Abstract

Aliphatic tricationic surfactants were prepared by the esterification reaction, followed by a quaternization reaction to protect oil well facilities from corrosion problems. Microelemental analysis and Fourier transform infrared and 1H NMR spectroscopic techniques were performed to explore the obtained motifs. The performance of these amphiphiles as inhibitors for metallic S90 steel corrosion in formation water was investigated through electrochemical tests (potentiodynamic polarization and electrochemical impedance spectroscopy). The results revealed significant inhibition effectiveness improvement with increasing concentrations of these amphiphiles. Its maximum inhibition efficiency reaches 93.07% at 250 ppm for the compound (AED). Potentiodynamic polarization graphs demonstrated that tricationic amphiphiles behave as mixed-type inhibitors. In addition, the adsorption of the tricationic surfactant at the S90 steel surface followed Langmuir isotherm. Atomic force microscopy revealed that a protective layer formed at the surface of S90 steel caused the inhibition of corrosion. During the inhibition procedure of S90 steel corrosion, theoretical research has been performed to validate electrochemical experiments and to clearly demonstrate the mechanism of these amphiphiles. Finally, quantum chemical calculations were calculated to achieve the justification for the obtained empirical results.
1. Introduction
Metallic equipment is likely to be exposed to corrosion and destructive attack through the environmental reaction in addition to the potential natural hazards associated with the handling and transportation of gas and oil.1 Corrosion can occur because of the presence of metallic equipment in any watery environment, which arises under many complex conditions in gas and petroleum output, refining, and pipeline networks.2 Several high-impurity products may be produced from crude oil and natural gas, which are inherently corrosive. Carbon dioxide, hydrogen sulfide, and formation water, which are described as highly corrosive media, can be formed in oil and gas fields.3 Notably, continuous removal of carbon dioxide, hydrogen sulfide, and formation water from gas and petroleum equipment leads to a potential effect of corrosion on the inner surfaces of this equipment over time. Moreover, the corrosion processes of metallurgical equipment play a remarkable role in the economics and safety of oil fields.4
This material degradation results in a loss of mechanical properties such as plasticity, strength, impact strength, and so forth. This results in loss of material, decreased density, and sometimes terminal failure.5 Finally, the ingredient can fully be disintegrated, and the installation must be changed when the production is completely finished. The severe effects of this corrosion process have become a global problem. As the corrosion cannot be prevented, monitoring the rate of corrosion could be the most cost-effective approach. Lately, the use of surfactants as inhibitors of corrosion was commonly utilized to secure metallic equipment against corrosive media. In addition, trimeric cationic amphiphiles have been found to be a new type of surfactant molecules.6 This kind of amphiphile received tremendous interest in industrial and academic laboratories. It comprises three monocationic amphiphile molecules which are chemically linked to each other through a spacer group.7,8 Generally, amphiphile corrosion inhibitors were exploited as a cost-effective process and as among the most efficient means of protecting S90 steel against serious corrosion assaults and became increasingly widespread in recent times.8,9
In general, the most commonly used corrosion inhibitors for metallic equipment in oilfields include heterocyclic atoms such as P, N, O, and S and aromatic rings in their structures that may be absorbed onto the surface of the metallic equipment and subsequent to create film blocking, which is an obstacle to the induced media for corrosion to attack the surface of the metal equipment, thus reducing corrosion.10−14 When choosing a form inhibitor, various considerations such as cost, availability, and toxicity should be taken into account. In environmental and health aspects, the prepared trimeric cationic amphiphiles are extremely important to inhibit corrosion in S90 steel because of its high inhibition capacity, ease of manufacture, low cost, small toxic effects, and biodegradability.8,15 The biodegradable character of these aliphatic tricationic surfactants can be ascribed to the existence of triester groups and ethylene oxide, and the structure of these compounds is an aliphatic.16−19 It is well known that aliphatic-based compounds are better biodegradable than aromatic-based compounds.18,20 The presence of ester linkages made these structures cleavable surfactants and have good biodegradability.21−23 In addition, the existence of ethylene oxide in the structure of aliphatic tri-cationic amphiphiles lead to the decrease in surfactant toxicities.16 Empirical studies are helpful to explain the process of inhibition, but it is often costly and time-consuming. Therefore, theoretical calculations have been employed to explore the mechanism of corrosion inhibition. Two main advantages could be easily obtainedvia the usage of theoretical parameters: first, chemicals of their different fractions and moieties could be immediately described based on only their molecular structure; and second, the suggested interaction mechanism could be immediately attributed to the chemical reactivity of the compounds.24 Furthermore, quantum chemical calculations can provide insights into the relation between corrosion inhibition efficiency (IE) and electronic properties of the compounds.25−28 The studied compounds having heteroatoms as oxygen, nitrogen, hydroxyl groups, and alkyl chains increase the electron density on the surfactants molecules. Consequently, the studied surfactants exposed to behave as good corrosion inhibitors.
The objective of this work is to synthesize three aliphatic tricationic amphiphiles based on esterification and quaternization reactions and to study their inhibition effectiveness as operational corrosion inhibitors for carbon steel in deep oil well formation water through polarization and electrochemical impedance spectroscopy (EIS) tests. Herein, our work is extended to discuss the impact of the structural factors of inhibitors on inhibition effectiveness. In addition, our work is also discussing its adsorption process on the carbon steel surface and thus relating the empirical data with theoretical calculations of these synthesized tricationic amphiphiles. Finally, atomic force microscopy (AFM) analysis was employed as an effective surface analysis tool to examine the morphology of S90 steel with and without inhibitor molecules.
This work gives a proof for the application of the newly synthesized aliphatic tricationic amphiphiles to oppose corrosive attacks produced by oil well formation water over S90 carbon steel samples used widely in oil fields. To the best of our knowledge, most studies focused on compounds bearing both aliphatic and aromatic moieties, but this study focused on aliphatic surfactants to assess their sole contribution to the aspired corrosion inhibition. This is due to the fact that aliphatic tricationic surfactants are more water-soluble and more biodegradable, so they are more convenient to be used.20,29 Driven by the placement of pertinent active moieties within the designed structure of the aliphatic tricationic surfactants showed high anticorrosion action in oil fields with low dosage compared to earlier researches recorded in Table 1. Consequently, aliphatic tricationic surfactants can serve as a good option to mitigate corrosion problem.
Table 1. Comparison of the Corrosion IE of the Prepared Trimeric Cationic Surfactants with Some Previous Studies.
| inhibitor name | inhibitor type | aggressive solution | optimum dose (mg/L) | max. inhibition (%) | refs |
|---|---|---|---|---|---|
| E11N8 | gemini amphiphile | oilfield produced water | 300 | 81.3 | (4) |
| l-arginine | amino acid | artificial reservoir water | 174 | 88.8 | (30) |
| TC18 | tricationic amphiphile | oilfield produced water | 300 | 89.9 | (8) |
| S1E | cationic surfactant | 0.5 M HCl | 500 | 90.8 | (31) |
| TRE | natural product | synthetic seawater | 150 | 86.3 | (23) |
| Gly-PASP | polymer | seawater | 250 | 83.1 | (32) |
| DHPB | cationic surfactant | 2 M HCl | 300 | 92.0 | (33) |
| (PASP/β-CD) | polymer | synthetic oil well water | 800 | 68.4 | (34) |
| AED | tricationic surfactant | formation water | 250 | 93.1 | this work |
2. Results and Discussion
2.1. Structure Confirmation of the Prepared Tricationic Aliphatic Amphiphiles
The structures of the synthesized compounds were elucidated via microelemental analysis, Fourier transform infrared (FTIR), and 1H NMR spectroscopy. Table 2 shows the achieved and computed percentages for the several components present in these prepared tricationic aliphatic amphiphiles, indicating their pureness. The FTIR spectrum of AEB as a representative example of other aliphatic tricationic surfactants showed the subsequent absorption bands: 3451 cm–1 is attributed to O–H stretching, 2962 cm–1 corresponds to CH3 stretching, 2875 cm–1 represents CH2 stretching, 1739 cm–1 corresponds to C=O, 1456 cm–1 is attributed to CH2 bending, 1378 cm–1 is attributed to CH3 bending, 1189 cm–1 corresponds to C–O stretching of ester, 1034 cm–1 corresponds to C–O stretching of alcohol, 1131 cm–1 corresponds to C–N+ stretching, 966 cm–1 corresponds to O–H bending, and 744 cm–1 is attributed to C–H rocking. The FTIR spectrum of tricationic aliphatic amphiphile (AEB) is displayed in Figure 1. The 1H NMR spectrum of AEB as an illustrative example of the other aliphatic tricationic amphiphiles in DMSO-d6 (Figure 2) showed signals at 0.91–0.94 ppm [t, 27H, (CH3(CH2)3N+)9], 1.16–1.35 ppm [m, 18H, (CH3CH2CH2CH2N–)9], 1.59 ppm [m, 18H, (CH3CH2CH2CH2N–)9], 2.74 ppm [s, 4H, (CH2C=O)2], 2.78–2.95 ppm [m, 18H, (CH3CH2CH2CH2N–)9], 3.64 ppm [t, 6H, (O=COCH2CH2OCH2CH2N+)3], 3.74 ppm [t, 6H, (O=COCH2CH2OCH2CH2N+)3], 4.02 ppm [t, 6H, (O=COCH2CH2OCH2CH2N+)3], 4.1 ppm [t, 6H, (O=COCH2CH2OCH2CH2N+)3], 4.2 ppm [s, 1H, OH]. The 1H NMR spectrum of the tricationic surfactant (AEB) is displayed in Figure 2. The synthesized surfactants were freely soluble in water because of the existence of extremely polar triquaternary ammonium groups and the presence of ethylene oxide and ester groups, which affected this extraordinary solubility property. This prepared compound was obtained as a yellowish brown liquid with yield = 82.9%. The results proved the anticipated structure of these prepared trimeric cationic aliphatic surfactants signified in Scheme 1.
Table 2. Elemental Analysis of the Prepared Tricationic Aliphatic Amphiphiles.
| C (%) |
H (%) |
Cl (%) |
N (%) |
||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|
| compound | mol formula | MW (g/mol) | yield (%) | calcd | found | calcd | found | calcd | found | calcd | found |
| AE | C18H29Cl3O10 | 511.77 | 89 | 42.25 | 42.06 | 5.71 | 5.66 | 20.78 | 20.63 | ||
| AEB | C54H110Cl3N3O10 | 1065.73 | 83 | 60.74 | 60.62 | 10.38 | 10.32 | 9.96 | 9.87 | 3.94 | 3.89 |
| AEO | C90H182Cl3N3O10 | 1572.81 | 75 | 68.73 | 68.59 | 11.66 | 11.61 | 6.76 | 6.65 | 2.67 | 2.63 |
| AED | C126H254Cl3N3O10 | 2077.78 | 66 | 72.84 | 72.68 | 12.32 | 12.27 | 5.12 | 5.01 | 2.02 | 1.98 |
Figure 1.
IR spectrum of compound (AEB).
Figure 2.
1H NMR spectrum of compound (AEB).
Scheme 1. Preparation of the Novel Aliphatic Tricationic Surfactants (AER).
In addition, the first stage in the mechanism is an esterification reaction of 2-(2-chloroethoxy)ethanol with citric acid to produce tris(2-(2-chloroethoxy)ethyl)2-hydroxypropane-1,2,3-tricarboxylate (AE), which was approved by the presence of the novel absorption band for ester groups at 1739.6 cm–1 and evanescence of either hydroxyl or carbonyl groups for acid at 2500–3300 and 1715 cm–1, respectively. The FTIR spectra revealed the subsequent absorption bands: 3493 cm–1 represents (O–H stretching alcohol group of citric acid), 2961 cm–1 is attributed to (CH3 stretching), 2876 cm–1 corresponds to (CH2 stretching), 1739 cm–1 is attributed to (C=O stretching of ester group), 1434 cm–1 is attributed to (CH2 bending), 1388 and 966 cm–1 are attributed to (O–H bending), 1300 cm–1 corresponds to (C–H wag terminal alkyl halides), 1192 and 1131 cm–1 correspond to (C–O stretching of ester), 1036 cm–1 corresponds to (C–O stretching of alcohol), and 746 and 664 cm–1 correspond to (C–Cl stretching). The FTIR spectrum of the prepared tri-chloro alkyl triester (AE) is presented in Figure S1 (Supporting Information). Besides, the 1H NMR spectrum of the synthesized tri-chloro alkyl triester (AE) in CDCl3 exhibited signals at δ = 2.835–2.984 ppm [m, 4H, (CH2C=O)2], 3.62–3.64 ppm [t, 6H, (OCH2CH2OCH2CH2Cl)3], 3.724 ppm [t, 6H, (OCH2CH2OCH2CH2Cl)3], 4.236 ppm [t, 6H, (OCH2CH2OCH2CH2Cl)3], 4.388 ppm [t, 6H, (OCH2CH2OCH2CH2Cl)3], 4.991 ppm [s, 1H, OH]. The 1H NMR spectrum of the prepared tri-chloro alkyl tri-ester (AE) is showed in Figure S2 (Supporting Information).
2.2. Corrosion Inhibition Evaluation
2.2.1. Potentiodynamic Polarization Studies
The potentiodynamic polarization (PP) curves of S90 type carbon steel in deep oil well produced water without and with various doses of the AED surfactant at 298 K (Figure 3). The AED scaffold is considered as a demonstrative sample for the functionality of these aliphatic tricationic amphiphiles. Other polarization curves of AEB and AEO are given in the Supporting Information (Figure S4). Electrochemical variables obtained from the extrapolation of Tafel curves are recorded in Table 3. The IEs (%) of these synthesized aliphatic tricationic amphiphiles were estimated as follows35,36
| 1 |
where Icorr° and Icorr are corrosion current densities without and with tricationic amphiphiles, respectively.
Figure 3.
Polarization curves for S90 steel in oil well produced water without and with different concentrations of AED at 298 K.
Table 3. Tafel Polarization Parameters for S90 Steel in Oil Well Produced Water without and with Different Concentrations of Aliphatic Tricationic Amphiphiles.
| inhibitor | concn (ppm) | –Ecorr (V) | icorr (μA/cm2) | corr. rate (μm/year) | –βc (V/dec) | βa (V/dec) | θ | IE (%) |
|---|---|---|---|---|---|---|---|---|
| Blank | 0.682 ± 0.09 | 26.09 ± 1.2 | 195.2 | 0.132 | 0.079 | |||
| AEB | 50 | 0.712 ± 0.11 | 13.161 ± 0.7 | 98.463 | 0.124 | 0.121 | 0.4956 | 49.56 |
| 100 | 0.678 ± 0.07 | 10.194 ± 0.9 | 76.259 | 0.152 | 0.127 | 0.6093 | 60.93 | |
| 150 | 0.691 ± 0.04 | 7.423 ± 0.6 | 55.542 | 0.186 | 0.102 | 0.7155 | 71.55 | |
| 200 | 0.809 ± 0.03 | 4.924 ± 0.5 | 36.827 | 0.041 | 0.153 | 0.8113 | 81.13 | |
| 250 | 0.790 ± 0.08 | 3.06 ± 0.3 | 22.904 | 0.052 | 0.115 | 0.8827 | 88.27 | |
| AEO | 50 | 0.796 ± 0.07 | 12.075 ± 1.1 | 90.341 | 0.064 | 0.168 | 0.5372 | 53.72 |
| 100 | 0.770 ± 0.12 | 9.716 ± 0.6 | 72.687 | 0.169 | 0.065 | 0.6276 | 62.76 | |
| 150 | 0.790 ± 0.05 | 7.185 ± 0.9 | 53.751 | 0.056 | 0.238 | 0.7246 | 72.46 | |
| 200 | 0.768 ± 0.09 | 4.371 ± 0.5 | 32.689 | 0.053 | 0.207 | 0.8325 | 83.25 | |
| 250 | 0.781 ± 0.06 | 2.454 ± 0.2 | 18.372 | 0.211 | 0.052 | 0.9059 | 90.59 | |
| AED | 50 | 0.792 ± 0.08 | 11.093 ± 0.9 | 82.993 | 0.128 | 0.084 | 0.5748 | 57.48 |
| 100 | 0.669 ± 0.05 | 8.781 ± 0.7 | 65.714 | 0.094 | 0.183 | 0.6634 | 66.34 | |
| 150 | 0.779 ± 0.03 | 6.427 ± 0.4 | 48.085 | 0.135 | 0.057 | 0.7537 | 75.37 | |
| 200 | 0.774 ± 0.10 | 4.129 ± 0.3 | 30.904 | 0.109 | 0.071 | 0.8417 | 84.17 | |
| 250 | 0.795 ± 0.07 | 1.807 ± 0.1 | 13.527 | 0.054 | 0.139 | 0.9307 | 93.07 |
As can be seen from Table 3, the values of Icorr diminish with the increase of the added concentration of these tricationic aliphatic amphiphiles that show excellent anticorrosion effectiveness for S90 steel in oil well produced water. As a result, the corrosion IE was boosted by increasing tricationic aliphatic amphiphile concentration until it approached its maximum value. This evidence revealed that the stable adsorbed layer of tricationic surfactant particles had been molded.37 In addition, the inhibition activity of tricationic aliphatic amphiphiles improved by increasing the alkyl chain length. For that reason, tricationic aliphatic amphiphiles could be organized according to their effectiveness in anticorrosion as follows: AED > AEO > AEB.
As shown in Figure 3, the values of corrosion potential (Ecorr) without and with surfactant inhibitors have been changed to more positive and negative potentials.38,39 From Table 3, the shift in Ecorr values was all less than 85 mV, indicating that the studied tricationic amphiphiles were mixed-type inhibitors because they could decrease cathodic and anodic reactions via creating the protective layer on the S90 steel surface in oil well formation water.40 Furthermore, Tafel slopes (βa and βc) with various concentrations of tricationic amphiphiles were approximately constant, indicating that the tricationic surfactants hindered the corrosion of S90 steel by inhibiting both anodic and cathodic reactions, without altering the corrosion reaction mechanism.41−43 The absorption of tricationic aliphatic surfactants on the S90 steel surface inhibits the corrosion process. This can be achieved through the interaction mechanism among the lone pairs of oxygen and nitrogen atoms that exist in ester and quaternary ammonium groups and the empty d-orbital of the S90 steel surface via coordination bonds. Consequent protection is obtained with the formation of a barrier layer and leads to corrosion obstruction.
2.2.2. Electrochemical Impedance Spectroscopy
Corrosion behavior of S90 steel displayed for oil well produced water in the absence and presence of several concentrations of the tricationic aliphatic surfactants was exhibited via EIS measurements. The Nyquist plot displayed in Figure 4 describes the impact of different doses of the AED surfactant, whereas AEB and AEO surfactants are summarized in the Supporting Information (Figure S5). The Bode plot for AED is exhibited in Figure 5, while AEB and AEO Bode plots have been summarized in the Supporting Information (Figure S6). Checking the previously aforementioned curves exposes that the impedance response of S90 steel in oil well produced water was concurrently moved after adding these tricationic aliphatic surfactants.
Figure 4.
Nyquist plots for S90 steel in oil well produced water without and with various doses of AED at 298 K.
Figure 5.
Bode plots for S90 steel in oil well produced water without and with various doses of AED at 298 K.
It is clear that two capacitive loops were observed without surfactant inhibitors because of the creation of insoluble oxide layer on the S90 steel surface.44,45 Furthermore, the impedance response of S90 steel in oil well produced water shows considerable change after adding the tricationic amphiphiles. This confirms the inhibition effect of these tricationic surfactants on the corrosion of S90 steel in oil well produced water. Additionally, the progressive growth of various tricationic surfactants in oil well produced water has a large increasing effect on the diameters of the capacitive loops. The experimental impedance spectra were simulated via an equivalent circuit model described in Figure 6. The model includes film resistance (Rf), solution resistance (Rs), double-layer capacitance (Cdl), film capacitance (Cf), and the charge-transfer resistance (Rct).46 All impedance spectra show two capacitive circuits consisting of a small circle at high frequencies and a big incomplete circle at low frequencies. The capacitance circuit has resulted from a frequency dispersion impact and charge transfer of electrochemical corrosion.47−49 The roughness and inhomogeneity on the surface of S90 steel electrode decreased in the higher frequency circle.4
Figure 6.

Equivalent circuit diagram for the studied tricationic surfactant inhibitors.
The diameter of the capacitance circuits elevated with increasing concentration of these aliphatic tricationic amphiphiles that approved the creation of the adsorbed protective layer on the S90 steel surface.49,50 Furthermore, Nyquist plots without and with tricationic surfactants were not noticeably changed. This emphasizes that the corrosion mechanism was not modified by adding tricationic surfactant inhibitors in oil well produced water.8 As seen in Table 4, the values of Rct and IE % increase, while the values of Cdl minimize with increasing dosage of tricationic amphiphiles. To the best of our knowledge, the tricationic surfactant molecules act via the adsorption process on the S90 steel surface in oil well produced water. The reason behind that is the diminishing effect on local dielectric constant and increasing density of the electrical double layer.49,51 In addition, Table 4 reveals that the Rf values boosted with the increase of the concentrations of tricationic amphiphiles, while the Cf values declined. The rational for the preceding manner might be caused by the gradual replacement of adsorbed water molecules by adsorbed tricationic surfactant particles, reducing the amount of S90 steel dissolution, thereby reducing severe corrosion attacks.52 A marginal rise in solution resistance (Rs) was observed with the growth in the added tricationic surfactant dosage. This could be attributed to a reduced cross section of solution exposed to a current passage owing to the large surface covering of the S90 carbon steel surface.53 A constant phase element (CPE) was inserted into a circuit in place of a pure Cdl to provide the best fit.54 Therefore, impedance using CPE can be expressed as follows
| 2 |
where Y0 is the magnitude of the CPE, −1 ≤ n ≤ 1. Therefore, in case of the high-frequency circles were a reduced semicircular look, 0.5 ≤ n ≤ 1. This might be caused by a frequency dispersion because of a nonhomogeneity surface of the metallic surface.55 Bode graphs showed that slopes of log Z versus log f lines were unequal −1. This perversion is caused by heterogeneity of the S90 steel surface. The values of IE (%) in Table 4 were computed as follows36,56
| 3 |
where Rct and Rct′ are charge-transfer resistances with and without tricationic amphiphiles, respectively. It should be noted that the effectiveness of inhibition was related to greater stability and hydrophobicity of the external films of inhibitors.42,49 EIS data proved that tricationic aliphatic amphiphiles could inhibit the corrosion of S90 steel, similar to that obtained from polarization measurements.
Table 4. EIS Parameters for S90 Carbon Steel in Oil Well Produced Water with and without Different Concentrations of Aliphatic Tricationic Amphiphiles.
| inhibitors | dose (ppm) | Rs (Ω cm2) | Cf (μs Ω–1 cm–2) | n1 | Rf (Ω cm2) | Cdl (μs Ω–1 cm–2) | n2 | Rct (Ω cm2) | IE (%) |
|---|---|---|---|---|---|---|---|---|---|
| blank | 8.1 ± 0.4 | 23.84 ± 0.9 | 0.86 ± 0.02 | 12.5 ± 0.3 | 2354.7 ± 19 | 0.94 ± 0.06 | 166.4 ± 0.9 | ||
| AEB | 50 | 8.57 ± 0.5 | 16.73 ± 1.1 | 0.88 ± 0.07 | 24.6 ± 0.7 | 1604.8 ± 27 | 0.74 ± 0.04 | 316.27 ± 0.4 | 47.39 |
| 100 | 10.92 ± 0.8 | 12.09 ± 0.7 | 0.81 ± 0.03 | 35.1 ± 1.1 | 1239.2 ± 14 | 0.92 ± 0.07 | 407.34 ± 0.5 | 59.15 | |
| 150 | 12.31 ± 0.3 | 11.21 ± 0.3 | 0.86 ± 0.04 | 44.7 ± 0.6 | 901.3 ± 9 | 0.79 ± 0.02 | 518.17 ± 1.3 | 67.89 | |
| 200 | 13.46 ± 0.9 | 7.85 ± 0.8 | 0.94 ± 0.05 | 57.3 ± 0.9 | 623.7 ± 15 | 0.83 ± 0.05 | 703.21 ± 0.9 | 76.34 | |
| 250 | 11.28 ± 0.7 | 6.29 ± 0.4 | 0.67 ± 0.01 | 68.4 ± 0.5 | 448.5 ± 7 | 0.86 ± 0.03 | 1048.52 ± 0.8 | 84.13 | |
| AEO | 50 | 9.17 ± 0.6 | 15.34 ± 0.7 | 0.85 ± 0.05 | 27.1 ± 0.3 | 1512.3 ± 21 | 0.79 ± 0.07 | 336.31 ± 0.7 | 50.52 |
| 100 | 10.42 ± 0.9 | 12.76 ± 0.3 | 0.89 ± 0.06 | 38.4 ± 0.6 | 1168.6 ± 17 | 0.86 ± 0.02 | 441.8 ± 0.3 | 62.34 | |
| 150 | 11.49 ± 0.5 | 10.46 ± 0.9 | 0.82 ± 0.01 | 47.2 ± 0.4 | 827.2 ± 11 | 0.84 ± 0.06 | 561.86 ± 0.9 | 70.38 | |
| 200 | 8.40 ± 0.7 | 7.25 ± 0.2 | 0.76 ± 0.03 | 60.7 ± 0.7 | 593.4 ± 5 | 0.91 ± 0.03 | 798.46 ± 0.2 | 79.16 | |
| 250 | 9.83 ± 0.3 | 5.46 ± 0.6 | 0.93 ± 0.07 | 71.9 ± 1.2 | 329.1 ± 9 | 0.82 ± 0.01 | 1307.2 ± 0.6 | 87.27 | |
| AED | 50 | 10.53 ± 0.3 | 13.07 ± 0.6 | 0.90 ± 0.04 | 29.3 ± 0.5 | 1429.4 ± 25 | 0.83 ± 0.06 | 367.34 ± 0.3 | 54.70 |
| 100 | 11.81 ± 0.6 | 11.35 ± 0.4 | 0.83 ± 0.01 | 43.5 ± 0.3 | 941.5 ± 12 | 0.76 ± 0.01 | 492.21 ± 0.8 | 66.19 | |
| 150 | 9.34 ± 0.8 | 8.57 ± 0.7 | 0.78 ± 0.03 | 51.6 ± 1.0 | 764.1 ± 16 | 0.92 ± 0.04 | 610.67 ± 0.4 | 72.75 | |
| 200 | 12.70 ± 0.4 | 6.89 ± 0.5 | 0.91 ± 0.07 | 63.2 ± 0.6 | 528.9 ± 7 | 0.88 ± 0.07 | 879.52 ± 0.6 | 81.08 | |
| 250 | 10.62 ± 0.6 | 4.92 ± 0.3 | 0.84 ± 0.05 | 75.7 ± 0.9 | 281.7 ± 4 | 0.90 ± 0.05 | 1723.81 ± 1.1 | 90.35 |
2.2.3. Adsorption Study
The interaction between tricationic surfactant inhibitors and the S90 steel surface in formation water was explained through adsorption isotherms. Tafel and EIS plots were employed to obtain the surface coverage data, which in turn was utilized to adapt numerous adsorption isotherms.57 The polarization and impedance empirical results analysis exhibited a straight line of slope approximating 1, suggesting the Langmuir isotherm model. A slight difference in unity was usually ascribed to the interaction of tricationic surfactants on the S90 steel surface. Langmuir isotherm was described via the following equation58
| 4 |
where Cinh represents the inhibitor concentration and Kads signifies the adsorptive equilibrium constant and has been estimated from the inverse of intercept. The values of Kads could be utilized to estimate the free energy of adsorption (ΔGads°) as follows49
| 5 |
where T represents the absolute temperature, R indicates the gas constant, and (55.5) represents the molar concentration of water in solution. Electrochemical measurements were employed to reinforce the accomplished results to validate the outcomes.59,60 Langmuir plots for the studied aliphatic tricationic amphiphiles are displayed in Figure 7. The regression coefficients (R2), Kads, and ΔGads° values are predestined and summarized in Table 5. It can be seen from Table 5 that the values of R2 were approximately unity, suggesting the Langmuir isotherm.8Kads values of tricationic amphiphiles recorded in Table 5 increased as follows: AED > AEO > AEB. The authors concluded from these achieved results that the increasing alkyl chain length strongly led to an increase of adsorption capability of the studied tricationic aliphatic surfactants. The ΔGads values of tricationic amphiphiles ranged from −33.88 to −36.13 kJ mol–1, so it is assumed that the adsorption process includes both physical and chemical adsorption mechanisms.58,61 The spontaneous adsorption for tricationic aliphatic amphiphiles and the stability of the protective film on the S90 steel surface in oil well produced water reflect the revealed negative values of ΔGads°.62
Figure 7.
