Journal Pre-proof Tetrazole-based organoselenium bi-functionalized corrosion inhibitors during oil well acidizing: Experimental, computational studies, and SRB bioassay A.M. Eid, S. Shaaban, K. Shalabi PII:
S0167-7322(19)33071-5
DOI:
https://doi.org/10.1016/j.molliq.2019.111980
Reference:
MOLLIQ 111980
To appear in:
Journal of Molecular Liquids
Received Date: 30 May 2019 Revised Date:
10 October 2019
Accepted Date: 19 October 2019
Please cite this article as: A.M. Eid, S. Shaaban, K. Shalabi, Tetrazole-based organoselenium bifunctionalized corrosion inhibitors during oil well acidizing: Experimental, computational studies, and SRB bioassay, Journal of Molecular Liquids (2019), doi: https://doi.org/10.1016/j.molliq.2019.111980. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier B.V.
Graphical Abstract Tetrazole-based organoselenium bi-functionalized corrosion inhibitors during oil well acidizing: experimental, computational studies and SRB bioassay
1 2 3
Tetrazole-based organoselenium bi-functionalized corrosion inhibitors during oil well acidizing: Experimental, computational studies, and SRB bioassay
4
A.M. Eid*a, S. Shaabanb,c and K. Shalabi*b a Chemistry Section, School of Distance Education, Universiti Sanis Malaysia, Malaysia *email:
[email protected];
[email protected] b Chemistry Department, Faculty of Science, Mansoura University, Mansoura, Egypt email:
[email protected];
[email protected] *email:
[email protected];
[email protected] c Chemistry Department, College of Science, King Faisal University, Al-Hofuf, Saudi Arabia.
5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50
Abstract Three tetrazole-based organoselenium compounds (TOS1, TOS2, and TOS3) were synthesized via Ugi and sequential Ugi/nucleophilic substitution approaches. The synthesized agents were evaluated against the corrosive effect of 10% HCl solution over J55 steel tubing samples as a simulation of the impact of a well-stimulating acid job. Chemical and electrochemical studies were conducted by means of various test methods, namely, weight loss, potentiodynamic polarization, electrochemical impedance spectroscopy, and electrochemical frequency modulation. Furthermore, surface examinations were also performed using scanning electron microscope and X-ray photoelectron spectroscopy, which confirmed the formation of protective layer of TOS molecules on J55 steel surface.The inhibition efficiencies (%IE) of TOS1 and TOS2 were up to 92.2% and 94.6%, respectively, whereas TOS3 manifested a lower %IE (89.1%). Moreover, the assessment of TOS1, TOS2, and TOS3 inhibitors as a potential biocidal agent was accounted using NACE Standard TM0194-14 (Field Monitoring of Bacterial Growth in Oil and Gas Systems) against well annulus water samples infected with sulfate-reducing bacteria. Interestingly, all of the tetrazole-based organoselenium compounds exhibited a strong biocidal effect toward microbial-induced corrosion. Finally, quantum chemical calculations and Monte Carlo simulations were performed to provide a rationalization for the inhibitory action of the tetrazole-based organoselenium compounds. Keywords: corrosion inhibition, organoselenium compounds, oil well acidizing, EIS, EFM, SEM, XPS, SRB, QM, Monte Carlo simulations. 1. Introduction Downhole corrosion problems are considered a major source of failures and casualties in the oil and gas industry [1,2]. Although J55 steel is the most commonly used metal for fabricating oil and gas well casing and downhole tubing, its low acid corrosion resistance makes it constantly exposed to corrosion impacts [3,4]. Acidizing of crude oil wells is an effective and commonly used stimulation technique for oil production recovery. Acidizing is used to improve the oil production of old or aging wells by pumping acid into the wellbore so as to enhance the permeability of production channels down the formation rocks by dissolving rubbles and sediments blocking their pores to restore the targeted flow and consequently maximize productivity. Acids are also employed to eliminate muds in newly drilled wells before starting up their production. Several acids are used in oil well acidizing treatment, such as hydrochloric acid (HCl), hydrofluoric acid (HF), acetic acid (CH3COOH), formic acid (HCOOH) and sulfamic acid (H2NSO3H) [5]. The most extensively used acid in such acidizing treatment is HCl with a concentration range of 5%–28% w/w [5]. This acidizing treatment, especially when HCl is utilized, exposes the oil well downhole tubing, which is commonly made of lowresistivity J55 steel to a severe corrosive environment. Within this context, a huge number of organic compounds were tested as inhibitors to mitigate J55 steel corrosion. In most cases, the proposed mechanism of action can be explicated by the ability of such molecules to be adsorbed on the metal surface [6–8]. Recently, such technique for corrosion inhibition has been optimized for commercial use in the oil and gas industry [9,10]. 1
51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102
Microorganisms play a key role in the transformation of organic and inorganic compounds, such as ammonium, nitrate, and other organic trace substances [11–15]. These microbial processes are in progress with natural terrestrial and aquatic ecosystems (e.g., nitrogen conversion in the soil), or they may come in different technical systems, such as the anaerobic degradation processes of sulfatereducing bacteria (SRB), which causes microbial-induced corrosion (MIC) [16–20]. The latter process can be clearly seen in oil and gas fields where anaerobic conditions are met in locations, such as well annulus [21–23]. Interestingly, several organic compounds were reported to counter the act of such microbial corrosion [24–26]. Among these compounds, organic selenides possess strong corrosion inhibition properties [27] in addition to their known antimicrobial activity [28–31]. This present work aims to explore the use of tetrazole-based organoselenium compounds, namely, 4,4'-diselanediylbis(N-(1-(1-(tert-butyl)-1H-tetrazol-5-yl)-2-methylpropyl)aniline) (TOS1), 4,4'diselane diylbis(N-((1-(tert-butyl)-1H-tetrazol-5-yl)(furan-2-yl)methyl)aniline) (TOS2), and 2-((4(((1-(tert-butyl)-1H-tetrazol-5-yl)(p-tolyl)methyl)amino)phenyl)selanyl)-3-methylnaphthalene-1,4dione (TOS3) as corrosion inhibitors for well tubing and casing protection during well stimulation acidizing process. Our aim is further extended to the utilization of organic selenides for the mitigation of the electrochemical and MIC. Chemical weight loss measurements and electrochemical techniques have been performed together with chemical bioassay on real-time field water samples as effective tools for the assessment of the biocidal efficiency. Theoretical quantum calculations and molecular dynamics simulations will be also conducted to confirm the experimental results. This study provides evidence of the optimization of the structural requirements of novel designed and prepared TOS compounds to counter the corrosive action of 10% HCl solution and SRB bacterial population over J55 steel samples, which may cause a synergistic corrosive effect to the oil well tubing during acidization [32]. Driven by the deployment of relevant active moieties within the designed structure, such multifunctional compounds showed high anti-corrosion and antibacterial actions with low dosage compared with those used in previous studies, which discussed only a single anti-corrosion act. Thus, TOS compounds can serve as a good option to mitigate such serious problem [4,5,33,34]. 2. Material and Experimental Techniques 2.1 Materials Specimens of J55 steel were supplied from a previous cut sample of oil well casing material from El Mansoura Petroleum Company’s production fields, where the chemical composition (wt.%) was provided as follows: C 0.34; Si 0.28; Mn 1.31; P 0.03; S 0.008; Cr 0.28; Ni 0.19; Cu 0. 15; Mo 0.005; and Fe balance. Specimens were shaped mechanically into small cubes of dimensions 1 × 1 × 0.5 cm with fused copper wires, which were added for their functionalization as the working electrode in the electrochemical measurements. Referring to the chemical weight loss testing, other specimens were shaped as coupons with dimensions of 2 × 2 × 0.2 cm. These specimens were polished by different grades of abrasive emery paper to ensure that the surface is smooth and were washed with distilled water afterward. Specimens were immersed in acetone to be completely degreased and then dried and kept in the desiccator before starting the experiment. Culture media for the bioassay of the sulfate-reducing bacteria growth were purchased from the Egyptian Petroleum Research Institute (EPRI) in compliance with the NACE Standard TMO194-04 (Standard Test Method Field Monitoring of Bacterial Growth in Oil and Gas Systems). The used media simulated the same NaCl concentration as that of the collected field samples. The tetrazole-based organoselenium compounds (TOS1, TOS2, and TOS3) were synthesized in accordance with our previous work [35]. All related characterization techniques can be found in the Supporting Information (S1). 2.2 Solutions Stock solutions were prepared for each of the tetrazole-based organoselenium compounds under investigation by dissolving each one in ethanol at first and then diluting by bi-distilled water to obtain different concentration ranges (from 10 ppm to 50 ppm). For the corrosive medium, HCl stock 2
103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135
solution of 37% was prepared to be diluted to 10% HCl. The percentage of ethanol to water was maintained to be of constant value all over the prepared solutions in the absence and presence of the compounds so that the effect of ethanol on the inhibition is revoked. 2.3 Chemical method (weight loss) Specimens of dimensions of 2 × 2 × 0.2 cm were cut from the source of the metal alloy under investigation (J55 steel). These specimens were weighed and then inserted into the corrosive media in the absence and presence of the inhibitor under investigation contained in a 100 ml beaker for 6 hrs with a time interval of 60 min. The loss in the weight is calculated after each equally separated time pause (60 min) using Eq. (1): M = − , (1) where M0 and Mt are the sample weight before the immersion and after a time interval of the sample placed in the solution contained in the beaker. The inhibition efficiency (IE%) is calculated by Eq. (2): % IE = θ × 100 = 1 -
× °
100, (2)
where θ denotes the surface coverage and resembles the weight loss in mg cm-2 in the presence of the inhibitor under study, whereas ° marks the weight loss in mg cm-2 without any presence of inhibitor species. This technique was employed in accordance with Standard ISO 9226:2012.
