Chapter 20

Monitoring Corrosion Using Vibrational Spectroscopic Techniques

Shengxi Li    Hawaii Corrosion Laboratory, Department of Mechanical Engineering, University of Hawaii at Manoa, Honolulu, Hawaii, USA

Abstract

Raman spectroscopy and Fourier transform infrared spectroscopy are two useful spectroscopic techniques for corrosion studies especially because they can provide real-time information on various corrosion systems. This chapter discusses the applications of these two techniques in the studies of aqueous corrosion, atmospheric corrosion, corrosion inhibitors, and corrosion protection coatings. Available techniques and instrumentation for each application will be provided as a guideline for related investigations. Various examples from literature are discussed to show the power of these two techniques in corrosion studies. Specifically, these two techniques found a wide application in corrosion protection coating systems, including monitoring the growth of coatings, studying their degradation in corrosive environments, and characterization of the corrosion products.

Keywords

Raman spectroscopy

FTIR

Aqueous corrosion

Atmospheric corrosion

Inhibitors

Coatings

Acknowledgments

The author is grateful for the support from College of Engineering, University of Hawaii at Manoa. The author is particularly grateful to Prof. Lloyd Hihara, director of the Hawaii Corrosion Laboratory at the Department of Mechanical Engineering, University of Hawaii at Manoa.

20.1 Introduction

Spectroscopic techniques such as Raman spectroscopy and Fourier transform infrared spectroscopy (FTIR) have been widely used in corrosion research. Both in situ and ex situ experiments using these two vibrational techniques provide valuable information about the corrosion system. Extensive literature search resulted in a considerable number of papers published on corrosion studies using Raman and IR spectroscopy. However, only in situ applications of the two techniques in corrosion studies will be covered here due to their ability to acquire real-time information of the corrosion systems that allows monitoring the corrosion process.

In the following sections, the underlying theory of the two vibrational techniques will be first given. Then, the available techniques and instrumentation for in situ studies will be discussed. Finally, various examples of applications are presented to demonstrate the power of these two techniques for corrosion science and related fields of research. The discussion of literature papers will be divided into four areas: general aqueous corrosion, atmospheric corrosion, corrosion inhibitors, and coatings.

20.2 Principles

20.2.1 Raman spectroscopy

When a monochromatic light is scattered from a molecule, most photons are elastically scattered with the same frequency (energy) as the incident photons. A small fraction of light is scattered at frequencies different from those of the incident photons, namely inelastic scattering. The elastic scattering is named Rayleigh scattering while the inelastic scattering is called Raman scattering, or Raman effect. The Raman effect was discovered in 1928 by V. C. Raman.

The Raman effect arises when incident photons interact with the electric dipole of a molecule. The photons excite the molecule from its ground state to a virtual energy state. The molecule emits a photon when it relaxes and returns to a different energy state. The energy difference between the two states results in the energy of the incident photons being shifted away from the excitation wavelength. If the final energy state of the molecule is higher than the initial state, then the emitted photon will be shifted to a lower frequency, which is designated as Stokes shift. If the final energy state is lower than the initial state, then the emitted photon will be shifted to a higher frequency, and this is designated as an anti-Stokes shift (Figure 20.1). The Stokes shift is what is usually observed in Raman spectroscopy.

f20-01-9780124114678
Figure 20.1 Schematic representation of the energy changes associated with Raman scattering, Rayleigh scattering, and IR absorption.

The Raman effect is based on molecular deformations in the electric field (E) generated by the laser beam. Assuming that the laser is fluctuating at frequency ν, E can be written as a function of the amplitude, E0, and the time, t:

E=E0cos(2πνt)

si1_e  (20.1)

The dipole moment (μ) of a molecule induced by this light is, therefore:

μ=αE=αE0cos(2πνt)

si2_e  (20.2)

Here, the proportionality constant α is called the polarizability of the molecule. The polarizability represents the ease with which the electron cloud around a molecule can be deformed.

Similarly, if the molecule is also vibrating or rotating in sinusoidal fashion with frequency νm, the nuclear displacement Δr can be written as:

Δr=r0cos(2πνmt)

si3_e  (20.3)

where r0 is the vibrational amplitude.

When a molecule is at its equilibrium nuclear geometry, the polarizability of the molecule is α0. While at some distance, the instantaneous polarization α is a linear function of the nuclear displacement, Δr:

a=a0+arΔr

si4_e  (20.4)

where the derivative (∂a/∂r) represents the change in polarizability with change in position.

Combining Equations (20.2)(20.4), we have:

μ=αE=αE0cos2πνt=α0E0cos2πνt+arr0E0cos2πνmtcos2πνt=α0E0cos2πνt+12arr0E0cos2πtννm+cos2πtν+νm

si5_e  (20.5)

The first term relates to the scattered photon that has the same frequency as the incident photon, and thus explains Rayleigh scattering. The second term relates to the scattered photons with increased frequency (ν − νm) and decreased frequency (ν + νm) as compared to the incident photon, representing Stokes shift and anti-Stokes shift, respectively.

Equation (20.5) also gives a general selection rule that governs Raman scattering. Notice the derivative (∂a/∂r) in the second term. If that derivative equals zero, the entire second term will be zero, and there is no Raman scattering. Therefore, the selection rule for Raman scattering can be given as:

ar0

si6_e  (20.6)

that is, a molecular motion will be Raman-active only if the motion occurs with a change in polarizability. The amount of the polarizability change determines the Raman scattering intensity.

20.2.2 IR spectroscopy

Infrared radiation encompasses a section of the electromagnetic spectrum with wavenumbers (ν¯si7_e) from 12,800 to 10 cm− 1 or wavelengths (λ) from 0.78 to 1000 μm. This infrared portion of the electromagnetic spectrum is usually divided into three subregions: near-IR region (12,800-4000 cm− 1), mid-IR region (4000-200 cm− 1), and far-IR region (200-10 cm− 1). The mid-IR region is the most commonly used region in IR spectroscopy.

Infrared spectroscopy is based on the interaction between IR radiation and molecules that results in the absorption of IR radiation by the molecules. At temperatures above absolute zero, all the atoms in molecules are in continuous vibration with respect to each other. When the frequency of a specific vibration equals the frequency of the incident IR radiation, the molecules absorb the radiation. The absorbed radiation matches the transition energy of the bond or group that vibrates in the molecule.

A molecule can vibrate in many ways, with each way called a vibrational mode. Simple diatomic molecules have only one bond and thus one vibrational mode that is not observed if the molecule is symmetrical. Polyatomic molecules have more bonds and correspondingly more vibrational modes. The total number of fundamental vibrational modes can be determined as follows. A polyatomic molecule with n atoms has 3n total degrees of freedom because each atom has three degrees of freedom. For a nonlinear molecule, 3 degrees of freedom are rotational and 3 are translational. Therefore, there are 3n − 6 degrees of freedom that are fundamental vibrations for nonlinear molecules. For linear molecules, because only 2 degrees are rotational, the total number of fundamental vibrations is 3n − 5. However, not all the vibrations can be observed in IR spectra. For a mode to be observed, changes must occur in the permanent dipole:

μr0

si8_e  (20.7)

where μ is the dipole moment and r is the normal coordinate. In general, the larger the dipole change, the stronger the intensity (I) of the band in an IR spectrum:

I=μ/r2

si9_e  (20.8)

A comparison of Raman and IR spectroscopy is shown in Table 20.1. The major difference between the two techniques lies in that Raman spectroscopy determines vibrations that cause a change in the polarizability of a molecule, while IR spectroscopy determines vibrations that cause a change in the dipole moment of a molecule. However, in theory, there are some vibrations that induce changes in both polarizability and dipole moment.

Table 20.1

Comparison of Raman and IR Spectroscopy

RamanIR
Originate from scattering of radiationOriginate from absorption of radiation
Require change in molecular polarizabilityRequire change in dipole moment
Water compatibleWater leads to intense absorption
Record by using a beam of monochromatic radiationRecord by using a beam of radiation having a large number of frequencies
Optics can be made of quartz or glassOptics are usually made of salts (e.g., NaCl, KBr, and CsI)
Homonuclear diatomic molecules are Raman activeHomonuclear diatomic molecules are IR inactive

20.3 Methods and Equipment

20.3.1 Raman spectroscopy

There are many variations of Raman spectroscopy that have been developed to enhance the sensitivity (e.g., surface-enhanced Raman), to improve the spatial resolution (Raman microscopy), or to acquire very specific information (resonance Raman). For in situ studies in corrosion science, normal Raman spectroscopy is sufficient in most cases, but sometimes surface enhanced Raman spectroscopy (SERS) is necessary. Therefore, other techniques (e.g., resonance Raman) are not further covered here.

20.3.1.1 Normal Raman spectroscopy

In situ Raman spectroscopic studies of a corrosion system generally do not require specially designed equipment. Cells with or without flow configurations can be made using any material. When an enclosed cell is necessary, a quartz or even glass window can be used for passing the laser beam to the sample.

