Skip to main content
ACS Omega logoLink to ACS Omega
. 2022 Nov 15;7(47):42975–42993. doi: 10.1021/acsomega.2c05192

Investigation of Sulfonium-Iodide-Based Ionic Liquids to Inhibit Corrosion of API 5L X52 Steel in Different Flow Regimes in Acid Medium

Víctor Díaz-Jiménez , Paulina Arellanes-Lozada , Natalya V Likhanova ‡,*, Octavio Olivares-Xometl †,*, Ernesto Chigo-Anota , Irina V Lijanova §, Giselle Gómez-Sánchez , F Verpoort
PMCID: PMC9713877  PMID: 36467960

Abstract

graphic file with name ao2c05192_0019.jpg

The present work deals with the corrosion inhibition mechanism of API 5L X52 steel in 1 M H2SO4 employing the ionic liquid (IL) decyl(dimethyl)sulfonium iodide [DDMS+I]. Such a mechanism was elicited by the polarization resistance (Rp), potentiodynamic polarization (PDP), and electrochemical impedance spectroscopy (EIS) techniques, both in stationary and dynamic states. The electrochemical results indicated that the corrosion inhibition was controlled by a charge transfer process and that the IL behaved as a mixed-type corrosion inhibitor (CI) with anodic preference. The experimental results revealed maximal inhibition efficiency (IE) rates up to 93% at 150 ppm in the stationary state, whereas in turbulent flow, the IE fell to 65% due to the formation of microvortexes that promoted higher desorption of IL molecules from the surface. The Gibbs free energy of adsorption (ΔG°ads) value of −34.89 kJ mol–1, obtained through the Langmuir isotherm, indicated the formation of an IL monolayer on the metal surface by combining physisorption and chemisorption. The surface analysis techniques confirmed the presence of FexOy, FeOOH, and IL on the surface and showed that corrosion damage diminished in the presence of IL. Furthermore, the quantum chemistry calculations (DFT) indicated that the iodide anion hosted most of the highest occupied molecular orbital (HOMO), which eased its adsorption on the anodic sites, preventing the deposition of sulfate ions on the electrode surface.

1. Introduction

In the oil industry, steel corrosion is regarded as one of the most serious problems to be dealt with due to the exposure of this metallic material to different conditions throughout a variety of industrial operations. In particular, during the extraction and transport of oil, the formation of acid media and flow conditions increase the corrosion rate, which provokes the degradation of metallic structures, and therefore, pipeline failures and damage of equipment and installations, which are part of different processes as a result of the shortening of the useful life of steel pieces.1,2 Diverse methods conceived to control corrosion at the industrial level have been put into practice; among these strategies, corrosion inhibitors (CIs) have been employed, which have been accepted willingly for their low cost and easy application with different alloys and corrosive media.3,4 CIs control the corrosion reactions that take place on the metal by an adsorption process that is affected by different factors such as the nature and surface charge of the metal, electronic structure, steric factors, aromaticity, functional groups with double and triple bonds, and high-density heteroatoms (N, O, S, and P) present in the inhibiting organic molecules.5,6

Inorganic compounds like chromates, nitrates, molybdates, phosphates, silicates, and arsenic have been used as CIs, which when combined with the oxide layer provide the metal surface with passivation protection.7 Notwithstanding, these compounds represent potential danger for the health, for they are highly toxic chemicals.8 Currently, environmental regulations concerning the use of chemical substances have limited the application of toxic compounds, which has led to the synthesis of low-toxicity CIs that are easy to handle and biodegradable, i.e., more environmentally friendly like the compounds known as ILs.9,10 These molecules are salts consisting of organic or inorganic anions and organic cations that display, in general, melting points below 100 °C, i.e., these are liquids at ambient temperature and possess unique features that stem from the arrangement and chemical distribution of their ions; in addition, the ILs have negligible vapor pressure, thermal stability, nonflammability, high ionic conductivity, and wide electrochemical stability spectrum.1113

In the literature, the inhibiting properties in acid medium of ILs with imidazolium,14,15 pyridinium,16,17 ammonium,1820 pyridazinium,21,22 and sulfonium2325 cations have been reported. These last compounds have been studied in sulfamic,23 phosphoric,24 and hydrochloric and sulfuric25 acids, reporting IE values from 84 to 98%. Their inhibition behavior has been attributed to the nature of the substituent groups: the presence of π electrons that are capable of interacting with the negative charge of the metal, aliphatic chains that promote hydrophobic properties, but mainly due to the capacity of S to accept electrons.

The iodide anion effect has been the subject matter of several studies, and it has been reported that the IE values of ILs with iodide anion have been higher than those displayed by other anions, as shown in Table 1. In this context, in 2008, Verma et al. analyzed ILs with the cation 2-hydroxyethyl-trimethyl-ammonium and three different anions, finding that the IL with iodide anion had higher IE values than those with acetate and chloride anions.26 The anion influence on the inhibition process was studied by Mashuga et al. comparing ILs based on 1-hexyl-3-methylimidazolium with four different anions. It was reported that the IE obtained with the IL featuring the iodide anion (79%) or trifluoromethanesulfonate (81%) was slightly higher than those values given by fluoro-substituted anions such as tetrafluoroborate (78%) or hexafluorophosphate (73%).27 In 2018, Azeez et al. suggested that the presence of the I anion as an electron donor increased the electron density of the cation heteroatoms (−C=N−) in ILs based on 1-alkyl-3-methylimidazolium iodide.28

Table 1. State of the Art of Sulfonium and Iodide Compounds Used as CIs.

1.

In addition, iodide has also been employed in acid corrosion as synergistic agent in inhibition processes of organic compounds. Cao et al. and Guo et al. showed that the IE of 1,1′[1,4-phenylenebis(methylene)] bis[3(carboxymethyl)imidazolium] chloride and l-tryptophane, respectively, increased significantly in the presence of KI as a synergistic agent.29,30 Cao et al. indicated that the addition of KI could provide better IE by decreasing the IL concentration.29 Moreover, Guo et al. suggested that iodide adsorbed on a metal surface tended to form joining bridges between organic cations and the positively charged iron surface, thus stabilizing the CI adsorption process;30 similar conclusions were reported by Feng et al.32 Additionally, Farag et al.31 indicated that iodide could also promote a change in potential of the metal surface toward more negative values, facilitating the adsorption of positively charged species. According to the aforementioned, the inhibition properties that distinguish iodide from the halide group are its higher capacity to donate electrons, easy polarization, high hydrophobicity, greater atomic radius, and low electronegativity.33

In addition to the effect exerted by the chemical structure on the corrosion inhibition mechanism, the medium conditions are essential factors that affect the performance of the CIs. Due to the multiple transfer phenomena of corrosive species and CI molecules in the metal–CI interphase, the proposal of a corrosion inhibition mechanism in the stationary state is complex. In addition, in the presence of flow, the adsorption mechanism is affected by factors such as sheer stress (τRDE) and formations of microvortexes that modify the adsorption and desorption of CI molecules and corrosive species. In contrast, the chemical structure of CIs is fundamental for the formation of stable films under flow conditions. For this reason, in the present work, a novel IL was synthesized, characterized, and evaluated in its inhibiting properties: decyl(dimethyl)sulfonium iodide [DDMS+I]. The IL was evaluated as CI of API 5L X52 steel under stationary and flow conditions [Reynolds numbers (NRE) from 1000 to 5000] in 1 M H2SO4 solution. To analyze the inhibiting effect exerted by [DDMS+I], the Rp, potentiodynamic polarization (PDP), and electrochemical impedance spectroscopy (EIS) electrochemical techniques were used. Afterward, the obtained data were processed by different adsorption isotherm models. Scanning electron microscopy/energy-dispersive X-ray spectroscopy (SEM/EDS), diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), and X-ray photoelectron spectroscopy (XPS) surface analyses complemented the electrochemical results. Finally, to describe the behavior of the [DDMS+I] molecule, molecular simulation at DFT B3LYP/MIDIx theory level was carried out.

2. Results and Discussion

2.1. Electrochemical Tests

2.1.1. EOCP as a Function of Time

Figure 1 shows the open-circuit potential (EOCP) values for API 5L X52 steel in 1 M H2SO4 in the absence and presence of different concentrations of [DDMS+I]. It can be observed that the EOCP values were displaced toward more positive values with increasing IL concentration with respect to the blank. These EOCP displacements because of IL addition indicate both the adsorption of IL molecules on the metal surface and the formation of an inhibiting film. Furthermore, it was found that the EOCP stability started from 800 s.

Figure 1.

Figure 1

EOCP plot as a function of time of API 5L X52 steel in 1 M H2SO4 with and without IL.

2.1.2. Polarization Resistance

Figure 2 shows the Rp behavior as a function of: (a) the [DDMS+I] concentrations and (b) different NRE. In general, it is observed that the slope of the current density (i) vs overpotential (η) diminished with the CI concentration (C), indicating that the presence of IL increased the Rp of the metal–electrolyte interphase due to the adsorption of inhibiting molecules from the formation of a layer on the steel surface in 1 M H2SO4. This inhibiting film produced resistance to the flow of electrons participating in the redox reactions, generating changes in the kinetics of the electrochemical reactions involved in the corrosion process. Furthermore, the slope increase at different NRE in the absence and presence of IL can be observed, which indicates that the diffusive processes under flow conditions modified the IL adsorption and interfacial properties of the metal–electrolyte system.

Figure 2.

Figure 2

Rp of API 5L X52 steel in 1 M H2SO4 with (a) different [DDMS+I] concentrations at the stationary state and (b) with 150 ppm of [DDMS+I] under flow conditions at 298 K.

Table 2 displays the Rp results and it can be seen that the maximal values were obtained at C above 100 ppm of CI and at the stationary state. The IERp was calculated from eq 13437

2.1.2. 1

where the superindexes In h and 0 refer to the experiments in the presence and absence of CI, respectively.

Table 2. Electrochemical Parameters Obtained by the Rp Technique of API 5L X52 Steel in 1 M H2SO4 with Different Concentrations of [DDMS+I].
system NRE C(ppm mM–1) Ecorr (mV) Rp(Ω cm2) IERp (%)
blank 0 0/0 466 ± 0 50 ± 1  
1000 0/0 437 ± 0 44 ± 0  
2000 0/0 444 ± 0 38 ± 2  
3000 0/0 436 ± 0 23 ± 1  
4000 0/0 439 ± 0 16 ± 0  
5000 0/0 439 ± 0 16 ± 0  
[DDMS+I] 0 25/0.076 444 ± 3 95 ± 2 47.14 ± 1.78
0 50/0.151 426 ± 4 371 ± 7 86.43 ± 0.45
0 75/0.227 429 ± 5 473 ± 24 89.30 ± 0.34
0 100/0.303 418 ± 6 699 ± 35 92.51 ± 0.22
0 125/0.378 420 ± 4 673 ± 18 92.42 ± 0.23
0 150/0.454 419 ± 3 674 ± 17 92.56 ± 0.22
1000 150/0.454 394 ± 1 317 ± 0 85.98 ± 0.01
2000 150/0.454 381 ± 1 278 ± 1 86.28 ± 0.06
3000 150/0.454 376 ± 0 151 ± 1 84.49 ± 0.13
4000 150/0.454 380 ± 4 87 ± 2 81.14 ± 0.48
5000 150/0.454 388 ± 3 48 ± 1 67.36 ± 0.79

2.1.3. Potentiodynamic Polarization

Figure 3a,b presents the PDP behavior as a function of C and different NRE. It is observed that the increase in C caused the displacement of the potentiodynamic curves toward low current density values with respect to the blank. Under flow conditions, the NRE increase intensified the corrosion process in the absence and presence of IL, displacing the potentiodynamic curves toward higher current density values, which indicates more serious surface damage in the presence of IL.

