Hany M. Abd El-Lateef*a and
Ahmed H. Tantawyb
aChemistry Department, Faculty of Science, Sohag University, 82524 Sohag, Egypt. E-mail: Hany_shubra@yahoo.co.uk; Fax: +20-93-4601159; Tel: +20-12-28-137-103
bChemistry Department, Faculty of Science, Benha University, 13518 Benha, Egypt. E-mail: daht1982@yahoo.com
First published on 13th January 2016
A new family of Schiff base cationic surfactants (CSSB) having various alkyl chain lengths were prepared and their chemical structure was elucidated by using different spectroscopic techniques (FTIR, 13C-NMR and 1H-NMR). The surface activity parameters of the prepared surfactants were measured to set the adsorption and micellization characteristics at the water/air interface. The corrosion inhibition capability of these surfactants was investigated on a carbon steel surface in 3.5% NaCl+ 0.5 M HCl solutions at different temperatures (30–60 °C) by potentiodynamic polarization and electrochemical impedance spectroscopy measurements. The results revealed that CSSB compounds inhibited corrosion of carbon steel in the investigated acidic chloride containing environment. It was found that the inhibition efficiency increases with an increase in inhibitor concentrations and decreases with increasing temperature. Polarization data indicated that the investigated compounds act as mixed-type inhibitors, and the adsorption isotherm basically obeys the Langmuir isotherm. The corrosion inhibition mechanism was discussed based on the potential of zero charge value. An SEM/EDX studies confirmed that CSSB inhibitors could form films by adsorption on the carbon steel surface. The theoretical predictions exhibit good agreement with empirical results.
Acid inhibitors have many important roles in the industrial processes as components in cleaning solutions and in pre-treatment composition.8 Organic compounds that contain nitrogen, oxygen or sulphur atoms, which can donate lone pairs of electrons are found to be particularly useful as inhibitors of steel corrosion via adsorption of the molecules on the metal surface, creating a barrier against corroding attack.9,10 Schiff bases compounds are one of the most widely used families of organic corrosion inhibitors.11 In general, they are prepared by the condensation reaction of carbonyl compounds with primary amines.12,13 Several Schiff bases have recently been investigated as corrosion inhibitors for various metals and alloys in aqueous media.14–17 Some research works revealed that the inhibition efficiency of Schiff bases is much greater than that of corresponding aldehydes and amines due to the presence of the azomethine group (–C
N–) in the molecule.18 Surfactants represent an important class of organic compounds, which are used widely in industry especially as corrosion inhibitors for metals in different acidic solutions.2,3,19–21 The surfactant inhibitors have many advantages such as low price, low toxicity, easy production and high inhibition efficiency.22,23 The most important action of inhibition is the adsorption of the surfactant functional group onto the metal surface. In acidic chloride solutions, the presence of Cl− ion either as a counter ion or in solution, it helps to increase the extent of adsorption due to the well-known synergistic effects.24
Recently, the effectiveness of an inhibitor molecule has been related to its spatial as well as electronic structure.25,26 Quantum chemical calculations have become an effective tool for investigating these parameters and are able to provide insight into the inhibitor–surface interaction.
The search for new, more efficient, more specific eco-surfactants compounds represents a major challenge and is of great interest in the area of environmental treatment. In the present study, some novel cationic surfactants contain Schiff base group was synthesized. The surface activity of the prepared surfactants was measured using surface tension. The inhibition performance of the investigated surfactants towards carbon steel in 3.5% NaCl+ 0.5 M HCl solution at different temperatures (30–60 °C) was evaluated by potentiodynamic polarization and electrochemical impedance spectroscopy (EIS) methods. SEM/EDX observations of the metal surface were performed in the absence and presence of the synthesized surfactants. Also, the correlations between the inhibition efficiency of the prepared inhibitors for the carbon steel corrosion and some quantum parameters have been discussed.
The corrosive medium 3.5% NaCl+ 0.5 M HCl (pH = 1.21), 3.5% NaCl (E. Merck), was prepared by dissolving of analytical grade NaCl in deionized water. Appropriate volume (0.5 M) of the acid was added to NaCl solution. The CSSB inhibitors are added to the 3.5% NaCl+ 0.5 M HCl at concentrations of 25, 50, 75, 100, 150 ppm by weight. All the CSSB inhibitors are soluble in bidistilled water.
Decyl (98%), tetradecyl (98%), and hexadecyl (97%), alcohols were obtained from M/s S.D. Fine chemicals Pvt. Ltd, India. Salicylaldehyde (97%), and tetrahydrofuran (99%), were purchased from AL-Nasr Chemical Company. 3-(N,N-Dimethylamino)-1-propylamine (99%), was obtained from Acros organics Company (Belgium) and used without further purification. Solvents (ethyl alcohol absolute (99%), and diethyl ether (99%)) are high grade and purchased from Algomhoria Chemical Co., Cairo, Egypt. All the solvent and reagents were used as received without further purification.
• Melting points of obtained compounds were determined by Gallenkamp.
• The FT-IR spectrum was recorded in KBr on a thermo nicolet iS10 FTIR spectrophotometer.
• The 1H-NMR (400 MHz) and 13C-NMR (75 MHz) spectra were measured in DMSO-d6 using FT-IR-ALPHA-BRUKER-Platinum-ATR.
• Tensiometer-K6 processor (krüss company, Germany) using the ring method.
![]() | ||
| Fig. 1 Synthetic procedure of Schiff base cationic surfactant compounds (CSSB-10, CSSB-14 and CSSB-16). | ||
–CH2–), 1.28 (s, 26H,
–), 1.68 (p, 2H,
CH2O), 4.06 (s, 2H,
–Cl), 4.2 (t, 2H,
–O).
N), also showed the disappearance of carbonyl band and –NH2 band.The polarization and impedance parameters such as corrosion current (Jcorr), corrosion potential (Ecorr), anodic Tafel slope (βa), cathodic Tafel slope (βc), double layer capacitance (Cdl) and polarization resistance (Rp), were computed from the polarization curves and Nyquist plots.27
The inhibition efficiency (PPDP) and surface coverage (θ) values were calculated from potentiodynamic polarization curves using the eqn (1),17
![]() | (1) |
The corrosion rate (CR) is calculated using the following equation:28
![]() | (2) |
From EIS measurements the double layer capacitance (Cdl) and the inhibition efficiency (PEIS) values were calculated from the following relations,18
| Cdl = Y0(ωm)n−1 | (3) |
![]() | (4) |
The electrical conductivity (K) of various solutions concentrations of synthesized surfactants were determined by an electrical conductivity meter (Type AD3000; EC/TDS and Temperature meter) at 30 °C.
![]() | (5) |
O of ester), and 1632 (imine group, –CH
N) (ESI, Fig. S1a†).1H-NMR-400 MHz (DMSO-d6) d ppm: 0.95 (t, 3H,
–CH2–), 1.42 (s, 14H,
–), 1.64 (m, 2H,
–CH2–O), 2.21 (t, 2H,
CH2–N), 3.32 (t, 2H,
–N+), 3.31 (s, 6H, 2
–N+), 3.73 (t, 2H,
–N
CH), 4.34 (t, 2H,
–O), 4.55 (t, 2H,
–CO), 6.95–7.68 (m, 4H, aromatic CH), 8.71 (s, 1H, ![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N), 13.13 (s, 1H, ![[O with combining low line]](https://www.rsc.org/images/entities/b_char_004f_0332.gif)
) (ESI, Fig. S1b†).
13C-NMR-400 MHz, δC (ppm) (DMSO): 14.12 [
–CH2], 22.25 [–
–CH2–N], 23.22 [
–CH3], 26.58 [–
–CH2CH2O], 28.31 [–
–CH2O], 30.26 [4(–
–) alkyl chain], 31.21 [
–CH2CH3], 51.31 [2
–N+], 53.62 [
–N], 61.27 [
–C
O], 63.45 [
–N+], 67.29 [
–O], 116–139 [![[5 with combining low line]](https://www.rsc.org/images/entities/b_char_0035_0332.gif)
, aromatic carbons], 160 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N–], 166.33 [CH![[double bond, length as m-dash]](https://www.rsc.org/images/entities/char_e001.gif)
–OH, aromatic carbon], 169.32 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
O, carbonyl carbon] (ESI, Fig. S1c†).
O of ester), and 1633 (imine group, CH
N) (ESI, Fig. S2a†).1H-NMR-400 MHz (DMSO-d6) d ppm: 0.91 (t, 3H,
–CH2–), 1.3 (s, 22H,
–), 1.6 (m, 2H,
–CH2–O), 2.1 (t, 2H,
CH2–N), 3.2 (t, 2H,
–N+), 3.3 (s, 6H, 2
–N+), 3.7 (t, 2H,
–N
CH), 4.2 (t, 2H,
–O), 4.5 (t, 2H,
–CO), 6.93–7.5 (m, 4H, aromatic CH), 8.63 (s, 1H, ![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N), 13.01 (s, 1H, ![[O with combining low line]](https://www.rsc.org/images/entities/b_char_004f_0332.gif)
) (ESI, Fig. S2b†).
13C-NMR-400 MHz, δC (ppm) (DMSO): 14.11 [
–CH2], 22.23 [–
–CH2–N], 23.12 [
–CH3], 26.48 [–
–CH2CH2O], 28.11 [–
–CH2O], 30.20 [8(–
–) alkyl chain], 31.21 [
–CH2CH3], 51.11 [2
–N+], 53.21 [
–N], 61.14 [
–C
O], 63.22 [
–N+], 67.16 [
–O], 116–134 [![[5 with combining low line]](https://www.rsc.org/images/entities/b_char_0035_0332.gif)
, aromatic carbons], 160 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N–], 165.23 [CH![[double bond, length as m-dash]](https://www.rsc.org/images/entities/char_e001.gif)
–OH, aromatic carbon], 168.22 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
O, carbonyl carbon] (ESI, Fig. S2c†).
O of ester), and 1633 (imine group, CH
N) (cf. Fig. 2a).
![]() | ||
| Fig. 2 FT-IR (a) 1H-NMR (b) and 13C-NMR (c) spectra of ((2-hydroxybenzylidene)amino)-N,N-dimethyl-N-(2-oxo-2-(hexadecyloxy)ethyl)propan-1-ammonium chloride [CSSB-16]. | ||
1H-NMR-400 MHz (DMSO-d6) d ppm: 0.92 (t, 3H,
–CH2–), 1.4 (s, 26H,
–), 1.61 (m, 2H,
–CH2–O), 2.11 (t, 2H,
CH2–N), 3.12 (t, 2H,
–N+), 3.33 (s, 6H, 2
–N+), 3.72 (t, 2H,
–N
CH), 4.32 (t, 2H,
–O), 4.45 (t, 2H,
–CO), 6.90–7.65 (m, 4H, aromatic CH), 8.73 (s, 1H, ![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N), 13.03 (s, 1H, ![[O with combining low line]](https://www.rsc.org/images/entities/b_char_004f_0332.gif)
) (cf. Fig. 2b).
13C-NMR-400 MHz, δC (ppm) (DMSO): 14.13 [
–CH2], 22.25 [–
–CH2–N], 23.22 [
–CH3], 26.58 [–
–CH2CH2O], 28.31 [–
–CH2O], 30.26 [11(–
–) alkyl chain], 31.23 [
–CH2CH3], 51.41 [2
–N+], 53.61 [
–N], 61.24 [
–C
O], 63.42 [
–N+], 67.26 [
–O], 117–137 [![[5 with combining low line]](https://www.rsc.org/images/entities/b_char_0035_0332.gif)
, aromatic carbons], 160 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
![[H with combining low line]](https://www.rsc.org/images/entities/b_char_0048_0332.gif)
N–], 166.13 [CH![[double bond, length as m-dash]](https://www.rsc.org/images/entities/char_e001.gif)
–OH, aromatic carbon], 169.12 [![[C with combining low line]](https://www.rsc.org/images/entities/b_char_0043_0332.gif)
O, carbonyl carbon] (cf. Fig. 2c).
C show a break at a concentration corresponding to the CMC of the three cationic surfactants. From Fig. 3a, the value of CMC and the surface tension at the CMC (γCMC) are listed in Table 1. It can be found that the CMC values of prepared Schiff base cationic surfactants decrease by increasing the hydrophobic chain length (Fig. 3 inset); this can be attributed to increasing the hydrophobicity and decreasing the solubility of the prepared compounds, so the free energy of system increase, this lead to the surfactant molecules aggregate into micelles, so the hydrophilic group is directed toward the solvent while the hydrophobic chain is directed toward the interior of micelle in a way to avoid energetically unfavorable contact with the aqueous media, thereby reducing the free energy of system. Therefore, by increasing the hydrophobic chain length, the tendency of surfactant molecule to form micelle raises thus CMC decreased as shown in Fig. 3 inset.
| Compounds | Surface tension measurements | Conductivity measurements | ||||||||
|---|---|---|---|---|---|---|---|---|---|---|
| CMC (mM L−1) | γCMC (mN m−1) | ΠCMC (mN m−1) | Γmax × 1011 (mol cm−2) | Amin (nm2) | CMC (mM L−1) | α | β | ΔGomic (kJ mol−1) | ΔGoads (kJ mol−1) | |
| CSSB-10 | 3.16 | 35.94 | 36.06 | 8.10 | 2.057 | 3.19 | 0.379 | 0.621 | −23.50 | −27.93 |
| CSSB-14 | 0.43 | 33.04 | 38.96 | 7.50 | 2.21 | 0.48 | 0.396 | 0.604 | −31.32 | −35.70 |
| CSSB-16 | 0.08 | 31.64 | 40.36 | 5.50 | 3.02 | 0.11 | 0.408 | 0.592 | −37.84 | −43.57 |
The measurements of specific conductivity (K) were performed for the synthesized Schiff base cationic surfactants at certain temperature (30 °C) in order to evaluate the CMC and the degree of counter ion dissociation, α. From Fig. 3b, the degree of counter ion dissociation (α) obtained from the ratio of the slopes above and below the break indicative of the CMC were calculated and listed in Table 1. Normally, the degree of counter ion binding (β) and the degree of counter ion dissociation have the following relationship: β = 1 − α.33 The β is an important parameter since it is the expression of how many counter ions are contained in the Stern layer to counterbalance the electrostatic force that opposes to the micelle formation.34 All the values of α and β are listed in Table 1, it noted that β values decrease and α values increase with increasing hydrophobic alkyl chain length at the certain temperature.35 Also the values of CMC have been determined using electrical conductivity, were agreed with those obtained using surface tension.
| ΠCMC = γ0 − γCMC | (6) |
![]() | (7) |
log
C is the slop of γ vs. log
C profile at the point CMC; and n is the number of ions that originate in solution by dissociation of the surfactant and whose concentration vary at the surface when changing the bulk solution concentration (assumed as 2 in our calculations). The data in Table 1, revel that by increasing length of hydrophobic moiety of prepared cationic surfactants, shift Γmax to lower concentrations. Meaning that, the surfactant molecules are directed towards the interface, decreasing the surface energy of their solutions.
![]() | (8) |
ΔGomic = (2 − α)RT ln CMC
| (9) |
| ΔGoads = ΔGomic − 0.0602πCMCAmin | (10) |
![]() | ||
| Fig. 5 Potential–time curves for carbon steel in 3.5% NaCl+ 0.5 M HCl solution in the absence and presence of 150 ppm of synthesized CSSB inhibitors at 60 °C. | ||
![]() | ||
| Fig. 6 EIS plots for carbon steel in 3.5% NaCl+ 0.5 M HCl in the absence and presence of various concentrations of CSSB-10 exemplified as: (a) Nyquist and (b) Bode and phase modulus at 60 °C. | ||
![]() | ||
| Fig. 7 EIS plots for carbon steel in 3.5% NaCl+ 0.5 M HCl in the absence and presence of various concentrations of CSSB-14 exemplified as: (a) Nyquist and (b) Bode and phase modulus at 60 °C. | ||
![]() | ||
| Fig. 8 EIS plots for carbon steel in 3.5% NaCl+ 0.5 M HCl in the absence and presence of various concentrations of CSSB-16 exemplified as: (a) Nyquist and (b) Bode and phase modulus at 60 °C. | ||
The impedance diagrams show a capacitive loop, whose size increased with the addition of the synthesized inhibitors. The capacitive loop was related to charge transfer in the corrosion process.43 The shape of the EIS diagrams of C steel in 3.5% NaCl+ 0.5 M HCl is similar to those found in the presence of studied CSSB inhibitors, suggesting similar mechanism for the corrosion inhibition of C steel by synthesized CSSB inhibitors. Fig. 6–8a displayed that, the impedance value in the presence of CSSB inhibitors is larger than in the absence of inhibitors and the value of impedance increases with increasing CSSB compounds concentrations. This means that the corrosion rate is reduced in the presence of the inhibitors and continued to decreasing upon increasing the concentration of the inhibitors.
Fig. 6–8b show the Bode impedance and Bode phase angle plots for C steel electrode after its immersion in 3.5% NaCl+ 0.5 M HCl solution for 50 min. The Bode plots for C steel immersed in 3.5% NaCl+ 0.5 M HCl solution with and without different concentrations of CSSB inhibitors exhibit one time constant, thus showing that the inhibitor systems behaves as a monolayer formation and the dissolution process is controlled by a charge transfer reaction, taking place at the steel/solution interface. As we can see (Fig. 6–8b) increasing the concentration of CSSB compounds in the investigated solution results in closer to 90 degree of the phase angle indicating excellent inhibitive behavior due to adsorption of more CSSB molecules on the steel surface at higher concentrations. On the other hand, the shift of phase angle indicates the change in the electrode interfacial structure upon addition of the inhibitors. The continuous increase in the phase angle shift in the presence of the inhibitor was obviously associated with the growth of inhibitor film and with the increase in surface coverage (θ; Table 2) on the C steel surface, resulting in higher inhibition efficiency. Fig. 6–8b show that the impedance values in the presence of studied inhibitors (CSSB-10, CSSB-14 and CSSB-16) is larger than in its absence and the value of impedance was increased with increasing the inhibitor concentrations. This means that the corrosion rate is decreased by addition of CSSB compounds and continued to decreasing upon increasing the concentration of the inhibitors.
| Inhibitors code | Cinh/ppm by weight | Rs/Ω cm2 | Rp/Ω cm2 | CPE | Cdl/μF cm−2 | θ | PEIS/% | |
|---|---|---|---|---|---|---|---|---|
| Y0/μF cm−2 | n | |||||||
| Absence | 0 | 1.05 | 13.2 ± 1.46 | 156.69 | 0.84 ± 0.03 | 97.01 ± 7.27 | — | — |
| CSSB-10 | 25 | 1.12 | 19.4 ± 2.17 | 90.28 | 0.87 ± 0.02 | 51.19 ± 3.58 | 0.319 | 31.95 |
| 50 | 1.16 | 27.9 ± 1.93 | 62.77 | 0.86 ± 0.03 | 34.07 ± 3.06 | 0.526 | 52.68 | |
| 75 | 1.15 | 41.3 ± 2.80 | 42.40 | 0.88 ± 0.02 | 25.11 ± 1.88 | 0.680 | 68.03 | |
| 100 | 1.17 | 69.4 ± 4.32 | 25.23 | 0.89 ± 0.04 | 15.61 ± 1.24 | 0.809 | 80.97 | |
| 150 | 1.23 | 118.8 ± 6.38 | 14.74 | 0.90 ± 0.03 | 9.52 ± 0.85 | 0.888 | 88.88 | |
| CSSB-14 | 25 | 1.16 | 26.8 ± 2.10 | 65.35 | 0.89 ± 0.02 | 40.43 ± 2.42 | 0.507 | 50.74 |
| 50 | 1.13 | 38.7 ± 2.96 | 45.25 | 0.88 ± 0.03 | 26.81 ± 2.43 | 0.658 | 65.89 | |
| 75 | 1.17 | 62.6 ± 3.27 | 27.97 | 0.91 ± 0.02 | 18.88 ± 1.51 | 0.789 | 78.91 | |
| 100 | 1.45 | 93.1 ± 6.85 | 18.81 | 0.91 ± 0.02 | 12.70 ± 0.99 | 0.858 | 85.82 | |
| 150 | 1.50 | 213.2 ± 9.50 | 8.20 | 0.92 ± 0.03 | 5.78 ± 0.55 | 0.938 | 93.80 | |
| CSSB-16 | 25 | 1.23 | 30.2 ± 3.01 | 57.99 | 0.91 ± 0.04 | 39.15 ± 3.13 | 0.562 | 56.29 |
| 50 | 1.31 | 49.1 ± 4.96 | 35.67 | 0.90 ± 0.02 | 23.06 ± 2.07 | 0.731 | 73.11 | |
| 75 | 1.28 | 72.7 ± 4.62 | 24.09 | 0.92 ± 0.03 | 16.99 ± 1.01 | 0.818 | 81.84 | |
| 100 | 1.52 | 119.3 ± 7.10 | 14.63 | 0.93 ± 0.03 | 10.78 ± 0.87 | 0.889 | 88.93 | |
| 150 | 1.61 | 294.7 ± 12.65 | 5.94 | 0.93 ± 0.02 | 4.37 ± 0.35 | 0.955 | 95.52 | |
Fig. 9a–c shows the comparison of the simulated spectrum and the experimental EIS data, using the Z-View impedance fitting software for carbon steel samples immersed in 3.5% NaCl+ 0.5 M HCl containing 150 ppm of CSSB-16. A good fit was obtained with the model used for all experimental data. The equivalent circuit used in the present study is shown in Fig. 9d. In this equivalent circuit, Rs is the solution resistance, CPE is the constant phase element and Rp, which corresponds to the diameter of Nyquist's plot, includes the diffuse layer resistance (Rd), charge transfer resistance (Rct), the resistance of inhibitor layer at the steel surface (Ri) and the accumulated species at the steel/solution interface (Ra) (Rp = Rd + Rct + Ri + Ra).44 Similar equivalent circuits were proposed by several authors.44–46 In order to obtain a more accurate and representative fit, the pure double layer capacitance (Cdl) was replaced by the constant phase element (CPE) which is defined as follows:44
| ZCPE = Y0−1(jω)−n | (11) |
The data shown in Table 2 reveal that, introducing of CSSB compounds to the aggressive solution increases Rp and PEIS values and lowers the values of Cdl and this effect is seen to be increased as the concentrations of inhibitors increase. This suggests that the inhibitors act via adsorption at the steel/acid interface. The decrease in Cdl value can be attributed to the increase in the thickness of the electrical double layer and/or a decrease in the local dielectric constant. Further the decrease in the Cdl values is caused by the replacement of H2O molecules by the adsorption of the CSSB molecules on the metal surface, which decreases the extent of steel dissolution.47 Also, the increase of Rp with rise in the inhibitor concentration, indicate that the charge transfer process is mainly controlling the corrosion process. The value of Cdl is always smaller in the presence of the inhibitor than in its absence, indicating the effective adsorption of the synthesized CSSB inhibitors.
Based on the Nyquist plots shown in Fig. 6–8a and Table 2, the values of Rp and PEIS for the investigated inhibitors increase in the order: CSSB-16 > CSSB-14 > CSSB-10 with values 95.52% > 93.80% > 88.88% at 150 ppm. When this is compared to the blank, the Cdl values show an appreciable decrease in the reverse order. The concentration range of CSSB compounds from 25 to 150 ppm leads to the change of the inhibition efficiency from 31.95% to 88.88%, 50.74 to 93.80% and 56.29 to 95.52% for CSSB-10, CSSB-14 and CSSB-16, respectively, indicating that the CSSB compounds performs as a high effective inhibitors for C steel corrosion in 3.5% NaCl+ 0.5 M HCl solution.
![]() | ||
| Fig. 10 Anodic and cathodic polarization curves for carbon steel in 3.5% NaCl+ 0.5 M HCl containing different concentrations of (a) CSSB-10, (b) CSSB-14 and (c) CSSB-16 obtained at 60 °C. | ||
| Inhibitors code | Cinh/ppm by weight | Jcorr (μA cm−2) | CR/mmpy | −Ecorr/mV (SCE) | βa/mV dec−1 | −βc/mV dec−1 | θ | PPDP/% |
|---|---|---|---|---|---|---|---|---|
| Blank | 0.0 | 2450 ± 212 | 28.90 ± 2.5 | 439 ± 4 | 102 | 186 | — | — |
| CSSB-10 | 25 | 1643.7 ± 127 | 19.39 ± 1.49 | 427 ± 5 | 108 | 183 | 0.329 | 32.91 |
| 50 | 1120.6 ± 132 | 13.22 ± 1.55 | 429 ± 3 | 108 | 195 | 0.542 | 54.26 | |
| 75 | 733.2 ± 92 | 8.65 ± 1.08 | 440 ± 4 | 105 | 198 | 0.700 | 70.07 | |
| 100 | 406.4 ± 45 | 4.791 ± 0.53 | 435 ± 2 | 106 | 175 | 0.834 | 83.41 | |
| 150 | 207.1 ± 28 | 2.44 ± 0.25 | 438 ± 7 | 110 | 190 | 0.915 | 91.55 | |
| CSSB-14 | 25 | 1169.6 ± 112 | 13.79 ± 1.32 | 438 ± 6 | 108 | 197 | 0.522 | 52.26 |
| 50 | 787.1 ± 90 | 9.28 ± 1.06 | 443 ± 3 | 112 | 191 | 0.678 | 67.87 | |
| 75 | 458.6 ± 67 | 5.41 ± 0.79 | 434 ± 4 | 108 | 185 | 0.812 | 81.28 | |
| 100 | 283.9 ± 25 | 3.349 ± 0.29 | 449 ± 2 | 108 | 187 | 0.884 | 88.41 | |
| 150 | 82.8 ± 8 | 0.97 ± 0.094 | 435 ± 3 | 111 | 190 | 0.966 | 96.62 | |
| CSSB-16 | 25 | 1029.5 ± 97 | 12.14 ± 1.14 | 436 ± 3 | 106 | 199 | 0.579 | 57.98 |
| 50 | 604.9 ± 76 | 7.13 ± 0.89 | 442 ± 5 | 109 | 182 | 0.753 | 75.31 | |
| 75 | 384.4 ± 29 | 4.53 ± 0.34 | 441 ± 2 | 108 | 191 | 0.843 | 84.31 | |
| 100 | 205.3 ± 12 | 2.42 ± 0.15 | 443 ± 2 | 107 | 193 | 0.916 | 91.62 | |
| 150 | 39.4 ± 4 | 0.46 ± 0.04 | 439 ± 4 | 108 | 197 | 0.983 | 98.39 |
As shown in Table 3, the corrosion rates (CR) as well as corrosion current densities (Jcorr) of C steel reduce remarkably in the presence of the CSSB inhibitors, indicating a great inhibition performance for C steel corrosion in 3.5% NaCl+ 0.5 M HCl solution at 60 °C. Also the data showed that, the increasing of inhibitor concentrations resulted in a decrease in Jcorr, CR and an increase in inhibition efficiency (PPDP), suggesting the adsorption of CSSB molecule at the metal surface to form a protective film.49 According to Jcorr and CR the inhibitory action of CSSB-16 are much better than those of CSSB-10 and CSSB-14. The best inhibition efficiency was about 91.55% for CSSB-10, 96.62% for CSSB-14 and 98.39% for CSSB-16 at optimal concentration 150 ppm.
The cathodic Tafel slope (βc) and the anodic Tafel slope (βa) of CSSB-10, CSSB-14 and CSSB-16 did not modify with increase in inhibitor concentration, indicating that the hydrogen evolution is activation-controlled and the inhibition mechanism does not change by addition of investigated surfactants.50 On the other hand, βc of blank is higher than 118 mV dec−1. It indicates that in the absence of inhibitors, the corrosion reaction is specified by substance diffusion and charge transfer. Moreover, the behavior of βc and βa demonstrates that the CSSB compounds prevent the corrosion by blocking effect of adsorbed inhibitive species at the steel surface.
The inhibitor can be considered as anodic or cathodic type, if the displacement in Ecorr is more than ±85 mV relating to the corrosion potential of the blank. If the change in Ecorr is less than ±85 mV, the corrosion inhibitor may be considered as a mixed type.51 According to the values of Ecorr listed in Table 3, the largest displacement in the Ecorr values was found 12, 10 and 5 mV for CSSB-14, CSSB-14 and CSSB-16, respectively (<85 mV), indicating that the CSSB compounds can be classified as a mixed-type of inhibitors by showing its inhibitory action on both metal dissolution and hydrogen evolution.52 Based on Table 3, increasing the inhibitor concentration resulted in an increase in the surface coverage (θ) of electrode surface due to an increase in the amount adsorbed of the inhibitor on C steel surface suggests that a more compact and stable adsorption layer can be formed on the steel surface at higher concentration, thus prevents the steel surface from being attacked by the corrosive medium.
The inhibition effect of the precursor Schiff base namely 2-((3-(dimethylamino)propylimino)methyl)phenol (compound II) on the C steel corrosion in 3.5% NaCl+ 0.5 M HCl was performed, in comparison with the inhibition performance of the synthesized CSSB inhibitors by potentiodynamic polarization measurements at 60 °C. It was found that in the presence of 150 ppm of all studied compounds, the inhibition efficiency of the precursor compound is 86.5%, while the values for CSSB-10, CSSB-14 and CSSB-16 are 91.55%, 96.62% and 98.38, respectively. The higher PPDP values of the CSSB compounds in a comparison with the precursor Schiff base compound is due to the surfactant molecule up to CMC forms thin film on the metal surface involved two inhibitive groups; one hydrophilic group and other water-insoluble hydrophobic group. But Schiff base adsorbed on the surface of steel via only one inhibitive group.
When we compared our study with the literature data, e.g. Negm et al.53 studied the inhibition effect of 4-diethyl amino benzaldehyde Schiff base cationic amphiphiles on carbon steel in different acidic media. They found that, in the presence of 200 ppm of inhibitors the inhibition efficiency ranged from 69.5% to 96.64% and from 94.77 to 98.33 in 2 N H2SO4 and 2 N HCl solutions, respectively. Hamitouche et al.54 also investigated the inhibition performance of some quaternary ammonium surfactants on carbon steel in 1.0 M HCl containing 720 ppm of inhibitors using the potentiodynamic polarization measurements. They showed that, the inhibition efficiency of these inhibitors was around 85%. As another study, Hegazy et al.55 measured the behaviors of inhibition of cationic and gemini surfactants on carbon steel in acidic solution. The inhibition efficiencies increased with the increase in inhibitor dose and reached 94% and 96% for cationic and gemini surfactants in 0.5 M H2SO4 solution, respectively. Moreover, Fouda et al.56 investigated the inhibition efficiencies of some cationic surfactants, namely: cetyl trimethyl ammonium bromide (CTAB) and dodecyl trimethyl ammonium chloride (DTAC) in 0.5 M HCl. They found inhibition efficiencies were around 86.5% and 87.1%, respectively. Our potentiodynamic polarization measurements improve and support the literature investigations taking into account the chemical and physical meaning of the inhibition characteristics of CSSB compounds.
Finally, these results support CSSB compounds as economically and environmentally friendly inhibitors. The inhibiting efficiency values got from Tafel polarization measurements are comparable and run parallel with those obtained from the EIS data. The EIS and Tafel polarization results shown in Tables 2 and 3, respectively, display a relatively higher inhibition efficiency of CSSB-16 compared with those of CSSB-10 and CSSB-14.
![]() | (12) |
mol−1], T is absolute temperature [K], and A is an Arrhenius pre-exponential factor. Fig. 12 represents the Arrhenius plot of the corrosion of C steel in 3.5% NaCl+ 0.5 M HCl solution (log
Jcorr as a function of
) with or without the presence of 150 ppm of the investigated surfactants. The slope of the line is (−Ea/2.303R) and the intercept of the line extrapolated gives log
A. The calculated Ea value of C steel in the absence of inhibitors was 10.01 kJ mol−1. However, in the presence of 150 ppm of CSSB-10, CSSB-14 and CSSB-14 compounds, the Ea values were increased to 13.12, 15.33 and 16.62 kJ mol−1, respectively. The decrease in PPDP with rise in temperature with an increase in Ea in presence of surfactants compared to the absence signifies physical adsorption.60
| CSSB(sol) + xH2O(ads) → CSSB(ads) + xH2O(sol) | (13) |
![]() | (14) |
271 and 23
474 M−1 for CSSB-10, CSSB-14 and CSSB-16, respectively. The high values of Kads reflected the high adsorption ability of these CSSB inhibitors on C steel surface in the investigated aggressive solution. But the adsorption of CSSB-16 inhibitor was more efficient than CSSB-10 and CSSB-14 compounds. Kads is also related to the standard free energy of adsorption (ΔGoads) by the following equation:17
ΔGoads = −RT ln(55.5Kads)
| (15) |
![]() | ||
| Fig. 13 Langmuir adsorption isotherms for the adsorption of CSSB inhibitors on carbon steel in 3.5% NaCl+ 0.5 M HCl solution at 60 °C. | ||
![]() | ||
| Fig. 14 SEM micrograph of carbon steel electrode surface after immersion in 3.5% NaCl+ 0.5 M HCl for 72 h (a) and EDX of the electrode at the same conditions (b). | ||
Fig. 14a shows the SEM image of C steel surface after immersion in uninhibited 3.5% NaCl+ 0.5 M HCl solution for 72 h. The micrographs showed that the C steel specimen exhibits a very rough surface and covered with thick porous oxide layer in the absence of inhibitors due to corrosive attack by the chloride acid solution. In aggressive solution-free inhibitor (uninhibited), the EDX spectra (Fig. 14b) of the tested C steel sample showed, in addition to the O and Cl, the atomic percentage of the alloying elements constituting each tested sample. The presence of the O may be due to the presence of Fe3O4 and α-FeOOH as corrosion products.63 The presence of Cl refers to the adsorption of Cl ions and subsequent formation of FeCl2 as one of the corrosion products. This indicated that the passive film is mainly Fe2O3 and/or FeCl2.
Fig. 15a describes the SEM of the C steel surface after 72 h of immersion in 3.5% NaCl+ 0.5 M HCl solution with the addition of 150 ppm CSSB-16 compound. It can be seen that the flakes in the surface of the specimens decreased as compared to that of the micrograph in Fig. 14a, where the surface of 15a is clearer as compared to that of 14a. The improvement in surface morphology of specimens is due to the decrease in the corroded areas caused by the inhibitor layer covering the electrode surface. In the inhibited solution, the EDX spectrum (Fig. 15b) showed additional characteristic peaks for the existence of N due to the adsorption of imine group. In addition, the contribution of C enhanced as a result of the C atoms of the adsorbed CSSB-16 inhibitor. This data show that a carbonaceous material containing N has covered the electrode surface. The detection of the N atom and the high contribution of the C observed in presence of additives, indicating the surface are covered with CSSB-16 inhibitor film. Further inspection of the EDX data demonstrated that the contribution of Cl and O elements were decreased relative to the specimens exposed to the free inhibitor solutions because of the covered inhibitor layer, which protects metal against corrosion. The percentages of the elements present on the analyzed surfaces with and without the titled inhibitor are calculated and tabulated in Fig. 14 and 15b (inset).
The synthesized CSSB compounds adsorbed to the C steel surface through Cl− ion and the quaternary nitrogen atom (N+). Cl− ion adsorbed on the anodic sites to minimize the anodic dissolution while N+ adsorbed on the cathodic sites to decrease the hydrogen evolution. The adsorption of CSSB compounds on anodic site occurred via lone pair of electrons of nitrogen atoms of azomethine group (–CH
N–) and π-electrons of aromatic ring which decreased the anodic dissolution of C steel.
In addition to the chemical adsorption, the surfactant molecules can also be adsorbed on the metal surface through electrostatic interaction between the charged surfactant molecules and the charged steel surface. The following mechanism is proposed for the corrosion of the C steel in the investigated chloride medium in the presence of inhibitors:
| Fe + Cl− ↔ (FeCl)ads− | (16) |
| CSSB + H+ ↔ CSSBH+ | (17) |
| (FeCl)ads− + CSSBH+ ↔ (FeCl−CSSBH+)ads | (18) |
In acidic chloride environment, CSSB is protonated at the nitrogen atom present in azomethine group (–CH
N–), and becomes a cation that exists in equilibrium with the corresponding molecular form as eqn (17). The charge on the steel surface can be measured from the potential of zero charge (EPZC) on the correlative scale (Ø), given by the equation:66
| Ø = Ecorr − EPZC | (19) |
In the present study, the obtained values of EPZC for C steel in acidic chloride solution in the presence of 150 ppm CSSB-10, CSSB-14 and CSSB-16 are −0.491, −0.509 and −0.487 V vs. SCE, respectively. It can be said that Ø potential is positive in these cases. From the above result, it follows that Cl− ions first adsorbed at the steel/solution interface at the Ecorr. After this first adsorption step, the C steel surface becomes negatively charged. Hence, the positively charged of CSSB cationic forms [CSSBH+] has been formed an electrostatic bond with the Cl− ions already adsorbed on steel surface as eqn (18). As a result, it can be concluded that CSSB compounds acts as an excellent mixed type acid corrosion inhibitors for C steel in the investigated acidic solution containing chloride.
![]() | ||
| Fig. 16 Optimized structures and frontier molecular orbitals (HOMO and LUMO) for the studied CSSB compounds, (a) CSSB-10, (b) CSSB-14 and (c) CSSB-16. | ||
| Quantum chemical parameters | CSSB-10 model | CSSB-14 model | CSSB-16 model |
|---|---|---|---|
| μ/Debye | 7.98 | 8.01 | 8.04 |
| EHOMO/eV | −4.9796 | −4.8708 | −4.7769 |
| ELUMO/eV | 0.81634 | 0.76191 | 0.7563 |
| ΔE/eV | −5.79545 | −5.63271 | −5.53320 |
| η/eV | 2.0811 | 2.054 | 2.0103 |
| σ/eV−1 | 0.480 | 0.486 | 0.497 |
| pi/eV | −2.92 | −2.90 | −2.82 |
| EA/eV | −0.81634 | −0.76191 | −0.7563 |
| EN/eV mol−1 | 4.209 | 4.201 | 4.108 |
| ECP/eV mol−1 | −4.209 | −4.201 | −4.108 |
The frontier molecular orbitals (HOMO and LUMO) perform significant roles in predicting the adsorption centers of the inhibitor molecules responsible of the interaction metallic surface/molecule.67–69 From Fig. 16, the HOMO is distributed over the imine group (–C
N–) due to the presence of lone electron pairs in the nitrogen atom, which indicates that the preferred active sites for an electrophilic attack are located within the region around nitrogen atom belonging to –C
N– group. Whereas, The LUMO is mainly located in the benzene ring, which reveals that the preferred active sites for a nucleophilic attack are located in π-electrons of benzene ring. These results suggest that the –C
N– benzene region in the inhibitor is the likely reactive site for the adsorption of molecule on the metal surface.
The energy of the highest occupied molecular orbital (EHOMO) is related to the electron donating ability of the compound; therefore, compounds with high values of EHOMO have a tendency to donate electrons to appropriate acceptor with low empty molecular orbital energy. On the other hand, the ELUMO indicates the electron accepting ability of the molecule, the lowest its value the higher the ability of accepting electrons. Therefore a decreasing ELUMO suggests better inhibition efficiency.70 From the results obtained, the expected trend for inhibition efficiency of the titled CSSB compounds is CSSB-16 > CSSB-14 > CSSB-10; this trend is in agreement with experimental data. The energy gap (ΔE) between ELUMO and EHOMO energies is another important factor in describing the molecular activity, so when the ΔE decreased, the inhibitor efficiency is improved.71 Therefore, the inhibition efficiency of the studied CSSB compounds is expected to increase with decreasing value of ΔE. This assertion is supported by the experimental results. The trend for the ΔE values follows the order CSSB-16 < CSSB-14 < CSSB-10 which suggests that CSSB-16 has the highest reactivity in comparison to the other inhibitors and would therefore likely interact strongly with the metal surface. From the above results, it can be seen that there is a strong correlation between the experimental inhibition efficiency values and the frontier molecular orbital energies (ELUMO, EHOMO and ΔE).
The dipole moment (μ) is an index that can also be used for the prediction of the direction of a corrosion inhibition process. μ is the measure of polarity in a bond and is related to the distribution of electrons in a molecule.72 The relationship between inhibition efficiencies and the dipole moment of similar molecules have often given results that are not univocal, i.e., in some cases the μ appears to decrease with increase in the inhibition efficiency of the inhibitors73 while in other systems the μ appears to increase with increasing inhibition efficiency.74 The dipole moments of CSSB-10, CSSB-14 and CSSB-16 are 7.98, 8.01 and 8.04 Debye, respectively, which are higher than that of H2O (μ = 1.88 Debye). The high μ values of these compounds probably indicate strong dipole–dipole interactions between inhibitor molecule and steel surface.75 Overall, the trend in the dipole moment follows the order; CSSB-16 > CSSB-14 > CSSB-10, which is in agreement with trends in the experimental inhibition efficiencies of the compounds.
The electron affinity (EA) of molecules is an intricate function of their electronic structure. EA is related directly to ELUMO. It is showed that from Table 4, there is a good correlation between the inhibition efficiency and the electron affinity for the investigated surfactant inhibitors. Electron affinity values for the three synthesized surfactants are negative, indicating that their inhibition performance may be related to the tendency of the molecules to be electrophilic. As the electron affinity increase along the compounds, the affinity of the compounds to accept electrons from the surface of metal into the inhibitor antibonding orbital increase and the energy given off increase. Then the inhibition efficiency increase indicating more protection for the steel surface. As a result, higher negative value of EA for CSSB-16 compound (−0.7563) indicates that the molecule strongly adsorb onto the metal surface and form a more protection layer on the C steel surface.76 One of the important calculated quantum chemical indices of the investigated inhibitors are electronic chemical potential (ECP) and electronegativity (EN). Electronegativity is related to the ability of the molecule to draw electron toward itself.77 The data in Table 4 displayed that, CSSB-16 has the highest probability (4.209 eV mol−1) to form a coordinating bond by accepting an electron from the steel surface (7 eV mol−1), and conversely, the electronic chemical potential (ECP) of the inhibitors is larger than the metallic Fe (−7 eV mol−1).
The chemical hardness and softness of a molecule are parameters that have been found to exhibit excellent relationship with the energy gap (ΔE). Chemical hardness (η) measures the resistance of an atom to a charge transfer; whereas, the chemical softness (σ), describes the capacity of an atom or group of atoms to receive electrons and calculated by the following equations:78
![]() | (20) |
![]() | (21) |
Generally, hard molecules are characterized with larger value of ΔE and are less reactive than soft molecules, which are characterized by small ΔE.78 Soft molecules are more reactive than hard ones because they could easily offer electrons to an acceptor. For the simplest transfer of electrons, adsorption could occur at the part of the molecule where σ has the highest value and η the lowest value.79 Therefore, as expected, CSSB-16 is the best inhibitor with the lowest chemical hardness. It was also found that CSSB-16 has the lowest electronic chemical potential (pi, −2.82 eV), which imply that CSSB-16 compound has better inhibition performance. From the above, experimental inhibition efficiencies of the studied CSSB compounds correlated excellently with EHOMO, ELUMO, ΔE, η, EA, EN, ECP and μ and both experimental and theoretical results point to the fact that CSSB-16 has the highest inhibition efficiency.
1. The CMC values of CSSB compounds decreased with increasing of the length of the hydrophobic group.
2. All synthesized surfactants show good inhibition performance for carbon steel in the investigated acidic chloride solution; the inhibition efficiency increases with an increase in inhibitor concentration and decreases with increasing temperature.
3. The potentiodynamic polarization curves show that the synthesized CSSB surfactants act basically as a mixed type inhibitors.
4. EIS results indicate that the resistance of the carbon steel electrode greatly increased in the presence of the synthesized surfactants and are considered as efficient inhibitors referring to their corrosion inhibition efficiency values which ranged 88.88–95.52% at 150 ppm.
5. The Langmuir adsorption isotherm exhibited the best fit to the experimental data with ΔGoads of −35.72, −38.18 and −38.98 kJ mol−1 for CSSB-10, CSSB-14 and CSSB-16, respectively, which indicate that the adsorption process involved both the chemical and physical adsorption.
6. SEM/EDX analyses indicate that carbon steel corrosion can be inhibited, obviously due to the adsorption of the synthesized CSSB compounds on the C steel surface.
7. The computed quantum chemical properties viz., ELUMO, EHOMO, the dipole moment (μ), ΔE (ELUMO − EHOMO), the chemical hardness (η), softness (σ), electron affinity (EA), electronegativity (EN), electronic chemical potential (ECP) and electronic chemical potential (pi) show good correlation with experimental inhibition efficiency.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra21626e |
| This journal is © The Royal Society of Chemistry 2016 |