Effect of pH on the Passivation of Carbon Steel by Sodium Borosilicate

Jun 4, 2017 - Exploring ChemRxiv and the Value of Preprint Servers with Reddit. On October 17, ACS Publication hosted a Reddit AMA with Darla Henderso...
0 downloads 8 Views 6MB Size
Article pubs.acs.org/IECR

Effect of pH on the Passivation of Carbon Steel by Sodium Borosilicate Controlled-Release Inhibitor in Simulated Recirculating Cooling Water Jun Cui,† Wenjiao Yuan,† Donghai Yuan,*,‡ and Yuansheng Pei*,† †

The Key Laboratory of Water and Sediment Sciences, Ministry of Education, School of Environment, Beijing Normal University, Beijing 100875, People’s Republic of China ‡ Key Laboratory of Urban Stormwater System and Water Environment, Beijing University of Civil Engineering and Architecture, Beijing 100044, People’s Republic of China S Supporting Information *

ABSTRACT: A sodium borosilicate controlled-release inhibitor has been prepared via calcination for suppressing carbon steel corrosion in simulated recirculating cooling water (SRCW). The synthesized inhibitor was mainly composed of B2O3 and SiO2, coupling with small quantities of NaB5O6(OH)4·3H2O and NaB5O8·5H2O crystals. The total boron concentration gradually increased to 250 mg L−1 during the 28 days of release (2 g inhibitor in 500 mL SRCW) at 50 °C. A higher controlledreleased rate existed in acidic environments (pH 5) than in neutral environments (pH 7) and alkaline environments (pH 9). The potentiodynamic polarization and electrochemical impedance spectroscopy results indicated that the inhibitor showed a better anticorrosion performance in neutral and alkaline environments than in acidic environments. In addition, a passive film with a Fe−O−B structure formed on the carbon steel surface to prevent the attack of the corrosive ions. A passivation-pittingrepassivation process was detected in the passive film by electrochemical noise, which demonstrated that the passive film exhibited a self-healing capacity. Moreover, a potential corrosion risk under the film was detected under the passive film in alkaline environment after 21 days of immersion. According to the experiment results and the analysis results of the point defect model, we believed that the anticorrosion performance of the sodium borosilicate inhibitor was in the following order: pH 7 > pH 9 > pH 5.



solution to maintain a relatively stable pH.14,15 A solution with high borate concentration (0.05 M H3BO3 + 0.075 M Na2B4O7· 10H2O, with a boron concentration of >2000 mg L−1) is widely used in the metal corrosion field, because it has excellent buffer performance.16,17 However, such a high boron concentration is impossible to be applied in the RCW system, because of increasing the treatment cost of the discharge and bringing pollution risk to the discharge receive region. Instead, this issue may be resolved by using a low borate concentration solution (boron concentrations of pH 5. However, according to the Nyquist results, a potential corrosion behavior under the passive film happened after 21 days of immersion in an alkaline environment. This type of local corrosion occurred with the growth of the passive film. Therefore, we considered that, long term, the inhibitor showed better anticorrosion performance in neutral environments than in alkaline environments. Mott−Schottky Measurements. From the Tafel and EIS analysis results, we inferred that a passive film was formed on the carbon steel surface. Previous studies based on PDM theory reported that the anticorrosion performance of the passive film is directly related to its semiconducting behavior.41,42 Hence, in this work, MS measurements were performed to evaluate the semiconductive property of the passive film in SRCW under different pH conditions with increasing inhibitor release time. Prior to the MS experiment, we assumed that the capacitance of the passive film in SRCW was mainly contributed by the space charge layer other than the Helmholtz layer, which determines the validity of the MS analysis.43 Figure 9 represents the C2 vs E plots for passive film semiconductive characterizations of all samples with increasing release time under different pH conditions. In this work, the passivation potential region derived from Tafel curve was selected as the potential test range of MS; the test frequency was chosen from the EIS results, when the capacitance was in a depletion condition (see Table 2). In addition, all the MS curves were obtained by sweeping at a high velocity to decrease the effect of film growth during the experiment. Figure 9 shows that almost all passive films exhibited a duplex structure, according to their n-type and p-type semiconducting behaviors, except for sample 1, after 28 days of immersion. Furthermore, the maximum peak value increased as the immersion time increased, resulting from the increasing capacitance effect of the passive film with the increasing boron concentration. Hence, two regions were distinguished, based on the passive film semiconductive property, and the corresponding parameters (donor density (Nd) and acceptor density (Na)) are tabulated in Table 3.44 Almost all samples showed two regions, except for sample 1 after 28 days of immersion; the positive slope is attributed to ntype behavior, while the negative slope is assigned to p-type

potential (V) Sample 1 −0.90 to −0.44 −0.70 to −0.14 −0.50 to −0.10 Sample 2 −0.80 to −0.24 −0.40 to 0.14 −0.20 to 0.60 Sample 3 −0.76 to −0.16 −0.40 to 0.20 −0.20 to 0.40

frequency (Hz) 0.37−71.96 0.01−100 000 0.05−372.76 0.19−37.27 37.27−1000 0.05−100 000 0.01−100 000 1.00−71.96 0.01−193.07

behavior. According to the PDM theory, cation vacancy, oxygen vacancy, and metal interstitial are the main defect types in the passive film. Cation vacancies are electron acceptors in the space charge region of the p-type passive film. In contrast, oxygen vacancies and metal interstitials are electron donors in the space charge region of the n-type passive film.45,46 Hence, the semiconductor property of the passive film depends on the electron location.17 When the number of electrons located in the conduction band of the passive film is more than the number of holes in the valence band, the passive film performs as an n-type conductor, such as the existence of the Fe2O3 and Fe(OH)3. Conversely, p-type semiconductor property is observed in layers composing of Fe3O4.47 Hence, the MS curves can be explained as follows. Fe2O3 as the main component in the passive film was formed on the surface of sample 1 at a relative low potential, leading to an n-type characteristic of the passive film. With the increasing measurement potential, the passive film was gradually dominated by the Fe3O4 in the vicinity of the turning potential. The semiconductive characterization of the passive film then changed from n-type to p-type, which was attributed to an inversion in the electronic character. Similarly, the conversion of the semiconductive characterization of sample 3 was attributed to the variation of the domination from Fe(OH)3 to Fe3O4. In addition, Nd and Na were calculated from the slope of the C2 vs E plots. Table 3 indicates that Nd and Na of all samples showed a decreasing trend with increasing immersion time, and the carrier density ranked in the following order of samples: 2 < 3 < 1. The results indicated that the number of the point defects decreased as the release time increased, and the capability of the passive film to hinder electron transport between carbon steel and SRCW ranked in the following order of samples: 2 < 3 < 1. The passive film formed at 0.22 V (sample 2) and 0.18 V (sample 3) after 28 days of immersion are due to the destruction of the Fe2O3 or Fe(OH)3 layer, which was regarded as a protective film. In contrast, no turning potential was observed from sample 1 after 28 days of immersion. The results revealed that the components of the passive film were not changed. In addition, in this work, the variation of the semiconductive characteristic might be related to the formation of corrosion products above or below the passive film. The corrosion products closely combined with the passive film, resulting in the variation of the semiconductor behavior of the passive film. Figure 9 and Tables 2 and 3 show that the test potential shifted toward the anodic region and the Na and Nd values sharply decreased. This is consistent with the theoretical prediction of the PDM, which suggests that the higher passivation potential is 7245

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research

Table 3. Evolution of the Donor and Acceptor Densities with Different pH Values Calculated from Mott−Schottky Experiment Sample 1 20

−3

Nd (× 10 cm ) Na (× 1020 cm−3)

Sample 2

Sample 3

1d

14 d

28 d

1d

14 d

28 d

1d

14 d

28 d

62.00

5.86 31.00

3.44

9.33 34.00

0.55 7.90

0.38 2.90

20.50 32.00

5.54 2.30

0.77 2.30

Figure 9. Mott−Schottky plots of passive film formed on carbon steel surface in SRCW with different pH values (a) pH 5, (b) pH 7, and (c) pH 9.

helpful to reduce Nd, stemming from promoting the formation of a highly ordered and less-defective passive layer.48 Compared with samples 1 and 3, sample 2 showed a critical potential at 0.04 after 28 days of immersion. This nonlinear phenomenon is attributed to the existence of a broad range of energies for Nd.43 In this work, Fe2O3 and Fe(OH)3 was supposed to be the main substance in n-type passive film in sample 2. Hence, the slope nonlinearity might be attributed to the influence of Fe3+ on the electronic structure of passive layers developing at potentials close to the valence band potential. Compared to samples 1 and 2, the MS curves of sample 3 after 14 and 28 days of immersion showed a larger shoulder, which revealed that the Fe3O4 content in passive film was higher in alkaline environments than in acidic and neutral environments. The thickness of the passive film formed on the carbon steel surface for all samples with increasing immersion time is presented in Figure 10. The thickness of the film was calculated

anticorrosion performance in a neutral environment than in acidic and alkaline environments. Moreover, the thickness of the passive film increased after 7 days of immersion. Combined with the EIS analysis, we confirmed that the passive film became integrated after 7 days of immersion. In this work, the trend of all MS curves differed from the previous results of a passive film on the carbon steel surface. We recognized that this phenomenon resulted from the existence of boron, which changed the semiconducting behavior of the passive film on the carbon steel surface. We inferred that a structure such as Fe−O−B formed on the carbon steel surface, to protect it from attack by aggressive ions. Considering the complexity of the SRCW and borate solution structure, the passive film and the corrosion product composition needs to be studied further. Electrochemical Noise Measurement. The dynamic process was characterized by EN for all samples under different pH conditions. Current and potential transient signals (with the DC trend removed) for all samples are presented in Figure 11. Both current and potential noises were immediately measured after immersing the electrode in the SRCW. For samples 2 and 3, the surface current and potential exhibited a clear decrease and increase, respectively, which indicated that a passive film was formed on the carbon steel surface under neutral and alkaline conditions, and not under acidic conditions. However, the current density decreased in sample 1, which resulted from the combination of a porous barrier layer and the incomplete passive film on the carbon steel surface. This structure resulted in the current oscillation in sample 1. Normally, the ECN signal shape reflects local corrosion processes. In this work, all samples showed a similar ECN signal shape, as shown in the inset of Figure 11. Various sharp peaks existed in the ECN curve; these ECN signal shapes were characterized by a fast increase, followed by an exponential decay, which was recognized as the occurrence of a carbon steel metastable pitting process, including pit initiation, rapid growth, and slow repassivation.50,51 The ECN results indicated that the inhibitor showed a passivation and repassivation function in the pH range from pH 5 to pH 9. In addition, with pitting corrosion occurring frequently, the pitting boundaries began to connect; then, general corrosion occurred,52 which is interpreted as the increase in pitting leading to pit connection in the passive film of sample 1 after 28 days of immersion.

Figure 10. Thickness variations of the passive film formed on the carbon steel surface for all samples with increasing immersion time.

based on the passive film capacitance calculated from EIS.49 Since the rust layer contributed little to the carbon steel surface capacitance, the calculation was semiquantitative for the evaluation of the passive film thickness. The passive film thickness was in the nanometer range and increased with immersion time. The passive film was thicker in sample 2 than in the others, which indicated that the inhibitor showed a better 7246

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research

Figure 11. ECN and EPN signals for the carbon steel in SRCW with (a) pH 5, (b) pH 7, and (c) pH 9. Two-dimensional visual presentation of discrete time wavelet transforms of ECN signals for the carbon steel in SRCW with (d) pH 5, (e) pH 7, and (f) pH 9.

It is widely recognized that pitting does not occur sequentially; thus, it is not accurate to use current and potential transients to characterize the pitting, because it could lead to the overlap of simultaneous formations of multiple pits.51,53 Hence, various methods have been proposed to deconvolve the time domain data. In this work, the wavelet transformation method was used to establish the relationship between the EN signal and variation in the carbon steel surface. The two-dimensional plots for all samples are shown in Figure 11. In this work, the rectangle represents the wavelet coefficient. Higher energies correspond to lighter colors. As can be seen from Figure 11d, levels 5−8 showed higher energies, which could be associated with the local corrosion, including intergranular corrosion and pit formation.53 Meanwhile, the rectangle color was lighter in the early stage of the experiment, which indicated that pitting dominated at the early stage and the intensity of pitting decreased with experiment time. In addition, Figures 11e and 11f show relative higher energies in levels 1−4, which is attributed to general corrosion; the sample in alkaline solution showed a lighter color than in neutral solution, which indicates that serious corrosion occurred in alkaline solution. The local and general corrosion were affected by the passivation progress. Hence, we believed that the passivation capacity of the inhibitor ranked in the following order: neutral > alkaline > acidic (environments).

Since the information on the corrosion mechanism was lost in the time domain curves, frequency domain curves were calculated by fast Fourier transformation. The PSD (I) curves that remove the DC trend are shown in Figure 12. All PSD (I) curves followed a similar trend. The PSD (I) curves decreased in the following order: samples 1 > 3 > 2. In addition, the electrochemical noise parameters obtained from the frequency domain curves are listed in Table 4. This table shows that the

Figure 12. PSD (I) plots associated with the samples in SRCW under different pH conditions in the presence of the inhibitor. 7247

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research

Table 4. Parameters Extracted from Frequency Domain Curves of the Carbon Steel in SRCW under Different pH Conditions sample 1 2 3

σV (V)

σI (A) −11

5.19 × 10 1.42 × 10−12 9.09 × 10−11

σV′ (V)

Rn (Ω cm2) −9

2.33 × 10 8.44 × 10−10 2.02 × 10−8

44.89 594.37 222.22

σI′ (A) −6

1.53 × 10 1.06 × 10−8 1.30 × 10−7

−10

1.00 × 10 8.67 × 10−12 1.52 × 10−11

f n (s−1 cm−2)

qn (C) −7

4.75 × 10 1.17 × 10−8 5.40 × 10−8

116.0 13.4 19.2

Figure 13. Schematic representation of the inhibitor anticorrosion performance in SRCW under different pH conditions: (a) pH 5, (b) pH 7, and (c) pH 9.

Figure 14. EDS results of the carbon steel surface in SRCW with (a) pH 5, (b) pH 7, and (c) pH 9. The detail XPS spectra of Fe 2p, B 2p, and O 1s on the carbon steel surface for (d) pH 5, (e) pH 7, and (f) pH 9 after 28 days of experiment, respectively.

7248

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research

work, PDM was applied to interpret the effect of pH on the growth and breakdown of a passive film on the carbon steel surface in SRCW. According to PDM theory, the enhancement of the cation vacancy flux (JCV) from the film/solution (F/S) interface to the metal/film (M/F) interface was the main cause for the occurrence of carbon steel corrosion. It is widely recognized that the cation vacancy (VM) appears at the M/F interface and moves to the metal surface under the action of an electric field. The VM combined with the Fe on the carbon steel surface to form FeM. After FeM was transferred from the M/F interface to the F/S interface, FeM was oxidized and entered the electrolyte. Since the VM could not be annihilated completely by cation injection into the film and reaction with Fe, the relationship between JCV and the cation vacancy annihilation flux (JM) was directly related to the VM concentration at the M/F interface. If JCV > JM, more VM would accumulate at the M/F interface, resulting in pitting. These pitting holes are supposed to be the weakest areas on the passive film, which are easily attacked by aggressive ions and result in the breakdown of the passive film.56−61 In SRCW, a relatively high Cl− concentration existed in the solution. Previous studies reported that Cl− showed a strong capacity to combine with oxygen vacancy (VO) formed with VM, based on the Schottky-pair theory;62 this promoted the production of VM at the F/S interface and contributed to the increase of JCV, and then the pitting occurred. In contrast, B would combine with O and Fe to form a passive film on the carbon steel surface. The Fe−O−B film hindered the VM transfer from the F/S interface to the M/F interface and reduced the JCV in neutral and alkaline environments, which resulted in the decrease of VM at the M/F interface. Hence, acceptable anticorrosion performance existed in neutral and alkaline environments. Moreover, in acidic environments, H+ could oxidize Fe in the Fe−O−B passive film, which resulted in the breakdown of the passive film.59 Meanwhile, the relatively high H+ concentration could restrain the combination of VO and H2O, which promoted the concentration of VO at the F/S interface. Hence, the inhibitor showed weak anticorrosion performance in acidic environments.

noise resistance (Rn) increased in the order of samples 1 < 3 < 2. The increase of Rn indicates that more inhibitor molecules adsorb on the carbon steel surface to suppress corrosion.54 In contrast, the charge of each electrochemical event (qn) showed the opposite trend. The higher qn values relate to more charge being transferred during the electrochemical event, indicating a higher current flux at the carbon steel surface.51 In addition, the event frequency (f n) as a novel index for detecting changes was calculated; a large f n value reflects the occurrence of general corrosion and a small f n value indicates local corrosion.55 The results indicated that the sample suffered stronger corrosion in sample 1 than in samples 2 and 3. The frequency domain analysis results are consistent with the EIS results, which indicated that the passivation effect was strongest in the neutral environment and weakest in the acidic environment. Mechanism Analysis. Figure 13 shows the growth process of a passive film on the carbon steel surface in SRCW containing the inhibitor. In the early stage, the film surfaces in samples 1, 2, and 3 were rough, slightly rough, and relatively smooth (with some precipitations), respectively. As less boron was released from the inhibitor in the early stage, the carbon steel surface was unprotected, which resulted in the occurrence of corrosion. At pH 5, the carbon steel was attacked by H+ and oxidized by NO3−. Then, Fe (electron donor) was oxidized to Fe3+, and H+ and NO3− (electron acceptors) were reduced to H2 and NO2−, respectively (assigned to the hydrogen evolution reaction). At pH 7, Fe was oxidized by O2 and NO3−, resulting in the formation of Fe3+, O2−, and NO2−. Meanwhile, O2− could react with the adsorbed atomic hydrogen, which resulted in depolarizing the carbon steel surface and permitting the continuous dissolution of Fe. At pH 9, Fe was oxidized by O2 and NO3−, and then reacted with OH−, which led to the formation of Fe(OH)3 that deposited on the passive film surface and provided a small amount of anticorrosion protection to the carbon steel by forming a deposition layer. After 28 days of immersion, the boron concentration was increased significantly and reached 250 mg L−1. The relative high boron concentration solution changed the carbon steel surface by forming a passive film. At pH 5, the passive film was formed as an inner layer, and the outer layer was formed by depositions that were mainly composed of Fe2O3 and FeOOH (see Figures 14a and 14d). Meanwhile, we found that the passive film was not intact under the continuous attack by the aggressive ions in acidic environments (Figure 13a), which contributed to the transfer of Fe from carbon steel to solution, resulting in the formation of various loose and porous pitting holes. At pH 7, a passive film was formed on the carbon steel surface, which provided an acceptable anticorrosion performance after 28 days of immersion. We inferred that the film was formed by reactions among Fe, B, and O (see Figures 14b and 14e). At pH 9, we found that corrosion behavior occurred below the passive film and the integrity of the passive film decreased slightly. In addition, it was noticed that the degree of corrosion below the passive film increased with the increase of immersion time in the late stage of the experiment, and the corrosion products contained Fe(OH)3 (Figures 14c and 14f). The results indicated that carbon steel suffered a potential corrosion risk in alkaline solution after 28 days of release of the sodium borosilicate inhibitor. From the analysis results, we recognized that the passivation mechanism of carbon steel by sodium borosilicate inhibitor was different under the different pH conditions. In recent years, PDM is regarded as a useful method to describe the growth and breakdown of passive films on the metal surfaces. Hence, in this



CONCLUSIONS In this work, a sodium borosilicate inhibitor was successfully synthesized to suppress carbon steel corrosion in simulated recirculating cooling water (SRCW), and the effect of pH on the anticorrosion performance of the inhibitor was analyzed by electrochemical methods coupled with surface analysis techniques. Moreover, we interpreted the passivation mechanism by a combination of point defect model (PDM) and experiment results, and the following results were obtained: (1) The inhibitor was composed of B2O3 and SiO2, coupled with small quantities of NaB5O6(OH)4·3H2O and NaB5O8· 5H2O crystals. In addition, the main elements released by the inhibitor were B and Na. Meanwhile, a higher controlled-released rate existed in acidic environments than in neutral and alkaline environments. (2) The inhibitor showed a long-term anticorrosion efficiency, and the efficiencies were >95% after 28 days of exposure in the pH range from pH 5 to pH 9. In acidic environments, a doublelayer structure, consisting of an incomplete passive film as the inner layer and a thick rust layer as the outer layer, formed on the carbon steel surface. In neutral environments, a single compact passive film with Fe−B−O structure formed on the carbon steel surface. In alkaline environments, a double-layer structure, 7249

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research consisting of an intact passive film as the outer layer and a thin rust layer as the inner layer, formed on the carbon steel surface. Hence, the anticorrosion capacity of the inhibitor was in the order of pH 7 > pH 9 > pH 5. (3) All passive films possessed a duplex semiconductor characteristic resulting from the variations in the content of Fe2O3, Fe(OH)3, and Fe3O4 in the passive film. In addition, a metastable pitting process, including pitting initiation, pitting core growth, and slow repassivation, occurred on the passive film surface in the presence of the inhibitor for all samples. (4) A Fe−O−B passive film was formed on the carbon steel surface to suppress corrosion in the presence of the inhibitor under different pH conditions. In addition, PDM results indicated that carbon steel corrosion was mainly induced by the combination of Cl− with VO in SRCW. This process led to the formation of VM and an increase in JCV, which resulted in the passive film detaching from the carbon steel surface after the accumulation of VM at the M/F interface.



(3) Theregowda, R. B.; Vidic, R.; Landis, A. E.; Dzombak, D. A.; Matthews, H. S. Integrating external costs with life cycle costs of emissions from tertiary treatment of municipal wastewater for reuse in cooling systems. J. Cleaner Prod. 2016, 112, 4733−4740. (4) Wang, F. H.; Hao, H. T.; Sun, R. F.; Li, S. Y.; Han, R. M.; Papelis, C.; Zhang, Y. Bench-scale and pilot-scale evaluation of coagulation pretreatment for wastewater reused by reverse osmosis in a petrochemical circulating cooling water system. Desalination 2014, 335 (1), 64−69. (5) Sun, J.; Feng, X.; Wang, Y.; Deng, C.; Chu, K. H. Pump network optimization for a cooling water system. Energy 2014, 67, 506−512. (6) Wan, Y.; Zhang, D.; Liu, H.; Li, Y.; Hou, B. Influence of sulphatereducing bacteria on environmental parameters and marine corrosion behavior of Q235 steel in aerobic conditions. Electrochim. Acta 2010, 55 (5), 1528−1534. (7) Maocheng, Y. A. N.; Jin, X. U.; Libao, Y. U; Tangqing, W. U.; Cheng, S. U. N.; Wei, K. E. EIS analysis on stress corrosion initiation of pipeline steel under disbonded coating in near-neutral pH simulated soil electrolyte. Corros. Sci. 2016, 110, 23−34. (8) Farsak, M.; Keleş, H.; Keleş, M. A new corrosion inhibitor for protection of low carbon steel in HCl solution. Corros. Sci. 2015, 98, 223−232. (9) Zhang, J.; Gong, X. L.; Yu, H. H.; Du, M. The inhibition mechanism of imidazoline phosphate inhibitor for Q235 steel in hydrochloric acid medium. Corros. Sci. 2011, 53 (10), 3324−3330. (10) Wang, H.; Gao, M.; Guo, Y.; Yang, Y.; Hu, R. A natural extract of tobacco rob as scale and corrosion inhibitor in artificial seawater. Desalination 2016, 398, 198−207. (11) Kim, Y. S.; Lee, S. H.; Son, M. Y.; Jung, Y. M.; Song, H. K.; Lee, H. Succinonitrile as a Corrosion Inhibitor of Copper Current Collectors for Overdischarge Protection of Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2014, 6 (3), 2039−2043. (12) Gerengi, H.; Mielniczek, M.; Gece, G.; Solomon, M. M. Experimental and Quantum Chemical Evaluation of 8-Hydroxyquinoline as a Corrosion Inhibitor for Copper in 0.1 M HCl. Ind. Eng. Chem. Res. 2016, 55 (36), 9614−9624. (13) Deyab, M. A.; Eddahaoui, K.; Essehli, R.; Rhadfi, T.; Benmokhtar, S.; Mele, G. Experimental evaluation of new inorganic phosphites as corrosion inhibitors for carbon steel in saline water from oil source wells. Desalination 2016, 383, 38−45. (14) Altaf, F.; Qureshi, R.; Ahmed, S. Surface protection of copper by azoles in borate buffers-voltammetric and impedance analysis. J. Electroanal. Chem. 2011, 659 (2), 134−142. (15) Sun, H.; Wu, X.; Han, E. H. Effects of temperature on the oxide film properties of 304 stainless steel in high temperature lithium borate buffer solution. Corros. Sci. 2009, 51 (12), 2840−2847. (16) Xu, H.; Wang, L.; Sun, D.; Yu, H. The passive oxide films growth on 316L stainless steel in borate buffer solution measured by real-time spectroscopic ellipsometry. Appl. Surf. Sci. 2015, 351, 367−373. (17) Lv, J. L.; Liang, T. X.; Wang, C.; Guo, W. L. Investigation of passive films formed on the surface of alloy 690 in borate buffer solution. J. Nucl. Mater. 2015, 465, 418−423. (18) Braunschweig, H.; Ye, Q.; Vargas, A.; Dewhurst, R. D.; Radacki, K.; Damme, A. Controlled homocatenation of boron on a transition metal. Nat. Chem. 2012, 4 (7), 563−567. (19) Jiang, M.; Zhu, J.; Chen, C.; Lu, Y.; Ge, Y.; Zhang, X. Poly(vinyl Alcohol) Borate Gel Polymer Electrolytes Prepared by Electrodeposition and Their Application in Electrochemical Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8 (5), 3473−81. (20) Gao, W.; Jing, Y.; Yang, J.; Zhou, Z.; Yang, D.; Sun, J.; Lin, J.; Cong, R.; Yang, T. Open-framework gallium borate with boric and metaboric acid molecules inside structural channels showing photocatalysis to water splitting. Inorg. Chem. 2014, 53 (5), 2364−6. (21) Pan, C. Y.; Zhong, L. J.; Zhao, F. H.; Luo, Y. Z.; Li, D. G. Cotemplating synthesis of a rare borate Ni(en) 3 .Hen. [B9O13(OH)4].H2O with a chiral polyanonic chain constructed from two different clusters. Inorg. Chem. 2015, 54 (2), 403−5. (22) Litters, S.; Kaifer, E.; Enders, M.; Himmel, H. J. A boron−boron coupling reaction between two ethyl cation analogues. Nat. Chem. 2013, 5 (12), 1029−1034.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.7b01433. Tafel measurement for carbon steel in SRCW after various immersion times at pH 5 (Figure S1), pH 7 (Figure S2), and pH 9 (Figure S3); corrosion data obtained from Tafel measurements (Table S1); Nyquist plots for carbon steel in SRCW in the absence (and presence) of the inhibitor after various immersion times at pH 5, pH 7, and pH 9 (Figure S4); Bode impedance plots for carbon steel in SRCW in the absence (and presence) of the inhibitor after various immersion times at pH 5, pH 7, and pH 9 (Figure S5); Bode phase-angle plots for carbon steel in SRCW in the absence (and presence) of the inhibitor after various immersion times at pH 5, pH 7, and pH 9 (Figure S6); corrosion data obtained from EIS measurements (Table S2) (PDF)



AUTHOR INFORMATION

Corresponding Authors

*Tel./Fax: +86-10-58801830. E-mail: [email protected]. *Tel./Fax: +86-10-68322124. E-mail: [email protected]. cn. ORCID

Yuansheng Pei: 0000-0002-0264-1974 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Natural Science Foundation of China (No. 51579009) and the Beijing Municipal Science and Technology Plan projects (No. Z161100001216009).



REFERENCES

(1) Rahmani, K.; Jadidian, R.; Haghtalab, S. Evaluation of inhibitors and biocides on the corrosion, scaling and biofouling control of carbon steel and copper−nickel alloys in a power plant cooling water system. Desalination 2016, 393, 174−185. (2) Cheng, Q.; Wang, S.; Yan, C. Sequential Monte Carlo simulation for robust optimal design of cooling water system with quantified uncertainty and reliability. Energy 2017, 118, 489. 7250

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research

nanoparticles and analysized by Mott-Schottky capacitance. Org. Electron. 2014, 15 (8), 1745−1752. (43) Gadala, I. M.; Alfantazi, A. A study of X100 pipeline steel passivation in mildly alkaline bicarbonate solutions using electrochemical impedance spectroscopy under potentiodynamic conditions and Mott−Schottky. Appl. Surf. Sci. 2015, 357, 356−368. (44) BenSalah, M.; Sabot, R.; Triki, E.; Dhouibi, L.; Refait, P.; Jeannin, M. Passivity of Sanicro28 (UNS N-08028) stainless steel in polluted phosphoric acid at different temperatures studied by electrochemical impedance spectroscopy and Mott−Schottky analysis. Corros. Sci. 2014, 86, 61−70. (45) Fattah-alhosseini, A.; Vafaeian, S. Comparison of electrochemical behavior between coarse-grained and fine-grained AISI 430 ferritic stainless steel by Mott−Schottky analysis and EIS measurements. J. Alloys Compd. 2015, 639, 301−307. (46) Taveira, L. V.; Montemor, M. F.; Da Cunha Belo, M.; Ferreira, M. G.; Dick, L. F. P. Influence of incorporated Mo and Nb on the Mott− Schottky behaviour of anodic films formed on AISI 304L. Corros. Sci. 2010, 52 (9), 2813−2818. (47) Feng, Z.; Cheng, X.; Dong, C.; Xu, L.; Li, X. Passivity of 316L stainless steel in borate buffer solution studied by Mott−Schottky analysis, atomic absorption spectrometry and X-ray photoelectron spectroscopy. Corros. Sci. 2010, 52 (11), 3646−3653. (48) Liu, L.; Xu, J.; Xie, Z. H.; Munroe, P. The roles of passive layers in regulating the electrochemical behavior of Ti5Si3-based nanocomposite films. J. Mater. Chem. A 2013, 1, 2064−2078. (49) Nicic, I.; Macdonald, D. D. The passivity of Type 316L stainless steel in borate buffer solution. J. Nucl. Mater. 2008, 379 (1−3), 54−58. (50) Mills, D.; Picton, P.; Mularczyk, L. Developments in the electrochemical noise method (ENM) to make it more practical for assessment of anti-corrosive coatings. Electrochim. Acta 2014, 124, 199− 205. (51) Ramezanzadeh, B.; Arman, S. Y.; Mehdipour, M.; Markhali, B. P. Analysis of electrochemical noise (ECN) data in time and frequency domain for comparison corrosion inhibition of some azole compounds on Cu in 1.0 M H2SO4 solution. Appl. Surf. Sci. 2014, 289, 129−140. (52) Rios, E. C.; Zimer, A. M.; Mendes, P. C. D.; Freitas, M. B. J.; de Castro, E. V. R.; Mascaro, L. H.; Pereira, E. C. Corrosion of AISI 1020 steel in crude oil studied by the electrochemical noise measurements. Fuel 2015, 150, 325−333. (53) Rios, E. C.; Zimer, A. M.; Pereira, E. C.; Mascaro, L. H. Analysis of AISI 1020 steel corrosion in seawater by coupling electrochemical noise and optical microscopy. Electrochim. Acta 2014, 124, 211−217. (54) Jamali, S. S.; Mills, D. J.; Sykes, J. M. Measuring electrochemical noise of a single working electrode for assessing corrosion resistance of polymer coated metals. Prog. Org. Coat. 2014, 77 (3), 733−741. (55) Casajús, P.; Winzer, N. Electrochemical noise analysis of the corrosion of high-purity Mg−Al alloys. Corros. Sci. 2015, 94, 316−326. (56) Fattah-alhosseini, A.; Soltani, F.; Shirsalimi, F.; Ezadi, B.; Attarzadeh, N. The semiconducting properties of passive films formed on AISI 316 L and AISI 321 stainless steels: A test of the point defect model (PDM). Corros. Sci. 2011, 53, 3186−3192. (57) Zhang, S. D.; Wu, J.; Qi, W. B.; Wang, J. Q. Effect of porosity defects on the long-term corrosion behaviour of Fe-based amorphous alloy coated mild steel. Corros. Sci. 2016, 110, 57−70. (58) Papadias, D. D.; Ahluwalia, R. K.; Thomson, J. K.; Meyer, H. M., III; Brady, M. P.; Wang, H.; Turner, J. A.; Mukundan, R.; Borup, R. Degradation of SS316L bipolar plates in simulated fuel cell environment: Corrosion rate, barrier film formation kinetics and contact resistance. J. Power Sources 2015, 273, 1237−1249. (59) Rougieux, F. E.; Grant, N. E.; Macdonald, D. Impact of Grown-in Point-defects on the Minority Carrier Lifetime in Czochralski-grown Silicon Wafers. Energy Procedia 2014, 60, 81−84 (special issue: Advanced Materials and Characterization Techniques for Solar Cells II). (60) Sazou, D.; Saltidou, K.; Pagitsas, M. Understanding the effect of bromides on the stability of titanium oxide films based on a point defect model. Electrochim. Acta 2012, 76, 48−61. (61) Geringer, J.; Macdonald, D. D. Modeling fretting-corrosion wear of 316L SS against poly(methyl methacrylate) with the Point Defect

(23) Xu, B.; Yang, W.; Liu, Y.; Yin, X.; Gong, W.; Chen, Y. Experimental and theoretical evaluation of two pyridinecarboxaldehyde thiosemicarbazone compounds as corrosion inhibitors for mild steel in hydrochloric acid solution. Corros. Sci. 2014, 78, 260−268. (24) Yusufali, C.; Kshirsagar, R. J.; Jagannath; Mishra, R. K.; Dutta, R. S.; Dey, G. K. Infrared and X-ray photoelectron spectroscopy studies on sodium borosilicate glass interacted with thermally oxidized aluminides formed on Alloy 690. J. Non-Cryst. Solids 2013, 366, 54−58. (25) Miura, Y.; Kusano, H.; Nanba, T.; Matsumoto, S. X-ray photoelectron spectroscopy of sodium borosilicate glasses. J. NonCryst. Solids 2001, 290 (1), 1−14. (26) Liu, H.; Xu, D.; Dao, A. Q.; Zhang, G.; Lv, Y.; Liu, H. Study of corrosion behavior and mechanism of carbon steel in the presence of Chlorella vulgaris. Corros. Sci. 2015, 101, 84−93. (27) Moradi, M.; Song, Z.; Tao, X. Introducing a novel bacterium, Vibrio neocaledonicus sp., with the highest corrosion inhibition efficiency. Electrochem. Commun. 2015, 51, 64−68. (28) Xu, H.; Sun, D.; Yu, H. Repassivation behavior of 316L stainless steel in borate buffer solution: Kinetics analysis of anodic dissolution and film formation. Appl. Surf. Sci. 2015, 357 (Part A), 204−213. (29) El-Moneim, A. A.; Akiyama, E.; Ismail, K. M.; Hashimoto, K. Corrosion behaviour of sputter-deposited Mg−Zr alloys in a borate buffer solution. Corros. Sci. 2011, 53 (9), 2988−2993. (30) Solomon, M. M.; Umoren, S. A. In-situ preparation, characterization and anticorrosion property of polypropylene glycol/silver nanoparticles composite for mild steel corrosion in acid solution. J. Colloid Interface Sci. 2016, 462, 29−41. (31) Shubha, H. N.; Venkatesha, T. V.; Vathsala, K.; Pavitra, M. K.; Punith Kumar, M. K. Preparation of Self Assembled Sodium Oleate Monolayer on Mild Steel and Its Corrosion Inhibition Behavior in Saline water. ACS Appl. Mater. Interfaces 2013, 5 (21), 10738−10744. (32) Zhang, W.; Ji, G.; Bu, A.; Zhang, B. Corrosion and Tribological Behavior of ZrO2 Films Prepared on Stainless Steel Surface by the Sol− Gel Method. ACS Appl. Mater. Interfaces 2015, 7 (51), 28264−28272. (33) Chong, A. L.; Mardel, J. I.; MacFarlane, D. R.; Forsyth, M.; Somers, A. E. Synergistic Corrosion Inhibition of Mild Steel in Aqueous Chloride Solutions by an Imidazolinium Carboxylate Salt. ACS Sustainable Chem. Eng. 2016, 4 (3), 1746−1755. (34) Withers, N. Main group chemistry: Boron and flat. Nat. Chem. 2012, 4, 147−147. (35) Sarkar, N.; Sahoo, G.; Das, R.; Prusty, G.; Sahu, D.; Swain, S. K. Anticorrosion Performance of Three-Dimensional Hierarchical PANI@ BN Nanohybrids. Ind. Eng. Chem. Res. 2016, 55 (11), 2921−2931. (36) Ahmed, M. G.; Kretschmer, I. E.; Kandiel, T. A.; Ahmed, A. Y.; Rashwan, F. A.; Bahnemann, D. W. A Facile Surface Passivation of Hematite Photoanodes with TiO2 Overlayers for Efficient Solar Water Splitting. ACS Appl. Mater. Interfaces 2015, 7 (43), 24053−24062. (37) Zhang, K.; Xu, B.; Yang, W.; Yin, X.; Liu, Y.; Chen, Y. Halogensubstituted imidazoline derivatives as corrosion inhibitors for mild steel in hydrochloric acid solution. Corros. Sci. 2015, 90, 284−295. (38) Sheng, X.; Cai, W.; Zhong, L.; Xie, D.; Zhang, X. Synthesis of Functionalized Graphene/Polyaniline Nanocomposites with Effective Synergistic Reinforcement on Anticorrosion. Ind. Eng. Chem. Res. 2016, 55 (31), 8576−8585. (39) Gopi, D.; Sherif, E.S. M.; Manivannan, V.; Rajeswari, D.; Surendiran, M.; Kavitha, L. Corrosion and Corrosion Inhibition of Mild Steel in Groundwater at Different Temperatures by Newly Synthesized Benzotriazole and Phosphono Derivatives. Ind. Eng. Chem. Res. 2014, 53, 4286−4294. (40) Umoren, S. A.; Gasem, Z. M.; Obot, I. B. Natural Products for Material Protection: Inhibition of Mild Steel Corrosion by Date Palm Seed Extracts in Acidic Media. Ind. Eng. Chem. Res. 2013, 52, 14855− 14865. (41) Xiong, D.; Zhang, Q.; Verma, S. K.; Bao, X. Q.; Li, H.; Zhao, X. Crystal structural, optical properties and Mott−Schottky plots of p-type Ca doped CuFeO2 nanoplates. Mater. Res. Bull. 2016, 83, 141−147. (42) Xiong, J.; Yang, J.; Yang, B.; Zhou, C.; Hu, X.; Xie, H.; Huang, H.; Gao, Y. Efficient and stable inverted polymer solar cells using TiO2 7251

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252

Article

Industrial & Engineering Chemistry Research Model: Fundamental theory, assessment, and outlook. Electrochim. Acta 2012, 79, 17−30. (62) Sharifi-Asl, S.; Taylor, M. L.; Lu, Z.; Engelhardt, G. R.; Kursten, B.; Macdonald, D. D. Modeling of the electrochemical impedance spectroscopic behavior of passive iron using a genetic algorithm approach. Electrochim. Acta 2013, 102, 161−173.

7252

DOI: 10.1021/acs.iecr.7b01433 Ind. Eng. Chem. Res. 2017, 56, 7239−7252