increase has an opposite connection with the inhibition effectiveness, the increase in temperature likely leads to the desorption of adsorbed ICS-10 and ICS-14 molecules from the metal interface.
The obvious activation energy (Ea) for the C-steel corrosion in molar hydrochloric acid in the absence and presence of 0.0005 M ICS-10 and ICS-14 was computed using an Arrhenius-type equation [51]:
log C R 2 C R 1 = E a 2.303 R 1 T 1 1 T 2
where R symbolizes the universal gas constant, a signifies the Arrhenius pre-exponential factor, T is the Kelvin temperature, and CR1 and CR2 are the corrosion rates computed after 24 h at temperatures T1 and T2, respectively. The values of the activation energies (Ea) were calculated and are given in Table 3. This value in the inhibited system is higher than the value obtained for the blank, which designates that ICS-10 and ICS-14 molecules are adsorbed on the C-steel surface, leading to this increase in the activation energy [52].
The adsorption heat (Qads) of the inhibitors was calculated using the next equation [53]:
Q a d s = 2.303 R log θ 2 1 θ 2 log θ 1 1 θ 1 × T 1 T 2 T 2 T 1
where θ1 and θ2 are the parts of the covered surface at temperatures T1 and T2, respectively. The Qads value was computed and is listed in Table 3. It has a negative value, so the inhibitor’s adsorption at the metal–acid interface is exothermic, and the amount of surface coverage is reduced with the temperature rise [54,55].

2.8. DFT Calculations

The optimized structure of the two inhibitors, ICS-10 and ICS-14, at the B3LYP level of theory and their frontier molecular orbitals are shown in Figure 6. The binding of the surfactant molecules to the steel interface increases with a higher HOMO energy value, which indicates the greater electron donation of the surfactant molecule to the empty d-orbital of the steel and a lower LUMO energy level, which refers to the ability of the surfactant to gain electrons from the d-orbital of the steel [56]. As seen in Table 4, if we compare the inhibitors in the gas phase, ICS-14 has a higher HOMO and lower LUMO energy value, indicating that ICS-14 shows higher inhibition efficiency than the ICS-10 inhibitor molecule. The adsorption ability of the inhibitor increases with a smaller energy gap. The smaller energy gap refers to higher inhibition efficiency and chemical reactivity [6,57]. As seen in Table 4, ICS-14 has a lower energy gap than the ICS-10 inhibitor molecule.
The HOMO electron distribution for the ICS-10 and ICS-14, as seen in Figure 6, were localized on a chloride ion, while the LUMO electron density was distributed on a phenylmethanimine moiety.
The dipole moment (µ) is an important quantum descriptor that reflects the global polarity of a molecule. The µ is associated with protection efficacy. The protection efficacy upsurges with the increment of the µ. As shown in Table 4, the ICS-14 inhibitor molecule has the highest dipole moment.
The electrons transferred number (ΔN) is well-intended. If the ΔN values are ˂3.6, the protection efficacy upsurges by increasing the capability of the surfactant molecules to contribute electrons to the steel interface [58,59]. As seen in Table 4, the ΔN values for the two compounds (ICS-10 and ICS-14) are positive, and the ICS-14 inhibitor shows a higher value than the ICS-10 inhibitor molecule, which indicates that the ICS-14 inhibitor has higher inhibition efficiency than the ICS-10 inhibitor.
The global softness (σ) and hardness (η) for the two ICS-10 and ICS-14 inhibitors were calculated, which determines the reactivity of the surfactant molecule. The inhibitor with a greater softness value and a lower hardness value is predictable as having the highest inhibition efficiency. Therefore, as seen in Table 4, the ICS-14 inhibitor shows higher inhibition efficiency than the ICS-10 surfactant molecule. The molecular electrostatic potential map (ESP) is a visual tool applied to determine the reactive places of the surfactant molecule. The red and yellow regions refer to a negative ESP. The more negatively charged atoms are, the more reactive the atoms are. As seen in Figure 6, the negative ESP is located on the chloride ion, nitrogen, and oxygen atoms of both inhibitors.

2.9. MC Simulation and the Mechanism of Inhibition

MC simulation can determine the most configurations of the surfactant molecules adsorbed on a steel interface. The two studied inhibitors, ICS-10 and ICS-14, were simulated in vacuum and gas phases. As seen in Figure 7, the two inhibitors adsorbed on the Fe (110) in a parallel adsorption configuration, and the adsorption energy values of the surfactants on the Fe (110) were found to be −180.22 and −208.84 kcal/mol for ICS-10 and ICS-14, respectively. The higher adsorption energy of the CS-14 inhibitor than the CS-10 inhibitor leads to the strong interaction of the CS-14 inhibitor on the Fe (110), leading to the production of a film that protects the C-steel surface against the aggressive environment (the HCl solution), leading to higher inhibition effectiveness.
It is widely recognized that the polar units found in polar carbon-based molecules operate as reaction positions for the molecules and speed up the process of adsorption by forming bonds between the polar atoms of the steel surface inhibitors. On the basis of the inhibitor’s orientation, shape, size, and electrical charge, the compound’s efficiency is mostly governed by the degree of adsorption [60,61]. Frequent factors that affect the protective effectiveness of carbon-based corrosion additives are the length of the carbon chain, size, and chemical composition of the organic inhibitor; the conjugated bonding and/or aromaticity in the compound; the type, nature, and some bonding groups or atoms in the compound; charges produced on the surface of the metal; the capability of a film to be multiple or single to procedure cross-linked or dense, and the ability to produce a complex with an efficient inhibitor atom with a metallic interface [62,63]. Taking into account the existing case, according to the presence of the π-electrons of the benzene ring, the electronegative N atom in its structure, and a double bond (-C=C-) (Scheme 1), the cationic imine-based surfactant may sensibly substantiate its great protection capacity and usage as an efficient corrosion additive. The π-electrons in the surfactant could not only localize the vacant d-orbital of the steel interface but also may receive the d-orbital electrons of the steel substrate to produce feedback of the steel–surfactant bond. The adsorption of the CS-14 inhibitor over the CS-10 inhibitor results in a strong interaction between the CS-14 inhibitor and C-steel, generating a layer that shields the C-steel surface against the aggressive environment.

