Next Article in Journal
Optical Biosensors Based on Silicon-On-Insulator Ring Resonators: A Review
Previous Article in Journal
Deltamethrin Residues in Milk and Cheese of Lactating Goats (Capra hircus)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Inhibition of Q235 Carbon Steel by Calcium Lignosulfonate and Sodium Molybdate in Carbonated Concrete Pore Solution

Beijing Key Laboratory of Electrochemical Process and Technology for Materials, Beijing University of Chemical Technology, Beijing 100029, China
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(3), 518; https://doi.org/10.3390/molecules24030518
Submission received: 4 January 2019 / Revised: 27 January 2019 / Accepted: 30 January 2019 / Published: 31 January 2019
(This article belongs to the Section Electrochemistry)

Abstract

:
The inhibition effect and mechanism of a compound calcium lignosulfonate (CLS) and sodium molybdate inhibitors for Q235 carbon steel in simulated carbonated concrete pore solution (pH 11.5) with 0.02 mol/L NaCl are studied using electrochemical and surface analysis techniques. The results show that in carbonated simulated concrete pore (SCP) solution CLS and Na2MoO4 show a synergistic inhibition effect. The compound inhibitor can be defined as mix-type inhibitor. With 400 ppm CLS plus 600 ppm Na2MoO4, the pitting potential moves positively about 200 mV, and the inhibition efficiency reaches 92.67%. After 24 h immersion, the IE% further increases up to 99.2%. The surface analysis results show that Na2MoO4 could promote stability of the passive film, and the insoluble molybdenum compounds and CaO/Ca(OH)2, together with adsorbed CLS, deposit on the steel surface, forming a complex film. The compounded film effectively inhibits corrosion of the steel.

1. Introduction

Reinforced concrete is the most widely used construction material [1,2,3], while corrosion of reinforcing steel is the main reason leading to the premature deterioration of reinforced concretes [4,5]. Corrosion of reinforcing steel is related to the carbonation of concrete [2,3,6] and the presence of chloride ions [7,8]. The carbonation of concrete is the result of CO2 diffusion and reaction with the hydrated product (Ca(OH)2), which leads to the composition change of the concrete pore solution and the decrease of pH value. Several techniques have been employed to reduce corrosion of reinforcing steel, among which corrosion inhibitor is the most practical method owing to its advantages of low cost, high efficiency, and wide applicability [1,2,3,9,10,11]. For example, Ca(NO2)2 is an inorganic corrosion inhibitor commonly used in reinforced concrete construction [12,13,14]. NO2 could help to form a uniform passive film on a steel surface [13,14] and prevent the steel from corroding. However, most of the commercial inhibitors remain unfriendly to the environment. In recent years, researchers have paid considerable attention towards the development of various more effective and greener inhibitors to prevent the corrosion of rebars [15,16].
Lignosulfonates (LS) are the by-products of the pulping waste liquor in acid sulfite pulp mills [17]. One of the most common applications of lignosulfonates is as dispersing agents and set-retarding agents in concretes [18]. As corrosion inhibitors, lignin-based compounds contain plenty of hydroxyl, carboxyl, or methoxyl groups, which can be adsorbed onto a metal surface by sharing their lone pair electron or π-electron with the free d-orbital of metal. In acidic solution, the inhibition efficiency of lignin and its modifications for steel depend on the number of carboxy groups in the macromolecule [19], the inhibitor concentration, and the environmental factors [20]. In neutral aqueous solutions with Cl, lignin monomers acted as mixed type inhibitors on steel corrosion [21] and the inhibitive performance of lignin derivatives depended on their functional groups [22]. Bishop et al. [23] have confirmed that the –SO3 and –OH groups present in LS can be adsorbed on a steel surface and generate an impermeable structure. Li et al. [24] revealed that LS adsorbed in the form of aggregates and the adsorption isotherm follows the Freundlich model at low pH values. In neutral solutions, LS increasingly adsorbed on steel in the form of well separated single molecules, which can be well fitted using the Langmuir equation. However, there is little research involving calcium lignosulfonate (CLS) as a corrosion inhibitor in alkaline solution. Our work [25] showed that CLS is a good corrosion inhibitor in saturated Ca(OH)2 solutions with Cl for carbon steel. The adsorption film on a steel surface is formed by both physisorption and chemisorption. Particularly, CLS could be preferentially adsorbed on the active sites on a steel surface using the sinapyl alcohol group, which is beneficial to the inhibition of pitting corrosion.
Molybdate is an environment-friendly anodic passivation inhibitor [26,27]. The inhibition mechanism of molybdates has been studied by many scholars, and competitive adsorption [28], adsorption film [27,29], and oxidation [30] are the dominant views. In order to enhance the inhibition effective of molybdates, some authors [31,32,33] tried to associate Na2MoO4 with organic inhibitors. It was reported [33] that the synergistic effect is present between molybdate and benzotriazole (BTA). The compounded inhibitor could promote the transformation of FeOOH and Fe2O3 in the passive film, and both passivation and pitting corrosion resistance were promoted.
On the other hand, most studies focusing on inhibitors for reinforcing steel were performed with high pH concrete pore solutions. The carbonation of concretes would directly affect the passivation film on reinforcing steel surfaces, but few studies were carried out in carbonated concrete environments. In the present work, the combined inhibition effect of CLS and molybdate on Q235 carbon steel in simulated carbonated concrete pore solution is studied with electrochemical measurements and surface analysis methods. Particularly, the synergistic effect between Na2MoO4 and CLS is paid attention to.

2. Results and Discussion

2.1. Inhibition Evaluation

2.1.1. Potentiodynamic Polarization Measurement

Figure 1a shows the polarization curves of Q235 carbon steel after immersion in test solutions without and with various concentrations of CLS. The polarization curves show typical passivation-pitting corrosion behavior. Figure 1b shows the electrochemical parameters obtained from the polarization curves.
The corrosion potential (Ecorr) of carbon steel in the test solution is about −290 mVSCE, and almost has almost no change as CLS concentration increases, indicating CLS acts as a mixed-type inhibitor [34,35]. The inhibition efficiency (IE%) increases as the CLS concentration increases, which is consistent with the results of other authors [17,25]. As an organic inhibitor, CLS could form an adsorption film on a steel surface [25], which has a barrier effect to prevent general corrosion. The pitting potential (Eb) is the potential at which the anodic current increases rapidly [36], and the localized corrosion starts above this potential. As the CLS concentration increases from 100 to 800 ppm, the Eb value increases from -146 mVSCE to 57 mVSCE, and then slightly decreases. This result indicates that the pitting corrosion susceptibility of carbon steel decreases as the inhibitor concentration increases. The max current density (ipeak), which is related to the growth rate of the corroded pits [37], remains almost unchanged as the CLS concentration increases. The repassivation potential (Epp) is the intersection potential of the forward and reverse scans, and the steel potential must be above Epp for existing areas of localized corrosion to propagate [38]. As the CLS concentration increases the Epp shows relatively smaller shifts to positive, compared with the obvious shifts of the Eb values. The results of imax and Epp reveal that, as a corrosion inhibitor, CLS has little effect on the repassivation process of carbon steel. Compared with the inhibition effect of CLS in pH 12.5 concrete pore solution [25], the inhibition ability of CLS in carbonated SCP solution decreases for both general corrosion and localized corrosion. Therefore, the combination of CLS with other inhibitors to improve the inhibition effect is necessary.
Figure 2 shows the polarization curves and the electrochemical parameters of Q235 steel in test solution with various concentrations of Na2MoO4. The Ecorr slightly increases with the increase of MoO42− concentration, indicating that Na2MoO4 acts as a mixed-type inhibitor [34,35] predominantly with anodic effectiveness [39]. Na2MoO4 is a moderate inhibitor for general corrosion, and the IE% increases as the Na2MoO4 concentration increases. The passive current density decreases obviously with the increase of the inhibitor concentration, indicating Na2MoO4 promoted passivation of the steel. Fe2(MoO4)3 complex is insoluble and protective in neutral and alkaline media [29], which could enhance the protection film formed on steel surfaces [39]. Refaey et al. [40] reported that the formation of a protective film played a critical role on the inhibition effect of molybdate. The increased Eb and passivation region (Eb-Ecorr) due to the increase of MoO42− concentration might be associated with the effect of molybdate reducing the number and magnitude of metastable pitting transients [39,41]. In other words, molybdate ions could block the active sites on the surface [42] and affect the nucleation of pits by deactivating or reducing the number of sites, leading to decreased pitting corrosion susceptibility [43]. As shown in Figure 3, on the reverse scan, a potential step, the pit transition potential (Eptp) [39,44,45], is detected for steel in the test solution with molybdate. The Eptp is a characteristic potential that is correlated with repassivation at the pit bottom [44,45], which might lead to the concentration gradients for mass transport and promote further pit nucleation. In the pit environment, a series of chemical reactions involving hydrolysis and polymerization of molybdates may occur as the pH value decreases [29,46]:
7MoO42−+8H+→Mo7O246−+4H2O
Mo7O24 has a chelate effect with iron(III) to form complexes, which could help repassivation of the pit. In Figure 2b, the Eptp potential increases as the molybdate concentration increases, which means that molybdate with relatively higher concentration could help to form a repassivation film at the pit bottom.
The mechanism of molybdate inhibition in carbonation SCP solution could be inferred. In the first step, the MoO42− ions competitively adsorb on the steel surface with Cl ions [47,48]. Then the passivation film could be enhanced by the adsorbed molybdate ions [49] and a precipitation film composed of oxidized molybdenum forms on the steel surface. The composite film could increase both the general corrosion and pitting corrosion resistances. However, once the pits occur, MoO42− cannot stop the pit growth and self-catalyzed corrosion occurs inside the pits which accelerates the growth of pits.
Figure 3 shows the cyclic potentiodynamic polarization (CPP) curves of Q235 carbon steel in carbonated SCP solution with different ratios of CLS and Na2MoO4 (total 1000 ppm), and the electrochemical parameters are shown in Table 1. The Ecorr slightly increases as the CLS ratio increases, which is the same as the result of molybdate compound with glycol [47]. The compound inhibitor acts as a mix-type inhibitor with predominantly anodic effectiveness [47]. The IE% slightly decreases as the CLS ratio increases. The synergistic parameter (S), which reveals the interaction relationship between CLS and Na2MoO4, is calculated using the following equation [50,51]:
S = 1 IE % ( 1 + 2 ) 1 IE % c
where IE%(1 + 2) = (IE%1 + IE%2) − (IE%1 × IE%2), and IE%1, IE%2, and IE%c are the inhibition efficiencies for CLS, Na2MoO4, and CLS/Na2MoO4 compound, respectively. The calculated S values are all above 1, indicating the synergistic behavior of the selected inhibitor combination [50], and the S value increases as the CLS concentration decreases, which means the synergistic effect between the two inhibitors decreases. The obtained Eb potentials for each compound inhibitor ratios are relatively close and obviously higher than the value without inhibitor. The compound with 400 ppm CLS and 600 ppm Na2MoO4 shows the highest Eb. The Epp significantly increases as the CLS ratio decreases, which means that a decrease of the CLS ratio in the compound inhibitor could promote the repassivation of carbon steel, while adding CLS or molybdate alone does not show this promoting effect. The difference between Eb and Epp represents the repassivation tendency of pits on steel surface, which decrease as the molybdate concentration increases. This result suggests that the high ratio of molybdate in the compound inhibitor would be beneficial to inhibit localized corrosion. It has been reported that pitting corrosion could not be inhibited and might even be promoted if the molybdate concentration is too low [52,53]. This might be attributed to the pH changes in the pit environment in which the pH cannot be raised enough to passivate the pit [27,52]. The above results show that the compound inhibitor effectively increases both the general corrosion resistance and the pitting corrosion resistance, and the optimum ratio is 400 ppm CLS compound with 600 ppm Na2MoO4.
The pre-filming time has an important role in the inhibition processes [54,55]. CLS could form an adsorption film on a steel surface after 10 h pre-filming in concrete pore solution, which could reduce the content of Cl adsorbed on the steel surface and prevent the local enrichment of Cl [25]. As shown in Figure 4a, for the samples immersed in solution without inhibitor, with extended immersion time, Ecorr decreased and icorr increased obviously, and the passivation behavior disappeared. While for the samples with compound inhibitor, shown in Figure 4b, the polarization curves show typical passivation-pitting behavior. As the pre-filming time increases the Ecorr increases, the icorr decreases, and the IE% (calculated by icorr) increases from 92% to 99%. This result reveals that the compound inhibitor could enhance the general corrosion resistance after pre-filming time. Eb for the inhibited steels obviously increases as the immersion time increases, while the Epp shows almost no shift, which suggests that in the carbonated solution the inhibitor could help Q235 steel maintain and improve the passivation state as the pre-filming time increases, especially for the first 16 h. The results confirm the good inhibition effect of the compound inhibitor after pre-filming time on both general corrosion and localized corrosion.

2.1.2. Electrochemical Impedance Spectroscopy Measurements

To investigate the inhibitive behavior of CLS, Na2MoO4, and the compound inhibitor, a series of EIS tests were performed in carbonated SCP solution with various concentrations of different inhibitors. As shown in Figure 5, a single depressed capacitive semicircle is presented in all the Nyquist plots, and the capacitive semicircle radius increases as the inhibitor concentration increases. This result reveals that the addition of inhibitor did not change the mechanism of the corrosion process but inhibited corrosion by forming protective film on the steel surface [56]. The inhibitive system could be described using the model of the solution/steel interface, as shown in Figure 5d. In the equivalent circuit, Rs is the solution resistance and Rct represents the charge-transfer resistance corresponding to the corrosion reaction at the metal substrate/solution interface. Rfilm and Cfilm represent the resistance and the capacitance of the protection film, respectively. The constant phase angle elements (CPE) is used in the model to compensate for the non-homogeneous electrode surface. The fitted parameters are shown in Figure 6.
Figure 5a shows the Nyquist plots of samples immersed in test solution with various concentrations of CLS. As the CLS concentration increases the diameter of the capacitive loop increases obviously. This phenomenon might be due to the increase of surface coverage by the adsorbed inhibitor film [56]. From the parameters in Figure 6, the Rct increases and the Rfilm shows almost no shift as the CLS concentration increases, which suggests that the corrosion reaction was inhibited by the adsorption CLS film on the steel surface. The Cdl decreases as the CLS concentration increases, indicating that the status of the steel/solution interface has been changed by the adsorption of CLS. Furthermore, the Cfilm decreases with increased CLS ratio in the compound inhibitor. According to the Helmholtz model, the capacitance is inversely proportional to the surface changes [50,56]:
C = ε0εS/d
where d is the film thickness, S is the surface area of working electrode, ε0 is the permittivity of air and ε the local dielectric constant [56,57,58]. The decrease of Cdl and Cfilm might be attributed to the replacement of the adsorbed water molecules at the metal surface by CLS molecules, which have a lower dielectric constant. The adsorbed CSL film and the passive film could form a double-layer barrier on the steel surface [25]. As the CLS concentration increased the adsorbed CLS molecules increased, resulting in a dense and homogeneous inhibition film on the steel surface.
From Figure 5b and Figure 6, it can be seen that the Cdl decreases and the Rct increases as the Na2MoO4 concentration increases, which means molybdate could increase the corrosion resistance to decrease the corrosion rate. As the molybdate concentration increases, the Cfilm value has almost no shift, and the Rfilm increases significantly. The EIS results confirm that a barrier layer, depending on the concentration of molybdate, blocked the adsorption of Cl [29]. In addition, molybdate could enhance the passivation film to protect the steel. Gong et al. [59] reported that molybdate enhances the self-healing ability of passive films, thereby leading to improvement of corrosion resistance.
After adding compound inhibitor, the Nyquist plots are shown in Figure 5c and the parameters are displayed in Figure 6. As the CLS ratio increases, the Cdl decreases slightly and the Rct almost has no change, again indicating that the compound inhibitor could increase the corrosion resistance. The Cfilm value significantly decreases and the Rfilm value decreases slightly. In the compound inhibitor, the barrier effect of the protective film increases as the CLS ratio increases, and the passive film is enhanced to prevent corrosion as the molybdate ratio increases. The results revealed the compound inhibitor has a better inhibition effect for carbon steel in test solution.
In order to confirm the effect of pre-filming time on inhibition of the compound inhibitor, EIS measurements for Q235 carbon steel in the test solution with compound inhibitor for various immersion times were carried out and the results are shown in Figure 7. As the immersion time increases the diameter of the capacitive loop increases, which means as the pre-filming time increases the inhibition effect increases, which might be due to the increase in the surface coverage [56] or in the thickness of the protection film. From the EIS parameters shown in Figure 7b, the Cfilm decreases obviously and the Rfilm slightly increases as the immersion time increases, which also suggests the formation of adsorption film on the steel surface. In addition, the decreased Cdl and increased Rct indicate that the corrosion reactions have been suppressed by the inhibitor. The above results confirm the results of the polarization measurements and suggest that after 16 h to 24 h pre-filming the compound inhibitor would show better inhibition effect.

2.2. Surface and Isotherm Adsorption Studies

2.2.1. X-ray Photoelectron Spectroscopy Measurement and Contact Angle

In order to determine the chemical composition of the inhibited surface, XPS analyses were used to characterize the Q235 carbon steel after 1 h immersion in carbonated SCP solutions with 800 ppm CLS, 800 ppm Na2MoO4, and 400 ppm CLS, plus 600 ppm Na2MoO4, respectively. The obtained high-resolution peaks for Ca 2p, C 1s, O 1s, S 1s, Fe 2p, and Mo 3d core levels are fitted and shown in Figure 8, Figure 9 and Figure 10.
In Figure 8a, the C 1s spectra for CLS and compound inhibitor treated steel surface shows four peaks. The peak at 284.8 eV and 287.9 eV can be attributed to the C-C/C-H bonds [35,44,60] and the C-O bonds [50,61] in CLS molecules, respectively. The peak at 286.4 eV may be assigned to the carbon atom bonded to S in C-S bonds [60]. This peak on the compound inhibitor treated surface shifts negatively to 286 eV, suggesting that the adsorbed CLS may accept the feedback electrons from Fe atoms on the surface. The last peak at 292.2 eV can be attributed to the benzene ring in the organic inhibitor [60,61]. This peak shifts positively to 293.4 eV on the compound inhibitor treated surface, indicating the aromatic ring of the CLS supplied electrons to Fe atoms [62]. Figure 9 presents the Ca 2p and S 2p spectra for the CLS and compound inhibitor treated sample. Ca and S are the characteristic elements in CLS, which could confirm the adsorption of CLS. The Ca2p spectrum is decomposed into two peaks at 351 eV and 347.4 eV, which are attributed to CaO/Ca(OH)2 [25,63] and Ca-O-S bonds [25] on the steel surface, respectively. This result suggests the CaO/Ca(OH)2 has been precipitated on the steel surface and could enhance the adsorption of CLS by forming Ca-O-S bonds [25] to form a more complete film on the surface. For S 2p, there are two peaks. The first one is the peak at 168.4 eV [25,50] attributed to Ca-O-S, and the co-adsorption of CLS and Ca could protect the surface from corrosion [25,60,64]. On the compound inhibitor treated surface, this peak is widened, which might be due to the M-S-O bonds, where M could be Ca2+, Fe2+, or Fe3+. The area ratio of M-O-S on the compound inhibitor treated sample is much larger than that on the CLS treated sample, which means molybdate could enhance the adsorption of LS2−. The other peak at 167 eV [25,50] can be attributed to S in the organic inhibitor. The above XPS results confirm the adsorption of CLS and reveal the precipitation of oxides or hydroxides of Ca and Mo, which could help to enhance the adsorption of the organic inhibitor.
As shown in Figure 9, the high-resolution Mo 3d spectra are detected on both the surface treated using Na2MoO4 and the compound inhibitor. The two peaks at 231.9 eV is attributed to Mo4+, such as MoO2 [62], and the peak at 235.1 eV is the peak of Mo6+ [33], which could confirm the precipitation of MoOx and adsorption of molybdate on the surface. The passive film formed by the reduction of molybdate mainly consists of iron oxide and molybdenum oxide. The reduction of Mo6+ could help to form a more stable film [47,65].
The O1s peaks for Q235 carbon steel immersed in test solutions with different inhibitors are shown in Figure 10a. The first peak at 530.1 eV corresponds to O2−, which could be attributed to the bond with Fe3+ in Fe2O3, Fe3O4, or MoO42− [33,35]. The second peak at 531.6 eV can be attributed to the metal oxides and hydroxides or H2O species on the steel surface [35,66]. The third peak may be assigned to the oxygen atoms of the O-C bond in the organic compounds from surface contamination and from the adsorbed water at 532.7 eV [60,62]. The deconvolution of the high-resolution Fe 2p3/2 spectra of the steel surface treated using different inhibitors show several peaks, which are shown in Figure 10b. The first component at 706.8 eV is attributed to metallic iron (Fe0) [35]. The peaks at 708.3 eV and 710.4 eV can be attributed to the Fe2+ present in FeO [35,39] and Fe3O4 [60], respectively. The peak at 711.5 eV is assigned to Fe3+, which can be associated to the ferric oxide/hydroxide species such as Fe2O3, Fe3O4, and FeOOH [50,60]. The peak at 714.2 eV is also assigned to Fe3+, which might be related to adsorption of the organic inhibitor [35,67]. In order to understand the effect of immersion time on the composition of passive film, XPS spectra of the samples immersed in the carbonated SCP solution with different inhibitors for different times were measured. The Fe3+/Fe2+ ratio could be used to evaluate the stability of the passive film on steel [29], as shown in Figure 11. As the immersion time increases, the Fe3+/Fe2+ ratio in the film obtained in solution without inhibitor decreases, indicating that the stability of the passive film decreases with time. For the steel treated using the CLS, the Fe3+/Fe2+ ratio slightly decreased in 30 h and then increased, which might be due to the adsorption of CLS replacing water molecules on the steel surface and reducing the passivation tendency. For the molybdate treated sample, the Fe3+/Fe2+ ratio increases as the immersion time increases, which confirms the previous result that molybdate could promote passivation of the surface. For the film formed in solution with the compound inhibitor, the Fe3+/Fe2+ ratio is the highest for different immersion times. This result also reveals that the combination of CLS and molybdate would further enhance the stability of the passive film.
The contact angle is an important parameter reflecting the wettability of materials [68,69]. The contact angles of Q235 steel surface treated using different inhibitors for 24 h are shown in Figure 12 and Table 2. The average contact angle for steel immersed in solution without inhibitor is 68.76°, which is consistent with references [68]. The inhibitor pretreatment induced significant changes in the contact angle. CLS causes the contact angle decrease obviously, which might be due to the fact that CLS is a surfactant and the adsorption film formed using CLS contains lots of methoxy, hydroxy, and other hydrophilic groups. The hydrophilic adsorption film could prevent the adsorption of aggressive anions and protect the steel surface [25]. Na2MoO4 could slightly increase the contact angle of the steel surface, which might be due to the Mo oxide precipitations decreasing the surface roughness [70]. The contact angle of the compound inhibitor treated steel is about 33.5°, which is between the values of the surface using single inhibitor treatment. The contact angle test results are consistent with the corrosion rate results and reveal the formation of the inhibition film on the steel surface.

2.2.2. Inhibition Mechanism of CLS and Na2MoO4 Compound Inhibitor

The inhibition mechanism for the compounded inhibitor could be inferred using the above results, as shown in Figure 13. CLS molecules dissociate instantly to lignosulfonate (LS2−) ions and calcium ions (Ca2+). The steel surface immersed in the SCP solution has extra positive charges, providing a certain driving force for the adsorption of negative (LS2−, MoO42−, Cl) ions [25,71]. The inhibitor ions (MoO42− and LS2−) would adsorb on the steel surface competitively against Cl to protect the passive film. The adsorbed molybdate may react with the iron matrix to form Fe2+ and MoO2, and the Fe2+ would be further oxidized to form Fe3+. Fe3+ could help to form more stable passivation film [29,39] or react with molybdate to form insoluble Fe2(MoO4)3. The deposited MoO2 on steel surface would block the adsorption of Cl and prevent the steel from corrosion. Ca2+ in alkaline solution can form Ca(OH)2 or CaO, which could precipitate on steel surfaces. Furthermore, the insoluble molybdenum compounds, together with CaO/Ca(OH)2, could precipitate on the passive film to form a complex film, further enhancing the protection of the film. The precipitated Ca or Mo compounds could also enhance the adsorption of CLS molecules by forming Metal-O-S bonds [25]. Finally, LS2− would adsorb on the outside film, forming a hydrophobic layer to protect the steel.

3. Materials and Methods

3.1. Materials and Test Solutions

The studied material was Q235 carbon steel with the following chemical composition (wt%): C 0.13, Si 0.29, Mn 0.76, S 0.11, Cu 0.08, and Fe balance. The sample size was 8 mm × 8 mm × 10 mm. The sample electrode was covered with epoxy resin, leaving a 0.4 cm2 area exposed to the test solution. The working surface for all samples was abraded with abrasive papers from 240# to 1000#, and cleaned with de-ionized water and ethanol.
The testing solution was composed of 0.0021 mol/L NaOH, 0.0042 mol/L KOH, and 0.02 mol/L NaCl, which was used to simulate the pore solution for a carbonated concrete environment [72]. The calcium lignosulfonate (CLS) and Na2MoO4 were analytical grade chemicals. The chemical structure of CLS is shown in Figure 14. The pH value of testing solutions was adjusted to 11.5, and all tests were carried out at room temperature (25 °C).

3.2. Electrochemical Measurements

Cyclic potentiodynamic polarization (CPP) curves and electrochemical impedance spectroscopy (EIS) were measured using CS350 electrochemical workstation (Corrtest Company, Wuhan, China). The polarization tests were started at a potential 300 mV below the open circuit potential (OCP) at a potential scanning rate of 0.1 mV/s in the anodic direction until the current density increased up to 0.02 mA/cm2, then the potential scanning was reversed at the same scanning rate until the test end. The corrosion potential (Ecorr) and corrosion current density (icorr) were obtained using Tafel plots. The inhibition efficiency (IE%) was calculated using Equation:
IE   ( % ) = i c o r r 0 i c o r r i c o r r 0 × 100
where i0corr and icorr are the corrosion current density values in the test solutions without and with the inhibitors, respectively. In addition, the pitting potential (Eb) is the critical potential above which pitting corrosion starts on steel surface [36], the repassivation potential (Epp) is the lowest potential for localized corrosion to propagate [38], and the pit transition potential (Eptp) is a characteristic potential that is correlated with the repassivation at the pit bottom [39,44,45].
The EIS measurements were performed under the corrosion potential with a potential perturbation of 15 mV and the scanning frequency range was 100 kHz‒0.01 Hz. The impedance data were fitted using ZSimpWin software. A three-electrode system was used in the electrochemical tests. A platinum electrode functioned as a counter electrode, the steel specimen worked as the working electrode, and the reference electrode was a saturated calomel reference electrode (SCE). Six parallel tests were run under each experimental condition.
An X-ray photoelectron spectroscopy (XPS) test was performed to analyze the surface composition with a Thermo Fisher ESCALAB 250 spectrometer (Qaltham, MA, USA). The samples were immersed in the carbonated SCP solution with different inhibitors for 24 h before XPS analysis. The binding energy values were calibrated using the C 1s peak at 248.8 eV.
To measure the wettability, the contact angle tests were performed on the inhibitor treated Q235 carbon steel surface using a JGW-36 contact angle test system (Chenghui Company, Chengde, China). During the test, the steel sample was mounted on the horizontal stage, the water drop (5 μL) was placed onto the steel surface using the syringe and the contact angle was measured using an integrated camera. For each sample, three distinct locations on the surface were measured and the average value was taken.

4. Conclusions

In carbonated SCP solution with 0.02 M NaCl, CLS could increase the pitting potential and decrease the corrosion current density for Q235 carbon steel, and the inhibition efficiency reaches 91.27% with 800 ppm CLS addition. Na2MoO4 also shows inhibition for both pitting and general corrosion, but the inhibition effect is obviously lower than that of CLS.
CLS and Na2MoO4 show a synergistic inhibition effect in the tested system. The compounded inhibitor can be defined as mix-type inhibitor. With 400 ppm CLS plus 600 ppm Na2MoO4, the pitting potential moves positively about 200 mV and the IE% reaches 92.67%. After 24 h immersion, the IE% further increases up to 99.2%.
XPS results confirm the adsorption of CLS and reveal the precipitation of oxides or hydroxides of Ca and Mo. Na2MoO4 could promote the stability of the passive film, and the insoluble molybdenum compounds and CaO/Ca(OH)2, together with the adsorbed CLS, deposit on the steel surface, forming a complex film. The compounded film effectively inhibits corrosion of the steel surface.

Author Contributions

Investigation, B.L.; Supervision, Y.Z.; Writing—original draft, B.L.; Writing—review & editing, Y.Z.

Funding

This research was funded by the National Natural Science Foundation of China grant 51210001.

Acknowledgments

The authors would like to thank the National Natural Science Foundation of China (Contract 51210001) for support for this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, H.; Liu, Y.; Chen, W.; Du, R.; Lin, C. Corrosion behavior of reinforcing steel in simulated concrete pore solutions: A scanning micro-reference electrode study. Electrochim. Acta 2009, 54, 4067–4072. [Google Scholar] [CrossRef]
  2. Cao, Y.; Dong, S.; Zheng, D.; Wang, J.; Zhang, X.; Du, R.; Song, G.; Lin, C. Multifunctional inhibition based on layered double hydroxides to comprehensively control corrosion of carbon steel in concrete. Corros. Sci. 2017, 126, 166–179. [Google Scholar] [CrossRef]
  3. Cao, F.; Wei, J.; Dong, J.; Ke, W. The corrosion inhibition effect of phytic acid on 20SiMn steel in simulated carbonated concrete pore solution. Corros. Sci. 2015, 100, 365–376. [Google Scholar] [CrossRef]
  4. Anonymous. Draft recommendation for repair strategies for concrete structures damaged by reinforcement corrosion. Mater. Struct. 1994, 27, 415–436. [Google Scholar] [CrossRef]
  5. Yang, Z.; Fischer, H.; Polder, R. Synthesis and characterization of modified hydrotalcites and their ion exchange characteristics in chloride-rich simulated concrete pore solution. Cem. Concr. Compos. 2014, 47, 87–93. [Google Scholar] [CrossRef]
  6. Huet, B.; L’Hostis, V.; Miserque, F.; Idrissi, H. Electrochemical behavior of mild steel in concrete: Influence of pH and carbonate content of concrete pore solution. Electrochim. Acta 2005, 51, 172–180. [Google Scholar] [CrossRef]
  7. Yang, Z.; Fischer, H.; Polder, R. Modified hydrotalcites as a new emerging class of smart additive of reinforced concrete for anticorrosion applications: A literature review. Mater. Corros. 2014, 64, 1066–1074. [Google Scholar] [CrossRef]
  8. Jin, M.; Xu, J.; Jiang, L.; Xu, Y.; Chu, H. Investigation on the performance characteristics of chloride selective electrode in concrete. Ionics 2015, 21, 2981–2992. [Google Scholar] [CrossRef]
  9. Hamdani, N.E.; Fdil, R.; Tourabi, M.; Jama, C.; Bentiss, F. Alkaloids extract of Retama monosperma (L.) Boiss. seeds used as novel eco-friendly inhibitor for carbon steel corrosion in 1 M HCl solution: Electrochemical and surface studies. Appl. Surf. Sci. 2015, 357, 1294–1305. [Google Scholar] [CrossRef]
  10. Faustin, M.; Maciuk, A.; Salvin, P.; Roos, C.; Lebrini, M. Corrosion inhibition of C38 steel by alkaloids extract of Geissospermum laeve in 1M hydrochloric acid: Electrochemical and phytochemical studies. Corros. Sci. 2015, 92, 287–300. [Google Scholar] [CrossRef]
  11. Heakal, E.T.; Elkholy, A.E. Gemini surfactants as corrosion inhibitors for carbon steel. J. Mol. Liq. 2017, 230, 395–407. [Google Scholar] [CrossRef]
  12. Abdulrahman, A.S.; Ismail, M.; Hussain, M.S. Corrosion inhibitors for steel reinforcement in concrete: A review. Sci. RES. Essays 2011, 6, 4152–4162. [Google Scholar] [CrossRef]
  13. Abdulrahman, A.S.; Ismail, M.; Hussain, M.S. Inhibiting sulphate attack on concrete by hydrophobic green plant extract. Adv. Mater. Res. 2011, 250, 3837–3843. [Google Scholar] [CrossRef]
  14. Abdulrahman, A.S.; Ismail, M.; Hussain, M.S. Inhibition of corrosion of mild steel in hydrochloric acid by bambusa arundinacea. Int. Rev. Mech. Eng. 2011, 5, 59–63. [Google Scholar]
  15. Tang, F.; Wang, X.; Xu, X.; Li, L. Phytic cid doped nanoparticles for green anticorrosion coatings. Colloids Surf. A 2010, 369, 101–105. [Google Scholar] [CrossRef]
  16. Mohamed, H.A. Eco-friendly zero VOC anticorrosive paints for steel protection. J. Appl. Polym. Sci. 2012, 125, 1790–1795. [Google Scholar] [CrossRef]
  17. Ouyang, X.; Qiu, X.; Lou, H.; Yang, D. Corrosion and Scale Inhibition Properties of Sodium Lignosulfonate and Its Potential Application in Recirculating Cooling Water System. Ind. Engin. Chem. Res. 2006, 45, 5716–5721. [Google Scholar] [CrossRef]
  18. Macias, A.; Goñi, S. Characterization of admixture as plasticizer or superplasticizer by deflocculation test. ACI Mater. J. 1999, 96, 40–46. [Google Scholar]
  19. Forostyan, Y.N.; Prosper, M. Inhibiting properties of lignin. Chem. Nat. Compd. 1983, 19, 358–359. [Google Scholar] [CrossRef]
  20. Yahya, S.; Othman, N.K.; Daud, A.R.; Jalar, A. The inhibition of carbon steel corrosion by lignin in HCl and H2SO4, Universiti-kebangsaan-malaysia, Faculty-of-science-and-technology. Am. Inst. Phys. 2013, 1571, 132–135. [Google Scholar]
  21. Akbarzadeh, E.; Ibrahim, M.N.M.; Rahim, A.A. Monomers of lignin as corrosion inhibitors for mild steel: Study of their behaviour by factorial experimental design. Br. Corros. J. 2012, 47, 302–311. [Google Scholar] [CrossRef]
  22. Abu-Dalo, M.A.; Al-Rawashdeh, N.A.F.; Ababneh, A. Evaluating the performance of sulfonated Kraft lignin agent as corrosion inhibitor for iron-based materials in water distribution systems. Desalination 2013, 313, 105–114. [Google Scholar] [CrossRef]
  23. Bishop, M.; Barron, A.R. Cement Hydration Inhibition with Sucrose, Tartaric Acid, and Lignosulfonate:  Analytical and Spectroscopic Study. Ind. Eng. Chem. Res. 2006, 45, 7042–7049. [Google Scholar] [CrossRef]
  24. Li, R.; Yang, D.; Guo, W.; Qiu, X. The adsorption mechanisms of sodium lignosulfonate on Al2O3 particles in aqueous solution. Adv. Chem. Eng. 2012, 550–553, 1120–1123. [Google Scholar] [CrossRef]
  25. Wang, Y.; Zuo, Y.; Zhao, X.; Zha, S. The adsorption and inhibition effect of calcium lignosulfonate on Q235 carbon steel in simulated concrete pore solution. Appl. Surf. Sci. 2016, 379, 98–110. [Google Scholar] [CrossRef]
  26. Emregül, K.C.; Aksüt, A.A. The effect of sodium molybdate on the pitting corrosion of aluminum. Corros. Sci. 2003, 45, 2415–2433. [Google Scholar] [CrossRef]
  27. Bai, W.; Li, X.; Mu, G.; Qu, Q. Molybdate and tungstate as corrosion inhibitors for cold rolling steel in hydrochloric acid solution. Clean. World 2006, 48, 445–459. [Google Scholar]
  28. Kumari, V.A.; Sreevalsan, K.; Shibli, S.M.A. Sodium molybdate for the effective protection of steel: A comprehensive review. Corros. Prev. Contro. 2001, 48, 83–109. [Google Scholar]
  29. Zhou, Y.; Zuo, Y. The inhibitive mechanisms of nitrite and molybdate anions on initiation and propagation of pitting corrosion for mild steel in chloride solution. Appl. Surf. Sci. 2015, 353, 924–932. [Google Scholar] [CrossRef]
  30. Cheng, T.; Lee, J.; Tsai, W. Passivation of titanium in molybdate-containing sulphuric acid solution. Electrochim. Acta 1991, 36, 2069–2076. [Google Scholar] [CrossRef]
  31. Atia, A.A.; Donia, A.M.; Awed, H.A. Synthesis of magnetic chelating resins functionalized with tetraethylenepentamine for adsorption of molybdate anions from aqueous solutions. J. Hazard. Mater. 2008, 155, 100–108. [Google Scholar] [CrossRef] [PubMed]
  32. Loto, R.T. Study of the synergistic effect of 2-methoxy-4-formylphenol and sodium molybdenum oxide on the corrosion inhibition of 3CR12 ferritic steel in dilute sulphuric acid. Result. Phys. 2017, 7, 769–776. [Google Scholar] [CrossRef]
  33. Zhou, Y.; Zuo, Y.; Lin, B. The compounded inhibition of sodium molybdate and benzotriazole on pitting corrosion of Q235 steel in NaCl+NaHCO3, solution. Mat. Chem. Phys. 2017, 192, 86–93. [Google Scholar] [CrossRef]
  34. Ferreira, E.S.; Giacomelli, C.; Giacomelli, F.C.; Spinelli, A. Evaluation of the inhibitor effect of l-ascorbic acid on the corrosion of mild steel. Mat. Chem. Phys. 2004, 83, 129–134. [Google Scholar] [CrossRef]
  35. Bouanis, M.; Tourabi, M.; Nyassi, A.; Zarrouk, A.; Jama, C.; Bentiss, F. Corrosion inhibition performance of 2,5-bis(4-dimethylaminophenyl)-1,3,4-oxadiazole for carbon steel in HCl solution: Gravimetric, electrochemical and XPS studies. Appl. Surf. Sci. 2016, 389, 952–966. [Google Scholar] [CrossRef]
  36. Standard test method for conducting cyclic potentiodynamic polarization measurements for localized corrosion susceptibility of iron-, nickel-, or cobalt-based alloys, ASTM-G61-86. In Annual Book of ASTM Standards; ASTM Internationa: Philadelphia, PA, USA, 1999; pp. 239–242.
  37. Tiziano, B.; Giampaolo, G.; Gabriella, R. Study of stainless steels corrosion in a strong acid mixture. Part 1: Cyclic potentiodynamic polarization curves examined by means of an analytical method. Corros. Sci. 2018, 130, 113–125. [Google Scholar]
  38. Blin, F.; Koutsoukos, P.; Klepetsianis, P.; Forsyth, M. The corrosion inhibition mechanism of new rare earth cinnamate compounds—Electrochemical studies. Electrochim. Acta 2007, 52, 6212–6220. [Google Scholar] [CrossRef]
  39. Ou, H.H.; Tran, Q.T.P.; Lin, P.H. A synergistic effect between gluconate and molybdate on corrosion inhibition of recirculating cooling water systems. Corros. Sci. 2018, 133, 231–239. [Google Scholar] [CrossRef]
  40. Refaey, S.A.M.; El-Rehim, S.S.A.; Taha, F.; Saleh, M.B.; Ahmed, R.A. Inhibition of chloride localized corrosion of mild steel by PO43−, CrO42−, MoO42−, and NO2 anions. Appl. Surf. Sci. 2000, 158, 190–196. [Google Scholar] [CrossRef]
  41. Eghbali, F.; Moayed, M.H.; Davoodi, A.; Ebrahimi, N. Critical pitting temperature (CPT) assessment of 2205 duplex stainless steel in 0.1 M NaCl at various molybdate concentrations. Corros. Sci. 2011, 53, 513–522. [Google Scholar] [CrossRef]
  42. Ilevbare, G.O.; Burstein, G.T. The inhibition of pitting corrosion of stainless steels by chromate and molybdate ions. Corros. Sci. 2003, 45, 1545–1569. [Google Scholar] [CrossRef]
  43. Moayed, M.H.; Newman, R.V. Aggressive effects of pitting ‘inhibitors’ on highly alloyed stainless steels. Corros. Sci. 2006, 48, 3513–3530. [Google Scholar] [CrossRef]
  44. Trasatti, S.P. Electrochemical approach to repassivation kinetics of Al alloys: Gaining insight into environmentally assisted cracking: Corrosion Reviews. Corros. Rev. 2015, 33, 373–393. [Google Scholar]
  45. Trueba, M. The repassivation response from single cycle anodic polarization: The case study of a sensitized Al-Mg alloy. Electrochim. Acta 2018, 259, 492–499. [Google Scholar] [CrossRef]
  46. Vukasovich, M.S.; Farr, J.P.G. Molybdate in corrosion inhibition—A review. Polyhedron 1986, 5, 551–559. [Google Scholar] [CrossRef]
  47. Tang, Y.; Zuo, Y.; Wang, J.; Zhao, X.; Niu, B.; Lin, B. The metastable pitting potential and its relation to the pitting potential for four materials in chloride solutions. Corros. Sci. 2014, 80, 111–119. [Google Scholar] [CrossRef]
  48. Danaee, I.; Khomami, M.N.; Attar, A.A. Corrosion behavior of AISI 4130 steel alloy in ethylene glycol–water mixture in presence of molybdate. Mat. Chem. Phys. 2012, 135, 658–667. [Google Scholar] [CrossRef]
  49. Samiento-Bustos, E.; Rodriguez, J.G.G.; Uruchurtu, J.; Dominguez-Patiño, G.; Salinas-Bravob, V.M. Effect of inorganic inhibitors on the corrosion behavior of 1018 carbon steel in the LiBr + ethylene glycol + H2O mixture. Corros. Sci. 2008, 50, 2296–2303. [Google Scholar] [CrossRef]
  50. Zhang, F.; Pan, J.; Claesson, P.M. Electrochemical and AFM studies of mussel adhesive protein (Mefp-1) as corrosion inhibitor for carbon steel. Electrochim. Acta 2011, 56, 1636–1645. [Google Scholar] [CrossRef]
  51. Zhang, Z.; Tian, N.; Zhang, W.; Huang, X.; Ruan, L.; Wu, L. Inhibition of carbon steel corrosion in phase-change-materials solution by methionine and proline. Corros. Sci. 2016, 111, 675–689. [Google Scholar] [CrossRef]
  52. Breslin, C.B.; Treacy, G.; Carroll, W.M. Studies on the passivation of aluminium in chromate and molybdate solutions. Corros. Sci. 1994, 36, 1143–1154. [Google Scholar] [CrossRef]
  53. Lopez-Garrity, O.; Frankel, G.S. Corrosion Inhibition of Aluminum Alloy 2024-T3 by Sodium Molybdate. Corrosion -Houston Tx- 2014, 70, 928–941. [Google Scholar] [CrossRef]
  54. Han, P.; Li, W.; Tian, H.; Gao, X.; Ding, R.; Xiong, C.; Chen, C. Designing and fabricating of time-depend self-strengthening inhibitor film: Synergistic inhibition of sodium dodecyl sulfate and 4-mercaptopyridine for mild steel. J. Mol. Liq. 2018, 268, 425–437. [Google Scholar] [CrossRef]
  55. Zhou, B.; Wang, Y.; Zuo, Y. Evolution of the corrosion process of AA 2024-T3 in an alkaline NaCl solution with sodium dodecylbenzenesulfonate and lanthanum chloride inhibitors. Appl. Surf. Sci. 2015, 357, 735–744. [Google Scholar] [CrossRef]
  56. Tourabi, M.; Nohair, K.; Nyassi, A. Thermodynamic characterization of metal dissolution and inhibitor adsorption processes in mild steel/3,5-bis(3,4-dimethoxyphenyl)-4-amino-1,2,4-triazole/hydrochloric acid system. J. Mater. Enviro. Sci. 2014, 5, 1133–1143. [Google Scholar]
  57. Gerónimo-López, C.; Vazquez-Arenas, J.; Picquart, M.; Picquart, M.; Gonzáleza, I. The energetic conditions determining the active dissolution of carbon steel during electrocoagulation in sulfate media. Electrochim. Acta 2014, 136, 146–156. [Google Scholar] [CrossRef]
  58. Hirschorn, B.; Orazem, M.E.; Tribollet, B.; Vivier, V.; Frateur, I.; Musiani, M. Determination of effective capacitance and film thickness from constant-phase-element parameters. Electrochim. Acta 2010, 55, 6218–6227. [Google Scholar] [CrossRef]
  59. Xun-jie, G.; Yu-chun, L.; Ke-ru, P. Electrochemical behavior of molybdate inhibitor in tap water. Corros. Sci. Prot. Technol. 2001, 13, 208–210. [Google Scholar]
  60. Frateur, I.; Carnot, A.; Zanna, S.; Marcus, P. Role of pH and calcium ions in the adsorption of an alkyl N-aminodimethylphosphonate on steel: An XPS study. Appl. Surf. Sci. 2006, 252, 2757–2769. [Google Scholar] [CrossRef]
  61. Olivares, O.; Likhanova, N.V.; Gómez, B.; Navarrete, J.; Llanos-Serrano, M.E.; Arce, E.; Hallen, J.M. Electrochemical and XPS studies of decylamides of α-amino acids adsorption on carbon steel in acidic environment. Appl. Surf. Sci. 2006, 252, 2894–2909. [Google Scholar] [CrossRef]
  62. Pang, X.; Ran, X.; Fei, K.; Xie, J.; Hou, B. Inhibiting Effect of Ciprofloxacin, Norfloxacin and Ofloxacin on Corrosion of Mild Steel in Hydrochloric Acid. Chin. J. Chem. Eng. 2010, 18, 337–345. [Google Scholar] [CrossRef]
  63. Dong, Z.H.; Shi, W.; Guo, X.P. Localized Corrosion Inhibition of Carbon Steel in Carbonated Concrete Pore Solutions Using Wire Beam Electrodes. Acta Phys. Chim. Sin. 2011, 27, 127–134. [Google Scholar]
  64. Nakayama, N. Inhibitory effects of nitrilotris (methylenephosphonic acid) on cathodic reactions of steels in saturated Ca(OH)2 solutions. Corros. Sci. 2000, 42, 1897–1920. [Google Scholar] [CrossRef]
  65. Clayton, C.R.; Lu, Y.C. A bipolar model of the passivity of stainless steels - III. The mechanism of MoO42− formation and incorporation. Corros. Sci. 1989, 29, 881–898. [Google Scholar] [CrossRef]
  66. Muñoz, A.I.; Antón, J.G.; Nuévalos, S.L.; Guiñón, J.L.; Herranz, V.P. Corrosion studies of austenitic and duplex stainless steels in aqueous lithium bromide solution at different temperatures. Corros. Sci. 2004, 46, 2955–2974. [Google Scholar]
  67. Temesghen, W.; Sherwood, P.M. Analytical utility of valence band X-ray photoelectron spectroscopy of iron and its oxides, with spectral interpretation by cluster and band structure calculations. Anal. Bioanal. Chem. 2002, 373, 601–608. [Google Scholar] [CrossRef] [PubMed]
  68. Sajid, H.U.; Kiran, R. Influence of corrosion and surface roughness on wettability of ASTM A 36 steels. J. Constr. Steel Res. 2018, 144, 310–326. [Google Scholar] [CrossRef]
  69. Datta, A.; Mukherjee, S. Structural and Morphological Evolution in Metal-Organic Films and Multilayers; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  70. Liang, C.; Lv, Z.; Zhu, Y.; Xu, S. Molybdate-based conversion treatment for improving the peeling strength between aluminum foil and polypropylene grafted with glycidyl methacrylate. Surf. Coat. Technol. 2014, 249, 1–5. [Google Scholar] [CrossRef]
  71. Rutland, M.W.; Parker, J.L. Surfaces Forces between Silica Surfaces in Cationic Surfactant Solutions: Adsorption and Bilayer Formation at Normal and High pH. Langmuir 1994, 10, 1110–1121. [Google Scholar] [CrossRef]
  72. Yu, H.; Chiang, K.T.K. Threshold chloride level and characteristics of reinforcement corrosion initiation in simulated concrete pore solutions. Construct. Build. Mat. 2012, 26, 723–729. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. (a) Polarization curves of Q235 carbon steel in simulated concrete pore (SCP) solution with various concentration of calcium lignosulfonate (CLS); (b) Electrochemical parameters based on the polarization curves.
Figure 1. (a) Polarization curves of Q235 carbon steel in simulated concrete pore (SCP) solution with various concentration of calcium lignosulfonate (CLS); (b) Electrochemical parameters based on the polarization curves.
Molecules 24 00518 g001
Figure 2. (a) Polarization curves of Q235 carbon steel in SCP solutions with various concentrations of Na2MoO4; (b) The electrochemical parameters based on the polarization curves.
Figure 2. (a) Polarization curves of Q235 carbon steel in SCP solutions with various concentrations of Na2MoO4; (b) The electrochemical parameters based on the polarization curves.
Molecules 24 00518 g002
Figure 3. Polarization curves in solutions with different ratios of CLS and Na2MoO4.
Figure 3. Polarization curves in solutions with different ratios of CLS and Na2MoO4.
Molecules 24 00518 g003
Figure 4. Polarization curves and parameters of Q235 steel immersed in carbonated SCP solution for various times, (a) without inhibitor, (b) with compound inhibitor, (c) electrochemical parameters obtained based on polarization curves.
Figure 4. Polarization curves and parameters of Q235 steel immersed in carbonated SCP solution for various times, (a) without inhibitor, (b) with compound inhibitor, (c) electrochemical parameters obtained based on polarization curves.
Molecules 24 00518 g004
Figure 5. Nyquist plots of Q235 carbon steel in test solutions, (a) with CLS; (b) with Na2MoO4; (c) with optimal ratio of CLS and Na2MoO4; (d) the equivalent circuit.
Figure 5. Nyquist plots of Q235 carbon steel in test solutions, (a) with CLS; (b) with Na2MoO4; (c) with optimal ratio of CLS and Na2MoO4; (d) the equivalent circuit.
Molecules 24 00518 g005
Figure 6. Electrochemical parameters of Q235 steel in carbonated SCP solution with different concentrations of inhibitor using electrochemical impedance spectroscopy (EIS) measurements, (a) Cdl and Cfilm; (b) Rct and Rfilm.
Figure 6. Electrochemical parameters of Q235 steel in carbonated SCP solution with different concentrations of inhibitor using electrochemical impedance spectroscopy (EIS) measurements, (a) Cdl and Cfilm; (b) Rct and Rfilm.
Molecules 24 00518 g006
Figure 7. Nyquist polts (a) and EIS parameters (b) of Q235 steel in test solution with different pretreatment filming time.
Figure 7. Nyquist polts (a) and EIS parameters (b) of Q235 steel in test solution with different pretreatment filming time.
Molecules 24 00518 g007
Figure 8. High-resolution XPS spectrum of Q235 steel immersed in test solution with CLS and CLS+Na2MoO4: (a) C 1s; (b) Ca 2p and S 2p.
Figure 8. High-resolution XPS spectrum of Q235 steel immersed in test solution with CLS and CLS+Na2MoO4: (a) C 1s; (b) Ca 2p and S 2p.
Molecules 24 00518 g008
Figure 9. High-resolution XPS spectrum of Mo 3d for Q235 carbon steel immersed in carbonation SCP solution with Na2MoO4 and CLS+Na2MoO4.
Figure 9. High-resolution XPS spectrum of Mo 3d for Q235 carbon steel immersed in carbonation SCP solution with Na2MoO4 and CLS+Na2MoO4.
Molecules 24 00518 g009
Figure 10. High-resolution XPS spectrum of Q235 carbon steel immersed in carbonation SCP solution with different inhibitors: (a) O 1s; (b) Fe 2p.
Figure 10. High-resolution XPS spectrum of Q235 carbon steel immersed in carbonation SCP solution with different inhibitors: (a) O 1s; (b) Fe 2p.
Molecules 24 00518 g010
Figure 11. The Fe3+/Fe2+ ratio for Q235 steel immersed in test solution with different inhibitors.
Figure 11. The Fe3+/Fe2+ ratio for Q235 steel immersed in test solution with different inhibitors.
Molecules 24 00518 g011
Figure 12. Typical images of water droplets on different inhibitor treated Q235 steel surfaces.
Figure 12. Typical images of water droplets on different inhibitor treated Q235 steel surfaces.
Molecules 24 00518 g012
Figure 13. Illustration of the passivation-adsorption structure.
Figure 13. Illustration of the passivation-adsorption structure.
Molecules 24 00518 g013
Figure 14. Chemical structure of calcium lignosulfonate (CLS).
Figure 14. Chemical structure of calcium lignosulfonate (CLS).
Molecules 24 00518 g014
Table 1. Electrochemical parameters of Q235 steel in carbonated SCP solution with different ratios of CLS and Na2MoO4 using cyclic potentiodynamic polarization (CPP) measurements.
Table 1. Electrochemical parameters of Q235 steel in carbonated SCP solution with different ratios of CLS and Na2MoO4 using cyclic potentiodynamic polarization (CPP) measurements.
InhibitorEcorrEbEppIE%S
CLSNa2MoO4mVSCEmVSCEmVSCE
00−298−171.2−584
200800−375−73.2−40793.886.89
400600−3383.9−39992.672.98
600400−310−43.8−42981.731.34
800200−364−15.3−57778.331.11
10000−304−0.3−56591.27
Table 2. Electrochemical parameters of Q235 steel in carbonated SCP solution with different ratios of CLS and Na2MoO4 using CPP measurements.
Table 2. Electrochemical parameters of Q235 steel in carbonated SCP solution with different ratios of CLS and Na2MoO4 using CPP measurements.
InhibitorContact Angle (Degrees)Average Contact Angle (Degrees)
Location 1Location 2Location 3
Without68.0066.7571.5468.76
CLS18.3218.5616.3717.75
Na2MoO485.8489.0079.0484.63
Compound30.1233.8436.4233.46

Share and Cite

MDPI and ACS Style

Lin, B.; Zuo, Y. Inhibition of Q235 Carbon Steel by Calcium Lignosulfonate and Sodium Molybdate in Carbonated Concrete Pore Solution. Molecules 2019, 24, 518. https://doi.org/10.3390/molecules24030518

AMA Style

Lin B, Zuo Y. Inhibition of Q235 Carbon Steel by Calcium Lignosulfonate and Sodium Molybdate in Carbonated Concrete Pore Solution. Molecules. 2019; 24(3):518. https://doi.org/10.3390/molecules24030518

Chicago/Turabian Style

Lin, Bing, and Yu Zuo. 2019. "Inhibition of Q235 Carbon Steel by Calcium Lignosulfonate and Sodium Molybdate in Carbonated Concrete Pore Solution" Molecules 24, no. 3: 518. https://doi.org/10.3390/molecules24030518

Article Metrics

Back to TopTop