Development of a Zwitterionic Compound Derived from β-Amino Acid

Oct 5, 2017 - (4) Koottungal, L. 2012 worldwide EOR survey. Oil Gas J. .... of mild steel valid up to 250 °C. Corrosion NACE conference & expo 16, pa...
0 downloads 0 Views 4MB Size
Research Article pubs.acs.org/journal/ascecg

Cite This: ACS Sustainable Chem. Eng. 2017, 5, 10396-10406

Development of a Zwitterionic Compound Derived from β‑Amino Acid as a Green Inhibitor for CO2 Corrosive Environments Alejandro Ramírez-Estrada,† Violeta Y. Mena-Cervantes,†,⊥ Ignacio Elizalde,† Arturo Manzo-Robledo,‡ Luis S. Zamudio-Rivera,§ David A. Nieto-Á lvarez,§ Fernando Farelas,∥ and Raúl Hernández-Altamirano*,†,⊥

Downloaded via FORDHAM UNIV on June 29, 2018 at 17:34:42 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Centro Mexicano para la Producción más Limpia, Instituto Politécnico Nacional, Av. Acueducto s/n, Col. La Laguna Ticomán, Ciudad de México, 07340, México ‡ Laboratorio de Electroquímica y Corrosión, DIQI-SEPI-Escuela Superior de Ingeniería Química e Industrias Extractivas (ESIQIE-IPN), UPALM, Instituto Politécnico Nacional, Col. Zacatenco, Edificio Z-5, 3er piso, Ciudad de México, C.P. 07738, México § Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte 152, Col. San Bartolo Atepehuacan, Gustavo A. Madero, Ciudad de México, 07730, México ∥ Institute for Corrosion and Multiphase of Technology, Department of Chemical and Biomolecular Engineering, Ohio University, Athens, Ohio 45701, United States ⊥ Laboratorio Nacional de Desarrollo y Aseguramiento de la Calidad de Biocombustibles (LaNDACBio), Instituto Politécnico Nacional, Ciudad de México, 07340, México ABSTRACT: Herein, the synthesis of three zwitterionic compounds (GZC-8, GZC-12, and GZC-18) derived from β-amino acids was carried out in good yield (>97%) through an ecofriendly, zero-waste synthesis procedure, starting from alkylamines (octylamine, dodecylamine, octadecylamine) and acrylic acid, under solventless conditions at ambient pressure and temperature. These compounds were characterized by Fourier transform infrared, two-dimensional nuclear magnetic resonance, electrospray ionization mass spectrometry, and elemental analyses and then evaluated as potential corrosion inhibitors for CO2 acidic environments characteristic of enhanced oil recovery (EOR) processes of the petroleum industry through electrochemical techniques such as open circuit potential (OCP), lineal polarization resistance (LRP), electrochemical impedance spectroscopy (EIS) at different immersion times, and potentiodynamic polarization after 24 h of immersion. From experimental measurements, inhibition efficiencies of 60, 89, and 90% were obtained for GZC-8, GZC-12, and GZC-18, respectively at a concentration of 25 ppm. In addition, acute toxicity of GZCs was evaluated on Photobacterium phosphoreum (Vibrio fischeri) by Microtox testing from which an EC50 value of 209, 188, and 43 ppm was obtained for GZC-8, GZC-12, and GZC-18, respectively, i.e., 2 orders of magnitude lower than typical ionic liquid inhibitors. From experimental results, it was found that the GZC-12 compound presented an adequate balance between inhibition efficiency and toxicity with maximum efficiency of 98.9% at a concentration of 75 ppm. Accordingly, GZC-12 can be classified as well-suited to human health and the environment and thus as a greener production alternative for oil and gas industry operations. KEYWORDS: Green zwitterionic compound, Green synthesis, Corrosion inhibitor, CO2 corrosion



INTRODUCTION Carbon, mild or low, steels are the most commonly used construction materials for pipelines and equipment in the oil and gas industry. Carbon dioxide (CO2) corrosion of carbon steel pipelines and equipment in the oil and gas industry has been given much attention in recent years because of an increased tendency to inject CO2 into oil wells as an enhanced oil recovery (EOR) process to reduce the viscosity of oil and increase the amount of oil recovered from the mineral formation.1−6 This process starts with the injection of CO2 into the oil producing formation,7 when CO2 contacts oil, © 2017 American Chemical Society

several mechanisms promote the mobilization of oil in immiscible and miscible processes improving the oil recovery.8,9 However, CO2 dissolution in water produces carbonic acid (H2CO3), decreasing the pH of the connate water.10 This acidic solution can cause failure on main downhole tubing, transition pipelines and separation units,11 and disruption of the oil/gas production. The consequences of failures caused by corrosion Received: July 19, 2017 Revised: September 28, 2017 Published: October 5, 2017 10396

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. Schematic representation of the chemical synthesis of GZC compounds.

the environmental impact of the synthetic procedures used to obtain such corrosion inhibitors must be also taken into account due to intensive use of toxic raw materials and solvents, as well as the production of byproducts and waste matter in general and the intensive use of energy.25−30 In this context, the aim of this work was to develop green zwitterionic compounds as environmentally friendly inhibitors for carbon steel corrosion in oilfield environments. This constitutes the first report on the synthesis and evaluation of nontoxic zwitterionic corrosion inhibitors as an alternative to typical amphiphilic and ionic liquid types, that can be obtained through a sustainable procedure. For such an approach, an ecofriendly synthesis route was designed to obtain three 3(alkylamino) propanoic acids in zwitterionic form, GZC-8, GZC-12, and GZC-18 and, also, to characterize them by Fourier transform infrared spectroscopy (FT-IR), nuclear magnetic resonance (NMR), electrospray ionization mass spectrometry (ESI-MS), and elemental analysis. Toxicity analysis for the three compounds was carried out by Microtox. Corrosion inhibition performance was evaluated by means of electrochemical techniques, including open circuit potential (OCP), linear polarization resistance (LRP), electrochemical impedance spectroscopic (EIS) at different immersion times, and potentiodynamic polarization after 24 h of immersion.

can include equipment replacement, production shutdown, environmental damage, and safety risks,12 resulting in severe economic losses.13 Since corrosion is an inevitable deterioration process, adequate control strategies are necessary. The use of organic inhibitors is one of the most economical and practical methods for protection of metal surfaces against corrosion in acid media.14 A large number of organic molecules have been developed and extensively studied as corrosion inhibitors.15 These molecules regularly have amphiphilic structure consisting of a polar headgroup which interacts with metal surface and a nonpolar tail group which provides hydrophobic metalprotecting film. In fact, the adsorption of amphiphilic molecules at the solid−liquid interface plays an important role in many technological and industrial applications, such as detergency, mineral flotation, solid-dispersion, biotechnology, oil recovery, and inhibition corrosion, among others.16,17 For instance, the inhibitory action of amphiphilic molecules in corrosive solutions is often related with its capacity to adhere and form a protective layer on metallic surface. This molecules can be anionic, cationic, or zwitterionic depending on the nature of their head groups.18 Polar headgroup contains heteroatoms such as phosphorus, nitrogen, sulfur, or oxygen and present multiple bonding between them or with carbon atom. These electron rich fragments are the adsorption sites of inhibitor molecule.19 While tail groups are constituted by hydrocarbon chains, frequently of saturated nature with a typical length between C6 and C18. The combination of these two molecular fragments actually regulate some macroscopic properties of the resulting substance such as polarity, solubility, film forming behavior, and toxicity. Thus, the challenge to develop novel and greener corrosion inhibitors is to find equilibrium between competitive technical performance and compatible toxic response of such chemical to human health and environment.20−22 However, despite the efforts, the high toxicity of some of these molecules, including recently developed ionic liquids (IL) such as 1-octyl-3-methylpyridinium bromide, precludes their use as corrosion inhibitors specifically in acidic mediums due to new generation of environmental regulations as their EC50 values are around 1 to 2 ppm even though can reach significant corrosion inhibiting efficiency (>80%).23,24 On the other hand,