Figure 3.

Figure 3

Potentiodynamic polarization of API 5L X52 steel in 1 M H2SO4 with (a) different [DDMS+I] concentrations at the stationary state and (b) with 150 ppm of [DDMS+I] under flow conditions at 298 K.

At the stationary state (Table 2), lower current density values in the presence of CI are reported; this fact can be attributed to the growing blocking of active sites as a function of the [DDMS+I] concentration, which replaced water molecules and aggressive ions such as H+, H3O+, O2–, and SO42– on the metallic substrate. The optimal adsorption process of the CI molecules on the metal surface occurred at a concentration of 100 ppm; in contrast, higher i values were obtained under flow conditions due to the wall shear stress (τRDE) increase. This behavior pattern was originated by the flow regime change, which provoked higher desorption rate of the CI film as NRE was increased and then, the progressive diminution of IE. The IEPDP was calculated by eq 234,35

2.1.3. 2

where the superindexes Inh and 0 refer to the experiments in the presence and absence of CI, respectively.

At the stationary state, the reported corrosion potential (Ecorr) displacement values with respect to the blank fall within the interval ranging from 43 to 65 mV, i.e., within the ±85 mV interval, which indicates that [DDMS+I] matches the behavior pattern of a mixed-type CI, retarding both the cathodic and anodic reactions occurring during the corrosion process38,39 with anodic preference due to the displacement toward positive potentials (Ecorr0 < EcorrInh).40 This phenomenon is related to the preferential adsorption of [I] and the contribution of the [DDMS+] cations on the anodic and cathodic sites, respectively, which diminished the attack of water molecules and aggressive ions in the medium. Under flow conditions, the IL anodic preference was not affected sensibly. Notwithstanding, the βc values under flow conditions, reported in Table 3, are slightly higher than those obtained at the stationary state. This increase suggests that the IL modified the kinetics of the reduction reactions, thus retarding the hydrogen evolution by CI adsorption. Then, it can be concluded that the contribution by the IL anions and cations to the inhibition process is the same.

Table 3. Electrochemical Parameters Obtained by PDP Curves of API 5L X52 steel in 1 M H2SO4 with Different Concentrations of [DDMS+I].
system NRE C(ppm mM–1) Ecorr (mV) βc(mV dec–1) icorr(μA cm–2) Vcorr(mm year–1) IEPDP (%)
blank 0 0/0 466 ± 1 108 ± 1 300.18 ± 4.23 3.48 ± 0.04  
1000 0/0 437 ± 0 164 ± 31 155.6 ± 3.05 6.25 ± 0.12  
2000 0/0 444 ± 0 141 ± 12 173.76 ± 19.31 6.98 ± 0.15  
3000 0/0 436 ± 0 126 ± 10 322.32 ± 3.63 12.94 ± 0.15  
4000 0/0 439 ± 0 108 ± 3 334.14 ± 17.77 14.35 ± 0.10  
5000 0/0 439 ± 0 121 ± 13 482.82 ± 10.78 19.39 ± 0.43  
[DDMS+I] 0 25/0.076 444 ± 3 128 ± 8 170.99 ± 4.71 1.98 ± 0.05 43.03 ± 1.76
0 50/0.151 426 ± 4 126 ± 2 64.42 ± 1.23 0.74 ± 0.01 78.53 ± 0.51
0 75/0.227 429 ± 5 104 ± 3 48.27 ± 2.80 0.56 ± 0.03 83.92 ± 0.96
0 100/0.303 418 ± 6 80 ± 6 27.11 ± 1.07 0.31 ± 0.01 90.97 ± 0.38
0 125/0.378 420 ± 4 78 ± 7 24.67 ± 0.54 0.28 ± 0.01 91.78 ± 0.21
0 150/0.454 419 ± 3 92 ± 10 23.89 ± 0.12 0.27 ± 0.01 92.04 ± 0.12
1000 150/0.454 394 ± 1 207 ± 15 24.47 ± 0.17 0.98 ± 0.01 84.27 ± 0.11
2000 150/0.454 381 ± 1 172 ± 21 26.25 ± 0.59 1.05 ± 0.01 84.89 ± 0.34
3000 150/0.454 376 ± 0 176 ± 4 50.91 ± 0.08 2.04 ± 0.01 84.20 ± 0.03
4000 150/0.454 380 ± 4 224 ± 1 74.56 ± 0.01 2.99 ± 0.01 77.69 ± 0.01
5000 150/0.454 388 ± 3 216 ± 14 117.06 ± 8.63 4.70 ± 0.14 75.76 ± 3.58

2.1.4. Electrochemical Impedance Spectroscopy

Figure 4 shows the Nyquist impedance curves at stationary and dynamic states and in the absence and presence of IL, where the formation of depreciated semicircles can be observed, which is attributed to the surface roughness and heterogeneity. In contrast, in the presence of [DDMS+I], an increase in the semicircle diameter as a function of C can be observed, which indicates that the corrosion process was controlled by the charge transfer resistance through the interface;41 this behavior pattern occurred at both stationary and dynamic states. In contrast, the diameter decrease in the semicircles as NRE increased indicates that the flow made the charge transfer and desorption rate of the species adsorbed on the metallic surface grow.42

Figure 4.

Figure 4

Nyquist impedance curves recorded for API 5L X52 steel in 1 M H2SO4 solution with (a) different [DDMS+I] concentrations at the stationary state and (b) with 150 ppm of [DDMS+I] under flow conditions at 298 K.

In the absence of IL and the stationary state, the formation of an inductive loop in the low-frequency region can be observed, which can be attributed to intermediate reactions related to the adsorption of water, aggressive ions (H+, H3O+, O2–, and SO42–) and/or insoluble solution corrosion products.43,44 This behavior pattern was also observed at the dynamic state in the absence and presence of IL, which was caused by diffusive processes of the electrolyte in the metal/IL interface. Nevertheless, in the presence of [DDMS+I] at the stationary state, the inductive loop disappeared, which suggests the stabilization of the charge transfer processes because of the formation of a more homogeneous and stable inhibiting film than the one produced at the dynamic state.

Figure 5 shows the Bode plots at stationary and dynamic states. In the presence of IL, a higher displacement of the |Z| values and phase angle than those obtained with the blank can be observed, which suggests a reduction of the adsorption of aggressive ions, thus stabilizing the surface. Furthermore, the diagrams show a maximal point for the phase angle at intermediate frequencies, which indicates the existence of a relaxation time constant (τdl) associated with the charge transfer resistance (Rct) and capacitance of the electric double layer (Cdl).43 At the stationary state, the presence of another relaxation time constant is also observed, which is related to a capacitive process due to the accumulation of corrosion products/CI, despite the presence of the IL, the corrosive medium ions are adsorbed on the surface causing their damage.

Figure 5.

Figure 5

Bode plots recorded for API 5L X52 steel in 1 M H2SO4 solution with (a) different [DDMS+I] concentrations at the stationary state and (b) with 150 ppm of [DDMS+I] under flow conditions at 298 K.

The equivalent electrical circuits (EECs) employed to fit the experimental data from the impedance spectra are shown in Figure 6.4548 What follows is the description of the physical interpretation of the EEC elements: Rs represents the solution resistance; Rct is the resistance to the charge transfer on the surface, which depends on the charge transfer between the electronic conduction region (metal) and the ionic conduction region (solution); CPEdl is a constant phase element (CPE) related to the Cdl, which suggests the accumulation of charge in that interphase; RL is an inductive resistance and L is an inductor and both elements represent the relaxation process of the intermediates involved in the oxidation reactions (H+, H3O+, O2–, and SO42–); and CPEf and Rf are the capacitance and resistance related to the formation of corrosion products and/or the IL film on the metal surface.

Figure 6.

Figure 6

Equivalent electrical circuits of the steel corrosion process in 1 M H2SO4 solution: (a) under stationary and flow conditions, (b) at different [DDMS+I] concentrations at the stationary state, and (c) with 150 ppm of [DDMS+I] under flow conditions.

The constant phase elements (CPEs) in Figure 6 represent the nonideal capacitances of the EIS spectra, which were calculated with eq 3

2.1.4. 3

where Y0 is the proportional factor and n is an empirical exponent between 0 and 1, which is related to the surface heterogeneity, where n values close to 1 are characteristic of homogeneous surfaces.49

Tables 4 and 5 show the EIS parameters at stationary and dynamic conditions, respectively. It is observed that the Rs values do not present variation (ΔRs ∼ 1 Ω·cm2) at different [DDMS+I] concentrations and flow conditions. Likewise, at the stationary state, the Rf and Cf values do not show significant changes in the inhibitor film–solution interphase with increasing [DDMS+I] concentration.

