Anticorrosion Studies
rMrCP20 was first expressed and purified using the same protocol as previously described.18 To examine the corrosion-inhibition properties of rMrCP20 protein, epoxy-embedded polished AH36 steel coupons were immersed in a pH 8.3 buffer containing 150 mM NaCl and 20 mM Tris(hydroxymethyl)aminomethane (Tris), in the absence or presence of rMrCP20 at different concentrations (see Materials and Methods). Figures 1a,b illustrate representative pictures of AH36 coupons at various time points and their respective immersion solutions after 24 h, at a range of rMrCP20 concentrations from 0 to 10 mg/mL. In the absence of rMrCP20, the coupons underwent severe corrosion, with surfaces covered with rust and other corrosion products, and the respective buffer solution turning into intense brown color. At a low concentration of ca. 0.1 mg/mL, rMrCP20 was previously found to accelerate the corrosion of steel coupon surfaces.19 At increasing concentration of rMrCP20, corrosion was reduced considerably and little to no corrosion was observed on coupons incubated in 5 mg/mL and 10 mg/mL of proteins (Fig. 1a), while their respective buffer solutions post-incubation had a light brown color. Randomized pitting corrosion events were observed across the coupon surfaces for corroded samples, with oxide products adhering to surfaces, as shown by representative scanning electron microscopy (SEM) images in Fig. 1c. Further analysis was performed using ImageJ software to estimate the corroded area and the results are shown in Fig. 1d. In the absence of rMrCP20, on average 91% of AH36 coupon surfaces was corroded after 24 h. However, at increasing protein concentration surface corrosion was reduced drastically, down to 2% at 10 mg/mL, thereby demonstrating a concentration-dependent corrosion inhibition effect of rMrCP20 on AH36 steel.
Weight loss measurements are routinely employed to evaluate the effectiveness of metal corrosion inhibitors.20 Fe content from the incubation solutions and respective coupon surfaces were measured by Inductively Coupled Plasma Optical Emission Spectrophotometer (ICP-OES) to reflect the total weight loss for each sample at concentrations of rMrCP20 from 0.1 mg/mL to 10 mg/mL, and to calculate the inhibition efficiency using Eqs. (1)–(2).21,22
\(CR=\frac{8.76 \times {10}^{4}\times \varDelta m}{\rho AT}\) (Eq. 1)
\(\eta \left(\%\right)=\theta \times 100=\left(1-\frac{{CR}_{1}}{{CR}_{0}}\right)\times 100\) (Eq. 2)
where CR is the corrosion rate (mm yr− 1), Δm is the total weight loss (g), ρ is density of the metal samples (7.86 g cm− 3), A is the surface area (cm2), and T is the exposure time (h). η is the corrosion inhibition efficiency (%), θ is degree of surface coverage, while CR0 and CR1 are the weight losses of the coupons in the absence and presence of rMrCP20, respectively. The corrosion rate calculated via Eq. (1) decreased significantly with increasing the concentration of rMrCP20 (Supplementary Table 1), with the lowest corrosion rate of 0.12 mm yr− 1 and a high corrosion-inhibition efficiency η 88.48% (Fig. 1e). The observed enhanced corrosion-inhibition efficiency is attributed to the strong adhesive properties of rMrCP20, whereby protein adsorption to the coupon surfaces –and hence the degree of coupon surface coverage– increased at higher protein concentrations.
Electrochemical Studies
To further investigate the impedance brought about by different concentrations of rMrCP20, electrochemical impedance spectroscopy (EIS) measurements were performed for AH36 samples after 24 h incubation in buffer solutions containing various concentration of rMrCP20 protein. A protein concentration of 5 mg/mL was selected as the optimal concentration for further characterization as it exhibited good corrosion inhibition performance, whereas increasing the amount of protein to 10 mg/mL did not significantly affect the results.
Figure 2a-c show the Nyquist plots (Fig. 2a), Bode modulus (Fig. 2b) and phase angle representations (Fig. 2c) from EIS measurements. The Nyquist plots show single semi-circle loops from high to mid-frequency range, indicating that the corrosion of AH36 was controlled by a charge transfer process (Fig. 2a).23–25 Tailing was observed at the lower frequency range of each Nyquist plot, and these imperfections can be attributed to the roughness and heterogeneity of the AH36 working electrode,24,26 distribution of active center,26,27 adsorption of inhibitor molecules28 and accumulation of corrosion products on the working electrode.27,28 The diameter of the Nyquist plot is directly correlated to the impedance of the system, hence providing a suitable comparison of the resistance of the Fe substrate exposed to different concentrations of rMrCP20. The diameters increased at higher concentration of rMrCP20, suggesting that rMrCP20 adsorbed onto AH36 surfaces increases the charge transfer resistance and hence imparts corrosion resistance properties.
Bode modulus plots (Fig. 2b) provide a more comprehensive representation of the impedance across the range of frequencies measured, while phase angle plots (Fig. 2c) indicate the phase shift across the range of frequencies and the plausible components in the respective equivalent electrical circuits (EEC). At higher protein concentrations, the Bode modulus plot showed an increase in impedance values at low frequencies, where corrosion events are usually observed, while the phase angle plots showed an increase in the maximum phase angle value. The higher impedance is attributed to the formation of more homogenous rMrCP20 protein films at higher protein concentrations. Furthermore, the diameter of the Nyquist semicircles also increased with exposure time at higher rMrCP20 concentration (Supplementary Fig. 1), which is caused by the formation of a mixed surface layer of corrosion products (iron oxide) and adsorbed proteins.28 The evolution of the surface film most likely involved complexation between rMrCP20 and Fe ions (Fe2+ and Fe3+), which blocked the charge transfer path. This observation also highlights the stability of the adsorbed rMrCP20 molecules on the steel surface.
To further interpret the EIS results, an equivalent electrical circuit (EEC) was chosen to extract the relevant electrochemical parameters (Fig. 2a inset). In the circuit, Rs represents the solution resistance, Rct the charge transfer resistance, CPEdl and CPEf the constant phase elements of the double layer and protein film respectively, and Rf the resistance of the adsorbed rMrCP20 film. The double layer capacitance (Cdl) values were calculated from the expression:23
\({C}_{dl}=\frac{{Y}_{0}{\omega }^{n-1}}{\text{s}\text{i}\text{n}\left[n\left(\frac{\pi }{2}\right)\right]}\) (Eq. 3)
where Y0 is the CPE constant, n is the CPE exponent, ω is the angular frequency (\(\omega =2\pi {f}_{\text{m}\text{a}\text{x}})\)in rad/s, and \({f}_{\text{m}\text{a}\text{x}}\) is the frequency at which the imaginary component of the impedance is highest. The corrosion inhibition efficiency (% η) was calculated using Eq. (4):29
\({\eta }_{EIS}\left(\%\right)=\frac{{R}_{ct}^{0}-{R}_{ct}^{{\prime }}}{{R}_{ct}^{0}}\times 100\) (Eq. 4)
where \({R}_{ct}^{0}\)and \({R}_{ct}^{{\prime }}\) are the charge transfer resistances in the presence and absence of rMrCP20, respectively. The values of the electrochemical parameters derived from the Nyquist plots are listed in Fig. 2d and demonstrate that adding rMrCP20 to the corrodent led to a reduction in Cdl and an increase in Rct and Rf, which became more apparent as the rMrCP20 concentration increased. As the concentration of rMrCP20 increased from 1 to 5 mg/mL, the Rct value increased from 926 Ω cm2 to 4721 Ω cm2, while the corresponding η value increased from 19.8–84.3%. The value of phase shift (n) for CPEdl also increased from 0.79 without rMrCP20 to 0.85 in the presence of 5 mg/mL rMrCP20, suggesting a decrease in the heterogeneity of the coupon surface arising from adsorption of rMrCP20. Cdl for the control sample was 757 µF cm− 2, which was significantly higher than the corresponding value for the system incubated in rMrCP20, indicating growth in the electrical double layer thickness. These results suggest that the adsorption of rMrCP20 at the metal/electrolyte interface impedes charge transfer due to greater resistance. Hence, at higher concentration of rMrCP20, the AH36 steel surface is better protected against corrosion.
To obtain further qualitative understanding of the corrosion reactions, potentiodynamic polarization measurements (PDP) measurements were carried out on AH36 samples after 24 h incubation in the respective solutions, and the polarization curves are shown in Fig. 2e. The polarization curves present two opposing reactions, the anodic branch corresponding to Fe dissolution and the cathodic branch due to hydrogen evolution. The obtained polarization parameters, namely corrosion potential (Ecorr), corrosion current density (icorr), anodic and cathodic Tafel slopes (\({\beta }_{a}\) and \({\beta }_{c}\)) measured by the Tafel extrapolation method are presented in Fig. 2f. The corrosion rate \(\left(CR\right)\) and corrosion inhibition efficiency \({(\eta }_{PDP})\) of rMrCP20 was calculated using Eqs. (5–6):29
\({\eta }_{PDP}\left(\%\right)=\left(1-\frac{{i}_{corr}^{0}}{{i}_{corr}^{{\prime }}}\right)\times 100\) (Eq. 5)
\(CR=3.17{E}^{-9}\times \frac{M}{nF\rho A}{i}_{corr}\) (Eq. 6)
where \({i}_{corr}^{0}\) and \({i}_{corr}^{{\prime }}\) are the corrosion current densities in the presence and absence of rMrCP20, respectively, 3.17E− 9 is conversion factor, the ratio M/n is the equivalent weight, F (96485 C mol− 1\()\) is the Faraday constant, \(\rho \left(7.86 g {cm}^{-3}\right)\) the density of the metal samples, and A is the area of the sample \(\left({cm}^{-3}\right)\). As shown in Fig. 2e, both the anodic and cathodic icorr decreased in the presence of rMrCP20, and Ecorr shifted towards the cathodic direction, which became more pronounced as the concentration of rMrCP20 increased. In addition, Fig. 2f shows the shift in ΔEcorr values to less than 85 mV, which indicates that rMrCP20 is a mixed-type inhibitor30,31 with a stronger effect on the cathodic reaction than on the anodic reaction. This result was complemented with cyclic voltammetry (CV) measurements (Supplementary Fig. 2). Since the peaks of the redox reactions measured by CV were not shifted at different concentrations of rMrCP20, the reaction mechanism appeared to remain consistent. In addition, the peak currents observed in CV scans for both the oxidation and reduction reactions progressively decreased with the addition of rMrCP20, further confirming the lowered occurrence of redox reactions, and hence the corrosion inhibition properties of rMrCP20. Furthermore, the peak currents from CV measurements remained constant in the presence of rMrCP20 with repetitive scans, whereas in the absence of the protein the peak currents were much greater for the second scan.
As shown in Fig. 2f, icorr decreased from 28.5 \(\mu A {cm}^{-2}\) without rMrCP20 down to 4.0 \(\mu A {cm}^{-2}\) with 5 mg/mL of rMrCP20. The lower values of icorr in the presence of rMrCP20 can be attributed to protein adsorption onto the sample’s surface and hence to the formation of a protective layer, which hinders the electron transfer process at the interface and reduces the rate of corrosion reactions. Figure 2f also illustrates the effect of rMrCP20 on the Tafel slopes values βc, which were independent of the rMrCP20 concentration, indicating that the presence of rMrCP20 can suppress the corrosion process by blocking reaction sites without affecting the kinetics of the cathodic reactions.32 Meanwhile, a gradual decline in βa was observed with increasing rMrCP20 concentration, from 267 mV dec− 1 in buffer to only 110 mV dec− 1 in the presence of 5 mg/mL rMrCP20, indicating a change in the iron dissolution mechanism. Iron dissolution is attributed to the formation of metal-ion protein complexes as the protein interacts with the steel surface, which was further investigated by attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR), QCM-D and MD simulations (discussed below). Accordingly, the corrosion-inhibition efficiency improved at increased rMrCP20 concentration, from 24% (1 mg/mL) to 86% (5 mg/mL), respectively. In summary, the corrosion inhibition values η obtained from all three methods (weight loss, EIS and PDP measurements) are consistent (Supplementary Fig. 3) and all indicate enhanced corrosion-inhibition activity as the concentration of rMrCP20 increased.
Adsorption of rMrCP20 and Metal ion – Protein Interaction studies
QCM-D measurements were carried out on Fe sensors to examine the adsorption properties of rMrCP20, followed by subsequent addition of FeCl3 analyte as a source of free Fe3+ to investigate the Fe3+/adsorbed protein layer interactions. The changes in resonance frequency (Δf) and dissipation (ΔD) of the sensor surface were measured simultaneously. Figure 3a shows the results obtained for the 5th overtone as a function of time for the following steps: flowing of rMrCP20, rinsing, flowing of FeCl3 solution and second rinsing (see Materials and Methods). To allow direct comparison of the Δf and ΔD shifts measured at different harmonics, the net shifts of the 5th and 11th overtones are displayed in Figs. 3b and c. Upon injection of rMrCP20, an exponential increase in Δf intensity was measured (Fig. 3a), followed by a plateau, indicative of a rapid initial adsorption of rMrCP20 onto the Fe sensor surface followed by saturation at the surface. Δf recorded was ca. -40.0 Hz and − 39.2 Hz upon adsorption saturation for the 5th and 11th overtones, respectively. The significant decrease in Δf was accompanied by a proportional increase in ΔD to 1.62 x 10− 6 and 1.55 x 10− 6 for 5th and 11th overtones, respectively. Fast initial adsorption rates suggests that rMrCP20 has a high affinity for the Fe sensor surface, which can be driven by complementary electrostatic and hydrophobic adsorption.33–35 The changes in Δf and ΔD due to rMrCP20 binding to the Fe sensor surface were generally similar amongst different overtones, indicating that the protein formed a compact and rigid layer.33 While the inversely proportionate changes in ΔD and Δf denote the strong binding kinetics of rMrCP20 on Fe substrate, ΔD decreased gradually while Δf plateaued, suggesting that the adsorbed protein layer more tightly packed over time. The changes of Δf was similar at different rMrCP20 concentrations, indicating that rMrCP20 formed a mono-adlayer (Supplementary Fig. 4c). Upon rinsing, a slight decrease in Δf intensity was detected due to the removal of excess unbound proteins, but most proteins remained strongly bound to the Fe surface.
The introduction of FeCl3 induced large shifts in the QCM-D signals, splitting the response across the different harmonics (Supplementary Figs .4a, b). The increase in Δf intensity indicates significant binding interaction between the analyte and adsorbed protein layer, which can occur via formation of Fe3+ metal-protein complexes and electrostatic interactions. The proportional increase in ΔD implies a loosen packing of the adlayer, which allows more ingression of water molecules that remained trapped within the adlayer.34 Upon the final rinsing step to remove any loosely bound material, only a slight decrease in Δf intensity was observed, suggesting that Fe3+ ions were strongly bound to the pre-adsorbed rMrCP20 layer, whereas the slight decrease in ΔD indicated stiffening of the surface layer.
The mode of interaction between rMrCP20 and the metal surface can be deduced from adsorption isotherms, which were obtained by adapting the results from weight loss measurements (Supplementary Table 1) to determine the adsorption characteristics of rMrCP20 on AH36 steel in buffer solution. The degree of surface coverage (θ) as function of concentration (C) of the protein was fitted into various adsorption isotherm models, including Langmuir, Temkin, Freundlich, Flory–Huggins, Frumkin and El-Awady to identify the best fit based on the obtained R2 values. The best fit was obtained for the Freundlich model (R2 = 0.96, Supplementary Table 2), which is generally applicable for multilayer adsorption events on heterogeneous surfaces, with the assumption of a large number of different types of binding sites acting simultaneously.36–38 The linear form of the Freundlich model can be written as: 36
\(log\theta =\frac{1}{n}log{C}_{inh}+log{K}_{ads}\left(7\right)\) (Eq. 7)
where Kads represents the Freundlich adsorption capacity (L/mg) and n describes the heterogeneity of the system related to the adsorption intensity. A larger n value connotes a more heterogeneous system, and n > 1 suggests a favorable adsorption process.39 Freundlich constants Kads and n values were \(1.33 \times {10}^{-2} L/mg\) and 2.18, respectively. The value of Kads was then used to calculate the standard free energy of adsorption (\({\varDelta G}_{ads}^{0})\):23
\({\varDelta G}_{ads}^{0}=-RT\text{ln}\left(1 \times {10}^{6}{K}_{ads}\right)\) (Eq. 8)
where R is the universal gas constant, T is the absolute temperature, and 106 is the concentration of water molecules expressed in ppm. The calculated \({\varDelta G}_{ads}^{0}\)value of -23.52 kJ/mol indicates spontaneous interaction between the protein and the surface of AH36 coupons, and is within the range for mixed adsorption involving both physisorption driven by electrostatic interactions as well as chemisorption caused by charge sharing or charge transfer from the protein molecules to the metal surface.25, 40, 41
To determine the thickness of the protein and metal-protein layers, we carried out nanoindentation (Fig. 4a) on samples prepared similarly to QCM-D studies but on gold sensor substrates. Samples were equilibrated in buffer, and a cube-corner tip was used to indent the samples via a displacement-controlled indentation of 200 nm in depth. Figure 4b shows the surface roughness profile of the samples and Fig. 4c the histogram of the average layer thickness, indicating significant difference between samples. The layer thickness was inferred from the slope change during the loading step of indentation cycles, as illustrated on load vs. displacement curves shown in Fig. 4d. The presence of a layer ca. 74 nm thick was consistently observed on the substrate when hydrated and was attributed to a thin oxide layer. Upon adsorption of rMrCP20, an indentation depth of ca. 82 nm was measured before the tip contacted the substrate, indicating that the protein formed a layer of ca. 8 nm (subtracting the thickness of the oxide layer). Subsequent indentation of the protein layer after interaction with FeCl3 showed a layer of thickness ca. 79 nm, indicating that the protein layer became slightly more compact to ca. 5 nm. The results corroborate with the QCM-D results, where the adsorbed rMrCP20 layer became more compact and increased in density upon interacting with FeCl3.
To assess the effect of Fe ions on rMrCP20 secondary structure and identify the type of binding interactions between Fe ions and the protein’s side chains, ATR-FTIR spectroscopy measurements were conducted (Fig. 5). With rMrCP20 kept at 5 mg/mL concentration, different concentrations of FeCl3 were added as indicated in Fig. 5a. The most significant changes in the spectra were observed at the protein:FeCl3 molar concentration ratio of 1:60. At low FeCl3 concentration, minimal changes were observed compared to rMrCP20-only spectra. The amide I band was further deconvoluted to obtain a semi-quantitative estimate of the secondary structural content, as shown in Fig. 5c. The initial β-sheet content of rMrCP20 protein was ~ 53% at 5 mg/mL. At the 1:60 protein:FeCl3 concentration ratio, the amide I maximum of rMrCP20 spectra shifted from 1641 cm− 1 to 1631 cm− 1 (Fig. 5b), indicating a transition towards anti-parallel β-sheet structures. Similarly, the appearance of 1553 cm− 1 in amide II confirm that FeCl3 at high concentrations induced changes in the protein’s conformation. The amide III peaks at 1262 cm− 1 and 1296 cm− 1 were likewise more noticeable at high FeCl3 content.
New peaks were noticeable at higher concentrations of FeCl3, indicating that new bonds formed between the protein and Fe3+. The appearance of 800 cm− 1 and 910 cm− 1 bands were assigned to the stretching of aromatic -CH groups and C = C, respectively, attributed to their interaction with Fe3+. The distinct splitting of the band centered at 1043 cm− 1 to 1038 cm− 1 and 1053 cm− 1 can be attributed to chelation of Fe3+ ions by the imidazole side chain of histidine (His).42 The band at 1403 cm− 1, assigned as S = O bond appeared in the presence of Fe3+, and was attributed to the oxidation of thiol functional groups of cysteine residues. The appearance of 1138 cm− 1, 1262 cm− 1 and 1296 cm− 1 bands were assigned to new interactions involving the protein’s side chain C-OH, C-O/-CH3 and amide bond C = O functional groups, respectively. The intensity of the COO− stretching band of acidic residues at 1398 cm− 1 observed for rMrCP20 decreased, concurrently with the appearance of a peak at 1437 cm− 1 upon introduction of high FeCl3 concentrations, is assigned to the formation of ionic bridges or coordination bonds between the carboxyl side chains and Fe3+.42,43