EXPERIMENTAL METHODOLOGY

Synthesis of GZC-8, GZC-12, and GZC-18 Compounds. An eco-friendly procedure31 was used to obtain the GZC-8, GZC-12 and GZC-18. The chemical synthesis according to Figure 1 consisted on a one-step, solventless reaction between alkylamines (octylamine, dodecylamine, octadecylamine) with acrylic acid (all reagent grade, Aldrich Chemical Co.) at atmospheric pressure and low temperature of 30 °C. Thus, 0.049 mol of alkylamines were added to a threenecked balloon flask of 50 mL at a temperature of 30 °C under vigorous stirring. Then 0.049 mol of acrylic acid was added slowly. The reactions were exothermic and the temperatures under these conditions increased gradually. The reactions were kept for 30 min and then mixtures were cooled to room temperature obtaining GZC-8, GZC-12, and GZC-18 with good yields (>97%). Chemical Characterization. Spectroscopic methods were employed to characterize the GZC synthesized compounds. IR spectra were recorded in the range of 4000 and 600 cm−1 on a PerkinElmer Frontier FT-IR spectrometer equipped with a diamond attenuated 10397

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering total reflectance (ATR) accessory. NMR was performed on a Bruker Avance III 750 MHz spectrometer with d-chloroform (CDCl3) as solvent. The 1H and 13C chemical shifts (δ) were quoted in ppm and measured relative to the internal SiMe4 (TMS), the coupling constants were quoted in hertz. Multiplicities are shown as the abbreviations: s (singlet), d (doublet), t (triplet), and m (multiplet). The ESI-MS measurement was carried out using a Bruker micrOTOF-Q II (Bruker Daltonics, Bremen, Germany) under a positive ion mode. Elemental analyses were performed on CHNS/O analysis in PerkinElmer Series II 2400. Acute Toxicity by Microtox Test. The acute toxicity method was carried out to analyze the toxicity of three GZC compounds. The toxic effect was determined as the concentration (EC50) of the sample causing a 50% reduction on the light emitted by the bacteria Photobacterium phosphoreum (Vibrio fischeri), after a predetermined exposure time (5 and 15 min) by triplicate. For the tests, a vial with bacteria was reconstituted in 1 mL of solution and maintained at 3 °C in an incubator well, and a serial dilution of the bacterial compounds was prepared in 2% saline water. The luminometer and supporting computer software (MicrotoxOmniR software) with a standard log− linear model were used to determine the EC50 value. All EC50 values were expressed as parts per million (ppm) with 95% confidence intervals and reported as the mean of three pseudoreplicates or true replicates; replicates are a statistical measurement of the test precision. Experimental Setup and Electrochemical Measurements. The experiments were carried out at atmospheric pressure in a 200 mL double-walled cylindrical electrochemical glass cell, at a constant temperature of 70 °C. The experimental setup consisted of a typical three electrode configuration, where a cylindrical electrode was used as working electrode (WE), a graphite bar was used as counter electrode (CE), and a Ag/AgCl electrode connected to the cell via a salt bridge throughout a Luggin capillary with a platinum tip served as reference electrode (RE). The glass cell was filled with 200 mL of a 1.0 M NaCl solution that was purged with CO2 during 30 min prior immersion of the WE (pH = 4.03). After 30 min, the solution pH was constant at 4.03 indicating its saturation with CO2. It is important to point out that the CO2 injection was maintained during the test with the aim of avoiding any oxygen contamination and to keep a CO2 saturated solution. The WE was made of carbon steel grade AISI-1018 (2.827 cm2 exposed area) with the following chemical composition (wt %): 0.16 C, 0.79 Mn, and 99.05 Fe. The electrodes were grounded with siliconcarbide sand paper until a 1200 grit surface finish was obtained, and then were cleaned in an ultrasonic bath, rinsed with acetone, and dried before immersion into the solution. The GZC-8, GZC-12, and GZC18 were prediluted with isopropanol, and then added to the cell at specific concentrations by triplicate. The electrochemical results were obtained with a computer controlled potentiostat/galvanostat Autolab. The corrosion potential (Ecorr) was monitored and recorded throughout the entire experiments. Before any electrochemical measurements, the coupons were precorroded for 30 min to reach steady-state conditions. The GZC solution was added after 1 h. The laboratory methodology of evaluation consisted of the following steps: (i) All three GZC compounds were characterized trough linear polarization resistance (LPR) measurements at 25 ppm and atmospheric pressure in the aforementioned electrochemical glass cell, at a constant temperature of 70 °C. The LPR measurements were taken during 24 h, at ±5 mV vs Ecorr, and a scan rate of 0.125 mV·s−1. (ii) Potentiodynamic polarization was measured after 24 h of immersion. Based on the results of Ecorr at 24 h, corrosion velocities were calculated and consequently corrosion inhibiting efficiencies for GZC-8, GZC-12, and GZC-18. (iii) By taking into account the corrosion inhibiting efficiencies for all three GZC compounds in addition to acute toxicity evaluation results, the best performance compound was selected to a next phase of characterization trough LPR

measurements at various concentrations of 25, 50, and 75 ppm from which it was possible to identify the optimum concentration for selected GZC compound. (iv) Finally, selected GZC compound was evaluated trough electrochemical impedance spectroscopy (EIS) at optimum concentration in order to elucidate a possible inhibiting mechanism. Measurements were carried out at selected immersion times at E = Ecorr under a sinusoidal excitation potential of 10 mV (rms), in the frequency range from 10 kHz to 0.1 Hz. Eight points per decade were recorded. Potentiodynamic sweeps were carried out in cathodic (−300 mV vs Ecorr) and anodic (+300 mV vs Ecorr) direction at scan rate of 0.166 mV·s−1, after 24 h of immersion.