Table 4. EIS Parameters Obtained for API 5L X52 Steel in 1 M H2SO4 without and with Different Concentrations of [DDMS+I] at the Stationary State.
    C(ppm mM–1)
parameter blank 25/0.076 50/0.151 75/0.227 100/0.303 125/0.378 150/0.454
χ2 0.158 ± 0.004 0.193 ± 0.010 0.214 ± 0.012 0.216 ± 0.012 0.188 ± 0.035 0.187 ± 0.018 0.186 ± 0.030
Rs(Ω·cm2) 0.68 ± 0.25 1.00 ± 0.09 1.17 ± 0.10 1.13 ± 0.14 1.27 ± 0.10 1.26 ± 0.23 1.16 ± 0.26
Rf(Ω·cm2) 1.47 ± 0.31 1.73 ± 0.10 1.91 ± 0.11 1.88 ± 0.01 1.98 ± 0.10 1.98 ± 0.18 1.87 ± 0.23
Cf(μF cm–2) 0.09 ± 0.03 0.31 ± 0.03 0.25 ± 0.03 0.26 ± 0.03 0.21 ± 0.02 0.18 ± 0.06 0.14 ± 0.04
RL(Ω·cm2) 277.07 ± 46.04            
L(H·cm2) 29.72 ± 2.84            
Rct(Ω·cm2) 47.86 ± 3.61 278.21 ± 9.04 353.77 ± 28.29 620.84 ± 23.96 841.84 ± 27.11 734.82 ± 17.05 759.31 ± 24.14
Y0·10–6 (S·sn cm–2) 281.05 ± 21.39 70.84 ± 2.80 80.22 ± 5.83 49.61 ± 6.51 39.16 ± 6.98 46.45 ± 7.49 62.55 ± 16.46
n 0.86 ± 0.01 0.88 ± 0.01 0.88 ± 0.01 0.89 ± 0.01 0.89 ± 0.01 0.89 ± 0.01 0.89 ± 0.01
Cdl(μF cm–2) 140.46 ± 4.08 41.42 ± 1.28 50.04 ± 3.58 31.87 ± 3.02 25.61 ± 3.13 29.87 ± 3.78 42.06 ± 7.89
τdl (ms) 6.72 ± 0.54 11.52 ± 0.52 17.70 ± 1.90 19.79 ± 2.02 21.56 ± 2.72 21.95 ± 2.82 31.94 ± 6.08
Rp(Ω·cm2) 42.96 ± 2.83 280.94 ± 9.04 356.85 ± 28.29 623.85 ± 23.96 845.09 ± 27.11 738.06 ± 17.05 762.34 ± 24.14
IEEIS (%)   84.70 ± 1.12 87.96 ± 1.24 93.11 ± 0.52 94.91 ± 0.37 94.18 ± 0.40 94.36 ± 0.41
Table 5. EIS Parameters Obtained for API 5L X52 Steel in 1 M H2SO4 with 150 ppm of [DDMS+I] in Flow Conditions.
regimen laminar
transitory
turbulent
NRE 1000 2000 3000 4000 5000
C(ppm mM–1) 0 150/0.454 0 150/0.454 0 150/0.454 0 150/0.454 0 150/0.454
χ2 0.158 ± 0.005 0.143 ± 0.005 0.196 ± 0.028 0.139 ± 0.014 0.121 ± 0.014 0.148 ± 0.006 0.075 ± 0.001 0.123 ± 0.004 0.056 ± 0.008 0.081 ± 0.002
Rs(Ω·cm2) 1.15 ± 0.08 0.55 ± 0.05 1.11 ± 0.18 0.91 ± 0.30 1.56 ± 0.02 0.78 ± 0.44 2.09 ± 0.51 1.59 ± 0.22 1.75 ± 0.14 1.03 ± 0.18
Rf(Ω·cm2) 1.91 ± 0.07   1.87 ± 0.16   2.28 ± 0.01   3.15 ± 0.40   3.93 ± 1.75  
Cf(μF cm–2) 0.38 ± 0.01   0.42 ± 0.10   0.37 ± 0.01   0.149 ± 0.05   0.149 ± 0.07  
RL(Ω·cm2) 130.77 ± 24.77 987.63 ± 73.62 108.86 ± 23.50 962.36 ± 88.12 91.09 ± 11.54 587.13 ± 55.43 90.24 ± 13.93 294.38 ± 39.74 74.28 ± 10.35 123.07 ± 18.89
L(H·cm2) 41.03 ± 0.71 266.48 ± 11.81 27.89 ± 5.23 230.37 ± 11.76 15.66 ± 1.92 124.71 ± 18.83 16.28 ± 2.94 57.79 ± 16.245 10.78 ± 1.56 40.05 ± 9.31
Rct(Ω·cm2) 28.71 ± 4.35 261.47 ± 4.83 26.78 ± 3.79 225.64 ± 20.8 22.02 ± 1.67 170.15 ± 7.57 22.69 ± 3.8 174.65 ± 6.79 17.37 ± 2.60 101.49 ± 7.49
Y0·10–6 (S·sn cm–2) 296.13 ± 1.11 55.23 ± 0.15 351.00 ± 13.57 73.55 ± 4.47 377.76 ± 21.09 88.52 ± 4.56 391.95 ± 6.06 137.23 ± 8.18 406.45 ± 3.69 158.92 ± 5.43
n 0.86 ± 0.01 0.83 ± 0.01 0.83 ± 0.01 0.807 ± 0.01 0.839 ± 0.01 0.819 ± 0.01 0.833 ± 0.01 0.815 ± 0.02 0.847 ± 0.01 0.83 ± 0.02
Cdl(μF cm–2) 136.40 ± 1.83 23.53 ± 1.67 137.5 ± 0.72 27.63 ± 0.34 150.65 ± 7.04 34.98 ± 4.66 152.17 ± 11.19 59.00 ± 2.33 165.94 ± 10.42 68.13 ± 6.78
τdl (ms) 3.92 ± 0.60 6.15 ± 0.45 3.68 ± 0.52 6.23 ± 0.58 3.32 ± 0.30 5.95 ± 0.84 3.45 ± 0.63 10.30 ± 0.57 2.88 ± 0.47 6.91 ± 0.86
Rp(Ω·cm2) 26.60 ± 3.03 207.28 ± 4.41 24.48 ± 2.62 183.69 ± 14.02 21.57 ± 1.16 132.70 ± 7.58 23.37 ± 2.57 111.21 ± 13.85 19.75 ± 2.47 56.65 ± 4.47
IEEIS (%)   87.17 ± 1.49   86.67 ± 1.75   83.74 ± 1.10   78.99 ± 2.59   65.14 ± 5.17

In contrast, the Rct values diminished as the NRE increased as shown by the diameter in the Nyquist spectra. However, in the presence of [DDMS+I] and independent of the NRE, it is observed that the Rct values grew significantly with the IL concentration, suggesting that the corrosion process was controlled by the charge transfer resistance through the interface by the IL adsorption.41 The electrical double layer of the steel–electrolyte system behaves electrically as a capacitor, i.e., there is an electric charge and discharge process that leads to the corrosion process by electron transfer. This process is represented with the Cdl electric element. At the stationary state, the Cdl values diminished with respect to the blank, indicating the formation of an IL homogeneous film that blocked the reaction sites distributed uniformly in the interphase.46 Notwithstanding, with the NRE increase, a gradual growth of the Cdl values is observed, which suggests the modification of the metal–solution interphase due to the reduction of the double layer thickness and/or to the increase of the dielectric constant,50 i.e., to the formation of a thinner and more heterogeneous CI film than at the stationary state, which can be attributed to the electrolyte diffusive processes in the metal/IL interphase as a consequence of the surface damage caused by the partial hydromechanical removal of the film.51

The impedance parameter (τdl) was calculated employing eq 4 and it is reported in Table 4

2.1.4. 4

It is observed that the τdl values increased with the CI concentration. This parameter measures the time required by the metal–solution interphase to return to an equilibrium state. This fact indicates that a higher number of IL molecules adsorbed on the surface makes the electric charge and discharge process slower, which favors the reduction of the corrosion rate.52 In contrast, the process dynamic state provokes the diminution of τdl, i.e., the acceleration of electron transfer (Table 5).

Likewise, it is observed that the RL and L values diminished with increasing NRE. This behavior pattern is observed at low frequencies in the Nyquist plots (Figure 3) and is associated with an increase in the desorption rate of the intermediate products of redox reactions (H+, H3O+, O2–, and SO42–) and adsorbed [DDMS+I] molecules.

At the stationary state (Table 4), the increase in CI concentration raised the n values, i.e., the surface homogeneity; this fact confirmed the diminution of the corrosion surface damage in the presence of an inhibitor. In contrast, under flow conditions, n values were diminished (Table 5), which indicated the increase in the corrosion rate by the desorption of inhibitor molecules from the surface as a consequence of the generated sheer stress.

The Rp values were calculated by the sum of the resistive elements corresponding to each EEC. It can be observed that the Rp values were increased by the IL concentration, improving the IEEIS calculated with eq 3. However, at the dynamic state, the reduction of the Rp values as a function of NRE is revealed, which exerted a negative effect on the IEEIS falling to 22% with the flow regime change.

2.1.5. Inhibition Efficiency

As it can be observed in Figure 7, [DDMS+I] presented IEs within the interval ranging from 40 to 94% at the stationary state, which can be associated with the competition of chemical species to occupy available surface active sites and suggests that on the cathodic and anodic sites occurred the preferential adsorption of [DDMS+] and [I] with respect to H+, H3O+, O2–, and SO42–. In addition, it is revealed that from the optimal CI concentration of 100 ppm, the maximal IE of 93% was reached.

Figure 7.

Figure 7

IE for API 5L X52 steel in 1 M H2SO4 (a) at different [DDMS+I] concentrations under the stationary state and (b) with 150 ppm of [DDMS+I] under flow conditions.

According to the obtained results, the IL adsorption diminished slightly in laminar regime (NRE = 1000–2000), which can be attributed to the increase in the transport of CI molecules toward the metal surface that kept the IL adsorption and formation of complexes,53,54 which in turn reduced the τRDE negative effect and IE in 8% with respect to the stationary state. The IE fell to 86% from the transitory regime (NRE ≥ 3000) due to the formation of microvortexes that promoted higher desorption of IL molecules from the surface, making it difficult to form a continuous CI film.55 This negative effect exerted by the increasing electrode rotation speed was also reported by Ismail et al. and Benmahammed et al. who observed the diminution of the IE values of organic CIs from 47 and 89% up to 3 and 60%, respectively.42,56

Meanwhile, in turbulent regime, this phenomenon was intensified, provoking the reduction of the IE down to values around 65%. The hydromechanical removal and formation of a nonuniform inhibitor film made the adsorption and orientation of IL molecules difficult, which also prompted the desorption of metal/IL/medium complexes and the adsorption of aggressive ions on newly available active sites, thus increasing the surface instability.57

The results confirm the influence of iodide on the inhibition process, which played a major role in the obtained IE of [DDMS+I]. Also, the presence of π orbitals in S with high capacity to donate electrons and the hydrophobic properties of the decyl chain contributed to the adsorption process as reported in the literature.58

2.2. Adsorption Isotherms

To theoretically describe the nature of the interactions between the IL and the steel surface during the inhibition process, the experimental data were fitted employing adsorption isotherm models such as Langmuir, Freundlich, Temkin, and Frumkin,49,59,60 where θ is the surface fraction covered by the CI (θ = IE/100), which was obtained by the average of the IE values by the Rp and PDP techniques, C is the IL concentration in mM, and Kads is the adsorption equilibrium constant.

The Langmuir isotherm of the H2SO4 – [DDMS+I] system is shown in Figure 8. The additional isotherm models are presented in Figure S1. The linear fitting of the experimental data produced a correlation coefficient (R2) of 0.988, a slope of 0.99, and an intercept of 0.035. This empirical model suggests the formation of a [DDMS+I] monolayer, where the adsorption occurred only on a finite number of defined sites, energetically identical and equivalent with neither lateral interactions nor steric hindrance among the CI molecules, where just one of them can occupy an active site on the surface.61 The Kads value was equal to 2.82 × 104 mM–1 and was calculated by the reciprocal of the y-intercept (Kads = 1/b) of the linear regression in Figure 8. This parameter allows the computation of the Gibbs free energy of adsorption (ΔG°ads) from eq 5 (62,63)

2.2. 5

where R is the universal gas constant (8.314 kJ mol–1·K–1), T is the system absolute temperature (K), and 55.5 is the water concentration in the solution. The ΔG°ads of [DDMS+I] was equal to −35.36 kJ mol–1, which suggests that the IL adsorption mechanism involved the combination of physisorption (electrostatic forces) and chemisorption (coordinate bonds),37,64 which implied a higher interaction between the IL and the steel surface. The adsorption process can be mainly related to the chemical properties of [I] due to its high ionic radius and electronegativity lower than other halides, and the boosting inhibition effect by forming intermediate bridges between the metallic substrate and the [DDMS+] cation, which promoted the production of a CI monolayer on the surface with a hydrophobic region (cation alkyl chains) toward the solution core.6567 Furthermore, iron presents affinity to sulfur heteroatoms, which promoted the formation of [DDMS+]–Fe complexes that were adsorbed directly on the metallic surface.68

Figure 8.

Figure 8

Langmuir isotherm for the API 5L X52 steel electrode in 1 M H2SO4 solution with [DDMS+I] at 298 K.

2.3. Surface Analysis Techniques

2.3.1. SEM

Figure 9a displays the micrographs of the steel surface after its immersion in the acid medium without IL. Surface damage and the palpable mass loss due to steel oxidation and the presence of corrosion products can be observed. In contrast, the metallic samples protected with 150 ppm of [DDMS+I], Figure 9b, present a homogeneous morphology, which confirms the adsorption of the IL molecule, thus blocking the active sites, diminishing the surface damage and mass loss of the metallic samples caused by the medium attack.

Figure 9.

Figure 9

SEM/EDS of API 5L X52 steel after 4 h of immersion in 1 M H2SO4 solution: (a, c) absence and (b, d) presence of 150 ppm of [DDMS+I] at 298 K.

Figure 9c shows the EDS spectrum of API 5L X52 steel in the absence of IL. Signals corresponding to S and O can be observed, which are characteristic elements of the corrosion products that are common in steel–H2SO4 systems like FexOy, FeOOH, and FeSO4. The low intensity of S and O peaks is related to the sample rinsing process after its retrieving from the corrosive medium, which favored the removal of excess corrosion products.

The EDS spectrum in the presence of IL, Figure 9d, displays the reduction of the O signal, which suggests the winding down of the production of corrosion products due to the formation of low-solubility complexes that limited the interaction between corrosive ions and the metallic surface. Likewise, a predominant Fe signal can be observed, which confirms the properties of [DDMS+I] as CI of steel in acid medium.

2.3.2. DRIFTS

Figure 10 shows the DRIFTS spectra of the steel surface attacked for 4 h in 1 M H2SO4 in the presence of 150 ppm of inhibitor. The sample has a characteristic absorption band of the Fe–O bond of Fe3O4 at 568 cm–1.69,70 The signal appeared at around 1130 cm–1, which corresponds to the alkylsulfonium group. The bands at ∼1643 cm–1 correspond to iron oxide and the alkylsulfonium group too, while the band at 1736 cm–1 corresponds to the remaining H2SO4.71,72 The O–H stretching bands from physisorbed molecular water can be observed at 3386 cm–1, while the signals at 2957 and 2854 cm–1 belong to the C–H stretching signal of the inhibitor aliphatic group and Fe3O4.73,74 The C–H scissoring of the alkyl chain signals appears at 1459 cm–1.

Figure 10.

Figure 10

DRIFTS spectra of the steel surface after immersion in 1 M H2SO4 for 4 h at 298 K in the presence of 150 ppm of inhibitor.

2.3.3. XPS

XPS technique was used to analyze the oxidation state of the elements present on the API 5L X52 steel surface attacked for 4 h in 1 M H2SO4 in the presence of 150 ppm of inhibitor. The general XPS spectra indicate the presence of Fe (2p), O (1s), C (1s) S (2p), and I (3d). Fe is attributed to the metal substrate. Oxygen and sulfur signals are related mainly to corrosion products and salts deposited on the metal surface due to the corrosive medium type. The high carbon signal (60.4%) is ascribed not only to adventitious carbon but also to the presence of CI on the metal surface. There is also an iodine signal, which can be associated with the corrosion inhibitor structure.

Figure 11a shows the high-resolution C 1s peak, which deconvoluted to three peaks. The main peak at 284.8 eV indicates the presence of C–C and C–H bonds attributed to the hydrocarbon alkyl chain bond,75,76 whereas the signal at 286.475,77 is related to C–S bonds present in the molecule; the signal at 288.6 eV is attributed to the presence of COO stemming from carbon impurities and substrate carbon.78,79

Figure 11.

Figure 11

XPS deconvoluted profiles of (a) C 1s, (b) S 2p, (c) O 1s, (d) Fe 2p, and (e) I 3d on the iron sample surface exposed to 1 M H2SO4 with 150 ppm of [DDMS+I] for 4 h at 298 K.

Figure 11b shows the high-resolution S 2p peak deconvoluted to four peaks. The S 2p3/2 binding energies at 162.0 and S 2p1 at 163.1 eV were associated with the S–C bond, while the peaks at 168.4 and 169.5 eV for S 2p3/2 and S 2p1, respectively, are attributed to sulfate species (SO42–), showing a clear correlation with the sulfur oxidation degree.75,8083

In the O 1s region, Figure 11c, the main peak at 531.9 eV corresponds to oxygen in the hydroxide form, which is present in oxyhydroxides (FeOOH),78,84 whereas the signal at 530.1 eV is associated with oxygen in oxidized form as ferrosoferric oxide (Fe3O4),85,86 while the peak at 532.9 eV is ascribed to oxygen in the sulfate group (SO42–).87

The Fe 2p spectrum exhibits mainly three peaks (Figure 11d): (1) at 706.8 eV (Fe 2p3/2) and 719.6 eV (Fe 2p1/2), corresponding to metal Fe0; (2) at 710.4 eV (Fe 2p3/2) and 723.7 eV (Fe 2p1/2), related to ferrosoferric oxide (Fe3O4);88 and (3) at 713.4 eV (Fe 2p3/2) and 727.2 eV (Fe 2p1/2), corresponding to oxyhydroxides (FeOOH).89

The two main signals located at 619.3 and 630.8 eV (Figure 11e) can be ascribed to the photoemission peaks from I 3d5 and I 3d3, respectively, and correspond to the iodide ion.90,91

2.4. Computer Simulation Analysis

2.4.1. Geometry

Figure 12 shows the optimal [DDMS+I] structure. The absence of imaginary frequencies confirms the system structural stability. To take the molecule to the ground state, neuter charge multiplicity 1 (singlet) was employed. In the optimal structure, the methyl and decyl groups were distributed outward to the plane with respect to sulfur (trigonal pyramidal shape). As for [I], it presented a preference to be located on the alkyl groups linked to sulfur, where the interaction with sulfur was minimized by the carbon and hydrogen atoms,92 being the most stable IL form. Regarding the decyl group, it displayed the characteristic linear conformation of a C ≥ 4 alkyl chain that has been reported in different ILs based on ammonium and imidazolium.9395

Figure 12.

Figure 12

Optimized structure of [DDMS+I] in vacuum obtained at B3LYP/MIDIx level. All atoms are at their relaxed positions. Gray: carbon, white: hydrogen, yellow: sulfur, and purple: iodine.

Table 6 reports the lengths and bond angles for each group of atoms. It is observed that the C–C, S–C, and C–H bonds did not present changes in vacuum and aqueous medium [employing COSMO (COnductor-like Screening MOdel) as the solvation model, which uses a dielectric constant of 78.4 to simulate the medium].96 In contrast, the space between sulfur and iodide shows slight distancing (∼0.33 Å), which increased the interaction capacity with other molecules. As for the bond angles of the different atom groups, they presented negligible variations when the medium was changed, suggesting that the distribution of the [DDMS+I] atoms was not affected.

Table 6. Bond Lengths and Angles Calculated for [DDMS+I].
  medium (Å)
bond vacuum water
C–C 1.54 1.54
S–C 1.84 1.84
C–H 1.10 1.10
S···I 4.20 4.53
C–C–C 59.25 59.25
C–S–C 39.41 39.41
H–C–H 112.15 112.15
H–H–H 60.4 60.4
S–C–C 110.28 110.28
C–C–C 112.26 112.26
H–C–H 107.85 107.85
H–H–H 59.99 59.99

2.4.2. Energetic Analysis

As observed in Figure 13a, the highest occupied molecular orbital (HOMO) works as an electron donor and is located in the [I] ion, whereas (b) the lowest unoccupied molecular orbital (LUMO) functions as an electron acceptor, since it is the cation empty internal orbital and, in this case, it extended itself from [S+] to the methyl groups and decyl (C5).

Figure 13.

Figure 13

(a) Highest occupied molecular orbital (HOMO) and (b) lowest unoccupied molecular orbital (LUMO) of [DDMS+I]. All are obtained at the optimized geometry in vacuum.

Regarding the molecular electrostatic potential (MEP) mapping, it is employed to determine reactive sites for electrophilic and nucleophilic attacks by visualizing the charge distribution, which is related to the molecule electron density. Figure 14 shows the MEP of [DDMS+I], where three different colors can be seen: the red color is associated with the IL most electronegative atom, the iodide anion, and electrophilic reactivity, i.e., with the most susceptible site to electrophilic attacks that cedes electrons to the steel surface to form coordinate bonds with the empty Fe orbital. The blue color located in the cation [S+] is related to the nucleophilic reactivity, which is the most susceptible site to nucleophilic attacks, i.e., the site that accepts electrons from other species. Finally, the green color indicates that both the methyl and decyl groups present neuter charge, i.e., their electron configurations are full.97 This behavior pattern has been observed in works on similar compounds like the one by Haque et al. on N-methyl-N,N,N-trioctylammonium chloride, where the blue region is located on the nitrogen atom of the cationic part, whereas the red region on the chloride ion.93

Figure 14.

Figure 14

MEP isosurface of [DDMS+I] obtained at the B3LYP/MIDIx level.

The MEP and MO analyses confirmed that the [I] anion worked as an electron donor and the [S+] cation as an electron acceptor, which represent sites associated with molecule adsorption processes.

2.4.3. Quantum Parameters

For the reactivity analysis of [DDMS+I], the HOMO (EHOMO) and LUMO (ELUMO) energy values were calculated as shown in Table 7. According to the theory, high EHOMO values are associated with a higher tendency to donate electrons to appropriate receiving molecules with low-energy empty molecular orbitals. In contrast, low ELUMO values suggest higher probability of accepting electrons on the metallic surface. The latter indicates that if a CI exhibits high EHOMO and low ELUMO, it will have better interaction with the metallic surface.42,98 As observed in Table 7, EHOMO is relatively higher and ELUMO is relatively lower in both media, which suggests that [DDMS+I] presents a strong tendency to donate and accept electrons.

Table 7. Dipole Moment, HOMO and LUMO Energies, and Energy Gap (ΔEL–H) of [DDMS+I].
system EHOMO (eV) ELUMO (eV) ΔEL–H (eV) μ (Debye)
vacuum 5.98 1.77 4.21 10.77
water 5.91 1.63 4.29 15.92

As for the formation of a transition state, it is due to the interaction between the MOs, i.e., to the energy gap between EHOMO and ELUMOEL–H = ELUMOEHOMO). This parameter is considered as a descriptor of the molecular activity, where low ΔEL–H values indicate higher polarization and an increase in the metal surface reactivity, which is associated with better CI behavior. Table 7 shows that [DDMS+I] presented a significant difference regarding ΔEL–H in both media, which facilitated the adsorption.93,99 The μ values for [DDMS+I] were equal to 10.77 and 15.92 Debye in vacuum and water, respectively, which resulted to be higher than those reported for water (1.85 Debye), suggesting the preference to replace the adsorbed water molecules on the metallic surface by the IL.93,100

2.5. Corrosion Inhibition Mechanism

As it is known, iodides exert an inhibitory effect on corrosion, which occurs mainly in the anodic sites of the metal surface, slowing down the dissolution of iron in the form of iron cations (Fe2+).101,102 Based on theoretical calculations, the iodide anion hosts the most part of HOMO, which attracts it to the anode sites.

Fe+2(OH)ads or Fe+3(OH)ads are formed at the anodic sites after adsorption of water and hydroxide anions (OH), which in turn form corrosion products like magnetite (Fe3O4) or through the path of “green rust” to form iron oxyhydroxides (FeOOH), usually lepidocrocite; the presence of these corrosion products was confirmed by XPS analysis. In the presence of corrosion inhibitor, very few corrosion products typical for the sulfuric acid medium such as rozenite and melanterite were detected,103 which means that I ions prevented the deposition of sulfate ions on the electrode surface, and hindered sulfate passivation, slowing down the formation on a larger scale of corrosion products that contain the sulfate anion in their structure.

However, since the CI is a mixed-type inhibitor, it acted not only on anodic sites, but also on cathodic sites (Figure 15). As the CI cation is a trialkyl sulfonium derivative, where LUMO is located, it accepted electrons. The sulfonium ion approached the cathode sites on the metal surface by competing with H3O+ and displacing water. Since the CI cation is much larger than H3O+, it removed the protons from the metal surface in the cathode zone and thus slowed down the hydrogen formation reaction. Based on the electrochemical analysis, the CI concentration of 50 ppm caused notable inhibition of the corrosive process (greater than 80%), while from the concentration of 75 ppm, there was no big change in the anticorrosive effect (>90%) in laminar and (>80%) transitory flows.

Figure 15.

Figure 15

Corrosion inhibition mechanism of API 5L X52 steel in 1 M H2SO4 containing [DDMS+I].

3. Conclusions

From the present work, the following conclusions were drawn.