3. Experimental

3.1. Materials

The metal used was carbon steel (C-Steel), which had the following structure (wt./wt.%): 0.05 Ni, 0.02 Cr, 0.0256 Si, 1.81 Mn, 0.09 P, 0.1 C, 0.001 V, 0.01 Mo, 0.03 Cu, and the remainder was iron.
Furthermore, 3-(N,N Dimethylamino)-1-propylamine (99%), 2-Chlorobenzaldehyde (97%), and Decyl (98%) were purchased from Acros Organics Company (Geel, Belgium). Hexadecyl (97%) alcohol was attained from M/s S.D. Fine Chemicals Pvt. Ltd. (Mumbai, India) and tetrahydrofuran (99%) were obtained from Alnasr-Chemical Company. Solvents (ethyl acetate, ethyl alcohol absolute (99%), and diethyl ether (99%)) were obtained from Algomhoria Chemical Co., Cairo, Egypt. All the utilized solvents and reagents were received without further purification.

3.2. Equipment and Instrumentation

The construction of the as-prepared imine-tethering cationic surfactants was chemically clarified using melting points (Gallen-Kamp, Cambridge, UK) and FTIR spectrum, performed in potassium bromide on (iS10-FTIR spectrophotometer, Paisley, UK) a Thermo-Nicolet, 1H, and 13C NMR spectra, and estimated in d6-DMSO using ALPHA-FT-IR-BRUKER-Pt-ATR. Using the ring method, a Tensiometer-K6 processor (KRÜSS-Company, Hamburg, Germany) was utilized to determine the surface tension of as-prepared cationic surfactants at 25 °C [64] in distilled water and 1.0 M of HCl. The electrical conductivity (K) of synthesized surfactants was identified by an electrical conductivity meter, AD3000 type (EC/TDS), and at a temperature of 25 °C.

3.3. Synthesis of the As-Prepared Cationic Surfactant

An amount of 10 mmol of 2-chlorobenzaldehyde was dissolved in 100 mL of ethyl alcohol and added in a single-neck flask, then mixed with 10 mmol of 3-(N,N-dimethylamino)-1-propyl amine. The solution mixture was refluxed for 7–8 h at 85 °C via thin-layer chromatography (TLC); the accomplishment of the reaction was detected, and the reaction was left to cool down at room temperature. A pale-yellow solid was obtained and purified by recrystallization in methanol to generate the target compound (1) with a yield reach of 90%. The purified compound was utilized directly in the next step.
Compound (1) (20 mmol) was dissolved in 100 mL of ethyl acetate and taken in two portions separately to react with 10 mmol of decyl- and tetradecyl-2-chloroacetates, stirred at 70 °C for 30 h in ethyl acetate as a solvent. The solid products were recrystallized from diethyl ether and dehydrated under a vacuum at 40 °C to give the target cationic imine surfactants (ICS-10 and ICS-14). The purification process was completed by removing the undesirable materials during washing with diethyl ether. The structure of the prepared cationic surfactants was elucidated via IR, 1H, and 13C NMR spectra.
ICS-10: Yellow, solid color, mp= 105 °C, and yield = 93%. FT-IR (KBr pellet) cm−1: 3382, 3013, 2952, 2923,1749, 1653, 1601, 1539, and 1252. 1H NMR-400 MHz (DMSO-d6) d ppm: 0.84(t, 3H, CH3-CH2−), 1.25(s, 14H, (CH2)7−), 1.64(m, 2H, CH2-CH2-O), 2.15(m, 2H, CH2CH2-N), 3.64(t, 2H, CH2-N+), 4.13(s, 6H, 2CH3-N+), 4.20(t, 2H, CH2-N=CH), 4.53(t, 2H, CH2-O), 4.56(t, 2H, CH2-CO), 6.94 (d, 1H, aromatic CH), 7.34(m, 2H, aromatic CH), 7.44(m, 1H, aromatic CH), and 8.60(s, 1H, CH=N). 13C NMR-400 MHz, δC (ppm) (DMSO-d6): 167, 165, 161, 132, 131, 118, 115, 66, 63, 61, 55, 51, 31, 29.46, 29.17, 29.06, 28, 25, 24, 22, and 14.01. (Supporting Information, Figure S1A–C).
ICS-14: Pale-yellow solid color, mp= 110 °C, and yield= 95%. FT-IR (KBr pellet) cm−1: 3268, 3063, 2919, 2850,1751, 1642, 1611, 1537, 1445, and 1254. 1H NMR-400 MHz (DMSO-d6) d ppm: 0.84(t, 3H, CH3-CH2−), 1.25(s, 20H, (CH2)10−), 1.63(m, 2H, CH2-CH2-O), 1.85(m, 2H, CH2CH2-N), 1.99(t, 2H, CH2-N+), 3.01(s, 6H, 2CH3-N+), 3.25(t, 2H, CH2-N=CH), 3.81(t, 2H, CH2 CH2-O), 4.24(t, 2H, CH2-O), 4.60(t, 2H, CH2-CO), 7.03(m, 1H, aromatic CH), 7.42 (s, 2H, aromatic CH), and 7.53(m, 1H, aromatic CH), 8.56(s, 1H, CH=N). 13C NMR (151 MHz, DMSO-d6) δ: 166.41, 164.67, 161.18, 144.56, 134.05, 127.79, 118.18, 115.52, 68.23, 63.64, 60.96, 54.27, 50.65, 31.76, 29.47, 29.41, 29.30, 29.17, 28.97, 26.22, 22.94, 22.56, 22.11, and 14.43 (Figure 8A–C).

3.4. Weight Loss (WL) Measurements

In the weight loss (WL) method, C-steel coupons with dimensions of 4.8 × 2 × 1.1 cm with an uncovered surface area of 34.16 cm2 were used. These coupons were mechanically brushed at Benha University with various grades of emery paper before being rinsed with distilled water, propanone, and distilled water. Before adding C-steel samples into the test solution, the weight of the samples was determined. The experimentation was replicated 3 times, with the average WL reported each time. The following equations were utilized to calculate the surface coverage (θ) and the inhibition capacity (IE/%):
I E / % = 1 Δ W s u r f Δ W f r e e × 100 = θ × 100
where ΔWsurf and ΔWfree reflect the weight change of C-steel samples per unit area with and without ICS-10 and ICS-14 surfactants, respectively. The next equation was used to measure the corrosion rate (CR) in g·hr−1.cm−2:
C R = Δ W t × A
where ΔW signifies the change in the weight of the specimen in grams, A is the sample area in cm2, and t denotes the time spent immersed in hours.

3.5. Electrochemical Measurements

Because corrosion occurs via electrochemical reactions, electrochemical techniques are ideal for the study of corrosion processes [65,66]. Electrochemical measurements were performed by a Meinsberg Potentiostat/Galvanostat electrochemical workstation (PS6, Hamburg, Germany), complemented with the EC-Lab software PS Remote, at Benha University. The Galvanostat/Potentiostat was committed in another way to a conventional cell with three-electrode systems. The counter and reference electrodes were a Pt-wire and a saturated calomel electrode (SCE), respectively. The working electrode was made of a C-steel that was encased in Araldite resin, with just 1.6 cm2 of the C-steel electrode apparent. The C-steel was rubbed with emery papers of up to a 2500 grade before being degreased with propanone and washed with distilled water. The potentiodynamic polarization (PDP) plots for C-steel corrosion in 1.0 M of HCl in the absence and presence of various amounts of ICS-10 and ICS-14 surfactants were determined. All experiments were conducted with a scan rate of 1.0 mV/s [67,68,69].

3.6. SEM Observations

The SEM examination was carried out at Mansoura University in Egypt [70] with the help of the JSM-6510LV. The C-steel specimen was scratched with emery papers of up to a 2500 grade and then maintained in 1.0 M of HCl for 24 h in the absence of 0.0005 M of ICS-10 and ICS-14 surfactants. After this period of inundation, the samples were washed with distilled H2O, properly desiccated, and placed in the spectrometer.

3.7. Computational Details

Full geometry optimization of the two inhibitors’ ICS10 and ICS14 molecules were studied without any constraints using DFT calculations with B3LYP functional [71,72,73] and 6–31 g(d,p) basis sets implemented in the GAMESS program [74,75], and the calculations were carried out in the gas phase. The results were evaluated to include the highest occupied molecular orbital (HOMO), the energy gap (ΔE), the lowest unoccupied molecular orbital (LUMO), the dipole moment (µ), hardness (η), global softness (σ), and the number of electrons transferred (ΔN). The η, σ, and ΔN were computed by:
η = E L U M O E H O M O 2
σ = 1 η
Δ N = ϕ χ i n h 2 ( η F e + η i n h )
where φ , χ i n h , η F e , a n d η i n h are the function work of Fe (110) (4.820 eV), the surfactant electronegativity, iron chemical hardness (0 eV), and surfactant chemical hardness, respectively.

3.8. MC Simulations

In this study, an Fe (110) surface was selected for the adsorption of a single inhibitor molecule. MC simulation was performed by the adsorption locator module. The Fe (110) surface was increased by constructing a supercell, 15 × 15, with 25 Å vacuum heights above the surface. Condensed-phase optimized molecular potentials for atomistic simulation studies (COMPASS) were utilized for the adsorption process. For the summation method, the Ewald method was set for the electrostatic and atom-based method for van der Waals. Additional information on the methodology of Monte Carlo simulations can be found in previous works [76]. The adsorption energy of the inhibitor was calculated by the following equation:
Eads = EcomplexEFeEinh
where Ecomplex, EFe, and Einh are the total energies of the inhibitor on the Fe (110) surface, the energy of the iron (110) surface, and the energy of the surfactant molecule, respectively.

4. Conclusions

Two cationic imine surfactants were synthesized, and we elucidated their chemical structures, which showed good surface-active properties. These prepared imine cationic surfactants (ICS-10 and ICS-14) exposed the capability of producing a defensive layer against the aggressive acidic medium at diverse temperatures via the adsorption of their molecules on the metal surface. Surface inspection through SEM displayed a significant inhibition of C-steel deterioration by the construction of a protecting film on the metal interface. Interestingly, the small negative values suggest that the adsorption process is spontaneous. There was no considerable shift in the potential corrosion values, inferring that these imine compounds act as mixed-type inhibitors. Theoretical calculations based on DFT are in agreement with the experimental findings that exhibited ICS-10 and ICS-14 as efficient corrosion additives, and the order of the inhibition capability for the investigated imine surfactants is as follows: ICS-14 (94.58%) > ICS-10 (91.53 %).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28114540/s1, Figure S1. IR, 1H, and 13C NMR spectra of ((2-chlorobenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(decyloxy)ethyl) propan-1-ammonium chloride [ICS-10]. Table S1: The effects of the addition of various concentrations of ICS-10 and ICS-14 on the weight of carbon steel in 1.0 M HCl solution. Table S2: The inhibition capacity comparisons for some conventional Surfactants.

Author Contributions

Conceptualization, A.H.T. and K.A.S.; Methodology, H.M.A.E.-L., A.H.T., K.A.S., S.E. and M.A.A.-R.; Software, Mohamed M.A.A.-R.; Validation, S.E. and Mohamed M.A.A.-R.; Formal analysis, H.M.A.E.-L., A.H.T., K.A.S., S.E. and Mohamed M.A.A.-R.; Investigation, H.M.A.E.-L., A.H.T., K.A.S., S.E. and Mohamed M.A.A.-R.; Data curation, S.E. and Mohamed M.A.A.-R.; Writing—original draft, H.M.A.E.-L., A.H.T., K.A.S., S.E. and Mohamed M.A.A.-R.; Writing—review & editing, A.H.T., K.A.S., S.E. and Mohamed M.A.A.-R.; Supervision, A.H.T.; Project administration, H.M.A.E.-L.; Funding acquisition, H.M.A.E.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research, King Faisal University, Saudi Arabia, through GRANT3455.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data generated in this work are available upon request from the corresponding author.

Acknowledgments

The authors acknowledge the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research at King Faisal University, Saudi Arabia, for financial support under the annual funding track (GRANT3455).