RESULTS AND DISCUSSION Synthesis and Chemical Characterization. GZ-8. A 6.33 g (0.049 mol) portion of octylamine was added to a threenecked balloon flask of 50 mL at a temperature of 30 °C under vigorous stirring. Then 3.53 g (0.049 mol) of acrylic acid was added slowly. The reaction was exothermic, and the temperature under these conditions increased gradually to 70−80 °C. The reaction was kept for 30 min, and then, the mixture was cooled to room temperature obtaining 9.7 g of 3-(octylamino) propanoic acid in zwitterionic form as a white solid with a yield of 98%. FT-IR spectra-absorption bands are observed at 3371 cm−1 (N+−H), 1642 cm−1 (asymmetric N−H bend), 1565 cm−1 (asymmetric −COO− stretch), and 1392 cm−1 (symmetric −COO− stretch). The respective NMR data were the following: 1H NMR (750 MHz, CDCl3, δ ppm); 0.846 (t, J = 6.7 Hz, 3H, H-11), 1.233 (m, 10H, H-6 to H-10), 1,720 (m, 2H, H-5), 2.509 (t, J = 5.6 Hz, 2H, H-2), 2.875 (t, J = 7.9 Hz, 2H, H-4), 3.045 (t, J = 5.56 Hz, 2H, H-3). 13C NMR (187.5 MHz, CDCl3, δ ppm); 176.918 (C-1), 46.954 (C-3), 44.527 (C-4), 31.783 (C-2), 31.665 (C-5), 29.075 (C-6, C-7), 26.721 (C-8), 26.165 (C-9), 22.536 (C-10), 14.024 (C-11). The ESIMS technique was employed to measure the molecular weight of GZC-8. Under the positive ion mode, the cationic [GZC-8 + H]+ (i.e., 1 mol of GZC-8) was binding with 1 mol of protons. The mass-to-charge ratio (m/z) of [GZC-8 + H]+ was 202.30, which matches well with the theoretical calculation result (201.30 + 1) of the molecule-ion weight for [GZC-8 + H]+. Then, the results of elemental analysis for GZC-8 are C11H23NO2 (calc) C 65.97, H 12.22, N 6.95; (found) C 65.62, H 12.18, N 6.93. GCZ-12. A 9.08 g (0.049 mol) portion of dodecylamine were added to a three-necked balloon flask of 50 mL at a temperature of 30 °C under vigorous stirring. Then 3.53 g (0.049 mol) of acrylic acid were added slowly. The reaction was exothermic and the temperature under these conditions increased gradually to 50−60 °C The reaction was kept for 30 min and then mixture was cooled to room temperature obtaining 12.4 g of 3-(dodecylamino) propanoic acid in zwitterionic form as a white solid with a yield of 97%. FT-IR spectra-absorption bands are observed at 3374 cm−1 (N+−H), 1643 cm−1 (asymmetric N−H bend),32 1568 cm−1 (asymmetric −COO − stretch), and 1394 cm −1 (symmetric −COO − stretch).33 The respective NMR data were the following: 1H NMR (750 MHz, CDCl3, δ ppm); 0.876 (t, J = 6.6 Hz, 3H, H15), 1.246 (m, 18H, H-6 to H-14), 1.740 (m, 2H, H-5), 2.519 (t, J = 6.0 Hz, 2H, H-2), 2.905 (t, J = 7.9 Hz, 2H, H-4), 3.059 (t, J = 6.0 Hz, 2H, H-3). 13C NMR (187.5 MHz, CDCl3, δ ppm); 176.904 (C-1), 47.050 (C-3), 44.594 (C-4), 31.901 (C5), 31.693 (C-2), 29.113−29.618 (C-6 to C-11), 26.753 (C12), 26.284 (C-13), 22.678 (C-14), 14.120 (C-15). These 10398

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Figure 2. 13C NMR spectrum of GZC-12.

employed to measure the molecular weight of GZC-18. Under the positive ion mode, the cationic [GZC-18 + H]+ (i.e., 1 mol of GZC-18) was binding with 1 mol of protons. The mass-tocharge ratio (m/z) of [GZC-18 + H]+ was 342.57, which matches well with the theoretical calculation result (341.57 + 1) of the molecule-ion weight for [GZC-18 + H]+. Then, the results of elemental analysis for GZC-18 are C21H43NO2 (calc) C 73.06, H 13.26, N 4.15; (found) C 72.97, H 13.42, N 4.14. Acute Toxicity Results. Microtox bioassay provides a convenient method to evaluate the adverse effect of oilfield chemicals on the aquatic micro-organism.35 Furthermore, the use of Vibrio fischeri based bioluminescence assay for toxicity measurement of polluted water has increased due to good reproducibility and sensibility.36 The light production is directly proportional to the metabolic activity of the bacterial population and any inhibition (such that caused by high levels of toxicity) of enzymatic activity gives a corresponding decrease in bioluminescence.37 Acute toxicity test results for GZC compounds at 5 and 15 min are listed in Table 1. The EC50 results were three and 2 orders of magnitude less toxic than conventional long chain quaternary ammonium salts (QAS) and ionic liquid (IL) amphiphiles. This observed behavior is attributed to the contribution of two main structural features of amphiphilic molecules (i) tail group effect and (ii) head polar group effect, which combined regulate lipophilicity and molecule interaction with biomolecules present in the cell. In the first case, the tail group lipophilicity is related to its ability to overpass the cell barriers, and thus at a fixed polar headgroup toxicity augment as lipophilicity of tail group increase. Even further, in the case of ionic liquids tail group is attached to formal cationic charge which can increase toxicity of such molecules at the level of reported for conventional long chain quaternary ammonium salts.23,38−40

assignments were based on Michael-type addition between alkyl-amine (dodecylamine) and α,β-unsaturated carbonyl compound (acrylic acid). The 13C NMR in Figure 2 shows two new broad signals at 47.050 and 31.693 ppm due to the respective amine quaternization and subsequent α-carbon formation.34 The ESI-MS technique was employed to measure the molecular weight of GZC-12. Under the positive ion mode, the cationic [GZC-12 + H]+ (i.e., 1 mol of GZC-12) was binding with 1 mol of protons. The mass-to-charge ratio (m/z) of [GZC-12 + H]+ was 258.49, which matches well with the theoretical calculation result (257.49 + 1) of the molecule-ion weight for [GZC-12 + H]+. Then, the results of elemental analysis for GZC-12 are C15H31NO2 (calc) C 67.42, H 12.45, N 5.26; (found) C 67.46, H 12.50, N 5.27. GCZ-18. A 13.21 g (0.049 mol) portion of octadecylamine were added to a three-necked balloon flask of 50 mL at a temperature of 30 °C under vigorous stirring. Then 3.53 g (0.049 mol) of acrylic acid were added slowly. The reaction was exothermic and the temperature under these conditions increased gradually to 45−55 °C The reaction was kept for 30 min and then mixture was cooled to room temperature obtaining 16.2 g of 3-(octadecylamino) propanoic acid in zwitterionic form as a white solid with a yield of 97%. FT-IR spectra-absorption-bands are observed at 3369 cm−1 (N+−H), 1645 cm−1 (asymmetric N−H bend), 1566 cm−1 (asymmetric −COO− stretch), and 1394 cm−1 (symmetric −COO− stretch). The respective NMR data were the following: 1H NMR (750 MHz, CDCl3, δ ppm); 0.874 (t, J = 6.5 Hz, 3H, H-21), 1.234 (m, 26H, H-6 to H-20), 1.733 (m, 2H, H-5), 2.518 (t, J = 6.1 Hz, 2H, H-2), 2.921 (t, J = 7.9 Hz, 2H, H-4), 3.065 (t, J = 6.2 Hz, 2H, H-3). 13C NMR (187.5 MHz, CDCl3, δ ppm); 176.918 (C-1), 47.113 (C-3), 44.534 (C-4), 31.921 (C-2), 31.721 (C5), 29.125−29.638 (C-6 to C-17), 26.728 (C-18), 26.218 (C19), 22.633 (C-20), 14.112 (C-21). The ESI-MS technique was 10399

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Since iron is the main component of carbon steels, the anodic reaction comprises the oxidation of iron into ferrous ions, eq 7:

Table 1. Acute Toxicity as EC50 for GZC Compounds and Selected Amphiphilic Chemicals chemical GZC-8 GZC-12 GZC-18 1-octyl-3-methylimidazolium bromide 1-octyl-3-methylpyridinium bromide Dodecyltrimethylammonium bromide Hexadecyltrimethylammonium bromide

EC50 5 min (ppm) 206 171 41

EC50 15 min (ppm)

Fe(s) + 2e− ← → Fe 2 +(aq)

209 188 43 1.1723

Numerous studies have been devoted to the study of CO2corrosion-electrochemistry, suggesting a multistep mechanism to explain the Fe dissolution process.10,48−53 These studies suggest that at pH ≤ 4 the reduction of H+ (eq 4) is one of the key cathodic processes. CO2 corrosion of carbon steels is usually accompanied by the formation of protective or nonprotective corrosion products. At high temperature and pressure and pH higher than 6, the formation of protective iron carbonate (FeCO3) is favored.54−57 At pH lower than 6, saturation of FeCO3 is more difficult achieved, rather a porous iron carbide layer (Fe3C) has been reported on the surface of steel samples exposed to CO2 corrosion.58−60 Fe3C is part of the steel makeup, being conductive and nobler than the ferrite phase (α-Fe). Therefore, reduction of hydrogen ions (eq 5) will occur on Fe3C sites, while the dissolution of the ferrite phase (eq 7) will take place, leaving behind a porous Fe3C layer. Figure 3 shows the effect of GZC compounds at 25 ppm concentration on the polarization curves for an AISI-1018 C-

1.7723 0.2438,47 0.6338,47

On the other hand, at a fixed tail group length, toxicity depends on the nature of headgroup interactions with biomolecules within the cell, mainly through two types of interaction: (1) disruption of cellular membranes by interaction with lipid components and (2) reaction of amphiphilic molecules with proteins essential to the cell-metabolism. Anionic surfactants have a clear biological activity due to their binding to bioactive macromolecules such as starch, peptides, enzymes, and DNA. The binding to proteins and peptides may change the folding of the polypeptide chain and the surface charge of a molecule.41 The most common cationictype surfactants are the quaternary ammonium compounds; these molecules can bind to inner membranes and disorganize them via their long alkyl chain.42 Thus, surfactant interaction with cellular membranes depends on chemical structure, charge, critical micellar concentration, and hydrophile−lipophile balance.43,44 In general, the cationic nature of molecules is associated with cytotoxic effects.45,46 This is why conventional QAS present EC50 values less than 1 ppm38,47 and ILs, such as 1-octyl-3methylimidazolium bromide and 1-octyl-3-methylpyridinium bromide, reach EC50 of 1.17 and 1.77 ppm, respectively.23 Effect of GZC Compounds on the CO2-CorrosionElectrochemistry. When water and CO2 are present, a small fraction of CO2 is hydrated, forming carbonic acid (H2CO3) according to eqs 1 and 2. H2CO3 will further dissociate into bicarbonate and hydrogen ions (eq 3). Finally, bicarbonate ions are dissociated into carbonate and hydrogen ions (eq 4): CO2(g) ← → CO2(aq)

(1)

CO2(aq) + H 2O(l) ← → H 2CO3(aq)

(2)

H 2CO3(aq) ← → HCO3−(aq) + H+(aq)

(3)

HCO3−(aq) ← → CO32 −(aq) + H+(aq)

(4)

Figure 3. Polarization curves for AISI-1018 C−steel electrode in 1.0 M NaCl solution saturated with CO2 at 70 °C, after 24 h of immersion, for GZC compounds at 25 ppm.

steel electrode in 1.0 M NaCl solution saturated with CO2 at 70 °C, measured after 24 h. The respective electrochemical parameters obtained by Tafel extrapolation are summarized in the Table 2. It can be observed that the addition of GZC compounds inhibited both anodic and cathodic reactions by decreasing the current densities and shifting the Ecorr toward more positive values. This suggests that GZC compounds acts as mixed-type corrosion-inhibitors. The corrosion rate (CR) values were calculated from icorr according to Faraday’s law.61 Whereas, icorr was estimated from polarization measures (showed in Table 2) based on the Tafel extrapolation method. Moreover, the percentage of inhibition efficiency (% IE) was calculated according to eq 8 from CR values for GZC compounds.

The electrochemical process of CO2 corrosion of carbon steels involves the reduction of hydrogen and the anodic dissolution of iron.48 Hydrogen ions from eqs 3 and 4 will be reduced at cathodic sites on the metal surface (eq 5). Water reduction (eq 6) is possible but only at very negative potentials: H+(aq) + e− ← → 1/2 H 2(g)

(5)

H 2O(l) + e− ← → OH−(aq) + 1/2 H 2(g)

(6)

(7)

10400

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Table 2. Electrochemical Parameters from Polarization Measurements for AISI-1018 C−Steel Electrode in a 1.0 M NaCl Solution Saturated with CO2 at 70 °C for GZC Compounds at 25 ppm, after 24 h Immersion compound

Ecorr V/(Ag/AgCl)

βa (mV·dec−1)

−βc (mV·dec−1)

ref GZC-8 GZC-12 GZC-18

−0.665 −0.625 −0.627 −0.647

68 65 68 63

136 274 226 152

icorr (A·cm−2) 2.76 1.06 3.51 2.72

× × × ×

10−4 10−4 10−5 10−5

CR (mmpy)

% IE

3.206 1.235 0.408 0.316

61.47 87.28 90.14

Table 3. Electrochemical Parameters from Polarization Measurements for AISI-1018 C−Steel Electrode in a 1.0 M NaCl Solution Saturated with CO2 at 70 °C for GZC-12 at Different Concentrations, after 24 h Immersion concentration

Ecorr V/(Ag/AgCl)

βa (mV·dec−1)

−βc (mV·dec−1)

icorr (A·cm−2)

CR (mmpy)

% IE

25 50 75

−0.627 −0.607 −0.597

68 87 69

226 216 212

3.51 × 10−5 3.96 × 10−6 2.95 × 10−6

0.408 0.046 0.034

87.28 98.57 98.93

%IE =

CR o − CR GZC × 100 CR o

(8)

Where, CRo and CRGZC are the respective corrosion rates without and with GZC compounds. The calculated CR without GZC was ca. 3.21 mmpy after 24 h of immersion. As can be observed from Table 2, GZC-18 exhibited the highest inhibition efficiency of 90.14%, whereas GZC-12 presented an efficiency of 87.28% and 61.47% for GZC-8. These results are in accordance to the lipophilic effect or contribution, i.e. a long hydrophobic substituent on an amphiphilic molecule augments corrosion protection in comparison to a molecule with smaller hydrophobic substituent. However, if acute toxicity results are taken into account it can be observed that GZC-18 does not result the best option to application given its EC50 value of 43 ppm, resulting four times more toxic compared to GZC-12 or GZC8. Given these results GZC-12 was selected as the best alternative to develop a green corrosion inhibitor as it represented a balance between high corrosion inhibiting efficiency and lower toxicity. Thus, it was considered to continue to further electrochemical evaluation. LPR for GZC-12 at Different Concentrations. Results for LPR of GZC-12 at different concentrations are presented in Table 3, where it can be observed that the maximum efficiency of 98.93%, was obtained at 75 ppm. In addition, Figure 4 shows the polarization curves of steel electrode from where it is observed that GZC-12 inhibit both anodic and cathodic reactions by decreasing the current densities and shifting the Ecorr toward more positive values and maximum effect was also measured at 75 ppm. Consequently, GZC-12 acts as a mixedtype corrosion-inhibitor due to its interaction with both anionic and cationic sites, probably due to its particular electrical charge distribution given its zwitterion nature. Figure 5a shows the change of the electrode potential (AISI1018 C−steel electrode) in a 1.0 M NaCl solution saturated with CO2 at 70 °C. In the absence of GZC-12 (reference), the corrosion potential (Ecorr) showed just small increase with time. In contrast, the addition of GZC-12 caused a shift of Ecorr to more positive values, a behavior that was enhanced with an increase of GZC-12 concentration. The change of polarization resistance (Rp) measured for 24 h is shown in Figure 5b. The corrosion rate is inversely proportional to the Rp, eq 9:

Figure 4. Polarization curves for AISI-1018 C−steel electrode in 1.0 M NaCl solution saturated with CO2 at 70 °C, after 24 h of immersion, for GZC-12 at different concentrations.

icorr =

B Rp

(9)

where icorr is the corrosion current density (A·cm−2), Rp is the polarization resistance (Ω·cm2), and B is a constant. It is possible to calculate B from the Tafel slopes according to eq 10:

B=

βa βc 2.303(βa + βc)

(10)

where ba and bc are the anodic and cathodic Tafel slopes, respectively. It is clear from Figure 5b that the addition of GZC-12 sharply increased the Rp values with an increase in concentration, so that the maximum effect was observed at 75 ppm. As observed in Figure 5, adsorption of GZC-12 onto the metal surface changed the Ecorr toward more noble values and increased the polarization resistance. This could be explained as follows. Corroded surfaces tend to spontaneously develop a surface charge, resulting in an electrical double-layer interaction with charged organic-ionic molecules (e.g., GZC). It has been shown that a carbon steel surface subjected to CO2 corrosion possess a positive charge.62 When the surface bears an opposite charge to the ionic molecules, adsorption tends to be maximized.63,64 The inhibition effect is attributed to the adsorption of the organic−ionic molecules via their functional ionic groups and 10401

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. Electrochemical performance during immersion of AISI-1018 C−steel electrode in 1.0 M NaCl solution saturated with CO2 at 70 °C and different GZC-12 concentrations: (a) Ecorr transient and (b) Rp variation.

Figure 6. Nyquist spectra of the AISI-1018 C−steel electrode immersed in 1.0 M NaCl solution saturated with CO2 at 70 °C and 75 ppm of GZC12: (a) evolution of the impedance response with time and (b) close-up at low frequencies (0.59−0.29 Hz) showing a 45° angle characteristic of a Warburg impedance.

performed on AISI-1018 C−steel samples exposed to a 1.0 M NaCl−CO2 solution at 70 °C. Nyquist and Bode plots at different immersion times are shown in Figures 6 and 7, respectively. In the absence of GZC-12 (at 1 h of immersion) the impedance response is characterized by one depressed capacitive semicircle at medium frequencies (MF) and an inductive loop at low frequencies (LF), see inset in Figure 6a. The capacitive response is associated with the capacitance of the double layer (Cdl) and the charge transfer process (Rct). The observed depressed semicircle, with the center under the real axis, is a typical behavior for solid-metal electrodes with frequency dispersion of the impedance data due to roughness and other inhomogeneities on the sample surface. At low frequencies, the inductive loop has been related to the adsorption of intermediate products during iron dissolution.65,66 The Bode representation of the impedance response for 1 h of immersion shows on the phase angle one maxima at MF (59 Hz) corresponding to the capacitive loop and a minimum at LF (0.3 Hz) corresponding to the inductive loop

their ability to aggregate and form a protective layer. The results shown in Figure 5b indicate that the addition of GZC-12 increases the interfacial resistance (Rp) magnitude with respect to the reference (no GZC-12). The Rp magnitude for the reference solution slightly decreased with time, showing the nonprotective nature of the corrosion products. Once the GZC-12 was added, the Rp magnitude increased with time, an indication of the protective nature of the film-formed GZC-12 compound. This observed behavior could be ascribed to the molecular structure of GZC-12 which is composed of a quaternary-ammonium group (N+), a C-12 saturated hydrophobic tail, and a hydrophilic carboxylate groups (COO−). It is well-known that the hydrophilic group of zwitterion compounds strongly prefers interaction with polar entities such as water, metals, and ions.17,63 Thus, these hydrophilic groups are adsorbed on the active sites of the metal surface.18 EIS Results for GZC-12 at 75 ppm. In order to get more insight about the adsorbed GZC-12 properties and its role on the corrosion inhibition process, EIS studies for 75 ppm were 10402

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

Figure 7. Bode spectra recorded for AISI-1018 C−steel electrode immersed in 1.0 M NaCl solution saturated with CO2 at 70 °C: (a) phase angle plot and (b) module plot.

Figure 8. Equivalent circuits used for fitting the experimental EIS results for GZC-12.

shown in the Nyquist plot, suggests a continuous displacement of water molecules and Cl−, that were adsorbed on the steel surface, by GZC molecules. The appearance of second time constant at high frequencies (1000 Hz) on the phase angle Body plot corroborates the presence of an inhibitor layer. This impedance trend was observed for 12 h of exposure. After 12 h of immersion a change in the corrosion mechanism was observed by the appearance of well-defined third time constant at LF (∼0.41 Hz). For example, the Nyquist plots in Figure 6b show Warburg-type impedance as can be seen indicating that corrosion processes are under diffusion control.68,69 The impedance response for up to 12 h of immersion was simulated using the EC shown in Figure 8b, where Rs is the electrolyte resistance, YGZC is a CPE representing the capacitance of the GZC-12-film, RGZC is the pore resistance of the outer layer of the GZC-12-film, Ydl is a CPE representing the double-layer capacitance, and Rct is the charge transfer resistance. The effect of diffusion on the impedance response, after 12 h of immersion was simulated with the EC shown in Figure 8c. Where ZW is the finite length Warburg impedance, representing the diffusion of electroactive species. ZW is defined by eq 13.70

observed on the Nyquist plot. The EIS data for the reference test were fitted with the equivalent circuit (EC) shown Figure 8a, where Rs is the electrolyte resistance, Ydl is a constant phase element (CPE) representing the double-layer capacitance, Rct is the charge transfer resistance, RL is the inductive resistance, and L is the inductance. As it is clear that a nonideal frequency response is observed, a CPE was used. The CPE is defined in eq 11 as67 ZCPE = Y o−1(jω)−n

(11)

In this equation, ZCPE is the impedance of the CPE, Yo is a factor proportional to the capacitance, j is −1 , and ω is the angular frequency. The depletion degree of the impedance loops depends on the value of n which can take values from 0 to 1. For example, for n = 0, CPE represents a resistance (ZCPE = R); whereas for n = 1, a capacitance is obtained (ZCPE = C). On the other hand, for n = 0.5, CPE represents a Warburg impedance (ZCPE = W); and for n = −1, an inductance (ZCPE = L). From the CPE (Yo) and Ri, the capacitance (Ci) can be calculated with eq 12. Ci = (YoiR i1 − ni)1/ ni

(12)

Z W = R W(TWiω)−n tanh[(TWiω)−n ]

Addition of 75 ppm of GZC-12 clearly influenced the impedance response of the system as it can be observed in Figure 6. A gradual increase of the capacitive loop with time,

(13)

Where RW represents the diffusion resistance, the exponent n is 0.5, and TW is defined as (eq 14): 10403

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

0.5 0.5 0.5 0.5

n (0−1)

4.7 9.4 26.8 34.5 40 1.7 4.7 5.3 5.7

W−T CGZC (μF·cm )

14.11 15.80 22.60 24.33 28.44 2.52 7.63 7.33 7.50 0.96 0.88 0.79 0.75 0.73 0.79 0.81 0.83 0.83

n (0−1) YGDZ (Ω ·s ·cm ) n (0−1)

10−5 10−5 10−4 10−4 10−4 10−5 10−5 10−5 10−5 × × × × × × × × × 2.07 4.55 1.07 1.43 1.77 3.37 5.24 4.12 4.14 225.61 157.38 141.72 79.76 50.92 22.979 160.72 106.87 100.21 84.38 0.79 0.72 0.69 0.66 0.63 0.64 0.69 0.66 0.67 0.68 10−4 10−4 10−4 10−4 10−5 10−5 10−4 10−4 10−4 10−4

Ydl (Ω ·s ·cm )

× × × × × × × × × ×

Rct (Ω·cm2)

−2

Cdl (μF·cm )

125 3130 3706 4258 4848 4966 4055 3058 2621 2271

CPEGZC

−1 n

−2

RGZC (Ω·cm2)

WGZC

−2

RW (Ω·cm2) 10404

−2

Rs (Ω·cm )

3.1 2.9 2.8 2.7 2.6 2.6 2.1 2.4 2.5 2.5

time (h)

1 3 4 6 8 9 12 16 20 23

CONCLUSIONS Three green zwitterionic compounds were synthesized through an ecofriendly, zero-waste route, i.e., in the absence of solvents under ambient conditions of pressure and temperature. Evaluation of corrosion inhibition performance of three GZC compounds by electrochemical techniques showed that they modify instantaneously the electrochemical corrosion behavior giving more positive potentials and enhancing the corrosion resistance starting from 25 ppm. In this context, it was found that GZC compounds acts as a mixed-type corrosion inhibitor by being adsorbed at the carbon steel surface due to their dual electrical charge distribution given their zwitterionic nature, forming a resistant protective layer due to the alkyl nonpolar tail group. GZC-18 exhibited the highest inhibition efficiency of 90.14%, whereas GZC-12 presented an efficiency of 87.28%, and 61.47% was shown for GZC-8. These results are in accordance to the lipophilic effect or contribution, i.e. a long hydrophobic substituent on an amphiphilic molecule augments corrosion protection in comparison to a molecule with smaller hydrophobic substituents. In addition, the EC50 results from the Microtox test were 209, 188, and 43 ppm for GZC-8, GZC-12, and GZC-18, respectively, showing that these molecules are 2 orders of magnitude less toxic than recently developed IL (ionic liquids) corrosion inhibitors and, thus, compatible with human health and the environment. This observed behavior might be linked to the effect of the particular formal charge distribution in the zwitterionic polar headgroup in GZC. From these results, GZC-12 was selected as the best option to develop a green corrosion inhibitor as it represented a balance between high corrosion inhibiting efficiency and lower

−1 n

Table 4. EIS Parameters Obtained by Fitting the Experimental Results Using the ECs Shown in Figure 8 for 75 ppm of GZC-12



4.77 1.92 1.73 1.15 8.54 5.02 1.84 1.56 1.56 1.43

Where T is the finite length diffusion element exponent (W− T), L the effective diffusion length or thickness, and D is the diffusion coefficient of the reactive species. The adsorption of inhibitor molecules on the metal surface was corroborate by a decrease of the Cdl with time (Table 4). The decrease of Cdl has been correlated to a decrease of the dielectric constant and to an increase of electrical double layer thickness due to the presence of an adsorbed inhibitor film.71−74 The increase on the RGZC means the formation of a barrier for the ions to diffuse and the increase of the RW and TW suggest the formation of a more compact inhibitor film. The described impedance response explains the high inhibition efficiency at 75 ppm. The kinetics of the corrosion processes was directly affected as it can be seen by an increase of the Rct which is inversely proportional to the corrosion rate. It is important to point out that the capacitance (CGZC) related to the inhibitor film slightly increased with time (1−12 h). This behavior could be related to dual charge nature of the GZC-12 molecule. Usually the inhibitor action of organic molecules is due to its adsorption on anodic and cathodic sites present over metallic surface. However, it is well-known that the inhibiting action of ionic corrosion inhibitors (e.g., anionic, cationic, or zwitterion compounds) involves blocking of anodic−cathodic metallic sites as well as the interaction with aggressive ions (e.g., H+, HO−, HS−,Cl−,SO4−2, HCO3−, NO3−), resulting in the formation of complex charged layers,17 and because of that, CGZC could be slightly increased.

0.65 0.84 1.17 1.49

(14)

CPEdl

L2 D

2

TW =

1370 2755 3320 3863

ACS Sustainable Chemistry & Engineering

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering

(11) Tawancy, H. M.; Al-Hadhrami, L. M.; Al-Yousef, F. K. Analysis of corroded elbow section of carbon steel piping system of an oil-gas separator vessel. Case Stud. Eng. Fail. Anal. 2013, 1, 6−14. (12) Sun, C.; et al. Investigate deepwater pipeline oil spill emergency repair methods. Aquat. Procedia 2015, 3, 191−196. (13) Koch, G.; Brongers, M. P. H.; Thompson, N. G.; Virmani, P.; Payer, J. H. Corrosion costs and preventive strategies in the United States. NACE Int. 2002, 1−12. (14) Fink, J. K. Oil field chemicals; Gulf Professional Publishing, 2015. (15) Arora, A.; Pandey, S. K. Review on materials for corrosion prevention in oil industry. SPE Int. Conf. Exhib. Oilfield Corr. 2012, 1− 11. (16) Vittal, R.; Gomathi, H.; Kim, K. J. Beneficial role of surfactants in electrochemistry and in the modification of electrodes. Adv. Colloid Interface Sci. 2006, 119, 55−68. (17) Verma, C.; Ebenso, E. E.; Quraishi, M. A. Ionic liquids as green and sustainable corrosion inhibitors for metals and alloys: An overview. J. Mol. Liq. 2017, 233, 403−414. (18) Malik, M. A.; Hashim, M. A.; Nabi, F.; AL-Thabaiti, S. A.; Khan, Z. Anti-corrosion ability of surfactants: A review. Int. J. Electrochem. Sci. 2011, 6, 1927−1948. (19) Finšgar, M.; Jackson, J. Application of corrosion inhibitors for steels in acidic media for the oil and gas industry: A review. Corros. Sci. 2014, 86, 17−41. (20) Asipita, S. A.; et al. Green Bambusa Arundinacea leaves extract as a sustainable corrosion inhibitor in steel reinforced concrete. J. Cleaner Prod. 2014, 67, 139−146. (21) Shubina, V.; Gaillet, L.; Chaussadent, T.; Meylheuc, T.; Creus, J. Biomolecules as a sustainable protection against corrosion of reinforced carbon steel in concrete. J. Cleaner Prod. 2016, 112, 666− 671. (22) Qiang, Y.; et al. Sodium dodecyl benzene sulfonate as a sustainable inhibitor for zinc corrosion in 26% NH4Cl solution. J. Cleaner Prod. 2017, 152, 17−25. (23) Docherty, K. M.; Kulpa, C. F., Jr. Toxicity and antimicrobial activity of imidazolium and pyridinium ionic liquids. Green Chem. 2005, 7, 185. (24) Zheng, X.; Zhang, S.; Li, W.; Gong, M.; Yin, L. Experimental and theoretical studies of two imidazolium-based ionic liquids as inhibitors for mild steel in sulfuric acid solution. Corros. Sci. 2015, 95, 168−179. (25) Sasson, Y.; Neumann, R. Handbook of Phase Transfer Catalysis 1997, DOI: 10.1007/978-94-009-0023-3. (26) Vanyúr, R.; Biczók, L.; Miskolczy, Z. Micelle formation of 1alkyl-3-methylimidazolium bromide ionic liquids in aqueous solution. Colloids Surf., A 2007, 299, 256−261. (27) Desimone, M. P.; Grundmeier, G.; Gordillo, G.; Simison, S. N. Amphiphilic amido-amine as an effective corrosion inhibitor for mild steel exposed to CO2 saturated solution: Polarization, EIS and PMIRRAS studies. Electrochim. Acta 2011, 56, 2990−2998. (28) Jawich, M. W. S.; Oweimreen, G. A.; Ali, S. A. Heptadecyl-tailed mono- and bis-imidazolines: A study of the newly synthesized compounds on the inhibition of mild steel corrosion in a carbon dioxide-saturated saline medium. Corros. Sci. 2012, 65, 104−112. (29) Yoo, S. H.; Kim, Y. W.; Chung, K.; Baik, S. Y.; Kim, J. S. Synthesis and corrosion inhibition behavior of imidazoline derivatives based on vegetable oil. Corros. Sci. 2012, 59, 42−54. (30) Pasasa, N. V. A.; Bundjali, B.; Wahyuningrum, D. The microwave assisted synthesis of 1-alkyl-3-methylimidazolium bromide as potential corrosion inhibitor toward carbon steel in 1 M HCl solution saturated with carbon dioxide. AIP Conf. Proc. 2014, 070016. (31) Hernández-Altamirano, R.; Zamudio-Rivera, L. S.; MenaCervantes, V. Y.; Beltrán-Conde, H. I.; Dominguez-Aguilar, M. A.; Martínez-Viramontes, J.; Estrada-Buendia, A. Amino and imino propionic acids, process of preparation and use. US Patent 20150112096, 2015). (32) Stuart, B. H. Infrared spectroscopy: fundamentals and applications 2004, DOI: 10.1002/0470011149.

toxicity. Further electrochemical evaluation showed that GZC12 exhibited excellent inhibition performance for a CO2 acidic environment, with a maximum efficiency of 98.93% at a concentration of 75 ppm. From EIS analysis, it was also found that after 12 h of exposure, GZC-12 affected both anodic and cathodic reactions as an indication of a change on the corrosion mechanism mode from charge transfer to diffusion control. This is the first report on the synthesis and evaluation of nontoxic zwitterionic corrosion inhibitors as an alternative to typical amphiphilic and ionic liquid types, that can be obtained through a sustainable procedure. The results from this investigation open a gate to the application of GCZ compounds and possible analogues as green corrosion inhibitors in different industries where these kinds of corrosive environments are found.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Telephone: +52 55 57296000, ext 52621. ORCID

Arturo Manzo-Robledo: 0000-0002-8570-4028 David A. Nieto-Á lvarez: 0000-0001-8500-1800 Raúl Hernández-Altamirano: 0000-0001-6685-2335 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was financially supported by FONDO DE INVESTIGACIÓ N PARA LA EDUCACIÓ N-CONACYT under the project No. 241262, SIP-IPN multidisciplinary project No. 2017-1864. We give special acknowledgment to Laboratorio Nacional de Desarrollo y Aseguramiento de la Calidad de Biocombustibles (LaNDACBio), CONACYT 280489, for supporting experimental measurements.



REFERENCES

(1) Green, D. W.; Willhite, G. P. Enhanced oil recovery; SPE Textbook Series; Society of Petroluem Engineers, 1988. (2) Alvarado, V.; Manrique, E. Enhanced oil recovery: An update review. Energies 2010, 3, 1529−1575. (3) Nuhi, M.; Abu Seer, T.; Al Tamimi, A. M.; Modarres, M.; Seibi, A. Reliability analysis for degradation effects of pitting corrosion in carbon steel pipes. Procedia Eng. 2011, 10, 1930−1935. (4) Koottungal, L. 2012 worldwide EOR survey. Oil Gas J. 2012, 110, 57−69. (5) Ilman, M. N.; Kusmono. Analysis of internal corrosion in subsea oil pipeline. Case Stud. Eng. Fail. Anal. 2014, 2, 1−8. (6) Yevtushenko, O.; et al. Corrosion behavior of steels for CO2 injection. Process Saf. Environ. Prot. 2014, 92, 108−118. (7) Chukwudeme, E. A.; Hamouda, A. A. Enhanced oil recovery (EOR) by miscible CO2 and water flooding of asphaltenic and nonasphaltenic oils. Energies 2009, 2, 714−737. (8) Pini, R.; Krevor, S.; Krause, M.; Benson, S. Capillary heterogeneity in sandstone rocks during CO2/water core-flooding experiments. Energy Procedia 2013, 37, 5473−5479. (9) Cerasi, P.; Kjøller, C.; Sigalas, L.; Bhuiyan, H.; Frykman, P. Mechanical effect of CO2 flooding of a sandstone specimen. Energy Procedia 2016, 86, 361−370. (10) Choi, Y. S.; Nesic, S. Determining the corrosive potential of CO2 transport pipeline in high pCO2-water environments. Int. J. Greenhouse Gas Control 2011, 5, 788−797. 10405

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406

Research Article

ACS Sustainable Chemistry & Engineering (33) Silverstein, M. R.; Webster, F. X.; Kiemle, D. J. Spectrometric identification of organic compounds; John Wiley & Sons, Inc., 2005. (34) Pretsch, E.; Bü hlmannn, P.; Badertscher, M. Structure determination of organic compounds 2009, DOI: 10.1007/978-3-54093810-1. (35) Grigson, S.; Cheong, C.; Way, E. Studies of produced water toxicity using luminescent marine bacteria. WIT Transactions on Biomedicine and Health 2006, 10, 111−121. (36) Parvez, S.; Venkataraman, C.; Mukherji, S. Toxicity assessment of organic pollutants: Reliability of bioluminescence inhibition assay and univariate QSAR models using freshly prepared Vibrio fischeri. Toxicol. In Vitro 2008, 22, 1806−1813. (37) Parvez, S.; Venkataraman, C.; Mukherji, S. A review on advantages of implementing luminescence inhibition test (Vibrio fischeri) for acute toxicity prediction of chemicals. Environ. Int. 2006, 32, 265−268. (38) Morán, M. C.; et al. ‘Green’ amino acid-based surfactant. Green Chem. 2004, 6, 233−240. (39) Ranke, J.; et al. Biological effects of imidazolium ionic liquids with varying chain lengths in acute Vibrio fischeri and WST-1 cell viability assays. Ecotoxicol. Environ. Saf. 2004, 58, 396−404. (40) Cho, C. W.; et al. Toxicity of imidazolium salt with anion bromide to a phytoplankton Selenastrum capricornutum: Effect of alkyl-chain length. Chemosphere 2007, 69, 1003−1007. (41) Cserhati, T.; Forgacs, E.; Oros, G. Biological activity and environmental impact of anionic surfactants. Environ. Int. 2002, 28, 337−348. (42) Zhang, C.; et al. Quaternary ammonium compounds (QACs): A review on occurrence, fate and toxicity in the environment. Sci. Total Environ. 2015, 518−519, 352−362. (43) Ying, G. G. Fate, behavior and effects of surfactants and their degradation products in the environment. Environ. Int. 2006, 32, 417− 431. (44) Masakorala, K.; Turner, A.; Brown, M. T. Toxicity of synthetic surfactants to the marine macroalga, Ulva lactuca. Water, Air, Soil Pollut. 2011, 218, 283−291. (45) Lv, H.; Zhang, S.; Wang, B.; Cui, S.; Yan, J. Toxicity of cationic lipids and cationic polymers in gene delivery. J. Controlled Release 2006, 114, 100−109. (46) Kim, Y.; Binauld, S.; Stenzel, M. H. Zwitterionic guanidinebased oligomers mimicking cell-penetrating peptides as a nontoxic alternative to cationic polymers to enhance the cellular uptake of micelles. Biomacromolecules 2012, 13, 3418−3426. (47) Ribo, J. M.; Kaiser, L. E. Photobacterium phosphoreum toxicity bioassay. I. Test procedures and applications. Toxic. Assess. 1987, 2, 305−323. (48) Nesic, S.; Nordsveen, M.; Nyborg, R.; Stangeland, A. A mechanistic model for carbon dioxide corrosion of mild steel in the presence of protective iron carbonate filmspart 2: A numerical experiment. Corrosion 2003, 59, 489−497. (49) Nešić, S. Key issues related to modelling of internal corrosion of oil and gas pipelines − A review. Corros. Sci. 2007, 49, 4308−4338. (50) Obuka.; et al. Review of corrosion kinetics and thermodynamics of CO2 and H2S corrosion effects and associated prediction/ evaluation on oil and gas pipeline system. Int. J. Sci. Technol. Res. 2012, 1, 156−162. (51) Tanupabrungsun, T.; Young, D.; Brown, B.; Nešić, S. Construction and verification of pourbaix diagrams for CO2 corrosion of mild steel valid up to 250 °C. Corrosion NACE conference & expo 16, paper no. 0001418, 2012. (52) Dugstad, A.Fundamental aspects of CO2 metal loss corrosion. Part I: mechanism. Corrosion NACE conference & expo; 2006; pp 1−10. (53) Khan, M. I.; Yasmin, T. Erosion-corrosion of low carbon (AISI 1008 Steel) ring gasket under dynamic high pressure CO2 environment. J. Fail. Anal. Prev. 2014, 14, 537−548. (54) Tanupabrungsun, T.; Brown, B.; Nesic, S. Effect of pH on CO2 Corrosion of Mild Steel at Elevated Temperatures. Corrosion NACE conference & expo; 2013; pp 1−11.

(55) Hassani, S.; et al. Wellbore integrinanoty and corrosion of low alloy and stainless steels in high pressure CO2 geologic storage environments: An experimental study. Int. J. Greenhouse Gas Control 2014, 23, 30−43. (56) Tavares, L. M.; Costa, E. M.; Andrade, J. J. D. O.; Hubler, R.; Huet, B. Effect of calcium carbonate on low carbon steel corrosion behavior in saline CO2 high pressure environments. Appl. Surf. Sci. 2015, 359, 143−152. (57) Navabzadeh Esmaeely, S. N.; Choi, Y.-S.; Young, D.; Nešic, S. Effect of Calcium on the Formation and Protectiveness of Iron Carbonate Layer in CO2 Corrosion. Corrosion 2013, 69, 912−920. (58) Crolet, J. L.; Thevenot, N.; Nesic, S. Role of conductive corrosion products in the protectiveness of corrosion layers. Corrosion 1998, 54, 194−203. (59) Eliyan, F. F.; Alfantazi, A. On the theory of CO2 corrosion reactions − Investigating their interrelation with the corrosion products and API-X100 steel microstructure. Corros. Sci. 2014, 85, 380−393. (60) Ochoa, N.; Vega, C.; Pébère, N.; Lacaze, J.; Brito, J. L. CO2 corrosion resistance of carbon steel in relation with microstructure changes. Mater. Chem. Phys. 2015, 156, 198−205. (61) ASTM G 102-89. Standard practice for calculation of corrosion rates and related information; ASTM, 2010; pp 1−7. (62) Belarbi, Z.; Farelas, F.; Singer, M.; Nesic, S. Role of amine in the mitigation of CO2 top of the line corrosion. Corrosion 2016, 72, 1300−1310. (63) Paria, S.; Khilar, K. C. A review on experimental studies of surfactant adsorption at the hydrophilic solid - water interface. Adv. Colloid Interface Sci. 2004, 110, 75−95. (64) Bockris, J. O. M.; Reddy, A. K. N.; Gamboa, M. Modern electrochemistry 2A: Fundamentals of electrodics 2000, DOI: 10.1007/ b113922. (65) Keddam, M.; Mattos, O. R.; Takenout, H. Reaction Model for Iron Dissolution Studied by Electrode Impedance. J. Electrochem. Soc. 1981, 128, 257. (66) das Chagas Almeida, T.; Bandeira, M. C. E.; Moreira, R. M.; Mattos, O. R. New Insights on the role of CO2 in the mechanism of carbon steel corrosion. Corros. Sci. 2017, 120, 239−250. (67) Liu, C.; Bi, Q.; Leyland, A.; Matthews, A. An electrochemical impedance spectroscopy study of the corrosion behaviour of PVD coated steels in 0.5 N NaCl aqueous solution: Part I. Establishment of equivalent circuits for EIS data modelling. Corros. Sci. 2003, 45, 1243− 1256. (68) Chen, Y.; Hong, T.; Gopal, M.; Jepson, W. P. EIS studies of a corrosion inhibitor behavior under multiphase flow conditions. Corros. Sci. 2000, 42, 979−990. (69) Heydari, M.; Javidi, M. Corrosion inhibition and adsorption behaviour of an amido-imidazoline derivative on API 5L X52 steel in CO2-saturated solution and synergistic effect of iodide ions. Corros. Sci. 2012, 61, 148−155. (70) Finšgar, M. 2-Mercaptobenzimidazole as a copper corrosion inhibitor: Part I. Long-term immersion, 3D-profilometry, and electrochemistry. Corros. Sci. 2013, 72, 82−89. (71) Liu, X.; Chen, S.; Ma, H.; Liu, G.; Shen, L. Protection of iron corrosion by stearic acid and stearic imidazoline self-assembled monolayers. Appl. Surf. Sci. 2006, 253, 814−820. (72) Farelas, F.; Ramirez, A. Carbon dioxide corrosion inhibition of carbon steels through bis-imidazoline and imidazoline compounds studied by EIS. Int. J. Electrochem. Sci. 2010, 5, 797−814. (73) Zhao, I.; Chen, G. The synergistic inhibition effect of oleicbased imidazoline and sodium benzoate on mild steel corrosion in a CO2-saturated brine solution. Electrochim. Acta 2012, 69, 247−255. (74) Singh, A.; et al. Gingko biloba fruit extract as an eco-friendly corrosion inhibitor for J55 steel in CO2 saturated 3.5% NaCl solution. J. Ind. Eng. Chem. 2015, 24, 219−228.

10406

DOI: 10.1021/acssuschemeng.7b02434 ACS Sustainable Chem. Eng. 2017, 5, 10396−10406