2.4. Electrochemical measurements All electrochemical measurements were performed using a three-compartment electrochemical cell. The three electrodes used in such cell were as follows: Pt sheet electrode with an exposed surface area of 1 cm2 as the counter electrode, saturated calomel electrode as the reference electrode, and a working electrode of a finely cut specimen of J55 steel with an exposed area of 1 cm2 and a fused copper wire for electrical connection. All the three electrodes were immersed in a plain corrosive 10% HCl solution, and the test was repeated afterward with the presence of different concentrations of the inhibitor under evaluation. The variance of the operating potential was scanned from −500 mV to 500 mV vs. open circuit potential (OCP). Then, the Tafel polarization curves were recorded at a scan rate of 1 mV/s. Corrosion current (icorr) and the corresponding corrosion potential (Ecorr) was calculated via extrapolation of the anodic and cathodic parts of the Tafel plots until they cross at a point of intersection. The efficiency of the corrosion inhibition process and surface coverage was computed by Eq. (3) [36]: % IE = θ × 100 1 -
× °
100, (3)
136 137 138 139 140 141 142 143 144
° where and are the corrosion current densities for the free acid solution and the inhibited solution, respectively, and θ represents the surface coverage. In electrochemical impedance spectroscopy (EIS), the system to be examined usually has a sinusoidal potential disturbance small amplitude of 10 mV and the system response, an alternating current, as a function of the frequency of the potential disturbance ranges between 0.2 Hz and 100 kHz. The evaluation of the diagrams is performed by a computer-assisted fitting of the acquired experimental measurements to establish an equivalent circuit. The charge transfer resistance values measured from Nyquist-fitted semi circuit curves are used to calculate the efficiency of the compounds as a potential corrosion inhibitor under evaluation using Eq. (4):
145
% IE = θ × 100 = 1 -
146 147 148 149 150
where R°ct is the charge transfer resistance for plain HCl solutions and Rct is the resistance of the solutions with different concentrations of inhibitor compounds. Electrochemical frequency modulation (EFM) is a newly introduced method used for measuring the rate of corrosion. However, this method outlines the corrosion current icorr and corrosion rate and determines the Tafel parameters (βa and βc). This method is also non-destructive, comparatively fast
°
× 100, (4)
3
151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202
and associated with low polarizations. EFM considers the fact that corrosion is a nonlinear process. Therefore, it generates a potential noise consisting of one or more sine wave responses, not just for the frequencies of the applied signal (2 and 5 Hz) but for other frequencies. For instance, in harmonic analysis, the current response will be at harmonic frequencies However, the so-called intermodulation frequencies can also be used. EFM has a great advantage as it can provide the causality factors CF2 and CF3, which in turn gives evidence for the confirmation of the validity of the obtained data. The efficiency of the corrosion inhibition process and surface coverage was computed by Eq. (3). For all the conducted electrochemical tests, the electrode was allowed to stand first for 30 min to stabilize the electrode potential and OCP was recorded. All the electrochemical evaluation techniques were accomplished using a Gamry PCI4-G750 Potentiostat/Galvanostat/ZRA with Gamry Framework™ ver. 6.33 for experimental control and data recording and Echem Analyst™ ver. 6.33 for data analysis, fitting, and plotting. All electrochemical tests were repeated thrice to maintain the repeatability of the acquired data in this study. The tabulated data of such tests represents the mean value ± standard deviation of the three test replications.
2.5. Quantum chemical (QM) parameters The molecular structures of the tetrazole-based organoselenium compounds were initially geometrically optimized with PM3 semi-empirical computational method as a starting point and then optimized with density functional theory (DFT) calculations. The energy levels of the orbitals were determined at the ground state to make statements about the molecular structure and stability properties [37]. All QM calculations were performed using Spartan 14 from Wavefunction, Inc., USA. DFT calculations were utilized by B3LYP/6-31G* hybrid functional module where solvation effects were treated in the calculations using the conductor-like polarizable continuum (CPCM) solvation model for aqueous solvation. 2.6. Monte Carlo (MC) simulations To explore the configuration of the adsorbed inhibitor molecule on the adsorbent metallic crystal surface, MC simulation methodology was employed by using BIOVIA Materials Studio 2017 from Dassault Systèmes SE, France. The J55 steel surface was simulated by the cleaving of an iron crystal surface in the selected planes of Fe (110, 111, 100), which represents all the main crystal growth faces of Fe crystal [38]. This imitated process was executed in a defined 3D emulated box (32.27 Å × 32.27 Å × 50.18 Å). Periodic boundary conditions were applied to this simulation so that all particles would be subjected to arbitrary homogenous loadings [39]. A slab of vacuum was constructed above all the cleaved planes of Fe (110, 111, 100) with a thickness of 20 Å. Tetrazole-based organoselenium compounds (TOS1, TOS2, and TOS3) were created, and their energy was optimized using Forcite module with a universal force field. After being geometrically optimized, the investigated molecules were listed as adsorbate, the corrosion system was established by loading the inhibitor molecules on the Fe (110, 111, 100) surfaces, and a state of minimum energy was achieved via simulated annealing algorithm. The universal force field was utilized rather than COMPASS as the latter has no assigned force field type for selenium. The non-bonded van der Waals interactions were computed with an atom-based cutoff, and the electrostatic ones were modeled with Ewald summation technique. Using such adsorption locator simulation can verify the energetics of the adsorption process by calculating the energy of interaction between the inhibitor molecule and the iron surface and, accordingly, the inhibition efficiencies of the tetrazole-based organoselenium compounds (TOS1, TOS2, and TOS3) under study. 2.7 SRB corrosion inhibition Infected water samples with SRB were collected from a selected wellhead annulus valve of El Mansoura Petroleum Company’s production well area. This particular location fulfills the static anaerobic conditions, which are ideal for SRB to colonize [40]. Standard culture media kits were 4
203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254
purchased from EPRI, which matches NACE Standard TM0194-14 [41]. The samples were measured for their salt composition right after collection to approximate the field condition and select the best culture media. Four sets of 200 ml sterile bottles were used in such experiment. The first set (blank) was composed of one bottle where field contaminated water was only used with added 5 ml of solvent to cancel the solvent effect. Considering the other three sets of sample bottles for TOS1, TOS2, and TOS3 compounds. Each of the latter mentioned sample sets was composed of two bottles for the used 20 ppm and 50 ppm concentrations under assaying. Stock solutions of 5 ml of the inhibitors under testing was added to each bottle, and then it was diluted by field contaminated water until complete at 200 ml. All bottles were thoroughly mixed, and 1 ml of mixture was withdrawn from each bottle and then injected in the sealed media to make a serial dilution. The injected culture media have been directly placed afterward in an incubator at a pre-set temperature of 40 °C to simulate the sample’s original temperature condition and prevent heat shocking effects on the bacteria. 2.8 Surface examinations using SEM and XPS The morphology of the J55 steel samples before and after immersion in 10% HCl in the absence and presence of 50 ppm TOS1, TOS2, and TOS3 for 24 hrs was investigated using the scanning electron microscopy (SEM) model: JEOL JSM-6390. SEM images were captured with two different magnifications of the metal surface: ×500 (bar = 50 µm) and ×2,000 (bar = 10 µm). X-ray photoelectron spectroscopy (XPS) measurements were conducted using ESCALAB 250Xi, Thermo Scientific, USA. 3. Result and Discussion 3.1 Design and synthesis of the TOS1–TOS3 bioactive inhibitors Motivated by the fact that organoselenium compounds are good nucleophiles [42], these compounds can, therefore, encourage the adsorption process over the metal surface by sharing their electron sets with metal [27]. In the context of the previous hypothesis, organoselenium compounds are accordingly supposed to display superior corrosion inhibition property; however, the anticorrosive properties of these compounds were barely discussed in the literature. Furthermore, tetrazole-based compounds have been intensively used as corrosion inhibitors [43]. Therefore, the aim of the present study is to investigate the corrosion IE% of tetrazoles and organoselenium-based compounds when they are combined in one molecule. The preparation of organoselenium compounds is generally a sophisticated task and usually requires the utilization of toxic/expensive reagents (e.g., potassium selenocyanate, sodium selenide, and disodium diselenides) [44]. A considerable advancement has been recently observed in the synthesis of various classes of organoselenium compounds, including organic selenides, diselenides, and selenocyanates, and selena-heterocyclic [45]. Recently, we have employed a sequential azido-Ugi and nucleophilic substitution approach for the synthesis of tetrazole-based symmetrical diselenide and selenoquinone compounds, as shown in Figure 1. The corresponding antioxidant properties of the synthesized compounds were previously evaluated using different redox assays (e.g., 2,2-diphenyl-1picrylhydrazyl, glutathione peroxidase-like activity, and bleomycin-dependent DNA damage assays) [35]. The compounds showed interesting antioxidant activities, which in turn needs more in-depth studies and additional experiments to investigate their exact antioxidant mode(s) of action [35]. Among the synthesized compounds, TOS1, TOS2, and TOS3 were the most promising and were therefore further selected to be tested as inhibitors against MIC for J55 steel. 3.2 Weight loss measurements Weight loss method is a simple and realistic approach to calculate the corrosion rate via a direct proportion between the cause and effect. The weight loss–time curve depicted in Figure 2 represents the etching of J55 steel sample using a blank of plain 10% HCl solution, and then varying concentration range of TOS2 was added to the solution at 25 °C. For the other two compounds TOS1 and TOS3, the curves are supplied in Supporting Information (S2). The variance of weight loss over 5
255 256 257 258 259 260 261 262 263 264 265 266
time was described to be linear. The comparison of the weight loss curves shows that the weight loss in the presence of the used inhibitors was diminishing with increasing the inhibitors concentration rather than using pare 10% HCl solution. This may promote that the inhibiting action of these molecules was attained by being adsorbed on the metal surface and then the formation of a protective layer to isolate the surface from the corrosive medium [46]. The IE% and surface coverage θ for TOS1, TOS2, and TOS3 compounds measured at 25 °C are manifested in Table 1. The data presented clearly shows the advancement of diselenide compounds (TOS1 and TOS2) IE% values over the mono-selenide derivative (TOS3) when using the same concentration range. This behavior may give an idea on the prevalence of diselenide compounds over the mono-selenide derivative as a corrosion inhibitor. 3.2.1 Effect of temperature
267 268 269 270 271 272 273 274 275 276 277
To quantify the effect of elevated temperature on the inhibition efficiencies of tetrazole-based organoselenium compounds under evaluation, a study of weight loss–time variation was conducted using an elevated temperature range. The enumerated data in Table 2 outline the dependence of IE% with the corresponding temperature for the TOS2 compound. For the other molecules (TOS1 and TOS3), the data are listed in Supporting Information (S3). Interpretation of such data listed in Table 2 reveals that the pattern of decrease in the inhibitor efficiency with increasing temperature assumes that the adsorption of TOS molecules on the J55 steel surface at these conditions follows a state of physical adsorption on the metal surface [47].
278 279 280 281 282 283 284
organoselenium inhibitor species can be acquired from identifying the type of adsorption isotherm. Different adsorption isotherms were found to describe the adsorption process of inhibitors on the steel surface: Langmuir [48], Temkin [49], Frumkin [50], and Freundlich [51]. In this study, between other tested isotherms, the Langmuir model seemed to be the best-fitted isotherm to describe the adsorption of tetrazole-based organoselenium compounds on the J55 steel surface. Langmuir isotherm was utilized to respectively determine the parameters, namely, adsorption equilibrium constant Kads and free energy of adsorption ∆G°ads using Eqs. (5) and (6), as follows:
285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301
3.2.2 Adsorption isotherm Further information on the interaction between the J55 steel surface and the tetrazole-based
=
!
"#
%$+ '. (5) !
K )*+ = ,.,, .
/∆1°234 56
. (6)
Figure 3 resembles a straight-line plot of C/θ vs. C as a typical graph for data fitting using the Langmuir model for all the TOS compounds under study. Equilibrium constant Kads values were determined from the intercepts of the plots, and free energy ∆G°ads values in kJ mol-1 were calculated using Eq. (6) and recorded in Table 3. The negative values of ∆G°ads ensure that the adsorption process is spontaneous and the adsorbed film formed over the J55 steel surface was stable, and its calculated value was approximately -20 kJ mol-1, indicating that the type of adsorption is physical adsorption [52]. The actuality of tetrazole-based organoselenium compounds was physically adsorbed on the metal surface in such a relatively high concentration of Cl- ions (10% HCl). It may play an important role to facilitate and stabilize their adherence to the metal surface. The presence of excess Cl- ions acts as a bridge between the positively charged metal surface and the protonated TOS molecules, which, to a great extent, suits their mechanism of action against well acidizing destructive effects [53].
6
302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347
3.4 Electrochemical impedance spectroscopy technique EIS technique offers a fairly decisive means for the evaluation of corrosion inhibitors [58–62]. The Nyquist diagrams (Figure 5) were outlined at the corresponding OCP for the used concentration range of TOS2 after the immersion of the J55 steel sample in 10% HCl for 30 min. The Nyquist diagrams were shaped like a typical high-frequency distorted semi-circle representing the impedance response with only one time constant. This is mainly due to the fact that only charge transfer resistant process contributes to the total impedance without any role in the diffusion process [63,64]. The distorted semi-circle curve shapes originated from the roughness and inhomogeneity of the surface of the working electrode [65–67]. In the semi-circle shape of the Nyquist plots, whenever the concentration of the inhibitors is increased, the diameter of the semi-circle is elevated, which indicates the increase of the corresponding inhibition power. EIS acquired data were also represented as a vector quantity in the Bode plots (Figure 6) for TOS2 by plotting the phase angle and log frequency that was depicted as only one maximum, i.e., one time constant, at middle frequencies. Thus, only one relaxation process may result from the charge transfer process between the metal–electrolyte interface. Such maximum may experience broadening with the addition of the inhibitor compounds, which confirms the construction of a protecting film along the metal surface. Moreover, Figure 6 represents a plotting of log Z vs. log frequency within the same bode plot, which shows an increased value of the total impedance in the presence of inhibitor compounds rather than that of their absence. Furthermore, the increment of such values is directly related to the increase in the inhibitor compound concentrations. Thus, the rate of corrosion decreases upon the increase of the inhibitor concentration as concluded from the Nyquist plots. The Bode and Nyquist plots for TOS2 and TOS3 are enclosed in Supporting Information (S5 and S6). Figure 7 portrays an electrical circuit (equivalent circuit) consisting of three components: a resistance (Rs) of the solution, a resistance (Rct) of the charge transfer, and a capacitor (CPE) that is called constant phase element emerging from the formed electrical double layer at the metal–electrolyte interface. This circuit was used to analyses the data acquired from the EIS spectra, particularly because it was the best fit with the experimental outcomes [68]. Eq. (7) describes the impedance as a function of CPE [69]:
348 349 350 351 352 353
Cdl = Y 7max n-1 , (7) where ωmax denotes the angular frequency when the imaginary component of the impedance at its maximum value, Y0 is the magnitude of CPE, and n is a CPE exponent that depends on the nature of the metal surface. The EIS data listed in Table 5 illustrate the increase in the Rct values in accordance with the increase in the inhibitor concentration as it forms a protective film on the metal/solution interface. On the
3.3 Potentiodynamic polarization measurements Tafel polarization plots for the corrosion process of J55 steel samples immersed in 10% HCl solution in the absence and presence of the tetrazole-based organoselenium compounds were inspected. Accordingly, Figure 4 shows that the anodic and cathodic sides of the curves were downshifted to a lower value of current density (i) by raising the concentration of TOS2. Likewise, Figure 4 represents the effect of adding various concentrations of TOS2 on the Tafel plots of tested J55 steel surface in 1 M HCl. For the other two compounds (TOS1 and TOS3) represented in Supporting Information (S4), all the obtained electrochemical parameters from the conducted experiment are manifested in Table 4. The values of icorr in Table 4 decrease as the concentration of TOS compounds increases, which indicate the decrease in corrosion rate and increase in IE%. However, for Ecorr, no significant change (approximately 40 mV) was observed, indicating that the used compounds are of a mixed-type corrosion inhibitor [54]. Tafel constants βa and βc for the TOS compounds were found to be altered with concentration increment. This finding confirms that the inhibitors hinder the cathodic discharge of hydrogen and anodic metal dissolution reaction by covering the exposed surface. Thus, they are considered as mixed-type inhibitors [55–57]. The IE% values show that the diselenide compounds (TOS1 and TOS2) have an increase over the mono-selenide (TOS3) derivative, which may elaborate the effect of selenium atom on the inhibition process.
7
354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405
contrary, the Cdl values decrease with the concentration, which can be explained by the thickening of the formed protective film overlaid by more adsorbed molecules that can be represented in Eq. (8) [70,71]: Cdl = εε /d , (8) where d is the thickness of the adsorbed film, ε is the dielectric constant of the adsorbed film, and ε0 is the vacuum permittivity. The data listed in Table 5 clearly confirm how the compounds under investigation (TOS1, TOS2, and TOS3) can serve as excellent corrosion inhibitors for the protection of the J55 steel metal surface with quite a superiority to the diselenide compounds (TOS1 and TOS2) over the mono-selenide compound (TOS3). These data were also found to be in line with those obtained from the other used techniques. 3.5 Electrochemical frequency modulation Utilizing EFM technique at intermodulation frequencies, parameters, such as corrosion current density (icorr) and anodic and cathodic Tafel parameters (βa and βc) can be deduced as a system response to the corrosion process [72]. Causality factors CF2 and CF3 acquired from the EFM technique offers a means to appraise the validity of the conducted experiment. The depicted curves from the EFM techniques represent two current response peaks appearing at 2 and 5 Hz of intermodulation frequencies, which were analyzed to give the Tafel parameters and the corrosion rate [73]. The EFM spectra for the evaluation of various ranges of the concentration of the inhibitor compound TOS2 against the corrosive action of 10% HCl on the J55 steel are represented in Figure 8. The EFM curves of TOS1 and TOS3 are enclosed in Supporting Information (S7). The acquired experimental data are listed in Table 6, including icorr, βa, βc, and IE% along with the causality factors at the applied concentration range of the compounds under evaluation at 25 °C. The values of icorr in Table 6 decrease as the concentration of tetrazole-based organoselenium compounds increases, which indicate the decrease of corrosion rate and increase in the IE%. Moreover, Tafel constants βa and βc for TOS compounds were found to alter with concentration increment, and therefore, they are considered mixed-type inhibitors. TOS1 and TOS2 scored increased IE% values in comparison with TOS3, which was consistent with previous methods. The obtained causality factor values CF2 and CF3 gave a valid index for the experiment in accordance with their theoretical values [74]. 3.6. QM parameters EHOMO can be regarded as a measure of the ability of molecules to donate electrons into vacant molecular orbitals of an appropriate acceptor. On the contrary, ELUMO estimates the tendency of a molecule to receive electrons from another. The lower the ELUMO value, the more the ability of the molecule to receive electrons [75]. By contrast, the higher the EHOMO value, the more electrondonating ability the molecule would exhibit toward the vacant d-orbital of the metallic iron at the surface, and consequently the more inhibition power it would acquire. The values listed in Table 7 display that the highest energy EHOMO in favor of diselenide compounds TOS1 and TOS2. Therefore, they will have the highest corrosion inhibition action [76,77]. Inspecting how the HOMO level is distributed over the structure, Figure 9 illustrates how it is mostly centered over the diselenide bond and somehow into the closer benzene rings. This finding indicates how the presence of two selenium atoms reinforces the chances for the molecule being adsorbed at the metal surface as they are proven to be preferred sites for the electrophilic attack. Thus, this may explain how the diselenide compounds TOS1 and TOS2 scored higher IE% rather than monoselenide derivative TOS3 as this claim was in line with the interpreted experimental results. The energy gap, ∆Egap, between the HOMO and LUMO levels can give a great index in understanding the molecular reactivity of the compounds under study. The lower the value of ∆Egap, the more reactive the molecule would be and the more inhibiting action the molecule would exhibit [78,79]. This arises because of the tendency to accept electrons; these molecules can offer as the relationship between the metal and adsorbed molecule is not limited to electron donation between the molecule and metal that unoccupied the d-orbital but also the electron reception from the metal to the anti-bonding orbital of the compound creating a back bonding in return. As shown in Table 7, TOS3 has the highest ∆Egap, 8
406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457
and TOS1 and TOS2 seem to have lower values. Thus, diselenide compounds are expected to have a stronger adherence to the metal surface when compared with the mono-selenide derivative. The results observed a relationship between the molecular area of the molecule and its IE%. The IE% increases as the molecular area of the molecule increases due to the increase of the contact area between the molecule and the surface of the metal. Table 7 displays that TOS2 has the highest molecular surface area and thus more IE% compared with TOS1 and TOS3. Furthermore, the obtained results were in a good agreement with those of previously used techniques. 3.7 MC simulations Figure 10 shows that the configured suited visualization of adsorbed tetrazole-based organoselenium molecules on the surface of J55 steel [Fe (110, 111, 100)] substrates obtained by the adsorption locator package using MC simulations and highlighted by the Materials Studio visualizer interface. The geometrically optimized molecules under study being loaded to the Fe (110, 111, 100) surfaces show that such molecules tend to have an almost planar orientation into the surface and thus maximize the surface coverage and in turn give less chances for acidic aggressive attacks toward the metal surface. Comparing the surface orientations of the molecules under study, TOS2 seems to have the superiority in such argument, which is in agreement with the results of the above techniques. The data listed in Table 8 represent all the outputs and descriptors computed by MC simulations, which include adsorption energy, rigid adsorption energy, and deformation energy (calculated as kcal mol-1). The adsorption energy values show that high energy is released (high value of a negative sign), which can be regarded as an act of high affinity of the adsorbed molecules species of H2O and inhibitor under study (TOS1, TOS2, and TOS3 in this case) toward the adsorbate surfaces of Fe (110, 111, 100). The outlined adsorption energy was mathematically calculated by the summation of the rigid adsorption energy and the deformation energy of the adsorbate molecules [80,81]. The rigid adsorption energy expresses the amount of released energy when the molecules initially loaded on the adsorbate surface before being geometrically relaxed and the deformation energy known as the energy released when the adsorbed-adsorbate components undergo relaxation on the surface [82]. The data in Table 8 show that the major differences between the adsorption cases of all molecules under study (TOS1, TOS2, and TOS3) refers to the deformation energy and not the rigid adsorption energy where solvated TOS2 scored the highest deformation energy (approximately -4972 kcal mol-1) for all the employed Fe crystal growth surfaces. This finding proves that the geometrical structural difference of the adsorbed molecules plays the most crucial role in the adsorption affinity and corrosion IE%, which was in the following order: TOS2 > TOS1 > TOS3. Moreover, dEads/dNi reports the adsorption energy, in kcal mol-1, of the individual species action of substrate–adsorbate configurations [83]. As manifested in Table 8, TOS2 has attained a maximum value of dEads/dNi in the simulation process, which indicates its highest contribution toward the total adsorption energy. Comparing all the adsorption energy data for the used Fe (110, 111, 100) surfaces, the rigid adsorption energy and consequently adsorption energy have scored the highest value in the case of Fe (110) surface rather than Fe (111) or Fe (100) as listed in Table 8. This finding may also explain the superiority of TOS compounds adherence to the surface because Fe (110) is responsible for more than 63% of the total Fe crystal growth surface area [84,85] and Fe (110) is a bulk terminated surface and has no surface reconstruction, no relaxations, and the closest pack structure of iron crystals [86]. The calculated relative surface coverage of the TOS compounds listed in Table 9 support this argument as coverage values were utmost in the case of Fe (110). This computational study upholds the claims of the chemical and electrochemical techniques for the superiority of TOS compounds as a corrosion inhibitor and proves the consistency of the all acquired data.