  • Electrochemical measurements showed that the charge transfer rate in the presence of CI was lower than the blank, which evidenced the formation of a protecting film.

  • It was observed that the IE values increased with the IL concentration. Furthermore, under hydrodynamic conditions, the IE values diminished by the desorption of IL molecules due to sheer stress.

  • ΔG°ads values suggested that the IL adsorption process was a combination of physisorption and chemisorption.

  • The surface analyses supported the [DDMS+I] inhibition process, exhibiting less damage of the metallic surface by IL adsorption.

  • The HOMO of [DDMS+I] suggested that the iodide anion had the capacity to donate electrons, which eased its adsorption on the anodic sites, thus preventing the depositing of sulfate ions on the electrode surface.

4. Methods

4.1. Synthesis and Characterization of Decyl(dimethyl)sulfonium Iodide

The molecular structure of the synthesized IL is depicted in Table 8. Iodomethane (≥99%, Sigma-Aldrich), n-decyl methyl sulfide (97%, Alfa Aesar), and trifluoroacetic acid (99%, Sigma-Aldrich) were employed in the synthesis. The compound was characterized by NMR. The 1H (300 MHz) spectrum was recorded on a JEOL Eclipse-300 equipment in DMSO-d6, and chemical shifts were expressed in ppm relative to tetramethylsilane as the internal standard. Decyl(dimethyl)sulfonium iodide was synthesized according to a similar previously described procedure.104

Table 8. Chemical Structure of the IL Evaluated as CI.

4.1.

4.1.1. Decyl(dimethyl)sulfonium Iodide [DDMS+I]

In total, 3.17 g (22.3 mmol) of methyl iodide, 1.60 g (8.47 mmol) of n-decyl methyl sulfide, and 3.12 g (27 mmol) of trifluoroacetic acid were mixed in a 25 mL round-bottom flask, which was closed with a stopper. The reaction was performed at room temperature without stirring. After 30 min, diethyl ether (100 mL) was added to precipitate the product as a white solid. Yield: 2.50 g (89%). 1H NMR (300 MHz, DMSO-d6): 3.34 (s, 6H), 3.88 (m, 2H), 1.91 (s, 2H), 1.60 (m, 2H), 1.33–1.20 (m, 14H), 0.85 (t, J = 6 Hz, 3H).

4.2. Materials and Test Solutions

API 5L X52 steel coupons with the exposed surface area of 0.289 cm2 were employed with the following chemical composition (wt %): C ≤ 0.28%, Mn ≤ 1.4%, P ≤ 0.030%, S 0.030%, V ≤ 0.15%, Nb 0.15%, Ti ≤ 0.15%, Cu 0.25%, Ni 0.25%, Cr 0.25%, Mo 0.15%, and Fe as the main element. Before each test, the metal samples were polished with SiC emery paper (from 400 to 1200). Afterward, they were degreased with acetone–ethanol and exposed to an ultrasonic bath to eliminate particles adhered to the surface. Finally, the polished samples were dried under nitrogen flow.105 1 M H2SO4 was the corrosive solution and it was prepared with analytic-grade acid and deionized water. [DDMS+I] was evaluated at concentrations ranging from 25 to 150 ppm.

4.3. Electrochemical Tests

Electrochemical tests were performed employing a Potentiostat/Galvanostat AutoLab apparatus model PGSTAT302N. The software NOVA 2.1.4 was used to obtain and analyze the experimental data. A glass electrochemical cell equipped with three electrodes was used: counter electrode (platinum 99%), working electrode (API 5L X52 steel), and reference electrode (Ag/AgCl). All of the tests were carried out at 25 ± 1 °C in aerated medium. In addition, the tests were run in triplicate and the reported results represent the average.

The working electrode was immersed in the electrolytic solution for 20 min until reaching the EOCP. The Rp measurements were carried out within a potential interval of ± 25 mV vs EOCP(106) whereas that of PDP was of ± 250 mV vs EOCP;107 both tests took place at a scanning rate of 0.1666 mV s–1. EIS tests were developed within a frequency interval ranging from 100 kHz to 10 mHz using a sinusoidal wave with 5 mV of amplitude after stabilizing the EOCP.108

The flow tests were performed with a rotating disc Metrohom RDE II 309.109 The NRE was calculated from eq 6110,111

4.3. 6

where U is the cylinder peripheral speed (m s–1), d is the electrode diameter (m), and v is the electrolyte kinematic viscosity (m2 s–1). The τRDE generated on the surface was calculated with eq 7112,113

4.3. 7

where ρ is the density (kg m–3), r is the cylinder radius (m), and ω is the electrode angular velocity (rad s–1). In Table 9, ω, NRE, and τRDE values under laminar, transitory, and turbulent hydrodynamic regime conditions, respectively, can be observed.

Table 9. Angular Velocity, NRE, and Shear Stress as Functions of the RDE Rotation Rate.

regimen rotation rate (rpm) ω (rad s–1) NRE τRDE (Pa)
laminar 343 35.9 1000 0.320
686 71.84 2000 1.040
transitory 1029 107.76 3000 2.073
turbulent 1372 143.68 4000 3.381
1715 179.6 5000 4.940

4.4. Surface Analysis

The samples employed for the surface analyses were prepared by following the methodology for electrochemical tests and polished with 1 μm alumina. The metal coupons were immersed in the corrosive medium in the absence and presence of 150 ppm of CI for 4 h at 25 °C. Afterward, the metal samples were retrieved from the medium and rinsed with deionized water and dried with nitrogen.105 The surface of API 5L X52 steel was analyzed by SEM/EDS on a JEOL-JSM-6300 microscope. The study of the treated metal surfaces was carried out using DRIFTS; these measurements were performed in situ using a Thermo Scientific Nicolet 560 Spectrometer in a series of spectra recorded with identical resolution (4 cm–1). The XPS analysis was performed with a K-Alpha Thermo Fisher Scientific spectrometer with monochromatic Al Kα (1486.6 eV) and vacuum pressure of 1 × 10–9 Torr. The pass energy values for the study and high-resolution spectra were set at 160 and 20 eV, respectively. The obtained spectra were referred to adventitious carbon (284.8 eV) and the peak fitting was performed using the software Thermo Avantage v.5.9915.

4.5. Computational Details

The IL corrosion inhibition performance was supported with first-principles energy calculations. The [DDMS+] cation was optimized structurally considering different positions of the [I] anion without symmetry restriction and singlet state (Multiplicity 1). The computations were developed with the density functional theory (DFT) by the Gaussian 09W software114 with B3LYP/MIDIx theory level.115,116 The entries were generated with the software Gauss View v6.0. The properties were obtained employing the lowest total energy configuration, whereas the CI theoretical performance was established through the analysis of the molecular orbitals and dipolar moment (μ),42,98 which were calculated from the optimized structure under standard temperature and pressure conditions.117

Acknowledgments

V.D.-J. and G.G.-S. gratefully acknowledge CONACYT for the scholarship granted to pursue postgraduate studies. Paulina Arellanes-Lozada and Octavio Olivares-Xometl thank CONACYT-Mexico and BUAP-VIEP (CA 256). This research work was supported by the high-performance computing system of PIDi-UTEM (SCC-PIDi-UTEM CONICYT-FONDEQUIP-EQM180180).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c05192.

  • Other analyzed models of adsorption isotherms for API 5L X52 steel in 1 M H2SO4 solution with [DDMS+I] (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao2c05192_si_001.pdf (94KB, pdf)

