Skip to main content

Mechanism of corrosion protection in chloride solution by an apple-based green inhibitor: experimental and theoretical studies

Abstract

Preservation of metals in infrastructures and other assets requires cost-effective and sustainable solutions such as green corrosion inhibitors. This study assesses an apple pomace-derived green inhibitor synthesized by an innovative zero-waste method. Electrochemical measurements revealed the high performance of this liquid extract in reducing the corrosion of carbon steel in NaCl brine. The chemical composition of this inhibitor was characterized by liquid chromatography mass spectroscopy (LC-MS) to shed light on the corrosion inhibition mechanism. Based on LC-MS analysis, the results of surface analysis were interpreted. Specifically, the major corrosion inhibitor agent in the apple pomace extract was determined as C26H50NO7P (1-Linoleoyl-sn-glycero-3-phosphocholine), which can adsorb onto the steel surface to form a barrier layer and serve as a blocker of active anodic sites. Further study showed that the apple extract adsorption follows the Langmuir isotherm, and physical adsorption is dominant (vs. chemical adsorption). Theoretical calculations using quantum chemistry proposed a physisorption mechanism for the protection of steel by C26H50NO7P molecules.

Graphical abstract

Introduction

Chloride-based deicers have been used extensively in winter road maintenance operations [24, 37]. Since these chemicals are water soluble and corrosive, they can pose a great risk on the integrity, performance, and service life of steel structures (and components) [7, 43]. They may induce premature failures, which have substantial implications on the reliability, resilience, serviceability, and environmental footprints of the structure [29, 69]. One of the cost-effective methods for preventing or reducing the corrosion damage entails the application of corrosion inhibitors [20, 59, 78]. This method has been widely employed by maintenance agencies to mitigate the corrosion risk of deicers [50]. The efficiency of corrosion inhibitors is defined by their physicochemical properties such as active functional groups, electronic structure, electron density, non-bonding (lone pair) electrons, and π-electrons [64, 66].

Traditional corrosion inhibitors contain chemical compounds that can have toxicological effects and pose harmful impacts on the receiving environment, especially surface waters [18, 30]. In recent years, green corrosion inhibitors have been introduced as alternatives to traditional inhibitors [13, 38, 40]. Bio-based materials have been used in the winter maintenance operations of roadways, either as freezing point depressant, or as green corrosion inhibitor [53]. These materials are mostly produced by fermentation and processing of desugared molasses, beet juice, corn, and other agricultural products [2, 16].

Recently, the authors have introduced an innovative zero-waste method for producing agro-based corrosion inhibitors as an additive to deicer/anti-icer chemicals. The major advantage of the chemicals prepared by this method over the commercial green chemicals used by the departments of transportation (e.g., beet juice) is that the former ones feature much lower chemical oxygen demand, which makes them more eco-friendly [31]. An example of inhibitors synthesized by the innovative method is a liquid extract made from peony leaf waste, exhibiting a corrosion inhibition efficiency of 65.8% for carbon steel in NaCl brine [51]. Due to relatively low corrosion inhibition efficiency (IE) of the peony leaf extract, the authors further explored a variety of other local agro-based wastes as feedstock (such as grass, dandelion leaf, blueberry pomace, cherry pomace, and apple pomace) to produce high-efficiency “green” corrosion inhibitors. After a preliminary study, the apple pomace seemed to be the most promising agricultural waste for producing green inhibitor among the tested wastes. Therefore, we explored this feedstock further. In addition, the amount of apple waste in Washington State is 27,794 tons of dry biomass per year [27]. Therefore, the current research may lay the foundation of converting a huge amount of waste to an eco-friendly product, which can notably benefit the society and the environment.

Apple pomace has been used as a food ingredient, and for producing jam and jelly [25, 41]. Some researchers have used commercial apple-derived pectin (C6H10O7) as a green inhibitor [79]. Pectin can be extracted from apple pomace using different acid extraction methods which produce solvent waste [42, 81]. However, no waste is produced during the extraction process of producing the apple pomace extract herein, which makes it a more environmentally friendly corrosion inhibitor than commercial pectin.

This work investigated the effect of an apple pomace liquid extract on the corrosion of C1010 mild steel in 3.5% NaCl brine. A zero-waste process developed by the authors was adopted to produce this “green” corrosion inhibitor. Chemical, electrochemical, and surface analyses were used to shed light on the corrosion behavior. Finally, using quantum chemical calculations we propose a mechanism of corrosion inhibition.

Experimental

Materials

The corrosion coupons were prepared from C1010 steel with a composition presented in Table 1. The exposed surface area of each coupon was 1 cm2. The coupons were polished by 60- to 1500-grit sandpaper sequentially, degreased by ethanol, and then washed with deionized water. In the next step, they were exposed to 3.5% NaCl brine over a 7-day period (168 h). The reason for choosing 3.5% NaCl brine for this research is that it is known that NaCl concentration of around 3% is the most corrosive concentration of NaCl brine for iron in aerated solutions at room temperature [11]. Therefore, the authors used this concentration to simulate a worst-case scenario. In the field, the concentration of 23% NaCl is the common as-applied concentration, for anti-icing or deicing applications. However, due to snow precipitation and ice melting conditions, the stormwater runoff tends to have much lower concentrations of NaCl.

Table 1 Composition of the steel samples

Different concentrations of the apple pomace extract (1%, 2%, and 3% v/v (volume of extract/total volume)) were employed in this research. These values corresponded to dry mass per volume of apple pomace extract (APE) at 2.61, 5.22, and 7.82 g/L, respectively. For drying the APE, it was oven dried at 103 °C for 24 h. The code names of the samples are shown in Table 2. While the range of 1–3% v/v is high relative to the range of ppm used by many researchers, it is much lower than 20% v/v that is common for using agro-based byproducts such as beet juice in winter maintenance operations. In addition, some researchers have reported the range of up to 4% v/v of alcohol-based inhibitors for protection of reinforcing steel [36].

Table 2 Code-names of the coupons in this work

The apple pomace extract was prepared using a chemical/biological process that has been described elsewhere in detail [51]. Briefly, in this process the apple pomace, which contained seed and peel, underwent two degradation steps: first, chemical degradation in an alkaline solution containing urea at around − 13 °C, then biological degradation by bacteria at room temperature. The resulting solution was heated at 85 °C until its volume halved. The concentrated solution was used as the corrosion inhibitor without further treatment or purification. All the chemicals used in this research were of analytical- grade purity.

It should be mentioned that the process used in this research is more complex than common processes of dissolving the agricultural feedstock in a solvent such as a mixture of water and alcohol, or acid [17, 76].

Mass spectrometry

The liquid chromatography–mass spectrometry (LC-MS) method was employed for in-depth chemical analysis of the APE. The analysis was conducted in the positive ion mode using Synapt G2-S mass spectrometer. Calculation of the normalized weight percent (wt.%) of each compound was conducted based on the normalized percent surface area of each peak with the aid of the ImageJ software.

Electrochemical measurement

A PARSTAT MC multichannel Potentiostat performed the corrosion measurements using a standard three-electrode system comprising the working electrode (corrosion coupon), a reference electrode (Ag/AgCl, KCl saturated), and a counter electrode (platinum mesh). The electrochemical impedance spectroscopy (EIS) measurements were carried out using the same system in the frequency range of 100 kHz – 10 mHz by superimposing AC signal of ±10 mV around the open-circuit potential (OCP). Tafel polarization test was conducted at a potential sweep rate of 0.167 mV/s in the range of − 250 mV to + 250 mV with respect to OCP. The VersaStudio software was used for running the electrochemical corrosion tests and calculating the Tafel polarization parameters, and ZSimpWin 3.60 software was employed for interpretation of the EIS data. The tests were at least duplicated to obtain an acceptable reproducibility. All of the electrochemical tests were performed at room temperature.

Surface analysis

The surface layer was examined via a JEOL JXA-8500F electron probe micro-analyzer (EPMA). The EPMA uses an electron probe to generate irradiation on the specimen surface, and then the characteristics of the produced X-ray are analyzed by wavelength-dispersive X-ray spectroscopy (WDS) techniques to qualify the element distributions on the surface. For each area, five elements (e.g. Fe, N, Cl, O, and P) were analyzed and their spatial distribution mapped.

In order to study the surface layer more in-depth, X-ray photoelectron spectroscopy (XPS) analysis was used. The analysis was carried out using an AXIS-165 multi-electron spectrometer from Kratos Analytical Inc. The data acquisition and data analysis were performed using the XPSPEAK 4.1 software.

Quantum chemical calculation

All the calculations were performed using molecular mechanics/MM+ level, and the semi-empirical calculations with PM3 method [74] at 25 °C by Hyperchem 7.52 software.

Results and discussion

Chemical composition of the liquid extract

Figure 1 and Table 3 show the results of LC-MS analysis for the APE. Based on the mass spectrum data, the APE molecules mainly include  C26H50NO7P and C31H43N5O (and other organic compounds), which consist of nitrogen, oxygen, and phosphorus elements. The organic mixtures containing N, P, and O elements have polar functions, so will adsorb onto the metallic surface as a barrier organic layer [19]. The adsorption of the polar atoms on the surface of metal may occur through coordination between the lone pair or π-electrons cloud and the metallic surface [23, 71].

Fig. 1
figure 1

LC-MS spectra of the APE

Table 3 Chemical constituents identified for APE using LC-MS analysis

Electrochemical characterization

OCP measurements

The OCP of steel coupons in the presence and the absence of the APE was measured after different periods of immersion in NaCl brine, as shown in Fig. 2. The OCP can be a function of corrosion potential and electrical resistivity of the protective layer formed on the sample [83]. It can be seen that the OCPs of the samples exposed to the solutions containing the apple pomace extract shifted to more noble values, which may be a sign of the more thermodynamically steady system [33]. This conclusion has been confirmed by considering the steady trend of the OCP for sample APE3 after the third day of the experiment (72 h). This effect is not the same for all concentrations; for instance, APE2 experiences a sharp decrease in OCP values after 72 h; even more than APE1. This behavior can be due to the probabilistic nature of corrosion which causes fluctuations of OCP [28].

Fig. 2
figure 2

OCP versus time for mild steel in NaCl brine containing 0 to 3% (v/v) APE

Tafel polarization results

Figure 3 presents the Tafel polarization plots of the samples after 7-day immersion in NaCl brines having different concentrations of APE. It is seen that addition of APE shifted the corrosion potential to a more noble potential and decreased the anodic current density. These can be ascribed to the adsorption of organic molecules on the anodic active sites of the steel surface [1, 14].

Fig. 3
figure 3

Tafel polarization curves for steel samples after 7-day exposure to NaCl brine containing 0 to 3% (v/v) apple pomace extract

Table 4 shows the electrochemical parameters derived from the Tafel polarization plots. By increasing the concentration of the APE, the corrosion current density (icorr) was notably decreased and the lowest icorr was obtained at 3% (v/v) apple pomace extract. The maximum inhibition efficiency (IE%) of APE calculated using icorr value was 83%, which is significantly more than that of pectin (78.7%) for X60 steel in 0.5 M HCl solution [79] and that of Sarang Semut (Myrmecodia Pendans) extract (65.21%) for low carbon steel in 3.5% NaCl solution [62].

Table 4 Electrochemical parameters fitted by using Tafel polarization data

By adding the extract at 3% (v/v), the corrosion potential (Ecorr) of the steel increased by more than 85 mV, which shows that APE is an anodic inhibitor and the corrosion inhibition is attributable to change in the activation energy [14, 21]. Based on Table 4, the values of anodic and cathodic Tafel slopes were 104.1 and 384.4 mV/dec, respectively for the APE0 sample. Both anodic and cathodic Tafel constants decreased with increases in the APE concentration.

EIS results

Figure 4 depicts the impedance Nyquist curves of steel samples in NaCl brines with various concentrations of APE. The capacitive arc radius increased by increasing the concentration of the apple pomace extract, which corresponds to improved corrosion resistance [15, 49]. The capacitive arcs for samples APE0 and APE1 have similar forms despite their different sizes, which shows that the inhibition mechanism is independent of  theextract concentration for these samples. However, the feature of the capacitive arc with a Warburg impedance emerged in samples APE2 and APE3. The observed arcs in all samples are non-ideal semi-circles which is likely due to the frequency dispersion caused by corrosion-induced non-homogeneity of the steel surface [8, 12]. In addition, adsorption of the barrier organic layer on the surface of metal can induce surface heterogeneity that can affect the shape of capacitive arc [9, 58].

Fig. 4
figure 4

Nyquist curves for steel samples after 7-day immersion in NaCl brine containing 0 to 3% (v/v) APE

The Bode curves in Fig. 5 reveal that by increasing the APE concentration, |Z| value at the frequency impedance of 10 mHz (Log f = − 2) increased. Since this value corresponds to the corrosion protection provided by the inhibitor [65], it shows that by increasing the inhibitor concentration, the corrosion protection was improved.

Fig. 5
figure 5

Bode curves for steel samples after 7-day immersion in NaCl brine containing 0 to 3% (v/v) APE

The equivalent electrical circuits used for studying the impedance spectra are shown in Fig. 6. The circuits consist of the following elements: Rs (solution resistance), Rf (corrosion film resistance), Qf (constant phase element (CPE) of film), Rct (charge transfer resistance of metal/solution interface), and Qdl (CPE of metal/electrolyte interface), and W (Warburg impedance).

Fig. 6
figure 6

Equivalent circuit models employed for assessing the EIS results

Some of the values of the fitted parameters based on the equivalent electrical circuit model are summarized in Table 5. In the equivalent electrical circuit, Qdl is the CPE of the electrical double layer at the metal/electrolyte interface. The CPE represents an non-ideal capacitor in the EIS results, and its impedance is stated as follows [73].

$$ {Z}_{CPE}=\frac{1}{Y_0{(jw)}^n} $$
(1)
Table 5 EIS parameters obtained by fitting the data to an equivalent circuit for the samples immersed for 7 days in NaCl brine

Where Y0 is the CPE magnitude, n is the phase shift due to surface roughness (− 1 ≤ n ≤ 1), and ω is the angular frequency. The values of Cdl, represented in Table 5, were measured as follows [44].

$$ {C}_{dl}={\left({Y}_0{R_{ct}}^{1-n}\ \right)}^{1/n} $$
(2)

The Cdl values decreased with increases in the concentration of APE, which can be ascribed to the reduction of double layer dielectric constant [10]. This means a thicker electrical double layer, which is caused by more displacement of corrosive species by APE molecules at the steel/electrolyte interface [57, 75]. By increasing the concentration of APE, the values of Rct increased which could translate to higher inhibition efficiencies [26]. In addition, Rf increased in the presence of higher concentrations of APE, which may pertain to the decrease in the number of ionically conducting paths. This implies a slowed-down migration rate of aggressive species (such as Cl ions) through the adsorbed protective layer towards the surface of steel. The Warburg impedance was observed in APE2 and APE3 samples, which shows a resistance due to the diffusion process over the porosity of the corrosion product film [32, 35].

The fitted curves for Nyquist and Bode curves of samples APE0 and APE3 are shown in Fig. 7. As can be seen, there is a good match between the modeled and experimental results. For modeling the electrochemical behavior of samples APE0 and APE3, the equivalent circuits presented in Fig. 6 (a) and 6 (b) were used, respectively. The maximum IE of APE calculated using EIS data was 98.8%, which was much more than that of pectin (77.1%) for X60 steel in 0.5 M HCl solution [79] and that of Myrmecodia Pendans extract (79.7%) for the corrosion of mild steel API 5 L Grade B in 3.5% NaCl [63].

Fig. 7
figure 7

Nyquist curves (a, c) and Bode curves (b, d) with the fitted curves for steel sample after 7-day immersion in NaCl brine containing 0% (v/v) of APE (APE0) and 3% (v/v) of APE (APE3)

The Nyquist curves of the APE3 sample at various immersion periods are shown in Fig. 8. The protection responses were modified over time, controlled by the charge transfer mechanism for all time periods. The biggest capacitance arc was observed at day 7, which showed the stability of the protective layer over time. The Bode curves depicted in Fig. 9 demonstrate the increasing of the |Z|10mHz with APE concentration, which shows improvement in anti-corrosion performance of the inhibitor and its stability over time. After five days of immersion, the Bode curve showed the fingerprint of two time constants, which can be explained by the changing of the protection mechanism from geometry blocking to energy effect [14]. This is in good agreement with Tafel results that linked the APE corrosion inhibition to changing of the activation energy.

Fig. 8
figure 8

Nyquist curves for steel samples immersed for different periods in NaCl brine containing 3% (v/v) APE

Fig. 9
figure 9

Bode curves for steel samples immersed for different periods in NaCl brine containing 3% (v/v) APE

Adsorption isotherm

The interaction of APE molecules with the steel surface can be determined by using an adsorption isotherm. In this regard, the surface coverage degree (θ) was assumed to be equivalent to inhibition efficiency (IE) measured by the EIS technique. The IE versus APE concentration was fitted to Langmuir, Freundlich, Flory-Huggins, and Temkin adsorption isotherms. The Langmuir isotherm exhibited the best fit to the data, with the correlation coefficient of 0.997 (Fig. 10). The Langmuir isotherm follows the following equation [54].

$$ \mathit{\log}\left(\frac{C}{\theta}\right)=l\mathrm{o} gC- logK $$
(3)
Fig. 10
figure 10

The isotherm for adsorption isotherm of APE (C, in g/L, dry mass per volume of corresponding APE)

Where C is the concentration of APE in g/L (dry mass/volume of extract solution), K is the adsorption coefficient (L/g). Classical adsorption isotherms (such as Langmuir) use the model of non-penetrable interface in which a solvent can be substituted by the molecules of an adsorbate [45].

The change in the free energy of adsorption (ΔG0ads) was measured as follows [54, 68].

$$ \Delta {G}_{ads}^0=-2.303 RTlog\left({C}_{solvent}K\right) $$
(4)

Where R designates the gas constant, T shows the absolute temperature, K is the adsorption coefficient (L/g), and Csolvent is the water concentration (1000 g/L). The Gibbs free energy was estimated to be − 16.33 kJ mol− 1. The negative value of ΔG0ads suggested the spontaneous occurrence of adsorption process of APE species onto the steel surface. The absolute value of Gibbs free energy was smaller than 20 kJ mol− 1, which suggests that the physical adsorption is dominant, rather than chemical adsorption [72]. In this condition, the electronic structure of the adsorbate is disturbed because of the electrostatic interaction between APE molecules and the steel surface [70].

Surface analyses

SEM micrographs

The SEM images of as-washed coupons are provided in Fig. 11. A rough surface was observed on the uninhibited sample (APE0) and smoother surfaces on the samples exposed to APE containing solutions. Other researchers for the uninhibited and inhibited samples in Cl containing media reported similar results [3]. The amount of metallic surface covered by the organic layer increased with increases in the concentration of APE. This is in good agreement with the results obtained by electrochemical measurements and adsorption isotherm.

Fig. 11
figure 11

SEM micrographs of as-washed coupons after 7-day immersion in NaCl brine with APE at (a) 0%, (b) 1%, (c) 2%, and (d) 3% (v/v)

EDS results

The main elements of the surface layer formed on the as-washed APE0 – APE3 steel samples are shown in Fig. 12. The main peaks for all samples corresponded to iron and oxygen, which suggests the formation of iron oxides on the surface. This is in good agreement with the results obtained by EIS measurement that suggested the formation of a corrosion product layer on the surface of steel. In the case of APE3, a relatively large peak related to phosphorus can be seen around 2 keV, which is attributed to the adsorption of inhibitor molecules onto the steel surface.

Fig. 12
figure 12

EDS results of the corrosion surface layer for (a) APE0, (b) APE1, (c) APE2, (d) APE3 steel samples

EPMA results

For more in-depth analysis of the corrosion surface layer, EPMA analysis was employed which enables simultaneous detection of different elements and mapping of their distribution on a defined surface area. The EPMA elemental maps for the surface layer formed on the APE0 and APE3 steel samples are shown in Fig. 13. On the steel surface exposed to uninhibited NaCl brine (APE0), there is little presence of N or P and the corrosion product layer consisted of an iron oxide/hydroxide layer containing Cl. The interface between the corroded surface (left) and the intact steel substrate (right) is clearly illustrated by the elemental maps of Fe, O and Cl. In contrast, the surface layer in NaCl brine with high inhibitor concentration (APE at 3%) clearly consisted of a phosphorous compound. These results are in very good agreement with the results obtained by EDS. Therefore, from these results and the LC-MS data it can be concluded that C26H50NO7P is the major corrosion inhibitor agent in the APE. The charge density of phosphates is more negative than that of amines [34]; as such, the bond of former with metallic surface is more stable.

Fig. 13
figure 13

EPMA maps of the corrosion surface layer for APE0 and APE3 steel samples. The size of analyzed area was set at 50 μm × 50 μm

XPS results

Figure 14 depicts the results of XPS analysis for WPE0 and WPE3 samples. In the plot associated with Fe2p in WPE0, there is a peak at 708.37 eV which is related to Fe3O4 [46]. The XPS analysis of O1s unveils two peaks at 530.65 eV and 528.8 eV, which are associated with Fe3O4 and Fe3O4 complexes, respectively [4, 5]. In the presence of APE (sample APE3), two peaks appeared in the Fe2p plot. The major peak seen at 709.66 eV comes from Fe2O3 [60]. The second peak at 708.63 is due to Fe3O4 [46]. Therefore, in the presence of APE, the major part of corrosion product layer is converted to Fe2O3 from Fe3O4. It is known that Fe2O3 is more corrosion-resistant than Fe3O4 [39]. As such, one of the reasons that APE3 has a less corrosion rate than APE0 is the catalytic effect of APE in converting Fe3O4 to Fe2O3 in corrosion surface layer. It is known that phosphates can catalyze the oxygen evolution at near neutral pH [77]. In addition, they can stabilize the corrosion product layer [47]. Therefore, C26H50NO7P played the main role in converting Fe3O4 to Fe2O3. Observation of the N1s plot shows a peak at 400.75 eV that indicates adsorption of an amine-containing compound on the steel surface [82], which may be C31H43N5O. In addition, the plot of P2p shows a peak at 134.72 eV that reveals an amine-phosphate based constituent [61], which may have resulted from the adsorption of C26H50NO7P compound.

Fig. 14
figure 14

XPS plots for APE0 and APE3 samples

Quantum chemical calculations

Theoretical studies have prospective applications in designing and developing of many organic inhibitors in the field of corrosion inhibition chemistry. The capability of the inhibitor for corrosion prevention bases on its spatial distribution of molecule and its electronic arrangement over frontier molecular orbitals. Molecular geometry systems of compounds C26H50NO7P and C31H43N5O as the main compounds of the APE were determined. The stable geometry of the compounds was obtained by PM3 method, which was carried out in Hyperchem software (Fig. 15).

Fig. 15
figure 15

Conformation structure for C26H50NO7P and C31H43N5O

The calculated chemical parameters for C26H50NO7P and C31H43N5O molecules are given in Table 6. It is seen that the dipole moments (μ) for C26H50NO7P and C31H43N5O molecules are 17.0 D and 6.6 D, respectively. The organic molecules with high dipole moment may form electrostatic interactions with the metal surface, that result in a strong adsorption on the surface of metal [55]. In this research, the value of μ for C26H50NO7P is more than twice of the value of μ associated with C31H43N5O. This can justify the better corrosion inhibition properties of C26H50NO7P relative to C31H43N5O. These results are also in good agreement with the results obtained by Verma et al. that showed D-glucose derivatives with the dipole moment of 8.17 D having more than 96% corrosion inhibition efficiency in 1 M HCl solution [80]. In addition, the μ value of C26H50NO7P (17.0 D) is more than that of pectin, 11.725 D [79], which justifies the better anti-corrosion performance of APE than pectin.

Table 6 Quantum calculation results of the APE obtained by using PM3 method

The energy gap (ΔE = ELUMO – EHOMO) between the ELUMO and EHOMO levels of inhibitor molecules is another important electronic parameter which is given in Table 6. The values of energy gap for C26H50NO7P and C31H43N5O molecules are 7.7548 eV and 8.3776 eV, respectively. Low absolute values of the energy gap means good inhibition efficiency of the organic inhibitor [56]. It is evident that C26H50NO7P has a lower energy gap than C31H43N5O. This confirms the EDS and EPMA results, which showed that C26H50NO7P is the main corrosion inhibitor agent in the APE. Similar observations have been reported by Aloysius et al. where VB7 with the ΔE of 3.769 eV showed stronger affinity towards the steel surface and better inhibition efficiency than VB1 with the ΔE of 6.555 eV [6].

Figure 16 displays the electrostatic potential maps for C26H50NO7P and C31H43N5O molecules. More dark red (negative) regions associated with nucleophilic reactivity in the potential map of C26H50NO7P and C31H43N5O indicate that these molecules are readily sharing free electron pairs with the metal surface (electrophilic agent) through electrostatic interactions. Also, the molecules with blue (positive) region have electrophilic reactivity [22].

Fig. 16
figure 16

Electrostatic potential maps of C26H50NO7P and C31H43N5O

Figure 17 displays the frontier orbitals including the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) for C26H50NO7P and C31H43N5O molecules. It can be seen in the HOMO of C26H50NO7P molecule that the isosurfaces have been localized via phosphate moiety. This moiety has donor sites that facilitate the adsorption on the metallic surface by O atoms. However, the side chain (trimethyl amine and ester group) did not contribute to the HOMO electronic distribution. On the other hand, LUMO of C26H50NO7P molecule plays the role of acceptor where the electronic distribution is localized on the O atoms of ester moiety and C atoms of C=C bonds. In addition, O atoms of phosphate moiety are able to coordinate with positive ions of the surrounding environment.

Fig. 17
figure 17

LUMO and HOMO distributions for C26H50NO7P and C31H43N5O

In Fig. 17, it can be seen that HOMO isosurfaces are localized in C31H43N5O via benzoimidazole and piperidine moieties. Benzoimidazole and piperidine moieties have donor sites of C31H43N5O that facilitate the adsorption on the metallic surface by N atoms. Nevertheless, the side chain (aryl groups and amide) has not any influence on the electronic distribution of the HOMO in C31H43N5O. In addition, N atoms of benzoimidazole and piperidine moieties have active sites to coordinate with ions, which enables them to be acceptor of electron. Finally, the theoretical studies revealed that the hydrocarbon groups attached to C26H50NO7P and C31H43N5O have no effect on LUMO or HOMO distributions and can just be as hydrophobic fragments.

It was observed that nitrogen and oxygen atoms in the studied molecules have high electron density. These sites are nominated to be the active sites to proper electrophiles and can be attached within the environment [48]. Based on this, a physisorption mechanism was proposed for adsorption of C26H50NO7P and C31H43N5O molecules on the metallic surface, as depicted in Fig. 18.

Fig. 18
figure 18

Physisorption mechanism of C26H50NO7P and C31H43N5O molecules on the steel surface

Mechanism of corrosion inhibition

The inhibition mechanism in the presence of APE was adsorption of an organic protective film, and chemical conversion of corrosion product layer to a more corrosion-resistant film. The adsorption type was Langmuir isotherm model in which a monolayer of organic molecules form a non-penetrable interface. In the Langmuir model, each active site on the metal surface can be related equally to an adsorbate molecule, and there is no interaction among the adsorbed molecules [67]. In addition, the saturation coverage occurs when all of the surface active cites are occupied by the adsorbate molecules [67]. According to the experimental results, the saturation behavior in the inhibitive characteristics occurred in 3% v/v of APE in this research, since the corrosion efficiency at this concentration of APE in close to 100%, thus it cannot increase considerably.

1-Linoleoyl-sn-glycero-3-phosphocholine (C26H50NO7P) has a major role in the inhibition properties of APE. The corrosion inhibitive behavior of C26H50NO7P can be related to the functional group of its molecular structure [52]. It was also responsible for converting Fe3O4 in corrosion product layer to a more corrosion-resistant iron oxide (Fe2O3). APE molecules may attach to the surface of steel though their high electron density sites [70]. The APE adsorption entails the interaction of non-bonding electrons on nitrogen and oxygen with the steel surface.

More APE was associated with higher Rct and Rf values which shows the formation of a more protective surface layer. On the other hand, more APE corresponded to  lower values of Cdl, due to displacement of corrosive agents by organic molecules via an adsorption process at the metal/media interface. The inhibition efficiency was improved by increasing of APE concentration, which indicates an active corrosion protection. The corrosion protection mechanism changed over time from geometry blocking to energy effect. APE molecules blocked the anodic active sites and shifted the corrosion potential to more passive values. The amount of this shift was more than 85 mV, and this suggests thatAPE is an anodic inhibitor and inhibition property is due to the change of activation energy.

Conclusions

This research assessed the corrosion behavior of C1010 steel in 3.5% NaCl brine in the presence of an apple pomace liquid extract. The corrosion mechanism in the presence of this extract was elucidated using electrochemical measurements, mass spectroscopy, and surface analyses and the following key findings have been obtained.

  • Apple pomace extract acted as a “green” corrosion inhibitor. The inhibition efficiency increased with increasing the concentration of the extract and over time.

  • An excellent inhibition efficiency of 98.8% was achieved at 3% of this green inhibitor after 7-day of immersion in NaCl media.

  • Apple pomace extract molecules adsorbed onto the surface of steel following the Langmuir isotherm and physical adsorption was dominant (vs. chemical adsorption).

  • The apple pomace extract blocked anodic active sites on the surface of steel, and converted Fe3O4 to Fe2O3, which is a more corrosion-resistant iron oxide.

  • The main corrosion inhibitor agent in the apple pomace extract was C26H50NO7P molecule.

  • The protection mechanism improved over time from geometry blocking to energy effect.

  • The quantum chemical calculations confirm the experimental results and suggest the physisorption mechanism of this green corrosion inhibitor.

Availability of data and materials

The datasets generated and/or analyzed during the current study are available in the Google Site repository, [https://sites.google.com/site/greensmartinfrastructure/ApplePomaceInhibitorDatasets.zip].

Abbreviations

LC-MS:

Liquid chromatography mass spectroscopy

IE:

Inhibition efficiency

APE:

Apple pomace extract

v/v:

Volume of extract/total volume

wt.:

Weight

EIS:

Electrochemical impedance spectroscopy

OCP:

Open-circuit potential

μ:

Dipole moments

EPMA:

Electron probe micro-analyzer

WDS:

Wavelength-dispersive X-ray spectroscopy

XPS:

X-ray photoelectron spectroscopy

i corr :

Corrosion current density

E corr :

Corrosion potential

W:

Warburg impedance

θ:

Surface coverage degree

ΔG0ads :

Free energy of adsorption

ΔE:

Energy gap

HOMO:

Highest occupied molecular orbital

LOMO:

Lowest unoccupied molecular orbital

References

  1. Abd El-Lateef HM, Abu-Dief AM, Abdel-Rahman LH, Sañudo EC, Aliaga-Alcalde N (2015) Electrochemical and theoretical quantum approaches on the inhibition of C1018 carbon steel corrosion in acidic medium containing chloride using some newly synthesized phenolic Schiff bases compounds. J Electroanal Chem 743:120–133. https://doi.org/10.1016/j.jelechem.2015.02.023

    Article  Google Scholar 

  2. Albright M (2005) Changes in water quality in an urban stream following the use of organically derived deicing products. Lake Reserv Manag 21:119–124. https://doi.org/10.1080/07438140509354419

    Article  Google Scholar 

  3. Al-Fakih AM, Abdallah HH, Aziz M (2019) Experimental and theoretical studies of the inhibition performance of two furan derivatives on mild steel corrosion in acidic medium. Mater Corros 70:135–148. https://doi.org/10.1002/maco.201810221

    Article  Google Scholar 

  4. Allen GC, Curtis MT, Hooper AJ, Tucker PM (1974) X-ray photoelectron spectroscopy of iron–oxygen systems. J Chem Soc Dalton Trans:1525–1530. https://doi.org/10.1039/DT9740001525

  5. Allen GC, Hallam KR (1996) Characterisation of the spinels MxCo1−xFe2O4 (M = Mn, Fe or Ni) using X-ray photoelectron spectroscopy. Appl Surf Sci 93:25–30. https://doi.org/10.1016/0169-4332(95)00186-7

    Article  Google Scholar 

  6. Aloysius A, Ramanathan R, Christy A, Baskaran S, Antony N (2017) Experimental and theoretical studies on the corrosion inhibition of vitamins – thiamine hydrochloride or biotin in corrosion of mild steel in aqueous chloride environment. Egypt J Pet. https://doi.org/10.1016/j.ejpe.2017.06.003

  7. Appleman BR (1987) Painting over soluble salts: a perspective. J Prot Coat Linings 4:68–82

    Google Scholar 

  8. Ashassi-Sorkhabi H, Asghari E (2008) Effect of hydrodynamic conditions on the inhibition performance of l-methionine as a “green” inhibitor. Electrochim Acta 54:162–167. https://doi.org/10.1016/j.electacta.2008.08.024

    Article  Google Scholar 

  9. Ashassi-Sorkhabi H, Seifzadeh D, Hosseini MG (2008) EN, EIS and polarization studies to evaluate the inhibition effect of 3H-phenothiazin-3-one, 7-dimethylamin on mild steel corrosion in 1 M HCl solution. Corros Sci 50:3363–3370. https://doi.org/10.1016/j.corsci.2008.09.022

    Article  Google Scholar 

  10. Behpour M, Ghoreishi SM, Khayatkashani M, Soltani N (2012) Green approach to corrosion inhibition of mild steel in two acidic solutions by the extract of Punica granatum peel and main constituents. Mater Chem Phys 131:621–633. https://doi.org/10.1016/j.matchemphys.2011.10.027

    Article  Google Scholar 

  11. Bin Ibrahim MF (2013) Effect of different sodium chloride (NaCl) concentration on corrosion of coated steel (bachelor of mechanical engineering). Universiti Malaysia Pahang. Pahang, Malaysia

  12. Bommersbach P, Alemany-Dumont C, Millet JP, Normand B (2005) Formation and behaviour study of an environment-friendly corrosion inhibitor by electrochemical methods. Electrochim Acta 51:1076–1084. https://doi.org/10.1016/j.electacta.2005.06.001

    Article  Google Scholar 

  13. Bozorg M, Shahrabi Farahani T, Neshati J, Chaghazardi Z, Mohammadi Ziarani G (2014) Myrtus Communis as green inhibitor of copper corrosion in sulfuric acid. Ind Eng Chem Res 53:4295–4303. https://doi.org/10.1021/ie404056w

    Article  Google Scholar 

  14. Cao C (1996) On electrochemical techniques for interface inhibitor research. Corros Sci 38:2073–2082. https://doi.org/10.1016/S0010-938X(96)00034-0

    Article  MathSciNet  Google Scholar 

  15. Chen Y, Xing W, Wang L, Chen L (2019) Experimental and electrochemical research of an efficient corrosion and scale inhibitor. Mater Basel Switz 12. https://doi.org/10.3390/ma12111821

  16. Cheng KC, Guthrie TF (1998) Liquid road deicing environment impact (498–0670). Insurance Corporation of British Columbia. North Vancouver

  17. Chevalier M, Robert F, Amusant N, Traisnel M, Roos C, Lebrini M (2014) Enhanced corrosion resistance of mild steel in 1M hydrochloric acid solution by alkaloids extract from Aniba rosaeodora plant: electrochemical, phytochemical and XPS studies. Electrochimica Acta, Electrochem Impedance Spectroscopy 131:96–105. https://doi.org/10.1016/j.electacta.2013.12.023

    Article  Google Scholar 

  18. Chigondo M, Chigondo F (2016) Recent natural corrosion inhibitors for mild steel: an overview. J Chem 2016:1–7. https://doi.org/10.1155/2016/6208937

    Article  Google Scholar 

  19. Dariva CG, Galio AF, 2014. Corrosion inhibitors – principles, mechanisms and applications, in: Aliofkhazraei, M. (Ed.), Developments in corrosion protection. InTechOpen, London

  20. Dong S, Yuan X, Chen S, Zhang L, Huang T (2018) A novel HPEI-based Hyperbranched scale and corrosion inhibitor: construction, performance, and inhibition mechanism. Ind Eng Chem Res 57:13952–13961. https://doi.org/10.1021/acs.iecr.8b03522

    Article  Google Scholar 

  21. Ebadi M, Basirun WJ, Khaledi H, Ali HM (2012) Corrosion inhibition properties of pyrazolylindolenine compounds on copper surface in acidic media. Chem Cent J 6:163. https://doi.org/10.1186/1752-153X-6-163

    Article  Google Scholar 

  22. Ebenso EE, Kabanda MM, Murulana LC, Singh AK, Shukla SK (2012) Electrochemical and quantum chemical investigation of some Azine and thiazine dyes as potential corrosion inhibitors for mild steel in hydrochloric acid solution. Ind Eng Chem Res 51:12940–12958. https://doi.org/10.1021/ie300965k

    Article  Google Scholar 

  23. El-Hafez GMA, Badawy WA (2013) The use of cysteine, N-acetyl cysteine and methionine as environmentally friendly corrosion inhibitors for cu–10Al–5Ni alloy in neutral chloride solutions. Electrochim Acta 108:860–866. https://doi.org/10.1016/j.electacta.2013.06.079

    Article  Google Scholar 

  24. Fay L, Nazari MH, Jungwirth S, Muthumani A (2015) Snow and ice control environmental best management practices. Environ Sustain Transp Infrastruct, Proc:147–161. https://doi.org/10.1061/9780784479285.013

  25. Fernandes PAR, Ferreira SS, Bastos R, Ferreira I, Cruz MT, Pinto A, Coelho E, Passos CP, Coimbra MA, Cardoso SM, Wessel DF (2019) Apple pomace extract as a sustainable food ingredient. Antioxidants 8. https://doi.org/10.3390/antiox8060189

  26. Fouda AS, Megahed HE, Fouad N, Elbahrawi NM (2016) Corrosion inhibition of carbon steel in 1 M hydrochloric acid solution by aqueous extract of Thevetia peruviana. J Bio- Tribo-Corros 2:1–13. https://doi.org/10.1007/s40735-016-0046-z

    Article  Google Scholar 

  27. Frear C, Zhao B, Fu G, Richardson M, Chen S, Fuchs MR (2005) Biomass inventory and bioenergy assessment (final report no. 05–07–047). Washington State Department of Ecology. Olympia

  28. Gabrielli C, Huet F, Keddam M, Oltra R (1990) A review of the probabilistic aspects of localized corrosion. CORROSION 46:266–278. https://doi.org/10.5006/1.3585102

    Article  Google Scholar 

  29. Honarvar Nazari M, Havens EA, Muthumani A, Shi X (2019) Effects of processed agro-residues on the performance of sodium chloride brine anti-Icer. ACS Sustain Chem Eng 7:13655–13667. https://doi.org/10.1021/acssuschemeng.8b06043

    Article  Google Scholar 

  30. Honarvar Nazari M, Laura F, Jungwirth S, Shi X (2015) Water quality implications and the toxicological effects of chloride-based deicers. Environ Sustain Transp Infrastruct, proceedings:272–292. https://doi.org/10.1061/9780784479285.022

  31. Honarvar Nazari M, Xianming S (2019) Developing renewable agro-based anti-Icers for sustainable winter road maintenance operations. J Mater Civ Eng 31:04019299. https://doi.org/10.1061/(ASCE)MT.1943-5533.0002963

    Article  Google Scholar 

  32. Hong T, Shi H, Wang H, Gopal M, Jepson WP (2000) EIS study of corrosion product film in pipelines. CORROSION2000. Presented at the CORROSION2000, NACE International, Orlando, p 16

    Google Scholar 

  33. Hsu C, Nazari MH, Li Q, Shi X (2019) Enhancing degradation and corrosion resistance of AZ31 magnesium alloy through hydrophobic coating. Mater Chem Phys. https://doi.org/10.1016/j.matchemphys.2018.12.106

  34. Hunter T (2012) Why nature chose phosphate to modify proteins. Philos Trans R Soc B Biol Sci 367:2513–2516. https://doi.org/10.1098/rstb.2012.0013

    Article  Google Scholar 

  35. Husain A, Kupwade-Patil K, Al-Aibani AF, Abdulsalam MF (2017) In situ electrochemical impedance characterization of cement paste with volcanic ash to examine early stage of hydration. Construct Build Mater 133:107–117. https://doi.org/10.1016/j.conbuildmat.2016.12.054

    Article  Google Scholar 

  36. Jamil HE, Shiri A, Boulif R, Bastos C, Montemor MF, Ferreira MGS (2004) Electrochemical behaviour of amino alcohol-based inhibitors used to control corrosion of reinforcing steel. Electrochim Acta 49:2753–2760. https://doi.org/10.1016/j.electacta.2004.01.041

    Article  Google Scholar 

  37. Jianlin W, Hui Y, Xianming S (2016) Effectiveness of products in managing metallic corrosion induced by cyclic deicer exposure: laboratory study using multielectrode Array sensors, electrochemical impedance, and laser Profilometer. J Mater Civ Eng 28:04015186. https://doi.org/10.1061/(ASCE)MT.1943-5533.0001449

    Article  Google Scholar 

  38. Kamal C, Sethuraman MG (2012) Caulerpin—a bis-indole alkaloid as a green inhibitor for the corrosion of mild steel in 1 M HCl solution from the marine alga Caulerpa racemosa. Ind Eng Chem Res 51:10399–10407. https://doi.org/10.1021/ie3010379

    Article  Google Scholar 

  39. Kim Y-S, Kim J-G (2017) Corrosion behavior of pipeline carbon steel under different iron oxide deposits in the district heating system. Metals 7:182. https://doi.org/10.3390/met7050182

    Article  Google Scholar 

  40. Liu Y, Zou C, Yan X, Xiao R, Wang T, Li M (2015) β-Cyclodextrin modified natural chitosan as a green inhibitor for carbon steel in acid solutions. Ind Eng Chem Res 54:5664–5672. https://doi.org/10.1021/acs.iecr.5b00930

    Article  Google Scholar 

  41. Lobo MG, Dorta E (2019) Chapter 19 - utilization and Management of Horticultural Waste. In: Yahia EM (ed) Postharvest Technology of Perishable Horticultural Commodities. Woodhead publishing, pp 639–666. https://doi.org/10.1016/B978-0-12-813276-0.00019-5

  42. Luo J, Xu Y, Fan Y (2019) Upgrading pectin production from apple pomace by acetic acid extraction. Appl Biochem Biotechnol 187:1300–1311. https://doi.org/10.1007/s12010-018-2893-1

    Article  Google Scholar 

  43. Ma F-Y (2012) Corrosive effects of chlorides on metals. Pitting Corros. https://doi.org/10.5772/32333

  44. Ma H, Cheng X, Li G, Chen S, Quan Z, Zhao S, Niu L (2000) The influence of hydrogen sulfide on corrosion of iron under different conditions. Corros Sci 42:1669–1683. https://doi.org/10.1016/S0010-938X(00)00003-2

    Article  Google Scholar 

  45. Markin VS, Volkova-Gugeshashvili MI, Volkov AG (2006) Adsorption at liquid interfaces: the generalized Langmuir isotherm and interfacial structure. J Phys Chem B 110:11415–11420. https://doi.org/10.1021/jp061818v

    Article  Google Scholar 

  46. McIntyre NS, Zetaruk DG (1977) X-ray photoelectron spectroscopic studies of iron oxides. Anal Chem 49:1521–1529. https://doi.org/10.1021/ac50019a016

    Article  Google Scholar 

  47. Mei L, Liao L, Wang Z, Xu C (2015) Interactions between phosphoric/tannic acid and different forms of FeOOH. Adv Mater Sci Eng 10. https://doi.org/10.1155/2015/250836

  48. Musa AY, Kadhum AAH, Mohamad AB, Rahoma AAB, Mesmari H (2010) Electrochemical and quantum chemical calculations on 4,4-dimethyloxazolidine-2-thione as inhibitor for mild steel corrosion in hydrochloric acid. J Mol Struct 969:233–237. https://doi.org/10.1016/j.molstruc.2010.02.051

    Article  Google Scholar 

  49. Nazari MH, Allahkaram SR (2010) The effect of acetic acid on the CO2 corrosion of grade X70 steel. Mater Des 31:4290–4295. https://doi.org/10.1016/j.matdes.2010.04.002

    Article  Google Scholar 

  50. Nazari MH, Shi X (2018) Vehicle risks of winter road operations and best management practices. In: Sustainable winter road operations. Wiley-Blackwell, pp 241–272. https://doi.org/10.1002/9781119185161.ch12

  51. Nazari MH, Shihab MS, Cao L, Havens EA, Shi X (2017) A peony-leaves-derived liquid corrosion inhibitor: protecting carbon steel from NaCl. Green Chem Lett Rev 10:359–379. https://doi.org/10.1080/17518253.2017.1388446

    Article  Google Scholar 

  52. Negm NA, Elkholy YM, Zahran MK, Tawfik SM (2010) Corrosion inhibition efficiency and surface activity of benzothiazol-3-ium cationic Schiff base derivatives in hydrochloric acid. Corros Sci 52:3523–3536. https://doi.org/10.1016/j.corsci.2010.07.001

    Article  Google Scholar 

  53. Nixon WA, Williams AD (2001) A guide for selecting anti-icing chemicals, Version 1.0 (IIHR Technical Report No. No. 420)

    Google Scholar 

  54. Nwabanne JT, Okafor VN (2012) Adsorption and thermodynamics study of the inhibition of corrosion of mild steel in H2SO4 medium using Vernonia amygdalina. J Miner Mater Charact Eng 11:885–890

    Google Scholar 

  55. Obot IB, Gasem ZM (2014) Theoretical evaluation of corrosion inhibition performance of some pyrazine derivatives. Corros Sci 83:359–366. https://doi.org/10.1016/j.corsci.2014.03.008

    Article  Google Scholar 

  56. Obot IB, Obi-Egbedi NO (2010) Theoretical study of benzimidazole and its derivatives and their potential activity as corrosion inhibitors. Corros Sci 52:657–660. https://doi.org/10.1016/j.corsci.2009.10.017

    Article  Google Scholar 

  57. Oguzie EE (2007) Corrosion inhibition of aluminium in acidic and alkaline media by Sansevieria trifasciata extract. Corros Sci 49:1527–1539. https://doi.org/10.1016/j.corsci.2006.08.009

    Article  Google Scholar 

  58. Osman MM, Omar AMA, Al-Sabagh AM (1997) Corrosion inhibition of benzyl triethanol ammonium chloride and its ethoxylate on steel in sulphuric acid solution. Mater Chem Phys 50:271–274. https://doi.org/10.1016/S0254-0584(97)01941-X

    Article  Google Scholar 

  59. Pandarinathan V, Lepková K, Bailey SI, Becker T, Gubner R (2014) Adsorption of corrosion inhibitor 1-Dodecylpyridinium chloride on carbon steel studied by in situ AFM and electrochemical methods. Ind Eng Chem Res 53:5858–5865. https://doi.org/10.1021/ie402784y

    Article  Google Scholar 

  60. Paparazzo E (1987) XPS and auger spectroscopy studies on mixtures of the oxides SiO2, Al2O3, Fe2O3 and Cr2O3. J Electron Spectrosc Relat Phenom 43:97–112. https://doi.org/10.1016/0368-2048(87)80022-1

    Article  Google Scholar 

  61. Pelavin M, Hendrickson DN, Hollander JM, Jolly WL (1970) Phosphorus 2p electron binding energies. Correlation with extended Hueckel charges. J Phys Chem 74:1116–1121. https://doi.org/10.1021/j100700a027

    Article  Google Scholar 

  62. Pradityana A, Sulistijono SA, Chyntara S (2014) Eco-friendly green inhibitor of mild steel in 3,5% NaCl solution by Sarang Semut (Myrmecodia Pendans) extract. AIP Conf Proc 1617:161–164. https://doi.org/10.1063/1.4897128

    Article  Google Scholar 

  63. Pradityana A, Sulistijono SA, Noerochim L, Susanti D (2016) Inhibition of corrosion of carbon steel in 3.5% NaCl solution by Myrmecodia Pendans extract. Int J Corros 6. https://doi.org/10.1155/2016/6058286

  64. Rani BEA, Basu BBJ (2012) Green inhibitors for corrosion protection of metals and alloys: an overview [WWW document]. Int J Corros. https://doi.org/10.1155/2012/380217

  65. Recloux I, Andreatta F, Druart M-E, Coelho LB, Cepek C, Cossement D, Fedrizzi L, Olivier M-G (2018) Stability of benzotriazole-based films against AA2024 aluminium alloy corrosion process in neutral chloride electrolyte. J Alloys Compd 735:2512–2522. https://doi.org/10.1016/j.jallcom.2017.11.346

    Article  Google Scholar 

  66. Ryl J, Brodowski M, Kowalski M, Lipinska W, Niedzialkowski P, Wysocka J (2019) Corrosion inhibition mechanism and efficiency differentiation of Dihydroxybenzene isomers towards aluminum alloy 5754 in alkaline media. Materials 12. https://doi.org/10.3390/ma12193067

  67. Sahu O, Singh N (2019) 13 - significance of bioadsorption process on textile industry wastewater. In: Shahid-ul-Islam BBS (ed) The impact and prospects of green chemistry for textile technology, The textile institute book series. Woodhead publishing, pp 367–416. https://doi.org/10.1016/B978-0-08-102491-1.00013-7

  68. Shah AM, Rahim AA, Hamid SA, Yahya S (2013) Green inhibitors for copper corrosion by mangrove tannin. Int J Electrochem Sci 8:2140–2153

    Google Scholar 

  69. Shi X, Fay L, Yang Z, Nguyen TA, Liu Y (2009) Corrosion of deicers to metals in transportation infrastructure: introduction and recent developments. Corrosion Rev 27:23–52. https://doi.org/10.1515/CORRREV.2009.27.1-2.23

    Article  Google Scholar 

  70. Shihab MS, Mahmood AF (2017) Experimental and theoretical study of some Npyridinium salt derivatives as corrosion inhibitors for mild-steel in 1 M H2SO4. Port. Electrochim Acta 35:39–51. https://doi.org/10.4152/pea.201701039

    Article  Google Scholar 

  71. Shihab MS, Nazari MH, Fay L (2016) Study of inhibition effect of pyridinium salt derivative on corrosion of C1010 carbon steel in saline solution. Prot Met Phys Chem Surf 52:714–720. https://doi.org/10.1134/S2070205116040213

    Article  Google Scholar 

  72. Sin HLY, Abdul Rahim A, Gan CY, Saad B, Salleh MI, Umeda M (2017) Aquilaria subintergra leaves extracts as sustainable mild steel corrosion inhibitors in HCl. Measurement 109:334–345. https://doi.org/10.1016/j.measurement.2017.05.045

    Article  Google Scholar 

  73. Somasundaran P (2004) Encyclopedia of surface and colloid science, 2004 update supplement. CRC Press. Boca Raton, FL, USA

  74. Stewart JJP (1989) Optimization of parameters for semiempirical methods I. Method J Comput Chem 10:209–220. https://doi.org/10.1002/jcc.540100208

    Article  Google Scholar 

  75. Suarez R, Gonzalez-Rodriguez J, Dominguez-Patino G, Martinez-Villafane A (2014) Use of Opuntia ficus extract as a corrosion inhibitor for carbon steel in acidic media. Anti-Corros Methods Mater 61. https://doi.org/10.1108/ACMM-01-2013-1238

  76. Suedile F, Robert F, Roos C, Lebrini M (2014) Corrosion inhibition of zinc by Mansoa alliacea plant extract in sodium chloride media: extraction, characterization and electrochemical studies. Electrochim Acta 133:631–638. https://doi.org/10.1016/j.electacta.2013.12.070

    Article  Google Scholar 

  77. Surendranath Y, Kanan MW, Nocera DG (2010) Mechanistic studies of the oxygen evolution reaction by a cobalt-phosphate catalyst at neutral pH. J Am Chem Soc 132:16501–16509. https://doi.org/10.1021/ja106102b

    Article  Google Scholar 

  78. Tao Z, He W, Wang S, Zhou G (2013) Electrochemical study of Cyproconazole as a novel corrosion inhibitor for copper in acidic solution. Ind Eng Chem Res 52:17891–17899. https://doi.org/10.1021/ie402693d

    Article  Google Scholar 

  79. Umoren SA, Obot IB, Madhankumar A, Gasem ZM (2015) Performance evaluation of pectin as ecofriendly corrosion inhibitor for X60 pipeline steel in acid medium: experimental and theoretical approaches. Carbohydr Polym 124:280–291. https://doi.org/10.1016/j.carbpol.2015.02.036

    Article  Google Scholar 

  80. Verma C, Quraishi MA, Kluza K, Makowska-Janusik M, Olasunkanmi LO, Ebenso EE (2017) Corrosion inhibition of mild steel in 1M HCl by D-glucose derivatives of dihydropyrido [2,3-d:6,5-d′] dipyrimidine-2, 4, 6, 8(1H,3H, 5H,7H)-tetraone. Sci Rep 7:44432. https://doi.org/10.1038/srep44432

    Article  Google Scholar 

  81. Wang X, Chen Q, Lü X (2014) Pectin extracted from apple pomace and citrus peel by subcritical water. Food Hydrocoll 38:129–137. https://doi.org/10.1016/j.foodhyd.2013.12.003

    Article  Google Scholar 

  82. Yoshida T, Sawada S (1974) X-ray photoelectron spectroscopy of EDTA. Bull Chem Soc Jpn 47:50–53. https://doi.org/10.1246/bcsj.47.50

    Article  Google Scholar 

  83. Zhang Y, Tian J, Zhong J, Shi X (2018) Thin nacre-biomimetic coating with super-anticorrosion performance. ACS Nano 12:10189–10200. https://doi.org/10.1021/acsnano.8b05183

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to acknowledge Dr. Girish Ganjyal (School of Food Science, Washington State University) for providing the apple pomace powder.

Funding

The authors thankfully acknowledge financial support by the USDOT (Tier 1 UTC CESTiCC) and the NSF (Award No. 1638384).

Author information

Authors and Affiliations

Authors

Contributions

Dr. M. Honarvar Nazari did final interpretation of the results of the electrochemical corrosion tests, conducted some of the laboratory tests and wrote the draft manuscript. Dr. M. S. Shihab conducted the quantum chemical calculations, provided the quantum chemical figures, and wrote an initial draft of the related section. E. A. Havens conducted most of the electrochemical tests and did an initial interpretation on the related data. Dr. X. Shi designed the scope of this work and advised the team throughout the process (from conceptualization to manuscript preparation). The author(s) read and approved the final manuscript.

Corresponding author

Correspondence to Xianming Shi.

Ethics declarations

Competing interests

The authors declare no competing interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Honarvar Nazari, M., Shihab, M.S., Havens, E.A. et al. Mechanism of corrosion protection in chloride solution by an apple-based green inhibitor: experimental and theoretical studies. J Infrastruct Preserv Resil 1, 7 (2020). https://doi.org/10.1186/s43065-020-00007-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s43065-020-00007-w

Keywords