9
458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509
3.8 SRB corrosion inhibition Biological corrosion is the consequence of the undesirable combination of the aggressive aqueous medium, susceptible material to corrosion with the suitable anaerobic condition, and bacterial strains whose presence is often unexpected [87]. SRB is a crucial cause of anaerobic biological corrosion of iron and steel. Although tetrazole-based organoselenium compounds are well-known antibacterial agents, their biocidal effect on MIC is poorly discussed in the literature. The potential biocides of TOS1, TOS2, and TOS3 for the inhibition of biological corrosion were evaluated using NACE Standard TMO194-04. This method is used extensively in the evaluation of bacterial growth rates in oil and gas wells, pipelines, and processing plant units, such as tanks and separators. In addition, this test can be used to determine the efficiency of currently used aldehyde-based biocides in the oil and gas industry [88]. Figure 11 represents the used synthetic culture media kits to quantify the bacterial growth rate for the used water sample retrieved from the well annulus under study in the presence and absence of specified concentrations of TOS1, TOS2, and TOS3. These media were encapsulated into sealed vials so that it can provide anaerobic condition and they were put into a dilution series. Using concentrations of 20 and 50 ppm, TOS1, TOS2, and TOS3 showed promising biocidal activity toward the SRB. This finding can be seen clearly in the reduction of the number blackened vials. Taking a closer look to the data listed in Table 10, TOS2 and TOS3 show the highest inhibition efficiencies (up to 80%) as they only score 102 cell/ml and 101 cell/ml in the 20 ppm and 50 ppm concentrations, respectively. The rise in the antibacterial behavior of compounds TOS2 and TOS3 against SRB may be attributed to the presence of furan and naphthoquinone moieties, respectively (Figure 12). This feature was identified by the fact that most of the furan- and naphthoquinone-based derivatives have strong antibacterial activity and have an active biocidal effect toward microbial corrosion [89–92]. 3.9 Surface morphology using SEM The scanning electron micrographs allow a detailed representation of the surface morphology. Before being immersed in 10% HCl (free), the J55 steel surface was described to be smooth and clear without any damage, as shown in Figure 13a. After being immersed in 10% HCl (blank) for 24 hours, the J55 steel surface suffered from severe changes in the surface structure (damage) due to the impact of corrosion exerted by HCl (Figure 13b) when compared with the free metal sample. Inspecting the micrographs on the addition of 50 ppm TOS1, TOS2, and TOS3 compounds for the same period of 24 hours (Figures 14a–14c), the metal surface appears to have a much smoother texture. The obtained micrographs support the mechanism of adsorption of the molecules on the surface of the metal, leading to the formation of a protective layer that shields the surface from corrosive attacks [93]. 3.10 X-ray Photoelectron Spectroscopy (XPS) The high-resolution XPS analysis was performed to confirm the adsorption of the TOS molecules on the J55 steel surface. The XPS spectra recorded for J55 steel surface corroded in 10% HCl were composed of C 1s, Cl 2p, Fe 2p, and O 1s elements as shown in Figures 15a–15d. In the presence of the TOS2 inhibitor, the XPS spectra involved the same elements (C 1s, Cl 2p, Fe 2p and, O 1s) along with N 1s core level and selenium (Se 3d) as shown in Figures 16a–16f. The binding energies (BE, eV) and the corresponding assignment of each peak component are summarized in Table 11 [94–102]. XPS spectrum of N 1s showed a single peak at 400.5 eV, which could be attributed to the neutral imine (-N=) and amine (-N-H) nitrogen atoms as previously reported [95]. Also, the deconvoluted spectra of Se showed four peaks of Se 3d at 54.0 eV, 54.5 eV, 55.4 eV and 56.7 eV which represents Se adsorption on Fe, Se-Se bond, elemental selenium (Se0) and oxidized selenium species, respectively [96–98]. The peaks of N and Se in the surface spectra of the inhibited sample approve the adsorption of the TOS2 inhibitor at the surface. Moreover, inspecting the C 1s spectra of J55 steel present in plain HCl, three characteristic peaks were recorded at binding energy values of about 284.7 eV, 286.5 eV and 288.2 eV. However, in the presence of TOS2, one additional peak was observed at 285.2 eV which corresponds to the C-C (alkyl group) 10
510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561
present in the inhibitor, which furtherly proved its adsorption on the surface [99]. According to the XPS results, we can confirm the adsorption of the TOS2 on the J55 steel surface in HCl solution. 4. Conclusions Three tetrazole-based organoselenium compounds, namely, TOS1, TOS2, and TOS3, were synthesized via azido-Ugi and sequential azido-Ugi/nucleophilic substitution reactions. The potential corrosion inhibition of TOS1, TOS2, and TOS3 was assessed against the acidizing process to stimulate oil well production. Different evaluation techniques were employed using chemical and electrochemical methods, quantum chemical calculations, and MC simulations, as well as surface morphology studies using SEM. TOS1, TOS2, and TOS3 were proven to serve as good corrosion inhibitors acting as mixed cathodic and anodic inhibitors for J55 steel corrosion in 10% HCl solution. The results obtained from chemical and electrochemical measurements showed that the IE% increases with the concentration of the inhibitors and decreases with increasing temperature. Furthermore, double-layer capacitances decrease with respect to the blank solution when TOS1, TOS2, and TOS3 were present. This fact proves the adsorption of the tetrazole-based organoselenium molecules on the J55 steel surface. As predicted, the adsorption of TOS1, TOS2, and TOS3 on the J55 steel surface obeyed the Langmuir adsorption isotherm and data showed that this adsorption was found to be physisorption. Also, the adsorption of the TOS compounds on the J55 steel surface in HCl solution was confirmed via XPS analysis. Remarkably, experimental, QM, and MC results confirmed this claim and the order of IE% for the tetrazole-based organoselenium compounds was TOS2 > TOS1 > TOS3. The presence of high chloride ion concentrations led to the stabilization of physically adsorbed TOS molecules by connecting the positively charged metal surface and the protonated organic inhibitor molecules. The latter acted in favor of the tetrazole-based organoselenium agents’ IE%. In addition, TOS1, TOS2, and TOS3 have evinced their dual functionality of their biocidal action toward the MIC. Therefore, these compounds can be utilized in mitigating corrosion caused by stagnant infected formation waters found in locations, such as annular spacing, or even locations within treatment plants, such as settling water in the bottom part of oil tanks. 5. References [1] M.B. Kermani, D. Harrop, The Impact of Corrosion on Oil and Gas Industry, SPE Prod. Facil. 11 (1996) 186–190. doi:10.2118/29784-pa. [2] R. Javaherdashti, How corrosion affects industry and life, Anti-Corrosion Methods Mater. 47 (2000) 30–34. doi:10.1108/00035590010310003. [3] T. Xu, Z. Jin, Y. Feng, S. Song, D. Wang, Study on the static and dynamic fracture mechanism of different casing-drilling steel grades, Mater. Charact. 67 (2012) 1–9. doi:10.1016/j.matchar.2012.02.016. [4] M.A. Migahed, I.F. Nassar, Corrosion inhibition of Tubing steel during acidization of oil and gas wells, Electrochim. Acta. 53 (2008) 2877–2882. doi:10.1016/j.electacta.2007.10.070. [5] K. Haruna, I.B. Obot, N.K. Ankah, A.A. Sorour, T.A. Saleh, Gelatin: A green corrosion inhibitor for carbon steel in oil well acidizing environment, J. Mol. Liq. 264 (2018) 515–525. doi:10.1016/j.molliq.2018.05.058. [6] A. Singh, K.R. Ansari, A. Kumar, W. Liu, C. Songsong, Y. Lin, Electrochemical, surface and quantum chemical studies of novel imidazole derivatives as corrosion inhibitors for J55 steel in sweet corrosive environment, J. Alloys Compd. 712 (2017) 121–133. doi:10.1016/j.jallcom.2017.04.072. [7] A. Singh, K.R. Ansari, J. Haque, P. Dohare, H. Lgaz, R. Salghi, M.A. Quraishi, Effect of electron donating functional groups on corrosion inhibition of mild steel in hydrochloric acid: Experimental and quantum chemical study, J. Taiwan Inst. Chem. Eng. 82 (2018) 233–251. doi:10.1016/j.jtice.2017.09.021. [8] E.B. Ituen, A.O. James, O. Akaranta, Fluvoxamine-based corrosion inhibitors for J55 steel in aggressive oil and gas well treatment fluids, Egypt. J. Pet. 26 (2017) 745–756. doi:10.1016/j.ejpe.2016.10.002. 11
562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612
[9] [10] [11]
[12] [13]
[14]
[15]
[16] [17]
[18] [19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
