Performance evaluation of triethanolamine as corrosion inhibitor for magnesium alloy in 3.5 wt% NaCl solution

Wei Shang, Chubin He, Yuqing Wen*, Yuanyuan Wang and Zhe Zhang
Guangxi Key Laboratory of Electrochemical and Magnetochemical Function Materials, College of Chemistry and Bioengineering, Guilin University of Technology, Guilin, 541004, China. E-mail: wenyuqing16@163.com

Received 17th September 2016 , Accepted 9th November 2016

First published on 22nd November 2016


Abstract

The corrosion inhibition performance of triethanolamine (TEA) with AZ91D magnesium alloy in 3.5 wt% NaCl solution has been studied using an immersion experiment, electrochemical impedance spectroscopy (EIS), potentiodynamic polarization, surface analysis, and quantum chemical methods. The results revealed that TEA was a good corrosion inhibitor. The polarization plots indicated that TEA served as a mixed-type inhibitor. The adsorption of TEA molecules was found to follow the Langmuir adsorption isotherm, which was a spontaneous, exothermic process of increased entropy. TEA could considerably improve the corrosion resistance of the Mg alloy in 3.5 wt% NaCl.


1. Introduction

As the lightest metal base material, magnesium alloy is a type of metal material that is frequently used in modern industry due to the advantages such as high chemical activity, high specific strength and specific stiffness, good machining stability, and recycling ability.1 Although magnesium alloys have such outstanding advantages, their poor corrosion resistance, caused by the chemical activity of magnesium, has limited their application.2–5

To promote the wider application of Mg alloy, an appropriate treatment can enhance its corrosion resistance, which has important practical significance. The use of a corrosion inhibitor is one of the most effective ways to slow the progress of the metal corrosion by adding it to the corrosive medium in small concentrations.6,7 Organic compounds are often used as corrosion inhibitors to protect media against corrosion, especially those containing a heteroatom with lone-pair electrons (e.g., N, O, P, and S), a benzene ring, or conjugated functional groups.8,9 The lone-pair electrons and π-electrons in these atoms and functional groups are the key to determining the corrosion inhibition performance of the compounds. Under the effects of electrostatic interactions or electrochemical action, these compounds can form a protective film on a metal surface, which can change the structure of the electric double layer and affect the process dynamics of metal corrosion, reducing the corrosion rate of the metal.10

The corrosion inhibitors applied to Mg alloy are mostly heavy metallic salts11,12 and heterocycles.13,14 Although these compounds are effective corrosion inhibitors for Mg alloys, their high toxicity and high cost has limited the production and applications of these types of corrosion inhibitors. Therefore, research on corrosion inhibitors for Mg alloys is very important.

As a common chemical reagent, triethanolamine (TEA) is economical and causes minimal pollution. TEA is a colorless transparent viscous liquid at room temperature. It can dissolve easily in water and ethanol. Besides, TEA has the properties of a tertiary amine and an alcohol, exhibiting significant coordination in synthetic applications. Furthermore, TEA compounds can also exhibit strong coordination ability.15–17 Zhang18 et al. studied the corrosion inhibition performance of TEA with respect to LaFe11.0Co0.7Si1.3 magnetic refrigeration materials in distilled water. It was considered that the TEA molecules were adsorbed on the metal surface by coordination bonding. Ding19 et al. studied the effects of the properties and microstructures of an anode oxide film for Mg alloy in an electrolyte containing TEA. It was indicated that the growth speed of the oxide film on the Mg alloy surface was rapid, due to the addition of TEA, and the film was composed of a surface layer and a transition layer. In a nutshell, these physical and chemical properties of TEA lead to its promising application for preventing corrosion in Mg alloys.

However, research on the corrosion inhibition performance of TEA for Mg alloys is limited, and the detailed study of the corrosion mechanism of TEA as an inhibitor for Mg alloy corrosion is still incomplete. In order to better understand the inhibition mechanism of TEA, this study has systematically studied the corrosion inhibition performance of TEA for Mg alloy in 3.5 wt% NaCl solution by a series of analytical techniques and methods. Electrochemical measurements were used to study the inhibition efficiency and adsorption behavior of the inhibitor. Immersion experiments were used to study the corrosion process with TEA of Mg alloy in 3.5 wt% NaCl solution. SEM and EDX analyses were used to demonstrate the formation of a TEA adsorption film on a Mg alloy surface. Finally, the theoretical parts of this study were carried out using quantum chemical calculations.

2. Experimental

2.1 Experimental material and sample preparation

The metal material used in the experiments was the AZ91D magnesium alloy. The Mg alloy was flaked before use, and the specification was 30 mm × 20 mm × 3 mm. The specific chemical compositions of the Mg alloy are listed in Table 1.
Table 1 AZ91D magnesium alloy chemical compositions and material analysis
Composition Mg Al Zn Cu Si Mn Ni Fe
Content (wt%) 90.8–89.12 8.5–9.5 0.45–0.9 <0.025 <0.05 0.17–0.4 <0.001 <0.004


Before each experiment, the AZ91D magnesium alloy substrates were polished mechanically with 600#, 1000#, and 1400# emery paper until their surfaces were smooth and bright, and then they were degreased and cleaned with ethanol and distilled water using an ultrasonic cleaner, and dried at room temperature. Distilled water was used for washing in each process.

The treated Mg alloy was filled with epoxy resin, with an exposed area of 1 cm2. After the drying step, the Mg alloy was immersed in 3.5 wt% NaCl solution containing different concentrations of TEA. After adsorption for a certain time, the Mg alloy was used for performance testing.

2.2 Performance testing

2.2.1 Electrochemical method. Electrochemical measurements were carried out using a CHI860 electrochemical workstation. The base electrolyte was 3.5 wt% NaCl solution without and with different concentrations of TEA. All experiments were conducted using a conventional three-electrode electrochemical cell with Pt as the counter electrode, a saturated calomel electrode (SCE) as the reference electrode, and the Mg alloy with an exposed area of 1 cm2 as the working electrode.

Before starting the experiments, the open-circuit potential (OCP) was studied at 298 K. The working electrodes were immersed in the test solution for a certain period of time until a steady potential was reached. After the open-circuit potential was stabilized, polarization curves were acquired with a scan rate of 1 mV s−1 in the potential range from EOCP −200 mV to EOCP +200 mV versus SCE at 298 K. Electrochemical impedance spectroscopy (EIS) was performed over a frequency range of 100 kHz to 0.1 Hz at the stable open-circuit potential for each test condition. The amplitude of the AC perturbation was 5 mV.

2.2.2 Surface morphology and component analysis. In order to better observe the TEA adsorption film on the Mg alloy surface, the immersion time was researched using polarization curves in 3.5 wt% NaCl containing 3 mL L−1 TEA at 298 K.

The Mg alloy specimens were prepared as described earlier, and then immersed in the 3.5 wt% NaCl without and with 3 mL L−1 TEA at 298 K for 1 h. Micromorphology analysis was used to demonstrate the formation of the TEA film on the Mg alloy surface with a JSM-5610LV scanning electron microscope. Component analysis of the TEA film was achieved by EDX.

2.2.3 Immersion experiment. The Mg alloy specimens were prepared as described earlier, and then immersed in the 3.5 wt% NaCl containing 3 mL L−1 TEA for 1 to 5 days. After that, these specimens were analyzed by EIS to study the corrosion process of the Mg alloy in the TEA medium, and to set up a corrosion model. EIS was recorded over a frequency range of 100 kHz to 0.1 Hz at the stable open-circuit potential for each test condition. The amplitude of the AC perturbation was 5 mV. All the above tests were carried out at 298 K, and the samples with an exposed area of 10 mm × 10 mm were filled with epoxy resin.
2.2.4 Quantum chemical calculations. The quantum chemical calculations were carried out using the Gaussian 03 programs.20 In this paper, the molecular optimization of TEA was performed with the density functional theory (DFT) method, B3LYP, combined with the 6-31G** (d, p) basis set for all atoms.21 There were several quantum chemical parameters, which are as follows: the energy of the highest occupied molecular orbital (EHOMO), the energy of the lowest unoccupied molecular orbital (ELUMO), ΔE = ELUMOEHOMO, and the dipole moment (μ).

3. Results and discussions

3.1 Electrochemical testing

3.1.1 Potentiodynamic polarization measurements. Polarization measurements could supply information about the type of corrosion inhibitors and the kinetics of corrosion reactions.22,23 The potentiodynamic polarization curves for Mg alloy in 3.5 wt% NaCl solutions in the absence and presence of TEA at different concentrations at 298 K are shown in Fig. 1. According to pertinent literature reports, the cathodic corrosion reaction of Mg alloy in NaCl solution was the hydrogen evolution reaction:24
 
4H2O + 4e → 4OH + 2H2 (1)

image file: c6ra23203e-f1.tif
Fig. 1 Polarization curves of AZ91D magnesium alloy in 3.5 wt% NaCl solutions in the absence and presence of TEA at different concentrations at 298 K.

The anodic corrosion reactions in NaCl solution represented the dissolution process of Mg:25,26

 
image file: c6ra23203e-t1.tif(2)

At the same time, the possibility that magnesium could be transformed to Mg2+ and be dissolved in the solution cannot be ruled out:

 
Mg → Mg2+ + 2e (3)

From what has been discussed above, the total corrosion reaction for Mg alloy in NaCl solution was:27

 
3Mg + 6H2O → 2Mg2+ + 4OH + 2H2 + Mg(OH)2 (4)

As shown in Fig. 1, the cathodic polarization currents of the Mg alloy increased and no passivation happened in the anodic branch in 3.5 wt% NaCl solution due to the occurrence of the cathodic and anodic processes. It was clear that the cathodic and anodic currents were decreased after addition of TEA, showing that the rates of Mg alloy dissolution and hydrogen evolution were decreased due to the addition of TEA.

Table 2 lists the parameters obtained from Fig. 1. As shown in Table 2, the corrosion potential shifted toward the positive direction due to the addition of TEA, showing that the anti-corrosion tendency was improved. Besides, it was clear that the corrosion current densities of Mg alloy exhibited an extreme concentration phenomenon as the TEA concentration increased. The corrosion current density was gradually reduced as the TEA concentration increased from 1 mL L−1 to 3 mL L−1; the minimum corrosion current density was observed when the concentration of the TEA was 3 mL L−1; the corrosion current density then increased with a further increase in the TEA concentration. In addition, the corrosion current density of the Mg alloy in the media without TEA was 9.039 × 10−6 A cm−2, and the corrosion current density of the Mg alloy in 3 mL L−1 TEA was 6.583 × 10−7 A cm−2. Comparing the above values, it was found that the current density was reduced by 1 order of magnitude. This phenomenon indicated that TEA could suppress the dissolution of Mg alloy and exhibited a good corrosion-inhibiting performance for Mg alloy. On the other hand, it can be seen in Fig. 1 and Table 2 that the cathodic reaction of Mg alloy corrosion was significantly inhibited in the presence of TEA. Moreover, the cathodic current–potential curves were similar to the almost parallel Tafel lines. This meant that the addition of TEA did not change the mechanism of hydrogen evolution.28 According to the anodic parts of the curves, the anodic reaction of Mg alloy corrosion was markedly suppressed in the presence of TEA. The inhibitor exhibited a stronger inhibitive effect on the anodic reaction than on the cathodic one, indicating that TEA acted as a mixed-type inhibitor to mainly inhibit the anodic reaction of Mg alloy in 3.5 wt% NaCl solution.28,29

Table 2 Potentiodynamic polarization curves of different concentrations: data analysis
Concentration (mL L−1) Ecorr (V vs. SCE) βc (V per decade) βa (V per decade) icorr (μA cm−2) η (%)
Bare −1.534 7.141 10.164 9.039
1 −1.510 7.546 10.118 3.293 63.57
2 −1.508 7.481 9.966 1.290 85.73
3 −1.488 7.343 9.130 0.6583 92.71
4 −1.496 7.684 10.084 2.513 72.20
5 −1.516 7.720 10.117 3.970 56.08


The inhibition efficiency, η (%), was calculated from eqn (5) and the results are listed in Table 2:

 
image file: c6ra23203e-t2.tif(5)
where ic is the current density of the Mg alloy surface without the inhibitor, and ii is the current density of the Mg alloy surface with the inhibitor. The obtained data revealed that the inhibition efficiency could exhibit an extreme phenomenon as the TEA concentration increased. Besides, the inhibition efficiency reached a maximum when the concentration of TEA was 3 mL L−1, and this value was 92.71%. This illustrated that TEA could be classified as an adsorptive corrosion inhibitor for Mg alloy, because TEA was adsorbed on the Mg alloy to form a protective film and block the active corrosion sites. In addition, the extreme concentration phenomenon could be explained as follows: firstly, TEA could adsorb on the active sites on the surface of the Mg alloy to inhibit corrosion of the alloy to some extent when the TEA concentration was low (e.g., 1 mL L−1, 2 mL L−1). However, TEA might be unevenly distributed on the Mg alloy surface due to the small molecular volume of TEA, so the Mg alloy surface could form a small part of the active area. This was easy to corrode in the corrosive medium, so the corrosion current could be increased. Secondly, when the TEA concentration was 3 mL L−1, TEA could evenly adsorb on the Mg alloy surface and cover its active sites. It could form a stable TEA adsorption film to effectively restrain the corrosion of Mg alloy. Therefore, the corrosion current was at a minimum. Thirdly, excess adsorption of TEA might occur, resulting in an agglomeration or shedding phenomenon under higher concentrations (e.g., 4 mL L−1, 5 mL L−1). The TEA film was unstable at that time, resulting in reduced protection for the Mg alloy. So, the corrosion current could also be increased.

3.1.2 Electrochemical impedance spectroscopy (EIS). The Nyquist plots for Mg alloy in 3.5 wt% NaCl solutions in the absence and presence of TEA at different concentrations at 298 K are shown in Fig. 2. Compared with the diagram, the EIS was markedly changed due to the adsorption of TEA. As shown in Fig. 2, the impedance diagram for the blank showed a capacitive arc at high frequency and an inductive arc at low frequency. The high-frequency semicircle was attributed to the charge transfer and double-layer capacitance. The inductive arc at low frequency was probably related to the solution of Mg alloy species due to local surface breakage or the low adsorption capacity of the surface film.
image file: c6ra23203e-f2.tif
Fig. 2 Nyquist impedance spectra of AZ91D magnesium alloy in 3.5 wt% NaCl solutions in the absence and presence of TEA at different concentrations at 298 K.

The EIS curves for different TEA concentrations were composed of double capacitive arcs. The capacitive arc at high frequency was related to charge transfer, and reflected the corrosion resistance of the sample. The capacitive arc at low frequency was caused by the adsorbing–desorbing process of the corrosion inhibitor on the electrode surface.30 In addition, compared with the blank sample, the semicircle diameter of the impedance arc at the high-frequency part was increased in the presence of TEA. This indicated that the charge-transfer resistance became dominant in the corrosion process due to the adsorption of TEA. Meanwhile, the capacitive arc at high frequency increased to a maximum when the concentration of the TEA was 3 mL L−1, and then decreased as the additive TEA concentration increased. This showed that an extreme concentration phenomenon occurred, which was in accord with the obtained results of the polarization curve.

The equivalent-circuit model was used to fit the capacitance part of all the EIS (shown in the Nyquist spectra). All the EIS were interpreted by means of the equivalent circuits, which are shown in Fig. 3. The fitting results were obtained with ZView2 software, including the solution resistance (Rs), the film resistance (Rcoat), the film capacitance (Qcoat), the charge-transfer resistance (Rct), the interface capacitance (Qdl), and the double-layer capacitance (CPEdl). The CPE was defined by the values of Q and n, which were used to make up for the inhomogeneities in the electrode surface and described the nature of the Nyquist semicircle.31 The admittance and impedance of CPE were defined by eqn (6):

 
image file: c6ra23203e-t3.tif(6)
where Y0 is the modulus, ω is the angular frequency, and n is the deviation parameter.


image file: c6ra23203e-f3.tif
Fig. 3 Simulation equivalent fitting circuit diagrams: (a) bare, (b) different concentrations of TEA.

The inhibition efficiencies η (%) of different TEA concentrations were calculated by the charge-transfer resistance according to eqn (7):32

 
image file: c6ra23203e-t4.tif(7)
where R0ct and Rct are the film resistances for the blank and different concentrations, respectively. The values of CPE capacitance were calculated by using eqn (8):33
 
Cdl = Y01/nRct(1−n)/n (8)
where Y0 and n are the magnitude of the CPE and deviation parameter.

These results are listed in Table 3. In Table 3, it can be seen that the Rcoat values of all TEA concentrations were greater than the value for the blank, which indicated that TEA could adsorb on the Mg alloy surface to form a protective film, which provided better protection for Mg alloy. Besides, the values of the Rcoat increased with the additive TEA concentrations, and then decreased. The value reached a maximum when the concentration of the TEA was 3 mL L−1. The reasons for this result are as follows. At low concentrations (e.g., 1 mL L−1), the TEA can still adsorb on the Mg alloy surface, but the formation of the protective film was flawed, so the corrosive ions could easily penetrate the protective film to corrode the Mg alloy. As the TEA concentration increased, the protective film became more complete, so the values increased (e.g., 2 mL L−1). Moreover, at high concentrations (e.g., 4 mL L−1, 5 mL L−1), excess adsorption of TEA might occur, resulting in an agglomeration or shedding phenomenon. Thus, film-free or broken areas could exist.14 The corrosive ions could adsorb more easily to the film-free or broken areas to form soluble ionic metal compounds.34,35 As the dissolution rates of the TEA film increased, the Rcoat values decreased. Under the optimum concentration (e.g., 3 mL L−1), TEA molecules could evenly adsorb on the Mg alloy surface to form a protective film, and continue to adsorb on the protective film to fill film-free or broken areas. So the value of the Rcoat reached a maximum at 3 mL L−1 TEA.

Table 3 Simulation parameters of equivalent circuit diagrams for different concentrations
Concentration (mL L−1) Rs (Ω cm2) Qcoat (μF cm−2) n Rcoat (Ω cm2) Qdl (μF cm−2) Rct (Ω cm2) η (%)
Bare 15.53 16.738 0.86761 609.6
1 21.5 16.873 0.85516 3844 316.36 103.53 84.14
2 27.31 11.445 0.89351 4433 109.39 1746 86.25
3 27.53 7.3509 0.91128 7881 601.04 2471 92.26
4 28.69 10.672 0.90424 4147 1052.3 1464 85.30
5 39.57 16.022 0.8549 2585 2588.3 588.6 76.42


The Cdl values of all TEA concentrations were smaller than the value for the blank. According to the Helmholtz model, the double-layer capacitance was inversely proportional to the surface changes:32,36

 
image file: c6ra23203e-t5.tif(9)
where d is the film thickness, S is the electrode surface area, ε0 is the permittivity of air, and ε is the local dielectric constant. Because of the formation of an effective TEA adsorption film, the active sites of the Mg alloy surface were replaced by TEA molecules. Meanwhile, the dielectric constant of TEA molecules was less than that of water. Thus, there were smaller double-layer capacitances in the TEA system. Another probable reason for the decrease in the Cdl values was that the surface area was reduced.32 At lower concentrations (e.g., 1 mL L−1, 2 mL L−1), the TEA protective film was flawed due to the lower TEA concentration, the alloy surface area exposed to the corrosive medium was increased, and reactive sites were also increased. The Mg alloy could therefore be corroded in the corrosive medium. However, the protective film became more even as the TEA concentration increased. Therefore, the Cdl values decreased. Similarly, at higher concentrations (e.g., 4 mL L−1, 5 mL L−1), the protective film might contain film-free or broken areas due to excess adsorption. The roughness of the film increased, and the film became unstable. The active areas of Mg alloy exposed to the corrosion medium were increased, so the Cdl values increased. Under the optimum concentration (e.g., 3 mL L−1), the protective film was stable, and TEA could fill defects in the protective film by reabsorption.14 Therefore, the alloy surface area exposed to the corrosive medium was reduced, together with the reactive sites, so the Cdl value was at a minimum.

It could be seen that the inhibition efficiencies were closely related to the TEA concentrations.37 There existed an optimal concentration of TEA, below which there was a decrease in the inhibition efficiency, and above which there was also a decrease in the inhibition efficiency. This was consistent with the results of the potentiodynamic polarization measurements. As shown in Table 3, the protection of the magnesium alloy was most effective at 3 mL L−1 TEA. Therefore, 3 mL L−1 TEA was the optimal concentration.

3.2 Surface morphology and component analysis

3.2.1 Effect of immersion time. In order to better observe the TEA adsorption film on the Mg alloy surface, the immersion time was researched by polarization curves in 3.5 wt% NaCl containing 3 mL L−1 TEA at 298 K. The results are shown in Fig. 4. Table 4 lists the corresponding parameters obtained from Fig. 4. In addition, the inhibition efficiencies were calculated using eqn (5). The values of the coverage (θ) were calculated using eqn (10):38
 
image file: c6ra23203e-t6.tif(10)
where ic is the current density of the Mg alloy surface without the inhibitor, and ii is the current density of the Mg alloy surface with the inhibitor. It should be noted that the inhibition efficiency η (%) is a function of θ: η (%) = θ × 100%.

image file: c6ra23203e-f4.tif
Fig. 4 Polarization curves of AZ91D magnesium alloy in 3.5 wt% NaCl solutions containing 3 mL L−1 TEA with different immersion times at 298 K.
Table 4 Potentiodynamic polarization curves of Fig. 4: data analysis
Immersion time (h) Ecorr (V vs. SCE) icorr (μA cm−2) θ η (%)
0.5 −1.488 0.9403 0.89597 89.60
1 −1.494 0.7257 0.91971 91.97
2 −1.478 1.223 0.86470 86.47
3 −1.471 2.324 0.74289 74.29
4 −1.472 4.141 0.54187 54.19


It can be seen from Fig. 4 and Table 4 that the current density was reduced before immersing for 1 hour, which indicated the formation of a TEA protective film. The current density was then increased, which indicated that the TEA film could be destroyed due to multilayer adsorption of TEA or the diffusion of the corrosive ions. As regards the coverage, the values increased with the immersion time, and then decreased. Similarly, the inhibition efficiency behaved in a similar way. According to Table 4, both the coverage of TEA and the inhibition efficiency were greatest after immersing for 1 hour, so the optimal immersion time was 1 hour, and the inhibition efficiency was 91.97%.

3.2.2 Morphological studies. The Mg alloy was immersed in the different TEA systems at 298 K for 1 hour. SEM was used to characterize the surface morphology, as shown in Fig. 5. The different TEA systems were 3.5 wt% NaCl solution without TEA (a), distilled water with 3 mL L−1 TEA (b), and 3.5 wt% NaCl solution with 3 mL L−1 TEA (c), as shown in Fig. 5.
image file: c6ra23203e-f5.tif
Fig. 5 SEM images of AZ91D magnesium alloy distilled in different TEA-systems for 1 h at 298 K: (a) 3.5 wt% NaCl solution without TEA, (b) distilled water with 3 mL L−1 TEA, (c) 3.5 wt% NaCl solution with 3 mL L−1 TEA.

Fig. 5(a) shows that NaCl adsorption could occur at the Mg alloy surface in the absence of inhibitor after immersion for 1 hour, and there may be corrosion products produced near the NaCl adsorption. It can be seen in Fig. 5(b) that a TEA adsorption film was produced on the Mg alloy surface and the surface was even with no obvious defects. It was inferred that TEA could effectively adsorb on the Mg alloy surface. The square area in Fig. 5(b) was magnified 5 times. It could be seen that the TEA particles had a slender shape, which ensured that TEA could adsorb in excess on the Mg alloy surface. As shown in Fig. 5(c), a TEA adsorption film was also observed on the Mg alloy surface, and the surface was even with no obvious defects. However, the white part of Fig. 5(c) may represent a reunion phenomenon of TEA, which indicated that TEA exhibited a slight reunion phenomenon at this concentration. This could also explain why the corrosion effect could be reduced under higher concentrations.

3.2.3 Component analysis. The components of the TEA adsorption film on the Mg alloy surface are shown in Fig. 6. It could be seen that, besides the matrix elements, Mg, Al, and Si, other elements, C, O, and N, were also presented on the surface. The elements C, O, and N belonged to the TEA. Besides, it can also be seen in Fig. 6 that each element was distributed evenly, which illustrated that the TEA was adsorbed homogeneously on the Mg alloy surface. According to the above, the TEA can be adsorbed on the Mg alloy surface, resulting in the formation of an effective passive film between the substrate and the corrosive medium.
image file: c6ra23203e-f6.tif
Fig. 6 Surface elemental distribution of sample.

3.3 Corrosion behavior study

In order to study the corrosion process of Mg alloy in the TEA medium and set up a corrosion model, EIS was studied for different immersion times. The Nyquist impedance spectra with different immersion times are shown in Fig. 7. Three spectra were selected from all the EIS (including 1 hour, 3 days, and 5 days) and were interpreted by means of the equivalent circuit, as shown in Fig. 8. After being immersed for 1 hour, a TEA adsorption film was formed on the Mg alloy surface, which can increase the charge-transfer resistance and effectively prevent the aggressive ions from accessing the substrate surface. Besides, the electron transfer process has occurred. Meanwhile, TEA could also fill the defects in the protective film by readsorption.14 Thus, the alloy surface area exposed to the corrosive medium was reduced, and reactive sites were also reduced. Therefore, the protection of the Mg alloy was most effective at this time. It can be seen from Fig. 7 that the semicircle diameters of the impedance arcs were reduced as the immersion time increased, which indicated that the polarization resistance decreased gradually, and the TEA film was damaged in the corrosive medium. Moreover, it is also shown in Fig. 7 that the EIS curves were composed of double capacitive arcs from 1 hour to 4 hours. Then, the impedance diagrams showed a capacitive arc at high frequency and an inductive arc at low frequency from the first day to the fifth day. And at low frequency, the impedance diagrams indicated an inductance contraction phenomenon from the fourth day to the fifth day. This indicated that the interfacial properties and surface state had undergone significant changes, involving the Mg alloy surface/TEA adsorption film/corrosive medium interface. The reason for this phenomenon might be that the chloride ions adsorbed to the local active sites of the TEA film, and formed soluble complex ions with metal ions.39,40 Therefore, the electrode surface was uneven and unstable, and the dissolution rate of the TEA film increased. Therefore, the polarization resistances decreased. Meanwhile, ion migration resistance in the film was increased.40 Local areas of the TEA film constantly thinned, forming corrosion-induced holes. Thus, the appearance of the inductive arc at low frequency in the equivalent circuit was because of chloride ion adsorption to the TEA film, forming soluble complex ions.35 Due to the autocatalytic process of pitting corrosion, thinning of the TEA film continued,40 the corrosion current increased, and the polarization resistance decreased. Thus, impedance arcs were seen at low frequency, and the inductance contraction phenomenon and the impedance arcs were weakened as the immersion time increased. After immersion for 120 hours, the TEA film could be detached because the degree of chloride erosion was increased. Many corrosive products collected on the surface, but the film was loosely packed and was of a poor quality.48
image file: c6ra23203e-f7.tif
Fig. 7 Nyquist impedance spectra with different immersion times.

image file: c6ra23203e-f8.tif
Fig. 8 Simulation equivalent fitting circuit diagram: (a) 1 hour, (b) 72 hours and 120 hours.

The equivalent circuit model was used to fit the capacitance part of all the EIS (shown in Nyquist spectra). The fitting results were obtained with ZView2 software, including the solution resistance (Rs), the film resistance (Rcoat), the film capacitance (Qcoat), the charge-transfer resistance (Rct), and the interface capacitance (Qdl).

These results are listed in Table 5. From Table 5, it can be seen that the values of the Rcoat gradually decreased as the immersion time increased from 1 hour to 120 hours, while the double-layer capacitance was changed remarkably. Several reasons for these results are listed as follows. First and foremost, the relationship between the immersion time and the coverage of TEA on the Mg alloy surface could be obtained according to the results of 3.2.1, and the fitting obtained through a function of two variables, as shown in Fig. 9. Fig. 9 shows that the correlation relation of the immersion time and the coverage coincided with the binary function because the correlation coefficient (R2) was very close to 1. So, it could be deduced from Fig. 9 that the coverage of TEA on the magnesium alloy surface was the greatest near 1 hour. In other words, the adsorption of TEA tended to a saturation state near 1 hour. So, the TEA adsorption film showed an excellent protection performance. Because of the formation of an effective TEA adsorption film, the active sites of the Mg alloy surface were replaced by TEA molecules. And the dielectric constant of the TEA molecules was less than that of the water. Thus, the film resistance was the largest and the double-layer capacitance was the smallest. Meanwhile, the TEA and corrosive ions still adsorbed constantly on the Mg alloy surface. The Mg alloy passed through a corrosion induction phase.

Table 5 Simulation parameters of equivalent circuit diagram with different immersion times
Immersion time (h) Rs (Ω cm2) Qcoat (μF cm−2) n Rcoat (Ω cm2) Qdl (μF cm−2) n Rct (Ω cm2)
1 25.55 8.3453 0.9242 12[thin space (1/6-em)]420 5.24 × 10−4 0.939 6695
72 33.96 47.218 0.7468 3464
120 24.27 21.59 0.8407 624.2



image file: c6ra23203e-f9.tif
Fig. 9 The relationship between the immersion times and the coverages of TEA.

In addition, the quality of the TEA film decreased as the immersion time increased, due to the multilayer adsorption of TEA. At that time, the corrosive ions more easily penetrated the TEA film to reach the Mg alloy surface, because the diffusion rate of the corrosive ions was faster than the formation of the TEA film. These factors were likely to lead to the loss of TEA film, so the film resistance was decreased. The alloy surface area exposed to the corrosive medium was increased due to the destruction of TEA film, so the film capacitance increased sharply. Then, the Mg alloy passed through a rapid local corrosion process.

Furthermore, after a period of corrosion, the corrosion products covered the Mg alloy surface and had little protection, so the film capacitance was decreased. However, the film was loosely packed, and was of a very poor quality. So, the corrosion reaction continued and the film resistance was further decreased. Then, the Mg alloy passed through a slow general corrosion process.

On the other hand, n is an apparent diffusion coefficient to characterize the electric double-layer capacitance of the electrolyte/Mg alloy interface. At the ideal capacitance, n = 1.41,42 In general, the rough porous surface causes the electric double-layer capacitance to deviate from the pure ideal capacitance.43 From Table 5, after immersion for 1 hour, n was most favorable, indicating that the surface was relatively uniform, while n was least favorable after immersion for 72 hours, possibly because the occurrence of pitting corrosion resulted in many corrosion holes on the Mg alloy surface. After immersion for 120 hours, n was increasing, rather than reducing, which could be because the corrosion products were formed at the defect.

3.4 Adsorption studies

3.4.1 Adsorption model. In order to investigate the interactions between the TEA and Mg alloy surface, the adsorption isotherms were studied. It was assumed that the adsorption of TEA molecules on the Mg alloy surface conformed to the Langmuir isotherm equation, eqn (11):
 
image file: c6ra23203e-t7.tif(11)
where Kads is the standard adsorption equilibrium constant, C is the inhibitor concentration, and θ is the surface coverage defined as η (%)/100 at different concentrations of TEA obtained from polarization measurements. Experientially, the applied assumption is that θη (corrosion inhibition efficiency).44 The corrosion inhibition efficiency and the surface coverage were calculated using eqn (5) and (12), respectively:
 
image file: c6ra23203e-t8.tif(12)
where η is the corrosion inhibition efficiency with different concentrations, and ηm is the maximum corrosion inhibition efficiency.

The plots of image file: c6ra23203e-t9.tif against C yielded straight lines, as shown in Fig. 10. Both the linear correlation coefficient (R2) and other parameters are given in Table 6 at different temperatures. Fig. 10 and Table 6 show that the correlation relation of image file: c6ra23203e-t10.tif and C coincided with the linear relationship, because the linear correlation coefficient (R2) was very close to 1, indicating that the adsorption of TEA obeyed the Langmuir adsorption isotherm at different temperatures; in other words, the TEA molecules formed a single-molecule adsorbed layer on the Mg alloy surface. Generally, a higher value of Kads is associated with strong adsorption.20 This indicated that the corrosion inhibition efficiency of TEA was higher: 98.26% (20 °C, 3 mL L−1) > 95.81% (30 °C, 3 mL L−1) > 92.55% (40 °C, 3 mL L−1) > 89.42% (50 °C, 3 mL L−1) > 86.01% (60 °C, 3 mL L−1). In this study, the value of Kads obeyed the order at different temperatures: 293 K (20 °C) > 303 K (30 °C) > 313 K (40 °C) > 323 K (50 °C) > 333 K (60 °C), which is in accordance with the order of the corrosion inhibition efficiency.


image file: c6ra23203e-f10.tif
Fig. 10 Relationship between image file: c6ra23203e-t11.tif and C at different temperatures.
Table 6 image file: c6ra23203e-t12.tifC linear fitting parameters at different temperatures
Temperatures (T) η (%) Kads R2 ln[thin space (1/6-em)]Kads
1 mL L−1 2 mL L−1 3 mL L−1 4 mL L−1 5 mL L−1
293 K (20 °C) 91.06 93.30 98.26 96.01 95.34 3333.33 0.999 8.112
303 K (30 °C) 89.42 91.06 95.81 92.16 88.17 1666.67 0.9957 7.419
313 K (40 °C) 87.06 89.31 92.55 87.49 84.24 1000 0.9956 6.908
323 K (50 °C) 85.45 87.57 89.42 83.81 80.29 666.67 0.9948 6.502
333 K (60 °C) 83.28 85.12 86.01 79.43 74.09 416.67 0.9907 6.032


3.4.2 Calculation of adsorption thermodynamic parameters. In order to further study the adsorption mechanism of TEA on the Mg alloy surface, the adsorption and thermodynamic behaviors were analyzed with the van't Hoff equation:45
 
image file: c6ra23203e-t13.tif(13)

The plots of ln[thin space (1/6-em)]Kads against image file: c6ra23203e-t14.tif yielded straight lines, as shown in Fig. 11. The adsorption enthalpy (ΔHadsθ) could be calculated from the straight slope. The standard Gibbs free energy of adsorption (ΔGadsθ) and the standard adsorption entropy (ΔSadsθ) could be calculated using eqn (14) and (15), respectively:45,46

 
ΔGadsθ = −RT[thin space (1/6-em)]ln(55.5Kads) (14)
 
ΔGadsθ = ΔHadsθTΔSadsθ (15)
where Kads is the standard adsorption equilibrium constant, the value 55.5 is the molecular concentration of water in solution in mol L−1, R is the gas equilibrium constant, and T is the temperature.


image file: c6ra23203e-f11.tif
Fig. 11 The relationship between ln[thin space (1/6-em)]K and 1/T.

The adsorption thermodynamic parameters of TEA at different temperatures are listed in Table 7. Eqn (14), was used to calculate the values of ΔGadsθ from −29.544 to −27.820 kJ mol−1. The obtained negative value of ΔGadsθ showed that the TEA was spontaneously adsorbed on the Mg alloy surface. Generally, for values of ΔGadsθ up to −20 kJ mol−1, the adsorption is regarded as physisorption, while values around −40 kJ mol−1 or smaller were associated with chemisorption.47 So, TEA was adsorbed on the Mg alloy surface in 3.5 wt% NaCl solutions at different temperatures by physical adsorption and chemical adsorption, because of the electrostatic interactions between the electric charges of TEA and the Mg alloy surface, and the formation of coordination bonds at the interface between the TEA and the vacant d-orbitals of magnesium atoms.45 Besides, the obtained negative value of ΔHadsθ (−41.308 kJ mol−1) indicated that the adsorption process was exothermic on the Mg alloy surface in 3.5 wt% NaCl solutions, so the effect of the corrosion inhibitor decreased when the temperature increased. ΔSadsθ > 0 indicated that TEA adsorption on the Mg alloy surface was driven by an increase in entropy.

Table 7 The adsorption thermodynamic parameters of TEA at different temperatures
Temperatures, K ΔHadsθ, kJ mol−1 ΔGadsθ, kJ mol−1 ΔSadsθ, J mol−1 K−1
293 (20 °C) −41.308 −29.544 86.734
303 (30 °C) −28.806 81.436
313 (40 °C) −28.427 77.624
323 (50 °C) −28.247 74.663
333 (60 °C) −27.820 71.139


3.5 Quantum chemical calculations

In order to clearly understand the adsorption mechanism of TEA on the Mg alloy surface, a quantum chemical calculation was considered to be an effective method, as this can provide several theoretical parameters. The structure of TEA was optimized, and it was ensured that the energy of the molecular system was minimum. The parameters of the geometry and Mulliken charge distribution of TEA were obtained after optimization, and are shown in Fig. 12. Fig. 12 shows that, after optimization, the TEA was a three-dimensional structure due to the existence of an induced effect on the TEA. From the Mulliken charge distribution, it could be seen that the negative charge resided mainly at the N and O atoms for TEA. Due to the existence of the induced effect on the TEA, the C atoms connecting with the N and O atoms had small amounts of negative charge.
image file: c6ra23203e-f12.tif
Fig. 12 The parameters of the geometry and Mulliken charge distribution of GO and TEA.

The Frontier molecular orbital theory48 considers that electron transfer can occur between the Frontier molecular orbitals when the reactants react with each other. Therefore, in order to analyze the adsorption process of TEA on the Mg alloy surface, the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of the TEA molecules must be considered, and these are shown in Fig. 13. It can be seen from Fig. 13 that the electron cloud of the HOMO orbit of the TEA is mainly distributed on the N and O atoms, which indicates that these atoms could be the major atoms to provide the electrons with the empty d-orbitals on the Mg alloy surface to form the coordinate bonds. The electron cloud of the LUMO orbital of the TEA focused on the C atoms, which indicated that the C atoms could accept electrons from the Mg alloy surface to form back-donating bonds.


image file: c6ra23203e-f13.tif
Fig. 13 The Frontier molecular orbitals of TEA.

The calculated quantum chemical data were also reported, including the energy of the highest occupied molecular orbital (EHOMO), the energy of the lowest unoccupied molecular orbital (ELUMO), the energy gap (ΔE = ELUMOEHOMO), the dipole moment (μ), the absolute electronegativity (χ), the global hardness (γ), and the fraction of the transferred electrons (ΔN). The Frontier molecular orbital energies, EHOMO and ELUMO, were associated with the ionization potential (I) and the electron affinity (A) of the magnesium atoms and the inhibitor molecules:49

 
I = −EHOMO and A = −ELUMO (16)

The absolute electronegativity (χ) and the global hardness (γ) of the inhibitor molecule were approximated by eqn (17) and (18):50

 
image file: c6ra23203e-t15.tif(17)
 
image file: c6ra23203e-t16.tif(18)

Therefore, the fraction of electrons transferred from the inhibitor to the metallic surface (ΔN) was calculated using eqn (19):51,52

 
image file: c6ra23203e-t17.tif(19)
where χMg and γMg are the absolute electronegativity and global hardness of the Mg atom, and χinh and γinh are the absolute electronegativity and global hardness of the TEA molecules, respectively. In order to calculate the fraction of the electrons transferred, the theoretical values of IMg (7.661 eV), AMg (−1.717 eV), χMg (2.972 eV), and γMg (4.689 eV) were used.53 According to eqn (16), the values of EHOMO and ELUMO of an Mg atom could be approximated to the negative values of the ionization potential and the electron affinity of an Mg atom, respectively. These results are presented in Table 8.

Table 8 Quantum chemical parameters calculated using the B3LYP method with a 6-31G** (d, p) basis set for TEA and Mg
  EHOMO (eV) ELUMO (eV) ΔE (eV) μ (Debye) χ γ ΔN
TEA −0.21173 0.06911 0.28084 3.8676 0.07131 0.14042 0.300316
Mg −7.661 1.717 9.378 1.31 2.972 4.689


EHOMO usually describes the electron-donating capability of the molecule, and ELUMO is related to the capability of the molecule to accept electrons.54,55 As shown in Table 8, EHOMO(TEA) > EHOMO(Mg), and ELUMO(Mg) > ELUMO(TEA), which demonstrated that the electron-donating capability of TEA was better than that of Mg atoms, and the capability of TEA to accept electrons was also better than that of Mg atoms. The energy gap (ΔE = ELUMOEHOMO) described the stability of the molecule.56,57 The stability of the molecule was inversely proportional to the energy gap; in other words, the smaller the molecular energy gap, the easier it was to take part in a chemical reaction. This showed that it was easier for TEA to provide the electrons with the empty orbital of the Mg alloy and accept the electrons from the Mg alloy surface to form back-donating bonds and a relatively stable adsorption layer on the Mg alloy surface. Besides, μTEA > μMg, and γTEA < γMg, which illustrated that TEA was activated and was easier to polarize and react with the Mg alloy in 3.5 wt% NaCl solution.58,59 Meanwhile, χTEA > χMg, which illustrated that TEA could more easily provide the electron to form the chemical bond with the Mg alloy surface.60 The ΔN value described the inhibition achieved from electron donation. If ΔN < 3.6, the inhibition efficiency was increased, with increasing ability to donate electrons to the metal surface.61,62 It could be observed from Table 8 that ΔNTEA > ΔNGO, and their ΔN was less than 3.6, which showed that TEA had a stronger ability to donate electrons, and the TEA adsorption film could effectively inhibit the corrosion of Mg alloy.

In order to confirm the active sites, the Fukui function of the TEA molecule was analyzed.63 The Mulliken charges and Fukui indices of selected atoms for the inhibitor TEA are listed in Table 9. The Fukui functions were calculated using eqn (20) and (21):64,65

 
f+ = q(N+1)q(N) (20)
 
f = q(N)q(N−1) (21)
where q(N+1), q(N), and q(N−1) are the charges of the atoms on the systems with N + 1, N, and N − 1 electrons, respectively; f+ is the electrophilic Fukui index, and f is the nucleophilic Fukui index.

Table 9 Mulliken charge distribution and Fukui indices of selected atoms for TEA inhibitor
  q(N) f+ f
N1 −0.444330 1.201373 −0.443282
C2 −0.045497 0.027492 −0.088348
C3 0.062546 −0.034712 0.024019
C8 −0.045589 0.027734 −0.089036
C9 0.062753 −0.034481 0.022357
C14 −0.045615 0.027539 −0.08862
C15 0.062735 −0.035405 0.023663
O20 −0.544190 0.557334 −0.520783
O22 −0.544205 0.557190 −0.520591
O24 −0.544267 0.557419 −0.520066


From Table 9, it is clear that N1 had more negative charge, f+ was larger, and the electron cloud was mainly distributed on the N atom, so N1 showed clear nucleophilicity and could more easily provide the electrons with empty d-orbitals to form strong chemical bonds on the Mg alloy surface. O20, O22, and O24 also had a negative charge, while their nucleophilic Fukui indices were less than those of N1. They could form weak chemical bonds. Although the C atoms had small amounts of negative charge, their nucleophilic Fukui indices were low, so these C atoms were not the nucleophilic active sites. Although the heteroatom on the TEA could provide the electrons with empty d-orbitals to form the coordinate bonds on the Mg alloy surface, the Mg alloy could also provide electrons to the TEA to form back-donating bonds. From Table 9, it can be seen that the electrophilic Fukui indices of the C3, C9, and C15 atoms were larger than those of the others, and the electron cloud of the LUMO orbit of the TEA focused on these C atoms, so these could be the electrophilic activity sites, forming back-donating bonds with the Mg alloy. To sum up, the TEA inhibitor achieved polycentric adsorption on the Mg alloy surface via chemical bonds and back-donating bonds, which prevented the corrosive medium from transferring to the substrate surface.

4. Conclusion

According to the experimental results and analysis, the results obtained lead to the following conclusions:

(1) The inhibition efficiencies were closely related to the TEA concentrations. The value of 3 mL L−1 TEA was the optimal concentration; lower concentrations caused a decrease in the corrosion current and the double-layer capacitance, and an increase in the film resistance with the TEA concentration; higher concentrations led to a decrease in the film resistance, and an increase in the corrosion current and the double-layer capacitance with the TEA concentration.

(2) The Langmuir isotherm equation could best describe the adsorption behavior of the TEA molecules on the Mg alloy surface, which was a spontaneous, exothermic process of increased entropy.

(3) The corrosion process with TEA of the Mg alloy went through four steps in 3.5 wt% NaCl solution, including a TEA adsorption process, a corrosion induction phase, a rapid local corrosion process, and a slow general corrosion process.

(4) TEA exhibited a good corrosion-inhibiting performance for the Mg alloy in 3.5 wt% NaCl solution by physical adsorption and chemical adsorption to form an evenly adsorbed film on the Mg alloy surface, because the TEA inhibitor could react with Mg alloy to form chemical bonds with impurity atoms, such as N and O atoms, and some C atoms could form back-donating bonds with the substrate.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (No. 51664011 and No. 51665010) and the Guangxi Natural Science Foundation of China (No. 2014GXNSFAA118335).

References

  1. N. G. Wang, R. C. Wang, C. Q. Peng, C. W. Hu, Y. Feng and B. Peng, Research progress of magnesium anodes and their applications in chemical power sources, Trans. Nonferrous Met. Soc. China, 2014, 24, 2427–2439 CrossRef CAS.
  2. D. Cao, L. Wu, G. Wang and Y. Lv, Electrochemical oxidation behavior of Mg–Li–Al–Ce–Zn and Mg–Li–Al–Ce–Zn–Mn in sodium chloride solution, J. Power Sources, 2008, 183, 799–804 CrossRef CAS.
  3. Y. Yang, F. Scenini and M. Curioni, A study on magnesium corrosion by real-time imaging and electrochemical methods: relationship between local processes and hydrogen evolution, Electrochim. Acta, 2016, 198, 174–184 CrossRef CAS.
  4. W. Yang, Z. J. Zhu, J. J. Wang, Y. C. Wu, T. Zhai and G. L. Song, Slow positron beam study of corrosion behavior of AM60B magnesium alloy in NaCl solution, Corros. Sci., 2016, 106, 271–280 CrossRef CAS.
  5. C. D. Gu, W. Yan, J. L. Zhang and J. P. Tu, Corrosion resistance of AZ31B magnesium alloy with a conversion coating produced from a choline chloride–urea based deep eutectic solvent, Corros. Sci., 2016, 106, 108–116 CrossRef CAS.
  6. H. A. Sorkhabi, D. Seifzadeh and M. G. HosseiniEN, EIS and polarization studies to evaluate the inhibition effect of 3H-phenothiazin-3-one, 7-dimethylamin on mild steel corrosion in 1 M HCl solution, Corros. Sci., 2008, 50, 3363–3370 CrossRef.
  7. H. Ashassi-Sorkhabi and D. Seifzadeh, Analysis of raw and trend removed EN data in time domain to evaluate corrosion inhibition effects of New Fuchsin dye on steel corrosion and comparison of results, J. Appl. Electrochem., 2008, 38, 1545–1552 CrossRef CAS.
  8. K. F. Khaled, Experimental, density function theory calculations and molecular dynamics simulations to investigate the adsorption of some thiourea derivatives on iron surface in nitric acid solutions, Appl. Surf. Sci., 2010, 256, 6753–6763 CrossRef CAS.
  9. C. Li, L. Li and C. Wang, Study of the inhibitive effect of mixed self-assembled monolayers on copper with SECM, Electrochim. Acta, 2014, 115, 531–536 CrossRef CAS.
  10. M. Mobin, M. Parveen and M. Z. A. Rafiquee, Synergistic effect of sodium dodecyl sulfate and cetyltrimethyl ammonium bromide on the corrosion inhibition behavior of l-methionine on mild steel in acidic medium, Arabian J. Chem., 2013, 24, 81–99 Search PubMed.
  11. Y. Ma, N. Liu, Y. S. Wang, J. S. Wang and H. X. Guo, Effect of chromate additive on corrosion resistance of MAO coatings on magnesium alloys, J. Chin. Ceram. Soc., 2011, 39, 1493–1497 CAS.
  12. Y. L. Liu, Z. Y. Yong, S. X. Zhou and L. M. Wu, Molybdate/phosphate composite conversion coating on magnesium alloy surface for corrosion protection, Appl. Surf. Sci., 2008, 255, 1672–1680 CrossRef.
  13. H. Gao, Q. Li, Y. Dai, F. Luo and H. X. Zhang, High efficiency corrosion inhibitor 8-hydroxyquinoline and its synergistic effect with sodium dodecylbenzenesulphonate on AZ91D magnesium alloy, Corros. Sci., 2010, 52, 1603–1609 CrossRef CAS.
  14. D. Seifzadeh, A. Bezaatpour and R. Asadpour Joghani, Corrosion inhibition effect of N,N′-bis(2-pyridylmethylidene)-1,2-diiminoethane on AZ91D magnesium alloy in acidic media, Trans. Nonferrous Met. Soc. China, 2014, 24, 3441–3451 CrossRef CAS.
  15. M. Du, X. S. Yin and H. Gong, Effects of triethanolamine on the morphology and phase of chemically deposited tin sulfide, Mater. Lett., 2015, 152, 40–44 CrossRef CAS.
  16. K. H. Kang and D. K. Lee, Synthesis of magnesium oxysulfate whiskers using triethanolamine as a morphology control agent, J. Ind. Eng. Chem., 2014, 20, 2580–2583 CrossRef CAS.
  17. J. M. Ashurov, A. B. Ibragimov and B. T. Ibragimov, Mixed-ligand complexes of Zn(II), Cd(II) and Cu(II) with triethanolamine and p-nitrobenzoic acid: syntheses and crystal structures, Polyhedron, 2015, 102, 441–446 CrossRef CAS.
  18. M. Zhang and Y. Long, Inhibition of triethanolamine for magnetic refrigeration material LaFe11.0Co0.7Si1.3 in distilled water, J. Funct. Mater., 2010, 41, 132–133 Search PubMed.
  19. Y. R. Ding, X. W. Guo, W. J. Ding and Y. P. Zhu, Effects of organic solution on performance and microstructure of oxide film of magnesium alloy, Surf. Technol., 2005, 34, 14–16 CAS.
  20. D. Daoud, T. Douadi, H. Hamani, S. Chafaa and M. Al-Noaimi, Corrosion inhibition of mild steel by two new S-heterocyclic compounds in 1 M HCl: experimental and computational study, Corros. Sci., 2015, 94, 21–37 CrossRef CAS.
  21. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr, T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian, Inc., Wallingford, CT, 2004.
  22. F. Mansfeld, Tafel slopes and corrosion rates obtained in the pre-Tafel region of polarization curves, Corros. Sci., 2005, 47, 3178–3186 CrossRef CAS.
  23. E. McCafferty, Validation of corrosion rates measured by the Tafel extrapolation method, Corros. Sci., 2005, 47, 3202–3215 CrossRef CAS.
  24. G. L. Song and A. Atrens, Corrosion mechanisms of magnesium alloys, Adv. Eng. Mater., 1999, 1, 11–33 CrossRef CAS.
  25. L. P. Xu, G. N. Yu, E. Zhang, F. Pan and K. Yang, In vivo corrosion behavior of Mg–Mn–Zn alloy for bone implant: application, J. Biomed. Mater. Res., Part A, 2007, 83, 703–711 CrossRef PubMed.
  26. F. Witte, J. Fischer, J. Nellesen, H.-A. Crostack, V. Kaese and A. Pisch, In vitro and in vivo corrosion measurements of magnesium alloys, Biomaterials, 2006, 27, 1013–1018 CrossRef CAS PubMed.
  27. F. Witte, N. Hort, C. Vogt, S. Cohen, K. U. Kainer, R. Willumeit and F. Feyeraben, Degradable biomaterials based on magnesium corrosion, Curr. Opin. Solid State Mater. Sci., 2008, 12, 63–72 CrossRef CAS.
  28. H. Ashassi-Sorkhabi, B. Shaabani and D. Seifzadeh, Corrosion inhibition of mild steel by some schiff base compounds in hydrochloric acid, Appl. Surf. Sci., 2005, 239, 154–164 CrossRef CAS.
  29. Z. C. Wang, F. Jia, L. Yu, Z. B. Qi, Y. Tang and G. L. Song, Direct electroless nickel–boron plating on AZ91D magnesium alloy, Surf. Coat. Technol., 2012, 206, 3676–3685 CrossRef CAS.
  30. X. H. Wang, J. H. Wang and C. W. Fu, Characterization of pitting corrosion of 7A60 aluminum alloy by EN and EIS techniques, Trans. Nonferrous Met. Soc. China, 2014, 24, 3907–3916 CrossRef CAS.
  31. H. Ashassi-Sorkhabi and E. Asghari, Effect of hydrodynamic conditions on the inhibition performance of L-methionine as a “green” inhibitor, Electrochim. Acta, 2008, 54, 162–167 CrossRef CAS.
  32. Z. Zhang, N. C. Tian, L. Z. Zhang and L. Wu, Inhibition of the corrosion of carbon steel in HCl solution by methionine and its derivatives, Corros. Sci., 2015, 98, 438–449 CrossRef CAS.
  33. B. Hirschorn, M. E. Orazem, B. Tribollet, V. Vivier, I. Frateur and M. Musiani, Determination of effective capacitance and film thickness from constant phase-element parameters, Electrochim. Acta, 2010, 55, 6218–6227 CrossRef CAS.
  34. H.-H. Strehblow, B. Titze and B. P. Loechel, The breakdown of passivity of iron and nickel by fluoride, Corros. Sci., 1979, 19, 1047–1057 CrossRef CAS.
  35. X. M. Bi and C. N. Cao, The influence of pH value and Cl concentration on the electrochemical behavior of Fe corrosion process in acid solution, J. Chin. Soc. Corros. Prot., 1983, 3, 200–218 Search PubMed.
  36. E. McCafferty and N. Hackerman, Double layer capacitance of iron and corrosion inhibition with polymethylene diamines, J. Electrochem. Soc., 1972, 119, 146–154 CrossRef CAS.
  37. R. Zhao, Y. M. Tang, J. P. Xiong and Y. Zuo, Inhibition effect of sodium diethyldithiocarbamate and thiourea on AZ91D magnesium alloy in 3.5% NaCl solution, Corros. Sci. Prot. Technol., 2011, 23, 251–255 CAS.
  38. A. Y. Musa, A. A. H. Kadhum, A. B. Mohamad, M. S. Takriff, A. R. Daud and S. K. Kamarudin, Adsorption isotherm mechanism of amino organic compounds as mild steel corrosion inhibitors by electrochemical measurement method, J. Cent. South Univ. Technol., 2010, 17, 34–39 CrossRef CAS.
  39. R. J. Chin and K. Nobe, Electrodissolution kinetics of iron in chloride solutions III. acidic solutions, J. Electrochem. Soc., 1972, 119, 1457–1461 CrossRef CAS.
  40. C. N. Cao, J. Wang and H. C. Lin, Effect of Cl ion on the impedance of passive-film-covered electrodes, J. Chin. Soc. Corros. Prot., 1989, 9, 261–269 Search PubMed.
  41. A. O. Yüce and G. Kardaş, Adsorption and inhibition effect of 2-thiohydantoin on mild steel corrosion in 0.1 M HCl, Corros. Sci., 2012, 58, 86–94 CrossRef.
  42. D. Veys-Renaux, E. Rocca, J. Martin and G. Henrion, Initial stages of AZ91 Mg alloy micro-arc anodizing: Growth mechanisms and effect on the corrosion resistance, Electrochim. Acta, 2014, 124, 36–45 CrossRef CAS.
  43. L. Y. Song and Z. Y. Chen, The role of UV illumination on the NaCl-induced atmospheric corrosion of Q235 carbon steel, Corros. Sci., 2014, 86, 318–325 CrossRef CAS.
  44. M. A. Deyab, Hydroxyethyl cellulose as efficient organic inhibitor of zinc–carbon battery corrosion in ammonium chloride solution: electrochemical and surface morphology studies, J. Power Sources, 2015, 280, 190–194 CrossRef CAS.
  45. M. A. Hegazy and I. Aiad, 1-Dodecyl-4-(((3-morpholinopropyl)imino)methyl)pyridin-1-ium bromide as a novel corrosion inhibitor for carbon steel during phosphoric acid production, J. Ind. Eng. Chem., 2015, 31, 91–99 CrossRef CAS.
  46. M. Outirite, M. Lagrenée, M. Lebrini, M. Traisnel, C. Jama, H. Vezin and F. Bentiss, ac impedance, X-ray photoelectron spectroscopy and density functional theory studies of 3,5-bis(n-pyridyl)-1,2,4-oxadiazoles as efficient corrosion inhibitors for carbon steel surface in hydrochloric acid solution, Electrochim. Acta, 2009, 55, 1670–1681 CrossRef.
  47. A. Y. Musa, A. A. H. Kadhum, A. B. Mohamad, M. S. Takriff, A. R. Daud and S. K. Kamarudin, Adsorption isotherm mechanism of amino organic compounds as mild steel corrosion inhibitors by electrochemical measurement method, J. Cent. South Univ. Technol., 2010, 17, 34–39 CrossRef CAS.
  48. K. F. Khaled, Corrosion control of copper in nitric acid solutions using some amino acids-a combined experimental and theoretical study, Corros. Sci., 2010, 52, 3225–3234 CrossRef CAS.
  49. F. Zhang, Y. Tang, Z. Cao, W. Jing, Z. Wu and Y. Chen, Performance and theoretical study on corrosion inhibition of 2-(4-pyridyl)-benzimidazole for mild steel in hydrochloric acid, Corros. Sci., 2012, 61, 1–9 CrossRef CAS.
  50. K. Zhang, B. Xu, W. Yang, X. Yin, Y. Liu and Y. Chen, Halogen-substituted imidazoline derivatives as corrosion inhibitors for mild steel in hydrochloric acid solution, Corros. Sci., 2015, 90, 284–295 CrossRef CAS.
  51. H. Zhao, X. Zhang, L. Ji, H. Hu and Q. Li, Quantitative structure–activity relationship model for amino acids as corrosion inhibitors based on the support vector machine and molecular design, Corros. Sci., 2014, 83, 261–271 CrossRef CAS.
  52. G. Xia, X. Jiang, L. Zhou, Y. Liao, M. Duan, H. Wang, Q. Pu and J. Zhou, Synergic effect of methyl acrylate and N-cetylpyridinium bromide in N-cetyl-3-(2-methoxycarbonylvinyl)pyridinium bromide molecule for X70 steel protection, Corros. Sci., 2015, 94, 224–236 CrossRef CAS.
  53. D. Yu, Z. D. Chen, F. Wang and S. Z. Li, Study on electronegativity and hardness of the elements by density functional theory, Acta Phys.–Chim. Sin., 2001, 17, 15–22 CAS.
  54. N. Khalil, Quantum chemical approach of corrosion inhibition, Electrochim. Acta, 2003, 48, 2635–2640 CrossRef CAS.
  55. M. Lashkari and M. R. Arshadi, DFT studies of pyridine corrosion inhibitors in electrical double layer: solvent, substrate, and electric field effects, Chem. Phys., 2004, 299, 131–137 CrossRef CAS.
  56. L. Herrag, B. Hammouti and S. Elkadiri, Adsorption properties and inhibition of mild steel corrosion in hydrochloric solution by some newly synthesized diamine derivatives: experimental and theoretical investigations, Corros. Sci., 2010, 52, 3042–3051 CrossRef CAS.
  57. I. B. Obot and N. O. Obi-Egbedi, Adsorption properties and inhibition of mild steel corrosion in sulphuric acid solution by ketoconazole: experimental and theoretical investigation, Corros. Sci., 2010, 52, 198–204 CrossRef CAS.
  58. Z. Zhang, N. Tian and L. Zhang, Inhibition of the corrosion of carbon steel in HCl solution by methionine and its derivatives, Corros. Sci., 2015, 98, 438–449 CrossRef CAS.
  59. R. G. Parr and R. G. Pearson, Absolute Hardness, Companion Parameter to Absolute Electronegativity, J. Am. Chem. Soc., 1983, 105, 7512–7516 CrossRef CAS.
  60. H. Tian, W. Li and K. Cao, Potent inhibition of copper corrosion in neutral chloride media by novel non-toxic thiadiazole derivatives, Corros. Sci., 2013, 73, 281–291 CrossRef CAS.
  61. M. A. Hegazy, A. M. Badawi and S. S. A. E. Rehim, Corrosion inhibition of carbon steel using novel N-(2-(2-mercaptoacetoxy)ethyl)-N,N-dimethyl dodecan-1-aminium bromide during acid pickling, Corros. Sci., 2013, 69, 110–122 CrossRef CAS.
  62. H. Ju, Z. P. Kai and Y. Li, Aminic nitrogen-bearing polydentate Schiff base compounds as corrosion inhibitors for iron in acidic media: a quantum chemical calculation, Corros. Sci., 2008, 50, 865–871 CrossRef CAS.
  63. R. Yıldız, An electrochemical and theoretical evaluation of 4,6-diamino-2-pyrimidinethiol as a corrosion inhibitor for mild steel in HCl solutions, Corros. Sci., 2015, 90, 544–553 CrossRef.
  64. B. Xu, W. N. Gong, K. G. Zhang, W. Z. Yang, Y. Liu, X. S. Yin, H. Shi and Y. Z. Chen, Theoretical prediction and experimental study of 1-butyl-2-(4-methylphenyl)benzimidazole as a novel corrosion inhibitor for mild steel in hydrochloric acid, J. Taiwan Inst. Chem. Eng., 2015, 51, 193–200 CrossRef CAS.
  65. N. O. Eddy, H. Momoh-Yahaya and E. E. Oguzie, Theoretical and experimental studies on the corrosion inhibition potentials of some purines for aluminum in 0.1 M HCl, J. Adv. Res., 2015, 6, 203–217 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.