Z. Zhang*ab,
N. C. Tiana,
X. D. Huanga,
W. Shanga and
L. Wuc
aGuangxi Key Laboratory of Electrochemical and Magneto-chemical Functional Materials, College of Chemistry and Bioengineering, Guilin University of Technology, Guilin 541004, PR China. E-mail: zhangzhe@glut.edu.cn; Tel: +86 15296001526
bDepartment of Chemical Engineering, Mid Sweden University, Sundsvall SE-85170, Sweden
cSchool of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, PR China
First published on 12th February 2016
The synergistic corrosion inhibition effect of indigo carmine and three kinds of cationic organic compounds on 1045 carbon steel (CS) in 0.5 M HCl solution is reported. Electrochemical measurements showed that these three cationic organic compounds combined with indigo carmine reduce the speed of corrosion on 1045 CS and act as effective inhibitors. The combination of indigo carmine with BAB resulted in the best synergistic corrosion inhibition effect (S = 17.14), and the best inhibition efficiency (95.0%). SEM images and XPS data of the corroded steel surfaces suggested that the indigo disulphonate anion and organic cation could be simultaneously adsorbed on the CS surface to inhibit the corrosion of iron. The synergistic inhibition mechanism was investigated by dynamic simulations using quantum chemistry.
As a derivative of the indigo molecule, the indigo carmine molecule is considered as a candidate for corrosion inhibition due to its better solubility. As shown in Table 1, the indigo carmine molecule contains N, O and S atoms and a conjugated aromatic nucleus, which suggest this molecule to be an efficient inhibitor. M. Abdeli et al.11 reported that indigo carmine could inhibit mild steel corrosion in 1 mol L−1 (M) HCl solution. The inhibition efficiency increased with the addition of a small amount of indigo carmine and then decreased when the concentration of indigo carmine was higher than 9.65 × 10−5 M. However, we found that the inhibition effect of indigo carmine is different on mild steel and 1045 carbon steel. The corrosion of 1045 carbon steel could be accelerated when the concentration of indigo carmine exceeded 1 × 10−4 M in our previous research.12 The carbon content and the purity of the iron are important factors for the corrosion-resistance of carbon steel, which is weakened as the carbon content increases, because the existence of carbon elements can enhance the electrochemical corrosion of carbon steel in a non-oxidizing acid medium.13–16
Consequently, the aim of the present work is to further verify the synergistic corrosion inhibition mechanism of indigo carmine and three kinds of cationic organic compounds for CS in 0.5 M HCl solution. Triethanolamine, benzyl trimethyl ammonium bromide and cetyltrimethylammonium bromide were selected as cationic organic compounds to eliminate the influence of hydrophobic properties, and their structural formulas are shown in Table 1. Electrochemical measurements were used to study the inhibition efficiency of each inhibitor. XPS and SEM techniques were used to examine the surface morphology of CS sheets with and without inhibitors in the 0.5 M HCl solution for 48 h. The synergistic corrosion inhibition mechanism and the correlation between molecular structure and inhibition efficiency were investigated using molecular dynamics simulations and quantum chemical calculations.
The adsorption configuration of the indigo carmine molecule and three other kinds of cationic organic molecules on the steel surface were dynamically simulated by using the Discover module of the Materials Studio 6.0 software from Accelrys Inc.22 The details for the setup of this molecular dynamics simulation are listed in Table 2. The details for the simulation system building process have previously been detailed in work by Zhe Zhang.12
Basic setup | Method | Convergence level | Maximum iteration | |
Smart minimizer | Ultra-fine | 20![]() |
||
Minimizer setup | Force field | Non-bond | Summation method | Cutoff distance |
Compass | vdW | Atom based | 9.5 Å | |
Dynamic simulation setup | Ensemble | Thermostat | Simulation temperature | Energy deviation |
NVT | Andersen | 298.0 K | 5000.0 kcal mol−1 | |
Dynamics time | Time step | Frame output | ||
2000.0 ps | 1 fs | Every 1000 steps |
The inhibitors aqueous solution layer and the confined water molecule layer, which were constructed using the amorphous cell module, contained 5H3O+ ions, 5Cl− ions, 500H2O molecules, 1 inhibitor molecule or 2 compounded inhibitor molecules and some corresponding inorganic ions in this simulated system to simulate the 0.5 M HCl solution. Due to the high tendency of TEA molecules be protonated in an acidic solution,17,23–25 TEAH was considered as the cationic organic compound in the dynamic simulation, and the protonation process is shown in Fig. 1. In order to study the interactions between the inhibitors and Fe (1 1 0) surface, 6 layers of iron atoms near the bottom were frozen. Before the molecular dynamics simulation, the constructed box could be optimized such that the total energy of the system was at a local minimum with respect to potential energy. Finally, the dynamic simulated process was carried out until both the temperature and the energy of the whole system reached equilibrium.
Fe → Fe2+ + 2e− | (1) |
2H+ + 2e− → H2↑ | (2) |
And the overall corrosion reaction equation appears as follows:
Fe + 2H+ → Fe2+ + H2↑ | (3) |
The anodic dissolution of iron occurs according to the following steps:30
Fe + Cl− → (FeCl−)ads | (4) |
(FeCl−)ads → (FeCl)ads + e− | (5) |
(FeCl)ads → (FeCl+) + e− | (6) |
(FeCl+) → Fe2+ + Cl− | (7) |
In addition, the hydrogen production occurs via two successive elementary events: to begin with, the initial discharge of hydrogen ions to adsorbed monoatomic hydrogen:31
H+ + e → H˙ | (8) |
Followed by the chemical (eqn (9)) or electrochemical (eqn (10)) recombination of monoatomic hydrogen to molecular hydrogen.
2H˙ → H2↑ | (9) |
H+ + H˙ + e → H2↑ | (10) |
To investigate the inhibition behavior of indigo carmine and other inhibitors (such as various cationic organic inhibitors and various composite inhibitors) of CS in 0.5 M HCl solution at 25 °C, EIS and potentiodynamic polarization measurements were carried out. The Nyquist plots and polarization curves of the WE measured in various test solutions are shown in Fig. 2–4. We could find that all of the Nyquist curves displayed a single capacitive arc, indicating that they each contained one time constant. And these capacitive arcs were the compressed type, which could be attributed to the heterogeneity or roughness of the WE surface.32 Additionally, the shapes of the Nyquist curves for all of the test inhibitors were similar to that of the blank solution. This indicated the corrosion mechanism was almost unchanged by the addition of the inhibitors.33
Thus, the EIS data could be simulated by ZView2 software with a simple equivalent circuit Rs (CPE, Rct) and the fitting model is shown in Fig. 5. The fitting data are shown in Table 3. In this equivalent circuit, Rs represents the solution resistance between the WE and SCE, and Rct represents the charge-transfer resistance. To more accurately fit the EIS of this study, CPE is a constant phase element to replace a double layer capacitance (Cdl), and the admittance and impedance of the CPE can be defined from the following formula:34,35
![]() | (11) |
Cdl = Y01/nRct(1−n)/n | (12) |
![]() | ||
Fig. 5 Equivalent electrical circuit diagram used to model the WE/solution interface in 0.5 M HCl solution. |
Inhibitors | Concentration (mol L−1) | Molar ratio | Rs (Ω) | CPE | Cdl (µF cm−2) | Rct (Ω cm2) | ηR (%) | |
---|---|---|---|---|---|---|---|---|
Y0 (µΩ−1 Sn cm−2) | n | |||||||
Blank sample | 0 | — | 6.98 | 27.77 | 0.852 | 9.47 | 74.0 | — |
Indigo | 0.00001 | — | 6.29 | 71.17 | 0.818 | 24.18 | 111.0 | 33.3 |
0.00005 | — | 6.78 | 77.51 | 0.830 | 26.39 | 68.2 | −8.5 | |
0.0001 | — | 6.83 | 97.66 | 0.831 | 34.35 | 60.5 | −22.4 | |
0.0005 | — | 6.23 | 160.69 | 0.829 | 58.42 | 46.2 | −60.3 | |
0.001 | — | 6.93 | 256.32 | 0.804 | 84.57 | 41.6 | −78.0 | |
0.005 | — | 6.63 | 268.25 | 0.806 | 84.58 | 31.0 | −138.9 | |
0.01 | — | 6.22 | 424.13 | 0.791 | 126.70 | 24.5 | −202.7 | |
TEA | 0.0002 | — | 6.76 | 263.43 | 0.809 | 83.77 | 29.5 | −151.2 |
0.001 | — | 7.05 | 210.00 | 0.808 | 64.00 | 31.6 | −134.4 | |
0.002 | — | 7.22 | 179.32 | 0.834 | 69.21 | 46.0 | −61.0 | |
0.01 | — | 6.65 | 78.43 | 0.835 | 29.81 | 94.2 | 21.4 | |
0.02 | — | 6.66 | 83.16 | 0.837 | 34.10 | 122.0 | 39.3 | |
BAB | 0.0002 | — | 6.63 | 256.32 | 0.804 | 78.70 | 31.0 | −138.9 |
0.001 | — | 6.78 | 199.46 | 0.824 | 72.66 | 44.5 | −66.1 | |
0.002 | — | 6.71 | 144.68 | 0.833 | 55.00 | 55.1 | −34.2 | |
0.01 | — | 6.50 | 44.37 | 0.865 | 20.42 | 159.5 | 53.6 | |
0.02 | — | 6.72 | 47.64 | 0.838 | 19.34 | 195.3 | 62.1 | |
CTAB | 0.00002 | — | 6.78 | 18.08 | 0.753 | 4.92 | 1055.0 | 93.0 |
0.0001 | — | 6.60 | 12.59 | 0.791 | 4.73 | 1938.0 | 96.2 | |
0.0002 | — | 5.68 | 11.59 | 0.875 | 7.01 | 2527.0 | 97.1 | |
0.001 | — | 6.43 | 12.09 | 0.820 | 5.36 | 2046.0 | 96.4 | |
Indigo/TEA | 0.0001 | 1![]() ![]() |
7.58 | 115.49 | 0.826 | 43.92 | 89.1 | 17.0 |
0.0005 | 1![]() ![]() |
6.92 | 44.10 | 0.848 | 18.42 | 172.6 | 57.1 | |
0.001 | 1![]() ![]() |
6.33 | 28.29 | 0.832 | 10.44 | 250.9 | 70.5 | |
0.005 | 1![]() ![]() |
7.00 | 25.87 | 0.854 | 11.37 | 313.8 | 76.4 | |
0.01 | 1![]() ![]() |
6.65 | 44.76 | 0.860 | 20.53 | 189.1 | 60.9 | |
Indigo/BAB | 0.0001 | 1![]() ![]() |
6.61 | 41.79 | 0.817 | 13.96 | 179.3 | 58.7 |
0.0005 | 1![]() ![]() |
6.68 | 29.07 | 0.835 | 13.06 | 594.9 | 87.6 | |
0.001 | 1![]() ![]() |
6.48 | 35.26 | 0.780 | 13.24 | 884.0 | 91.6 | |
0.005 | 1![]() ![]() |
6.19 | 31.55 | 0.823 | 15.91 | 1301.0 | 94.3 | |
0.01 | 1![]() ![]() |
6.50 | 27.98 | 0.825 | 14.37 | 1546.0 | 95.2 | |
Indigo/CTAB | 0.00001 | 1![]() ![]() |
6.51 | 7.88 | 0.810 | 2.73 | 1399.0 | 94.7 |
0.00005 | 1![]() ![]() |
6.74 | 9.48 | 0.779 | 3.85 | 4387.0 | 98.3 | |
0.0001 | 1![]() ![]() |
6.69 | 7.71 | 0.791 | 2.92 | 3290.0 | 97.8 | |
0.0005 | 1![]() ![]() |
6.52 | 8.43 | 0.805 | 3.19 | 2177.0 | 96.6 |
The corrosion inhibition efficiencies (ηR) of impedance have been estimated by Rct according to the following formula:
![]() | (13) |
In addition, the polarization curves for the WE were recorded and the results are shown in Fig. 2–4. The polarization parameters, fitted using the built-in CHI760e software, are listed in Table 4 and include: the corrosion potential (Ecorr), the cathodic and anodic Tafel slopes βc and βa (V dec−1), and the corrosion current density (icorr). These values were obtained via processed polarization curves by Tafel extrapolation. The protection efficiencies (ηi) of corrosion current density were calculated using eqn (14) and the results are listed in Table 4:
![]() | (14) |
Inhibitors | Concentration (mol L−1) | Molar ratio | Ecorr (V vs. SCE) | −βc (V dec−1) | βa (V dec−1) | icorr (µA cm−2) | ηi (%) | S |
---|---|---|---|---|---|---|---|---|
Blank sample | 0 | — | −0.515 | 0.157 | 0.165 | 664.6 | — | — |
Indigo | 0.00001 | — | −0.513 | 0.153 | 0.157 | 373.2 | 43.8 | — |
0.00005 | — | −0.511 | 0.154 | 0.157 | 765.5 | −15.2 | — | |
0.0001 | — | −0.508 | 0.160 | 0.165 | 877.1 | −32.0 | — | |
0.0005 | — | −0.504 | 0.163 | 0.169 | 994.0 | −49.6 | — | |
0.001 | — | −0.503 | 0.162 | 0.164 | 1025.0 | −54.2 | — | |
0.005 | — | −0.501 | 0.170 | 0.163 | 1462.0 | −120.0 | — | |
0.01 | — | −0.502 | 0.183 | 0.192 | 2694.0 | −305.4 | — | |
TEA | 0.0002 | — | −0.506 | 0.172 | 0.175 | 1451.0 | −118.3 | — |
0.001 | — | −0.504 | 0.173 | 0.172 | 1386.0 | −108.5 | — | |
0.002 | — | −0.506 | 0.163 | 0.163 | 1037.0 | −56.0 | — | |
0.01 | — | −0.509 | 0.147 | 0.145 | 472.9 | 28.8 | — | |
0.02 | — | −0.509 | 0.143 | 0.142 | 437.5 | 34.2 | — | |
BAB | 0.0002 | — | −0.501 | 0.170 | 0.164 | 1462.0 | −120.0 | — |
0.001 | — | −0.502 | 0.165 | 0.161 | 1008.0 | −51.7 | — | |
0.002 | — | −0.504 | 0.155 | 0.148 | 798.1 | −20.1 | — | |
0.01 | — | −0.503 | 0.147 | 0.140 | 367.1 | 44.8 | — | |
0.02 | — | −0.505 | 0.145 | 0.132 | 280.2 | 57.8 | — | |
CTAB | 0.00002 | — | −0.525 | 0.158 | 0.150 | 51.7 | 92.2 | — |
0.0001 | — | −0.524 | 0.155 | 0.145 | 24.0 | 96.4 | — | |
0.0002 | — | −0.524 | 0.148 | 0.120 | 16.4 | 97.5 | — | |
0.001 | — | −0.528 | 0.156 | 0.128 | 18.4 | 97.2 | — | |
Indigo/TEA | 0.0001 | 1![]() ![]() |
−0.512 | 0.146 | 0.147 | 478.9 | 27.9 | 4.00 |
0.0005 | 1![]() ![]() |
−0.513 | 0.143 | 0.146 | 349.8 | 47.4 | 5.93 | |
0.001 | 1![]() ![]() |
−0.524 | 0.188 | 0.160 | 185.8 | 72.0 | 8.61 | |
0.005 | 1![]() ![]() |
−0.522 | 0.190 | 0.161 | 176.2 | 73.5 | 5.90 | |
0.01 | 1![]() ![]() |
−0.523 | 0.182 | 0.164 | 240.4 | 63.8 | 7.38 | |
Indigo/BAB | 0.0001 | 1![]() ![]() |
−0.511 | 0.141 | 0.130 | 286.9 | 56.8 | 15.79 |
0.0005 | 1![]() ![]() |
−0.512 | 0.133 | 0.114 | 80.4 | 87.9 | 17.14 | |
0.001 | 1![]() ![]() |
−0.523 | 0.125 | 0.098 | 56.7 | 91.5 | 16.31 | |
0.005 | 1![]() ![]() |
−0.522 | 0.117 | 0.090 | 42.5 | 93.6 | 12.03 | |
0.01 | 1![]() ![]() |
−0.523 | 0.118 | 0.091 | 33.2 | 95.0 | 12.44 | |
Indigo/CTAB | 0.00001 | 1![]() ![]() |
−0.521 | 0.157 | 0.134 | 30.9 | 95.3 | 0.94 |
0.00005 | 1![]() ![]() |
−0.523 | 0.142 | 0.109 | 9.9 | 98.5 | 3.63 | |
0.0001 | 1![]() ![]() |
−0.524 | 0.139 | 0.112 | 13.2 | 98.0 | 1.92 | |
0.0005 | 1![]() ![]() |
−0.520 | 0.150 | 0.109 | 21.8 | 96.7 | 1.86 |
As shown in Table 3, the Cdl value increased from 9.49 to 126.70 µF cm−2 with an increase in the concentration of indigo carmine. According to the Helmholtz model, this result is reasonable, as Cdl is inversely proportional to the surface charge and directly proportional to the local dielectric constant of the WE surface:37,41,42
![]() | (15) |
From Fig. 2b, we can find that the anodic and cathodic current densities all changed with an increase in concentration. With the exception of the corrosion current density in the presence of the 1 × 10−5 M indigo carmine sample, which was lower than that of the blank sample, the icorr values increased with an increase in the concentration of the indigo carmine samples, and were higher than the value for the blank sample (Table 4). This means that indigo carmine could inhibit the corrosion of CS in 0.5 M HCl solution at a concentration of about 1 × 10−5 M. When the concentration of indigo carmine was above 5 × 10−5 M, however, it could accelerate the corrosion of CS in 0.5 M HCl solution. Comparing the cathodic and anodic Tafel slopes (βc and βa, respectively) of indigo carmine (Table 4), we found that the increase in the value of βc was bigger than the increase in the value of βa, which means that hydrogen reduction in the cathodic region was further accelerated. This indicated that the adsorbed indigo disulphonate ion inducing the discharge of H+ on the CS surface is the major factor in indigo carmine’s acceleration of CS corrosion in 0.5 M HCl solution.
Based on the Cdl values of these three kinds of cationic organic inhibitors, they could be placed in the order TEA > BAB ≫ CTAB. There was little difference in the Cdl values for TEA and BAB, but both were higher than the Cdl value for CTAB. It is worth noting that the Cdl values for CTAB were all lower than that of the blank sample. The local dielectric constant (ε) of the CS surface with adsorbed TEAH or benzyl trimethyl ammonium ions was much higher than the ε of the CS surface with adsorbed cetyltrimethylammonium ions, because the hydrophobicity of the cetyltrimethylammonium ion is better than that of TEAH and the benzyl trimethyl ammonium ion. From the values of n in Table 3, we could find that the n of each cationic organic inhibitor increased in the range of 0.8 to 0.9 with an increase in concentration. The reason may be that the surface of the working electrodes is non-homogeneous, and the self-assembled films improve the heterogeneity. In summary, these three kinds of cationic organic inhibitors can all become adsorbed on the CS surface and form films with different properties. Their corrosion inhibition effects follow the order of CTAB ≫ BAB > TEA.
As shown in Table 4, the icorr values for TEA and BAB were higher than that of the blank sample when their concentrations were under 0.01 M. When the concentrations of TEA and BAB were above 0.01 M, their icorr values were lower than that of the blank sample. Furthermore, their icorr values decreased as their concentrations increased. This means that TEA and BAB could inhibit the corrosion of CS in 0.5 M HCl solution when their concentrations are high enough. The icorr values for CTAB were all lower than that of the blank sample, which indicated that CTAB could effectively inhibit the corrosion of CS. These results are consistent with the result from the EIS measurements. From the polarization curves in Fig. 3, we could find that the anodic and cathodic current densities obviously decreased when these three kinds of cationic organic molecules played the role of inhibitor. The oxidative dissolution of the working electrodes in the anodic region and hydrogen reduction in the cathodic region were weaker in the Tafel plots, which indicated that the anodic dissolution of CS and hydrogen evolution were both suppressed. In general, the inhibitors can be classified as anodic or cathodic types when the change in the Ecorr value is greater than 85 mV.44–47 Obviously, TEA and BAB could be classified as mixed type inhibitors due to the slight shifts in the Ecorr values. Although the Ecorr values with CTAB were all shifted in the negative direction, the maximal displacement was also less than 85 mV. So, CTAB can also be classified as a mixed type inhibitor.
As shown in Table 3, the Cdl values for these three kinds of compound inhibitors were much smaller than those of the single indigo carmine or corresponding single cationic organic inhibitors. In particular, the Cdl values for indigo carmine/CTAB were smaller than that of the blank sample. This result indicated that the local dielectric constant of the CS surface modified with various compound inhibitors was decreased due to the simultaneous adsorption of the indigo disulphonate anion and other organic cation on the CS surface, equilibrating the charge distribution of the CS. The n of each compound inhibitor changed in the range of 0.78 to 0.86 with no obvious regularity, as shown in Table 3. This may be because the surface heterogeneity of CS was changed with no obvious regularity due to the simultaneous adsorption of two oppositely charged molecules on the CS surface. But this change of n was acceptable. Thus, we inferred that the indigo carmine and the cationic organic inhibitors could adsorb simultaneously on the CS surface, and formed a more stable and excellent film to protect CS from corrosion in 0.5 M HCl solution.
From the polarization curves shown in Fig. 4, we could find that the anodic and cathodic current densities obviously decreased. And the icorr values for these three kinds of compound inhibitors decreased with an increase in their concentrations, as shown in Table 4. The oxidative dissolution of the working electrodes in the anodic region and the hydrogen reduction in the cathodic region were both weaker in the Tafel plots, which indicated that the anodic dissolution of CS and hydrogen evolution were both suppressed. Additionally, the Tafel slopes βc and βa for indigo carmine/TEA increased with an increase in their concentrations, and the Tafel slopes βc and βa for indigo carmine/BAB and indigo carmine/CTAB both decreased with increasing concentration, as shown in Table 4. The reason may be that the adsorption capability of indigo carmine is stronger than that of TEA, thus, more and more TEA molecules could be replaced by indigo carmine molecules as the concentration increased. So, the Tafel slopes change for indigo carmine/TEA was consistent with the Tafel slopes change for indigo carmine. Additionally, the change in βa was more than for βc for both indigo carmine/BAB and indigo carmine/CTAB compound inhibitors, which indicated that the anodic dissolution inhibition was stronger than the hydrogen evolution inhibition. However, in these experiments, the largest displacement of Ecorr (ΔEcorr) was less than 20 mV (vs. SCE, Table 4). This displacement indicates that these three kinds of compound inhibitors all act as mixed-type inhibitors by inhibiting both hydrogen evolution and the CS dissolution reaction. In summary, these three kinds of compound inhibitors could effectively inhibit the corrosion of CS in 0.5 M HCl solution, and their protection efficiencies followed the order of indigo carmine/CTAB > indigo carmine/BAB > indigo carmine/TEA. The maximum protection efficiency was 98.5% for the indigo carmine/CTAB compound inhibitor.
![]() | (16) |
In summary, the synergistic inhibition mechanism of the present work can be described as follows:
Firstly, the single indigo carmine could strongly adsorb on the CS surface through nitrogen atoms, sulfur atoms, oxygen atoms and aromatic ring groups. However, it was unable to protect the iron from corrosion in 0.5 M HCl solution, because the adsorption of indigo disulphonate ions on the CS surface could change the charge distribution of CS, resulting in its surface becoming negatively charged. This negatively charged surface would induce the discharge of H+ and accelerate corrosion (as shown in Fig. 6a). Secondly, the single cationic organics could inhibit the corrosion of CS in 0.5 M HCl by forming a complete and dense adsorption film when their concentrations were high enough (for instance, CTAB could adsorb on the steel surface and form a dense hydrophobic film). If the cationic organics adsorption film was broken, Cl− ions would induce pitting corrosion of the steel surface, as Cl− ions were gathered around the positively charged steel surface (as shown in Fig. 6b). As such, the single cationic organics could be described as “dangerous inhibitors”. Finally, the indigo disulphonate ions and organic cationic ions could be adsorbed evenly on the CS surface and, gradually, a dense film formed when the indigo carmine combined with the cationic organics. This co-adsorption would not break the charge distribution of the CS (as shown in Fig. 6c). Thus, the inhibition effect of the compound inhibitors is better than that of the single indigo carmine or other cationic organics inhibitors.
The morphologies of the CS surfaces immersed in the corrosion solutions in the absence and presence of various inhibitors are displayed in SEM micrographs (on Fig. 7). It can be observed that the CS surface was corroded badly in the 0.5 M HCl solution (Fig. 7b). Porous structures could be found on the corroded surface, and the multiple holes may have been caused by the Cl− ions erosion process.28,49 In contrast, a flatter and cracked structure was found on the CS surface that was corroded in 0.5 M HCl solution containing indigo carmine (Fig. 7c). This cracked structure indicated that the CS surface may have been corroded by the hydrogen evolution reaction.50 Additionally, the corroded surface showed a network-like pattern with a non-uniform distribution and irregular micro-cracks, which means that the indigo disulphonate anions were adsorbed on the CS surface and partly inhibited the corrosion. However, the negatively charged surface induced the discharge of H+ and caused the hydrogen evolution reaction on uncovered areas. It can be observed that the CS surface experienced pitting corrosion to different extents in the 0.5 M HCl solutions containing TEA, BAB and CTAB, and the corrosion process induced a porous structure in all cases (Fig. 7d, f and h). The damage to the CS followed the order TEA > BAB > CTAB. However, the damage to the CS sheet immersed in corrosion solutions containing BAB or CTAB was less than that to the CS surface immersed in the 0.5 M HCl solution. This means that the single cationic organic inhibitors (TEA, BAB and CTAB) could adsorb on the CS surface and protect it from corrosion. However, if the adsorption films of cationic organic inhibitors were not complete enough or broken, the CS could experience pitting corrosion in the corrosion solution containing cationic organic inhibitors. In contrast, the damage to CS sheets immersed in corrosion solutions containing compound inhibitors was always less than that to CS sheets immersed in solutions containing single cationic organic inhibitors (Fig. 7e, g and i). Comparing Fig. 7f and g, we could find that the pitting corrosion on the CS surface was obviously inhibited by the indigo carmine/BAB compound inhibitor. The inhibition effect of these three kinds of compound inhibitors followed the order: indigo carmine/CTAB > indigo carmine/BAB > indigo carmine/TEA, based on the surface morphology. This result is consistent with the electrochemical measurements.
![]() | ||
Fig. 9 The frontier molecular orbital (EHOMO, ELUMO) surfaces for the inhibitor molecules indigo carmine (a), CTAB (b), TEA (c), TEAH (d) and BAB (e). |
Assembly molecules | EHOMO (eV) | ELUMO (eV) | ΔE (eV) | µ (Debye) | χ = (I + A)/2 | γ = (I − A)/2 | ΔN |
---|---|---|---|---|---|---|---|
Indigo carmine | −0.719 | 1.493 | 2.212 | 0.350 | −0.387 | 1.106 | 3.340 |
TEA | −5.765 | 1.844 | 7.609 | 3.222 | 1.960 | 3.804 | 0.662 |
TEAH | −11.349 | −3.468 | 7.881 | 1.1766 | 7.409 | 3.941 | −0.052 |
BAB | −10.364 | −4.215 | 6.148 | 5.957 | 7.290 | 3.074 | −0.047 |
CTAB | −8.956 | −3.082 | 5.874 | 38.397 | 6.019 | 2.937 | 0.167 |
EHOMO is usually associated with the ability of a molecule to donate electrons, and a higher EHOMO molecule is more likely to donate electrons to a suitable acceptor with low energy. However, the ELUMO is associated with the ability of a molecule to accept electrons, and a lower EHOMO molecule is more likely to accept electrons. The energy gap (ΔE = ELUMO − EHOMO), is a significant quantum chemical parameter that represents the adsorption reactivity tendency between an organic molecule and metal surface. Additionally, the adsorption energy will increase as the energy gap (ΔE = ELUMO − EHOMO) decreases.56,57 Fig. 9a and c show the surfaces of the highest occupied molecular orbital covering the sulfonic groups of indigo carmine and the nitrogen atom of TEA, because the sulfonic groups and nitrogen atom are good electron-donating groups. This indicates that the indigo carmine and TEA molecules could adsorb on the iron surface through coordination. Due to the ionization of BAB and CTAB molecules in acid solution, the highest occupied molecular orbital surfaces cover the long carbon chain of CTAB and the benzyl group of BAB, not the ammonium group (Fig. 9b and e). This means that the capability of CTAB and BAB to donate electrons was much poorer. Comparing TEA with TEAH, we could find that the coverage area of the highest occupied molecular orbital was changed from the nitrogen atom to oxygen atoms (Fig. 9c and d). It therefore appears that protonation could change the molecular hybridization orbital. Table 5 shows that the EHOMO of these inhibitor molecules followed in the order of: indigo carmine ≫ TEA > CTAB > BAB > TEAH. This confirms that the capability of these inhibitor molecules to donate electrons followed the order of indigo carmine ≫ TEA > CTAB > BAB > TEAH. Fig. 9a shows the surfaces of the lowest unoccupied molecular orbital covered the whole indigo carmine molecule, which means that indigo carmine’s ability to accept electrons is much better. However, the EHOMO in Table 5 followed the order of BAB > TEAH > CTAB ≫ indigo carmine > TEA, due to the ionization and the protonation of TEA, BAB and CTAB. Ordering by the value of ΔE, as shown in Table 5, results in indigo carmine < CTAB < BAB < TEA < TEAH, which means that the adsorption capability of these molecules follows the order indigo carmine > CTAB > BAB > TEA > TEAH.
According to Koopmans’s theorem, EHOMO and ELUMO are related to the ionization potential (I) and the electron affinity (A) of the inhibitor molecules and the metal atoms,58,59 which are estimated by: I = −EHOMO and A = −ELUMO, respectively.60 Then, the values of χ and γ for the inhibitor molecules were calculated from eqn (17) and (18):61
![]() | (17) |
![]() | (18) |
Thus, the change in the number of electrons transferred (ΔN) is calculated by eqn (19):23,62
![]() | (19) |
Eadsorption = Etotal − (Esurface+solution + Einhibitor+solution) + Esolution | (20) |
Ebinding = −Eadsorption | (21) |
Inhibitors | Eadsorption (eV) | Ebinding (eV) | Diffusion coefficient/(10−9 m2 s−1) | ηmax (%) | |
---|---|---|---|---|---|
H3O+ | Cl− | ||||
Indigo carmine | −7.481 | 7.481 | 0.13737 | 0.082842 | −202.7 |
TEA | −3.439 | 3.439 | 0.098373 | 0.123883 | 39.3 |
Indigo carmine/TEA | −11.333 | 11.333 | 0.081077 | 0.062967 | 76.4 |
BAB | −3.220 | 3.220 | 0.064103 | 0.134607 | 62.1 |
Indigo carmine/BAB | −10.660 | 10.660 | 0.06294 | 0.08728 | 95.2 |
CTAB | −1.605 | 1.605 | 0.083885 | 0.090855 | 97.1 |
Indigo carmine/CTAB | −9.256 | 9.256 | 0.04267 | 0.045315 | 98.3 |
TEAH | −3.181 | 3.181 | 0.059927 | 0.134815 | — |
Indigo carmine/TEAH | −10.791 | 10.791 | 0.100543 | 0.061465 | — |
![]() | ||
Fig. 10 The equilibrium configuration of indigo carmine adsorbed on an Fe (110) surface, and the radial distribution function between various main elements of indigo carmine molecule and Fe surface. |
It can be observed from Fig. 10 that the indigo carmine molecules adsorbed on the Fe (1 1 0) surface have a planar structure. And the Ebinding value for the indigo carmine molecule is 7.481 eV (Table 6), which is much bigger than the Ebinding value for the three other kinds of cationic organic molecules. This means that the adsorption capability of the indigo carmine molecule is much stronger than that of the three other kinds of cationic organic molecules. The basic molecule–molecule interaction types can be discerned by the typical bond lengths. For instance, the typical bond lengths for a van der Waals interaction are 5–10 Å, metal complexation bond lengths are 2–3 Å and H bond lengths are 2–3.5 Å, approximately.65,66 In general, the radial distribution function is an effective method for estimating the bond length. The x-axis value for the first peak always gives the bond length, and the peak area is identified as the coordination number. The chemisorption bond length is within the limits of 1–3.5 Å, while physisorption bond lengths are longer than 3.5 Å. In the equilibrium configuration of indigo carmine, the radial distribution function of the C, N, O and S atoms shows that the bond lengths of C–Fe, N–Fe, O–Fe and S–Fe are all less than 3.5 Å, notably, the O–Fe bond length is less than 3.0 Å (Fig. 10). This indicates that all of the atoms mentioned above could be coordinately adsorbed on the iron surface by donating π-electrons to the unoccupied d-orbital of iron to form coordinate bonds, as well as by accepting electrons from the d-orbital of iron to form anti-bonding orbital coordinate bonds.17,23,67 In consideration of the frontier molecular orbital of indigo carmine in Fig. 9a, we could infer that the sulfonic acid groups are the main active sites for adsorption between the indigo carmine molecule and the iron surface, and the existence of indoxyl groups could enhance this adsorption. Thus, we could conclude that the adsorption type of indigo carmine was chemisorption, and the adsorption intensity is strong.
From Fig. 11, we can find that most of the cationic organic molecules and their compound inhibitor molecules adsorbed on the Fe (1 1 0) surface have a planar structure (Fig. 11a, c and g). The CTAB molecules adsorbed on the Fe (1 1 0) surface have a vertical structure, because the long-chain part of CTAB’s structure is a hydrophobic group. In addition, the Ebinding for the single inhibitor molecules followed the order of indigo carmine ≫ TEA > BAB > TEAH > CTAB (Table 6). With the exception of CTAB’s position, this order of Ebinding is consistent with the order of the ΔN for indigo carmine and the three other kinds of cationic organic inhibitor molecules (indigo carmine ≫ TEA > CTAB > BAB > TEAH). This may be because the CTAB molecule can be adsorbed on the iron surface by its amino group and not the long carbon chain, but the surfaces of the highest occupied molecular orbital covered the long carbon group. Comparing the adsorption equilibrium configurations of TEA and TEAH molecules, the protonated amino group of TEAH molecule (–[NH2–R′]+) was far away from the iron surface (Fig. 11a and g), which possibly because of the electron-deficient nature of the protonated N atom. Thus the Ebinding for the TEAH molecule is smaller than that for the TEA molecule. Additionally, part of the TEA molecule could combine with indigo carmine due to the electrostatic attraction between the sulfonate group and H atom present in hydroxyl group of the TEA molecule. Thus, some of the TEAH molecules in the TEAH/indigo carmine compound inhibitor mixture will be replaced by TEA molecules. This means that the effect of TEAH molecule adsorption in balancing the charge on the CS surface will not be effective, due to the competitive adsorption of TEA molecules. That is why the corrosion inhibition effect of the indigo carmine/TEA compound inhibitor is not good. In order to analyze the synergy between the indigo carmine molecule and cationic organic molecules, we have compared the Ebinding values of the compound inhibitors and single inhibitors in Table 6. We can find that the Ebinding values of the compound inhibitors are bigger than the sums of the Ebinding values of the individual inhibitors (for example, the Ebinding for indigo carmine/TEA is bigger than sum of the Ebinding values for the single indigo carmine molecule and single TEA). This means that the adsorption capability of the compound inhibitors can be enhanced by a synergistic effect.
In order to investigate the mechanism behind the synergy of indigo carmine and the cationic organic compound inhibitors with molecular dynamics simulations, the diffusion coefficient (D) can be used to describe the migration rate of a corrosive species (H3O+ ions and Cl− ions) in the simulation system. A small diffusion coefficient indicates greater hindrance of the simulation system and the weak reactivity of H3O+ ions and Cl− ions, which reflects a high corrosion inhibition efficiency. The diffusion coefficient is defined as follows:66,68,69
![]() | (22) |
![]() | (23) |
From Fig. 12 and Table 6, we can find that the D values for the H3O+ ions and Cl− ions in the compound inhibitors simulation system were smaller than the D values for the H3O+ ions in the single indigo carmine inhibitor simulation system and the Cl− ions in the single cationic organic inhibitors system. This means that the migration rates of the corrosive species (H3O+ ions and Cl− ions) in the simulation system are both inhibited by the indigo disulphonate anion and other organic cation being simultaneously adsorbed on the iron surface and balancing the charge of the iron surface.
Also, we have investigated the surface concentration profiles of the corrosive species (H3O+ ions and Cl− ions) in the dynamic simulation systems along the normal direction (0 0 1), and the results are shown in Fig. 13. It can be observed from Fig. 13 that the distribution of H3O+ ions and Cl− ions in the compound inhibitors simulation system is farther from the iron surface than the distribution of H3O+ ions and Cl− ions in the single indigo carmine simulation system and single cationic organic inhibitors simulation system, respectively. This means that the compound inhibitors consisting of indigo carmine and cationic organics could balance the charge of the iron surface, and keep the corrosive species (H3O+ ions and Cl− ions) away from the iron surface. Thus, we can conclude that indigo carmine could cooperate with the cationic organic compounds to synergistically inhibit the corrosion of iron in 0.5 M HCl solution.
XPS spectra and SEM images revealing the morphologies of the CS samples corroborated the electrochemical measurements and showed evidence of the chemisorption of indigo carmine and the three kinds of cationic organic compounds on the CS surface. The micro-cracks and porosity featured in the micrographs may be used to discern hydrogen evolution corrosion and pitting corrosion caused by Cl− ions, respectively.
Additionally, quantum chemical calculations and molecular dynamics simulations adequately confirmed that indigo carmine could cooperate with the cationic organic compounds to synergistically inhibit the corrosion of iron in 0.5 M HCl solution. Using the diffusion coefficient and the surface concentration profile are useful methods of investigating the synergistic corrosion inhibition effect in acid solution.
This journal is © The Royal Society of Chemistry 2016 |