Langmuir adsorption plots of (a) AEB, (b) AEO, (c) AED on S90 steel in oil well produced water containing different concentrations at 25 °C.
Table 5. Adsorption Parameters Obtained from Polarization and EIS Results for the Studied Aliphatic Tricationic Surfactants in Oil Well Produced Water at 25 °C.
| technique | inhibitor | regression coefficient (R2) | slope | intercept | Kads (L mol–1) | –ΔGads° (kJ mol–1) |
|---|---|---|---|---|---|---|
| PP | AEB | 0.9851 | 0.8941 | 0.0626 | 15,974.44 | –33.93 |
| AEO | 0.9777 | 0.8934 | 0.0389 | 25,706.94 | –35.11 | |
| AED | 0.9802 | 0.9003 | 0.0258 | 38,759.69 | –36.13 | |
| EIS | AEB | 0.9857 | 0.9525 | 0.0639 | 15,649.45 | –33.88 |
| AEO | 0.9848 | 0.9345 | 0.0395 | 25,316.46 | –35.07 | |
| AED | 0.9821 | 0.9324 | 0.0263 | 38,022.81 | –36.08 |
2.2.4. AFM Studies
The surface morphology of S90 steel samples has been performed via AFM. Figure 8a–c describes the three-dimensional (3D) AFM micrographs of the polished S90 steel surface and the S90 steel surface in oil well produced water without and with 250 ppm of AED. In Figure 8a, it can be seen that the S90 steel surface before immersion was a very smooth and homogeneous surface. Figure 8b describes the S90 steel surface in formation water solution, which displays a roughly deteriorated and roughed surface. Figure 8c represents the S90 steel surface in oil well produced water containing 250 ppm of the AED inhibitor in which the surface has become smoother compared to Figure 8b. From the AFM analysis, the root-mean-square roughness (Rq), average surface roughness (Ra), and a maximum peak-to-valley (P–V) height have been summarized in Table 6.
Figure 8.
3D AFM images for S90 type carbon steel surface in formation water: (a) before immersion, (b) in formation water (blank), and (c) in formation water containing 250 ppm of AED.
Table 6. AFM Results Obtained for S90 Steel before Immersion and after Immersion in Oil Well Produced Water in the Absence and Presence of the 250 ppm AED Surfactant.
| samples | average roughness, Ra (nm) | root-mean-square, Rq (nm) | maximum-peak-to valley height, (P–V) (nm) |
|---|---|---|---|
| free | 59.109 | 76.89 | 656.72 |
| formation water | 300.88 | 370.71 | 2471.5 |
| 250 ppm AED | 101.84 | 126.31 | 919.39 |
The Ra of S90 steel surface before immersion and S90 steel in corrosive solution was about 59.109 and 300.88 nm, respectively. The average surface roughness has been significantly reduced to 101.84 nm by adding 250 ppm AED to the corrosive solution. This signalizes that the AED surfactant adsorbed on the S90 steel surface enhanced the surface morphology, which becomes uniform via decreasing corrosion attack.49,63 The Rq values of the AED inhibitor declined from 370.71 to 126.31 nm in the corrosive solution. In addition, the (P–V) values declined from 2471.5 to 919.39 nm in the corrosive solution.64 These factors proved that the S90 steel surface has become smoother and therefore less damaged because of the prime inhibitive action of AED.
2.2.5. Computational Study
The efficient use of computational science in corrosion inhibition studies depends mainly on the continuing progress of hardware and software developments. The corrosion inhibition impacts of these novel synthesized organic inhibitors can be examined by such quantum chemical calculations. These calculations represent a valuable way of ensuring a particular relationship among different molecular structures and their inhibition effectiveness.25,65 The optimized molecular structure, with the sketched frontier molecular orbital regional electron density distribution for the studied inhibitors, is presented in Figure 9. In addition, several theoretical parameters that were obtained from such calculations, comprising the energy of the highest occupied molecular orbital (EHOMO), energy of the LUMO (ELUMO), energy gap (ΔEL–H), global hardness (η), softness (σ), ionization potential (I), electron affinity (A), dipole moment (μ), electronegativity (X), the chemical potential (π), and the number of transferred electrons from the surfactant to the S90 steel surface (ΔN), are recorded in Table 7.
Figure 9.
Optimized structures and HOMO and LUMO of aliphatic tricationic surfactant molecules utilized by the DFT method.
Table 7. Quantum Chemical Parameters for the Tricationic Surfactant Inhibitors.
| inhibitor | EHOMO (eV) | ELUMO (eV) | ΔE (eV) | A (eV) | I (eV) | μ (debye) | X (eV) | η (eV) | σ (eV)−1 | ΔN |
|---|---|---|---|---|---|---|---|---|---|---|
| AEB | –4.349 | –0.835 | 3.514 | 0.835 | 4.349 | 3.162 | 2.592 | 1.757 | 0.569 | 1.2544 |
| AEO | –3.732 | –1.274 | 2.458 | 1.274 | 3.732 | 5.419 | 2.503 | 1.229 | 0.814 | 1.8295 |
| AED | –3.057 | –1.681 | 1.376 | 1.681 | 3.057 | 6.703 | 2.369 | 0.688 | 1.453 | 3.3656 |
Based on the density functional theory (DFT), some theoretical parameters were responsible for inhibiting the corrosion by inhibitors and directly affected the interaction between the molecules of the inhibitor and the metallic surface of the studied S90 steel, that is, EHOMO was associated to the capability of inhibitors to provide electrons to the S90 carbon steel surface to bond with it. The highest EHOMO values of the AED inhibitor (−3.057 eV) demonstrate that the AED inhibitor has the highest ability for electron donation to the unoccupied d orbital of the Fe surface with smaller energy, which improves the adsorption procedure than that of AEB and AEO inhibitors. On the other hand, ELUMO displays the inhibitor’s ability to receive an electron from the S90 d-orbital. The AED inhibitor with the lowest ELUMO values (−1.681 eV) has the capability to receive electrons easily, with an outstanding IE.66 For that reason, the predictable trend for inhibition effectiveness was AED > AEO > AEB, which was consistent with these impedance and Tafel results.
Besides, the effectiveness of the prepared inhibitor molecules was the energy gap (ΔE = ELUMO – EHOMO), and it is proved by another significant factor. The AED inhibitor with the lowest ΔE value (1.376 eV) scored the highest corrosion IE because of the decrease in the required energy to get rid of an electron from the occupied orbital. Energy gap has often been connected with softness and hardness, that is to say hard particles have a great energy gap value and are therefore less responsive to those associating with the carbon steel surface than soft particles. This is due to the fact that hard molecules might have difficulty in donating the electrons to the S90 carbon steel surface, where it acts as an acceptor.67 The AED inhibitor with the lowest energy gap value suggests a soft–soft interaction between this surfactant and the S90 steel surface, which leads to high corrosion IE.
The polarity and the distribution of electrons in the molecule are strongly described by the dipole moment (μ). The dipole moment (μ) arises from the unequal distribution of charges above particle’s atoms. The AED inhibitor with the highest μ values (6.703 debye) considers the highest corrosion inhibitor owing to its strong interaction with the surface of carbon steel via dipole–dipole interaction and supports the accumulation of AED molecules on the S90 steel surface.68 The number of transferable electrons (ΔN) from the adsorbed molecule to the S90 steel surface can be regarded as the additional technique that is used to link quantum chemical parameters to the corrosion IE. This can be computed as follows58,69
| 6 |
where η and X symbols donate the absolute hardness and absolute electronegativity for both the inhibitor and the S90 steel, respectively. These two parameters were linked to ionization potential (I = −EHOMO) and electron affinity (A = −ELUMO) as presented in the following equations68−71
| 7 |
| 8 |
where π denotes the chemical optional and denotes the softness. The theoretical values for XFe and ηFe were 7.0 and 0.0 eV mol–1, respectively. Based on Lukovits et al.,70 if the ΔN value is less than (3.6), the effectiveness of corrosion inhibition rises straightly with ΔN, indicating that the AED inhibitor was the most absorbed molecule effectively on the surface of carbon steel. ΔN values were positive number, so it was anticipated that the inhibitors to be electron donors to the Fe atoms at the carbon steel surface acted as the electron acceptor.58,72 The results in Table 7 revealed that the σ and ΔN values were in the following order (AEB < AEO < AED). In addition, the inhibitor AED with the lowest electronegativity value has the highest IE, as shown in Table 7, because of the high availability of electrons that such a good corrosion inhibitor offers to the Fe atoms at the S90 steel surface. In conclusion and as shown by the previously conducted quantum chemical parameter data, the corrosion IE increases in the following order (AEB < AEO < AED). The data in Table 7, indicating that these inhibitors (as electron donor) have a robust capability to be coordinated to the d-orbitals of Fe atoms at the S90 steel surface (as electron acceptor) via donation and acceptance of electron pairs, were compatible with the electrochemical testing data.
2.2.6. Mechanism of Corrosion Inhibition of the Aliphatic Tricationic Surfactant
Aliphatic tricationic surfactants exhibited excellent corrosion inhibition properties owing to the adsorption of these amphiphiles on the S90 steel surface over either physisorption or chemisorption types, as illustrated in Figure 10. Triquaternary nitrogen N+, alkyl chains, and oxygen atoms are the active centers for the adsorption mechanism in the synthesized inhibitors (AEB, AEO, and AED). Consequent adsorption of Cl– ions on the metal surface led to the acquisition of negative charges as chloride ion dosage of approximately 90,000 ppm in the formation water solution.49,73 The physisorption involves the first Coulombic interaction among chloride ions, and then electrostatic interaction happened among the positively charged triquaternary N atoms on the inhibitors and the negatively charged chloride ions on the S90 carbon steel surface to afford a barrier film, inhibiting the S90 steel surface from reacting with corrosive species (physical adsorption). Furthermore, the existence of chloride anions might improve the possibility of the adsorption mechanism on the S90 steel surface. The chemisorption74 involves coordination bond formed between the unoccupied 3d-orbitals of Fe on the S90 carbon steel surface and the nonbonding electrons located on heteroatoms (O, N) of inhibitors to form a coordinate type bond, as shown in Figure 10. Moreover, the presence of tri-alkyl chain (hydrophobic chain length) could improve the adsorption process on the S90 steel surface, maintaining the corrosive ions (oxygen and chloride) outside the surface of S90 carbon steel.75,76 The previous discussion was consistent with the experimental values of ΔGads°.
Figure 10.
Corrosion inhibition mechanism scheme for the synthesized aliphatic tricationic surfactant inhibitors (AEB) in oil well produced water.
3. Conclusions
The main conclusions of this work may be stated in the following points:
-
1.
Aliphatic tricationic surfactants have been prepared with a two-step reaction, and the obtained materials were investigated through microelemental analysis, FTIR, and 1H NMR.
-
2.
The prepared aliphatic tricationic surfactants have excellent anticorrosion performance for S90 steel in oil well produced water, and the effectiveness of corrosion inhibition increases as follows: AEB < AEO < AED.
-
3.
Tafel curves indicated that the prepared aliphatic tricationic surfactants impede the cathodic and anodic reactions and therefore classified as mixed-type inhibitors. EIS measurements revealed that the increase in the inhibitors dosage is accompanied by an increase in both Rct and IE % values and decline in Cdl value.
-
4.
The adsorption of the prepared tricationic aliphatic amphiphiles on the S90 steel surface followed the Langmuir isotherm and blocking of its active sites. The negative sign of ΔGads° revealed that the adsorption of tricationic surfactants was a spontaneous process. ΔGads values reveal that the adsorption involves both chemisorption and physisorption mechanisms.
-
5.
AFM images showed that a good protective layer was created on the S90 carbon steel surface via the adsorption of tricationic surfactants. This shielded layer separates the S90 steel surface from corrosion attacks.
-
6.
Theoretical calculations indicated that the inhibition effectiveness of surfactants was increased with a growth in EHOMO and a reduction in the energy band gap (ΔE). The data obtained from theoretical calculations were well compatible with electrochemical empirical findings.
-
7.
The adsorption of these tricationic surfactants on the S90 steel surface could happen directly by coordination bonds among the lone pair of electrons located on heteroatoms (O, N) of these inhibitors (as electron donor) and unoccupied d-orbitals of Fe on the carbon steel surface (as electron acceptor) or by electrostatic interaction between the positive charge in the triquaternary N atoms on inhibitors and the negative charge chloride ions on the S90 carbon steel surface.
4. Experimental Techniques and Materials
4.1. Instrumentation and Materials
Materials used in this work without further purification are as follows: citric acid, 2-(2-chloroethoxy)ethanol, tridodecyl amine, trioctyl amine, tributyl amine, xylene, acetone, and p-toluenesulfonic acid were obtained from Alfa Aesar, Germany.
Characterizations have been performed with a Mattson 5000 FTIR spectrophotometer (KBr discs), Bruker AVANCE III 500 MHz NMR, Vario elementar analyzer, Autolab PGSTAT 30, and AFM (Nanosurf Flex AFM).
4.2. Synthesis of Aliphatic Tricationic Surfactants
The synthetic routes of the novel tri-quaternary ammonium tricationic surfactants based on citric acid and 2-(2-chloroethoxy)ethanol consist of two steps, as illustrated in Scheme 1.
-
(i)
Esterification reaction77 of 2-hydroxypropane-1,2,3-tricarboxylic acid (0.025 mol) and 2-(2-chloroethoxy)ethanol (0.075 mol) in the presence of xylene and p-toluenesulfonic acid at 140 °C to produce AE. The reactants were continuously refluxed with stirring up to an azeotropic amount of H2O (1.35 mL), which was collected in the Dean–Stark trap. The reaction was continued until no further water was generated. After the solvent was removed under reduced pressure, trichloro alkyl triester (AE) was obtained as a yellowish precipitate.
-
(ii)
Quaternization reaction78 for tri-chloro alkyl triester (AE) obtained from the former stage (5.1177 g, 10 mmol) with different tertiary nitrogen atoms, namely, tri-n-butyl amine (5.5608 g, 30 mmol), tri-n-octyl amine (10.6104 g, 30 mmol), and tri-n-dodecyl amine (15.66 g, 30 mmol) individually in acetone at 70 °C for 30 h to obtain one of the following compounds: N1,N1,N1,N17,N17,N17-hexabutyl-9-hydroxy-7,11-dioxo-9-((2-(2-(tributyl ammonio)ethoxy)ethoxy)carbonyl)-3,6,12,15-tetraoxaheptadecane-1,17-diaminium chloride [compound AEB], N1,N1,N1,N17,N17,N17-hexaoctyl-9-hydroxy-7,11-dioxo-9-((2-(2-(trioctylammonio)ethoxy)ethoxy)carbonyl)-3,6,12,15-tetraoxaheptadecane-1,17-diaminium chloride [compound AEO]. and N1,N1,N1,N17,N17,N17-hexadodectyl-9-hydroxy-7,11-dioxo-9-((2-(2-(tridodectylammonio)ethoxy)ethoxy)carbonyl)-3,6,12,15-tetraoxaheptadecane-1,17-diaminium chloride [compound AED].
The reaction mixture was left to cool. The resulting compounds were washed with diethyl ether, recrystallized with acetone, and dried under reduced pressure to produce the tri-quaternary ammonium tricationic surfactants (Scheme 1). All tri-quaternary ammonium tricationic surfactants were soluble in water.
4.3. Corrosion Measurements
Corrosion studies were performed via a S90 steel electrode with the following chemical composition (wt %) of C, 0.0777; Mn, 0.435; Si, 0.0814; S, 0.0105; Al, 0.0052; Nb, 0.0023; Cu, 0.001; Cr, 0.0088; Ni, 0.0099; N, 0.0063; Ti, 0.0013; and remaining Fe in the tested produced water from oil wells. The physical characteristics and chemical composition of this formation water from oil source wells in Egypt are presented in Table 8.
Table 8. Characteristics of the Oil Well Formation Water.
| physical properties | |
|---|---|
| property | values |
| density | 1.10853 g/cm3 |
| turbidity | 391 FAU |
| specific gravity | 1.10959 |
| pH | 6.0 |
| salinity as NaCl | 148,500 mg/L |
| conductivity | 33.28 × 10–2 μs/cm |
| iron | 10.0 mg/L |
| total hardness | 50,600 mg/L |
| chemical properties | |
|---|---|
| ionic species | values (mg/L) |
| Na+ | 35,600 |
| K+ | 11,250 |
| Mg2+ | 5000 |
| Ca2+ | 12,000 |
| Ba2+ | 0.2 |
| F– | 63.27 |
| Cl– | 90,000 |
| Br– | 125.21 |
| HCO3– | 300 |
| SO42– | 700 |
| TDS | 120,000 |
Electrochemical tests (PP and EIS tests) were carried out via an Autolab PGSTAT 30. Electrochemical experiments were conducted via a three-electrode corrosion cell made from PERSPEX glass.79 The carbon steel specimen with 0.4 cm2 was utilized as the working electrode, a platinum wire as the counter electrode, and Ag/AgCl in 3 M KCl as a reference electrode. Carbon steel electrode had been polished by numerous emery papers (grade 400–2500), washed with ethanol and distilled water, and then dried before testing. EIS tests were performed by AC signals (10 mV) peak-to-peak at the open-circuit potential (OCP) in a frequency range from 100 kHz to 20 mHz. Polarization experiments were conducted after the EIS measurement at a scan rate of 1 mV s–1 and sweeping the electrode potential ±350 mV versus OCP. Each test was run multiple times to guarantee data accuracy.4,58
4.4. Surface Analysis
Morphological investigation was performed utilizing Nanosurf Flex AFM for S90 steel samples, with dimensions 10 × 10 × 2 mm. The S90 steel surface was pretreated in the identical way of the working electrode and immersed for 65 days in the profound oilfield produced water solution without and with 250 ppm of AED at 24 ± 1 °C. Specimens were collected and then dried for use with AFM. The specific scan area in the photographs was 10 × 10 μm. In this analysis, the 3D morphology of surface films gave different film roughness parameters, that is, square mean value.49,80
4.5. Computational Studies
The molecular structure optimization of the studied tricationic amphiphiles was executed using the semiempirical PM6 computational method by Gaussian 09 software package81 to estimate the maximum and the minimum values of the computed potential energy curve, with no presence of imaginary frequencies. The molecules were first sketched using a built-in graphical user interface of GaussView 5.0. Frontier molecular orbital (HOMO) and (LUMO) energies were calculated in order to predict the abilities of the molecule to donate/accept electrons. Other quantum chemical descriptors were evaluated in order to assess the adsorption abilities of the molecules and to specify the adsorption centers and were correlated to the EHOMO and ELUMO of the molecule according to Koopman’s theorem,82 for instance, the ionization potential (I), the electron affinity (A), dipole moment (μ), chemical potential (π), electronegativity (X), hardness (η), and softness (σ).
Acknowledgments
The authors express their appreciation to the Egyptian Petroleum Research Institute (EPRI) for the financial support for this study.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c03432.
FTIR and 1H NMR spectra of the synthesized trichloro alkyl triester (AE); polarization curves for S90 steel in oil well produced water without and with different concentrations of AEB and AEO at 298 K; Nyquist graphs for S90 steel in formation water without and with different concentrations of AEB and AEO at 298 K; Bode graphs for S90 steel in formation water without and with different concentrations of AEB and AEO at 298 K; and EOCP versus time curves for S90 steel in formation water without and with different concentrations of the prepared trimeric cationic surfactants at 25 °C (PDF)
Author Contributions
∥ M.A.E.-M. and M.M.S. contributed equally to the article.
The authors declare no competing financial interest.
Supplementary Material
References
- Roberge P. R.Handbook of Corrosion Engineering; McGraw-Hill: New York, 2000. [Google Scholar]
- Bedair M. A.; El-Sabbah M. M. B.; Fouda A. S.; Elaryian H. M. Synthesis, electrochemical and quantum chemical studies of some prepared surfactants based on azodye and Schiff base as corrosion inhibitors for steel in acid medium. Corros. Sci. 2017, 128, 54–72. 10.1016/j.corsci.2017.09.016. [DOI] [Google Scholar]
- Kermani M. B.; Smith L. M.. CO2 Corrosion Control in Oil and Gas Production: Design Considerations; The Institute of Materials, European Federation of Corrosion Publications: London, 1997. [Google Scholar]
- Migahed M. A.; Shaban M. M.; Fadda A. A.; Ali T. A.; Negm N. A. Synthesis of some quaternary ammonium gemini surfactants and evaluation of their performance as corrosion inhibitors for carbon steel in oil well formation water containing sulfide ions. RSC Adv. 2015, 5, 104480–104492. 10.1039/c5ra15112k. [DOI] [Google Scholar]
- Popoola L. T.; Grema A. S.; Latinwo G. K.; Gutti B.; Balogun A. S. Corrosion problems during oil and gas production and its mitigation. Int. J. Ind. Chem. 2013, 4, 35. 10.1186/2228-5547-4-35. [DOI] [Google Scholar]
- Zana R. Dimeric and oligomeric surfactants. Behavior at interfaces and in aqueous solution: a review. Adv. Colloid Interface Sci. 2002, 97, 205–253. 10.1016/s0001-8686(01)00069-0. [DOI] [PubMed] [Google Scholar]
- Danino D.; Talmon Y.; Levy H.; Beinert G.; Zana R. Branched threadlike micelles in an aqueous solution of a trimeric surfactant. Science 1995, 269, 1420–1421. 10.1126/science.269.5229.1420. [DOI] [PubMed] [Google Scholar]
- Negm N. A.; Migahed M. A.; Farag R. K.; Fadda A. A.; Awad M. K.; Shaban M. M. High performance corrosion inhibition of novel tricationic surfactants on carbon steel in formation water: Electrochemical and computational evaluations. J. Mol. Liq. 2018, 262, 363–375. 10.1016/j.molliq.2018.04.092. [DOI] [Google Scholar]
- Zhu Y.; Free M. L.; Woollam R.; Durnie W. A review of surfactants as corrosion inhibitors and associated modeling. Prog. Mater. Sci. 2017, 90, 159–223. 10.1016/j.pmatsci.2017.07.006. [DOI] [Google Scholar]
- Abdallah M.; Al-Tass H. M.; AL Jahdaly B. A.; Fouda A. S. Inhibition properties and adsorption behavior of 5-arylazothiazole derivatives on 1018 carbon steel in 0.5 M H2SO4 solution. J. Mol. Liq. 2016, 216, 590–597. 10.1016/j.molliq.2016.01.077. [DOI] [Google Scholar]
- Kaczerewska O.; Leiva-Garcia R.; Akid R.; Brycki B.; Kowalczyk I.; Pospieszny T. Effectiveness of O-bridged cationic gemini surfactants as corrosion inhibitors for stainless steel in 3 M HCl: Experimental and theoretical studies. J. Mol. Liq. 2018, 249, 1113–1124. 10.1016/j.molliq.2017.11.142. [DOI] [Google Scholar]
- Touir R.; Belakhmima R. A.; Touhami M. E.; Lakhrissi L.; El Fayed M.; Lakhrissi B.; Essassi E. M. Comparative inhibition study of mild steel corrosion in hydrochloric acid by benzimidazole derivatives. J. Mater. Environ. Sci. 2013, 4, 921–930. [Google Scholar]
- El Defrawy A. M.; Abdallah M.; Al-Fahemi J. H. Electrochemical and theoretical investigation for some pyrazolone derivatives as inhibitors for the corrosion of C-steel in 0.5 M hydrochloric acid. J. Mol. Liq. 2019, 288, 110994. 10.1016/j.molliq.2019.110994. [DOI] [Google Scholar]
- Migahed M. A.; Al-Sabagh A. M. Beneficial role of surfactants as corrosion inhibitors in petroleum industry: a review article,. Chem. Eng. Commun. 2009, 196, 1054–1075. 10.1080/00986440902897095. [DOI] [Google Scholar]
- El-Tabei A. S.; Hegazy M. A.; Bedair A. H.; Sadeq M. A. Synthesis and inhibition effect of a novel tri-cationic surfactant on carbon steel corrosion in 0.5 M H2SO4 solution. J. Surfactants Deterg. 2014, 17, 341–352. 10.1007/s11743-013-1524-7. [DOI] [Google Scholar]
- Patoczka J.; Pulliam G. W. Biodegradation and secondary effluent toxicity of ethoxylated surfactants. Water Res. 1990, 24, 965–972. 10.1016/0043-1354(90)90118-p. [DOI] [Google Scholar]
- Fatma N.; Panda M.; Ansari W. H. Environment-friendly ester bonded gemini surfactant: mixed micellization of 14-E2-14 with ionic and nonionic conventional surfactants. J. Mol. Liq. 2015, 211, 247–255. 10.1016/j.molliq.2015.06.064. [DOI] [Google Scholar]
- Al-Mohammed N. N.; Hussen R. S. D.; Alias Y.; Abdullah Z. Tris-imidazolium and benzimidazolium ionic liquids: a new class of biodegradable surfactants. RSC Adv. 2015, 5, 2869–2881. 10.1039/c4ra14027c. [DOI] [Google Scholar]
- Tatsumi T.; Zhang W.; Kida T.; Nakatsuji Y.; Ono D.; Takeda T.; Ikeda I. Novel hydrolyzable and biodegradable cationic gemini surfactants: bis (ester-ammonium) dichloride having a butenylene or a butynylene spacer. J. Surfactants Deterg. 2001, 4, 279–285. 10.1007/s11743-001-0180-6. [DOI] [Google Scholar]
- Al-Mohammed N. N.; Hussen R. S. D.; Ali T. H.; Alias Y.; Abdullah Z. Tetrakis-imidazolium and benzimidazolium ionic liquids: a new class of biodegradable surfactants. RSC Adv. 2015, 5, 21865–21876. 10.1039/c4ra16811a. [DOI] [Google Scholar]
- Tehrani-Bagha A. R.; Oskarsson H.; van Ginkel C. G.; Holmberg K. Cationic ester-containing gemini surfactants: chemical hydrolysis and biodegradation. J. Colloid Interface Sci. 2007, 312, 444–452. 10.1016/j.jcis.2007.03.044. [DOI] [PubMed] [Google Scholar]
- Garcia M. T.; Ribosa I.; Kowalczyk I.; Pakiet M.; Brycki B. Biodegradability and aquatic toxicity of new cleavable betainate cationic oligomeric surfactants. J. Hazard. Mater. 2019, 371, 108–114. 10.1016/j.jhazmat.2019.03.005. [DOI] [PubMed] [Google Scholar]
- Tehrani-Bagha A. R.; Holmberg K. Cationic ester-containing gemini surfactants: physical–chemical properties. Langmuir 2010, 26, 9276–9282. 10.1021/la1001336. [DOI] [PubMed] [Google Scholar]
- Madkour L. H.; Elshamy I. H. Experimental and computational studies on the inhibition performances of benzimidazole and its derivatives for the corrosion of copper in nitric acid. Int. J. Ind. Chem. 2016, 7, 195–221. 10.1007/s40090-015-0070-8. [DOI] [Google Scholar]
- Gece G. The use of quantum chemical methods in corrosion inhibitor studies. Corros. Sci. 2008, 50, 2981–2992. 10.1016/j.corsci.2008.08.043. [DOI] [Google Scholar]
- Han T.; Guo J.; Zhao Q.; Wu Y.; Zhang Y. Enhanced corrosion inhibition of carbon steel by pyridyl gemini surfactants with different alkyl chains. Mater. Chem. Phys. 2020, 240, 122156. 10.1016/j.matchemphys.2019.122156. [DOI] [Google Scholar]
- Zhou T.; Yuan J.; Zhang Z.; Xin X.; Xu G. The comparison of imidazolium Gemini surfactant [C14-4-C14im] Br2 and its corresponding monomer as corrosion inhibitors for A3 carbon steel in hydrochloric acid solutions: Experimental and quantum chemical studies. Colloids Surf., A 2019, 575, 57–65. 10.1016/j.colsurfa.2019.05.004. [DOI] [Google Scholar]
- Abdallah M.; Gad E. A. M.; Sobhi M.; Al-Fahemi J. H.; Alfakeer M. M. Performance of tramadol drug as a safe inhibitor for aluminum corrosion in 1.0 M HCl solution and understanding mechanism of inhibition using DFT. Egypt. J. Pet. 2019, 28, 173–181. 10.1016/j.ejpe.2019.02.003. [DOI] [Google Scholar]
- Banno T.; Toshima K.; Kawada K.; Matsumura S. Synthesis and properties of gemini-type cationic surfactants containing carbonate linkages in the linker moiety directed toward green and sustainable chemistry. J. Surfactants Deterg. 2009, 12, 249–259. 10.1007/s11743-009-1119-5. [DOI] [Google Scholar]
- Khaled K. F.; Abdel-Shafi N. S.; Al-Mubarak N. A.; Alonazi M. S. L-arginine as corrosion and scale inhibitor of steel in synthetic reservoir water. Int. J. Electrochem. Sci. 2016, 11, 2433–2446. [Google Scholar]
- Negm N. A.; Al Sabagh A. M.; Migahed M. A.; Bary H. M. A.; El Din H. M. Effectiveness of some diquaternary ammonium surfactants as corrosion inhibitors for carbon steel in 0.5 M HCl solution. Corros. Sci. 2010, 52, 2122–2132. 10.1016/j.corsci.2010.02.044. [DOI] [Google Scholar]
- Migahed M. A.; Rashwan S. M.; Kamel M. M.; Habib R. E. Synthesis, characterization of polyaspartic acid-glycine adduct and evaluation of their performance as scale and corrosion inhibitor in desalination water plants. J. Mol. Liq. 2016, 224, 849–858. 10.1016/j.molliq.2016.10.091. [DOI] [Google Scholar]
- 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. 10.1016/j.molliq.2019.111480. [DOI] [Google Scholar]
- Fu L.; Lv J.; Zhou L.; Li Z.; Tang M.; Li J. Study on corrosion and scale inhibition mechanism of polyaspartic acid grafted β-cyclodextrin. Mater. Lett. 2020, 264, 127276. 10.1016/j.matlet.2019.127276. [DOI] [Google Scholar]
- Bayol E.; Gürten T.; Gürten A. A.; Erbil M. Interactions of some Schiff base compounds with mild steel surface in hydrochloric acid solution. Mater. Chem. Phys. 2008, 112, 624–630. 10.1016/j.matchemphys.2008.06.012. [DOI] [Google Scholar]
- Hussien B. M.; Migahed M. A.; Shaban M. M.; Negm N. A.; Moawad Z.. Corrosion Control of X-60 Type Carbon Steel in Petroleum Formation Water Under High Pressure of CO2 at High Temperature. In Offshore Mediterranean Conference and Exhibition, Offshore Mediterranean Conference, 2017; OMC-2017-595.
- El-Lateef H. M. A.; Soliman K. A.; Tantawy A. H. Novel synthesized Schiff base-based cationic gemini surfactants: electrochemical investigation, theoretical modeling and applicability as biodegradable inhibitors for mild steel against acidic corrosion. J. Mol. Liq. 2017, 232, 478–498. 10.1016/j.molliq.2017.02.105. [DOI] [Google Scholar]
- Fiori-Bimbi M. V.; Alvarez P. E.; Vaca H.; Gervasi C. A. Corrosion inhibition of mild steel in HCL solution by pectin. Corros. Sci. 2015, 92, 192–199. 10.1016/j.corsci.2014.12.002. [DOI] [Google Scholar]
- Wang H.; Gao M.; Guo Y.; Yang Y.; Hu R. A natural extract of tobacco rob as scale and corrosion inhibitor in artificial seawater. Desalination 2016, 398, 198–207. 10.1016/j.desal.2016.07.035. [DOI] [Google Scholar]
- Luo X.; Pan X.; Yuan S.; Du S.; Zhang C.; Liu Y. Corrosion inhibition of mild steel in simulated seawater solution by a green eco-friendly mixture of glucomannan (GL) and bisquaternary ammonium salt (BQAS). Corros. Sci. 2017, 125, 139–151. 10.1016/j.corsci.2017.06.013. [DOI] [Google Scholar]
- 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]
- Rodríguez-Gómez F. J.; Valdelamar M. P.; Vazquez A. E.; Perez P. D. V.; Mata R.; Miralrio A.; Castro M. Mycophenolic acid as a corrosion inhibitor of carbon steel in 3% wt. NaCl solution. An experimental and theoretical study. J. Mol. Struct. 2019, 1183, 168–181. 10.1016/j.molstruc.2018.12.035. [DOI] [Google Scholar]
- Hussein B. M.; Migahed M. A.; Shaban M. M.; Negm N. A.; Moawad Z.. Synthesis and Inhibition Performance of Diquaternary Ammonium Gemini Surfactants on Carbon Steel Pipelines Corrosion in Gas Field. In Offshore Mediterranean Conference & Exhibition, Offshore Mediterranean Conference, 2017; OMC-2017-731.
- Zhang G. A.; Cheng Y. F. Corrosion of X65 steel in CO2-saturated oilfield formation water in the absence and presence of acetic acid. Corros. Sci. 2009, 51, 1589–1595. 10.1016/j.corsci.2009.04.004. [DOI] [Google Scholar]
- Eliyan F. F.; Mahdi E.-S.; Alfantazi A. Electrochemical evaluation of the corrosion behaviour of API-X100 pipeline steel in aerated bicarbonate solutions. Corros. Sci. 2012, 58, 181–191. 10.1016/j.corsci.2012.01.015. [DOI] [Google Scholar]
- Jović V. D.; Jović B. M. EIS and differential capacitance measurements onto single crystal faces in different solutions: Part I: Ag (111) in 0.01 M NaCl. J. Electroanal. Chem. 2003, 541, 1–11. 10.1016/S0022-0728(02)01309-8. [DOI] [Google Scholar]
- Ashassi-Sorkhabi H.; Ghalebsaz-Jeddi N.; Hashemzadeh F.; Jahani H. Corrosion inhibition of carbon steel in hydrochloric acid by some polyethylene glycols. Electrochim. Acta 2006, 51, 3848–3854. 10.1016/j.electacta.2005.11.002. [DOI] [Google Scholar]
- Gonçalves R. S.; Azambuja D. S.; Lucho A. M. S. Electrochemical studies of propargyl alcohol as corrosion inhibitor for nickel, copper, and copper/nickel (55/45) alloy. Corros. Sci. 2002, 44, 467–479. 10.1016/s0010-938x(01)00069-5. [DOI] [Google Scholar]
- Shaban M. M.; Eid A. M.; Farag R. K.; Negm N. A.; Fadda A. A.; Migahed M. A. Novel trimeric cationic pyrdinium surfactants as bi-functional corrosion inhibitors and antiscalants for API 5L X70 carbon steel against oilfield formation water. J. Mol. Liq. 2020, 305, 112817. 10.1016/j.molliq.2020.112817. [DOI] [Google Scholar]
- Liu G.; Xue M.; Yang H. Polyether copolymer as an environmentally friendly scale and corrosion inhibitor in seawater. Desalination 2017, 419, 133–140. 10.1016/j.desal.2017.06.017. [DOI] [Google Scholar]
- Fu J.; Zang H.; Wang Y.; Li S.; Chen T.; Liu X. Experimental and theoretical study on the inhibition performances of quinoxaline and its derivatives for the corrosion of mild steel in hydrochloric acid. Ind. Eng. Chem. Res. 2012, 51, 6377–6386. 10.1021/ie202832e. [DOI] [Google Scholar]
- Sherif E.-S. M.; Park S.-M. Effects of 1, 4-naphthoquinone on aluminum corrosion in 0.50 M sodium chloride solutions. Electrochim. Acta 2006, 51, 1313–1321. 10.1016/j.electacta.2005.06.018. [DOI] [Google Scholar]
- Dkhireche N.; Dahami A.; Rochdi A.; Hmimou J.; Touir R.; Touhami M. E.; El Bakri M.; El Hallaoui A.; Anouar A.; Takenouti H. Corrosion and scale inhibition of low carbon steel in cooling water system by 2-propargyl-5-o-hydroxyphenyltetrazole. J. Ind. Eng. Chem. 2013, 19, 1996–2003. 10.1016/j.jiec.2013.03.012. [DOI] [Google Scholar]
- Hassan H. H.; Abdelghani E.; Amin M. A. Inhibition of mild steel corrosion in hydrochloric acid solution by triazole derivatives: Part I. Polarization and EIS studies. Electrochim. Acta 2007, 52, 6359–6366. 10.1016/j.electacta.2007.04.046. [DOI] [Google Scholar]
- Macdonald J. R.; Johanson W. B. In Theory in Impedance Spectroscopy; Macdonald J. R., Ed.; John Wiley & Sons: New York, 1987. [Google Scholar]
- Fouda A. S.; Mostafa H. A.; Heakal F.; Elewady G. Y. Synergistic influence of iodide ions on the inhibition of corrosion of C-steel in sulphuric acid by some aliphatic amines. Corros. Sci. 2005, 47, 1988–2004. 10.1016/j.corsci.2004.09.008. [DOI] [Google Scholar]
- Lebrini M.; Robert F.; Vezin H.; Roos C. Electrochemical and quantum chemical studies of some indole derivatives as corrosion inhibitors for C38 steel in molar hydrochloric acid. Corros. Sci. 2010, 52, 3367–3376. 10.1016/j.corsci.2010.06.009. [DOI] [Google Scholar]
- Migahed M. A.; Zaki E. G.; Shaban M. M. Corrosion control in the tubing steel of oil wells during matrix acidizing operations. RSC Adv. 2016, 6, 71384–71396. 10.1039/c6ra12835a. [DOI] [Google Scholar]
- Li Y.; Zhang S.; Ding Q.; Qin B.; Hu L. Versatile 4, 6-dimethyl-2-mercaptopyrimidine based ionic liquids as high-performance corrosion inhibitors and lubricants. J. Mol. Liq. 2019, 284, 577–585. 10.1016/j.molliq.2019.04.042. [DOI] [Google Scholar]
- Tasić Ž. Z.; Mihajlović M. B. P.; Radovanović M. B.; Simonović A. T.; Antonijević M. M. Cephradine as corrosion inhibitor for copper in 0.9% NaCl solution. J. Mol. Struct. 2018, 1159, 46–54. 10.1016/j.molstruc.2018.01.031. [DOI] [Google Scholar]
- Migahed M. A.; Aly R. O.; Al-Sabagh A. M. Impact of gamma-ray-pre-irradiation on the efficiency of corrosion inhibition of some novel polymeric surfactants. Corros. Sci. 2004, 46, 2503–2516. 10.1016/j.corsci.2004.01.013. [DOI] [Google Scholar]
- Ding Y.-S.; Zha M.; Zhang J.; Wang S.-S. Synthesis, characterization and properties of geminal imidazolium ionic liquids. Colloids Surf., A 2007, 298, 201–205. 10.1016/j.colsurfa.2006.10.063. [DOI] [Google Scholar]
- Chidiebere M. A.; Oguzie E. E.; Liu L.; Li Y.; Wang F. Ascorbic acid as corrosion inhibitor for Q235 mild steel in acidic environments. J. Ind. Eng. Chem. 2015, 26, 182–192. 10.1016/j.jiec.2014.11.029. [DOI] [Google Scholar]
- Fouda A. S.; Ismail M. A.; EL-ewady G. Y.; Abousalem A. S. Evaluation of 4-amidinophenyl-2, 2′-bithiophene and its aza-analogue as novel corrosion inhibitors for CS in acidic media: experimental and theoretical study. J. Mol. Liq. 2017, 240, 372–388. 10.1016/j.molliq.2017.05.089. [DOI] [Google Scholar]
- Bentiss F.; Lagrenée M.; Elmehdi B.; Mernari B.; Traisnel M.; Vezin H. Electrochemical and quantum chemical studies of 3, 5-di (n-tolyl)-4-amino-1, 2, 4-triazole adsorption on mild steel in acidic media. Corrosion 2002, 58, 399–407. 10.5006/1.3277629. [DOI] [Google Scholar]
- Khalil N. N. Quantum chemical approach of corrosion inhibition. Electrochim. Acta 2003, 48, 2635–2640. 10.1016/s0013-4686(03)00307-4. [DOI] [Google Scholar]
- Bedair M. A.; Soliman S. A.; Metwally M. S. Synthesis and characterization of some nonionic surfactants as corrosion inhibitors for steel in 1.0 M HCl (Experimental and computational study). J. Ind. Eng. Chem. 2016, 41, 10–22. 10.1016/j.jiec.2016.07.005. [DOI] [Google Scholar]
- Zaky R. R.; Ibrahim K. M.; El-Nadar H. M. A.; Abo-Zeid S. M. Bivalent transition metal complexes of (E)-3-(2-benzylidenehydrazinyl)-3-oxo-N-(p-tolyl) propanamide: Spectroscopic, computational, biological activity studies. Spectrochim. Acta, Part A 2015, 150, 40–53. 10.1016/j.saa.2015.04.107. [DOI] [PubMed] [Google Scholar]
- Padmanabhan J.; Parthasarathi R.; Subramanian V.; Chattaraj P. K. Electrophilicity-based charge transfer descriptor. J. Phys. Chem. A 2007, 111, 1358–1361. 10.1021/jp0649549. [DOI] [PubMed] [Google Scholar]
- Lukovits I.; Kálmán E.; Zucchi F. Corrosion inhibitors—correlation between electronic structure and efficiency. Corrosion 2001, 57, 3–8. 10.5006/1.3290328. [DOI] [Google Scholar]
- El Ashry E. S. H.; El Nemr A.; Ragab S. Quantitative structure activity relationships of some pyridine derivatives as corrosion inhibitors of steel in acidic medium. J. Mol. Model. 2012, 18, 1173–1188. 10.1007/s00894-011-1148-7. [DOI] [PubMed] [Google Scholar]
- Sastri V. S.; Perumareddi J. R. Molecular orbital theoretical studies of some organic corrosion inhibitors. Corrosion 1997, 53, 617–622. 10.5006/1.3290294. [DOI] [Google Scholar]
- Ahamad I.; Prasad R.; Quraishi M. A. Adsorption and inhibitive properties of some new Mannich bases of Isatin derivatives on corrosion of mild steel in acidic media. Corros. Sci. 2010, 52, 1472–1481. 10.1016/j.corsci.2010.01.015. [DOI] [Google Scholar]
- Wan K.; Feng P.; Hou B.; Li Y. Enhanced corrosion inhibition properties of carboxymethyl hydroxypropyl chitosan for mild steel in 1.0 M HCl solution. RSC Adv. 2016, 6, 77515–77524. 10.1039/c6ra12975g. [DOI] [Google Scholar]
- Negm N. A.; Elkholy Y. M.; Zahran M. K.; Tawfik S. M. Corrosion inhibition efficiency and surface activity of benzothiazol-3-ium cationic Schiff base derivatives in hydrochloric acid. Corros. Sci. 2010, 52, 3523–3536. 10.1016/j.corsci.2010.07.001. [DOI] [Google Scholar]
- El Achouri M.; Infante M. R.; Izquierdo F.; Kertit S.; Gouttaya H. M.; Nciri B. Synthesis of some cationic gemini surfactants and their inhibitive effect on iron corrosion in hydrochloric acid medium. Corros. Sci. 2001, 43, 19–35. 10.1016/s0010-938x(00)00063-9. [DOI] [Google Scholar]
- Zhuang L.-H.; Yu K.-H.; Wang G.-W.; Yao C. Synthesis and properties of novel ester-containing gemini imidazolium surfactants. J. Colloid Interface Sci. 2013, 408, 94–100. 10.1016/j.jcis.2013.07.029. [DOI] [PubMed] [Google Scholar]
- Migahed M. A.; Negm N. A.; Shaban M. M.; Ali T. A.; Fadda A. A. Synthesis, characterization, surface and biological activity of diquaternary cationic surfactants containing ester linkage. J. Surfactants Deterg. 2016, 19, 119–128. 10.1007/s11743-015-1749-8. [DOI] [Google Scholar]
- ASTM . ASTM G3-89. Standard practice for conventions applicable to electrochemical measurements in corrosion testing. Annual Book of ASTM Standards; ASTM International: West Conshohocken, 2006. [Google Scholar]
- Yadav M.; Sinha R. R.; Sarkar T. K.; Bahadur I.; Ebenso E. E. Application of new isonicotinamides as a corrosion inhibitor on mild steel in acidic medium: Electrochemical, SEM, EDX, AFM and DFT investigations. J. Mol. Liq. 2015, 212, 686–698. 10.1016/j.molliq.2015.09.047. [DOI] [Google Scholar]
- Gece G.; Bilgiç S. Quantum chemical study of some cyclic nitrogen compounds as corrosion inhibitors of steel in NaCl media. Corros. Sci. 2009, 51, 1876–1878. 10.1016/j.corsci.2009.04.003. [DOI] [Google Scholar]
- Mishra A.; Verma C.; Lgaz H.; Srivastava V.; Quraishi M. A.; Ebenso E. E. Synthesis, characterization and corrosion inhibition studies of N-phenyl-benzamides on the acidic corrosion of mild steel: Experimental and computational studies. J. Mol. Liq. 2018, 251, 317–332. 10.1016/j.molliq.2017.12.011. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.