20.3.1.2 SERS

Surface-enhanced Raman spectroscopy is a technique that offers orders of magnitude increases in Raman scattering by molecules adsorbed on rough metal surfaces or nanostructures. Typical metals used for Raman scattering enhancement are Ag, Au, and Cu. To prepare the roughed surfaces, a conventional approach is to electrochemically rough the substrate through etching. Recently, surfaces are usually prepared by coating metallic nanoparticles on various substrates. There are two primary theories proposed for SERS: the electromagnetic theory and the chemical theory. The electromagnetic theory is based on the enhancement in the electric field provided by localized surface plasmons. The chemical theory proposes the formation of charge-transfer complexes and thus making it only applicable to chemisorbed species on metal surfaces.

20.3.2 IR spectroscopy

There are two primary experimental techniques in IR spectroscopy, transmission and reflection techniques. For in situ studies of corrosion phenomenon, reflectance spectroscopy is the commonly used technique. Among the various types of reflectance techniques, specular reflection, infrared reflection-absorption spectroscopy (IRAS), and attenuated total reflection (ATR) are well suitable for in situ studies.

20.3.2.1 Cells with optical windows for IR path

Because of the intense absorption of IR light by solvent (water), IR spectroscopy cannot be used as readily as Raman spectroscopy for in situ monitoring corrosion phenomenon in aqueous environments. Therefore, specific cells have to be designed with the aim of minimizing radiation losses by solvent absorption. One approach is by depositing very thin layers of electrolyte or small droplets on metal surfaces, which represents the common configuration for atmospheric corrosion studies. Figure 20.2 shows two typical experimental setups for in situ studies of atmospheric corrosion, with the one in Figure 20.2a for specular reflectance study and the one in Figure 20.2b for IRAS. For the IRAS studies, grazing angle incidence is usually used.

f20-02-9780124114678
Figure 20.2 Experimental configurations for in situ IR spectroscopic studies: 1.

For studies in bulk solutions, IR reflectance spectra can still be obtained by reducing the distance between the sample and the IR window (Figure 20.2c). The IR transparent window is pressed against the sample, creating an extremely thin solution layer in between them. One problem with this configuration is the possible depletion of solution species due to electrochemical activities (e.g., corrosion) in the confined environment. To solve this problem, a flow cell can be designed to replenish solution species as reaction proceeds.

For all three configurations in Figure 20.2, the IR windows are in direct contact with liquid electrolyte or high-humid environment. Therefore, non-water-soluble IR window materials have to be used, such as zinc selenide (ZnSe) and calcium fluoride (CaF2).

20.3.2.2 ATR

When an IR radiation enters from optically denser ATR crystals (refractive index n1) to the optically rarer samples (refractive index n2) at an angle exceeding the critical angle (θc), total internal reflection will occur inside the crystal. The critical angle is defined as:

θc=sin1n2n1

si10_e  (20.9)

This internal reflectance creates an evanescent wave that extends beyond the surface of the crystal into the sample that is in contact with the crystal. The evanescent wave protrudes only a few microns (0.5-5 μm) beyond the crystal surface and into the sample.

Some common ATR crystals are ZnSe, thallium bromide-thallium iodide (KRS-5), germanium (Ge), or silicon (Si). Prisms (Figure 20.3a) and hemispheres (Figure 20.3b) made of ATR crystals, for example, ZnSe, can be used for ATR-FTIR studies. In these configurations, IR beam is reflected once inside the crystal. Multiple internal reflection can be achieved by using the configuration shown in Figure 20.3c. In all three configurations, samples have to be in close contact with the crystals or can be plated on the crystals as thin films.

f20-03-9780124114678
Figure 20.3 Experimental configurations for in situ IR spectroscopic studies: 2 ATR techniques.

20.4 Applications of In Situ Raman Spectroscopy in Corrosion Science

In situ Raman spectroscopy has several major advantages for corrosion studies especially in aqueous environments. First, water does not interfere with Raman acquisition in aqueous environments. Second, the laser beam does not perturb the corrosion system under the protection of water, that is, without altering the corrosion species. This enables the study of those unstable compounds such as green rust, which would have transformed to other phases in the ambient. Third, the high surface roughness caused by corrosion does not hinder Raman acquisition, but instead enhances Raman scattering.

20.4.1 Aqueous corrosion

The corrosion of metallic materials in aqueous environments is one of the most prevalent forms of corrosion and has been extensively studied over the past few decades. In situ Raman spectroscopy is well suited for monitoring the corrosion process in metal/liquid interface. A great number of in situ Raman spectroscopic studies have been conducted to characterize the anodic corrosion films or monitor the formation of corrosion products on metal surfaces in corrosive environments.

20.4.1.1 Anodic oxide film formation

The passive oxide films on metals enable the self-protection of metals against corrosion in aqueous environments. A series of work has focused on in situ Raman spectroscopic identification of oxide films grown on various metal surfaces, including Fe, Ni, Co, Ag, Ti, Pb, and stainless steel. Most of the oxide films studied were either formed by anodic polarization115 or by exposing the metals to hydrothermal environments.1620

At anodic potentials, oxide films grow on metal surfaces; while at cathodic potentials, the oxide films dissolve. During this process, the formation, transformation (if any), and dissolution of the oxide films could be followed using in situ Raman spectroscopy. For example, according to in situ Raman data, the oxide film on Ni was found to be Ni2O3 (477 and 555 cm− 1) in 0.05 M NaOH solution2 and NiO (500 and 555 cm− 1) in concentrated (1-10 M) H2SO4 solution.3 On Cu in 0.1 M NaOH, Cu2O was detected for potentials from 50 to about 500 mV during anodic sweeps, but was converted to Cu(OH)2 at more anodic potentials (> 637 mV), as evidenced by the appearance of the band at ~ 488 cm− 1.6,7 The oxide film on Ti in acidic sulfuric solution was found to be anatase-type TiO2 (145, 400, 515, and 640 cm− 1), which changed from amorphous to crystalline as the potential exceeded a particular value.8 The protective oxide films also grow on Fe polarized in alkaline solutions and were determined to consist mainly of magnetite (Fe3O4) (550 and 670 cm− 1) in the inner layer.1113,15 An outer layer consisting of goethite (α-FeOOH) or lepidocrocite (γ-FeOOH) was also detected. The in situ capabilities of Raman spectroscopy were also extended to monitoring the surface films on iron and steel in high-temperature environments,1619 and various phases of iron oxides and oxyhydroxides were detected depending on the temperature.

20.4.1.2 General aqueous corrosion

Similar to monitoring the oxide film formation using in situ Raman spectroscopy, the general aqueous corrosion has been widely studied using this approach in order to identify the corrosion products formed during the corrosion process. Most of the work has been focused on iron and steel (e.g., galvanized steel, carbon steel, and stainless steel). Earlier work has characterized the corrosion products formed on galvanized steel during exposure to corrosive solutions.21,22 Zinc hydroxycarbonates and hydroxychlorides were successfully identified as the two primary corrosion products. For the corrosion of iron and steel in corrosive media, especially chlorinated solutions, nonstable green rust has been detected as the corrosion products generated during pitting.2330 Two major Raman bands at ~ 433 and 507 cm− 1 for the green rust were assigned to the Fe2 +glyph_sbndOH and Fe3 +glyph_sbndOH stretching modes, respectively. Depending on the corrosive media, the green rust can have different ions integrated in it, such as Cl, CO32 −, SO42 −, and HCOO. Other types of rust, such as Fe3O4 and γ-FeOOH, were also detected together with green rust. Siderite (FeCO3) is another corrosion product that forms on steel in concentrated carbonate/bicarbonate solutions, simulating groundwater.31,32

Other corrosion phenomena of metals that have been investigated using in situ Raman spectroscopy include molybdenum33 and tin34 in NaCl and KOH solutions, zirconium-niobium alloy35 and Ni-based alloy 60036 under hydrothermal conditions, and Pt-Ni alloys in HCl solution.37

20.4.2 Atmospheric corrosion

Atmospheric corrosion proceeds under thin electrolyte layers or small droplets that form by condensation or absorption of water by airborne sea salt particles. The small amount of electrolyte facilitates the use of Raman spectroscopy to monitor the corrosion process in situ. Corrosion of zinc caused by predeposited NaCl particles with a density of ~ 0.4 μg/cm2 serves as an example.38 Zinc chloride (ZnCl2) with a Raman band at 290 cm− 1 was detected on the Zn surface 2 h after the sample was exposed to high humidity (84%). After 6 h exposure, when ZnCl2 in the surface layer reached saturation, simonkollite (ZnCl2(Zn[OH]2)4) with major Raman bands at 255 and 390 cm− 1 was detected. The corrosion process under a single droplet was also studied for Zn39 and carbon steel.4042 The Raman spectra from Zn under a seawater droplet showed the formation of ZnO, a sulfate-containing and a carbonate-containing compound.39 The Raman spectra from carbon steel under single NaCl or Na2SO4 droplets indicated green rust and γ-FeOOH as the initial rust formation during sea salt droplets-induced marine atmospheric corrosion.4042 The spatial distribution of the corrosion reaction products was also revealed by in situ Raman spectroscopy, as shown in Figure 20.4.41 Ferrous chloride (FeCl2) and also possibly its chloride ions (245 and 286 cm− 1) were detected inside the corrosion initiation site or the corrosion pit, as the corrosion reaction products (Figure 20.4a). Green rust (GR1(Cl)) (425 and 501 cm− 1) was identified as the corrosion product formed in the vicinity immediately outside of the pit, while γ-FeOOH (246, 377, 525, and 645 cm− 1) was detected in regions outside of the GR region (Figure 20.4b and d). In the transitional region from GR to γ-FeOOH, Fe3O4 (666 cm− 1) also formed together with GR and γ-FeOOH (Figure 20.4c).

f20-04-9780124114678
Figure 20.4 In situ Raman spectra from different locations on steel under the NaCl droplet after 30 min corrosion: (a) corrosion initiation site or the corrosion pit, (b) region surrounding the corrosion initiation site, (c) rust cluster formed in the transitional region from GR to lepidocrocite, and (d) yellowish rust cluster that formed far away from the corrosion initiation site. Reproduced with permission from Ref. [41].

20.4.3 Corrosion inhibitors

Corrosion inhibitors are chemical compounds that are added to the surroundings (liquids or gases) of metallic materials to decrease their corrosion. Inhibitors work by the formation of a coating on metals, which prevents access of corrosive substance to the metal surfaces. The inhibitor molecules either attach to the metal or react with the surface to form a thin, adherent layer. In both cases, in situ Raman spectroscopy has been employed to study the adsorption mechanism and the film formation mechanism.

Benzotriazole (BTAH) has been used as an effective corrosion inhibitor for Cu and its alloys since the 1950s,43 and was also extended to other metals. The high corrosion inhibition efficiency of BTAH is attributed to the formation of a compact, polymerlike metal-BTAH complex on the metal surface.4446 The structure and composition of this complex can be revealed by in situ SERS, which provides information on even only a few monolayers of adsorbed species. Most of the in situ SERS studies on BTAH as corrosion inhibitors have been focused on Cu and Ag due to their strong Raman-enhancing effect. The inhibition effect of BTAH on Cu has been studied in various environments, such as halide solutions,44,4648 sulfuric acid media,4952 organic solution,53 and ionic liquids.54

Earlier studies showed that, in near neutral KCl solution, the complex formed by the reaction between Cu and BTAH was [Cu(I)BTA] according to in situ Raman data. In KCl/acid solution, the Raman spectra obtained at less negative potentials resembled that of [Cu(I)BTA] complex but, at more positive potentials the spectra were similar to that of [Cu(I)-ClBTAH]4.44 A recent study reported that, in neutral KCl solution, the initially formed complex was [CuI(BTA)]n and might have transformed to [CuI(BTAH)]4 at more negative potentials. This was evidenced by the redshifts of the in-plane trigonal breathing mode from 1041 cm− 1 at − 0.5 V to 1021 cm − 1 at − 1.1 V and the increase in intensity of the NH in-plane bending mode (1144 cm− 1) at more negative potentials (Figure 20.5).46 Therefore, it is probably necessary to take into account the effect of the presence of the [CuI(BTAH)]4 or [Cu(I)-ClBTAH]4 complexes in the corrosion inhibition of Cu in mild acidic KCl solution containing BTAH. In relatively strong acidic environment (e.g., H2SO4), [Cu(I)BTA]n complex was found as the film formed on the surface of Cu and its alloys.

f20-05-9780124114678
Figure 20.5 SERS spectra of Benzotriazole obtained from 0.1 M KCl at a Cu electrode at potentials indicated. Acquisition time, 10 s; accumulations, 2. Reprinted with permission from Ref. [46]. Copyright 2002 American Chemical Society.

Similar in situ SERS results were observed on Ag in either halide media or acetonitrile solution with BTAH.45,55 In addition, in situ SERS studies were also extended to Fe, Ni, and Co by proper electrochemical roughing procedures.46,56

Other inhibitors that have been studied using in situ SERS include triethylstibine for Fe and Ni,57 propargyl alcohol for Fe,58 2-mercaptobenzothiazole (MBT) for Cu,59 phytic acid (IP6) for Ag,60 volatile corrosion inhibitors for carbon steel,61 salicylate for Cu,62 and benzyldimethylphenylammonium chloride (BDMPAC) for low carbon steel,63 methimazole (MMI) for Cu,64 2-amino-5-mercapto-1,3,4-thiadiazole (AMT) for Ag65 and Co,66 3-amino-5-mercapto-1,2,4-triazole (AMTA) for Fe,67 and 2-amino-5-(4-pyridinyl)-1,3,4-thiadiazole (4-APTD) for Cu.68

20.4.4 Coatings

Coatings including both organic and inorganic coatings are an effective approach to achieve corrosion protections for metals. In situ Raman spectroscopy is a powerful tool for the study of corrosion-protection coatings. First, the formation of coatings on metallic materials could be monitored using in situ Raman spectroscopy. Second, the changes and degradation of coatings especially organic coatings upon exposure to corrosive environments can be studied using in situ Raman spectroscopy. Last, the corrosion products underneath the coatings can be characterized through in situ Raman spectroscopy provided that the coatings are thin and transparent.

20.4.4.1 Conversion coatings

Conversion coatings refer to the coatings produced by a chemical or electrochemical treatment of the metallic surface using coating solutions. During the conversion process, part of the metal surface is converted into a protective surface layer. Examples of conversion coatings include chromate conversion coating (CCC), phosphate conversion coating (PCC), and black oxide coating.

In situ Raman spectroscopy was successfully used to study CCC on aluminium alloys. Given that the Raman scattering from water is quite week, Raman spectroscopy could be used to monitor the formation process of CCC in aqueous solution.69 A band at approximately 858 cm− 1 appeared following the addition of Alodine solution (or K2CrO4, K2Cr2O7) to fresh Cr(OH)3 precipitate, indicating the formation of CCC. By monitoring the change of the band intensity at 858 cm− 1, the growth rates of the CCC as well as their dependence on coating bath composition were studied.70 It was found that the redox mediation action by Fe(CN)6− 3/− 4 is the mechanism for the acceleration of CCC formation.

In situ Raman spectroscopy was also used to monitor the release of chromate (CrO42 −) from CCC and its migration to neighboring regions of exposed alloy to protect them.7173 The chromate species redeposited as a film on fresh alloy surface, making the initially untreated alloy less active toward corrosion. The redeposition process occurred more rapidly in or near pits, showing the self-healing properties of CCC films.74 Besides pitting corrosion, filiform corrosion was also effectively prevented by CCC films.75

PCC was also studied using in situ Raman spectroscopy. The dissolution of phosphate layers on zinc-coated steel in dilute NaOH solution was monitored by measuring the decay of the 996 cm− 1 (PO43 −) band intensity, from which the rate of phosphate loss could be obtained.76

20.4.4.2 Polymeric coatings

Raman spectroscopy is especially suitable for studying polymers thanks to its high sensitivity towards the structure and conformation of the polymer backbone. Therefore, in situ Raman spectroscopy could be used to follow the growth of polymeric coatings on metals and their degradation in corrosive environments. In addition, this technique enables monitoring the corrosion product formation under the coatings provided that the coatings are very thin and transparent. The thin coating also reduces the time frame for corrosion to occur to match with that of Raman acquisition. Another requirement is the absence of fluorescence caused by the epoxy coating. Finally, it is better if the Raman bands of the coating do not interfere with those from the corrosion products.77

20.4.4.2.1 Epoxy coating

The earliest applications of in situ Raman spectroscopy in the study of polymer coatings were discussed in a series of papers on the corrosion of painted galvanized steel.7780 After being immersed in NaCl solutions for a certain time, the epoxy-polyamide (15-20 μm) coated galvanized steel developed two clearly different regions, with and without blisters. Two types of blisters were observed, black and large blisters and white and small blisters.77 Raman spectra of the corrosion products inside the blisters could be obtained directly through the thin paint layer. The Raman spectra from the nonblistered regions corresponded to epoxy-polyamide resin coating with characteristic bands at 570 and 637 cm− 1. The Raman spectra from inside the black blisters showed the presence of zinc hydroxychloride with sharp bands at 260 and 390 cm− 1, while those from the white blisters indicated the existence of amorphous zinc oxide with a prominent band at 569 cm− 1.7880 It was believed that the black blisters correspond to the cathodic zones and the white blisters are anodic. This study was also extended to the case of zinc-coated steel samples with chromate or phosphate conversion layers.79

20.4.4.2.2 Electronically conducting polymer coating

Recently, there has been an increasing interest in the use of electronically conducting polymers (ECPs) for corrosion resistant coatings. Some of the ECPs that have been investigated for corrosion protection include polyaniline (PANI), polypyrrole (PPy), and polythiophene (PT).

The use of PANI for steel corrosion protection was first reported in the early 1980s.81,82 Subsequently, a substantial amount of research has been devoted to the use of PANI for corrosion protection of various metals and alloys. The conventional approach to obtain the protective PANI layers was by electrochemical polymerization. In situ Raman spectroscopy has been extensively used to study the formation mechanisms of the PANI films,83,84 normally on platinum electrodes. Even though not directly related to corrosion study, these studies laid the foundation for studying the protective PANI coating on metals using in situ Raman spectroscopy.

In situ Raman spectroscopy combined with cyclic voltammetry was employed to investigate the interaction between the passive layer on iron and PANI coating in sulfate medium.85 The oxidation state of the PANI film on iron, which changed after voltammetric cycling, could be followed. After one redox cycle (0 to − 1.0 V), the PANI film deprotonation was verified by the depletion of the band at 1330 cm− 1 (Cglyph_sbndN+ stretching). Also, the depletion of benzenoid units was evident due to the decrease of the 1620 cm− 1 band (Cglyph_sbndC ring stretching) and the depletion of 1185 cm− 1 band (quinine ring Cglyph_sbndH bending).85 After further cycling, the Raman spectra evolved toward a less oxidized form of PANI with a predominance of benzenoid rings (1185 and 1620 cm− 1). The reduced PANI on iron is no longer able to provide enough electrons to maintain iron in its passive state, and the passive layers break down. They also analyzed the PANI layer potentiostatically polymerized in phosphoric/metanilic solution as compared to the PANI layer grown in inorganic acid (HCl) by collecting Raman spectra at different polarization potentials (e.g., from 150 to − 700 mV/SSE).86 They found that the film prepared in metanilic acid had a more oxidized and less protonated state than that grown in inorganic acid solution, as evidenced by the strong Raman bands at 1330 cm− 1 (Cglyph_sbndN) and 1500 cm− 1 (C = N) observed from the PANI film formed in metanilic acid. Therefore, the reduction of the PANI film is more difficult in the presence of metanilic acid, resulting in higher protective properties.

Another strategy for protecting metals using PANI is by forming epoxy and acrylic blends. The acrylic blend formed by PANI, poly(methyl methacrylate) (PMMA), and camphorsulfonic acid (CSA) is an example,8790 which had a low percolation threshold, a high conductivity, and enhanced mechanical properties as compared to PANI. Raman spectra of the PANI-PMMA-CSA blend indicated characteristic features of “secondary doped” PANI,91 with a significant increase of the band at 1333 cm− 1 corresponding to the C = N mode. The ratio of the band at 1333 cm− 1 and that at 1599 cm− 1 corresponding to the aromatic Cglyph_sbndC mode is equal to ~ 1, indicating that the film is in the emeraldine state. However, it diminished to ~ 0.7 after the coating (on iron) was immersed in H2SO4 solution for 20 days, suggesting the reduction of the polymeric coating.87 This ratio varies for different metal substrates, which indicates that the extent of the redox reaction in the film depends on the metals that have different reducing power (Figure 20.6).88 Raman data also showed the presence of a second layer formed by counter ions released from PANI and various metal cations.

f20-06-9780124114678
Figure 20.6 Intensity ratio of Raman bands located at 1333 and 1599 cm− 1 for PMMA20% PANI–CSA blend coating on different metals. Spectra were taken after 12 days in 1 mol L− 1 H2SO4 electrolytic solution. Reprinted from Ref. [88]. Copyright 2005, with permission from Elsevier.

Another ECP that has been widely used for corrosion protection coatings is PPy. The formation or the electropolymerization of the PPy films on various metals could be followed by using in situ Raman spectroscopy.9296 Raman bands characteristic of the oxidized form of PPy at 927, 1086, 1238, 1368, and 1605 cm− 1 were observed during the growth of PPy film, and those typical of the reduced form of PPy at 988, 1038, 1260, 1313, and 1564 cm− 1 were detected during corrosion in corrosive medium.94

Take the thin PPy film on iron electrode and immersed in a 0.1 M K2SO4 solution (pH = 4) as an example.97 Raman spectra were recorded during immersion and at different OCPs to monitor the redox states of the PPy and composite films. The Raman data revealed that the PPy film was in the oxidized form in the original state (Eoc = 0 V) and was protective to the iron substrate. Then, the OCP dropped to − 0.4 V during protection and finally to − 1.0 V after corrosion. During this process, the Raman bands at 1606 cm− 1 (inter-ring C = C) and 1385 cm− 1 (ring C = C) corresponding to the oxidized PPy both shifted towards lower wavenumbers, ~ 1570 and 1315 cm− 1, respectively, for the reduced form of PPy (Figure 20.7a). The bands for the defect modes at 929, 1410, and 1240 cm− 1 also decreased. The Raman bands which increased significantly during the OCP measurement are at 985 and 1050 cm− 1, which correspond to the reduced form of PPy. As compared to the PPy film alone, the PPy-PDAN composite film showed better ability to protect iron because the Raman bands that are typical of the oxidized form of PPy (929, 1240, 1410, and 1610 cm− 1) decreased more slowly upon reduction (Figure 20.7b).97

f20-07-9780124114678
Figure 20.7 In situ Raman spectra of (a) PPy/Fe electrode, (b) composite/Fe electrode in 0.1 M K2SO4 (pH 4), recorded at different Eoc: 0, − 0.4, and − 1 V. Reproduced with permission from Ref. [97].

20.5 Applications of In Situ FTIR in Corrosion Science

FTIR has been widely used for corrosion studies, both in situ and ex situ. An extensive treatment of both in situ and ex situ application of FTIR in corrosion studies is found in Ref. [98] In this chapter, only in situ application of FTIR for corrosion studies will be covered.

20.5.1 Aqueous corrosion

It is difficult to probe corrosion phenomenon in bulk solutions using in situ FTIR due to intense absorption of IR radiation by water. However, by reducing the distance between the sample surface and the IR window, the corrosion process on the metal surface could be followed. The electrochemically controlled oxidation and reduction of lead in sulfuric acid was successfully monitored using in situ IRAS.99 The growth and reduction of lead sulfate (PbSO4) was followed by monitoring the band at 631 cm− 1, the integrated area of which showed a linear relationship with the consumed charge in the potential region where PbSO4 forms. At more anodic potentials, the formation of lead dioxide (PbO2) was observed as evidenced by an increased absorption band at 5200 cm− 1. Similar experiments were conducted to characterize the corrosion products formed on Ni electrodes in aqueous solutions containing pseudohalide ion series (OCN, SCN, and SeCN).100 Ni corroded in all three solutions with the formation of Ni(II) cyanate complex, Ni(II) thiocyanate complex, and Ni(II) selenocyanate complex depending on the ion species in the solution. In the Ni/OCN system, the in situ data obtained using subtractively normalized interfacial Fourier transform infrared spectroscopy (SNIFTIRS) also showed the formation of CO2 (2343 cm− 1), which was believed to be the ultimate product of electrolysis of the cyanate ion.

Other in situ FTIR studies of corrosion-related phenomena include anion adsorption on colloidal chromium (III) oxide hydroxide film simulating passive stainless steel surface,101 characterization of the anodic film on iron in neutral phosphate solution (FeIIIPO4),102 and monitoring of the transport of corrosive species through the mortar layer.103

20.5.2 Atmospheric corrosion

Given that atmospheric corrosion occurs under extremely thin electrolyte layers, there is less absorption of IR radiation by water. Therefore, a great amount of work has been conducted to study atmospheric corrosion using in situ IRAS techniques. In situ IRAS has been developed to study the corrosion of metals (e.g., Cu, Zn, Al, and Mg) exposed to humid air104 or humid air mixed with corrosive gases, such as SO2,104110 NO2,107,109111 O3,109 and SO3.112 The corrosive environments were generated by passing air with humidity and certain levels of corrosive gases through the sample chamber. The interaction of metals with corrosive environments and the formation of corrosion products could be followed using in situ IRAS. For example, the initial corrosion products on Cu exposed to humid air (90% RH) were identified as copper(I) oxide while that formed in humid air containing SO2 (0.23 ppm) was sulfite.104 Similarly, the corrosion products formed on Ni,105 Zn,105 and Al112 exposed to SO2-containing humid air were also characterized using in situ IRAS. As complementary to in situ IRAS, quartz crystal microbalance (QCM)106,109,111 and atomic force microscopy (AFM)108,109 were the other two in situ techniques used for the study of humid air-induced atmospheric corrosion, to obtain the mass changes and morphological information, respectively.

The influence of organic constituents such as acetic acid, acetaldehyde, and formic acid on the atmospheric corrosion of metals, especially in indoor environments, was also studied using in situ IRAS.113117 Zinc acetate was detected as the major corrosion products in humid environments with sub-ppm concentration of acetic acid and acetaldehyde,113,114 while zinc formate formed on zinc exposed to environments containing formic acid.115

Atmospheric corrosion of metals is greatly enhanced by NaCl deposition, which can originate from either airborne sea salt particles or deicing salts. NaCl particle- induced atmospheric corrosion has been studied using in situ IRAS for Zn,118,119 Al,120 Fe,121 Cu,122,123 and Mg.124 For Zn with predeposited NaCl particles and then exposed to high humidity (> 90%), a surface film containing ZnO, Zn5(OH)8Cl2·H2O, and zinc Zn5(OH)6(CO3)2 was detected by in situ IRAS.118 The rate of the corrosion process could also be evaluated from the rate of surface film formation obtained from the IRAS measurements. In situ IRAS also detected carbonate ions (1390 cm− 1) in the thin solution films in the secondary spreading region close to the original NaCl droplet.119 Similar secondary spreading effects were observed on Cu and studied using in situ IRAS.122,123 Another important application of in situ IRAS in atmospheric corrosion studies is the investigation of filiform corrosion of Al120 and Fe.121 The movement of the active filament head on coated Al surface could be followed with in situ FTIR microspectroscopy using the characteristic IR band at ~ 2500 cm− 1 from Al(H2O)63 +, the corrosion product in the filament.120 Notice that the coating on Al must be thin and transparent in the spectral region of interest.

20.5.3 Corrosion inhibitors

In situ FTIR is a powerful tool for the investigation of the mechanism of corrosion inhibitors. In combination with electrochemical techniques, SNIFTIRS was used to study the formation of BTAH on Cu(100) surface.125 The Cu(I)BTA complex was detected on Cu surface at potentials >−0.3 V as evidenced by the negative-going bands at 1119 and 1155 cm− 1. The bands assigned to BTAH in the solution were observed as the positive-going bands at 1014, 1217, 1248, 1268, and 1308 cm− 1. The effects of anions (Cl and SO42 −) in the supporting electrolyte on the formation and decomposition of the Cu(I)BTA film on Cu were also investigated using in situ FTIR.126 It was found that the Cu(I)BTA complex was formed more easily in Cl-containing solution than in sulfate or bisulfate solutions. Other corrosion inhibitors that have been studied using in situ FTIR include BDMPAC, glutamic acid, triethylenetetramine (TETA), sodium tartrate, and sodium benzoate.127

20.5.4 Coatings

In situ FTIR studies of corrosion protection coating systems have been limited to polymeric coatings. Generally, two techniques can be used: FTIR-reflection spectroscopy and ATR-FTIR. While FTIR-reflection spectroscopy was employed to monitor the growth of organic coatings on metal surfaces, ATR was used mostly to study the transport of water and some ionic species through the organic coatings.

20.5.4.1 FTIR-reflection spectroscopy

Electrochemical cell with CaF2 windows was designed to study the formation of polymer films on metal surfaces.128,129 For the poly(thiophene-3-methanol) (PTOH) film on platinum electrodes, in situ FTIR spectra indicated that the perchlorate anions entered the PTOH film during the oxidation process as evidenced by the bands in the region of 1000-1500 cm− 1. When the film was reduced, the anions were expelled from the film because the band for the perchlorate anions decreased.128 In situ FTIR was also used to study the growth of poly(o-phenylenediamine) (PPD) film on stainless steel, and the results suggested a ladder-type structure with alternating pyrazine and phenazine rings.129 Similarly, in situ FTIR could be used to detect the solution species, that is, CO2, during the formation of a semi-passivating layer on Cu.62 The CO2 was believed to be from the decarboxilation of the salicylate anion.

20.5.4.2 ATR-FTIR

The adhesion between the polymeric coating and the metal substrate plays an important role in the protective properties of the coatings. Water that transports through the coating to the metal/coating interfaces is a major cause of degradation and delamination of the coating. In addition, water also promotes corrosion of the metal substrate.

ATR-FTIR has been extensively used to study the transport of water and/or ionic species through polymer films to metal/polymer interfaces.130139 FTIR in the multiple internal reflection mode (FTIR-MIR) was used to obtain information on the water layer at the organic coating/substrate interface.130132 The effects of water evident in the 3000-3650 and 1625-1645 cm− 1 regions were observed after the coating was exposed to water for a certain time. These features increased in intensity with time while those for the coating decreased. The fact that the intensity change of 3400 cm− 1 is proportionally related to the total amount of water in the coating enables a quantitative analysis of the water layer. By comparing the intensity changes of the water OH stretching band as a function of time, the rates at which water entered the coating/substrate interfacial region for three coating systems can be determined and compared.132 Normal ATR-FTIR using ZnSe crystal as substrate was also employed to study the transport kinetics of both water and inhibitor anion (HPO42 −).133,134

In a series of papers, ATR-FTIR with the Kretschmann configuration was applied for in situ studies of the transport of water and ionic species (thiocyanate ions) through a polymer film to metal/polymer interfaces, such as aluminium/polymer interface and conversion coated-zinc/polymer interface. The fast increase of the stretching and bending vibrations of water at 3450 and 1650 cm− 1 suggested a fast sorption of water in the coating and the aluminium/coating interface (Figure 20.8).135 Another band at 950 cm− 1 due to the vibrations of Al-O groups indicated the formation of corrosion products—aluminium oxide or hydroxide at the aluminium surface. In the thicyanate solution, a band at 2075 cm− 1 due to the C = N vibration of the thicyanate ion was detected after prolonged exposure. As complementary to ATR-FTIR, which provided information about a specific region, electrochemical impedance spectroscopy (EIS) was employed to take into account the whole system.137,138 The combination of ATR-FTIR and EIS was also extended to investigate the hidden interface between a conversion-coated zinc surface and a polymer coating upon exposure to an electrolyte.139

f20-08-9780124114678
Figure 20.8 ATR-FTIR spectra of a ZnSe element coated with a thin aluminium film and a polymer film in contact with ultra pure deionized water. The exposure times were 40, 163 min, 26 and 50 h, respectively. Reprinted from Ref. [135]. Copyright 2006, with permission from Elsevier.

20.6 Conclusion

This chapter demonstrated the potential of Raman and IR spectroscopy for in situ studies of corrosion-related phenomena. Both techniques provide valuable real-time information on various corrosion systems. However, literature survey showed that more in situ studies were conducted using Raman spectroscopy than IR spectroscopy because of less interference from water for Raman acquisition. In addition, special sample preparation and experimental setup are required for in situ FTIR studies, especially in aqueous solutions. Therefore, challenges exist for IR spectroscopy to be applicable to more corrosion-related in situ studies.

Among the several applications discussed in this chapter, in situ study of corrosion protection coating systems is a prevalent subject, especially in recent years. The studies have been focused on all the important aspects in coating systems, which include in situ monitoring the growth of coatings, their degradation in corrosive environments, and characterization of the corrosion products underneath the coating if coating delamination occurs.

References

1 Thibeau RJ, Brown CW, Goldfarb AZ, et al. Infrared and Raman spectroscopy of aqueous corrosion films on lead. J Electrochem Soc. 1980;127(1):37–44.

2 Melendres CA, Xu S. In situ laser Raman spectroscopic study of anodic corrosion films on nickel and cobalt. J Electrochem Soc. 1984;131(10):2239–2243.

3 Delichere P, Hugot-Le Goff A, Yu N. Identification by in situ Raman spectroscopy of the films grown during the polarization of nickel in sulfuric solutions. J Electrochem Soc. 1986;133(10):2106–2107.

4 Hugot-Le Goff A, Pallotta C. In situ Raman spectroscopy for the study of iron passivity in relation to solution composition. J Electrochem Soc. 1985;132(11):2805–2806.

5 Thanos ICG. In situ Raman and other studies of electrochemically oxidized iron and iron-9% chromium alloy. Electrochim Acta. 1986;31(7):811–820.

6 Hamilton JC, Farmer JC, Anderson RJ. In situ Raman spectroscopy of anodic films formed on copper and silver in sodium hydroxide solution. J Electrochem Soc. 1986;133(4):739–745.

7 Smith JM, Wren JC, Odziemkowski M, et al. The electrochemical response of preoxidized copper in aqueous sulfide solutions. J Electrochem Soc. 2007;154(8):C431–C438.

8 Ohtsuka T, Guo J, Sato N. Raman spectra of the anodic oxide film on titanium in acidic sulfate and neutral phosphate solutions. J Electrochem Soc. 1986;133(12):2473–2476.

9 Thierry D, Persson D, Leygraf C, et al. In-situ Raman spectroscopy combined with x-ray photoelectron spectroscopy and nuclear microanalysis for studies of anodic corrosion film formation on iron-chromium single crystals. J Electrochem Soc. 1988;135(2):305–310.

10 McMahon JJ, Ruther W, Melendres CA. In situ laser Raman spectroelectrochemical study of the corrosion of lead in dilute disodium sulfate solution at high temperature. J Electrochem Soc. 1988;135(3):557–562.

11 Melendres CA, Camillone III N, Tipton T. Laser Raman spectroelectrochemical studies of anodic corrosion and film formation on iron in phosphate solutions. Electrochim Acta. 1989;34(2):281–286.

12 Hugot-Le Goff A, Flis J, Boucherit N, et al. Use of Raman spectroscopy and rotating split ring disk electrode for identification of surface layers on iron in 1 M sodium hydroxide. J Electrochem Soc. 1990;137(9):2684–2690.

13 Johnston C. In situ laser Raman microprobe spectroscopy of corroding iron electrode surfaces. Vib Spectrosc. 1990;1(1):87–96.

14 Baek WC, Kang T, Sohn HJ, et al. In situ surface enhanced Raman spectroscopic study on the effect of dissolved oxygen on the corrosion film on low carbon steel in 0.01 M NaCl solution. Electrochim Acta. 2001;46(15):2321–2325.

15 Joiret S, Keddam M, Novoa XR, et al. Use of EIS, ring-disk electrode, EQCM and Raman spectroscopy to study the film of oxides formed on iron in 1 M NaOH. Cem Concr Compos. 2002;24(1):7–15.

16 Farrow RL, Nagelberg AS. Raman spectroscopy of surface oxides at elevated temperatures. Appl Phys Lett. 1980;36(12):945–947.

17 Maslar JE, Hurst WS, Bowers WJ, et al. In situ Raman spectroscopic investigation of aqueous iron corrosion at elevated temperatures and pressures. J Electrochem Soc. 2000;147(7):2532–2542.

18 Maslar JE, Hurst WS, Bowers Jr. WJ, et al. In situ Raman spectroscopic investigation of stainless steel hydrothermal corrosion. Corrosion. 2002;58(9):739–747.

19 Kumai CS, Devine TM. Oxidation of iron in 288°C, oxygen-containing water. Corrosion. 2005;61(3):201–218.

20 Maslar JE, Hurst WS, Bowers Jr. WJ, et al. In situ Raman spectroscopic investigation of nickel hydrothermal corrosion. Corrosion. 2002;58(3):225–231.

21 Bernard MC, Hugot-Le Goff A, Massinon D, et al. In situ Raman identification of corrosion products on galvanized steel sheets. Mater Sci Forum. 1992;111–112:617–620.

22 Bernard MC, Hugot-Le Goff A, Phillips N. In situ Raman study of the corrosion of zinc-coated steel in the presence of chloride. I. Characterization and stability of zinc corrosion products. J Electrochem Soc. 1995;142(7):2162–2167.

23 Boucherit N, Hugot-Le Goff A, Joiret S. Raman studies of corrosion films grown on Fe and Fe-6Mo in pitting conditions. Corros Sci. 1991;32(5–6):497–507.

24 Bocherit N, Hugot-Le Goff A, Joiret S. In situ Raman identification of stainless steels pitting corrosion films. Mater Sci Forum. 1992;111–112:581–587.

25 Boucherit N, Hugot-Le Goff A. Localized corrosion processes in iron and steels studied by in situ Raman spectroscopy. Faraday Discuss. 1992;94:137–147.

26 Bonin PML, Odziemkowski MS, Reardon EJ, et al. In situ identification of carbonate-containing green rust on iron electrodes in dolutions simulating groundwater. J Solution Chem. 2000;29(10):1061–1074.

27 Simard S, Odziemkowski M, Irish DE, et al. In situ micro-Raman spectroscopy to investigate pitting corrosion product of 1024 mild steel in phosphate and bicarbonate solutions containing chloride and sulfate ions. J Appl Electrochem. 2001;31(8):913–920.

28 Reffass M, Sabot R, Jeannin M, et al. Effects of NO2 ions on localised corrosion of steel in NaHCO3 + NaCl electrolytes. Electrochim Acta. 2007;52(27):7599–7606.

29 Reffass M, Sabot R, Jeannin M, et al. Effects of phosphate species on localised corrosion of steel in NaHCO3 + NaCl electrolytes. Electrochim Acta. 2009;54(18):4389–4396.

30 Barchiche C, Sabot R, Jeannin M, et al. Corrosion of carbon steel in sodium methanoate solutions. Electrochim Acta. 2010;55(6):1940–1947.

31 Lee CT, Qin Z, Odziemkowski M, et al. The influence of groundwater anions on the impedance behaviour of carbon steel corroding under anoxic conditions. Electrochim Acta. 2006;51(8–9):1558–1568.

32 Lee CT, Odziemkowski MS, Shoesmith DW. An in situ Raman-electrochemical investigation of carbon steel corrosion in Na2CO3∕NaHCO3, Na2SO4, and NaCl solutions. J Electrochem Soc. 2006;153(2):B33–B41.

33 Wang K, Li Y-S, He P. In situ identification of surface species on molybdenum in different media. Electrochim Acta. 1998;43(16–17):2459–2467.

34 Huang BX, Tornatore P, Li Y-S. IR and Raman spectroelectrochemical studies of corrosion films on tin. Electrochim Acta. 2000;46(5):671–679.

35 Maslar JE, Hurst WS, Bowers WJ, et al. In situ Raman spectroscopic investigation of zirconium-niobium alloy corrosion under hydrothermal conditions. J Nucl Mater. 2001;298(3):239–247.

36 Maslar JE, Hurst WS, Bowers WJ, et al. Alloy 600 aqueous corrosion at elevated temperatures and pressures: an in situ Raman spectroscopic investigation. J Electrochem Soc. 2009;156(3):C103–C113.

37 Chen S, Wu S, Zheng J, et al. Spectroscopic and morphological studies on the electrooxidation of Pt–Ni alloys in HCl solution. J Electroanal Chem. 2009;628(1–2):55–59.

38 Ohtsuka T, Matsuda M. In situ Raman spectroscopy for corrosion products of zinc in humidified atmosphere in the presence of sodium chloride precipitate. Corrosion. 2003;59(5):407–413.

39 Cole IS, Muster TH, Lau D, et al. Products formed during the interaction of seawater droplets with zinc surfaces. J Electrochem Soc. 2010;157(6):C213–C222.

40 Li S. Marine atmospheric corrosion initiation and corrosion products characterization. In: Mechanical engineering. Honolulu: University of Hawai'i at Manoa; 2010:205.

41 Li S, Hihara LH. In situ Raman spectroscopic study of NaCl particle-induced marine atmospheric corrosion of carbon steel. J Electrochem Soc. 2012;159(4):C147–C154.

42 Li S, Hihara LH. In situ Raman spectroscopic identification of rust formation in Evans’ droplet experiments. Electrochem Commun. 2012;18:48–50.

43 Allam N, Nazeer A, Ashour E. A review of the effects of benzotriazole on the corrosion of copper and copper alloys in clean and polluted environments. J Appl Electrochem. 2009;39(7):961–969.

44 Rubim J, Gutz IGR, Sala O, et al. Surface enhanced Raman spectra of benzotriazole adsorbed on a copper electrode. J Mol Struct. 1983;100:571–583.

45 Rubim JC, Gutz IGR, Sala O. Surface-enhanced Raman spectra of benzotriazole adsorbed on a silver electrode. J Mol Struct. 1983;101:1–6.

46 Cao PG, Yao JL, Zheng JW, et al. Comparative study of inhibition effects of Benzotriazole for metals in neutral solutions as observed with surface-enhanced Raman spectroscopy. Langmuir. 2002;18(1):100–104.

47 Kester JJ, Furtak TE, Bevolo AJ. Surface enhanced Raman scattering in corrosion science: benzotriazole on copper. J Electrochem Soc. 1982;129(8):1716–1719.

48 Thierry D, Leygraf C. Simultaneous Raman spectroscopy and electrochemical studies of corrosion inhibiting molecules on copper. J Electrochem Soc. 1985;132(5):1009–1014.

49 Da Costa S.L.F.A., Agostinho SML, Chagas HC, et al. Study of the inhibiting action of benzotriazole on copper corrosion in deaerated sulfuric acid containing ferric ions by the rotating disc electrode, fluorescence, and Raman spectroscopies. Corrosion. 1987;43(3):149–153.

50 Rubim JC. Surface enhanced Raman scattering (SERS) from benzotriazole adsorbed on brass electrodes. Chem Phys Lett. 1990;167(3):209–214.

51 Villamil RFV, Corio P, Agostinho SML, et al. Effect of sodium dodecylsulfate on copper corrosion in sulfuric acid media in the absence and presence of benzotriazole. J Electroanal Chem. 1999;472(2):112–119.

52 Maciel JM, Jaimes R.F.V.V., Corio P, et al. The characterisation of the protective film formed by benzotriazole on the 90/10 copper–nickel alloy surface in H2SO4 media. Corros Sci. 2008;50(3):879–886.

53 Yao J-L, Yuan Y-X, Gu R-A. Negative role of triphenylphosphine in the inhibition of benzotriazole at the Cu surface studied by surface-enhanced Raman spectroscopy. J Electroanal Chem. 2004;573(2):255–261.

54 Costa LAF, Breyer HS, Rubim JC. Surface-enhanced Raman scattering (SERS) on copper electrodes in 1-n-butyl-3-methylimidazolium tetrafluoroborate (BMI.BF4): the adsorption of benzotriazole (BTAH). Vib Spectrosc. 2010;54(2):103–106.

55 Yuan YX, Yang FZ, Morag CH, et al. The effect of triphenylphosphane on corrosion inhibition of benzotriazole at Ag electrode monitored by SERS in nonaqueous solution. Spectrochim Acta A Mol Biomol Spectrosc. 2013;105:184–191.

56 Gallant D, Pézolet M, Simard S. Inhibition of cobalt active dissolution by benzotriazole in slightly alkaline bicarbonate aqueous media. Electrochim Acta. 2007;52(15):4927–4941.

57 Saito N, Nishihara H, Aramaki K. The mechanism for corrosion protective film formation on iron and nickel in acid solutions with organo-antimony compounds. Corros Sci. 1992;33(8):1253–1265.

58 Aramaki K, Fujioka E. Surface-enhanced Raman scattering spectroscopy studies on the inhibition mechanism of propargyl alcohol for iron corrosion in hydrochloric acid. Corrosion. 1996;52(2):83–91.

59 Marconato JC, Bulhões LO, Temperini ML. A spectroelectrochemical study of the inhibition of the electrode process on copper by 2-mercaptobenzothiazole in ethanolic solutions. Electrochim Acta. 1997;43(7):771–780.

60 Yang H-F, Feng J, Liu Y-L, et al. Electrochemical and surface enhanced Raman scattering spectroelectrochemical study of phytic acid on the silver electrode. J Phys Chem B. 2004;108(45):17412–17417.

61 Tormoen G, Burket J, Dante JF, et al. Monitoring the adsorption of volatile corrosion inhibitors in real time with surface-enhanced Raman spectroscopy. Corrosion. 2006;62(12):1082–1091.

62 Batista EA, Temperini MLA. An in situ SERS and FTIRAS study of salicylate interaction with copper electrode. J Solid State Electrochem. 2007;11(11):1559–1565.

63 Bozzini B, Romanello V, Mele C, et al. A SERS investigation of carbon steel in contact with aqueous solutions containing BenzylDiMethylPhenylAmmonium Chloride. Mater Corros. 2007;58(1):20–24.

64 Pan Y-C, Wen Y, Xue L-Y, et al. Adsorption behavior of methimazole monolayers on a copper surface and its corrosion inhibition. J Phys Chem C. 2012;116(5):3532–3538.

65 Yang H, Sun X, Zhu J, et al. Surface enhanced Raman scattering, in situ spectro-electrochemical, and electrochemical impedance spectroscopic investigations of 2-amino-5-mercapto-1,3,4-thiadiazole monolayers at a silver electrode. J Phys Chem C. 2007;111(22):7986–7991.

66 Huo S-J, Zhu Q, Chu C-S, et al. Anticorrosive behavior of AMT on cobalt electrode: from electrochemical methods to surface-enhanced vibrational spectroscopy study. J Phys Chem C. 2012;116(38):20269–20280.

67 Sherif E.-S.M., Erasmus RM, Comins JD. In situ Raman spectroscopy and electrochemical techniques for studying corrosion and corrosion inhibition of iron in sodium chloride solutions. Electrochim Acta. 2010;55(11):3657–3663.

68 Pan Y-C, Wen Y, Guo X-Y, et al. 2-Amino-5-(4-pyridinyl)-1,3,4-thiadiazole monolayers on copper surface: observation of the relationship between its corrosion inhibition and adsorption structure. Corros Sci. 2013;73:274–280.

69 Xia L, McCreery RL. Chemistry of a chromate conversion coating on aluminum alloy AA2024-T3 probed by vibrational spectroscopy. J Electrochem Soc. 1998;145(9):3083–3089.

70 Xia L, McCreery RL. Structure and function of ferricyanide in the formation of chromate conversion coatings on aluminum aircraft alloy. J Electrochem Soc. 1999;146(10):3696–3701.

71 Zhao J, Frankel G, McCreery RL. Corrosion protection of untreated AA-2024-T3 in chloride solution by a chromate conversion coating monitored with Raman spectroscopy. J Electrochem Soc. 1998;145(7):2258–2264.

72 Zhao J, Xia L, Sehgal A, et al. Effects of chromate and chromate conversion coatings on corrosion of aluminum alloy 2024-T3. Surf Coat Technol. 2001;140(1):51–57.

73 Chidambaram D, Halada GP, Clayton CR. Spectroscopic elucidation of the repassivation of active sites on aluminum by chromate conversion coating. Electrochem Solid-State Lett. 2004;7(9):B31–B33.

74 Ramsey JD, McCreery RL. In situ Raman microscopy of chromate effects on corrosion pits in aluminum alloy. J Electrochem Soc. 1999;146(11):4076–4081.

75 Le BN, Joiret S, Thierry D, et al. The role of chromate conversion coating in the filiform corrosion of coated aluminum alloys. J Electrochem Soc. 2003;150(12):B561–B566.

76 Tomandl A, Wolpers M, Ogle K. The alkaline stability of phosphate coatings II: in situ Raman spectroscopy. Corros Sci. 2004;46(4):997–1011.

77 Hugot-Le Goff A, Bernard MC, Phillips N, et al. Contributions of Raman spectroscopy and electrochemical impedance to the understanding of the underpaint corrosion process of zinc-coated steel sheets. Mater Sci Forum. 1995;192–194:779–787.

78 Thierry D, Massinon D, Hugot-Le Goff A. In situ determination of corrosion products formed on painted galvanized steel by Raman spectroscopy. J Electrochem Soc. 1991;138(3):879–880.

79 Bernard MC, Hugot-Le Goff A, Phillips N. In situ Raman study of the corrosion of zinc-coated steel in the presence of chloride. II Mechanisms of underpaint corrosion and role of the conversion layers. J Electrochem Soc. 1995;142(7):2167–2170.

80 Bernard MC, Hugot-Le Goff A, Massinon D, et al. Underpaint corrosion of zinc-coated steel sheet studied by in situ raman spectroscopy. Corros Sci. 1993;35(5–8):1339–1349.

81 Mengoli G, Munari MT, Bianco P, et al. Anodic synthesis of polyaniline coatings onto Fe sheets. J Appl Polym Sci. 1981;26(12):4247–4257.

82 DeBerry DW. Modification of the electrochemical and corrosion behavior of stainless steels with an electroactive coating. J Electrochem Soc. 1985;132(5):1022–1026.

83 Hugot-Le Goff A, Bernard MC. Protonation and oxidation processes in polyaniline thin films studied by optical multichannel analysis and in situ Raman spectroscopy. Synth Met. 1993;60(2):115–131.

84 Quillard S, Berrada K, Louarn G, et al. In situ Raman spectroscopic studies of the electrochemical behavior of polyaniline. New J Chem. 1995;19(4):365–374.

85 Bernard MC, Hugot-Le Goff A, Joiret S, et al. Polyaniline layer for iron protection in sulfate medium. J Electrochem Soc. 1999;146(3):995–998.

86 Bernard MC, Joiret S, Hugot-Le Goff A, et al. Protection of iron against corrosion using a polyaniline layer II. Spectroscopic analysis of the layer grown in phosphoric/metanilic solution. J Electrochem Soc. 2001;148(8):B299–B303.

87 de Souza S, da Silva JEP, de Torresi SIC, et al. Polyaniline based acrylic blends for iron corrosion protection. Electrochem Solid-State Lett. 2001;4(8):B27–B30.

88 Torresi RM, de Souza S, da Silva JEP, et al. Galvanic coupling between metal substrate and polyaniline acrylic blends: corrosion protection mechanism. Electrochim Acta. 2005;50(11):2213–2218.

89 Seegmiller JC, Pereira da Silva JE, Buttry DA, et al. Mechanism of action of corrosion protection coating for AA2024-T3 based on poly(aniline)-poly(methylmethacrylate) blend. J Electrochem Soc. 2005;152(2):B45–B53.

90 da Silva JEP, de Torresi SIC, Torresi RM. Polyaniline/poly(methylmethacrylate) blends for corrosion protection: the effect of passivating dopants on different metals. Prog Org Coat. 2007;58(1):33–39.

91 da Silva JEP, Temperini MLA, de Torresi SIC. Secondary doping of polyaniline studied by resonance Raman spectroscopy. Electrochim Acta. 1999;44(12):1887–1891.

92 Bukowska J, Jackowska K. In situ Raman studies of polypyrrole and polythiophene films on Pt electrodes. Synth Met. 1990;35(1–2):143–150.

93 Ohtsuka T, Wakabayashi T, Einaga H. Optical characterization of polypyrrole-polytungstate anion composite films. Synth Met. 1996;79(3):235–239.

94 Nguyen Thi Le H, Bernard MC, Garcia-Renaud B, et al. Raman spectroscopy analysis of polypyrrole films as protective coatings on iron. Synth Met. 2004;140(2–3):287–293.

95 Van ST, Joiret S, Deslouis C, et al. In situ Raman spectroscopy and spectroscopic ellipsometry analysis of the iron/polypyrrole interface. J Phys Chem C. 2007;111(39):14400–14409.

96 Sheng N, Ueda M, Ohtsuka T. The formation of polypyrrole film on zinc-coated AZ91D alloy under constant current characterized by Raman spectroscopy. Prog Org Coat. 2013;76(2–3):328–334.

97 Nguyen TD, Pham MC, Piro B, et al. Conducting polymers and corrosion PPy-PPy-PDAN composite films. J Electrochem Soc. 2004;151(6):B325–B330.

98 Leygraf C, Johnson M. Infrared spectroscopy. In: Marcus P, Mansfeld F, eds. Analytical methods in corrosion science and engineering. Boca Raton: CRC Press; 2006:237–268.

99 Trettenhahn GLJ, Nauer GE, Neckel A. In situ external reflection absorption FTIR spectroscopy on lead electrodes in sulfuric acid. Electrochim Acta. 1996;41(9):1435–1441.

100 Mucalo MR, Li Q. In situ infrared spectroelectrochemical studies of the corrosion of a nickel electrode as a function of applied potential in cyanate, thiocyanate, and selenocyanate solutions. J Colloid Interface Sci. 2004;269(2):370–380.

101 Degenhardt J, McQuillan AJ. In situ ATR-FTIR spectroscopic study of adsorption of perchlorate, sulfate, and thiosulfate ions onto chromium(III) oxide hydroxide thin films. Langmuir. 1999;15(13):4595–4602.

102 Borras CA, Romagnoli R, Lezna RO. In-situ spectroelectrochemistry (UV-visible and infrared) of anodic films on iron in neutral phosphate solutions. Electrochim Acta. 2000;45(11):1717–1725.

103 Lin J, Lin C, Lin Z, et al. In situ measurement of the transport processes of corrosive species through a mortar layer by FTIR-MIR. Cem Concr Res. 2012;42(1):95–98.

104 Persson D, Leygraf C. In situ infrared reflection absorption spectroscopy for studies of atmospheric corrosion. J Electrochem Soc. 1993;140(5):1256–1260.

105 Persson D, Leygraf C. Initial interaction of sulfur dioxide with water covered metal surfaces: an in situ IRAS study. J Electrochem Soc. 1995;142(5):1459–1468.

106 Itoh J, Sasaki T, Seo M, et al. In situ simultaneous measurement with IR-RAS and QCM for investigation of corrosion of copper in a gaseous environment. Corros Sci. 1997;39(1):193–197.

107 Faguy PW, Richmond WN, Jackson RS, et al. Real-time polarization modulation in situ infrared spectroscopy applied to the study of atmospheric corrosion. Appl Spectrosc. 1998;52(4):557–564.

108 Wadsak M, Aastrup T, Odnevall WI, et al. Multianalytical in situ investigation of the initial atmospheric corrosion of bronze. Corros Sci. 2002;44(4):791–802.

109 Aastrup T, Wadsak M, Leygraf C, et al. In situ studies of the initial atmospheric corrosion of copper. Influence of humidity, sulfur dioxide, ozone, and nitrogen dioxide. J Electrochem Soc. 2000;147(7):2543–2551.

110 Kleber C, Kattner J, Frank J, et al. Design and application of a new cell for in situ infrared reflection-absorption spectroscopy investigations of metal-atmosphere interfaces. Appl Spectrosc. 2003;57(1):88–92.

111 Aastrup T, Leygraf C. Simultaneous infrared reflection absorption spectroscopy and quartz crystal microbalance measurements for in situ studies of the metal/atmosphere interface. J Electrochem Soc. 1997;144(9):2986–2990.

112 Dai Q, Freedman A, Robinson GN. Sulfuric acid-induced corrosion of aluminum surfaces. J Electrochem Soc. 1995;142(12):4063–4069.

113 Johnson CM, Tyrode E, Leygraf C. Atmospheric corrosion of zinc by organic constituents I. The role of zinc/water and water/air interfaces studied by infrared reflection/absorption spectroscopy and vibrational sum frequency spectroscopy. J Electrochem Soc. 2006;153(3):B113–B120.

114 Johnson CM, Leygraf C. Atmospheric corrosion of zinc by organic constituents. II. Reaction routes for zinc-acetate formation. J Electrochem Soc. 2006;153(12):B542–B546.

115 Johnson CM, Leygraf C. Atmospheric corrosion of zinc by organic constituents. III. An infrared reflection-absorption spectroscopy study of the influence of formic acid. J Electrochem Soc. 2006;153(12):B547–B550.

116 Hedberg J, Baldelli S, Leygraf C, et al. Molecular structural information of the atmospheric corrosion of zinc studied by vibrational spectroscopy techniques. Part I. Experimental approach. J Electrochem Soc. 2010;157(10):C357–C362.

117 Hedberg J, Baldelli S, Leygraf C. Molecular structural information of the atmospheric corrosion of zinc studied by vibrational spectroscopy techniques. II. Two and three-dimensional growth of reaction products induced by formic and acetic acid. J Electrochem Soc. 2010;157(10):C363–C373.

118 Persson D, Axelsen S, Zou F, et al. Simultaneous in situ infrared reflection absorption spectroscopy and Kelvin probe measurements during atmospheric corrosion. Electrochem Solid-State Lett. 2001;4(2):B7–10.

119 Chen ZY, Persson D, Leygraf C. Initial NaCl-particle induced atmospheric corrosion of zinc—effect of CO2 and SO2. Corros Sci. 2008;50(1):111–123.

120 LeBozec N, Persson D, Thierry D. In situ studies of the initiation and propagation of filiform corrosion on aluminum. J Electrochem Soc. 2004;151(7):B440–B445.

121 Weissenrieder J, Leygraf C. In situ studies of filiform corrosion of iron. J Electrochem Soc. 2004;151(3):B165–B171.

122 Chen ZY, Persson D, Nazarov A, et al. In situ studies of the effect of CO2 on the initial NaCl-induced atmospheric corrosion of copper. J Electrochem Soc. 2005;152(9):B342–B351.

123 Chen ZY, Persson D, Leygraf C. In situ studies of the effect of SO2 on the initial NaCl-induced atmospheric corrosion of copper. J Electrochem Soc. 2005;152(12):B526–B533.

124 Jönsson M, Persson D, Thierry D. Corrosion product formation during NaCl induced atmospheric corrosion of magnesium alloy AZ91D. Corros Sci. 2007;49(3):1540–1558.

125 Vogt MR, Nichols RJ, Magnussen OM, et al. Benzotriazole adsorption and inhibition of Cu(100) corrosion in HCl: a combined in-situ STM and in-situ FTIR spectroscopy study. J Phys Chem B. 1998;102(30):5859–5865.

126 Biggin ME, Gewirth AA. Infrared studies of benzotriazole on copper electrode surfaces. Role of chloride in promoting reversibility. J Electrochem Soc. 2001;148(5):C339–C347.

127 Bozzini B, Mele C, Romanello V. An in situ FT-IR evaluation of candidate organic corrosion inhibitors for carbon steel in contact with alkaline aqueous solutions. Mater Corros. 2007;58(5):362–368.

128 Pohjakallio M, Sundholm G, Talonen P, et al. Characterization of the redox processes of poly (thiophene-3-methanol) by voltammetry, in situ optical beam deflection and Fourier transform IR techniques. J Electroanal Chem. 1995;396(1–2):339–348.

129 D’Elia LF, Ortiz RL, Marquez OP, et al. Electrochemical deposition of poly(o-phenylenediamine) films on type 304 stainless steel. J Electrochem Soc. 2001;148(4):C297–C300.

130 Nguyen T, Byrd E, Lin C. A spectroscopic technique for in situ measurement of water at the coating/metal interface. J Adhes Sci Technol. 1991;5(9):697–709.

131 Nguyen T, Bentz D, Byrd E. A study of water at the organic coating/substrate interface. J Coat Technol. 1994;66(834):39–50.

132 Nguyen T, Byrd E, Bentz D, et al. In situ measurement of water at the organic coating/substrate interface. Prog Org Coat. 1996;27(1–4):181–193.

133 Philippe L, Sammon C, Lyon SB, et al. An FTIR/ATR in situ study of sorption and transport in corrosion protective organic coatings: 1. Water sorption and the role of inhibitor anions. Prog Org Coat. 2004;49(4):302–314.

134 Philippe L, Sammon C, Lyon SB, et al. An FTIR/ATR in situ study of sorption and transport in corrosion protective organic coatings: paper 2. The effects of temperature and isotopic dilution. Prog Org Coat. 2004;49(4):315–323.

135 Öhman M, Persson D, Leygraf C. In situ ATR-FTIR studies of the aluminium/polymer interface upon exposure to water and electrolyte. Prog Org Coat. 2006;57(1):78–88.

136 Wapner K, Stratmann M, Grundmeier G. In situ infrared spectroscopic and scanning Kelvin probe measurements of water and ion transport at polymer/metal interfaces. Electrochim Acta. 2006;51(16):3303–3315.

137 Öhman M, Persson D. An integrated in situ ATR-FTIR and EIS set-up to study buried metal–polymer interfaces exposed to an electrolyte solution. Electrochim Acta. 2007;52(16):5159–5171.

138 Öhman M, Persson D, Leygraf C. A spectroelectrochemical study of metal/polymer interfaces by simultaneous in situ ATR-FTIR and EIS. Electrochem Solid-State Lett. 2007;10(4):C27–C30.

139 Öhman M, Persson D, Jacobsson D. In situ studies of conversion coated zinc/polymer surfaces during exposure to corrosive conditions. Prog Org Coat. 2011;70(1):16–22.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset