Skip to main content
ACS Omega logoLink to ACS Omega
. 2021 May 27;6(22):14525–14532. doi: 10.1021/acsomega.1c01422

Ethanedihydrazide as a Corrosion Inhibitor for Iron in 3.5% NaCl Solutions

Ayman H Ahmed †,‡,*, El-Sayed M Sherif §, Hany S Abdo §,, Ehab Said Gad †,
PMCID: PMC8190908  PMID: 34124475

Abstract

graphic file with name ao1c01422_0008.jpg

Corrosion of iron in sodium chloride (3.5% wt) solutions and its inhibition by ethanedihydrazide (EH) have been reported. Electrochemical impedance spectroscopy (EIS), cyclic potentiodynamic polarization (CPP), and change of current with time at −475 mV (Ag/AgCl) measurements were employed in this study. Scanning electron microscopy (SEM) and energy-dispersive X-ray (EDX) techniques were utilized to report surface morphology and elemental analysis, respectively. The presence of 5 × 10–5 M EH was found to inhibit the corrosion of iron, and the effect of inhibition profoundly increased with an increase in EH concentration up to 1 × 10–4 M and further to 5 × 10–4 M. The low values of corrosion current and high corrosion resistance, which were obtained from the EIS, CPP, and change of current with time experiments, affirmed the adequacy of EH as a corrosion inhibitor for iron. Surface investigations demonstrated that the chloride solution without EH molecules causes severe corrosion, while the coexistence of EH within the chloride solution greatly minimizes the acuteness of chloride, particularly pitting corrosion.

1. Introduction

Corrosion of metals is a major problem since it costs many billion dollars annually. The physiochemical interaction between a metal and its environment leads to variations in the properties of the metal, which may affect the function of the metal. Therefore, inhibition of metal corrosion is of paramount importance owing to the utilization of metals in various fields of innovation. Although metals are not completely safe from corrosion, we can prevent corrosion to some extent in metals. For this reason, corrosion inhibitors are used to interact with a metal surface and hence prevent metal corrosion. Iron metal and its numerous alloys are considered the most important materials as they have numerous practical applications. Iron is used to make alloys in addition to the manufacture of ships, rails, bridges, pipes, boilers, reinforcement steel, etc. Interestingly, iron corrosion causes a loss of one-fourth of world’s production in a year and this is an economic catastrophe.

According to the literature, many organic and inorganic compounds are utilized as inhibitors in controlling the corrosion of metals in 3.5% NaCl. Among these inorganic inhibitors, sodium molybdate, cerium nitrate, and Ce(III, IV) ammonium nitrate have been employed as corrosion inhibitors for steel pipelines,1 mild steel (MS),2 and AA2024 aluminum alloy,3 respectively. In the context of preservation of different metals in 3.5% NaCl brine solution using organic inhibitors, many substances have been evaluated. Macedo et al. studied the inhibition mechanism of imidazole and some of its derivatives onto an iron surface in 3.5% NaCl medium.4 The outcomes showed that the considered compounds act as anodic corrosion inhibitors for iron in saline medium. Aslam et al. investigated the inhibitory behavior of N,N′-didodecyl-N,N′-digluconamideethylenediamine gemini surfactant on mild steel (MS) corrosion in 3.5% NaCl at 30–60 °C.5 The compound mitigated corrosion, and the degree of inhibition was dependent on concentration and temperature. Moreover, 3-amino-5-mercapto-1,2,4-triazole,6 5-(3-aminophenyl)-tetrazole,7 1,1′-thiocarbonyldiimidazole,8 sodium 2-(4-(benzo[d]thiazol-2-ylthio)-6-(hexylamino)-1,3,5-triazin-2-ylamino) ethanesulfonate,9 hexamine,10 (3-amino-1,2,4-triazole-5-thiol and 1,1′-thiocarbonyldiimidazole),11 (2-sulfhydryl)-(5-phenmethyl)-(1-(4-phenol)-methanimine)-triazole,12 poly(ethyleneimine),13 1-butyl-3-methylimidazolium chloride,14 and cetylpyridinium chloride15 have been examined as corrosion inhibitors for different metals, especially iron species, in the same medium.

Although the previously mentioned inhibitors have an excellent corrosion inhibition efficiency, their high cost encourages researchers to look for alternative sources. For this reason, plant extracts/natural materials are chosen to restrain corrosion of metals because of their low cost, safety, and easy accessibility; moreover, they are environmentally friendly and biodegradable. Othman et al. examined the inhibition efficiency of rice straw extract for steel in 3.5% NaCl.16 Suarez et al. investigated the inhibition effect of mango extract as a corrosion inhibitor on carbon steel by varying the pH of the electrolytic medium.17 Furthermore, the corrosion inhibition performance of apple pomace,18Elettaria cardamomum,19 and Myrmecodia pendans(20) for C1010 steel, mild steel, and carbon steel, respectively, was evaluated in 3.5% NaCl medium and the accomplishments proved the effectiveness of these materials.