References

  1. Bhardwaj N.; Sharma P.; Singh K.; Rana D.; Kumar V. Phyllanthus emblica seed extract as corrosion inhibitor for stainless steel used in petroleum industry (SS-410) in acidic medium. Chem. Phys. Impact 2021, 3, 100038 10.1016/j.chphi.2021.100038. [DOI] [Google Scholar]
  2. Bhardwaj N.; Sharma P.; Kumar V. Corrosion Inhibition Property and Adsorption Behavior of P. tremula Leaf Extract in Acidic Media for Steel used in Petroleum Industry (SS-410). Prot. Met. Phys. Chem. Surf. 2021, 57, 1076–1084. 10.1134/S2070205121050051. [DOI] [Google Scholar]
  3. Wan S.; Wei H.; Quan R.; Luo Z.; Wang H.; Liao B.; Guo X. Soybean extract firstly used as a green corrosion inhibitor with high efficacy and yield for carbon steel in acidic medium. Ind. Crops Prod. 2022, 187, 115354 10.1016/j.indcrop.2022.115354. [DOI] [Google Scholar]
  4. Zhang W.; Zhang D.; Li X.; Li C.; Gao L. Excellent performance of dodecyl dimethyl betaine and calcium gluconate as hybrid corrosion inhibitors for Al alloy in alkaline solution. Corros. Sci. 2022, 207, 110556 10.1016/j.corsci.2022.110556. [DOI] [Google Scholar]
  5. Yang L.; Fan H.; Yan R.; Zhang J.; Liu S.; Huang X.; Zhang D. N-substituted methyl ethylenediamine derivatives as corrosion inhibitors for carbon steel in 1 M hydrochloride acid. J. Mol. Struct. 2022, 1270, 133975 10.1016/j.molstruc.2022.133975. [DOI] [Google Scholar]
  6. Gholivand K.; Sarmadi-babaee L.; Faraghi M.; Badalkhani-khamseh F.; Fallah N. Heteroatom-containing phosphoramides as carbon steel corrosion inhibitors: Density functional theory and molecular dynamics simulations. Chem. Phys. Impact 2022, 5, 100099 10.1016/j.chphi.2022.100099. [DOI] [Google Scholar]
  7. Verma D. K.; Aslam R.; Aslam J.; Quraishi M. A.; Ebenso E. E.; Verma C. Computational Modeling: Theoretical Predictive Tools for Designing of Potential Organic Corrosion Inhibitors. J. Mol. Struct. 2021, 1236, 130294 10.1016/j.molstruc.2021.130294. [DOI] [Google Scholar]
  8. Zhang Q.; Zhang R.; Wu R.; Luo Y.; Guo L.; He Z. Green and high-efficiency corrosion inhibitors for metals: a review. J. Adhes. Sci. Technol. 2022, 1–24. 10.1080/01694243.2022.2082746. [DOI] [Google Scholar]
  9. Al-Rashed O.; Nazeer A. A. Effectiveness of Some Novel Ionic Liquids on Mild Steel Corrosion Protection in Acidic Environment: Experimental and Theoretical Inspections. Materials 2022, 15, 2326 10.3390/ma15062326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Verma C.; Alrefaee S. H.; Quraishi M. A.; Ebenso E. E.; Hussain C. M. Recent developments in sustainable corrosion inhibition using ionic liquids: A review. J. Mol. Liq. 2021, 321, 114484 10.1016/j.molliq.2020.114484. [DOI] [Google Scholar]
  11. Salim R.; Hajjaji F. E. L.; Ech-chihbi E.; Titi A.; Messali M.; Nahle A.; Kaya S.; El Ibrahimi B.; Taleb M. Experimental and theoretical approach for novel imidazolium ionic liquids as Smart Corrosion inhibitors for mild steel in 1.0 M hydrochloric acid. Arab. J. Chem. 2022, 15, 103967 10.1016/j.arabjc.2022.103967. [DOI] [Google Scholar]
  12. Tian G.; Yuan K. Adsorption and inhibition behavior of imidazolium tetrafluoroborate derivatives as green corrosion inhibitors for carbon steel. J. Mol. Model. 2021, 27, 195. 10.1007/s00894-021-04794-1. [DOI] [PubMed] [Google Scholar]
  13. Zunita M.; Kevin Y. J. Results in Engineering Ionic liquids as corrosion inhibitor: From research and development to commercialization. Results Eng. 2022, 15, 100562 10.1016/j.rineng.2022.100562. [DOI] [Google Scholar]
  14. Bhaskaran; Pancharatna P. D.; Lata S.; Singh G. Imidazolium based ionic liquid as an efficient and green corrosion constraint for mild steel at acidic pH levels. J. Mol. Liq. 2019, 278, 467–476. 10.1016/j.molliq.2019.01.068. [DOI] [Google Scholar]
  15. Gurjar S.; Sharma S. K.; Sharma A.; Ratnani S. Performance of imidazolium based ionic liquids as corrosion inhibitors in acidic medium: A review. Appl. Surf. Sci. Adv. 2021, 6, 100170 10.1016/j.apsadv.2021.100170. [DOI] [Google Scholar]
  16. El-Hajjaji F.; Salim R.; Taleb M.; Benhiba F.; Rezk N.; Singh-Chauhan D.; Quraishi M. A. Pyridinium-based ionic liquids as novel eco-friendly corrosion inhibitors for mild steel in molar hydrochloric acid: Experimental & computational approach. Surf. Interfaces 2021, 22, 100881 10.1016/j.surfin.2020.100881. [DOI] [Google Scholar]
  17. Hamidi Z.; Mosavian S. Y.; Sabbaghi N.; Karimi-Zarchi M. A.; Noroozifar M. Cross-linked poly(N-alkyl-4-vinylpyridinium) iodides as new eco-friendly inhibitors for corrosion study of St-37 steel in 1 M H2SO4. Iran. Polym. J. 2020, 29, 225–239. 10.1007/s13726-020-00787-8. [DOI] [Google Scholar]
  18. Wang S.; Sun J.; Shan B.; Fan W.; Ding R.; Yang J.; Zhao X. Performance of dodecyl dimethyl benzyl ammonium chloride as bactericide and corrosion inhibitor for 7B04 aluminum alloy in an aircraft fuel system. Arab. J. Chem. 2022, 15, 103926 10.1016/j.arabjc.2022.103926. [DOI] [Google Scholar]
  19. Zheng T.; Liu J.; Wang M.; Liu Q.; Wang J.; Chong Y.; Jia G. Synergistic corrosion inhibition effects of quaternary ammonium salt cationic surfactants and thiourea on Q235 steel in sulfuric acid: Experimental and theoretical research. Corros. Sci. 2022, 199, 110199 10.1016/j.corsci.2022.110199. [DOI] [Google Scholar]
  20. Kannan P.; Varghese A.; Palanisamy K.; Abousalem A. S. Evaluating prolonged corrosion inhibition performance of benzyltributylammonium tetrachloroaluminate ionic liquid using electrochemical analysis and Monte Carlo simulation. J. Mol. Liq. 2020, 297, 111855 10.1016/j.molliq.2019.111855. [DOI] [Google Scholar]
  21. El-Hajjaji F.; Salim R.; Messali M.; Hammouti B.; Chauhan D. S.; Almutairi S. M.; Quraishi M. A. Electrochemical Studies on New Pyridazinium Derivatives as Corrosion Inhibitors of Carbon Steel in Acidic Medium. J. Bio- Tribo-Corros. 2019, 5, 1–13. 10.1007/s40735-018-0195-3. [DOI] [Google Scholar]
  22. Gurjar S.; Ratnani S.; Kandwal P.; Tiwari K. K.; Sharma A.; Sharma S. K. Experimental and theoretical studies of 1-Benzyl pyridazinium bromide as green inhibitor for mild steel corrosion. Adv. Electr. Eng. Electron. Energy 2022, 2, 100054 10.1016/j.prime.2022.100054. [DOI] [Google Scholar]
  23. Deyab M. A. Sulfonium-based ionic liquid as an anticorrosive agent for thermal desalination units. J. Mol. Liq. 2019, 296, 111742 10.1016/j.molliq.2019.111742. [DOI] [Google Scholar]
  24. Arab S. T.; Al-Turkustani A. M. Corrosion Inhibition of Steel in Phosphoric acid by Phenacyldimethyl Sulfonium Bromide and some of its p-Substituted Derivatives. Port. Electrochim. Acta 2006, 24, 53–69. 10.4152/pea.200601053. [DOI] [Google Scholar]
  25. Shein A. B.; Nedugov A. N. Examination of Trialkyl-Substituted Sulfonium, Selenonium, and Telluronium Salts as Inhibitors of Acid Corrosion of Iron and Steel. Prot. Met. 2000, 36, 240–243. 10.1007/BF02758398. [DOI] [Google Scholar]
  26. Verma C.; Obot I. B.; Bahadur I.; Sherif E.-S. M.; Ebenso E. E. Choline based ionic liquids as sustainable corrosion inhibitors on mild steel surface in acidic medium: Gravimetric, electrochemical, surface morphology, DFT and Monte Carlo simulation studies. Appl. Surf. Sci. 2018, 457, 134–149. 10.1016/j.apsusc.2018.06.035. [DOI] [Google Scholar]
  27. Mashuga M.; Olasunkanmi L.; Adekunle A.; Yesudass S.; Kabanda M.; Ebenso E. Adsorption, Thermodynamic and Quantum Chemical Studies of 1-hexyl-3-methylimidazolium Based Ionic Liquids as Corrosion Inhibitors for Mild Steel in HCl. Materials 2015, 8, 3607–3632. 10.3390/ma8063607. [DOI] [Google Scholar]
  28. Azeez F. A.; Al-Rashed O. A.; Nazeer A. A. Controlling of mild-steel corrosion in acidic solution using environmentally friendly ionic liquid inhibitors: Effect of alkyl chain. J. Mol. Liq. 2018, 265, 654–663. 10.1016/j.molliq.2018.05.093. [DOI] [Google Scholar]
  29. Cao S.; Liu D.; Ding H.; Wang J.; Lu H.; Gui J. Corrosion inhibition effects of a novel ionic liquid with and without potassium iodide for carbon steel in 0.5 M HCl solution: An experimental study and theoretical calculation. J. Mol. Liq. 2019, 275, 729–740. 10.1016/j.molliq.2018.11.115. [DOI] [Google Scholar]
  30. Guo L. Synergistic Effect of Potassium Iodide with L-Tryptophane on the Corrosion Inhibition of Mild Steel: A Combined Electrochemical and Theoretical Study. Int. J. Electrochem. Sci. 2017, 166–177. 10.20964/2017.01.04. [DOI] [Google Scholar]
  31. Farag A. A.; Hegazy M. A. Synergistic inhibition effect of potassium iodide and novel Schiff bases on X65 steel corrosion in 0.5M H2SO4. Corros. Sci. 2013, 74, 168–177. 10.1016/j.corsci.2013.04.039. [DOI] [Google Scholar]
  32. Feng L.; Zhang S.; Qiang Y.; Xu S.; Tan B.; Chen S. The synergistic corrosion inhibition study of different chain lengths ionic liquids as green inhibitors for X70 steel in acidic medium. Mater. Chem. Phys. 2018, 215, 229–241. 10.1016/j.matchemphys.2018.04.054. [DOI] [Google Scholar]
  33. Qiu Y.; Tu X.; Lu X.; Yang J. A novel insight into synergistic corrosion inhibition of fluoride and DL-malate as a green hybrid inhibitor for magnesium alloy. Corros. Sci. 2022, 199, 110177 10.1016/j.corsci.2022.110177. [DOI] [Google Scholar]
  34. Sastri V. S.; Ghali E.; Elboujdaini M.. Corrosion Prevention and Protection Practical Solutions; John Wiley & Sons Ltd, 2007. [Google Scholar]
  35. Papavinasam S.; Doiron A.; Shen G.; Revie R. W.. Prediction of Inhibitor Behaviour in the Field From Data in the Laboratory; NACE Corrosion 2004: New Orleans, Louisiana, 2004. [Google Scholar]
  36. Morales-Gil P.; Walczak M. S.; Camargo C. R.; Cottis R. A.; Romero J. M.; Lindsay R. Corrosion inhibition of carbon-steel with 2-mercaptobenzimidazole in hydrochloric acid. Corros. Sci. 2015, 101, 47–55. 10.1016/j.corsci.2015.08.032. [DOI] [Google Scholar]
  37. Scendo M.; Uznanska J. The Effect of Ionic Liquids on the Corrosion Inhibition of Copper in Acidic Chloride Solutions. Int. J. Corros. 2011, 2011, 718626 10.1155/2011/718626. [DOI] [Google Scholar]
  38. Zhang K.; Yang W.; Xu B.; Liu Y.; Yin X.; Chen Y. Corrosion inhibition of mild steel by bromide-substituted imidazoline in hydrochloric acid. J. Taiwan Inst. Chem. Eng. 2015, 57, 167–174. 10.1016/j.jtice.2015.05.021. [DOI] [Google Scholar]
  39. Kamali-Ardakani E.; Kowsari E.; Ehsani A. Imidazolium-derived polymeric ionic liquid as a green inhibitor for corrosion inhibition of mild steel in 1.0 M HCl: Experimental and computational study. Colloids Surf., A 2020, 586, 124195 10.1016/j.colsurfa.2019.124195. [DOI] [Google Scholar]
  40. Wang X.; Yang H.; Wang F. An investigation of benzimidazole derivative as corrosion inhibitor for mild steel in different concentration HCl solutions. Corros. Sci. 2011, 53, 113–121. 10.1016/j.corsci.2010.09.029. [DOI] [Google Scholar]
  41. Aldana-González J.; Espinoza-Vázquez A.; Romero-Romo M.; Uruchurtu-Chavarin J.; Palomar-Pardavé M. Electrochemical evaluation of cephalothin as corrosion inhibitor for API 5L X52 steel immersed in an acid medium. Arab. J. Chem. 2015, 12, 3244–3253. 10.1016/j.arabjc.2015.08.033. [DOI] [Google Scholar]
  42. Ismail A.; Irshad H. M.; Zeino A.; Toor I. H. Electrochemical Corrosion Performance of Aromatic Functionalized Imidazole Inhibitor Under Hydrodynamic Conditions on API X65 Carbon Steel in 1 M HCl Solution. Arab. J. Sci. Eng. 2019, 44, 5877–5888. 10.1007/s13369-019-03745-6. [DOI] [Google Scholar]
  43. Burduhos-Nergis D.-P.; Vizureanu P.; Sandu A. V.; Bejinariu C. Evaluation of the Corrosion Resistance of Phosphate Coatings Deposited on the Surface of the Carbon Steel Used for Carabiners Manufacturing. Appl. Sci. 2020, 10, 2753. 10.3390/app10082753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Xu L.-N.; Zhu J.-y.; Lu M.-x.; Zhang L.; Chang W. Electrochemical impedance spectroscopy study on the corrosion of the weld zone of 3Cr steel welded joints in CO2 environments. Int. J. Miner., Metall. Mater. 2015, 22, 500–509. 10.1007/s12613-015-1099-6. [DOI] [Google Scholar]
  45. Tan Z.; Yang L.; Zhang D.; Wang Z.; Cheng F.; Zhang M.; Jin Y. Development mechanism of internal local corrosion of X80 pipeline steel. J. Mater. Sci. Technol. 2020, 49, 186–201. 10.1016/j.jmst.2019.10.023. [DOI] [Google Scholar]
  46. Mo M.; Zhao W.; Chen Z.; Liua E.; Xuea Q. Corrosion inhibition of functional graphene reinforced polyurethane nanocomposite coatings with regular textures. RSC Adv. 2016, 6, 7780–7790. 10.1039/C5RA24823J. [DOI] [Google Scholar]
  47. Bojang A. A.; Wu H. S. Characterization of Electrode Performance in Enzymatic Biofuel Cells Using Cyclic Voltammetry and Electrochemical Impedance Spectroscopy. Catalysts 2020, 10, 782 10.3390/catal10070782. [DOI] [Google Scholar]
  48. Delgado M. C.; García-Galvan F. R.; Barranco V.; Feliu-Batlle S.. A Measuring Approach to Assess the Corrosion Rate of Magnesium Alloys Using Electrochemical Impedance Spectroscopy. In Magnesium Alloys, Aliofkhazraei M., Ed.; IntechOpen: London, 2017; pp 129–160. [Google Scholar]
  49. Hsissou R.; Benhiba F.; Dagdag O.; El Bouchti M.; Nouneh K.; Assouag M.; Briche S.; Zarrouk A.; Elharfi A. Development and potential performance of prepolymer in corrosion inhibition for carbon steel in 1.0 M HCl: Outlooks from experimental and computational investigations. J. Colloid Interface Sci. 2020, 574, 43–60. 10.1016/j.jcis.2020.04.022. [DOI] [PubMed] [Google Scholar]
  50. Chafiq M.; Chaouiki A.; Al-Hadeethi M. R.; Ali I. H.; Mohamed S. K.; Toumiat K.; Salghi R. Naproxen-Based Hydrazones as Effective Corrosion Inhibitors for Mild Steel in 1.0 M HCl. Coatings 2020, 10, 1–17. 10.3390/coatings10070700. [DOI] [Google Scholar]
  51. Barmatov E.; Hughes T. L. Effect of corrosion products and turbulent flow on inhibition efficiency of propargyl alcohol on AISI 1018 mild carbon steel in 4 M hydrochloric acid. Corros. Sci. 2017, 123, 170–181. 10.1016/j.corsci.2017.04.020. [DOI] [Google Scholar]
  52. Singh A.; Ansari K. R.; Chauhan D. S.; Quraishi M. A.; Lgaz H.; Chung I.-M. Comprehensive investigation of steel corrosion inhibition at macro/micro level by ecofriendly green corrosion inhibitor in 15% HCl medium. J. Colloid Interface Sci. 2020, 560, 225–236. 10.1016/j.jcis.2019.10.040. [DOI] [PubMed] [Google Scholar]
  53. Rodriguez-Clemente E.; Gonzalez-Rodriguez J. G.; Valladares-Cisneros M. G.; Chacon-Nava J. G. Evaluation of Allium sativum as corrosion inhibitor for carbon steel in sulphuric acid under hydrodynamic conditions. Green Chem. Lett. Rev. 2015, 8, 49–58. 10.1080/17518253.2015.1071435. [DOI] [Google Scholar]
  54. Vakili-Azghandi M.; Davoodi A.; Farzi G. A.; Kosari A. Water-base acrylic terpolymer as a corrosion inhibitor for SAE1018 in simulated sour petroleum solution in stagnant and hydrodynamic conditions. Corros. Sci. 2012, 64, 44–54. 10.1016/j.corsci.2012.07.003. [DOI] [Google Scholar]
  55. Zhao J.; Liu Y.; Yang X.; He X.; Wang L.; Xiong D.; Gu Y. Corrosion Behavior of Pipeline Steel in Oilfield Produced Water under Dynamic Corrosion System. J. Wuhan Univ. Technol. 2022, 37, 677–691. 10.1007/s11595-022-2582-3. [DOI] [Google Scholar]
  56. Benmahammed I.; Douadi T.; Issaadi S.; Al-Noaimi M.; Chafaa S. Heterocyclic Schiff bases as corrosion inhibitors for carbon steel in 1 M HCl solution: hydrodynamic and synergetic effect. J. Dispersion Sci. Technol. 2020, 41, 1002–1021. 10.1080/01932691.2019.1614038. [DOI] [Google Scholar]
  57. Obot I.-B.; Onyeachu I.-B.; Umoren S. A. Alternative corrosion inhibitor formulation for carbon steel in CO2-saturated brine solution under high turbulent flow condition for use in oil and gas transportation pipelines. Corros. Sci. 2019, 159, 108140 10.1016/j.corsci.2019.108140. [DOI] [Google Scholar]
  58. Guo L.; Obot I. B.; Zheng X.; Shen X.; Qiang Y.; Kaya S.; Kaya C. Theoretical insight into an empirical rule about organic corrosion inhibitors containing nitrogen, oxygen, and sulfur atoms. Appl. Surf. Sci. 2017, 406, 301–306. 10.1016/j.apsusc.2017.02.134. [DOI] [Google Scholar]
  59. Hsissou R.; Abbout S.; Safi Z.; Benhiba F.; Wazzan N.; Guo L.; Nouneh K.; Briche S.; Erramli H.; Ebn Touhami M.; et al. Synthesis and anticorrosive properties of epoxy polymer for CS in [1 M] HCl solution: Electrochemical, AFM, DFT and MD simulations. Constr. Build. Mater. 2021, 270, 121454 10.1016/j.conbuildmat.2020.121454. [DOI] [Google Scholar]
  60. Hsissou R.; Abbout S.; Seghiri R.; Rehioui M.; Berisha A.; Erramli H.; Assouag M.; Elharfi A. Evaluation of corrosion inhibition performance of phosphorus polymer for carbon steel in [1 M] HCl: Computational studies (DFT, MC and MD simulations). J. Mater. Res. Technol. 2020, 9, 2691–2703. 10.1016/j.jmrt.2020.01.002. [DOI] [Google Scholar]
  61. Fernandes C. M.; Alvarez L. X.; Escarpini-dos-Santos N.; Maldonado-Barrios A. C.; Ponzio E. A. Green synthesis of 1-benzyl-4-phenyl-1H-1,2,3-triazole, its application as corrosion inhibitor for mild steel in acidic medium and new approach of classical electrochemical analyses. Corros. Sci. 2019, 149, 185–194. 10.1016/j.corsci.2019.01.019. [DOI] [Google Scholar]
  62. Kowsari E.; Payami M.; Amini R.; Ramezanzadeh B.; Javanbakht M. Task-specific ionic liquid as a new green inhibitor of mild steel corrosion. Appl. Surf. Sci. 2014, 289, 478–486. 10.1016/j.apsusc.2013.11.017. [DOI] [Google Scholar]
  63. Tang Y.; Zhang F.; Hu S.; Cao Z.; Wu Z.; Jing W. Novel benzimidazole derivatives as corrosion inhibitors of mild steel in the acidic media. Part I: Gravimetric, electrochemical, SEM and XPS studies. Corros. Sci. 2013, 74, 271–282. 10.1016/j.corsci.2013.04.053. [DOI] [Google Scholar]
  64. Okafor P. C.; Zheng Y. Synergistic inhibition behaviour of methylbenzyl quaternary imidazoline derivative and iodide ions on mild steel in H2SO4 solutions. Corros. Sci. 2009, 51, 850–859. 10.1016/j.corsci.2009.01.027. [DOI] [Google Scholar]
  65. Umoren S. A.; Ebenso E. E.; Okafor P. C.; Ekpe U. J.; Ogbobe O. Effect of halide ions on the corrosion inhibition of aluminium in alkaline medium using polyvinyl alcohol. J. Appl. Polym. Sci. 2007, 103, 2810–2816. 10.1002/app.25446. [DOI] [Google Scholar]
  66. Bentiss F.; Lebrini M.; Traisnel M.; Lagrenée M. Synergistic effect of iodide ions on inhibitive performance of 2,5-bis(4-methoxyphenyl)-1,3,4-thiadiazole during corrosion of mild steel in 0.5 M sulfuric acid solution. J. Appl. Electrochem. 2009, 39, 1399–1407. 10.1007/s10800-009-9810-9. [DOI] [Google Scholar]
  67. Sastri V. S.Green Corrosion Inhibitors: Theory and Practice; John Wiley & Sons, Inc., 2011. [Google Scholar]
  68. Godínez L. A.; Meas Y.; Ortega-Borges R.; Corona A. Corrosion inhibitors. Rev. Metal. 2003, 39, 140–158. [Google Scholar]
  69. Cha Y.-J.; Kim M.-J.; Choa Y.-H.; Kim J.; Nam B.; Lee J.; Kim D. H.; Kim K. H. Synthesis and Characterizations of Surface-Coated Superparamagnetic Magnetite Nanoparticles. IEEE Trans. Magn. 2010, 46, 443–446. 10.1109/TMAG.2009.2033205. [DOI] [Google Scholar]
  70. Raghu M. S.; Kumar K. Y.; Prashanth M. K.; Prasanna B. P.; Vinuth R.; Pradeep-Kumar C. B. Adsorption and antimicrobial studies of chemically bonded magnetic graphene oxide-Fe3O4 nanocomposite for water purification. J. Water Process. Eng. 2017, 17, 22–31. 10.1016/j.jwpe.2017.03.001. [DOI] [Google Scholar]
  71. Lachkar-Zamouri O.; Brahim K.; Bennour-Mrad F.; Khattech I. Attack of Tunisian Phosphate Ore by a Mixture of Sulfuric and Phosphoric Acid: Thermochemical Study by Means of Differential Reaction Calorimetry. Adv. Mater. Phys. Chem. 2018, 8, 411–427. 10.4236/ampc.2018.810028. [DOI] [Google Scholar]
  72. Kurniawan C.; Eko A. S.; Ayu Y. S.; Sihite P. T. A.; Ginting M.; Simamora P.; Sebayang P.. Synthesis and Characterization of Magnetic Elastomer based PEG-Coated Fe3O4 from Natural Iron Sand. In IOP Conference Series. Materials Science and Engineering; IOP Publishing, 2017. [Google Scholar]
  73. Chirita M.; Banica R.; Ieta A.; Grozescu I. Superparamagnetic Unusual Behavior of Micrometric Magnetite Monodisperse Monocrystals Synthesized by Fe-EDTA Thermal Decomposition. Part. Sci. Technol. 2012, 30, 354–363. 10.1080/02726351.2011.585220. [DOI] [Google Scholar]
  74. Nguyen Q.; Quyen D.; Hoang T. A new route of emulsifier-free emulsion polymerization for the preparation of polymer coated magnetite nanoparticles. Mater. Sci.-Pol. 2014, 32, 264–271. 10.2478/s13536-013-0172-y. [DOI] [Google Scholar]
  75. Hou Y.; Li J.; Gao X.; Wen Z.; Yuan C.; Chen J. 3D Dual-Confined Sulfur Encapsulated in Porous Carbon Nanosheets and Wrapped with Graphene Aerogels as Cathode for Advanced Lithium Sulfur Batteries. Nanoscale 2016, 8, 8228–8235. 10.1039/C5NR09037G. [DOI] [PubMed] [Google Scholar]
  76. Liu H.; Xu C.-Y.; Du Y.; Ma F.-X.; Li Y.; Yu J.; Zhen L. Ultrathin Co9s8 nanosheets vertically aligned on N,s/rGo for low voltage electrolytic water in alkaline media. Sci. Rep. 2019, 9, 1951 10.1038/s41598-018-35831-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Hummel H.-U.; Förner W. Investigation of the Charge Distribution of Free and Bound 1,1-Dicyanoethylene-2,2-dithiolate. An ab initio HF-MO and XPS Study. J. Inorg. Gen. Chem. 1986, 540, 300–306. 10.1002/zaac.19865400933. [DOI] [Google Scholar]
  78. Ning L.; Wang D.; Wang L.; Wu L.; Yang J.; Wang X.; Ma H.; Feng S.; Lu H. Interesting Corrosion Inhibition Performance and Mechanism of Two Silanes Containing Multiple Phosphate Group. Silicon 2020, 12, 1455–1468. 10.1007/s12633-019-00236-z. [DOI] [Google Scholar]
  79. Orfanoudaki T.; Skodras G.; Dolios I.; Sakellaropoulos G. P. Production of carbon molecular sieves by plasma treated activated carbon fibers. Fuel 2003, 82, 2045–2049. 10.1016/S0016-2361(03)00172-8. [DOI] [Google Scholar]
  80. Moura C. A. d. S.; Belmonte G. K.; Reddy P. G.; Gonslaves K. E.; Weibel D. E. EUV photofragmentation study of hybrid nonchemically amplified resists containing antimony as an absorption enhancer. RSC Adv. 2018, 8, 10930. 10.1039/C7RA12934C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Belmonte G. K.; Moura C. A. d. S.; Reddy P. G.; Gonsalves K. E.; Weibel D. E. EUV photofragmentation and oxidation of a polyarylene—Sulfonium resist: XPS and NEXAFS study. J. Photochem. Photobiol., A 2018, 364, 373–381. 10.1016/j.jphotochem.2018.06.005. [DOI] [Google Scholar]
  82. Gassman P. G.; Callstrom M. R.; et al. XPS Data for Linear Three-Center, Four-Electron Bonding in Sulfur Species. J. Am. Chem. Soc. 1988, 110, 8724–8725. 10.1021/ja00234a040. [DOI] [Google Scholar]
  83. Kelemen S. R.; George G. N.; Gorbaty M. L. Direct determination and quantification of organic sulfur forms by X-ray photoelectron spectroscopy (XPS) and sulfur k-edge absorption spectroscopy. Fuel Process. Technol. 1990, 24, 425–429. 10.1016/0378-3820(90)90082-4. [DOI] [Google Scholar]
  84. Zhang Q. H.; Hou B. S.; Xu N.; Xiong W.; Liu H. F.; Zhang G. A. Effective inhibition on the corrosion of X65 carbon steel in the oilfield produced water by two Schiff bases. J. Mol. Liq. 2019, 285, 223–236. 10.1016/j.molliq.2019.04.072. [DOI] [Google Scholar]
  85. Allen G. C.; Curtis M. T.; Hooper A. J.; Tucker P. M. X-Ray Photoelectron Spectroscopy of Iron-Oxygen Systems. J. Chem. Soc., Dalton Trans. 1974, 52, 1445–1451. 10.1039/DT9740001525. [DOI] [Google Scholar]
  86. Wagner C. D.; Zatko D. A.; Raymond R. H. Use of the oxygen KLL Auger lines in identification of surface chemical states by electron spectroscopy for chemical analysis. Anal. Chem. 1980, 52, 1445. 10.1021/ac50059a017. [DOI] [Google Scholar]
  87. Siriwardane R. V.; Cook J. M. Interactions of SO2 with Sodium Deposited on CaO. J. Colloid Interface Sci. 1986, 114, 525. 10.1016/0021-9797(86)90438-8. [DOI] [Google Scholar]
  88. Hallam P. M.; Gómez-Mingot M.; Kampouris D. K.; Banks C. E. Facile synthetic fabrication of iron oxide particles and novel hydrogen superoxide supercapacitors. RSC Adv. 2012, 2, 6672. 10.1039/C2RA01139E. [DOI] [Google Scholar]
  89. Puech L.; Dubarry C.; Ravel G.; de-Vito E. Modeling of iron oxide deposition by reactive ion beam sputtering. J. Appl. Phys. 2010, 107, 054908 10.1063/1.3327431. [DOI] [Google Scholar]
  90. Luo J.; Du X.; Gao F.; Yang Y.; Hao X.; Li S.; Hao X.; Tang K.; Guan G. Iodide ion trapping polypyrrole film: Selective capture of iodide ions by electrochemically switched ion extraction (ESIE) process. Chem. Eng. J. 2019, 380, 122529 10.1016/j.cej.2019.122529. [DOI] [Google Scholar]
  91. Qian B.; Wang J.; Zheng M.; Hou B. Synergistic effect of polyaspartic acid and iodide ion on corrosion inhibition of mild steel in H2SO4. Corros. Sci. 2013, 75, 184–192. 10.1016/j.corsci.2013.06.001. [DOI] [Google Scholar]
  92. Chesnut D. B. Structure and bonding in some S-methylsulfonium halides. Heteroat. Chem. 2005, 16, 263–270. 10.1002/hc.20087. [DOI] [Google Scholar]
  93. Haque J.; Srivastava V.; Verma C.; Lgaz H.; Salghi R.; Quraishi M. A. N-Methyl-N,N,N-trioctylammonium chloride as a novel and green corrosion inhibitor for mild steel in an acid chloride medium: electrochemical, DFT and MD studies. New J. Chem. 2017, 41, 13647–13662. 10.1039/C7NJ02254A. [DOI] [Google Scholar]
  94. Zuriaga-Monroy C.; Oviedo-Roa R.; Montiel-Sánchez L. E.; Vega-Paz A.; Marín-Cruz J.; Martínez-Magadán J.-M. Theoretical Study of the Aliphatic-Chain Length’s Electronic Effect on the Corrosion Inhibition Activity of Methylimidazole-Based Ionic Liquids. Ind. Eng. Chem. Res. 2016, 55, 3506–3516. 10.1021/acs.iecr.5b03884. [DOI] [Google Scholar]
  95. Kurnia K. A.; Neves C. M. S. S.; Freire M. G.; Santos L. M. N. B. F.; Coutinho J. A. P. Comprehensive Study on the Impact of the Cation Alkyl Side Chain Length on the Solubility of Water in Ionic Liquids. J. Mol. Liq. 2015, 210, 264–271. 10.1016/j.molliq.2015.03.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Sampathkumar S.; Subramaniam V. A density functional theory study on the water aggregation behaviour of fatty acid-based anionic surface active ionic liquids. Struct. Chem. 2022, 33, 961–972. 10.1007/s11224-022-01910-6. [DOI] [Google Scholar]
  97. Znini M.; Ansari A.; Costa J.; Senhaji O.; Paolini J.; Majidi L. Experimental, Quantum Chemical and Molecular Dynamic Simulations Studies on the Corrosion Inhibition of C38 Steel in 1M HCl by Anethum graveolens Essential Oil. Anal. Bioanal. Electrochem. 2019, 11, 1426–1451. [Google Scholar]
  98. 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. 10.1016/j.corsci.2013.09.009. [DOI] [Google Scholar]
  99. Zarrouk A.; Zarrok H.; Salghi R.; Hammouti B.; Al-Deyab S. S.; Touzani R.; Bouachrin M.; Warad I.; Hadda T. B. A Theoretical Investigation on the Corrosion Inhibition of Copper by Quinoxaline Derivatives in Nitric Acid Solution. Int. J. Electrochem. Sci 2012, 7, 6353–6364. 10.1016/j.jscs.2011.09.011. [DOI] [Google Scholar]
  100. Yilmaz N.; Fitoz A.; Ergun ÿ.; Emregül K. C. A combined electrochemical and theoretical study into the effect of 2-((thiazole-2-ylimino)methyl)phenol as a corrosion inhibitor for mild steel in a highly acidic environment. Corros. Sci. 2016, 111, 110–120. 10.1016/j.corsci.2016.05.002. [DOI] [Google Scholar]
  101. Musa A. Y.; Mohamad A. B.; Kadhum A. A. H.; Takriff M. S.; Tien L. T. Synergistic effect of potassium iodide with phthalazone on the corrosion inhibition of mild steel in 1.0M HCl. Corros. Sci. 2011, 53, 3672–3677. 10.1016/j.corsci.2011.07.010. [DOI] [Google Scholar]
  102. Umoren S. A.; Eduok U. M. Application of carbohydrate polymers as corrosion inhibitors for metal substrates in different media: A review. Carbohydr. Polym. 2016, 140, 314–341. 10.1016/j.carbpol.2015.12.038. [DOI] [PubMed] [Google Scholar]
  103. Likhanova N. V.; Domínguez-Aguilar M. A.; Olivares-Xometl O.; Nava-Entzana N.; Arce E.; Dorantes H. The effect of ionic liquids with imidazolium and pyridinium cations on the corrosion inhibition of mild steel in acidic environment. Corros. Sci. 2010, 52, 2088–2097. 10.1016/j.corsci.2010.02.030. [DOI] [Google Scholar]
  104. Sokka I.; Fischer A.; Kloo L. Optimization of the synthesis of non-symmetrical alkyl dimethyl sulfonium halides. Polyhedron 2007, 26, 4893–4898. 10.1016/j.poly.2007.06.032. [DOI] [Google Scholar]
  105. ASTM G1-03-(2017). Preparing, Cleaning and Evaluating Corrosion Test Specimens; West Conshohocken: PA, 2017. 10.1520/G0001-03R11. [DOI] [Google Scholar]
  106. ASTM G59-97-(2009). Conducting Potentiodynamic Polarization Resistance Measurements; West Conshohocken: PA, 2009. 10.1520/G0059-97R09. [DOI] [Google Scholar]
  107. ASTM G3-14-(2019). Conventions Applicable to Electrochemical Measurements in Corrosion Testing; West Conshohocken: PA, 2019. 10.1520/G0003-14R19. [DOI] [Google Scholar]
  108. ASTM G106-89-(2015). Verification of Algorithm and Equipment for Electrochemical Impedance Measurements; West Conshohocken: PA, 2015. 10.1520/G0106-89R15. [DOI] [Google Scholar]
  109. ASTM G185-06-(2016). Evaluating and Qualifying Oil Field and Refinery Corrosion Inhibitors Using the Rotating Cylinder Electrode; West Conshohocken: PA, 2016. 10.1520/g0185-06r16. [DOI] [Google Scholar]
  110. Childs P. R. N.Rotating Flow; Elsevier Ltd, 2011. 10.1016/C2009-0-30534-6. [DOI] [Google Scholar]
  111. Ibrahim I. M.; Jai J.; Daud M.; Hashim M. A. Inhibition effect of fatty amides with secondary compound on carbon steel corrosion in hydrodynamic condition. IOP Conf. Ser.: Mater. Sci. Eng. 2018, 334, 012048 10.1088/1757-899x/334/1/012048. [DOI] [Google Scholar]
  112. ASTM G170-06-(2012). Evaluating and Qualifying Oilfield and Refinery Corrosion Inhibitors in the Laboratory; West Conshohocken: PA, 2012. 10.1520/g0170-06r12. [DOI] [Google Scholar]
  113. Barmatov E.; Hughes T.; Nagl M. Efficiency of film-forming corrosion inhibitors in strong hydrochloric acid under laminar and turbulent flow conditions. Corros. Sci. 2015, 92, 85–94. 10.1016/j.corsci.2014.11.038. [DOI] [Google Scholar]
  114. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; et al. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2013.
  115. Easton R. E.; Giesen D. J.; Welch A.; Cramer C. J.; Truhlar D. G. The MIDI! basis set for quantum mechanical calculations of molecular geometries and partial charges. Theor. Chim. Acta 1996, 93, 281–301. 10.1007/BF01127507. [DOI] [Google Scholar]
  116. Thompson J. D.; Winget P.; Truhlar D. G. MIDIX basis set for the lithium atom: Accurate geometries and atomic partial charges for lithium compounds with minimal computational cost. PhysChemComm 2001, 4, 72–77. 10.1039/b105076c. [DOI] [Google Scholar]
  117. Singh A.; Ansari K. R.; Quraishi M. A.; Lin Y.. Investigation of Corrosion Inhibitors Adsorption on Metals Using Density Functional Theory and Molecular Dynamics Simulation. In Corrosion Inhibitors, Singh A., Ed.; IntechOpen, 2019. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao2c05192_si_001.pdf (94KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES