Development of a Zwitterionic Compound Derived from β-Amino Acid

Oct 5, 2017 - The corrosion potential (Ecorr) was monitored and recorded throughout the entire experiments. Before any electrochemical measurements, t...
1 downloads 15 Views 1MB Size
Subscriber access provided by UNIVERSITY OF CONNECTICUT

Article

Development of a Zwitterionic Compound derived from #Amino Acid as Green Inhibitor for CO2 Corrosive Environments Alejandro Ramírez-Estrada, Violeta Y. Mena-Cervantes, Ignacio Elizalde, Arturo Manzo-Robledo, Luis S. Zamudio-Rivera, David Aaron Nieto-Alvarez, Fernando Farelas, and Raúl Hernández-Altamirano ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b02434 • Publication Date (Web): 05 Oct 2017 Downloaded from http://pubs.acs.org on October 6, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Development of a Zwitterionic Compound derived from β-Amino Acid as Green Inhibitor for CO2 Corrosive Environments Alejandro Ramírez-Estrada1, Violeta Y. Mena-Cervantes1,5, Ignacio Elizalde1, Arturo Manzo-Robledo2, Luis S. Zamudio-Rivera3, David A. Nieto-Álvarez 3, Fernando Farelas4 and Raúl Hernández-Altamirano*1,5 1

Centro Mexicano para la Producción más Limpia, Instituto Politécnico Nacional, Av. Acueducto s/n, Col. La Laguna Ticomán, Ciudad de México, 07340, México. 2 Instituto Politécnico Nacional, Laboratorio de Electroquímica y Corrosión, DIQI-SEPI-Escuela Superior de Ingeniería. Química e Industrias Extractivas (ESIQIE-IPN), UPALM, Col. Zacatenco, Edificio Z-5, 3er piso, C.P. 07738, Ciudad de México. 3 Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte 152, Col. San Bartolo Atepehuacan, Gustavo A. Madero, Ciudad de México, 07730, México. 4 Institute for Corrosion and Multiphase of Technology, Department of Chemical and Biomolecular Engineering, Ohio University, Athens, OH 45701, United States. 5 Laboratorio Nacional de Desarrollo y Aseguramiento de la Calidad de Biocombustibles (LaNDACBio). (*) corresponding author: [email protected], telephone: +52 55 57296000, Ext. 52621

Abstract Herein, the synthesis of three zwitterionic compounds (GZC-8, GZC-12 and GZC-18) derived from βAmino 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 condition at ambient pressure and temperature. These compounds were characterized by FT-IR, 2D NMR, EIS-MS, elemental analysis and then evaluated as potential corrosion inhibitors for CO2 acidic environments characteristic of Enhanced Oil Recovery (EOR) processes of petroleum industry through electrochemical techniques such as open circuit potential (OCP), lineal polarization resistance (LRP), electrochemical impedance spectroscopic (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® test from which an EC50 value of 209, 188 and 43 ppm was obtained for GZC-8, GZC-12 and GZC-18, respectively, i.e., two orders of magnitude lower than typical ionic liquid inhibitors. From experimental results, it was found that 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

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 2 of 25

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 formation1–6. This process starts with the injection of CO2 into the oil producing formation7, when CO2 contacts oil, several mechanisms promote the mobilization of oil in immiscible and miscible processes improving the oil recovery8,9. However, CO2 dissolution in water produces carbonic acid (H2CO3), decreasing the pH of the connate water10. This acidic solution can cause failure on main downhole tubing, transition pipelines and separation units11, and disruption of the oil/gas production. The consequences of failures caused by corrosion can include equipment replacement, production shutdown, environmental damage and safety risks12, resulting in severe economic losses13. 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 media14. A large number of organic molecules have been developed and extensively studied as corrosion inhibitors15. These molecules regularly have amphiphilic structure consisting of a polar head group which interacts with metal surface and a non-polar tail group which provides hydrophobic metal-protecting 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, inhibition corrosion, among others16,17. 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 groups18. Polar head group contains heteroatoms such as phosphorus, nitrogen, sulphur 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-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 environment20–22. However, despite the efforts, the high toxicity of some of these molecules, including recently developed ionic liquids (IL) such as 1-Octyl-3-methyl pyridinium 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 another hand, 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. 2 ACS Paragon Plus Environment

Page 3 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

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 non-toxic 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 eco-friendly synthesis route was designed for the obtaining of three 3-(alkylamino) propanoic acids in zwitterionic form, GZC-8, GZC-12 and GZC-18. Also characterized 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 24h of immersion.

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 the chemical synthesis of the 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 three-necked balloon flask of 50 mL at a temperature of 30ºC 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 minutes 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 Perkin Elmer Frontier FT-IR spectrometer equipped with a diamond attenuated 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 Hz. 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 Perkin Elmer Series II 2400. Acute toxicity by Microtox® test

3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 4 of 25

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 3ºC 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 percent per mL 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 200mL 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 1.0M NaCl solution that was purged with CO2 during 30 min prior immersion of the WE (pH=4.03). After 30 minutes, 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 (2.827cm2 exposed area) with the following chemical composition (wt. %): 0.16 C, 0.79 Mn and 99.05 Fe. The electrodes were grounded with silicon-carbide sand paper until a 1200 grit surface finish was obtained, then were cleaned in an ultrasonic bath, rinsed with acetone and dried before immersed into the solution. The GZC-8, GZC-12 and GZC-18 were pre-diluted 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 pre-corroded for 30 minutes to reach steady-state conditions. The GZC solution was added after 1h. The laboratory methodology of evaluation consisted on the following steps: i)

ii)

All three GZC compounds were characterized trough Lineal Polarization Resistance (LPR) measurements at 25 ppm and atmospheric pressure in aforementioned electrochemical glass cell, at a constant temperature of 70°C. The Lineal Polarization Resistance (LPR) measurements were taken during 24h, at ±5 mV vs Ecorr and a scan rate of 0.125 mV•s-1. 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. 4 ACS Paragon Plus Environment

Page 5 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

iii)

iv)

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 y 75 ppm from which it was possible to identify the optimum concentration for selected GZC compound. 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 (-300mV vs Ecorr) and anodic (+300mV vs Ecorr) direction at scan rate of 0.166 mV•s-1, after 24h of immersion

Results and discussion Synthesis and Chemical characterization GZ-8: 6.33 g (0.049 mol) of octylamine were added to a three-necked balloon flask of 50 mL at a temperature of 30ºC under vigorous stirring. Then 3.53g (0.049 mol) of acrylic acid were added slowly. The reaction was exothermic and the temperature under these conditions increased gradually to 70-80°C The reaction was kept for 30 minutes and then 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 3371cm-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: 1 H NMR (750MHz, 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 (187.5MHz, CDCl3, δ ppm); 176.918 (C-1), 46.954 (C-3), 44.527 (C-4), 31.783 (C2), 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 ESI–MS technique was employed to measure the molecular weight of GZC-8. Under the positive ion mode, the cationic [GZC-8 + H]+, (i.e. one mole of GZC-8) was binding with one mole 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: 9.08 g (0.049 mol) of dodecylamine were added to a three-necked balloon flask of 50 mL at a temperature of 30ºC under vigorous stirring. Then 3.53g (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 minutes 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 3374cm-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 5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

were: 1H-NMR (750MHz, CDCl3, δ ppm); 0.876 (t, J = 6.6 Hz, 3H, H-15), 1.246 (m, 18H, H-6 to H14), 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 (187.5MHz, CDCl3, δ ppm); 176.904 (C-1), 47.050 (C-3), 44.594 (C-4), 31.901 (C-5), 31.693 (C-2), 29.113-29.618 (C-6 to C-11), 26.753 (C-12), 26.284 (C-13), 22.678 (C-14), 14.120 (C-15). These 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 formation34. 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. one mole of GZC-12) was binding with one mole 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: 13.21 g (0.049 mol) of octadecylamine were added to a three-necked balloon flask of 50 mL at a temperature of 30ºC under vigorous stirring. Then 3.53g (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 minutes 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 3369cm-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: 1H NMR (750MHz, 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 (187.5MHz, CDCl3, δ ppm); 176.918 (C-1), 47.113 (C-3), 44.534 (C-4), 31.921 (C-2), 31.721 (C-5), 29.125-29.638 (C-6 to C-17), 26.728 (C-18), 26.218 (C-19), 22.633 (C-20), 14.112 (C-21). The ESI–MS technique was employed to measure the molecular weight of GZC-18. Under the positive ion mode, the cationic [GZC-18 + H]+, (i.e. one mole of GZC-18) was binding with one mole of protons. The mass-to-charge 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-organism35. Furthermore, the use of Vibrio fischeri based bioluminescence assay for toxicity measurement of polluted water has increased due to good reproducibility and sensibility36. 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 bioluminescence37.

6 ACS Paragon Plus Environment

Page 7 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Acute toxicity test results for GZC compounds at 5 and 15 min are listed in Table 1. The EC50 results were three and two 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 head group 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 salts23,38–40. On another hand, at a fixed tail group length, toxicity depends on the nature of head group 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 molecule41. The most common cationic-type surfactants are the quaternary ammonium compounds; these molecules can bind to inner membranes and disorganize them via their long alkyl chain42. Thus surfactant interaction with cellular membranes depends on chemical structure, charge, critical micellar concentration and hydrophile-lipophile balance43,44. In general, the cationic nature of molecules is associated with cytotoxic effects45,46. 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-corrosion-electrochemistry When water and CO2 are present, a small fraction of CO2 is hydrated, forming carbonic acid (H2CO3) according to equations 1 and 2. H2CO3 will further dissociate into bicarbonate and hydrogen ions (equation 3). Finally, bicarbonate ions are dissociated into carbonate and hydrogen ions (equation 4): CO2(g)  CO2(aq) CO2(aq) + H2O(l)  H2CO3(aq) H2CO3(aq)  HCO3-(aq) + H+(aq) HCO3-(aq)  CO32-(aq)+ H+(aq)

(1) (2) (3) (4)

The electrochemical process of CO2 corrosion of carbon steels involves the reduction of hydrogen and the anodic dissolution of iron48. Hydrogen ions from equations 3 and 4 will be reduced at cathodic sites on the metal surface (equation 5). Water reduction (equation 6) is possible but only at very negative potentials: H+(aq) + e-  ½ H2(g)

(5) 7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

H2O(l) + e-  OH-(aq) + ½ H2(g)

Page 8 of 25

(6)

Since iron is the main component of carbon steels, the anodic reaction comprises the oxidation of iron into ferrous ions, equation 7: Fe(s) + 2e-  Fe2+(aq)

(7)

Numerous studies have been devoted to the study of CO2 corrosion-electrochemistry, suggesting a multistep mechanism to explain the Fe dissolution process10,48–53. These studies suggest that at pH≤4 the reduction of H+ (equation 4) is one of the key cathodic processes. CO2 corrosion of carbon steels is usually accompanied by the formation of protective or non-protective corrosion products. At high temperature and pressure and pH higher than 6, the formation of protective iron carbonate (FeCO3) is favored54–57. 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 corrosion58-60. Fe3C is part of the steel makeup, being conductive and nobler than the ferrite phase (α-Fe). Therefore, reduction of hydrogen ions (equation 5) will occur on Fe3C sites, while the dissolution of the ferrite phase (equation 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-steel electrode in 1.0M 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 law61. Whereas, the 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 equation 8 from CR values for GZC compounds.

%IE =

CR  − CR X 100 CR 

(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 GZC-8. 8 ACS Paragon Plus Environment

Page 9 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

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 mixed-type 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 (AISI-1018 C-steel electrode) in a 1.0M 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 on the Ecorr to more positive values, 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, equation 9:  =

 

(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 equation 10: =

  2.303 +  

(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 charge62. When the surface bears an opposite charge to the ionic molecules, adsorption tends to be maximized63, 64. The inhibition effect is attributed to the adsorption of the organic-ionic molecules via their functional ionic-groups and their ability to aggregate and form a protective layer. The results shown in Figure 5b 9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

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 non-protective 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 ions17,63. Thus, these hydrophilic groups are adsorbed on the active-sites of the metal surface18. 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 performed on AISI-1018 C-steel samples exposed to a 1.0M NaCl-CO2 solution at 70°C. Nyquist and Bode plots at different immersion times are shown in Figure 6 and Figure 7 respectively. In the absence of GZC-12 (at 1h 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 to 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 dissolution65,66. 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 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 non-ideal frequency-response is observed, a constant phase element (CPE) was used. The CPE is defined in equation (11) as67: 

!"

= #$% &'$(

(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 represent 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 the equation 12. %$(- %/(-

*+ = ,#+ +

.

(12) 10

ACS Paragon Plus Environment

Page 11 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

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, 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 is under diffusion control68,69. 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, 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 equation (13)70. 0 = 1 21 '$( 345ℎ720 '$( 8

(13)

Where Rw represents the diffusion resistance, the exponent n is 0.5 and Tw is defined as (equation 14): 20 =

9: ;

(14)

Where T is the finite length diffusion element exponent (W-T), L the effective diffusion length or thickness and D 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 film71-74. 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-layers17 and because of that CGZC could be slightly increased.

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 12 of 25

Conclusions Three Green Zwitterionic Compounds were synthetized trough an eco-friendly zero waste route, i.e., in 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 carbon steel surface due to their dual electrical charge distribution given its zwitterion nature, forming a resistant protective layer due to the alkyl non- polar tail group. 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. In addition, the EC50 result from Microtox® test was 209, 188 and 43 ppm for GZC-8, GZC-12 and GZC-18, respectively, showing that these molecules are two orders of magnitude less toxic than recently developed IL (ionic liquids) corrosion inhibitors and thus, compatible with human health and environment. This observed behavior might be linked to the effect of particular formal charge distribution in zwitterionic polar head group at GZC. From these results, GZC-12 was selected as the best option to develop green corrosion inhibitor as it represented a balance between high corrosion inhibiting efficiency and lower toxicity. Further electrochemical evaluation showed that GZC-12 exhibited excellent inhibition performance for 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 12h of exposure, GZC-12 affected both anodic and cathodic reactions as an indication of a change on the corrosion mechanism mode from charged transfer to diffusion control. This is the first report on the synthesis and evaluation of non-toxic 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. Acknowledgements

12 ACS Paragon Plus Environment

Page 13 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

This research was financially supported by FONDO DE INVESTIGACIÓN PARA LA EDUCACIÓN-CONACYT under the project No. 241262, SIP-IPN multidisciplinary project No. 2017-1864. Special acknowledge 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, 1988).

2.

Alvarado, V. & Manrique, E. Enhanced oil recovery: An update review. Energies 3, 1529–1575 (2010). https://doi.org/10.3390/en3091529

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. 10, 1930–1935 (2011). https://doi.org/10.1016/j.proeng.2011.04.320

4.

Koottungal, L. 2012 worldwide EOR survey. Oil Gas J. 110, 57–69 (2012).

5.

Ilman, M. N. Analysis of internal corrosion in subsea oil pipeline. Case Stud. Eng. Fail. Anal. 2, 1–8 (2014). https://doi.org/10.1016/j.csefa.2013.12.003

6.

Yevtushenko, O. et al. Corrosion behavior of steels for CO2 injection. Process Saf. Environ. Prot. 92, 108–118 (2014). https://doi.org/10.1016/j.psep.2013.07.002

7.

Chukwudeme, E. A. & Hamouda, A. A. Enhanced oil recovery (EOR) by miscible CO2 and water flooding of asphaltenic and non-asphaltenic oils. Energies 2, 714–737 (2009). https://doi.org/10.3390/en20300714

8.

Pini, R., Krevor, S., Krause, M. & Benson, S. Capillary heterogeneity in sandstone rocks during CO2/water core-flooding experiments. Energy Procedia 37, 5473–5479 (2013). https://doi.org/10.1016/j.egypro.2013.06.467

9.

Cerasi, P., Kjøller, C., Sigalas, L., Bhuiyan, H. & Peter, F. Mechanical effect of CO2 flooding of a sandstone specimen. Energy Procedia 86, 361–370 (2016). https://doi.org/10.1016/j.egypro.2016.01.037

10.

Choi, Y. S. & Nesic, S. Determining the corrosive potential of CO2 transport pipeline in high pCO2-water environments. Int. J. Greenh. Gas Control 5, 788–797 (2011). https://doi.org/10.1016/j.ijggc.2010.11.008

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. 1, 6–14 (2013). https://doi.org/10.1016/j.csefa.2012.11.001

12.

Sun, C. et al. Investigate deepwater pipeline oil spill emergency repair methods. Aquat. Procedia 3, 191–196 (2015). https://doi.org/10.1016/j.aqpro.2015.02.210

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. 1–12 (2002).

14.

Fink, J. K. Oil field chemicals. Gulf Professional PublishingG 1, (Gulf Professional Publishing, 13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

2015). 15.

Arora, A. & Pandey, S. K. Review on materials for corrosion prevention in oil industry. in SPE International Conference and Exhibition on Oilfield corrosion 1–11 (2012). https://doi.org/10.2118/155211-MS

16.

Vittal, R., Gomathi, H. & Kim, K. J. Beneficial role of surfactants in electrochemistry and in the modification of electrodes. Adv. Colloid Interface Sci. 119, 55–68 (2006). https://doi.org/10.1016/j.cis.2005.09.004

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. 233, 403–414 (2017). https://doi.org/10.1016/j.molliq.2017.02.111

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. 6, 1927–1948 (2011).

19.

Finšgar, M. & Jackson, J. Application of corrosion inhibitors for steels in acidic media for the oil and gas industry: A review. Corros. Sci. 86, 17–41 (2014). https://doi.org/10.1016/j.corsci.2014.04.044

20.

Asipita, S. A. et al. Green Bambusa Arundinacea leaves extract as a sustainable corrosion inhibitor in steel reinforced concrete. J. Clean. Prod. 67, 139–146 (2014). https://doi.org/10.1016/j.jclepro.2013.12.033

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. Clean. Prod. 112, 666–671 (2016). https://doi.org/10.1016/j.jclepro.2015.07.124

22.

Qiang, Y. et al. Sodium dodecyl benzene sulfonate as a sustainable inhibitor for zinc corrosion in 26% NH4Cl solution. J. Clean. Prod. 152, 17–25 (2017). https://doi.org/10.1016/j.jclepro.2017.03.104

23.

Docherty, K. M. & Kulpa, Jr., C. F. Toxicity and antimicrobial activity of imidazolium and pyridinium ionic liquids. Green Chem. 7, 185 (2005).https://doi.org/10.1039/B419172B

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. 95, 168–179 (2015). https://doi.org/10.1016/j.corsci.2015.03.012

25.

Sasson, Y. & Neumann, R. in Handbook of Phase Transfer Catalysis 563 pp. (Blackie Academic & Professional, 1997). https://doi.org/10.1007/978-94-009-0023-3

26.

Vanyúr, R., Biczók, L. & Miskolczy, Z. Micelle formation of 1-alkyl-3-methylimidazolium bromide ionic liquids in aqueous solution. Colloids Surfaces A Physicochem. Eng. Asp. 299, 256– 261 (2007). https://doi.org/10.1016/j.colsurfa.2006.11.049

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 PM-IRRAS studies. Electrochim. Acta 56, 2990–2998 (2011). https://doi.org/10.1016/j.electacta.2011.01.009

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. 65, 104–112 (2012). 14 ACS Paragon Plus Environment

Page 15 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

https://doi.org/10.1016/j.corsci.2012.08.001 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. 59, 42–54 (2012). https://doi.org/10.1016/j.corsci.2012.02.011

30.

Pasasa, N. V. A., Bundjali, B. & Wahyuningrum, D. The microwave assisted synthesis of 1-alkyl3-methylimidazolium bromide as potential corrosion inhibitor toward carbon steel in 1 M HCl solution saturated with carbon dioxide. AIP Conf. Procedingsedings 1677, 17 (2015). http://doi.org/10.1063/1.4930720

31.

Hernández-Altamirano, R., Zamudio-Rivera, L. S., Mena-Cervantes, V. Y., Beltrán-Conde, H. I., Dominguez-Aguilar, M. A., Martínez-Viramontes, J., & Estrada-Buendia, A. Amino and imino propionic acids, process af preparation and use. (2015). US Patent 20150112096.

32.

Stuart, B. H. Infrared spectroscopy: fundamentals and applications. (John Willey & Sons, Ltd., 2004). https://doi.org/10.1002/0470011149

33.

Silverstein, M. R., Webster, F. X. & Kiemle, D. J. Spectrometric identification of organic compounds. (John Wiley & Sons, Inc., 2005).

34.

Pretsch, E., Bühlmannn, P. & Badertscher, M. Structure determination of organic compounds. (Springer-Verlag, 2009). http://dx.doi.org/10.1007/978-3-540-93810-1

35.

Grigson, S., Cheong, C. & Way, E. Studies of produced water toxicity using luminescent marine bacteria. WIT Transactions on Biomedicine and Health 10, 111–121 (2006). https://doi.org/10.2495/ETOX060111

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. Vitr. 22, 1806–1813 (2008). https://doi.org/10.1016/j.tiv.2008.07.011

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. 32, 265–268 (2006). https://doi.org/10.1016/j.envint.2005.08.022

38.

Morán, M. C. et al. ‘Green’ amino acid-based surfactants. Green Chem. 6, 233–240 (2004). https://doi.org/10.1039/B400293H

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. 58, 396–404 (2004). https://doi.org/10.1016/S0147-6513(03)00105-2

40.

Cho, C. W. et al. Toxicity of imidazolium salt with anion bromide to a phytoplankton Selenastrum capricornutum: Effect of alkyl-chain length. Chemosphere 69, 1003–1007 (2007). https://doi.org/10.1016/j.chemosphere.2007.06.023

41.

Cserha, T. Forga, E. Oros, G. Biological activity and environmental impact of anionic surfactants. Environ. Int. 28, 337–348 (2002). https://doi.org/10.1016/S0160-4120(02)00032-6

42.

Zhang, C. et al. Quaternary ammonium compounds (QACs): A review on occurrence, fate and toxicity in the environment. Sci. Total Environ. 518–519, 352–362 (2015). https://doi.org/10.1016/j.scitotenv.2015.03.007 15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 16 of 25

43.

Ying, G. G. Fate, behavior and effects of surfactants and their degradation products in the environment. Environ. Int. 32, 417–431 (2006). https://doi.org/10.1016/j.envint.2005.07.004

44.

Masakorala, K., Turner, A. & Brown, M. T. Toxicity of synthetic surfactants to the marine macroalga, Ulva lactuca. Water. Air. Soil Pollut. 218, 283–291 (2011). https://doi.org/10.1007/s11270-010-0641-4

45.

Lv, H., Zhang, S., Wang, B., Cui, S. & Yan, J. Toxicity of cationic lipids and cationic polymers in gene delivery. J. Control. Release 114, 100–109 (2006). https://doi.org/10.1016/j.jconrel.2006.04.014

46.

Kim, Y., Binauld, S. & Stenzel, M. H. Zwitterionic guanidine-based oligomers mimicking cellpenetrating peptides as a nontoxic alternative to cationic polymers to enhance the cellular uptake of micelles. Biomacromolecules 13, 3418–3426 (2012). https://doi.org/10.1021/bm301351e

47.

Ribo, J. M. & Kaiser, L. E. Photobacterium phosphoreum toxicity bioassay. I. Test procedures and applications. Environ. Toxicol. 2, 305–323 (1987). https://doi.org/10.1002/tox.2540020307

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. Corros. Sci. 59, 489–497 (2003). https://doi.org/10.5006/1.3277576

49.

Nešić, S. Key issues related to modelling of internal corrosion of oil and gas pipelines – A review. Corros. Sci. 49, 4308–4338 (2007). https://doi.org/10.1016/j.corsci.2007.06.006

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. 1, 156–162 (2012).

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. in Corrosion NACE conference & expo 16 (Paper No. 0001418, 2012).

52.

Dugstad, A. Fundamental aspects of CO2 metal loss corrosion. Part I: mechanism. in Corrosion NACE conference & expo 1–10 (2006).

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. 14, 537–548 (2014). https://doi.org/10.1007/s1166

54.

Tanupabrungsun, T., Brown, B. & Nesic, S. Effect of pH on CO2 Corrosion of Mild Steel at Elevated Temperatures. in Corrosion NACE conference & expo 1–11 (2013).

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. Greenh. Gas Control 23, 30–43 (2014). https://doi.org/10.1016/j.ijggc.2014.01.016

56.

Tavares, L. M., Costa, E. M. Da, 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. 359, 143–152 (2015). https://doi.org/10.1016/j.apsusc.2015.10.075

57.

Esmaeely, S. N., Choi, Y.-S., Young, D. & Neši, S. Effect of Calcium on the Formation and Protectiveness of Iron Carbonate Layer in CO2 Corrosion. Corrosion 69, 912–920 (2013). http://dx.doi.org/10.5006/0942 16 ACS Paragon Plus Environment

Page 17 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

58.

Crolet, J. L., Thevenot, N. & Nesic, S. Role of conductive corrosion products in the protectiveness of corrosion layers. Corrosion 54, 194–203 (1998). https://doi.org/10.5006/1.3284844

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. 85, 380–393 (2014). https://doi.org/10.1016/j.corsci.2014.04.055

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. 156, 198–205 (2015). https://doi.org/10.1016/j.matchemphys.2015.02.047

61.

ASTM G 102-89. Standard practice for calculation of corrosion rates and related information. ASTM 1–7 (2010).

62.

Belarbi, Z., Farelas, F., Singer, M. & Nesic, S. Role of amine in the mitigation of CO2 top of the line corrosion. NACE - Int. Corros. Conf. Ser. 2, 1300–1310 (2016). https://doi.org/10.5006/2121

63.

Paria, S. & Khilar, K. C. A review on experimental studies of surfactant adsorption at the hydrophilic solid - water interface. Adv. Colloid Interface Sci. 110, 75–95 (2004). https://doi.org/10.1016/j.cis.2004.03.001

64.

Bockris, J. O. M., Reddy, A. K. N. & Gamboa, M. Modern electrochemistry 2A: Fundamentals of electrodics. (Springer US, 2000). https://doi.org/10.1007/b113922

65.

Keddam, M., Mattos, O. R. & Takenout, H. Reaction Model for Iron Dissolution Studied by Electrode Impedance. J. Electrochem. Soc. 128, (1981). https://doi.org/10.1149/1.2127401

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. 120, 239–250 (2017). https://doi.org/10.1016/j.corsci.2017.02.016

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. 45, 1243–1256 (2003). https://doi.org/10.1016/S0010-938X(02)00214-7

68.

Chen, Y., Hong, T., Gopal, M. & Jepson, W. P. EIS studies of a corrosion inhibitor behavior under multiphase flow conditions. Corros. Sci. 42, 979–990 (2000). https://doi.org/10.1016/S0010-938X(99)00127-4

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. 61, 148–155 (2012). https://doi.org/10.1016/j.corsci.2012.04.034

70.

Finšgar, M. 2-Mercaptobenzimidazole as a copper corrosion inhibitor: Part I. Long-term immersion, 3D-profilometry, and electrochemistry. Corros. Sci. 72, 82–89 (2013). https://doi.org/10.1016/j.corsci.2013.03.011

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. 253, 814–820 (2006). https://doi.org/10.1016/j.apsusc.2006.01.038

72.

Farelas, F. & Ramirez, A. Carbon dioxide corrosion inhibition of carbon steels through bisimidazoline and imidazoline compounds studied by EIS. Int. J. Electrochem. Sci. 5, 797–814 17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

(2010). 73.

Zhao, I. & Chen, G. The synergistic inhibition effect of oleic-based imidazoline and sodium benzoate on mild steel corrosion in a CO2-saturated brine solution. Electrochim. Acta 69, 247– 255 (2012). https://doi.org/10.1016/j.electacta.2012.02.101

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. 24, 219–228 (2014). https://doi.org/10.1016/j.jiec.2014.09.034 Tables

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) 209 188 43

-

1.1723

-

1.7723

-

0.2438,47

-

0.6338,47

Table 2. Electrochemical parameters from polarization measurements for AISI-1018 C-steel electrode in a 1.0M NaCl solution saturated with CO2 at 70°C for GZC compounds at 25 ppm, after 24h immersion.

Compound

Ecorr V/(Ag/AgCl)

βa (mV•dec-1)

-βc (mV•dec-1)

icorr (A•cm-2)

CR (mmpy)

%IE

Reference

-0.665

68

136

2.76 X 10-4

3.206

-

274

1.06 X 10

-4

1.235

61.47

3.51 X 10

-5

0.408

87.28

2.72 X 10

-5

0.316

90.14

GZC-8 GZC-12 GZC-18

-0.625 -0.627 -0.647

65 68 63

226 152

Table 3. Electrochemical parameters from polarization measurements for AISI-1018 C-steel electrode in a 1.0M NaCl solution saturated with CO2 at 70°C for GZC-12 at different concentrations, after 24h immersion.

Concentration

Ecorr

βa

-βc

icorr

CR

%IE

18 ACS Paragon Plus Environment

Page 19 of 25

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

V/(Ag/AgCl) 25

(mV•dec-1)

-0.627

(mV•dec-1)

68

226

(A•cm-2)

(mmpy)

3.51 X 10

-5

0.408

87.28

-6

0.046

98.57

0.034

98.93

50

-0.607

87

216

3.96 X 10

75

-0.597

69

212

2.95 X 10-6

Table 4. EIS parameters obtained by fitting the experimental results using the ECs shown in Figure 8 for 75 ppm of GZC-12. CPEdl Time (h)

Rs (Ω•cm2)

1 3 4 6 8 9 12 16 20 23

3.1 2.9 2.8 2.7 2.6 2.6 2.1 2.4 2.5 2.5

Rct (Ω•cm2)

CPEGZC

Ydl (Ω-1•sn•cm-2)

n (0-1)

Cdl (µF•cm-2)

4.77x10-4 1.92x10-4 1.73x10-4 1.15x10-4 8.54x10-5 5.02x10-5 1.84x10-4 1.56x10-4 1.56x10-4 1.43x10-4

0.79 0.72 0.69 0.66 0.63 0.64 0.69 0.66 0.67 0.68

225.61 157.38 141.72 79.76 50.92 22.979 160.72 106.87 100.21 84.38

125 3130 3706 4258 4848 4966 4055 3058 2621 2271

RGZC (Ω•cm2)

Rw (Ω•cm2)

WGZC

YGZC (Ω-1•sn•cm-2)

n (0-1)

CGZC (µF•cm-2)

2.07x10-5 4.55x10-5 1.07x10-4 1.43x10-4 1.77x10-4 3.37x10-5 5.24x10-5 4.12x10-5 4.14x10-5

0.96 0.88 0.79 0.75 0.73 0.79 0.81 0.83 0.83

14.11 15.80 22.60 24.33 28.44 2.52 7.63 7.33 7.50

4.7 9.4 26.8 34.5 40 1.7 4.7 5.3 5.7

W-T

n (0-1)

0.65 0.84 1.17 1.49

0.5 0.5 0.5 0.5

1370 2755 3320 3863

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

Figures

3

R

2

R

1

R: 5

GZC-8

4

GZC-12

4

GZC-18

4

9

7 6

5

7 6

5

11

9

11

9 8

13 12

10

8

7 6

11 10

8

10

15 14

15

13 12

14

17 16

18

19

20

21

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

20 ACS Paragon Plus Environment

Page 21 of 25

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

-0.3 -0.4 -0.5 E / V(Ag/AgCl)

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

-0.6 Reference 25 ppm GZC-8 25 ppm GZC-12 25 ppm GZC-18

-0.7 -0.8 -0.9 -1.0 -7 10

-6

10

-5

10

10

-4

-3

10

-2

10

-1

10

0

10

-2 i / Ahcm

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

21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

-0.3 -0.4

E / V(Ag/AgCl)

-0.5 -0.6

Reference 25 ppm 50 ppm 75 ppm

-0.7 -0.8 -0.9 -1.0 -7 10

-6

10

-5

10

-4

-3

10

-2

10

10

-1

10

0

10

-2 i / Ahcm

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

-0.58

8000

-0.59

(b)

(a)

after addition

7000

after addition

-0.60 6000

-0.61 -0.62

E / V (Ag/AgCl) corr

5000 2

-0.63 -0.64

Rp / Ωh cm

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

Reference 25 ppm 50 ppm 75 ppm

-0.65 -0.66

Reference 25 ppm 50 ppm 75 ppm

4000 3000

-0.67

2000

-0.68

1000

-0.69 0

-0.70 0

4

8

12

16

20

24

0

4

t/h

8

12

16

20

24

t/h

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

22 ACS Paragon Plus Environment

Page 23 of 25

3250

2000

1500

(a)

0.1 Hz

0.42 Hz

2750 2500

20 40 60 80 100 120

2250 2000

0.1 Hz

3.45 Hz

1000

1h 3h 4h 6h 8h 9h 12 h 16 h 20 h 23 h Fit

3000

2

2

2500

4.92 Hz

|Z| / Ωh cm

1h 60 3h 40 4h 20 6h 8h 0 9h 0 0.2 Hz 12 h 0.41 Hz 16 h 20 h 23 h 0.84 Hz Fit

3000

-Zimag / Ωh cm

(b)

0.1 Hz 0.2 Hz

1750 0.41 Hz

1500 0.59 Hz

500

1250

20.3 Hz

0 0

500

1000 1500 2000 2 Zreal / Ω hcm

2500

1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 Frequency / Hz

3000

Figure 6. Nyquist spectra of the AISI-1018 C-steel electrode immersed in 1.0M NaCl solution saturated with CO2 at 70°C and 75 ppm of GZC-12; a) Evolution of the impedance response with time and b) Zoom in at low frequencies (0.59 to 0.29 Hz) showing a 45° angle characteristic of a Warburg impedance. 4

70

10

(a)

(b)

1h 3h 4h 6h 8h 9h 12 h 16 h 20 h 23 h Fit

60 3

50

10

40 30 20 10 0

2

1h 3h 4h 6h 8h 9h 12 h 16 h 20 h 23 h Fit

|Z| / Ωh cm

-Phase angle / Degree

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

2

10

1

10

0

10 -1

10

0

10

1

10

2

10

3

10

4

10

-1

10

0

10

1

10

2

10

3

10

4

10

Frequency / Hz

Frequency / Hz

Figure 7. Bode spectra recorded for AISI-1018 C-steel electrode immersed in 1.0M NaCl solution saturated with CO2 at 70°C; a) Phase angle plot and b) Module plot.

23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 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 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

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

24 ACS Paragon Plus Environment

Page 25 of 25

Zwitterionic Compounds derived from β-Amino Acids as Green Inhibitors for CO2 Corrosive Environments R

-0.3

+

-0.4 -0.5 E / V(Ag/AgCl)

1 2 3 4 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

ACS Sustainable Chemistry & Engineering

Where R:

-0.6 Reference 25 ppm GZC-8 25 ppm GZC-12 25 ppm GZC-18

-0.7 -0.8

GZC-8 -0.9 -1.0 -7 10

GZC-12

-6

10

10

-5

-4

10

10

i / Acm

-3

-2

10

10

-1

GZC-18

One Pot Chemical Synthesis     

Steel without inhibitor

Steel with Green Corrosion Inhibitor (GZC-12)

Chemical Synthesis of Conventional Corrosion Inhibitors:

Ecofriendly Zero-waste Solventless reaction Low energy consumption Non-toxic products

× × ×

ACS Paragon Plus Environment

Intensive use of energy Waste generation Toxic products

0

10

-2