Conflicts of Interest

The authors declare no conflict of interest. The authors declare that they have no known competing interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Adway, A.I.; Abbas, M.A.; Zakaria, K. New Schiff base cationic surfactants as corrosion inhibitors for carbon steel in acidic medium: Weight loss, electrochemical and SEM characterization techniques. Res. Chem. Intermed. 2016, 42, 3385. [Google Scholar] [CrossRef]
  2. Yan, C.; Jia, J.; Liao, C.; Wu, S.; Xu, G. Rare earth separation in China. Tsinghua Sci. Technol. 2006, 11, 241–247. [Google Scholar] [CrossRef]
  3. Sathiyanarayanan, S.; Marikkannu, C.; Palaniswamy, N. Corrosion inhibition influence of tetramines for mild steel in 1M HCl. Appl. Surf. Sci 2005, 241, 477–484. [Google Scholar] [CrossRef]
  4. Trabanelli, G. 1991 Whitney Award Lecture: Inhibitors-An Old Remedy for a New Challenge. Corrosion 1991, 47, 410–419. [Google Scholar] [CrossRef]
  5. Lagrenee, M.; Mernari, B.; Bouanis, M.; Traisnel, M.; Bentiss, F. Study of the mechanism and inhibiting efficiency of 3,5-bis (4-methylthiophenyl)-4H-1, 2, 4-triazole on mild steel corrosion in acidic media. Corros. Sci. 2002, 44, 573–588. [Google Scholar] [CrossRef]
  6. Kosari, A.; Moayed, M.H.; Davoodi, A.; Parvizi, R.; Momeni, M.; Eshghi, H.; Moradi, H. Electrochemical and quantum chemical assessment of two organic compounds from pyridine derivatives as corrosion inhibitors for mild steel in HCl solution under stagnant condition and hydrodynamic flow. Corros. Sci. 2014, 78, 138–150. [Google Scholar] [CrossRef]
  7. Tawfik, S.M. Corrosion inhibition efficiency and adsorption behavior of N,N-dimethyl-4-(((1-methyl-2-phenyl-2,3-dihydro-1H-pyrazol-4-yl)imino)methyl)-N-alkylbenzenaminium bromide surfactant at carbon steel/hydrochloric acid interface. J. Mol. Liq. 2015, 207, 185. [Google Scholar] [CrossRef]
  8. Thomas, J.G.N. Some New Fundamental Aspects in Corrosion Inhibition; Corrosion Inhibitors: Ferrara, Italy, 1980; pp. 453–470. [Google Scholar]
  9. Attia, M.M.; Soliman, K.A.; Eid, S.; Mabrouk, E.M. Experimental and theoretical study on some azo chromotropic acid dyes compounds as inhibitor for carbon steel corrosion in sulfuric acid. J. Iran. Chem. Soc. 2021, 19, 655–664. [Google Scholar] [CrossRef]
  10. Eid, S. Expired Desloratidine Drug as Inhibitor for Corrosion of Carbon Steel Pipeline in Hydrochloric acid Solution. Int. J. Electrochem. Sci. 2021, 16, 150852. [Google Scholar] [CrossRef]
  11. Moussa, M.N.H.; El-Far, A.A.; El-Shafei, A.A. The use of water-soluble hydrazones as inhibitors for the corrosion of C-steel in acidic medium. Mater. Chem. Phys. 2007, 105, 105–113. [Google Scholar] [CrossRef]
  12. Wagdy, E.D.; Eid, S.; Zaher, A.A.; El-Etre, A.Y. Inhibition of carbon steel corrosion in aqueous solutions using some fatty amido-cationic surfactant. J. Basic Environ. Sci. 2016, 3, 55–64. [Google Scholar]
  13. Ruso, J.M.; Pérez, A.G.; Prieto, G.; Sarmiento, F. Study of the interactions between lysozyme and a fully-fluorinated surfactant in aqueous solution at different surfactant–protein ratios. Int. J. Biol. 2003, 33, 67–73. [Google Scholar] [CrossRef]
  14. Betih, M.A.; El-Henawy, S.B.; Al-Sabagh, A.M.; Negm, N.A.; Mahmoud, T. Experimental evaluation of cationic-Schiff base surfactants based on 5-chloromethyl salicylaldehyde for improving crude oil recovery and bactericide. J. Mol. Liq. 2020, 316, 113862. [Google Scholar] [CrossRef]
  15. Rajput, S.M.; Mondal, K.; Kuddushi, M.; Aswal, D.R.; Malek, N.I. Formation of hydrotropic drug/gemini surfactant based catanionic vesicles as efficient nano drug delivery vehicles. Colloids Interface Sci. Commun. 2020, 37, 100273. [Google Scholar] [CrossRef]
  16. Pakiet, M.; Kowalczyk, I.; Leiva Garcia, R.; Akid, R.; Brycki, B. Cationic clevelable surfactants as highly efficient corrosion inhibitors of stainless steel AISI 304: Electrochemical study. J. Mol. Liq. 2020, 315, 113675. [Google Scholar] [CrossRef]
  17. Shalabi, K.; Helmy, A.M.; El-Askalany, A.H.; Shahba, M.M. New pyridinium bromide mono-cationic surfactant as corrosion inhibitor for carbon steel during chemical cleaning: Experimental and theoretical studies. J. Mol. Liq. 2019, 293, 111480. [Google Scholar] [CrossRef]
  18. Abd El-Lateef, H.M.; Shalabi, K.; Tantawy, A.H. Corrosion inhibition and adsorption features of novel bioactive cationic surfactants bearing benzenesulphonamide on C1018-steel under sweet conditions: Combined modeling and experimental approaches. J. Mol. Liq. 2020, 320, 114564. [Google Scholar] [CrossRef]
  19. Godec, R.F.; Dolecek, V. Effect of sodium dodecylsulfate on the corrosion of copper in sulphuric acid media. Colloids Surf. A 2004, 244, 73–76. [Google Scholar] [CrossRef]
  20. Alfakeer, M.; Abdallah, M.; Abdel Hameed, R.S. Propoxylated Fatty Esters as Safe Inhibitors for Corrosion of Zinc in Hydrochloric Acid. Prot. Met. Phys. Chem. Surf. 2020, 56, 225–232. [Google Scholar] [CrossRef]
  21. Tang, F.; Wang, X.; Xu, X.; Li, L. Phytic acid doped nanoparticles for green anticorrosion coatings. Colloids Surf. A 2010, 369, 101–105. [Google Scholar] [CrossRef]
  22. Shi, Z.Z.; Song, C.X.; Ying Kan, Y.; Fan, X.S.; Zhang, Y. Inhibition Corrosion Effect of Extract from Basella rubra on Carbon Steel in HCl Solution. Int. J. Electrochem. Sci. 2020, 15, 4032–4055. [Google Scholar] [CrossRef]
  23. Fergachi, O.; Benhiba, F.; Rbaa, M.; Ouakki, M.; Galai, M.; Touir, R.; Lakhrissi, B.; Oudda, H.; Touhami, M.E. Corrosion Inhibition of Ordinary Steel in 5.0 M HCl Medium by Benzimidazole Derivatives: Electrochemical, UV–Visible Spectrometry, and DFT Calculations. J. Bio- Tribo-Corros. 2019, 5, 21. [Google Scholar] [CrossRef]
  24. Abd El-Lateef, H.M.; El-Beltagi, H.S.; Mohamed, M.E.M.; Kandeel, M.; Bakir, E.; Toghan, A.; Shalabi, K.; Tantawy, A.H.; Khalaf, M.M. Novel Natural Surfactant-Based Fatty Acids and Their Corrosion-Inhibitive Characteristics for Carbon Steel-Induced Sweet Corrosion: Detailed Practical and Computational Explorations. Front. Mater. 2022, 9, 843438. [Google Scholar] [CrossRef]
  25. Cerchiaro, G.; Aquilano, K.; Filomeni, G.; Rotilio, G.; Ciriolo, M.R.; Ferreira, A.M.D.C. Isatin-Schiff base copper(II) complexes and their influence on cellular viability. J. Inorg. Biochem. 2005, 99, 1433–1440. [Google Scholar] [CrossRef]
  26. Vancoa, J.; Svajlenova, O.; Racanskac, E.; Muselıka, J.; Valentova, J. Antiradical activity of different copper(II) Schiff base complexes and their effect on alloxan-induced diabetes. J. Trace Elem. Med. Biol. 2004, 18, 155–161. [Google Scholar] [CrossRef] [PubMed]
  27. Tebbji, K.; Hammouti, B.; Oudda, H.; Ramdani, A.; Benkadour, M. The inhibitive effect of bipyrazolic derivatives on the corrosion of steel in hydrochloric acid solution. Appl. Surf. Sci. 2005, 252, 1378–1385. [Google Scholar] [CrossRef]
  28. Prakash, M.; Moon, A.P.; Mondal, K.; Shekhar, S. Effect of machining configurations on the electrochemical response of mild steel in 3.5% NaCl solution. J. Mater. Eng. Perform. 2015, 24, 3643–3650. [Google Scholar] [CrossRef]
  29. Sorkhabia, H.A.; Shaabanib, B.; Seifzadeha, D. Corrosion inhibition of mild steel by some Schiff base compounds in hydrochloric acid. Appl. Surf. Sci. 2005, 239, 154–164. [Google Scholar] [CrossRef]
  30. Behpour, M.; Ghoreishi, S.M.; Salavati-Niasari, M.; Ebrahimi, B. Evaluating two new synthesized S–N Schiff bases on the corrosion of copper in 15% hydrochloric acid. Mater. Chem. Phys. 2008, 107, 153–157. [Google Scholar] [CrossRef]
  31. El Kacimi, Y.; Touir, R.; Galai, M.; Belakhmima, R.A.; Zarrouk, A.; Alaoui, K.; Harcharras, M.; El Kafssaoui, H.; Ebn Touhami, M. Effect of silicon and phosphorus contents in steel on its corrosion inhibition in 5 M HCl solution in the presence of Cetyltrimethylammonium/KI. J. Mater. Environ. Sci. 2016, 7, 371–381. Available online: https://www.jmaterenvironsci.com/Document/vol7/vol7_N1/40-JMES-2155-2015-El%20Kacimi.pdf (accessed on 15 January 2016).
  32. Dahmani, K.; Galai, M.; Ouakki, M.; Cherkaoui, M.; Touir, R.; Erkan, S.; Kaya, S.; El Ibrahimi, B. Quantum chemical and molecular dynamic simulation studies for the identification of the extracted cinnamon essential oil constituent responsible for copper corrosion inhibition in acidified 3.0 wt% NaCl medium. Inorg. Chem. Commun. 2021, 124, 108409. [Google Scholar] [CrossRef]
  33. Rbaa, M.; Galai, M.; Abousalem, A.S.; Lakhrissi, B.; Ebn Touhami, M.; Warad, I.; Zarrouk, A. Synthetic, spectroscopic characterization, empirical and theoretical investigations on the corrosion inhibition characteristics of mild steel in molar hydrochloric acid by three novel 8-hydroxyquinoline derivatives. Ionics 2020, 26, 503–522. [Google Scholar] [CrossRef]
  34. Negm, N.A.; Zaki, M.F.; Salem, M.A.I. Synthesis and Evaluation of 4-Diethyl Amino Benzaldehyde Schiff Base Cationic Amphiphiles as Corrosion Inhibitors for Carbon Steel in Different Acidic Media. J. Surfact. Deterg. 2009, 12, 321–329. [Google Scholar] [CrossRef]
  35. Kamboj, R.; Singh, S.; Bhadani, A.; Kataria, H.; Kaur, G. Gemini Imidazolium Surfactants: Synthesis and Their Biophysiochemical Study. Langmuir 2012, 28, 11969–11978. [Google Scholar] [CrossRef] [PubMed]
  36. Quagliotto, P.; Viscardi, G.; Barolo, C.; Barni, E.; Bellinvia, S.; Fisicaro, E.; Compari, C. Gemini Pyridinium Surfactants:  Synthesis and Conductometric Study of a Novel Class of Amphiphiles. J. Org. Chem. 2003, 68, 7651. [Google Scholar] [CrossRef] [PubMed]
  37. Zhong, X.; Guo, J.; Feng, L.; Xu, X.; Zhu, D. Cationic Gemini surfactants based on adamantane: Synthesis, surface activity and aggregation properties. Colloids Surf. A 2014, 441, 572–580. [Google Scholar] [CrossRef]
  38. Mousli, R.; Tazerouti, A. Synthesis and Some Surface Properties of Glycine-Based Surfactants. J. Surfactants Deterg. 2011, 14, 65–72. [Google Scholar] [CrossRef]
  39. Tantawy, A.H.; Soliman, K.A.; Abd El-Lateef, H.M. Novel synthesized cationic surfactants based on natural piper nigrum as sustainable-green inhibitors for steel pipeline corrosion in CO2-3.5% NaCl: DFT, Monte Carlo simulations and experimental approaches. J. Clean. Prod. 2020, 250, 119510. [Google Scholar] [CrossRef]
  40. de Queiroz Baddini, A.L.; Cardoso, S.P.; Hollauer, E.; Gomes, J.A.D.C.P. Statistical analysis of a corrosion inhibitor family on three steel surfaces (duplex, super-13 and carbon) in hydrochloric acid solutions. Electrochim. Acta 2007, 53, 434. [Google Scholar] [CrossRef]
  41. Negm, N.A.; Kandile, N.G.; Aiad, I.A.; Mohammad, M.A. New eco-friendly cationic surfactants: Synthesis, characterization and applicability as corrosion inhibitors for carbon steel in 1 N HCl. Colloids Surf. A 2011, 391, 224–233. [Google Scholar] [CrossRef]
  42. Abboud, Y.; Abourriche, A.; Saffaj, T.; Berrada, M.; Charrouf, M.; Bennamara, A.; Al Himidi, N.; Hannache, H. 2, 3-Quinoxalinedione as a novel corrosion inhibitor for mild steel in 1 M HCl. Mater. Chem. Phys. 2007, 105, 1–5. [Google Scholar] [CrossRef]
  43. Fouda, A.S.; Elewady, Y.A.; Abd El-Aziz, H.K.; Ahmed, A.M. Corrosion Inhibition of Carbon Steel in 0.5 M HCl Solution Using Cationic Surfactants. Int. J. Electrochem. Sci. 2012, 7, 10456–10475. [Google Scholar]
  44. Abd El-Lateef, H.M.; Tantawy, A.H. Synthesis and evaluation of novel series of Schiff base cationic surfactants as corrosion inhibitors for carbon steel in acidic/chloride media: Experimental and theoretical investigations. RSC Adv. 2016, 6, 8681–8700. [Google Scholar] [CrossRef]
  45. Hamitouche, H.; Khelifa, A.; Kouache, A.; Moulay, S. Petroleum quaternary ammonium surfactants mixture synthesized from light naphtha as corrosion inhibitors for carbon steel in 1 m HCl. Corros. Rev. 2013, 31, 61. [Google Scholar] [CrossRef]
  46. Abiola, O.K.; Otaigbe, J. Effect of common water contaminants on the corrosion of aluminium alloys in ethylene glycol–water solution. Corros. Sci. 2008, 50, 242–247. [Google Scholar] [CrossRef]
  47. Abdallah, M.; Kamar, E.M.; El-Etre, A.Y.; Eid, S. Gelatin as Corrosion Inhibitor for Aluminum and Aluminum Silicon Alloys in Sodium Hydroxide Solutions. Prot. Met. Phys. Chem. 2016, 52, 140–148. [Google Scholar] [CrossRef]
  48. Al Otaibi, N.; Hammud, H.H. Corrosion Inhibition Using Harmal Leaf Extract as an Eco-Friendly Corrosion Inhibitor. Molecules 2021, 26, 7024. [Google Scholar] [CrossRef]
  49. Seyam, D.F.; Tantawy, A.H.; Eid, S.; El-Etre, A.Y. Study of the inhibition effect of two novel synthesized amidoamine- based cationic surfactants on aluminum corrosion in 0.5 M HCl solution. J. Surfact. Deterg. 2021, 25, 133–143. [Google Scholar] [CrossRef]
  50. Frumkin, A. Surface tension curves of higher fatty acids and the equation of condition of the surface layer. Z. Phys. Chem. 1925, 116, 466–484. [Google Scholar] [CrossRef]
  51. Eddy, N.O.; Awe, F.; Ebenso, E.E. Adsorption and Inhibitive Properties of Ethanol Extracts of Leaves of Solanum Melongena for the Corrosion of Mild Steel in 0.1 M HCl. Int. J. Electrochem. Sci. 2010, 5, 1996–2011. Available online: http://www.electrochemsci.org/papers/vol5/5121996.pdf (accessed on 7 May 2023).
  52. Eid, S.; Hassan, W.M.I. Chemical and theoretical studies for corrosion inhibition of magnesium in hydrochloric acid by tween 80 surfactant. Int. J. Electrochem. Sci. 2015, 10, 8017–8027. Available online: http://www.electrochemsci.org/papers/vol10/101008017.pdf (accessed on 7 May 2023).
  53. Eddy, N.O.; Ita, B.I. Theoretical and experimental studies on the inhibition potentials of aromatic oxaldehydes for the corrosion of mild steel in 0.1 M HCl. J. Mol. Model. 2011, 17, 633–647. [Google Scholar] [CrossRef]
  54. Ateya, B.G.; El-Anadauli, B.E.; El-Nizamy, F.M. The adsorption of thiourea on mild steel. Corros. Sci. 1984, 24, 509–515. [Google Scholar] [CrossRef]
  55. Emranuzzaman, T.K.; Vishwanatham, S.; Udayabhanu, G. Synergistic effects of formaldehyde and alcoholic extract of plant leaves for protection of N80 steel in 15%HCl. Corros. Eng. Sci. Technol. 2004, 39, 327. [Google Scholar] [CrossRef]
  56. Haque, J.; Srivastava, V.; Quraishi, M.A.; Chauhan, D.S.; Lgaz, H.; Chung, I.M. Polar group substituted imidazolium zwitterions as eco-friendly corrosion inhibitors for mild steel in acid solution. Corros. Sci. 2020, 172, 108665. [Google Scholar] [CrossRef]
  57. Zhang, W.; Ma, Y.; Chen, L.J.; Wang, L.; Wu, Y.-C.; Li, H.-J. Aloe polysaccharide as an eco-friendly corrosion inhibitor for mild steel in simulated acidic oilfield water: Experimental and theoretical approaches. J. Mol. Liq. 2020, 307, 112950. [Google Scholar] [CrossRef]
  58. Dagdag, O.; Hsissou, R.; Berisha, A.; Erramli, H.; Hamed, O.; Jodeh, S.; El Harfi, A. Polymeric-Based Epoxy Cured with a Polyaminoamide as an Anticorrosive Coating for Aluminum 2024-T3 Surface: Experimental Studies Supported by Computational Modeling. J. Bio- Tribo-Corros. 2019, 5, 58. [Google Scholar] [CrossRef]
  59. Hsissou, R.; Abbout, S.; Berisha, A.; Berradi, M.; Assouag, M.; Hajjaji, N.; Elharfi, A. Experimental, DFT and molecular dynamics simulation on the inhibition performance of the DGDCBA epoxy polymer against the corrosion of the E24 carbon steel in 1.0 M HCl solution. J. Mol. Struct. 2019, 1182, 340. [Google Scholar] [CrossRef]
  60. Abd El-Lateef, H.M.; Shalabi, K.; Sayed, A.R.; Gomha, S.M.; Bakir, E.M. The novel polythiadiazole polymer and its composite with α-Al(OH)3 as inhibitors for steel alloy corrosion in molar H2SO4:Experimental and computational evaluations. J. Ind. Eng. Chem. 2022, 105, 238–250. [Google Scholar] [CrossRef]
  61. El-Lateef, H.M.A.; Abdallah, Z.A.; Ahmed, M.S.M. Solvent-free synthesis and corrosion inhibition performance of Ethyl 2-(1,2,3,6-tetrahydro-6-oxo-2-thioxopyrimidin-4-yl)ethanoate on carbon steel in pickling acids: Experimental, quantum chemical and Monte Carlo simulation studies. J. Mol. Liq. 2019, 296, 111800. [Google Scholar] [CrossRef]
  62. Alnajjar, A.O.; Abd El-Lateef, H.M.; Khalaf, M.M.; Mohamed, I.M.A. Steel protection in acidified 3.5% NaCl by novel hybrid composite of CoCrO3/polyaniline: Chemical fabrication, physicochemical properties, and corrosion inhibition performance. Constr. Build. Mater. 2022, 317, 125918. [Google Scholar] [CrossRef]
  63. Abd El-Lateef, H.M.; Shalabi, K.; Abdelhamid, A.A. 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. 2021, 334, 116081. [Google Scholar] [CrossRef]
  64. Abd El-Lateef, H.M.; Shalabi, K.; Tantawy, A.H. Corrosion inhibition of carbon steel in hydrochloric acid solution using newly synthesized urea-based cationic fluorosurfactants: Experimental and computational investigations. N. J. Chem. 2020, 44, 17791. [Google Scholar] [CrossRef]
  65. Dao, V.-D.; Vu, N.H.; Choi, H.-S. All day Limnobium laevigatum inspired nano generator self-driven via water evaporation. J. Power Sources 2020, 448, 227388. [Google Scholar] [CrossRef]
  66. Dao, V.-D. An experimental exploration of generating electricity from nature-inspired hierarchical evaporator: The role of electrode materials. Sci. Total Environ. 2021, 759, 143490. [Google Scholar] [CrossRef] [PubMed]
  67. El-Etre, A.Y. Inhibition of acid corrosion of carbon steel using aqueous extract of olive leaves. J. Colloid Interface Sci. 2007, 314, 578–583. [Google Scholar] [CrossRef]
  68. El-Etre, A.; El-Karim, I.G.; Ibrahim, S.; El-Kattan, F. Preparation of Amino nictino nitril and its application as corrosion inhibitor for carbon steel. J. Basic Environ. Sci. 2017, 4, 226–235. [Google Scholar]
  69. Sobhi, M.; Eid, S. Chemical, electrochemical and morphology studies on Methyl hydroxyethyl cellulose as green inhibitor for corrosion of copper in hydrochloric acid solutions. Prot. Met. Phys. Chem. Surf. 2018, 54, 893–898. [Google Scholar] [CrossRef]
  70. Abdallah, M.; El-Etre, A.Y.; Kamar, E.M.; Eid, S. Licorice root extract as corrosion inhibitor for tin in sodium chloride solutions. J. Environ. Sci. 2020, 7, 191–194. [Google Scholar]
  71. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results obtained with the correlation energy density functionals of becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200–206. Available online: http://www.lct.jussieu.fr/pagesperso/savin/papers/lyp/MieSavStoPre-89.pdf (accessed on 7 May 2023). [CrossRef]
  74. Perri, M.J.; Weber, S.H. Web-Based Job Submission Interface for the GAMESS Computational Chemistry Program. J. Chem. Educ. 2014, 91, 2206–2208. [Google Scholar] [CrossRef]
  75. Barca, G.M.J.; Bertoni, C.; Carrington, L.; Datta, D.; De Silva, N.; Deustua, J.E.; Włoch, M.; Xu, P.; Zahariev, F.; Gordon, M.S. Recent developments in the general atomic and molecular electronic structure system. J. Chem. Phys. 2020, 152, 154102. [Google Scholar] [CrossRef] [Green Version]
  76. Abdallah, M.; Soliman, K.A.; Alfattani, R.; Al-Gorair, A.S.; Fawzy, A.; Ibrahim, M.A. Insight of corrosion mitigation performance of SABIC iron in 0.5 M HCl solution by tryptophan and histidine: Experimental and computational approaches. Int. J. Hydrog. Energy 2022, 47, 12782–12797. [Google Scholar] [CrossRef]
Scheme 1. Synthetic path for the synthesis of two novel Imine-tethering cationic Surfactants (ICS-10 and ICS-14). Reagents and conditions: (a) Schiff’s base formation; absolute ethanol, 7-8 h at 85 °C, 93%; (b) quaternization reaction; ethyl acetate as a solvent at 70 °C for 30 h; 77-81%.
Scheme 1. Synthetic path for the synthesis of two novel Imine-tethering cationic Surfactants (ICS-10 and ICS-14). Reagents and conditions: (a) Schiff’s base formation; absolute ethanol, 7-8 h at 85 °C, 93%; (b) quaternization reaction; ethyl acetate as a solvent at 70 °C for 30 h; 77-81%.
Molecules 28 04540 sch001
Figure 1. Variation of surface tension (A,B) in HCl solution and aqueous medium, respectively, and (C) specific conductivity against the concentration of the imine-tethering cationic surfactants (ICS-10 and ICS-14) at 25 °C.
Figure 1. Variation of surface tension (A,B) in HCl solution and aqueous medium, respectively, and (C) specific conductivity against the concentration of the imine-tethering cationic surfactants (ICS-10 and ICS-14) at 25 °C.
Molecules 28 04540 g001
Figure 2. The effects of the addition of various concentrations of imine-tethering cationic surfactants (ICS-10 and ICS-14) on the weight loss (A) and the inhibition capacity (B) of carbon steel in the 1.0 M HCl solution.
Figure 2. The effects of the addition of various concentrations of imine-tethering cationic surfactants (ICS-10 and ICS-14) on the weight loss (A) and the inhibition capacity (B) of carbon steel in the 1.0 M HCl solution.
Molecules 28 04540 g002
Figure 3. PDP curves for C-steel electrode in the 1.0 M HCl solution containing different concentrations of (A) ICS-10 and (B) ICS-14 at a scan rate of 1.0 mV/s.
Figure 3. PDP curves for C-steel electrode in the 1.0 M HCl solution containing different concentrations of (A) ICS-10 and (B) ICS-14 at a scan rate of 1.0 mV/s.
Molecules 28 04540 g003
Figure 4. SEM for C-steel coupons after immersion in (A) blank, 1.0 M of HCl, (B) 1 M of HCl + 0.0005 M of ICS-10, and (C) 1 M of HCl + 0.0005 M of ICS-14.
Figure 4. SEM for C-steel coupons after immersion in (A) blank, 1.0 M of HCl, (B) 1 M of HCl + 0.0005 M of ICS-10, and (C) 1 M of HCl + 0.0005 M of ICS-14.
Molecules 28 04540 g004
Figure 5. Langmuir’s isotherm curves for the adsorption of (A) ICS-10 and (B) ICS-14 on the C-steel surface in 1.0 M of HCl.
Figure 5. Langmuir’s isotherm curves for the adsorption of (A) ICS-10 and (B) ICS-14 on the C-steel surface in 1.0 M of HCl.
Molecules 28 04540 g005
Figure 6. The optimized structure, frontier molecular orbitals (HOMO and LUMO), and ESP for ICS-10 and ICS-14 molecules.
Figure 6. The optimized structure, frontier molecular orbitals (HOMO and LUMO), and ESP for ICS-10 and ICS-14 molecules.
Molecules 28 04540 g006
Figure 7. Top and side view of adsorbed (A) ICS-10 and (B) ICS-14 inhibitors on Fe (110) surface.
Figure 7. Top and side view of adsorbed (A) ICS-10 and (B) ICS-14 inhibitors on Fe (110) surface.
Molecules 28 04540 g007
Figure 8. (AC) IR, 1H, and 13C NMR spectra of ((2-chlorobenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(tetradecyloxy) ethyl) propan-1-ammonium chloride [ICS-14].
Figure 8. (AC) IR, 1H, and 13C NMR spectra of ((2-chlorobenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(tetradecyloxy) ethyl) propan-1-ammonium chloride [ICS-14].
Molecules 28 04540 g008
Table 1. Surface-active properties of the prepared cationic imine surfactants (ICS-10 and ICS-14) from the surface tension and conductivity measurements at 25 °C.
Table 1. Surface-active properties of the prepared cationic imine surfactants (ICS-10 and ICS-14) from the surface tension and conductivity measurements at 25 °C.
Comp.Surface Tension MeasurementsConductivity Measurements
CMC a
(mM L−1)
CMC b
(mM L−1)
γCMC
(mN/m)
ΠCMC
(mN/m)
Γmax × 1011
(mol cm−2)
Amin
(nm2)
CMC
(mM L−1)
αβ Δ G m i c o (KJ mol−1) Δ G a d s o (KJ mol−1)
ICS-101.340.9536.1735.837.311.881.170.440.56−26.66−30.72
ICS-140.830.7934.4637.546.112.680.860.490.51−26.53−32.29
a CMC values obtained from surface tension measurements in distilled water; b CMC values obtained from surface tension measurements in 1.0 M of HCl.
Table 2. Corrosion parameters obtained from PDP of C-steel electrode in 1 M HCl solution containing different concentrations of ICS-10 and ICS-14 surfactants.
Table 2. Corrosion parameters obtained from PDP of C-steel electrode in 1 M HCl solution containing different concentrations of ICS-10 and ICS-14 surfactants.
Inhibitor CodeInh. Conc.
(M)
βa
mV·Decade−1
βc
mV·Decade−1
Ecor
mV (SCE)
icor
µA/cm2
I.E.%θ
-063214−379720--
ICS-105 × 10−662221−37336749.020.4902
1 × 10−556237−37234651.940.5194
5 × 10−550208−37417076.390.7639
1× 10−446242−37110285.830.8583
5 × 10−455262−3806191.530.9153
ICS-145 × 10−651220−36420371.810.7181
1 × 10−550212−38412682.500.8250
5 × 10−545212−3778488.330.8833
1× 10−439188−3864793.470.9347
5 × 10−443188−3973994.580.9458
Table 3. The values of the corrosion rate, activation energy, and heat of adsorption of carbon steel in 1 M of HCl and 0.0005 M of ICS-10 and ICS-14.
Table 3. The values of the corrosion rate, activation energy, and heat of adsorption of carbon steel in 1 M of HCl and 0.0005 M of ICS-10 and ICS-14.
SolutionCR1
293 K
CR2
333 K
θ1
293 K
θ2
333 K
Ea
kJ·mol−1
Qads
kJ·mol−1
Blank1.671 × 10−44.018 × 10−3--64.498-
ICS-101.504 × 10−53.221 × 10−30.9100.198108.858−75.276
ICS-149.758 × 10−62.87 × 10−30.9410.285115.290−74.980
Table 4. Quantum parameters of ICS-10 and ICS-14 inhibitor molecules.
Table 4. Quantum parameters of ICS-10 and ICS-14 inhibitor molecules.
ParametersICS-10ICS-14
EHOMO (eV)−4.71−4.69
ELUMO (eV)−1.33−1.34
ΔE (eV)3.383.35
μ (Debye)13.2313.31
η ( e V ) 1.691.68
σ ( e V ) 1 0.590.60
ΔN (e)0.530.54
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abd El-Lateef, H.M.; Tantawy, A.H.; Soliman, K.A.; Eid, S.; Abo-Riya, M.A. Novel Imine-Tethering Cationic Surfactants: Synthesis, Surface Activity, and Investigation of the Corrosion Mitigation Impact on Carbon Steel in Acidic Chloride Medium via Various Techniques. Molecules 2023, 28, 4540. https://doi.org/10.3390/molecules28114540

AMA Style

Abd El-Lateef HM, Tantawy AH, Soliman KA, Eid S, Abo-Riya MA. Novel Imine-Tethering Cationic Surfactants: Synthesis, Surface Activity, and Investigation of the Corrosion Mitigation Impact on Carbon Steel in Acidic Chloride Medium via Various Techniques. Molecules. 2023; 28(11):4540. https://doi.org/10.3390/molecules28114540

Chicago/Turabian Style

Abd El-Lateef, Hany M., Ahmed H. Tantawy, Kamal A. Soliman, Salah Eid, and Mohamed A. Abo-Riya. 2023. "Novel Imine-Tethering Cationic Surfactants: Synthesis, Surface Activity, and Investigation of the Corrosion Mitigation Impact on Carbon Steel in Acidic Chloride Medium via Various Techniques" Molecules 28, no. 11: 4540. https://doi.org/10.3390/molecules28114540

Article Metrics

Back to TopTop