It is reported that efficient inhibitors are those compounds that contain heteroatoms like O, N, S, and P, which cause physical/chemical adsorption on metal surfaces.21,22 The inhibition performance of these compounds is enhanced in the order O < N < S < P.2325 In continuation of previous work on examining some N- and O-containing organic inhibitors,2629 the authors have considered the corrosion inhibiting behavior of a hydrazide, namely, ethanedihydrazide (EH) (Figure 1), on iron corrosion in an NaCl solution. Although some reports have documented the use of hydrazides as corrosion impeding agents for some iron species such as N′[(1E)-(4-hydroxyphenyl)methylene]isonicotinohydrazide and N′-[(1E)-(4-hydroxy-3-methoxyphenyl)methylene]isonicotinohydrazide,30 hydroxyphenylhydrazide,31 sulfonohydrazide derivatives,32 and pyrazine carbohydrazide,33 the corrosion inhibition ability of ethanedihydrazide for iron in neutral chloride medium (3.5% NaCl) has not been achieved up to now. It is noteworthy that the molecular structures of many reported inhibitors that slow down the corrosion rate of metals in a 3.5% NaCl environment often contain N and/or O heteroatoms and π-bonds (particularly as NH, C=O, and NH2 functional groups) analogous to the molecular structure of the studied inhibitor. Surely, the inhibition property of EH is attributed to its adequate chemical structure. The lone pairs of electrons on the heteroatoms (N, O) and the existence of π bonds in the entire structure, besides its ability to chelate with metals,34 are assumed to be significant factors that induce the adsorption of this molecule on an iron surface.

Figure 1.

Figure 1

Structure of ethanedihydrazide (EH).

In view of the inhibitive impact of organic compounds containing N and O heteroatoms, we examined the effect of ethanedihydrazide (EH) on iron corrosion in aerated 3.5% NaCl solutions employing various electrochemical strategies. The cathodic, anodic, and corrosion currents, corrosion potential, corrosion rate, and corrosion resistance of iron were assessed from the polarization data. Electrochemical impedance spectroscopy (EIS) experiments were performed to report Nyquist plots as well as the values of polarization resistances. The absolute flow currents and the influence of tested samples on corrosion inhibition were determined from the chronoamperometric current versus time runs. Scanning electron microscopy (SEM) analysis was performed to portray the surface morphology of iron metal after corrosion process. Energy-dispersive X-ray (EDX) techniques were used to detect the composition of the adsorbed layer on the iron surface in the absence and presence of the EH inhibitor.

2. Experimental Section

2.1. Chemicals and the Electrochemical Cell

Ethanedihydrazide (EH) was supplied by Sigma-Aldrich. NaCl salt (99% purity) was delivered by Merck. A high-purity iron rod (99.99%) with an area of 0.25 cm2 was purchased from Goodfellow (Ermine Business Park, Huntingdon, England) and was used as the working electrode in this work. Electrochemical measurements were performed in a traditional three-electrode electrochemical cell that can accommodate 0.250 L of solution. A Pt sheet and a silver–silver chloride (Ag/AgCl) electrode were employed as the counter and reference electrodes, respectively. The iron working electrode was prepared and employed to obtain corrosion results as reported in previous studies.35,36 EH powder was dissolved in dimethylformamide (DMF) to prepare a stock solution of 0.01 M EH. A stock solution of 7% NaCl was prepared by dissolving 70 g of powdered sodium chloride (NaCl) in 1000 mL of bidistilled water. The desired chloride solutions without and with EH molecules were synthesized from the stock solutions by dilution.

2.2. Surface Electrochemical Strategies and Characterization

The electrochemical experiments were measured using a PGSTAT302N Autolab that was delivered by Metrohm. The EIS data were obtained at the open-circuit potential value with frequency ranging from 100 MHz to 100 mHz. The impedance measurements were obtained via applying a ±5 mV amplitude sinusoidal wave perturbation. The polarization investigations (cyclic potentiodynamic polarization (CPP)) were carried out by scanning the potential from −1200 to +100 mV (Ag/AgCl) with a scanning rate of 1.66 mV/s. The current–time experiments were performed for the iron rod by stepping the potential to −475 mV for 1 h. The SEM micrographs and EDX investigations were collected using a JEOL microscope with an EDX unit attached; the machine operated at 15 kV. All experiments were performed in triplicate on a newly polished surface of iron with a fresh test solution at a temperature of 25 ± 3 °C.

3. Results and Discussion

3.1. Electrochemical Impedance Spectroscopy

The EIS technique was employed to report corrosion and corrosion inhibition processes.37,38 Nyquist plots for the iron rod after (1) 1 and (2) 48 h immersion in (a) 3.5% blank NaCl solution and 3.5% NaCl solutions with (b) 5 × 10–5, (c) 1 × 10–4, and (d) 5 × 10–4 M EH are shown in Figure 2. It was observed that the spectrum obtained for iron in the blank chloride solution (Figure 2a) had the lowest diameter for the obtained semicircle. Mixing 5 × 10–5 M EH with chloride solutions (Figure 2b) resulted in a wider diameter of the plotted semicircle; this effect was enhanced on adding higher concentrations of EH, 1 × 10–4 M and further 5 × 10–4 M.

Figure 2.

Figure 2

Nyquist plots for the iron rod after (1) 1 h and (2) 48 h immersion in (a) 3.5% blank NaCl solution and 3.5% NaCl solutions with (b) 5 × 10–5 M EH, (c) 1 × 10–4 M EH, and (d) 5 × 10–4 M EH.

The experimental data were fitted to an equivalent circuit that is exhibited in Figure 3. The symbols of the circuit can be defined as follows: a solution resistance (RS), constant phase elements (CPEs, Q), and a polarization resistance (RP). The values of these parameters were determined for the iron rod in all tested solutions, in addition to the inhibition efficiency percentage (IE/%) for the EH compound (shown in Table 1). The IE% values were obtained as follows3944

3.1. 1

Here, RPo and RP are the polarization resistances for the iron rod in chloride solutions without and with EH inhibitor molecules present, respectively.

Figure 3.

Figure 3

Circuit model that fits the impedance data.

Table 1. Values of EIS Parameters and IE% for Iron in 3.5% NaCl without and with EH.

    Q (CPEs)
   
parameter (solution) Rs (Ω cm2) YQ (F/cm2) n RP (Ω cm2) IE (%)
3.5% NaCl + 0.0 M EH (1 h) 9.65 7.45 × 10–6 0.86 655  
3.5% NaCl + 5 × 10–5 M EH (1 h) 10.23 6.55 × 10–6 0.89 1231 46.82
3.5% NaCl + 1 × 10–4 M EH (1 h) 10.91 6.02 × 10–6 0.91 1974 66.84
3.5% NaCl + 5 × 10–4 M EH (1 h) 11.54 5.72 × 10–6 0.93 3259 79.91
3.5% NaCl + 0.0 M EH (48 h) 10.13 7.23 × 10–6 0.82 691  
3.5% NaCl + 5 × 10–5 M EH (48 h) 10.86 6.77 × 10–6 0.86 1484 53.43
3.5% NaCl + 1 × 10–4 M EH (48 h) 11.44 5.33 × 10–6 0.89 2137 67.71
3.5% NaCl + 5 × 10–4 M EH (48 h) 11.97 4.42 × 10–6 0.91 3632 81.04

The Nyquist plots in Figure 2 reveal that the addition and increment of EH concentration minimize the corrosion impact of NaCl solution via inhibiting the deterioration of the iron rod. This can be deduced from the breadths of the semicircle in the Nyquist plots for 3.5% NaCl solution with EH. The values recorded in Table 1 also ascertain that the presence of EH increased all resistances (RS and RP) for iron, and this influence was boosted by augmentation of EH concentration. The presence of Q (CPEs) with the values of its “n” exponent varying between 0.8 and 0.9, indicated that Q can be considered as twofold layer capacitors with some porosity.

The impedance (ZCPE) and admittance (YCPE) of a CPE are defined as follows45

3.1. 2
3.1. 3

where Y0 is the modulus, ω is the angular frequency, j is the imaginary unit, and n is the phase.

The values of YQ (listed in Table 1) decrease when EH is present in the chloride solution and with increasing EH concentration. This can be attributed to the reduction of the capacitive impacts via the reduction in the local dielectric constant along with the growth in the thickness of the electrical double layer. EIS results subsequently boost the claim that the adsorption of EH molecules increases the inhibition of the iron surface; the adsorption process increases with the increment in EH concentrations. This suggestion is also supported by the increase of the IE% values with the increase of EH concentrations. It is worth mentioning that prolonging the immersion time to 48 h increases the inhibition efficiency of EH, particularly at a high concentration (5 × 10–4 M).

3.2. Cyclic Polarization (CPP) Data

Polarization, CPP, investigations were performed to report the corrosion reactions for the iron rod in the chloride solutions without and with EH added. Figure 4 displays the polarization measurements obtained for the iron rod after (1) 1 and (2) 48 h exposure in (a) 3.5% blank NaCl solutions and 3.5% NaCl solutions with (b) 5 × 10–5, (c) 1 × 10–4, and (d) 5 × 10–4 M EH. The corrosion current values (jcorr), corrosion potential (Ecorr), anodic, βa, and cathodic βc slopes, corrosion rate (Rcorr), and polarization resistance (RP) were obtained as reported in earlier studies.2527 The values of all of these parameters are listed in Table 2. The values of IE% were also calculated from polarization curves as per the following equation6

3.2. 4

Here, jcorri and jcorr0 are the corrosion current densities for the iron rod in the chloride solution with and without EH molecules, respectively.

Figure 4.

Figure 4

CPP curves for the iron rod after (1) 1 h and (2) 48 h immersion in (a) 3.5% blank NaCl solution and 3.5% NaCl solutions with (b) 5 × 10–5 M EH, (c) 1 × 10–4 M EH, and (d) 5 × 10–4 M EH.

Table 2. Data Obtained from Figure 4 for the Iron Rod in the Different Examined Solutions.

solution βc (mV/dec) Ecorr (mV) βa (mV/dec) jcorr (μA/cm2) RP (Ω cm2) Rcorr (mmpy) IE (%)
3.5% NaCl + 0.0 M EH (1 h) 80 –945 115 3.2 641.03 0.0823  
3.5% NaCl + 5 × 10–5 M EH (1 h) 75 –930 120 2.0 1003.3 0.0514 37.55
3.5% NaCl + 1 × 10–4 M EH (1 h) 70 –945 125 1.5 1300.6 0.0386 53.10
3.5% NaCl + 5 × 10–4 M EH (1 h) 65 –925 135 1.0 1766.3 0.0257 68.77
3.5% NaCl + 0.0 M EH (48 h) 70 –940 110 2.1 885.7 0.0540  
3.5% NaCl + 5 × 10–5 M EH (48 h) 65 –925 105 0.75 2327.4 0.0193 64.26
3.5% NaCl + 1 × 10–4 M EH (48 h) 75 –870 115 0.55 3588.5 0.0141 73.89
3.5% NaCl + 5 × 10–4 M EH (48 h) 65 –830 95 0.40 4195.0 0.0103 80.93

Figure 4 clarifies that the current in the cathodic side decreased until it reached the values of jcorr, after which the current increased in the anodic branch. The cathodic reaction for iron at these conditions was reported68 to be a reduction of oxygen as follows

3.2. 5

On the other hand, the reaction in the anodic branch was reported to be the active dissolution of metallic iron to a ferrous cation, followed by the appearance of a passive region due to iron hydroxide and then oxide layer formation as follows7

3.2. 6
3.2. 7

Owing to the instability of the formed Fe(OH)2, it is transformed to an iron oxide, which leads to the appearance of a long passive region. This region is due to the slow increase in current values accompanied by the increment of the potential in a less negative direction. Transformation of the hydroxide into an oxide can be shown by the following reaction6

3.2. 8

The current abruptly increases again due to the breakdown of the formed oxide film on the surface during passivation, which motivates the occurrence of pitting corrosion as well. The following reactions may take place with positive potential increment and attack of the chloride ions68,46

3.2. 9
3.2. 10
3.2. 11
3.2. 12

In reality, the presence of a hysteresis loop confirms the occurrence of pitting corrosion, where the currents in the forward direction are less than their values in the back scan direction. It is well known that the severity of pitting corrosion increases with increasing size of the hysteresis loop.

The polarization curves (Figure 4) and Table 2 indicate that 5 × 10–5 M EH reduced the cathodic and anodic currents and the values of jcorr and Rcorr, which increased the RP value. This effect increases on increasing the EH concentration to 1 × 10–4 M and further to 5 × 10–4 M. The highest IE% value was obtained at 5 × 10–4 M. Prolonging the immersion to 48 h before measurement promotes the EH efficiency, as seen in Table 2. Moreover, an increase in immersion time minimizes the degree of corrosion as a result of the thickening of an oxide film and/or the formation of a corrosion product layer. Whenever the thickness of the formed film increases, the degree of corrosion by the chloride solution decreases. In addition, the availability of EH in the solutions leads to the adsorption of its molecules onto the external surface of the iron rod, thus preventing corrosion. This is emphasized in the data listed in Table 2, as jcorr and Rcorr diminished, while the values of IE% and RP increased with increase in EH concentration. Increment of EH concentration also decreases the pitting corrosion degree. In comparison to other organic compounds, EH can be regarded as a good corrosion inhibitor.68,45,4750

3.3. Change of Current with Time Measurements

The change of current versus time after 1 and 48 h for the chloride solution without/with EH was performed. These experiments were carried out to report the influence of EH concentration on the pitting corrosion of the iron rod. Figure 5a shows the current–time curves collected at −475 mV (Ag/AgCl) for the iron rod in 3.5% NaCl solutions containing (1) 0.0 EH, (2) 5 × 10–5 EH, (3) 1 × 10–4 EH, and (4) 5 × 10–4 M EH after 1 h immersion. The same experiments were performed after 48 h, and the curves are plotted in Figure 5b. It was observed that the initial current obtained for the iron rod in all solutions, both after 1 h (Figure 5a) and 48 h (Figure 5b), recorded low values. Perhaps, this occurred because of oxide film formation during the immersion of the iron electrode prior to application of the potential. An increase in the time of potential application enhances the output current for all samples, confirming the occurrence of pitting corrosion. The most elevated current values were obtained for the iron rod in solution containing no EH. Besides, the continuous increment of the current values, which were recorded for the iron rod in 3.5% NaCl blank solution, indicated that both uniform and pitting corrosion have occurred. Here, the dissolution of iron into ferrous ions happens as per eq 6. Furthermore, at the applied potential value, the surface of the iron rod reacts with Cl of the solution to form FeCl2 as in eq 9. The FeCl2 surface film transfers to the FeCl3 surface film. Mixed films of FeCl2 and FeCl3 move to the interface and finally to the solution via eqs 1012, respectively.

Figure 5.

Figure 5

Current–time curves collected at −475 mV (Ag/AgCl) for the iron rod immersed for (a) 1 h and (b) 48 h in 3.5% NaCl solutions containing (1) 0.0 EH, (2) 5 × 10–5 M EH, (3) 1 × 10–4 M EH, and (4) 5 × 10–4 M EH.

At 5 × 10–5 M EH (Figure 5a, curve 2), the absolute value of the obtained current decreased due to the increase in the degree of uniform corrosion. Furthermore, an increase in EH concentration to 1 × 10–4 M (Figure 5a, curve 3) resulted in lower currents, particularly with expansion of the potential application period. Increment in the concentration of EH to 5 × 10–4 M (Figure 5a curve 4) led to the decrease of the current to the least values. This affirmed that the presence of EH molecules restrained the consumption of iron metal, and this impact was improved through an increase in the concentrations of EH in the chloride solution.

After 48 h exposure (Figure 5b), the change of current showed almost similar behavior to those acquired after 1 h (Figure 5a), where there was a steady increase in its value with time for the iron rod in all solutions. The most elevated currents were registered for the iron rod in the NaCl blank solution, while the presence of EH subsequently diminished the absolute current, which reduced further on increasing EH concentration. This is on the grounds that EH molecules get adsorbed on the outer surface of iron and diminish the corrosive attack of chloride solution.

3.4. SEM/EDX Analysis

To explore the morphology and the elements existing on the outer surface of the iron rod after being corroded in the chloride solution in the presence and absence of the EH compound, SEM analysis and EDX investigation were performed. Figure 6 displays the SEM images and EDX spectra that were obtained after the iron rod was immersed in (a) 3.5% NaCl and (b) 3.5% NaCl + 1 × 10–4 M EH solutions for 48 h before stepping the potential to −475 mV (Ag/AgCl) for 60 min. The SEM graph exhibits so many wide pits that resulted from the harsh effect of the chloride ions along with the active applied potential value. The element contents (atomic %) on the iron surface recorded by EDX analysis were 75.44% Fe, 14.26% O, 8.11% C, 1.26% Cl, and 0.93% Na. The highest percentage is obtained for Fe, followed by O, which suggests that the surface may develop an iron oxide film as per eqs 7 and 8. The presence of low and almost equal amounts of Cl and Na suggests that some NaCl salt may be adsorbed on the iron surface.

Figure 6.

Figure 6

SEM micrographs and EDX spectra collected after holding the potential at −475 mV for 1 h (Ag/AgCl) for the iron surface that was immersed in (a) 3.5% NaCl and (b) 3.5% NaCl + 1 × 10–4 M EH solutions for 48 h before measurement.

The SEM image shown in Figure 6b clearly portrays that the outer surface is completely covered with an inhibitor layer and/or corrosion products, in addition to the absence of any pits. The atomic percentages of the discovered elements on the surface of the iron rod (given by EDX analysis) were as follows: 56.98% Fe, 35.40% O, 4.65% C, 1.55% Cl, and 1.42% Na. The percentage of Fe here is excessively less, while that of O is too high in contrast to that revealed by Figure 6a. The surface is more passivated because of the presence of EH molecules. The low percentage of Na and Cl proves the presence of a thin adsorbed layer of NaCl salt on the surface of the iron rod. The SEM images and EDX profile are in good agreement with the impedance, CPP, and chronoamperometric electrochemical estimations.

4. Conclusions

The corrosion of iron in 3.5% NaCl solution and its inhibition by ethanedihydrazide (EH) molecules have been reported and the outcomes can be concluded as follows:

  • EIS data indicated that the molecules of EH can inhibit corrosion of iron via enhancing its corrosion resistance

  • CPP measurements revealed that EH mitigates corrosion of iron in NaCl solution by suppressing the values of jcorr and Rcorr and magnifying the RP values for iron

  • Chronoamperometric results obtained at −475 mV (Ag/AgCl) showed that the presence of EH decreases the possibility of pitting corrosion and reduces the values of absolute currents with the increase in time

  • The inhibitive impact of EH was found to increase with the increase in its concentration from 5 × 10–5 to 1 × 10–4 M and further to 5 × 10–4 M

  • Prolonging the exposure time from 1 to 48 h prevents corrosion of iron in NaCl solutions and increases the inhibition efficiency of EH, particularly at its high concentrations

  • SEM/EDX outcomes agree well with the data obtained by the electrochemical techniques; all data clearly presented that the degree of pitting corrosion is higher in the case of NaCl blank solution, while the availability of EH decreases the occurrence of pitting.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at Jouf University for funding this work through research grant No (DSR2020-05-412).

The authors declare no competing financial interest.

References

  1. Tristijanto H.; Ilman M. N.; TriIswanto P. Corrosion inhibition of welded of X–52 steel pipelines by sodium molybdate in 3.5% NaCl solution. Egypt. J. Pet. 2020, 29, 155–162. 10.1016/j.ejpe.2020.02.001. [DOI] [Google Scholar]
  2. Dastgheib A.; Attar M. M.; Zarebidaki A. Evaluation of corrosion inhibition of mild steel in 3.5wt% NaCl solution by cerium nitrate. Met. Mater. Int. 2020, 26, 1634–2642. 10.1007/s12540-019-00432-x. [DOI] [Google Scholar]
  3. Matter E. A.; Kozhukharov S.; Machkova M.; Kozhukharov V. Electrochemical studies on the corrosion inhibition of AA2024 aluminium alloy by rare earth ammonium nitrates in 3.5% NaCl solutions. Mater. Corros. 2013, 64, 408–414. 10.1002/maco.201106349. [DOI] [Google Scholar]
  4. Macedo M. C. S. S.; Barcia O. E.; Da Silva E. C.; Mendes J. D.; Mattos O. R. Corrosion inhibition of iron in NaCl 3.5% by some derivatives imidazole. J. Electrochem. Soc. 2012, 159, C160 10.1149/2.048204jes. [DOI] [Google Scholar]
  5. Aslam R.; Mobin M.; Aslam J.; Lgaz H. Sugar based N, N′didodecyl - N, N′ digluconamideethylenediamine gemini surfactant as corrosion inhibitor for mild steel in 3.5% NaCl solution-effect of synergistic KI additive. Sci. Rep, 2018, 8, 3690 10.1038/s41598-018-21175-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Sherif E.-S. M.; Erasmus R. M.; Comins J. D. In situ raman spectroscopy and electrochemical techniques for studying corrosion and corrosion inhibition of iron in sodium chloride solutions. Electrochim. Acta 2010, 55, 3657–3663. 10.1016/j.electacta.2010.01.117. [DOI] [Google Scholar]
  7. Sherif E.-S. M. Effects of 5-(3-aminophenyl)-tetrazole on the inhibition of unalloyed iron corrosion in aerated 3.5% sodium chloride solutions as a corrosion inhibitor. Mater. Chem. Phys. 2011, 129, 961–967. 10.1016/j.matchemphys.2011.05.043. [DOI] [Google Scholar]
  8. Sherif E.-S. M. Corrosion and corrosion inhibition of pure iron in neutral chloride solutions by 1,1’-thiocarbonyldiimidazole. Int. J. Electrochem. Sci. 2011, 6, 3077–3092. [Google Scholar]
  9. Zhu H.; Li X.; Lu X.; Chen X.; Li J.; Han X.; Ma X.; Hu Z. Intra-/inter-molecular synergistic inhibition effect of sulfonate surfactant and 2-benzothiazolethiol on carbon steel corrosion in 3.5% NaCl solution. Corros. Sci. 2021, 182, 109291 10.1016/j.corsci.2021.109291. [DOI] [Google Scholar]
  10. Fayomi O. S. I.; Akande I. G. Corrosion mitigation of aluminium in 3.65% NaCl medium using hexamine. J. Bio. Tribo. Corros. 2019, 5, 23 10.1007/s40735-018-0214-4. [DOI] [Google Scholar]
  11. Sherif E.-S. M. corrosion inhibition in chloride solutions of iron by 3-amino-1,2,4-triazole- 5-thiol and 1,1’-thiocarbonyldiimidazole. Int. J. Electrochem. Sci. 2012, 7, 4834–4846. [Google Scholar]
  12. Zhang Z.; Wang F.; Liu Y.; Wu S.; Li W.; Sun W.; Guod D.; Jiang J. Molecule adsorption and corrosion mechanism of steel under protection of inhibitor in a simulated concrete solution with 3.5% NaCl. RSC Adv. 2018, 8, 20648. 10.1039/C8RA03235A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Shahzad K.; Sliem M. H.; Shakoor R. A.; Radwan A. B.; Kahraman R.; Umer M. A.; Manzoor U.; Abdullah A. M. Electrochemical and thermodynamic study on the corrosion performance of API X120 steel in 3.5% NaCl solution. Sci. Rep. 2020, 10, 4314 10.1038/s41598-020-61139-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Velrani S.; Jeyaprabha B.; Prakash P. Inhibition of mild steel corrosion in 3.5% NaCl medium using 1-butyl-3-methylimidazolium chloride. Int. J. Innovative Sci. Eng. Technol. 2014, 1, 57–69. [Google Scholar]
  15. ArockiaSelvi J.; Kamaraj P.; Arthanareeswari M.; PushpaMalini T.; Mohanapriya S.; Subasree N. Effect of cetylpyridinium chloride on corrosion inhibition of mild steel in chloride environment. Mater. Today: Proc. 2019, 14, 264–270. 10.1016/j.matpr.2019.04.146. [DOI] [Google Scholar]
  16. Othman N. K.; Yahya S.; Ismail M. C. Corrosion inhibition of steel in 3.5%NaCl by rice straw extract. J. Ind. Eng. Chem. 2019, 70, 299–310. 10.1016/j.jiec.2018.10.030. [DOI] [Google Scholar]
  17. Suarez M.; Delgado F.; Yanez F.; León J. B.; Prinzio A. D.; González W. A. Study of the electrochemical behavior of mango extract as a corrosion inhibitor of carbon steels by varying the pH of the electrolytic medium. Acta Microsc. 2020, 29, 10–18. [Google Scholar]
  18. Nazari M. H.; Shihab M. S.; Havens E. A.; Shi X. Mechanism of corrosion protection in chloride solution by an apple-based green inhibitor: experimental and theoretical studies. J. Infrastruct. Preserv. Resilience 2020, 1, 1–19. 10.1186/s43065-020-00006-x. [DOI] [Google Scholar]
  19. Shyamvarnan B.; Shanmugapriya S.; Selvi J. A.; Kamaraj P.; Mohankumar R. Corrosion inhibition effect of elettaria cardamomum extract on mild steel in 3.5% NaCl medium. Mater. Today: Proc. 2021, 40, S192–S197. 10.1016/j.matpr.2020.09.085. [DOI] [Google Scholar]
  20. Pradityana A.; Sulistijono; Shahab A.; Noerochim L.; Susanti D. Inhibition of corrosion of carbon steel in 3.5% nacl solution by myrmecodia pendans extract. Int. J. Corros. 2016, 2016, 6058286 10.1155/2016/6058286. [DOI] [Google Scholar]
  21. Dahmani M.; Et - Touhami A.; Al – Deyab S. S.; Hammouti B.; Bouyanzer A. Corrosion inhibition of C38 steel in 1M HCl: a comparative study of black pepper extract and its isolated piperine. Int. J. Electrochem. Sci. 2010, 5, 1060–1069. [Google Scholar]
  22. Amin M. A.; Abd El – Rehim S. S.; El - Sherbini E. E.; Bayoumi R. S. The inhibition of low carbon steel corrosion in hydrochloric acid solutions by succinicacid part I. weight loss, polarization, EIS, PZC, EDX and ESM studies. Electrochim. Acta 2007, 52, 3588–3599. 10.1016/j.electacta.2006.10.019. [DOI] [Google Scholar]
  23. Thomas J. G. N.5th European Symposium on Corrosion Inhibitors; University of Ferrara: Ferrara, Italy, 1981; p 453. [Google Scholar]
  24. Tadros A. B.; Abdel-Naby Y. Inhibition of the acid corrosion of steel by 4-amino-3-hydrazino-5-thio-1,2,4-triazoles. J. Electroanal. Chem. 1988, 246, 433–439. 10.1016/0022-0728(88)80178-5. [DOI] [Google Scholar]
  25. Subramanyam N. C.; Sheshadri B. S.; Mayanna S. M. Thiourea and substituted thioureas as corrosion inhibitors for aluminum in sodium nitrite solution. Corros. Sci. 1993, 34, 563–571. 10.1016/0010-938X(93)90272-I. [DOI] [Google Scholar]
  26. Sherif E.-S. M.; Ahmed A. H. Synthesizing new hydrazone derivatives and studying their effects on the inhibition of copper corrosion in sodium chloride solutions. Synth. React. Inorg., Met.-Org., Nano-Met. Chem. 2010, 40, 365–372. 10.1080/15533174.2010.492546. [DOI] [Google Scholar]
  27. Ahmed A. H.; Hassan A. M.; Gumaa H. A.; Mohamed B. H.; Eraky A. M.; Omran A. A. Copper(II)-oxaloyldihydrazone complexes: Physico-chemical studies; energy band gap and inhibition evaluation of free oxaloyldihydrazones toward the corrosion of copper metal in acidic medium. Arabian J. Chem. 2019, 12, 4287–4302. 10.1016/j.arabjc.2016.05.015. [DOI] [Google Scholar]
  28. Ahmed A. H.; Hassan A. M.; Gumaa H. A.; Mohamed B. H.; Eraky A. M. Physicochemical studies on some selected oxaloyldihydrazones and their novel palladium(II) complexes along with using oxaloyldihydrazones as corrosion resistants. Inorg. Nano-Met. Chem. 2017, 47, 1652–1663. [Google Scholar]
  29. Ahmed A. H.; Hassan A. M.; Gumaa H. A.; Mohamed B. H.; Eraky A. M. Mn2+-complexes of N,O-dihydrazone: structural studies, indirect band gap energy and corrosion inhibition on aluminum in acidic medium. J. Chil. Chem. Soc. 2018, 63, 4159–4168. 10.4067/S0717-97072018000404180. [DOI] [Google Scholar]
  30. Shanbhag A. V.; Venkatesha T. V.; Prabhu R. A.; Kalkhambkar R. G.; Kulkarni G. M. Corrosion inhibition of mild steel in acidic medium using hydrazide derivatives. J. Appl. Electrochem. 2008, 38, 279–287. 10.1007/s10800-007-9436-8. [DOI] [Google Scholar]
  31. Singh A. K.; Chugh B.; Singh M.; Thakur S.; Pani B.; Guo L.; Kaya S.; Serdaroglu G. Hydroxy phenyl hydrazides and their role as corrosion impeding agent: A detail experimental and theoretical study. J. Mol. Liq. 2021, 330, 115605 10.1016/j.molliq.2021.115605. [DOI] [Google Scholar]
  32. Ichchou I.; Larabi L.; Rouabhi H.; Harek Y.; Fellah A. Electrochemical evaluation and DFT calculations of aromatic sulfonohydrazides as corrosion inhibitors for XC38 carbon steel in acidic media. J. Mol. Struct. 2019, 1198, 126898 10.1016/j.molstruc.2019.126898. [DOI] [Google Scholar]
  33. Chugh B.; Singh A. K.; Chaouiki A.; Salghi R.; Thakur S.; Pani B. A comprehensive study about anti-corrosion behavior of pyrazine carbohydrazide: Gravimetric, electrochemical, surface and theoretical study. J. Mol. Liq. 2020, 299, 112160 10.1016/j.molliq.2019.112160. [DOI] [Google Scholar]
  34. Kumbhar C. G.; Sadasivan N. Co(II), Ni(II) and Cu(1I) complexes of oxalic, malonic and terephthalic dihydrazides. Def. Sci. J. 1982, 35, 113–117. 10.14429/dsj.32.620. [DOI] [Google Scholar]
  35. Sherif E.-S. M.; Abdo H. S.; Latief F. H.; Alharthi N. H.; Zein El Abedin S. Fabrication of Ti–Al–Cu new alloys by inductive sintering, characterization, and corrosion evaluation. J. Mater. Res. Technol. 2019, 8, 4302–4311. 10.1016/j.jmrt.2019.07.040. [DOI] [Google Scholar]
  36. Sherif E.-S. M.; Latief F. H.; Abdo H. S.; Alharthi N. H. Electrochemical and spectroscopic study on the corrosion of ti–5al and ti–5al–5cu in chloride solutions. Met. Mater. Int. 2019, 25, 1511–1520. 10.1007/s12540-019-00306-2. [DOI] [Google Scholar]
  37. Feliu S. Jr Electrochemical impedance spectroscopy for the measurement of the corrosion rate of magnesium alloys: brief review and challenges. Metals 2020, 10, 775 10.3390/met10060775. [DOI] [Google Scholar]
  38. Uwaya G. E.; Fayemi O. E.; Sherif E. M.; Junaedi H.; Ebenso E. E. Synthesis, electrochemical studies, and antimicrobial properties of Fe3O4 nanoparticles from callistemon viminalis plant extracts. Materials 2020, 13, 4894 10.3390/ma13214894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Okpara E. C.; Fayemi O. E.; Sherif E. M.; Junaedi H.; Ebenso E. E. Green wastes mediated zinc oxide nanoparticles: synthesis, characterization and electrochemical studies. Materials 2020, 13, 4241 10.3390/ma13194241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kuriakose N.; Kakkassery J. T.; Raphael V. P.; Shanmughan S. K. electrochemical impedance spectroscopy and potentiodynamic polarization analysis on anticorrosive activity of thiophene-2-carbaldehyde derivative in acid medium. Indian J. Mater. Sci. 2014, 2014, 124065 10.1155/2014/124065. [DOI] [Google Scholar]
  41. Sherif E. M.; Park S.-M. Inhibition of copper corrosion in 3.0% NaCl by n-phenyl-1,4-phenylenediamine. J. Electrochem. Soc. 2005, 152, B428 10.1149/1.2018254. [DOI] [Google Scholar]
  42. Sherif E. M.; Park S.-M. Effects of 2-amino-5-ethylthio-1,3,4-thiadiazole on copper corrosion as a corrosion inhibitor in aerated acidic pickling solutions. Electrochim. Acta 2006, 51, 6556–6562. 10.1016/j.electacta.2006.04.047. [DOI] [Google Scholar]
  43. Sherif E. M.; Park S.-M. Inhibition of copper corrosion in acidic pickling solutions by N-phenyl-1,4-phenylenediamine. Electrochim. Acta 2006, 51, 4665–4673. 10.1016/j.electacta.2006.01.007. [DOI] [Google Scholar]
  44. Sherif E. M.; Park S.-M. 2-Amino-5-ethyl-1,3,4-thiadiazole as a corrosion inhibitor for copper in 3.0% nacl solutions. Corros. Sci. 2006, 48, 4065–4079. 10.1016/j.corsci.2006.03.011. [DOI] [Google Scholar]
  45. Zhang Z.; Chen S.; Li Y.; Li S.; Wang L. A study of the inhibition of iron corrosion by imidazole and its derivatives self-assembled films. Corros. Sci. 2009, 51, 291–300. 10.1016/j.corsci.2008.10.040. [DOI] [Google Scholar]
  46. Li W.; Nobe K.; Pearlstein A. J. Potential/current oscillations and anodic film characteristics of iron in concentrated chloride solutions. Corros. Sci. 1990, 31, 615–620. 10.1016/0010-938X(90)90170-A. [DOI] [Google Scholar]
  47. Sabirneeza A. A. F.; Geethanjali R.; Subhashini S. Polymeric corrosion inhibitors for iron and its alloys: a review. Chem. Eng. Commun. 2015, 202, 232–244. 10.1080/00986445.2014.934448. [DOI] [Google Scholar]
  48. Allaoui M.; Rahim O.; Sekhri L. Electrochemical study on corrosion inhibition of iron in acidic medium by moringa oleifera extract. Orient. J. Chem. 2017, 33, 637–646. 10.13005/ojc/330211. [DOI] [Google Scholar]
  49. El-Haddad M. A. M.; Radwan A. B.; Sliem M. H.; Hassan W. M. I.; Abdullah A. M. Highly efcient eco-friendly corrosion inhibitor for mild steel in 5M HCl at elevated temperatures: experimental & molecular dynamics study. Sci. Rep. 2019, 9, 3695 10.1038/s41598-019-40149-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Shainy K. M.; Ammal P. R.; Unni K. N.; Benjamin S.; Joseph A. surface interaction and corrosion inhibition of mild steel in hydrochloric acid using pyoverdine, an eco-friendly biomolecule. J. Bio. Tribo. Corros. 2016, 2, 20 10.1007/s40735-016-0050-3. [DOI] [Google Scholar]

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

RESOURCES