Startseite Using electrochemical testing and modeling to assess the efficiency of a water-soluble inhibitor on the corrosion of 6061Al-15 % (v) SiC (p) composite
Artikel Öffentlich zugänglich

Using electrochemical testing and modeling to assess the efficiency of a water-soluble inhibitor on the corrosion of 6061Al-15 % (v) SiC (p) composite

  • Maithili Krishnananda , Prakasha Shetty ORCID logo EMAIL logo , Preethi Kumari und Sneha Kagatikar
Veröffentlicht/Copyright: 19. Mai 2023

Abstract

The corrosion behaviour of 6061Al-15 % (v) SiC (P) composite was investigated in a 0.5 M HCl medium using electrochemical techniques. A water-soluble inhibitor, N-(1-morpholinobenzyl) semicarbazide (NMSc), was synthesised to test its corrosion inhibition activity on 6061Al-15 % (v) SiC (P) composite. The inhibitor showed cathodic-type behaviour and 95.42 % inhibition efficiency at 2.56 mM concentration and 303 K temperature. The thermodynamic results revealed the inhibitor’s physisorption, which followed Langmuir’s isotherm model. A drastic reduction in corrosion current density in the inhibited medium indicates that the inhibitor effectively controls the deterioration of the composite in the HCl medium. A significant rise in polarisation resistance on increasing inhibitor concentration offers greater resistance for the charge transfer process, ensuring better control over the composite corrosion. The surface analysis by scanning electron microscopy (SEM) and atomic force microscopy (AFM) confirms the inhibitor film formation on the composite. The quantum chemical theoretical calculations supported the electrochemical results.

1 Introduction

Aluminium and its alloys find application in various fields because of their low density, lightweight, good thermal conductivity and corrosion resistance properties (Davis 1993). They resist corrosion in the air because of the protective oxide film formation on the surface (Bockris and Minevski 1993; de Wit and Lenderink 1996). Hence, they find wide domestic and industrial applications. However, Al alloys have a low strength-to-weight ratio, making them prone to failure in high-load applications. It is controllable by adding a suitable particulate material (SiC, graphite or Al2O3), resulting in an Al alloy composite (Rohatgi 1993). One such composite material can be produced using an Al alloy as the matrix phase and SiC particulate as the dispersed phase (Pardo et al. 2005; Reddy and Zitoun 2010). Their unique properties (high strength-to-weight ratio, elastic modulus, surface durability and impact strength) are useful in aviation, automobile and military applications (Singla et al. 2009). However, adding SiC as a reinforcing material leads to surface discontinuity, generating more active corrosion sites (Bobic et al. 2010; Pinto et al. 2009; Trowsdale et al. 1996). Hydrochloric acid solutions are widely employed in automobile and aero industries as pickling agents for Al-based alloys and composites (Sheasby and Pinner 2001). Al alloy composites usually undergo pitting corrosion in an aggressive acid medium (Bhat et al. 1991; Shetty 2022). In such cases, it is advantageous to use inhibitors to control corrosion effectively. Therefore, selecting a suitable material as a corrosion inhibitor is essential.

Organic compounds having heteroatoms (N, O and S) and aromatic rings facilitate their adsorption on the metal surface, acting as efficient corrosion inhibitors (Shetty 2020, 2022). Because of their excellent inhibition activity and simple synthesis with low-cost starting materials, Mannich bases are gaining importance as inhibitors for controlling the corrosion of metals such as mild steel (Ahamad et al. 2010; Jeeva et al. 2017; Quraishi et al. 2008; Singh and Quraishi 2010) in acid medium. Hence, in the present work, a simple Mannich base, N-(1- morpholinobenzyl) semicarbazide (NMSc), was synthesised and tested for its corrosion inhibition activity on 6061-15 % (v) SiC (P) composite in 0.5 M HCl. The inhibitor compound, NMSc, is easy to prepare and readily soluble in a 0.5 M HCl medium without adding an organic solvent. Hence, it is an economical and environmentally friendly corrosion inhibitor. NMSc exhibited excellent inhibition performance (IE of 95 %).

NMSc’s inhibitory performance is compared with inhibitors of 6061Al-SiC composite previously reported. Some of these inhibitors are not freely soluble in the hydrochloric acid medium and hence require the usage of a suitable organic solvent (ethanol/acetone) to prepare the solution of the inhibitor (Kini et al. 2010, 2011; Shetty et al. 2020), which can lead to environmental pollution. Some inhibitors are soluble in a hydrochloric acid medium (Charitha and Rao 2017, 2018a, 2018b, 2020). However, most of these inhibitors’ inhibition performance is not significant, and some are useful only for a lower concentration of hydrochloric acid media (˂0.5 M HCl).

2 Materials and methods

2.1 Materials

6061Al-15 % (v) SiC (p) composite (represented hereafter words as Al alloy-SiC composite) was used as the specimen material in this work. The composite was prepared using the stir-casting technique by reinforcing 6061Al alloy with SiC particulate (purity 99.9 % and size 23 μm). 6061Al base alloy used has the composition: Mg (1.0 wt%); Si (0.60); Cu (0.25); Cr (0.25); and Al (97.9). The test coupons were prepared using a rod of Al alloy-SiC composite and then embedded with epoxy resin under cold conditions, leaving an uncovered surface of 1 cm2 at one end. The specimen’s exposed surface was abraded with emery papers (grades 200 to 1500) and finely on a disc polisher. Then the sample was washed using deionised water, and acetone was rinsed and dried.

2.2 Preparation of medium

Hydrochloric acid solutions (0.5 M) were prepared from 37 % hydrochloric acid (AR grade) on dilution using double-distilled water and standardised by a volumetric method.

2.3 Synthesis and characterisation of NMSc

N-(1-morpholinobenzyl) semicarbazide (NMSc) was synthesised as per the reported procedure (Viswanathan 2016). Semicarbazide hydrochloride (1.11 g, 10 mM) was dissolved in 20 mL ethanol and neutralised with a dropwise addition of NH3. To this solution, morpholine (1 mL, 10 mM) and benzaldehyde (1 mL, 10 mM) were added dropwise on constant stirring. A colourless product precipitated was filtered. The separated product was recrystallised from ethanol. The molecular structure of NMSc is shown in Figure 1. The recrystallized sample’s IR spectra were recorded by a Shimadzu FTIR 8400S spectrophotometer in a 4000–400 cm−1 frequency range using KBr. 1H-NMR spectra were recorded in DMSO‑d6 by a Bruker 400 MHz NMR spectrophotometer.

Figure 1: 
						Molecular structure of N-(1-morpholinobenzyl) semicarbazide (NMSc).
Figure 1:

Molecular structure of N-(1-morpholinobenzyl) semicarbazide (NMSc).

Colourless crystalline solid, yield: 80 %; C12H18N4O2: m.p: 216 ± 1 °C; IR (KBr) [cm−1]: 1141 (C–O str.), 1600 (C=O str.), 1643 (C=N str.), 3028 (Ar. CH str.), and 3288 (NH str.). 1H-NMR (400 MHz, DMSO‑d6) (⸹ppm): 2.5 (4H, –N–CH2), 3.74 (4H, –O–CH2), 6.23 (1H, NH), 6.32–7.72 (5H, Ar. H), 6.7 (1H, –CH), 7.83 (2H, –NH2), 10.94 (1H, –N=C–OH). Elemental analysis, Calcd. (Found) for C12H18N4O2: C, 57.14 (57.28); H, 6.88 (6.92); N, 21.62 (21.52).

2.4 Electrochemical techniques

Electrochemical experiments were conducted with CH Instruments (CH600D-series, U.S. model). A three-electrode cell containing a saturated calomel electrode, the Al alloy-SiC composite specimen’s working electrode and an auxiliary electrode of Pt were used. Experiments were conducted with three trials, and the average value was considered for the calculation. Initially, the three-electrode system dipped in 100 mL of 0.5 M HCl solution in a 250 mL beaker was allowed to attain the steady state open circuit potential (OCP). The experiment was conducted at 303, 313 and 323 K temperatures using a water bath connected to a calibrated thermostat. The electrochemical impedance tests were performed at OCP in the 10 kHz–0.01 Hz frequency range by applying an AC signal of 10 mV amplitude. Nyquist impedance data plots were recorded and analysed using the Zsimpwin program. At OCP, the potentiodynamic polarisation (PDP) tests were performed by varying potential from −250 to +250 mV with a 1 mV/s scanning rate. From the recorded Tafel plots, Ecorr (corrosion potential, V), icorr (corrosion current density, mA/cm2), the anodic (βa) and cathodic (βc) slopes were obtained.

2.5 Specimen’s surface analysis

Al alloy-SiC specimen’s surface morphology was analysed after immersing it in 0.5 M hydrochloric acid solution without and with NMSc for 3 h using an SEM (JEOL JSM-6380L model). Pre-and post-inhibition roughness of the specimen surface was recorded using AFM (IB342-Innova model).

2.6 UV-visible spectral analysis

UV-visible spectra of NMSc (0.01 mM) in 0.5 M HCl solution were taken by a UV-visible spectrophotometer (1800 Shimadzu). The Al alloy-SiC composite test coupon was immersed in this solution for 3 h. After removing the test coupon, the absorption spectra of the solution were recorded.

2.7 Theoretical studies

The theoretical calculations by density functional theory (DFT) were made with the help of Schrodinger’s software with B3LYP functions and a 631G** basis set. The optimised structure of the inhibitor molecule, HOMO (highest occupied molecular orbital) and LUMO (lowest unoccupied molecular orbital) were recorded. Different theoretical parameters were obtained, including energies of HOMO and LUMO, energy band gap, the fraction of electrons transferred, hardness and softness values.

3 Results and discussion

3.1 Electrochemical results

3.1.1 PDP measurements

The PDP plots for Al alloy-SiC composite in 0.5 M HCl medium in the absence and presence of NMSc at 303 K are shown in Figure 2. Polarisation parameters obtained from these plots include corrosion potential (Ecorr, V), corrosion current density (icorr, mA/cm2), anodic slope (βa, V dec−1) and cathodic slope (βc, V dec−1).

Figure 2: 
							The PDP plots for Al alloy-SiC in acid and inhibited acid solutions at 303 K.
Figure 2:

The PDP plots for Al alloy-SiC in acid and inhibited acid solutions at 303 K.

The corrosion rate (CR, mmy−1) of the composite specimen was given by Equation (1).

(1)CR(mmy1)=3270×M×icorrd×Z

where 3270 is the unit conversion constant, Z is the number of electrons transferred per metal (3 for Al), M is the atomic mass (9.15) and d (2.66 g/cm3) is the density of the metal (Fontana 1987).

The inhibition efficiency (IE (%)) of NMSc was determined according to Equation (2).

(2)IE(%)=icorricorr(inh)icorr×100

where icorr represents the corrosion current density in acid, while icorr(inh) is in an inhibited acid medium (Fontanna 1987).

The PDP parameters obtained (Table 1) indicate the decrease in icorr and CR of the composite specimen in the inhibited acid medium because of the NMSc’s adsorption on the composite surface. According to Table 1, the IE of NMSc increases with increasing concentration and decreases with increasing medium temperature. It presumably indicates the physisorption of NMSc molecules on the composite surface (Nagalaxmi et al. 2020). NMSc showed a maximum IE of 95.42 % at 2.56 mM concentration and 303 K. The metal oxidation process is represented by the anodic curve of the potentiodynamic polarisation plot (Figure 1), while the cathodic curve represents the hydrogen evolution phenomenon.

Table 1:

PDP results for Al alloy-SiC in the blank and inhibited medium at varying temperatures.

T (K) C Inh (mM) E corr (V vs. SCE) i corr (mA/cm2) β a (mV dec−1) β c (mV dec−1) CR (mm y−1) IE (%)
303 0.0 −0.676 9.440 4.694 4.861 103.3
0.01 −0.682 6.135 5.025 5.660 67.08 35.01
0.04 −0.685 5.180 4.939 6.241 56.64 45.13
0.16 −0.692 2.711 5.279 6.893 29.64 71.28
0.64 −0.695 1.466 5.492 7.509 16.03 84.47
2.56 −0.712 0.4319 8.879 7.845 4.72 95.42
313 0.0 −0.676 13.15 4.689 4.612 143.8
0.01 −0.691 11.05 4.731 5.216 120.8 15.97
0.04 −0.691 8.024 4.703 5.653 87.73 39.00
0.16 −0.695 5.403 4.921 6.137 59.08 58.91
0.64 −0.696 4.478 5.129 6.197 48.97 65.95
2.56 −0.696 3.020 5.300 6.407 33.02 77.03
323 0.0 −0.692 21.30 4.653 4.007 232.9
0.01 −0.704 18.93 4.599 4.543 207.1 11.13
0.04 −0.704 16.71 4.565 4.840 182.8 21.55
0.16 −0.704 12.59 4.626 5.208 137.6 40.89
0.64 −0.703 10.45 4.630 5.368 114.2 50.94
2.56 −0.705 6.017 5.032 5.984 65.8 71.75

It is clear from Table 1 that there is a slight variation in the Tafel slopes (βa and βc) in inhibited acid medium, indicating the NMSc’s adsorption, which blocks the surface available for corrosion reducing the corrosion rate without altering the corrosion mechanism (Pinto et al. 2011; Yurt et al. 2006). However, a significant variation in the βa value of the anodic curve corresponding to a 2.56 mM concentration of NMSc at 303 K is observed (Table 1). The plateau of this anodic curve is unclear (Figure 1). It may be due to the kinetic barrier effect occurring because of either the partial breakdown or the formation of porosity on the passivated layer. Because of the passivation of the metal in the range of −0.7 to −0.6 V, a considerable change in the anodic Tafel slope (βa) was observed (Charitha and Rao 2018b). Since the anodic current plateau is unclear, the corrosion current densities were calculated by extrapolating a cathodic slope.

As reported elsewhere (Li et al. 2008), an inhibitor can be classified as an anodic or cathodic type if the shift in Ecorr value in the inhibited medium is more than ±85 mV than that in the blank medium. In this work, the observed change in Ecorr value is – 0.36 mV indicating that NMSc acts as a mixed-type inhibitor showing predominant control on cathodic reaction.

3.1.2 EIS measurements

The impedance plots (Figure 3a) obtained are depressed semi-circular, indicating the charge transfer nature of the corrosion phenomenon (Lenderink et al. 1993). There is an increase in the diameter of impedance curves in the inhibited acid medium, which indicates that corrosion inhibition occurs through NMSc adsorption at the composite surface. The impedance plot (Figure 3a) contains an inductive loop and a depressed capacitive loop at the lower and higher frequency regions. The in-homogeneities on the composite surface results in frequency dispersion (Metikos-Hukovic et al. 2002). A similar impedance plot was reported earlier on Al alloy/composite corrosion in HCl solution (Kumari et al. 2020; Noor 2009).

Figure 3: 
							Al alloy-SiC composite’s (a) impedance plot at 303K and (b) stimulating plot with the equivalent circuit.
Figure 3:

Al alloy-SiC composite’s (a) impedance plot at 303K and (b) stimulating plot with the equivalent circuit.

The capacitive loop at the higher frequency regions could be accountable for the charge transfer during corrosion and oxide layer formation covering the composite surface (Metikos-Hukovic and Babic 1998). Brett 1992, reported that the capacitive loop attributes to interfacial reactions. At the metal oxide/electrolyte interface, Al+ ions are formed, which migrate through the interface and oxidise to Al3+ ions with simultaneous formation of OH or O2− ions. The inductive loop observed at a lower frequency may be due to the surface relaxation process, indicating the inhibitor adsorption on the oxide layer (Abd El Rehim et al. 2002). The capacitive loop’s diameter increases as the concentration of NMSc rises, revealing the reduction in the relaxation process and an increase in the corrosion resistance of the composite due to NMSc adsorption (Umoren et al. 2010).

The EIS data obtained for the Al alloy-SiC composite in the presence of an inhibitor was checked for fitment to an appropriate equivalent circuit (Figure 3b) utilizing Zsimpwin software (version 3.21). It is evident from Table 2 that the values of χ2 are in the order of 10−3, which authenticates the selected equivalent circuit. The equivalent circuit depicted in Figure 3b comprises elements, Rct (charge transfer resistance), Q (constant phase element, CPE), Rs (solution resistance) and RL and L (inductive elements). The elements Q, Rct and RL are parallel in the equivalent circuit, but L and RL are in series. CPE (Q) is utilised as a substitute for the ideal capacitive element (C) in simulation analysis to account for depressed capacitive loops. The non-uniformity of the composite electrode surface leads to frequency dispersion, resulting in depressed semicircle impedance plots and deviations from the ideal behaviour (Jüttner 1990). The real value of Cdl (double layer capacitance) in the existence of frequency dispersion can be calculated using Equation (3) (Babic et al. 1998).

(3)Cdl=Qdl(ωmax)n1

In this relation, Qdl is the CPE constant, ωmax = 2πfmax, fmax indicates the frequency where the −Z″ reaches the maximum and n is a CPE exponent determined by the morphology of the composite electrode surface. Its value lies in the range −1 ≤ n ≤ 1, and CPE shows an ideal behaviour as a capacitor when n = 1.

Table 2:

EIS results for Al alloy-SiC in the blank and inhibited solutions at different temperatures.

T (K) C Inh (mM) R L (Ω cm2) R ct (Ω cm2) R p (Ω cm2) n χ 2 (×10−3) C dl (μF cm−2) IE (%)
303 Blank 19.27 7.62 5.45 0.85 4.789 3056
0.01 25.64 12.86 8.56 0.83 4.436 1361 36.33
0.04 27.45 17.40 10.65 0.82 2.396 1113 48.83
0.16 75.34 37.46 25.02 0.81 3.655 376 78.22
0.64 85.98 61.19 35.75 0.81 6.066 179 84.76
2.56 222.18 98.78 68.38 0.80 2.602 91.23 92.03
313 Blank 15.67 6.35 4.52 0.85 8.226 5069
0.01 17.98 9.85 6.36 0.82 5.348 3638 28.93
0.04 20.12 13.19 7.97 0.82 3.676 2285 43.29
0.16 25.67 20.95 11.54 0.81 3.693 1355 60.83
0.64 35.96 28.79 15.99 0.81 3.547 938 71.73
2.56 48.56 45.75 23.56 0.81 3.248 555 80.81
323 Blank 11.45 4.93 3.44 0.83 4.300 10,339
0.01 13.98 6.00 4.20 0.82 4.930 8104 18.09
0.04 16.54 6.76 4.80 0.82 4.954 6756 28.33
0.16 21.57 8.77 6.24 0.82 6.626 3895 44.87
0.64 26.65 10.54 7.55 0.82 5.427 2993 54.44
2.56 34.54 19.14 12.31 0.82 7.589 1291 72.05

The relation, Equation (4), was used to calculate the polarisation resistance (Rp) value (Noor 2009)

(4)Rp=RLRctRL+Rct

The inhibition efficiency of NMSc (IE, %) were obtained using Equation (5) (Noor 2009)

(5)IE(%)=Rp(inh)RpRp(inh)×100

R p(inh) refers to polarisation resistance in the inhibited solution and Rp in the acid solution. Table 2 displays the parameters obtained from the EIS measurement for the Al alloy-SiC composite in a 0.5 M HCl medium. It is evident from Table 2 that polarisation resistance (Rp) and charge transfer resistance (Rct) increase with a rise in NMSc concentration, indicating that the formation of adsorbed inhibitor film resists the charge transfer process, controlling corrosion of the composite in the acid medium. These results suggest that NMSc undergoes physisorption. The electrical double layer developed at the charged composite surface-medium interface could be taken as a capacitor. By replacing the initially adsorbed water molecules and other ions on the surface, NMSc adsorption on the composite electrode reduces its electrical capacity. The decreased capacity could be attributed to forming of a protective film on the composite electrode surface (Kedam et al. 1981). The decrease in Cdl values with increasing NMSc concentration is primarily due to the improved thickness of the electrical double-layer at the composite/acid medium interface (Bessone et al. 1992).

3.2 Kinetic parameters

The inhibition efficiency of NMSc decreased with a temperature rise of the medium, probably because of the short time lag between the adsorption/desorption process of NMSc molecules on the surface of the composite that occurs with the temperature rise. It reveals the physisorption of NMSc molecules on the composite surface (Nagalaxmi et al. 2020). A maximum IE of 95.42 % was displayed at 2.56 mM NMSc and 303 K. The effect of temperature on the extent of NMSc adsorption on the composite surface helps to obtain kinetic data like Ea (activation energy), ΔHa (enthalpy of activation) and ΔSa (entropy of activation).

E a value in the acid and the inhibited acid solution was obtained from Equation (6) (Schorr and Yahalom 1972)

(6)ln(CR)=BEaRT

B indicates the Arrhenius constant, R refers to the gas constant (8.314 J K−1 mol−1) and T denotes the temperature in K.

The Arrhenius plot for Al alloy-SiC composite in the blank and inhibited solution (Figure 4a) depicts a straight line, and its slope equals (−Ea/R). Hence, Ea values in acid and inhibited acid solutions were calculated. With the help of the transition state equation [Equation (6)], ΔHa (Enthalpy of activation) and ΔSa (Entropy of activation) values were calculated (Abd El-Rehim et al. 1999).

(7)CR=RTNhexp(SaR)exp(HaRT)

where h refers to Planck’s constant and N denotes Avogadro’s number.

Figure 4: 
						Al alloy-SiC composite’s  (a) Arrhenius plot and  (b) ln (CR/T) vs. 1/T plot.
Figure 4:

Al alloy-SiC composite’s  (a) Arrhenius plot and  (b) ln (CR/T) vs. 1/T plot.

Figure 4b depicts a straight-line graph of ln(CR/T) versus 1/T for Al alloy-SiC composite. The slope of the straight line is (−ΔHa/R), and the intercept is [ln(R/Nh) + ∆Sa/R]. As a result, ΔHa and ΔSa values were computed. Table 3 shows the activation parameters for Al alloy-SiC composite in acid and inhibited acid solutions.

Table 3:

Activation parameters for Al alloy-SiC in the blank and inhibited medium.

C Inh (mM) E a (kJ mol−1) R 2 ΔH a (kJ mol−1) ΔS a (J mol−1 K−1)
0 33.04 0.984 30.43 −106.11
0.01 45.86 0.999 43.26 −67.11
0.04 47.52 0.974 44.91 −63.51
0.16 62.38 0.994 59.78 −19.67
0.64 79.99 0.996 77.39 33.88
2.56 107.69 0.940 105.09 116.44

In the presence of NMSc, activation energy increased, resulting in a decrease in corrosion rates. It has been proposed that organic inhibitor adsorption can influence corrosion rate by either decreasing the available reaction area (geometric blocking effect) or changing the activation energy of anodic or cathodic reactions occurring in the inhibitor-free surface during the inhibited corrosion process (Fouda et al. 2010).

The increase in the Ea value in the inhibited acid medium indicates the possibility of NMSc physisorption on the composite surface (Obot et al. 2009). The negative ∆Sa value suggests that the activated-complex state of the rate-determining step involves association rather than dissociation, resulting in less disorder in the reaction system (Gomma and Wahdan 1995).

3.3 Adsorption and thermodynamic behaviour

Knowing the adsorption mode of NMSc on the composite surface helps to understand the inhibition mechanism for Al alloy-SiC composite corrosion. The surface coverage (θ) values at various NMSc concentrations were computed from Equation (8).

(8)Ө=IE100

The θ values were applied to Langmuir, Temkin and Freundlich adsorption isotherms to confirm the fitment. The finest fitment of the data was obtained with the Langmuir adsorption model. Based on the Langmuir adsorption isotherm, θ is related to the inhibitor concentration (Cinh) (Obot et al. 2009),

(9)Cinhθ=1Kads+Cinh

where Kads indicates the adsorption equilibrium constant.

Figure 5a represents the plot of Langmuir adsorption isotherm for NMSc on Alalloy-SiC composite in 0.5 M HCl at different temperatures. The Cinh/θ versus Cinh plot gives a straight line with an intercept at 1/Kads and regression coefficients (R2) close to one, suggesting that NMSc follows the adsorption isotherm of Langmuir (Jeeva et al. 2017). It reveals the potential adsorption of NMSc molecules, creating a protective barrier on the composite surface. Besides, the Langmuir isotherm shows a protective monolayer film formation on the composite surface with a fixed number of equivalent adsorption sites. Each site holds an adsorbate with no interaction between the inhibitor molecules. The higher value of Kads (Table 4) indicates the stronger adsorption of MBS, particularly at a lower temperature (Tang et al. 2006).

Figure 5: 
						Al alloy-SiC-NMSc adsorption system’s (a) Langmuir  adsorption isotherm plot and (b) ΔG°ads vs. T plot.
Figure 5:

Al alloy-SiC-NMSc adsorption system’s (a) Langmuir  adsorption isotherm plot and (b) ΔG°ads vs. T plot.

Table 4:

Thermodynamic parameters for Al alloy-SiC/NMSc adsorption system.

T (K) K ads (mM−1) R 2 ΔG°ads (kJ mol−1) ΔHads (kJ mol−1) ΔS ads (kJ mol−1 K1)
303 21,236.07 0.999 −37.72
313 13,132.48 0.999 −35.90 −92.82 −0.1818
323 5481.40 0.996 −34.08

The standard free energy of adsorption (ΔG°ads) was obtained from Equation (10) (Umoren et al. 2010).

(10)Kads=155.5e(G°adsRT)

where 55.5 mol/L is the water concentration.

For the adsorption process, ΔG°ads values are linked to ΔH°ads (standard enthalpy) and ΔS°ads (standard entropy) as per Equation (11).

(11)Gads°=Hads°TSads°

Figure 5b depicts a straight line obtained from the ΔG°ads versus T plot. The slope provided the ΔS°ads value for the adsorption process, and the intercept provided the ΔH°ads value. Table 4 displays the thermodynamic outcomes.

Generally, ΔG°ads ≤ −20 kJ mol−1 indicate the inhibitor’s electrostatic interactions with the metal surface leading to physisorption. When ΔG°ads ≥ −40 kJ mol−1, the charge transfer from the inhibitor to the metal surface occurs, leading to chemisorption (Li et al. 2008). In this work, the values of ΔG°ads (Table 4) lie between 34 and 38 kJ mol−1, implying physico-chemical adsorption of NMSc. The ΔG°ads (Table 4) are negative, indicating that NMSc adsorption is spontaneous. Quraishi et al. 2000 reported that ΔH°ads value is less than −41.86 kJ mol−1 implying electrostatic interaction and that approaching −100 kJ mol−1 means electron-pair interaction of the inhibitor. The ΔH°ads value for NMSc is −92.82 kJ mol−1, which is closer to −100 kJ mol−1, suggesting its chemisorption. The negative sign of ΔS°ads value indicates the decrease in disordering occurs during the inhibition process (Bentiss et al. 2001).

3.4 Inhibition mechanism for Al alloy-SiC composite corrosion

NMSc’s inhibition performance on Al alloy-SiC composite corrosion primarily depends on the extent of its adsorption. The strength of inhibitor adsorption depends on its structure, electrochemical potential and temperature. The chemical structure of the inhibitor will play a major, including the number and charge density of the active adsorption centres, the size of the molecules and the type of adsorption. As per one concept, the organic inhibitor molecules could adsorb at the metal/medium interface by replacing the previously adsorbed water molecules, as indicated below (Solmaz 2014)

(12)

In this equation, Inh (aq) and Inh (ads) referred to inhibitors in an aqueous medium and adsorbed on the metal, respectively, and ‘n’ denotes the number of replaced water molecules. As a result, the adsorption of NMSc molecules creates a protective film, inhibiting Al alloy-SiC composite corrosion. The thermodynamic results showed that NMSc adsorption occurs through the physico-chemical adsorption process.

The aluminium metal undergoes dissolution reactions in the hydrochloric acid medium as follows (Bereket and Pinarbasi 2004):

In the anodic region:

(13)
(14)

In the cathodic region:

(15)
(16)

The Al alloy-SiC composite surface is positively charged in an acidic medium because the pHZCh value (pH at zero charge potential) equals 9.1 for Al (Bereket and Pinarbasi 2004). As a result of the attraction of electrostatic forces, the Cl ions in the acid medium can adsorb at the composite/acid solution interface. The amino group in NMSc is typically protonated (NMSc-H+) in a strongly acidic medium, such as the one used in this study. Furthermore, NMSc-H+ can be attracted towards the negatively charged chloride ions that are primarily adsorbed on the composite electrode surface. As a result, the physical adsorption of protonated NMSc can occur on the composite surface. The chemisorption of NMSc may also happen by the donor-acceptor kind interactions of N and O atoms, or π bonds of the aromatic ring in NMSc with empty d-orbitals of Al, resulting in coordination bond formation (Mansfeld 1987). The protective layer formed by NMSc can protect the composite surface by isolating it from direct contact with the corrosive acid medium, thereby controlling the composite’s deterioration.

3.5 Surface analysis of Al alloy-SiC specimen

SEM images of the Al alloy-SiC specimen immersed in acid and inhibited acid solutions are presented in Figure 6a and b. Many micro pits/holes were observed on the corroded specimen surface immersed in 0.5 M HCl solution (Figure 6a). However, a uniform surface with very few micro pits/holes were observed on the specimen surface, immersed in 0.5 M HCl containing 2.56 mM NMSc (Figure 6b). It could be the result of NMSc’s adsorption on the composite surface.

Figure 6: 
						SEM pictures of Al alloy-SiC specimen immersed in (a) acid and (b) inhibited acid solution.
Figure 6:

SEM pictures of Al alloy-SiC specimen immersed in (a) acid and (b) inhibited acid solution.

The AFM images of the Al alloy-SiC composite dipped in acid medium and inhibited acid medium with 2.56 mM NMSc are shown in Figure 7a and b. The Ra (average surface roughness) and Rq (root mean square roughness) values for the corroded composite specimens are 709 nm and 914 nm, respectively. Similarly, these values for inhibited composite samples are 248 nm and 176 nm, respectively. These results reveal that inhibited surface is more uniform/smoother than the fully corroded surface, suggesting the NMSc adsorption, which leads to a protective film formation. Thus, the composite material’s deterioration was kept under control.

Figure 7: 
						AFM images of Al alloy-SiC sample dipped in (a) acid and (b) inhibited acid solution.
Figure 7:

AFM images of Al alloy-SiC sample dipped in (a) acid and (b) inhibited acid solution.

2.6 UV-visible spectral analysis on NMSc adsorption

UV-visible spectrum is one of the valuable tools to support NMSc adsorption onto Al alloy-SiC composite. The UV-visible spectra of the inhibited acid solution containing 0.01 mM NMSc were recorded (Figure 8). The spectra showed two absorption peaks at 226 nm and 256 nm associated with ππ* and nπ* transitions. Then, the composite specimen was immersed in the above solution for 3 h, and spectra of the left solution were recorded. Figure 8 shows that the second spectra showed the same two peaks with a reduced intensity compared to the first spectra. The decrease in intensity of the second spectra reveals the adsorption of NMSc at the surface of the composite specimen.

Figure 8: 
						Absorption spectra of the inhibited acid solution containing 0.01 mM NMSc before and after the immersion of composite coupon.
Figure 8:

Absorption spectra of the inhibited acid solution containing 0.01 mM NMSc before and after the immersion of composite coupon.

3.7 Quantum chemical calculations

The theoretical studies using density functional theory (DFT) were carried out to find the influence of the inhibitor structure on its inhibition activity. Electron-donating or accepting properties of organic inhibitor molecules at the metal surface mainly control its inhibition activity. The amino group in NMSc can undergo protonation in an acid medium like the one (0.5 M HCl) used in this investigation. As a result, the participation of neutral and protonated NMSc in the inhibition of Al alloy-SiC composite corrosion in a 0.5 M HCl medium could be influenced. Therefore, the DFT studies were performed on neutral and protonated NMSc using the 631G** basis set. The parameters calculated were EHOMO (energy of highest occupied molecular orbital), ELUMO (energy of lowest unoccupied molecular orbital), ∆Eg (energy gap), η (hardness), σ (softness), ΔN (the fraction of electron transferred) and ω (electrophilicity index) (Chang-Guo et al. 2003).

The Mulliken charge indicates the active sites for adsorption in the inhibitor molecule and the possible interactions with the metal surface. The heteroatoms with high negative Mulliken charges on atoms in neutral and protonated NMSc (Figure 9a and b) are the preferable active sites for the donor–acceptor interactions. In the neutral NMSc, atoms such as N3 (−0.953), O4 (−0.65), N7 (−0.523), N2 (−0.478) and O11 (−0.466), while in protonated NMSc atoms like N3 (−0.628), N7 (−0.477), O11 (−0.471) and N2 (−0.378) showed higher negative charges (Figure 9a and b). These high negative charge atoms in neutral and protonated NMSc can act as active adsorption centres.

Figure 9: 
						Mulliken charges on (a) neutral and (b) protonated NMSc.
Figure 9:

Mulliken charges on (a) neutral and (b) protonated NMSc.

The optimised structures of neutral and protonated MBS and the corresponding HOMO and LUMO are depicted in Figure 10.

Figure 10: 
						Neutral NMSc’s (a) optimized structure, (b) HOMO, (c) LUMO, and protonated NMSc’s (d) optimized structure, (e) HOMO, (f) LUMO.
Figure 10:

Neutral NMSc’s (a) optimized structure, (b) HOMO, (c) LUMO, and protonated NMSc’s (d) optimized structure, (e) HOMO, (f) LUMO.

E HOMO value indicates an inhibitor molecule’s probability of donating electrons, whereas ELUMO values express the ability to accept electrons. The neutral NMSc showed a higher EHOMO value (−6.2662 eV) than the protonated NMSc (−8.9071 eV). Hence, the neutral NMSc can readily donate electrons to the Al atom’s empty d-orbital, leading to a coordinate-type bond. The ELUMO value of protonated NMSc is lower (−4.4787 eV) than its neutral form (−0.3379 eV). It suggests that the protonated NMSc can easily accept electrons from the Al atom leading to a back-donating type bond (Obot and Obi-Egbedi 2008).

Generally, the lower the energy bandgap (ΔEg) between the HOMO and LUMO higher is chemical reactivity shown by the inhibitor and hence exhibits good inhibition activity (Arslan et al. 2009). In this work, the observed ΔEg value for protonated NMSc (4.4284 eV) is lower than its neutral form (5.9283 eV). Therefore, the protonated form can show higher reactivity compared to neutral NMSc. Thus, protonated NMSc contributes more towards the inhibition performance. These results support the physisorption of NMSc on the composite, as explained earlier under the experimental results. The protonated NMSc showed a lower η value (2–2142 eV) and higher σ value (0.4516 eV) compared to its neutral form (η = 2.9641 eV and σ = 0.3374 eV). It reveals the stronger chemical reactivity of protonated NMSc, contributing more towards the inhibition performance (Masoud et al. 2010). It again confirms the possible physisorption of protonated NMSc on the composite surface.

The fraction of electrons transferred (ΔN) from the inhibitor to the composite electrode surface was computed by Equation (17) (Pearson 1990).

(17)ΔN=χAlχInh2(ηAl+ηInh)=ɸAlχInh2ηInh

where ηAl is the hardness of Al and ηInh is the hardness of the inhibitor, while χAl and χInh are the electronegativity values of Al and inhibitor. The theoretical electronegativity value of Al (ɸAl) is taken as 4.28 eV (Kokalj and Kovacevic 2011), and the hardness of Al (ηAl) as zero (Pearson 1990). The positive value of ΔN indicates the higher electron-donating nature, whereas the negative value implies the stronger electron-accepting trend of the inhibitor involved (Lukovits et al. 2001). Therefore neutral NMSc (ΔN = 0.1650) readily donates electrons, whereas protonated NMSc (ΔN = −0.5449) readily accepts electrons from the metal surface.

The electrophilicity index (ω) value for the inhibitor is given by Equation (18) (Parr et al. 1999).

(18)ω=µ22η

The electronic chemical potential (µ) is given by Equation (19).

(19)µ=I+A2=χ

An inhibitor with a low value of ω acts as a good nucleophile, whereas a high value of ω shows a good electrophile nature (Obot and Obi-Egbedi 2010). The protonated form of NMSc showed a higher ω value (10.1041 eV), indicating its higher tendency to accept electrons. Its neutral form exhibited a lower ω value (1.8392 eV), showing a higher electron-donating ability. It results in stronger adsorption of both neutral and protonated NMSc on the composite, displaying a good inhibition performance.

3.8 Comparison of reported Al alloy-SiC composite inhibitors with NMSc

The corrosion inhibition performance of NMSc is compared to that of other reported inhibitors against Al alloy-SiC composite corrosion (Table 5). Among the reported inhibitors of Al alloy-SiC composite corrosion in a hydrochloric acid medium, NMSc evinced the highest IE of 95 % at 2.56 mM and 303 K. Most of the inhibitors showed maximum IE at 303 K, whereas others at 323 K. Some inhibitors require organic solvent (Ethanol/Acetone) to prepare their solution (Kini et al. 2010, 2011; Shetty et al. 2020), and others are soluble in the aqueous hydrochloric acid media (Charitha and Rao 2017, 2018a, 2018b, 2020). However, some of these inhibitors are effective only at lower media concentrations (0.025/0.05 M).

Table 5:

Comparison of NMSc with other reported inhibitors of AA-SiC composite.

Inhibitors (optimum concentration) Medium (optimum temperature) IE (%) Remarks References
WL PDP EIS
Ethyl-2-phenyl hydrozono-3-oxobutyrate (150 ppm) 0.5 M HCl (303 K) 81.2 81.9 IE decreases with rise in temperature Kini et al. (2010)
3-Chloro-1-benzothiophene-2-cabohydrazide (6.6 × 10−4 M) 0.5 M HCl (303 K) 78.7 86.0 IE decreases with rise in temperature Kini et al. (2011)
Starch (0.8 g  L─1) 0.025 M HCl (323 K) 83.44 83.44 IE increases with rise in temperature Charitha and Rao (2016)
Dextran (0.4 g L−1) 1 M HCl (303 K) 91.3 90.24 IE decreases with rise in temperature Charitha and Rao (2017)
Insulin (1 g L−1) 0.05 M HCl (303 K) 88.8 89.2 IE decreases with rise in temperature Charitha and Rao (2018a)
Pullulan (1 g L−1) 0.025 M HCl (303 K) 89.6 88.5 IE decreases with rise in temperature Charitha and Rao (2018b)
Pectin (1 g L−1) 0.025 M HCl (303 K) 74.5 75.4 IE increases with the rise in temperature Charitha and Rao (2020)
4-Hydroxy-N′-[3-phenylprop-2-en-1-ylidene] benzo hydrazide (1 × 10−3 M) 0.5 M HCl (303 K) 85.7 83.1 IE decreases with a rise in temperature Shetty et al. (2020)
NMSc (2.56 mM) 0.5 M HCl (303 K) 95.42 90.61 IE decreases with a rise in temperature Present work

4 Conclusions

  1. MBS is explored as an effective corrosion inhibitor of Al alloy-SiC composite in an acid medium.

  2. The inhibition efficiency of NMSc increased by increasing its concentration and decreasing the medium temperature.

  3. NMSc showed a maximum efficiency of 95.42 % at 2.56 mM and 303 K.

  4. NMSc showed mixed inhibitor behaviour by influencing the anodic as well as cathodic reactions.

  5. The inhibition process was caused by mixed-type adsorption of NMSc molecules, following Langmuir’s adsorption model.

  6. The adsorption of NMSc on the composite was confirmed by surface morphological studies conducted using SEM and AFM.

  7. The experimental results for MBS obtained by the PDP technique agree with those obtained by the EIS technique.

  8. The theoretical studies using DFT supported the experimental results.


Corresponding author: Prakasha Shetty, Department of Chemistry, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal576104, Karnataka, India, E-mail:

Acknowledgements

Maithili Krishnananda is grateful to the Manipal Institute of Technology, MAHE, Manipal, Karnataka, India, for the lab facilities.

  1. Author contribution: Maithili Krishnananda: experimental work, collection of data and initial draft of the manuscript; Prakasha Shetty: supervision and review of the manuscript; Preethi Kumari P: methodology; Sneha Kagatikar: performance of theoretical studies.

  2. Research funding: This research received no specific grant from the public, commercial or not-for-profit funding agencies.

  3. Conflicts of interest: The authors declare that they have no conflicts of interest regarding this article.

References

Abd El-Rehim, S.S., Ibrahim, M.A.M., and Khaled, K.F. (1999). 4-Aminoantipyrine as an inhibitor of mild steel corrosion in HCl solution. J. Appl. Electrochem. 29: 593–599.10.1023/A:1003450818083Suche in Google Scholar

Abd El Rehim, S.S., Hassan, H.H., and Amin, M.A. (2002). The corrosion inhibition study of sodium dodecyl benzene sulphonate to aluminium and its alloys in 1.0 M HCl solution. Mater. Chem. Phys. 78: 337–348, https://doi.org/10.1016/s0254-0584(01)00602-2.Suche in Google Scholar

Ahamad, I., Prasad, R., and Quraishi, M.A. (2010). Adsorption and inhibitive properties of some new Mannich base of isatin derivatives on corrosion of mild steel in acidic media. Corros. Sci. 52: 1472–1481, https://doi.org/10.1016/j.corsci.2010.01.015.Suche in Google Scholar

Arslan, T., Kandemirli, F., Ebenso, E.E., Love, I., and Alemu, H. (2009). Quantum chemical studies on the corrosion inhibition of some sulphonamides on mild steel in acidic medium. Corros. Sci. 51: 35–47, https://doi.org/10.1016/j.corsci.2008.10.016.Suche in Google Scholar

Babic, R., Metikos, M., and Grubac, Z. (1998). Corrosion protection of aluminium in acidic chloride solutions with nontoxic inhibitors. J. Appl. Electrochem. 28: 433–439.Suche in Google Scholar

Bentiss, F., Traisnel, M., and Lagrene, M. (2001). Influence of 2,5-bis(4-dimethylaminophenyl)-1,3,4-thiadiazole on corrosion inhibition of mild steel in acidic media. J. Appl. Electrochem. 33: 41–48.Suche in Google Scholar

Bereket, G. and Pinarbasi, A. (2004). Electrochemical thermodynamic and kinetic studies of the behaviour of aluminium in hydrochloric acid containing various benzotriazole derivatives. Corros. Eng. Sci. Technol. 39: 308–312, https://doi.org/10.1179/174327804x13136.Suche in Google Scholar

Bessone, J.B., Salinas, D.R., Mayer, C., Ebert, M., and Lorenz, W.J. (1992). An EIS study of aluminium barrier-type oxide films formed in different media. Electrochim. Acta 37: 2283–2290, https://doi.org/10.1016/0013-4686(92)85124-4.Suche in Google Scholar

Bhat, M.S.N., Surappa, M.K., and Nayak, H.V.S. (1991). Corrosion behaviour of silicon carbide particle reinforced 6061/Al alloy composites. J. Mater. Sci. 26: 4991–4006, https://doi.org/10.1007/bf00549882.Suche in Google Scholar

Bobic, B., Mitrovic, S., Babic, M., and Bobic, I. (2010). Corrosion of metal-matrix composites with aluminium alloy substrate. New York: McGraw-Hill.Suche in Google Scholar

Bockris, J.O.’M. and Minevski, L.V. (1993). On the mechanism of the passivity of aluminium and aluminium alloys. J. Electroanal. Chem. 349: 375–414, https://doi.org/10.1016/0022-0728(93)80186-l.Suche in Google Scholar

Brett, C.M.A. (1992). On the electrochemical behavior of aluminium in acidic chloride solution. Corros. Sci. 33: 203–210, https://doi.org/10.1016/0010-938x(92)90145-s.Suche in Google Scholar

Chang-Guo, Z., Nichols, J.A., and Dixon, D.A. (2003). Ionization potential, electron affinity, electronegativity, hardness, and electron excitation energy: molecular properties from density functional theory orbital energies. J. Phys. Chem. A107: 4184–4195.10.1021/jp0225774Suche in Google Scholar

Charitha, B.P. and Rao, P. (2016). Protection of 6061 Al-15%(v) SiC(P) composite from corrosion by a biopolymer and surface morphology studies. Protect. Met. Phys. Chem. Surface 52: 704–713, https://doi.org/10.1134/s2070205116040079.Suche in Google Scholar

Charitha, B.P. and Rao, P. (2017). An eco-friendly approach for corrosion control of 6061 Al-15%(v) SiC(P) composite and its base alloy. Chin. J. Chem. Eng. 25: 363–372, https://doi.org/10.1016/j.cjche.2016.08.007.Suche in Google Scholar

Charitha, B.P. and Rao, P. (2018a). Pullulan as a potent green inhibitor for corrosion mitigation of aluminium composite: electrochemical and surface studies. Int. J. Biol. Macromol. 112: 461–472, https://doi.org/10.1016/j.ijbiomac.2018.01.218.Suche in Google Scholar PubMed

Charitha, B.P. and Rao, P. (2018b). Environmentally benign green inhibitor to attenuate acid corrosion of 6061 aluminium-15 %(v) SiC(P) composite. J. Ind. Eng. Chem. 58: 357–368.10.1016/j.jiec.2017.09.049Suche in Google Scholar

Charitha, B.P. and Rao, P. (2020). Pectin as a potential green inhibitor for corrosion control of 6061Al–15%(V) SiC(P) composite in acid medium: electrochemical and surface studies. J. Fail. Anal. Prev. 20: 1684–1696, https://doi.org/10.1007/s11668-020-00973-z.Suche in Google Scholar

Davis, J.R. (1993). ASM specialty handbook: aluminium and aluminium alloys. Metals Park, Ohio, USA: ASM International.Suche in Google Scholar

de Wit, J.H.W. and Lenderink, H.J.W. (1996). Electrochemical impedance spectroscopy as a tool to obtain mechanistic information on the passive behaviour of aluminium. Electrochim. Acta 41: 1111–1119, https://doi.org/10.1016/0013-4686(95)00462-9.Suche in Google Scholar

Fontana, M.G. (1987). Corrosion engineering. New York: McGraw Hill, 1987.Suche in Google Scholar

Fouda, A.S., El Ewady, G.Y., Mostafa, H.A., and El-Toukhee, Y.M. (2010). Tertiary ketonic Mannich bases as corrosion inhibitors for aluminium dissolution in acidic solutions. Chem. Eng. Commun. 197: 366–376, https://doi.org/10.1080/00986440903089015.Suche in Google Scholar

Gomma, M.K. and Wahdan, M.H. (1995). Schiff bases as corrosion inhibitors for aluminium in hydrochloric acid solution. Mater. Chem. Phys. 39: 209–213, https://doi.org/10.1016/0254-0584(94)01436-k.Suche in Google Scholar

Jeeva, M., Prabhu, G.V., and Rajesh, C.M. (2017). Inhibition effect of nicotinamide and its Mannich base derivatives on mild steel corrosion in HCl. J. Mater. Sci. 52: 12861–12888, https://doi.org/10.1007/s10853-017-1401-2.Suche in Google Scholar

Jüttner, K. (1990). Electrochemical impedance spectroscopy (EIS) of corrosion processes on inhomogeneous surfaces. Electrochim. Acta 35: 1501–1508, https://doi.org/10.1016/0013-4686(90)80004-8.Suche in Google Scholar

Kedam, M., Mattos, O.R., and Takenouti, H. (1981). Reaction model for iron dissolution studied by electrode impedance. I. Experimental results and reaction model. J. Electrochem. Soc. 128: 257, https://doi.org/10.1149/1.2127401.Suche in Google Scholar

Kini, U.A., Shetty, P., Shetty, S.D., Isloor, A.M., and Herle, R. (2010). The inhibition action of ethyl-2-phenyl hydrozono-3-oxobutyrate on the corrosion of 6061 Al alloy/SiCp composite in hydrochloric acid medium. J. Chil. Chem. Soc. 55: 56–60, https://doi.org/10.4067/s0717-97072010000100014.Suche in Google Scholar

Kini, U.A., Shetty, P., Shetty, S.D., and Isloor, A.M. (2011). Corrosion inhibition of 6061 aluminium alloy/SiCp composite in hydrochloric acid medium using 3-chloro-1-benzothiophene-2-carbohydrazide. Indian J. Chem. Technol. 18: 439–445.10.4067/S0717-97072010000100014Suche in Google Scholar

Kokalj, A. and Kovacevic, N. (2011). On the consistent use of electrophilicity index and HSAB-based electron transfer and its associated change of energy parameters. Chem. Phys. Lett. 507: 81–184, https://doi.org/10.1016/j.cplett.2011.03.045.Suche in Google Scholar

Kumari, P.P., Shetty, P., Nagalaxmi, and Dhanya, S. (2020). Effect of cysteine as environmentally friendly inhibitor on AA6061-T6 corrosion in 0.5 M HCl: electrochemical and surface studies. Surf. Eng. Appl. Electrochem. 56: 624–634, https://doi.org/10.3103/s1068375520050087.Suche in Google Scholar

Lenderink, H.J.W., Linden, M.V.D., and De Wit, J.H.W. (1993). Corrosion of aluminium in acidic and neutral solutions. Electrochim. Acta 38: 1989–1992, https://doi.org/10.1016/0013-4686(93)80329-x.Suche in Google Scholar

Li, W., He, Q., Zhang, S., Pel, C., and Hou, B. (2008). Some new triazole derivatives as inhibitors for mild steel corrosion in acid medium. J. Appl. Electrochem. 38: 289–295, https://doi.org/10.1007/s10800-007-9437-7.Suche in Google Scholar

Lukovits, L., Kalman, E., and Zucchi, F. (2001). Corrosion inhibitors: correlation between electronic structure and efficiency. Corrosion 57: 3–8, https://doi.org/10.5006/1.3290328.Suche in Google Scholar

Mansfeld, F. (1987). Corrosion mechanisms. New York, USA: Marcel Dekker, Inc.10.5006/1.3290328Suche in Google Scholar

Masoud, M.S., Awad, M.K., Shaker, M.A., and El-Tahawy, M.M.T. (2010). The role of structural chemistry in the inhibitive performance of some amino pyrimidines on the corrosion of steel. Corros. Sci. 52: 2387–2396, https://doi.org/10.1016/j.corsci.2010.04.011.Suche in Google Scholar

Metikos-Hukovic, M. and Babic, R. (1998). Corrosion protection of aluminium in acidic chloride solutions with nontoxic inhibitors. J. Appl. Electrochem. 28: 433–439.10.1016/j.corsci.2010.04.011Suche in Google Scholar

Metikos-Hukovic, M., Babic, R., and Grubac, Z. (2002). The study of aluminium corrosion in acidic solution with nontoxic inhibitors. J. Appl. Electrochem. 32: 35–41.10.1023/A:1003200808093Suche in Google Scholar

Nagalaxmi, Shetty, P., and Kumari, P.P. (2020). Inhibitive action of glutathione reduced on the deterioration of AA6061 in 0.5M HCl. Tribol. Ind. 42: 214–224, https://doi.org/10.24874/ti.785.10.19.02.Suche in Google Scholar

Noor, E.A. (2009). Evaluation of inhibitive action of some quaternary N-heterocyclic compounds on the corrosion of Al–Cu alloy in hydrochloric acid. Mater. Chem. Phys. 114: 533–541, https://doi.org/10.1016/j.matchemphys.2008.09.065.Suche in Google Scholar

Obot, I.B. and Obi-Egbedi, N.O. (2008). Fluconazole as an inhibitor for aluminium corrosion in 0.1 M HCl. Colloids Surf. A. 330: 207–212, https://doi.org/10.1016/j.colsurfa.2008.07.058.Suche in Google Scholar

Obot, I.B. and Obi-Egbedi, N.O. (2010). Adsorption properties and inhibition of mild steel corrosion in sulphuric acid solution by ketoconazole: experimental and theoretical investigation. Corros. Sci. 52: 198–204, https://doi.org/10.1016/j.corsci.2009.09.002.Suche in Google Scholar

Obot, I.B., Obi-Egbedi, N.O., and Umoren, S.A. (2009). The synergistic inhibitive effect and some quantum chemical parameters of 2,3-diaminonaphthalene and iodide ions on the hydrochloric acid corrosion of aluminium. Corros. Sci. 51: 276–282, https://doi.org/10.1016/j.corsci.2008.11.013.Suche in Google Scholar

Pardo, A., Merino, M.C., Merino, S., Viejo, F., Carboneras, M., and Arrabal, R. (2005). Influence of reinforcement proportion and matrix composition on pitting corrosion behaviour of cast aluminium matrix composites (A3xx.X/SiCp). Corros. Sci. 47: 1750–1764, https://doi.org/10.1016/j.corsci.2004.08.010.Suche in Google Scholar

Parr, R.G., Szentpaly, L.V., and Liu, S. (1999). Electrophilicity index. J. Am. Chem. Soc. 121: 1922–1924, https://doi.org/10.1021/ja983494x.Suche in Google Scholar

Pearson, R.G. (1990). Hard and soft acids and bases-the evolution of a chemical concept. Coord. Chem. Rev. 100: 403–425, https://doi.org/10.1016/0010-8545(90)85016-l.Suche in Google Scholar

Pinto, G.M., Nayak, J., and Shetty, A.N. (2009). Corrosion behaviour of 6061 Al - 15vol. Pct. SiC composite and its base alloy in a mixture of 1:1 hydrochloric and sulphuric acid medium. Int. J. Electrochem. Sci. 4: 1452–1468.10.1016/0010-8545(90)85016-LSuche in Google Scholar

Pinto, G.M., Nayak, J., and Shetty, A.N. (2011). Corrosion inhibition of 6061 Al–15 vol. pct. SiC(p) composite and its base alloy in a mixture of sulphuric acid and hydrochloric acid by 4-(N,N-dimethyl amino) benzaldehyde thiosemicarbazone. Mater. Chem. Phys. 125: 628–640, https://doi.org/10.1016/j.matchemphys.2010.10.006.Suche in Google Scholar

Quraishi, M.A., Rawat, J., and Ajmal, M. (2000). Dithiobiurets: a novel class of acid corrosion inhibitors for mild steel. J. Appl. Electrochem. 30: 745–751, https://doi.org/10.1023/a:1004099412974.10.1016/j.matchemphys.2010.10.006Suche in Google Scholar

Quraishi, M.A., Ahamad, I., Singh, A.K., Shukla, S.K., Lal, B., and Singh, V. (2008). N- (piperidinomethyl)-3-[(pyridylidene)amino]isatin: new and effective acid corrosion inhibitor for mild steel. Mater. Chem. Phys. 112: 1035–1039, https://doi.org/10.1016/j.matchemphys.2008.07.011.Suche in Google Scholar

Reddy, A.C. and Zitoun, E. (2010). Matrix Al-alloys for silicon carbide particle reinforced metal matrix composites. Indian J. Sci. Technol. 3: 1184–1187, https://doi.org/10.17485/ijst/2010/v3i12.8.Suche in Google Scholar

Rohatgi, P.K. (1993). Metal matrix composite. Defence Sci. J. 43: 323–349, https://doi.org/10.14429/dsj.43.4336.Suche in Google Scholar

Schorr, M. and Yahalom, J. (1972). The significance of the energy of activation for the dissolution reaction of metal in acids. Corros. Sci. 12: 867–868, https://doi.org/10.1016/s0010-938x(72)80015-5.Suche in Google Scholar

Sheasby, P.G. and Pinner, R. (2001). The surface treatment and finishing of aluminium and its alloys, 6th ed. USA: ASM International and Finishing Publications Ltd.10.1016/S0010-938X(72)80015-5Suche in Google Scholar

Singh, A.K. and Quraishi, M.A. (2010). Inhibiting effects of 5-substituted isatin-based Mannich bases on the corrosion of mild steel in hydrochloric acid solution. J. Appl. Electrochem. 40: 1293–1306, https://doi.org/10.1007/s10800-010-0079-9.Suche in Google Scholar

Singla, D.M., Dwivedi, D., Singh, L., and Chawla, K. (2009). Development of aluminium based silicon carbide particulate metal matrix composite. J. Miner. Mater. Char. Eng. 8: 455–467, https://doi.org/10.4236/jmmce.2009.86040.Suche in Google Scholar

Shetty, D., Kumari, P.P., Rao, S.A., and Shetty, P. (2020). Anticorrosion behaviour of a hydrazide derivative on 6061 Al-15%(v) SiC(P) composite in acid medium: experimental and theoretical calculations. J. Bio- Tribo-Corros. 59: 1–15, https://doi.org/10.1007/s40735-020-00356-9.Suche in Google Scholar

Shetty, P. (2020). Schiff bases: an overview of their corrosion inhibition activity in acid media against mild steel. Chem. Eng. Commun. 207: 985–1029, https://doi.org/10.1080/00986445.2019.1630387.Suche in Google Scholar

Shetty, P. (2022). Corrosion inhibitors for AA6061 and AA6061-SiC composite in aggressive media - a review. Corros. Rev. 40: 543–560, https://doi.org/10.1515/corrrev-2021-0084.Suche in Google Scholar

Solmaz, R. (2014). Investigation of adsorption and corrosion inhibition of mild steel in hydrochloric acid solution by 5-(4-dimethylaminobenzylidene) rhodanine. Corros. Sci. 79: 169–176, https://doi.org/10.1016/j.corsci.2013.11.001.Suche in Google Scholar

Tang, L., Li, X., Si, Y., Mu, G., and Liu, G. (2006). The synergistic inhibition between 8-hydroxy quinoline and chloride ion for the corrosion of cold rolled steel in 0.5 M sulfuric acid. Mater. Chem. Phys. 95: 29–38, https://doi.org/10.1016/j.matchemphys.2005.03.064.Suche in Google Scholar

Trowsdale, A.J., Noble, B., Haris, S.J., Gibbins, I.S.R., Thomson, G.E., and Wood, G.C. (1996). The influence of silicon carbide reinforcement on the pitting behavior of aluminium. Corros. Sci. 38: 177–191, https://doi.org/10.1016/0010-938x(96)00098-4.Suche in Google Scholar

Umoren, S.A., Li, Y., and Wang, F.H. (2010). Effect of polyacrylic acid on the corrosion behaviour of aluminium in sulphuric acid solution. J. Solid State Electrochem. 14: 2293–2305, https://doi.org/10.1007/s10008-010-1064-2.Suche in Google Scholar

Viswanathan, M. (2016). Synthesis, spectral, X-ray diffraction and antibacterial studies of oxovanadium (IV) and dioxouranium(VI) complexes of N-(1-morpholino benzyl) semicarbazide. Asian J. Chem. 28: 1621–1623, https://doi.org/10.14233/ajchem.2016.19786.Suche in Google Scholar

Yurt, A., Ulutas, S., and Dal, H. (2006). Electrochemical and theoretical investigation on the corrosion of aluminium in acidic solution containing some Schiff bases. Appl. Surf. Sci. 253: 919–925, https://doi.org/10.1016/j.apsusc.2006.01.026.Suche in Google Scholar

Received: 2022-09-01
Accepted: 2023-04-11
Published Online: 2023-05-19
Published in Print: 2023-10-26

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 25.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2022-0078/html
Button zum nach oben scrollen