J. Brzeszcz, A. Turkiewicz, Corrosion inhibitors–application in oil industry, Nafta-Gaz. 71 (2015). H. De Reus, E.L.J.A. Hendriksen, M.J.J.S. Thomas, Field Corrosivity Toolbox to Optimize Corrosion Control, Corros. 2003. (2003) 10. https://doi.org/. D. V Murphy, E.A. Stockdale, P.C. Brookes, K.W.T. Goulding, Impact of Microorganisms on Chemical Transformations in Soil, Soil Biol. Fertil. (n.d.) 37–59. doi:10.1007/978-1-40206619-1_3. T.B. Parkin, E.C. Berry, Microbial nitrogen transformations in earthworm burrows, Soil Biol. Biochem. 31 (1999) 1765–1771. doi:10.1016/s0038-0717(99)00085-1. K.-H. Lee, Y.-F. Wang, Y. Wang, J.-D. Gu, J.J. Jiao, Abundance and Diversity of Aerobic/Anaerobic Ammonia/Ammonium-Oxidizing Microorganisms in an Ammonium-Rich Aquitard in the Pearl River Delta of South China, Microb. Ecol. 76 (2016) 81–91. doi:10.1007/s00248-016-0815-8. H. Qu, R. Ma, D. Liu, J. Gao, F. Wang, Z. Zhou, P. Wang, Environmental behavior of the chiral insecticide fipronil: Enantioselective toxicity, distribution and transformation in aquatic ecosystem, Water Res. 105 (2016) 138–146. doi:10.1016/j.watres.2016.08.063. K. Kowalska, E. Felis, Identification of selected microorganisms from activated sludge capable of benzothiazole and benzotriazole transformation, Acta Biochim. Pol. 62 (2015) 935– 939. doi:10.18388/abp.2015_1167. R. Javaherdashti, Microbiologically Influenced Corrosion (MIC), Eng. Mater. Process. (2016) 29–79. doi:10.1007/978-3-319-44306-5_4. Y. Li, D. Xu, C. Chen, X. Li, R. Jia, D. Zhang, W. Sand, F. Wang, T. Gu, Anaerobic microbiologically influenced corrosion mechanisms interpreted using bioenergetics and bioelectrochemistry: A review, J. Mater. Sci. Technol. 34 (2018) 1713–1718. doi:10.1016/j.jmst.2018.02.023. G. Voordouw, J. Voordouw, T. Pinnock, Effect of Biocides and Corrosion Inhibitors on SRBMediated MIC Under Flow Conditions, Corros. 2017. (2017) 8. https://doi.org/. T. Gu, R. Jia, T. Unsal, D. Xu, Toward a better understanding of microbiologically influenced corrosion caused by sulfate reducing bacteria, J. Mater. Sci. Technol. 35 (2019) 631–636. doi:10.1016/j.jmst.2018.10.026. B. Anandkumar, R.P. George, S. Maruthamuthu, N. Parvathavarthini, U.K. Mudali, Corrosion characteristics of sulfate-reducing bacteria (SRB) and the role of molecular biology in SRB studies: an overview, Corros. Rev. 34 (2016). doi:10.1515/corrrev-2015-0055. M.H. Thomsen, T.L. Skovhus, M. Mashietti, Application of natural antimicrobial compounds for reservoir souring and MIC prevention in offshore oil and gas production systems, in: DHRTC Danish Hydrocarb. Res. Technol. Cent. Technol. Conf. 2017, 2017. J. Larsen, L. Holmkvist, K. Sørensen, T.L. Skovhus, Identification And Quantification of Microorganisms Involved In Downhole MIC In a Dan Oil Producing Well, Corros. 2011. (2011) 9. https://doi.org/. T.L. Skovhus, R.B. Eckert, E. Rodrigues, Management and control of microbiologically influenced corrosion (MIC) in the oil and gas industry—Overview and a North Sea case study, J. Biotechnol. 256 (2017) 31–45. doi:10.1016/j.jbiotec.2017.07.003. H.U. Nwankwo, L.O. Olasunkanmi, E.E. Ebenso, Experimental, quantum chemical and molecular dynamic simulations studies on the corrosion inhibition of mild steel by some carbazole derivatives, Sci. Rep. 7 (2017). doi:10.1038/s41598-017-02446-0. D. Djouahra-Fahem, Y. Angar, L.M. Gana, F. Khoukhi, S. Kebbouche-Gana, A Comprehensive Study on Crude Methanolic Extract of Daphne gnidium L. as Effective Corrosion Inhibitors of Mild Steel Induced by SRB Consortium, J. Bio-and Tribo-Corrosion. 5 (2019) 18. I. EO, J.C. Orji, C.O. Nweke, Corrosion and Fungal Growth Inhibiting Effects of Piper guineense Extracts, J. Appl. Environ. Microbiol. 6 (2018) 37–41. 12
613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662
[27]
[28]
[29]
[30]
[31]
[32] [33]
[34]
[35]
[36] [37]
[38]
[39] [40] [41] [42] [43] [44]
[45]
A.H. El-Askalany, S.I. Mostafa, K. Shalabi, A.M. Eid, S. Shaaban, Novel tetrazole-based symmetrical diselenides as corrosion inhibitors for N80 carbon steel in 1 M HCl solutions: Experimental and theoretical studies, J. Mol. Liq. 223 (2016) 497–508. doi:10.1016/j.molliq.2016.08.088. T.H.D. Nguyen, B. Vardhanabhuti, M. Lin, A. Mustapha, Antibacterial properties of selenium nanoparticles and their toxicity to Caco-2 cells, Food Control. 77 (2017) 17–24. doi:10.1016/j.foodcont.2017.01.018. S. Michelle, W. Steve, W. Thomas, Efficacy and mechanism of selenium nanoparticles as antibacterial agents, Front. Bioeng. Biotechnol. 4 (2016). doi:10.3389/conf.fbioe.2016.01.03040. M. Jawad Nasim, W. Ali, E. Domínguez-Álvarez, E.N. da Silva Júnior, R.S.Z. Saleem, C. Jacob, Chapter 10. Reactive Selenium Species: Redox Modulation, Antioxidant, Antimicrobial and Anticancer Activities, Organoselenium Compd. Biol. Med. (n.d.) 277–302. doi:10.1039/9781788011907-00277. N. Srivastava, M. Mukhopadhyay, Green synthesis and structural characterization of selenium nanoparticles and assessment of their antimicrobial property, Bioprocess Biosyst. Eng. 38 (2015) 1723–1730. doi:10.1007/s00449-015-1413-8. S.D. Zhu, J.F. Wei, R. Cai, Z.Q. Bai, G.S. Zhou, Corrosion failure analysis of high strength grade super 13Cr-110 tubing string, Eng. Fail. Anal. 18 (2011) 2222–2231. I.B. Obot, N.K. Ankah, A.A. Sorour, Z.M. Gasem, K. Haruna, 8-Hydroxyquinoline as an alternative green and sustainable acidizing oilfield corrosion inhibitor, Sustain. Mater. Technol. 14 (2017) 1–10. S.A. Umoren, M.M. Solomon, S.A. Ali, H.D.M. Dafalla, Synthesis, characterization, and utilization of a diallylmethylamine-based cyclopolymer for corrosion mitigation in simulated acidizing environment, Mater. Sci. Eng. C. 100 (2019) 897–914. S. Shaaban, A. Negm, A.M. Ashmawy, D.M. Ahmed, L.A. Wessjohann, Combinatorial synthesis, in silico , molecular and biochemical studies of tetrazole-derived organic selenides with increased selectivity against hepatocellular carcinoma, Eur. J. Med. Chem. 122 (2016) 55–71. doi:10.1016/j.ejmech.2016.06.005. R.G. Parr, R.A. Donnelly, M. Levy, W.E. Palke, Electronegativity: The density functional viewpoint, J. Chem. Phys. 68 (1978) 3801–3807. doi:10.1063/1.436185. Quantum chemical studies on inhibition activities of 2,3-dihydroxypropyl-sulfanyl derivative on carbon steel in acidic media, Int. J. Corros. Scale Inhib. 7 (2018). doi:10.17675/2305-68942018-7-4-2. S. Kaya, L. Guo, C. Kaya, B. Tüzün, I.B. Obot, R. Touir, N. Islam, Quantum chemical and molecular dynamic simulation studies for the prediction of inhibition efficiencies of some piperidine derivatives on the corrosion of iron, J. Taiwan Inst. Chem. Eng. 65 (2016) 522–529. F. Radjai, C. Voivret, Periodic boundary conditions, (n.d.). S. Okabe, W.L. Jones, W. Lee, W.G. Characklis, Anaerobic SRB biofilms in industrial water systems: a process analysis, in: Biofouling Biocorrosion Ind. Water Syst., 1993: pp. 189–204. N. International, Field Monitoring of Bacterial Growth in Oil and Gas Systems, (2014) 32. https://doi.org/. S. Shaaban, M.A. Arafat, W.S. Hamama, Vistas in the domain of organoselenocyanates, 2014. L. V Myznikov, A. Hrabalek, G.I. Koldobskii, Drugs in the tetrazole series. (Review), Chem. Heterocycl. Compd. 43 (2007) 1–9. doi:10.1007/s10593-007-0001-5. W.S. Hamama, G.G. El-Bana, S. Shaaban, O.M.O. Habib, H.H. Zoorob, Advances in the domain of 4-amino-3-mercapto-1,2,4-triazine-5-ones, RSC Adv. 6 (2016) 24010–24049. doi:10.1039/c5ra26433b. S. Shaaban, E. Gaffer, M. Alshahd, S.S. Elmorsy, Cytotoxic symmetrical thiazole diselenides with increased selectivity against MCF-7 breast cancer cells, Int. J. Res. Dev. Pharm. Life Sci. 13
663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712
[46]
[47]
[48]
[49] [50] [51] [52]
[53]
[54]
[55]
[56]
[57]
[58]
[59] [60] [61] [62] [63]
4 (2015) 1654–1668. K. Shalabi, A.M. Helmy, A.H. El-Askalany, M.M. Shahba, New pyridinium bromide monocationic surfactant as corrosion inhibitor for carbon steel during chemical cleaning: Experimental and theoretical studies, J. Mol. Liq. 293 (2019) 111480. S.E. Nataraja, T. V Venkatesha, K. Manjunatha, B. Poojary, M.K. Pavithra, H.C. Tandon, Inhibition of the corrosion of steel in hydrochloric acid solution by some organic molecules containing the methylthiophenyl moiety, Corros. Sci. 53 (2011) 2651–2659. doi:10.1016/j.corsci.2011.05.004. A. Anejjar, R. Salghi, A. Zarrouk, O. Benali, H. Zarrok, B. Hammouti, E.E. Ebenso, Inhibition of carbon steel corrosion in 1 M HCl medium by potassium thiocyanate, J. Assoc. Arab Univ. Basic Appl. Sci. 15 (2014) 21–27. doi:10.1016/j.jaubas.2013.06.004. W. Durnie, Development of a Structure-Activity Relationship for Oil Field Corrosion Inhibitors, J. Electrochem. Soc. 146 (1999) 1751. doi:10.1149/1.1391837. M. Christov, A. Popova, Adsorption characteristics of corrosion inhibitors from corrosion rate measurements, Corros. Sci. 46 (2004) 1613–1620. doi:10.1016/j.corsci.2003.10.013. H. Otmačić, E. Stupnišek-Lisac, Copper corrosion inhibitors in near neutral media, Electrochim. Acta. 48 (2003) 985–991. doi:10.1016/s0013-4686(02)00811-3. C. Verma, P. Singh, M.A. Quraishi, A thermodynamical, electrochemical and surface investigation of Bis (indolyl) methanes as Green corrosion inhibitors for mild steel in 1 M hydrochloric acid solution, J. Assoc. Arab Univ. Basic Appl. Sci. 21 (2016) 24–30. doi:10.1016/j.jaubas.2015.04.003. C. Jeyaprabha, S. Sathiyanarayanan, G. Venkatachari, Influence of halide ions on the adsorption of diphenylamine on iron in 0.5M H2SO4 solutions, Electrochim. Acta. 51 (2006) 4080–4088. doi:10.1016/j.electacta.2005.11.026. F.M. Donahue, K. Nobe, Closure to “Discussion of ‘Theory of Organic Corrosion Inhibitors— Adsorption and Linear Free Energy Relationships ’ [F. M. Donahue and K. Nobe (pp. 886– 891, Vol. 112, No. 9)],” J. Electrochem. Soc. 113 (1966) 634. doi:10.1149/1.2424060. F. Bentiss, M. Lebrini, H. Vezin, F. Chai, M. Traisnel, M. Lagrené, Enhanced corrosion resistance of carbon steel in normal sulfuric acid medium by some macrocyclic polyether compounds containing a 1,3,4-thiadiazole moiety: AC impedance and computational studies, Corros. Sci. 51 (2009) 2165–2173. doi:10.1016/j.corsci.2009.05.049. D.K. Singh, S. Kumar, G. Udayabhanu, R.P. John, 4(N,N-dimethylamino) benzaldehyde nicotinic hydrazone as corrosion inhibitor for mild steel in 1M HCl solution: An experimental and theoretical study, J. Mol. Liq. 216 (2016) 738–746. doi:10.1016/j.molliq.2016.02.012. M.N.H. Moussa, A.A. El-Far, A.A. El-Shafei, The use of water-soluble hydrazones as inhibitors for the corrosion of C-steel in acidic medium, Mater. Chem. Phys. 105 (2007) 105– 113. doi:10.1016/j.matchemphys.2007.04.007. M. Benabdellah, R. Touzani, A. Aouniti, A. Dafali, S. El Kadiri, B. Hammouti, M. Benkaddour, Inhibitive action of some bipyrazolic compounds on the corrosion of steel in 1M HCl, Mater. Chem. Phys. 105 (2007) 373–379. doi:10.1016/j.matchemphys.2007.05.001. D.C. Silverman, J.E. Carrico, Electrochemical Impedance Technique — A Practical Tool for Corrosion Prediction, CORROSION. 44 (1988) 280–287. doi:10.5006/1.3583938. W.J. Lorenz, F. Mansfeld, Determination of corrosion rates by electrochemical DC and AC methods, Corros. Sci. 21 (1981) 647–672. doi:10.1016/0010-938x(81)90015-9. D.D. Macdonald, M.C.H. McKubre, Impedance Measurements in Electrochemical Systems, Mod. Asp. Electrochem. (1982) 61–150. doi:10.1007/978-1-4615-7458-3_2. F. Mansfeld, Recording and Analysis of AC Impedance Data for Corrosion Studies, CORROSION. 37 (1981) 301–307. doi:10.5006/1.3621688. C. Gabrielli, Identification of Electrochemical Processes by Frequency Response Analysis: Technical Report Number 004/83, Schlumberger Instruments, 1984. 14
713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762
[64]
[65]
[66]
[67]
[68] [69] [70] [71] [72] [73]
[74]
[75] [76]
[77]
[78]
[79]
[80]
[81]
[82]
S.F. Mertens, C. Xhoffer, B.C. De Cooman, E. Temmerman, Short-Term Deterioration of Polymer-Coated 55% Al-Zn — Part 1: Behavior of Thin Polymer Films, CORROSION. 53 (1997) 381–388. doi:10.5006/1.3280481. G. Trabanelli, C. Monticelli, V. Grassi, A. Frignani, Electrochemical study on inhibitors of rebar corrosion in carbonated concrete, Cem. Concr. Res. 35 (2005) 1804–1813. doi:10.1016/j.cemconres.2004.12.010. S.K. Shukla, M.A. Quraishi, Cefotaxime sodium: A new and efficient corrosion inhibitor for mild steel in hydrochloric acid solution, Corros. Sci. 51 (2009) 1007–1011. doi:10.1016/j.corsci.2009.02.024. W.R. Fawcett, Z. Kováčová, A.J. Motheo, C.A. Foss, Application of the ac admittance technique to double-layer studies on polycrystalline gold electrodes, J. Electroanal. Chem. 326 (1992) 91–103. doi:10.1016/0022-0728(92)80505-x. E. Barsoukov, J.R. Macdonald, Impedance spectroscopy: theory, experiment, and applications, John Wiley & Sons, 2018. J.R. Macdonald, Impedance spectroscopy, Ann. Biomed. Eng. 20 (1992) 289–305. doi:10.1007/bf02368532. O.L. RIGGS, R.A.Y.M. HURD, Temperature Coefficient of Corrosion Inhibition, CORROSION. 23 (1967) 252–260. doi:10.5006/0010-9312-23.8.252. E. McCafferty, N. Hackerman, Double Layer Capacitance of Iron and Corrosion Inhibition with Polymethylene Diamines, J. Electrochem. Soc. 119 (1972) 146. doi:10.1149/1.2404150. E. McCafferty, Introduction to Corrosion Science, (2010). doi:10.1007/978-1-4419-0455-3. S.S. Abdel-Rehim, K.F. Khaled, N.S. Abd-Elshafi, Electrochemical frequency modulation as a new technique for monitoring corrosion inhibition of iron in acid media by new thiourea derivative, Electrochim. Acta. 51 (2006) 3269–3277. doi:10.1016/j.electacta.2005.09.018. R.W. Bosch, J. Hubrecht, W.F. Bogaerts, B.C. Syrett, Electrochemical Frequency Modulation: A New Electrochemical Technique for Online Corrosion Monitoring, CORROSION. 57 (2001) 60–70. doi:10.5006/1.3290331. E. Kuş, F. Mansfeld, An evaluation of the electrochemical frequency modulation (EFM) technique, Corros. Sci. 48 (2006) 965–979. doi:10.1016/j.corsci.2005.02.023. G. Gao, C. Liang, Electrochemical and DFT studies of β-amino-alcohols as corrosion inhibitors for brass, Electrochim. Acta. 52 (2007) 4554–4559. doi:10.1016/j.electacta.2006.12.058. Y. Feng, S. Chen, W. Guo, Y. Zhang, G. Liu, Inhibition of iron corrosion by 5,10,15,20tetraphenylporphyrin and 5,10,15,20-tetra-(4-chlorophenyl)porphyrin adlayers in 0.5M H2SO4 solutions, J. Electroanal. Chem. 602 (2007) 115–122. doi:10.1016/j.jelechem.2006.12.016. G. Gece, S. Bilgiç, Quantum chemical study of some cyclic nitrogen compounds as corrosion inhibitors of steel in NaCl media, Corros. Sci. 51 (2009) 1876–1878. doi:10.1016/j.corsci.2009.04.003. M. Özcan, İ. Dehri, M. Erbil, Organic sulphur-containing compounds as corrosion inhibitors for mild steel in acidic media: correlation between inhibition efficiency and chemical structure, Appl. Surf. Sci. 236 (2004) 155–164. doi:10.1016/j.apsusc.2004.04.017. L.H. Madkour, S. Kaya, I.B. Obot, Computational, Monte Carlo simulation and experimental studies of some arylazotriazoles (AATR) and their copper complexes in corrosion inhibition process, J. Mol. Liq. 260 (2018) 351–374. doi:10.1016/j.molliq.2018.01.055. R. Hsissou, S. Abbout, A. Berisha, M. Berradi, M. Assouag, N. Hajjaji, A. Elharfi, Experimental, DFT and molecular dynamics simulation on the inhibition performance of the DGDCBA epoxy polymer against the corrosion of the E24 carbon steel in 1.0 M HCl solution, J. Mol. Struct. 1182 (2019) 340–351. doi:10.1016/j.molstruc.2018.12.030. Y.M. Abdallah, K. Shalabi, N.M. Bayoumy, Eco-friendly synthesis, biological activity and evaluation of some new pyridopyrimidinone derivatives as corrosion inhibitors for API 5L 15
763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812
X52 carbon steel in 5% sulfamic acid medium, J. Mol. Struct. 1171 (2018) 658–671. doi:10.1016/j.molstruc.2018.06.045. [83] A.S. Fouda, G.Y. Elewady, K. Shalabi, H.K. Abd El-Aziz, Alcamines as corrosion inhibitors for reinforced steel and their effect on cement based materials and mortar performance, RSC Adv. 5 (2015) 36957–36968. doi:10.1039/c5ra00717h. [84] L. Guo, I.B. Obot, X. Zheng, X. Shen, Y. Qiang, S. Kaya, C. Kaya, Theoretical insight into an empirical rule about organic corrosion inhibitors containing nitrogen, oxygen, and sulfur atoms, Appl. Surf. Sci. 406 (2017) 301–306. [85] L. Guo, X. Ren, Y. Zhou, S. Xu, Y. Gong, S. Zhang, Theoretical evaluation of the corrosion inhibition performance of 1, 3-thiazole and its amino derivatives, Arab. J. Chem. 10 (2017) 121–130. [86] D.E. Jiang, E.A. Carter, Adsorption and diffusion energetics of hydrogen atoms on Fe(110) from first principles, Surf. Sci. 547 (2003) 85–98. doi:10.1016/j.susc.2003.10.007. [87] J.E.O. Ovri, S. Okeahialam, O. Onyemaobi, Microbial corrosion of mild and medium carbon steels, J. Eng. Sci. Technol. 8 (2013) 639–653. [88] J. Fink, Oil field chemicals, Elsevier, 2003. [89] D.A. Wright, R. Dawson, S.J. Cutler, H.G. Cutler, C.E. Orano-Dawson, E. Graneli, Naphthoquinones as broad spectrum biocides for treatment of ship’s ballast water: Toxicity to phytoplankton and bacteria, Water Res. 41 (2007) 1294–1302. doi:10.1016/j.watres.2006.11.051. [90] E. CHELOSSI, M. FAIMALI, Comparative assessment of antimicrobial efficacy of new potential biocides for treatment of cooling and ballast waters, Sci. Total Environ. 356 (2006) 1–10. doi:10.1016/j.scitotenv.2005.03.018. [91] E. Loğoğlu, M. Yilmaz, H. Katircioğlu, M. Yakut, S. Mercan, Synthesis and biological activity studies of furan derivatives, Med. Chem. Res. 19 (2009) 490–497. doi:10.1007/s00044-0099206-8. [92] E. Lukevits, L. Demicheva, Biological activity of furan derivatives (review), Chem. Heterocycl. Compd. 29 (1993) 243–267. doi:10.1007/bf00531499. [93] T.W. Pickthall, A. Carlysle, T. Duncan, Predictive internal corrosion monitoring on a crude oil pipeline: A case study, in: Corros. 2005, NACE International, 2005. [94] N.Z.N. Hashim, K. Kassim, H.M. Zaki, A.I. Alharthi, Z. Embong, XPS and DFT investigations of corrosion inhibition of substituted benzylidene Schiff bases on mild steel in hydrochloric acid, Appl. Surf. Sci. 476 (2019) 861–877. [95] A. Meneguzzi, C.A. Ferreira, M.C. Pham, M. Delamar, P.C. Lacaze, Electrochemical synthesis and characterization of poly (5-amino-1-naphthol) on mild steel electrodes for corrosion protection, Electrochim. Acta. 44 (1999) 2149–2156. [96] S. Cao, Y. Chen, L. Kang, Z. Lin, W.-F. Fu, Enhanced photocatalytic H 2-evolution by immobilizing CdS nanocrystals on ultrathin Co 0.85 Se/RGO–PEI nanosheets, J. Mater. Chem. A. 3 (2015) 18711–18717. [97] R. Golovchak, A. Kovalskiy, A.C. Miller, H. Jain, O. Shpotyuk, Structure of Se-rich As-Se glasses by high-resolution x-ray photoelectron spectroscopy, Phys. Rev. B. 76 (2007) 125208. [98] T. Jiang, Y. Tong, A. Bendounan, F. Nicolas, S. Kubsky, V.A. Esaulov, Selenium and benzeneselenol interaction with Cu (111), RSC Adv. 6 (2016) 84627–84633. [99] X. Wang, W. Yang, F. Li, Y. Xue, R. Liu, Y. Hao, In situ microwave-assisted synthesis of porous N-TiO2/g-C3N4 heterojunctions with enhanced visible-light photocatalytic properties, Ind. Eng. Chem. Res. 52 (2013) 17140–17150. [100] A.S. Fouda, A.A. Nazeer, Assessment of begonia extract as new eco-friendly inhibitor for lowcarbon-steel corrosion in acidic environment, J. Bio-and Tribo-Corrosion. 4 (2018) 7. [101] K. Boumhara, M. Tabyaoui, C. Jama, F. Bentiss, Artemisia Mesatlantica essential oil as green inhibitor for carbon steel corrosion in 1 M HCl solution: Electrochemical and XPS 16
813 814 815 816 817
investigations, J. Ind. Eng. Chem. 29 (2015) 146–155. [102] P. Singh, V. Srivastava, M.A. Quraishi, Novel quinoline derivatives as green corrosion inhibitors for mild steel in acidic medium: electrochemical, SEM, AFM, and XPS studies, J. Mol. Liq. 216 (2016) 164–173.
17
Table 1. weight loss data of the J55 steel corrosion in 10% HCl solution without and with added fixed concentration range of tetrazole-based organoselenium compounds under evaluation at 25°C.
Inhibitors
Conc. ppm
C.R. mg cm-2 min-1
θ
I.E %
10% HCl
--
0.0635
--
--
10
0.0328
0.484
48.4
20
0.0243
0.618
61.8
30 40
0.0158 0.0098
0.752 0.846
75.2 84.6
50
0.0069
0.891
89.1
10
0.0308
0.516
51.6
20
0.0204
0.678
67.8
30 40
0.0135 0.0081
0.787 0.873
78.7 87.3
50
0.0053
0.917
91.7
10
0.0363
0.429
42.9
20
0.0255
0.598
59.8
30
0.0163
0.744
74.4
40 50
0.0108 0.0079
0.831 0.876
83.1 87.6
TOS1
TOS2
TOS3
Table 2. weight loss data of the J55 steel corrosion in 10% HCl solution without and with added fixed concentration range of TOS2 compound under evaluation at 25, 35 and 45 °C. Conc. ppm
10% HCl
10
20
30
40
50
Temp °C
C.R. mg cm-2 min-1
θ
I.E %
25
0.0635
--
--
35
0.0825
--
--
45
0.0995
--
--
25
0.0308
0.516
51.6
35
0.0450
0.454
45.4
45
0.0625
0.372
37.2
25
0.0204
0.678
67.8
35 45
0.0323
0.609
60.9
0.0435
0.563
56.3
25
0.0135
0.787
78.7
35 45
0.0249 0.0345
0.698 0.653
69.8 65.3
25
0.0081
0.873
87.3
35
0.0163
0.803
80.3
45
0.0278
0.721
72.1
25
0.0053
0.917
91.7
35
0.0148
0.821
82.1
45
0.0220
0.779
77.9
Table 3. Langmuir adsorption isotherm outputs for dissolution of J55 steel in 10% HCl solution in the absence and presence of different concentrations of tetrazole-based organoselenium compounds at 25°C Inhibitor
Kads
- ∆G°ads
g L-1
kJ mol-1
TOS1
74.13
20.62
TOS2
87.03
21.02
TOS3
61.28
20.15
Table 4. Potentiodynamic data for J55 steel corrosion in 10% HCl solution without and with added fixed concentration range of tetrazole-based organoselenium compounds under evaluation at 25°C. (standard deviation represented by the error values)
Inhibitors 10% HCl
TOS1
TOS2
TOS3
icorr µA cm-2
-Ecorr mV vs. SCE
βa mV dec-1
βc mV dec-1
CR mpy
θ
% IE
1090 ± 7
455 ± 4
68 ± 1
140 ± 4
499.1 ± 0.5
--
--
10
622 ± 8
432 ± 7
120 ± 3
163 ± 2
284.8 ± 0.5
0.429
42.9
20
462 ± 2
427 ± 4
129 ± 4
177 ± 6
211.5 ± 0.8
0.576
57.6
30
309 ± 5
425 ± 11
119 ± 2
187 ± 2
141.5 ± 0.7
0.717
71.7
40
152 ± 10
430 ± 2
97.5 ± 5
140 ± 3
69.6 ± 0.5
0.861
86.1
50
97 ± 10
435 ± 5
125 ± 4
190 ± 4
44.4 ± 0.3
0.911
91.1
10
543 ± 5
428 ± 6
88 ± 7
130 ± 9
248.6 ± 0.5
0.502
50.2
20
411 ± 6
420 ± 4
99 ± 4
144 ± 7
188.2 ± 0.9
0.623
62.3
30
270 ± 7
415 ± 3
93 ± 3
128 ± 3
123.6 ± 0.8
0.752
75.2
40
123 ± 8
422 ± 4
81 ± 1
113 ± 4
56.3 ± 0.3
0.887
88.7
50
76 ± 7
431 ± 9
77 ± 8
115 ± 7
34.8 ± 0.6
0.930
93.0
10
683 ± 4
443 ± 7
96 ± 2
155 ± 4
312.7 ± 0.7
0.373
37.3
20
575 ± 3
440 ± 3
106 ± 3
181 ± 7
263.3 ± 0.5
0.472
47.2
30
397 ± 5
444 ± 11
113 ± 7
162 ± 6
181.8 ± 0.2
0.636
63.6
40
193 ± 5
453 ± 3
91 ± 9
127 ± 5
88.4 ± 0.4
0.823
82.3
50
120 ± 5
446 ± 9
99 ± 4
188 ± 10
54.9 ± 0.5
0.890
89.0
Conc. ppm --
Table 5. EIS data for J55 steel corrosion in 10% HCl solution without and with added fixed concentration range of inhibitor compounds under evaluation at 25°C. (standard deviation represented by the error values)
Inhibitors
Conc. ppm
Rct ohm cm2
Y0 ×10-4 s ohm-1cm-2
n
Cdl mF cm-2
θ
% IE
10% HCl
--
16.5 ± 0.7
68.72 ± 0.85
0.886 ± 0.005
5.191 ± 0.012
--
--
10 20 30 40 50 10 20 30 40 50
31.8 ± 0.8 43.7 ± 0.7 64.8 ± 0.4 100.1 ± 0.9 165.8 ± 0.8 37.2 ± 0.8 52.1 ± 0.4 75.6 ± 0.6 116.5 ± 0.5 214.6 ± 0.2
21.91 ± 0.45 16.60 ± 0.82 13.14 ± 0.73 9.96 ± 0.13 6.58 ± 0.41 10.11 ± 0.69 5.07 ± 0.36 1.42 ± 0.77 1.39 ± 0.96 1.02 ± 0.08
0.798 ± 0.008 0.846 ± 0.007 0.831 ± 0.010 0.806 ± 0.007 0.819 ± 0.005 0.859 ± 0.007 0.851 ± 0.008 0.798 ± 0.011 0.758 ± 0.005 0.771 ± 0.007
1.112 ± 0.009 1.024 ± 0.018 0.794 ± 0.025 0.572 ± 0.007 0.403 ± 0.007 0.589 ± 0.011 0.268 ± 0.008 0.045 ± 0.022 0.037 ± 0.004 0.033 ± 0.005
0.483 0.623 0.746 0.836 0.901 0.557 0.684 0.782 0.859 0.923
48.3 62.3 74.6 83.6 90.1 55.7 68.4 78.2 85.9 92.3
10
29.0 ± 0.3
22.81 ± 0.95
0.789 ± 0.009
1.102 ± 0.008
0.432
43.2
20 30
40.9 ± 0.5 54.2 ± 0.9
15.22 ± 0.11 10.90 ± 0.08
0.786 ± 0.003 0.789 ± 0.012
0.714 ± 0.006 0.513 ± 0.004
0.598 0.696
59.8 69.6
40
89.5 ± 0.7
8.04 ± 0.03
0.776 ± 0.008
0.376 ± 0.007
0.816
81.6
50
135.7 ± 0.5
5.74 ± 0.05
0.775 ± 0.006
0.273 ± 0.004
0.879
87.9
TOS1
TOS2
TOS3
n
Table 6. EFM data for corrosion ofJ55 steel in 10% HCl solution in the absence and presence of different concentrations of tetrazole-based organoselenium compounds at 25°C. (standard deviation represented by the error values)
Inhibitors
Conc. ppm
10% HCl
--
TOS1
TOS2
TOS3
icorr, µA cm-2
βa, mV dec-1
βc, mVdec-1
C.R. mpy
CF-2
CF-3
θ
% IE
1185±
78± 5
110± 3
541± 8
1.83± 0.02
3.20± 0.08
--
--
11
10
655± 8
92± 7
180± 2
299± 4
1.99± 0.05
3.22± 0.04
0.447
44.7
20
489± 5
115± 8
178± 7
223± 7
2.03± 0.04
3.15± 0.09
0.587
58.7
30
297± 9
97± 5
157± 8
135± 6
1.94± 0.09
3.07± 0.07
0.749
74.9
40
132± 7
132± 2
170± 5
60± 3
1.89± 0.08
2.88± 0.05
0.889
88.9
50
92± 7
93± 4
150± 8
42± 8
2.09± 0.03
2.89± 0.07
0.922
92.2
10
554± 5
111± 8
187± 4
253± 7
1.94± 0.04
2.84± 0.06
0.532
53.2
20
421± 5
133± 9
196± 6
192± 5
2.01± 0.02
2.78± 0.04
0.645
64.5
30
263± 7
131± 8
183± 4
120± 5
1.79± 0.08
3.08± 0.06
0.778
77.8
40
110± 3
111± 7
149± 7
50± 2
2.01± 0.03
2.83± 0.07
0.907
90.7
50
64± 3
122± 7
145± 3
29± 3
1.76± 0.07
2.92± 0.05
0.946
94.6
10
716± 8
119± 5
198± 4
327± 9
1.79± 0.04
2.99± 0.06
0.396
39.6
20
600± 8
110± 7
184± 3
274± 8
2.08± 0.06
2.93± 0.07
0.494
49.4
30
414± 6
98± 7
194± 8
189± 8
2.01± 0.04
2.84± 0.04
0.651
65.1
40
265± 5
100± 8
167± 9
121± 6
1.91± 0.08
2.87± 0.08
0.776
77.6
50
129± 6
89± 9
162± 7
59± 3
2.06± 0.04
3.09± 0.05
0.891
89.1
Table 7. The calculated quantum chemical parameters for tetrazole-based organoselenium compounds. Parameter
TOS1
TOS2
TOS3
E HOMO (eV)
-8.73
-8.59
-9.05
E LUMO (eV)
-3.25
-3.20
-1.46
5.48
5.39
7.59
659.51
778.52
541.4
∆E (eV) 2
Molecular area (Å )
Table 8. Data and descriptors calculated by the Mont Carlo simulation for adsorption of tetrazole-based organoselenium inhibitors on Fe (110, 100, 111).
Structure
Fe (110)
Fe (100)
Fe (111)
Inhibitor
Adsorption energy, kcal mol-1
Rigid adsorption energy, kcal mol-1
Deformation energy, kcal mol-1
TOS: dEads/dNi, kcal mol-1
Water: dEads/dNi, kcal mol-1
TOS1
-5160.08
-360.25
-4799.83
-1575.99
-18.18
TOS2
-5332.07
-360.64
-4971.43
-1745.68
-18.62
TOS3
-4266.26
-352.23
-3914.03
-678.59
-18.21
TOS1
-5067.20
-267.38
-4799.82
-1561.01
-17.35
TOS2
-5237.33
-264.87
-4972.45
-1725.29
-17.47
TOS3
-4173.64
-258.94
-3914.69
-665.38
-17.46
TOS1
-4968.23
-167.91
-4800.32
-1537.18
-16.75
TOS2
-5144.14 -4079.67
-169.66 -165.65
-4974.49 -3914.02
-1710.73 -640.63
-16.90 -16.70
TOS3
Table 9. Calculated relative surface coverage data for TOS compounds over Fe (110, 100, 111) surfaces. Relative surface coverage Structure
TOS1
TOS2
TOS3
Fe (110)
0.322
0.388
0.248
Fe (100)
0.264
0.280
0.180
Fe (111)
0.215
0.248
0.198
Table 10. SRB media count results using concentrations of 20 and 50 ppm of the inhibitor under evaluation and incubated at 40o C for 28 Days.
Inhibitor conc. ppm
inhibitors
SRB Bacterial Count cell/ml
Reduction in bacterial growth cell/ml
Efficiency
--
Blank
105
--
--
TOS1
103
102
40%
TOS2
102
103
60%
TOS3
10
2
10
3
60%
TOS1
10
2
10
3
60%
TOS2
101
104
80%
TOS3
101
104
80%
20
50
Table 11. Binding energies (eV), and their assignments for the major core lines observed for
Core element
C 1s
Cl 2p
Fe2p
O 1s
N 1s
10% HCl
10% HCl + 50 ppm TOS2
BE, eV
Assignments
BE, eV
Assignments
284.7 286.5 288.2 -198.9 200.6 711.1 714.7 718.9 724.3 727.8 732.3 -530.3 531.4 532.4
C-C, C-H C-O C+-O -Cl 2p3/2 Cl 2p1/2 metallic iron Fe 2p3/2 satellite of Fe3+ Fe 2p1/2 of Fe2+ Fe 2p1/2 of Fe3+ Fe 2p1/2 of Fe2+ FeO and Fe2O3 FeO and Fe2O3 FeOOH
284.7 285.2 286.1 288.5 199.2 200.8 711.0 713.7 716.9 720.0 724.4 727.6 730.8 532.9 533.7 --
--
--
400.5
-----
54.0 54.5 55.4 56.7
C-C, C-H C-C C-O, C-N C+-O, C-N+ Cl 2p3/2 Cl 2p1/2 metallic iron Fe3+ Fe 2p3/2 of Fe2+ satellite of Fe3+ Fe 2p1/2 of Fe2+ Fe 2p1/2 of Fe3+ Fe 2p1/2 of Fe2+ FeOOH adsorbed water molecules -neutral imine (-N=) and amine (-N-H) Se adsorption on metal Se-Se Elemental Se0 Oxidized selenium species
--Se 3d --TOS2 treated J55 steel surface.
Fig. 1. Synthesis of tetrazole-based organoselenium (TOS1, TOS2 and TOS3) via azido-Ugi
and sequential nucleophilic substitution (SN) Methodology. 25
10% HCl 10 ppm 20 ppm 30 ppm 40 ppm 50 ppm
Wt-loss, mg cm-2
20
15
10
5
0 0
60
120
180
240
300
360
time, min Fig. 2. Wt. loss-time curve for corrosion of J55 steel in 10% HCl in the absence and presence of TOS2 at 25 ˚C.
0.060
TOS1 R2=0.9897 TOS2 R2=0.9972 TOS3 R2=0.9956
0.055 0.050
C/θ
0.045 0.040 0.035 0.030 0.025 0.020 0.015 0.01
0.02
0.03
0.04
0.05
C, gL-1 Fig. 3. Model of Langmuir isotherm for the adsorption of tetrazole-based organoselenium (TOS1, TOS2 and TOS3) on the surface of J55 steel in 10% HCl at 25 ˚C.
10 1
log i, Acm-2
0.1 0.01 0.001 10% HCl 10 ppm 20 ppm 30 ppm 40 ppm 50 ppm
1E-4 1E-5 1E-6 1E-7 1E-8 -1.0
-0.8
-0.6
-0.4
-0.2
0.0
0.2
E vs SCE, V Fig. 4. Potentiodynamic polarization curves for the corrosion of J55 steel in 10% HCl solution in the absence and presence of different concentrations of TOS2 at 25˚C.
250 225
10% HCl 10 ppm 20 ppm 30 ppm 40 ppm 50 ppm Fitting
200
Zimage, ohm cm2
175 150 125 100 75 50 25 0 -25 -25
0
25
50
75
100 125 150
175
200 225 250
2
Zreal, ohm cm
Fig. 5. Nyquist plots recorded for the corrosion of J55 steel in 10% HCl solution in the absence and presence of different concentrations of TOS2 at 25˚C.
40
20
0
10
1
0.01
-20 10% HCl 10 ppm 20 ppm 30 ppm 40 ppm 50 ppm Fitting 0.1
1
θ, ο
log Z, ohm cm2
100
-40
-60
10
100
1000
-80 10000 1000001000000
log Freq, Hz Fig. 6. Bode plots for the corrosion of J55 steel in 10% HCl solution in the absence and presence of different concentrations of TOS2 at 25˚C.
Fig. 7. Electrical circuit (equivalent circuit) for EIS data Fitting.
0.001
1E-4
10% HCl
10 ppm
log i, Acm-2
log i, Acm-2
1E-4 1E-5 1E-6
1E-6
1E-7
1E-7 1E-8 -0.2
1E-5
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1E-8 -0.2
1.6
0.0
0.2
0.4
Freq, Hz
1.0
1.2
1.4
1E-4
20 ppm log i, Acm-2
log i, Acm-2
0.8
1.6
Freq, Hz
1E-4
1E-5
1E-6
30 ppm
1E-5
1E-6
1E-7
1E-7
1E-8 -0.2
0.6
1E-8 0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
-0.2
0.0
0.2
0.4
Freq, Hz
0.6
0.8
1.0
1.2
1.4
1.6
Freq, Hz
1E-4
40 ppm
1E-6 1E-7 1E-8 -0.2
50 ppm
1E-5
log i, Acm-2
log i, Acm-2
1E-5
1E-6
1E-7
1E-8 0.0
0.2
0.4
0.6
0.8
Freq, Hz
1.0
1.2
1.4
1.6
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Freq, Hz
Fig. 8. Intermodulation spectrums for the corrosion of J55 steel in 10% HCl solution in the absence and presence of different concentrations of TOS2 at 25˚C.
1.4
1.6
TOS2
TOS3
LUMO
HOMO
Optimized Moleclar Structures
TOS1
Fig. 9. Optimized structures, HOMO and LUMO of the tetrazole-based organoselenium molecules (TOS1, TOS2 and TOS3) optimized with DFT method.
TOS2
TOS3
Fe (111)
Fe (100)
Fe (110)
TOS1
Fig. 10. The most suitable configuration for adsorption of the tetrazole-based organoselenium molecules (TOS1, TOS2 and TOS3) on Fe (110, 100, 111) substrates obtained by adsorption locator module.
Fig 11. Used synthetic culture media kits to quantify the bacterial growth rate in the used industrial water sample (a) blank (b) using 20 ppm of TOS1 (c) using 20 ppm of TOS2 (d) using 20 ppm of TOS3 (e) using 50 ppm of TOS1(f) using 50 ppm of TOS2(e) using 50 ppm of TOS3
Fig 12. Naphthoquinone moiety in TOS3 and furan moiety in TOS2 respective structures
10 µm
(b) Blank
(a) Free
50 µm
Fig 13. SEM micrographs for metal surface of J55 steel specimen before (free) and after immersed in 10% HCl (blank) for 24 hours.
10 µm
(c) TOS3
(b) TOS2
(a) TOS1
50 µm
Fig. 14. SEM micrographs for metal surface of J55 steel specimen immersed in 10% HCl with added 50 ppm of tetrazole-based organoselenium (TOS1, TOS2 and TOS3) under evaluation for 24 hours.
0
2
4
6
8
290
289
288
287 286 285 Binding Energy, eV
284
283
Cl 2p3/2
198.9 eV 200.6 eV Fitting Background
C 1s
Intensity, (a.u)
Intensity, (a.u)
284.7 eV 286.5 eV 288.2 eV Fitting Background
10
Cl 2p1/2
282
202
200 198 Binding Energy, eV
(a)
735
730
4
6
8
Fe 2p1/2
10
O1s
Intensity, (a.u)
Intensity, (a.u)
2
530.3 eV 531.4 eV 532.4 Fitting Background
Fe 2p3/2
536 740
196
(b) 0
711.1 eV 714.7 eV 718.9 eV 724.3 eV 727.8 eV 732.3 eV Fitting Background
Cl 2p
725
720
715
710
705
700
534
532
530
528
Binding Energy, eV
Binding Energy, eV
(c)
(d)
Fig 15. X-ray photoelectron deconvoluted profiles of (a) C 1s, (b) Cl 2p, (c) Fe 2p, and (d) O 1s for J55 steel in 10% HCl.
526
0
292
Intensity, (a.u)
Intensity, (a.u)
4
290
288
286
6
199.2 eV 200.8 eV Fitting Background
C 1s
284.7 eV 285.2 eV 286.1 eV 288.5 eV Fitting Background
294
2
284
282
204
203
202
201
200
730
196
4
6
8
10
532.9 eV 533.7 eV Fitting Background
Fe 2p3/2
O1s
Fe 2p1/2
725 720 Binding Energy, eV
715
710
536
535
534
533
(e)
531
530
529
54
53
52
(d) N1s
399.5
399.0
398.5
Se 3d5/2
54.0 eV 54.5 eV 55.4 eV 56.7 eV Fitting Background
Intensity, (a.u) 401.0 400.5 400.0 Binding Energy, eV
532
Binding Energy, eV
Intensity, (a.u)
401.5
197
Intensity, (a.u)
Intensity (a.u.)
Fe 2p
400.5 eV (overlays with fitting) Background
402.0
198
(b) 2
(c)
402.5
199
Binding Energy, eV
0
735
Cl 2p
Cl 2p1/2
(a)
740
10
Cl 2p3/2
Binding Energy, eV
711.0 eV 713.7 eV 716.9 eV 720.0 eV 724.4 eV 727.6 eV 730.8 eV Fitting Background
8
Se 3d3/2
59
58
57
56
55
Binding Energy, eV
(f)
Fig 16. X-ray photoelectron deconvoluted profiles of (a) C 1s, (b) Cl 2p, (c) Fe 2p, (d) O 1s, (e) N 1s and (f) Se 3d for J55 steel in 10% HCl with 50 ppm TOS2.
Research Highlights Tetrazole-based organoselenium bi-functionalized corrosion inhibitors during oil well acidizing: experimental, computational studies and SRB bioassay • Three organoselenium compounds (TOS) were tested as corrosion and SRB • • •
inhibitors. TOS compounds are good mixed-type inhibitors for J55 steel corrosion in 10% HCl. Surface examinations confirmed the adsorption of TOS compounds on J55 steel. TOS compounds showed good biocidal action toward microbial induced corrosion
•
Computational studies were in a good agreement with experimental work.
(MIC).
Declaration of interests The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: