trivalent chromium conversion coating formation on aluminium

13
Trivalent chromium conversion coating formation on aluminium J.-T. Qi, T. Hashimoto, J.R. Walton, X. Zhou, P. Skeldon , G.E. Thompson School of Materials, The University of Manchester, Manchester M13 9PL, England, UK abstract article info Article history: Received 4 July 2015 Revised 11 September 2015 Accepted in revised form 12 September 2015 Available online 16 September 2015 Keywords: Aluminium TEM RBS XPS Raman spectroscopy Trivalent chromium coating The formation of a trivalent conversion coating on aluminium has been investigated using analytical electron microscopy, atomic force microscopy, ion beam analysis, glow discharge optical emission spectroscopy, Raman spectroscopy and X-ray photoelectron spectroscopy. The coating is shown to comprise a chromium- and zirconium-rich outer layer and an aluminium-rich inner layer. Zirconium and chromium are present in chemical states consistent with ZrO 2 , ZrF 4 , Cr(OH) 3 , Cr 2 (SO 4 ) 3 , CrF 3 and CrO 3 or CrO 4 2. However, negligible amounts of Cr(VI) species occurred in coatings formed in de-aerated solution. Electrochemical impedance spectroscopy revealed that the inner layer provides the main corrosion protection during short-term tests in 0.1 M sodium sulphate solution at room temperature. © 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 1. Introduction Chromate conversion coatings are widely used as protective coatings for high-strength aluminium alloys, such as AA 2024-T3, which are susceptible to pitting corrosion [1,2]. However, due to the toxicity and high disposal costs of Cr(VI) compounds, alterna- tive treatments are being sought, amongst which trivalent chromi- um conversion (TCC) coatings are considered promising [24]. A TCC coating bath generally contains ZrF 6 2 , Cr 3+ and SO 4 2 constituents in an aqueous solution with a pH of 3.84.0 and coating is undertaken at a temperature of 40 °C [3,5]. The bath is similar to those used for the formation of zirconium-based conver- sion coatings, but with an addition of a trivalent chromium salt [6]. The resultant coatings on aluminium alloys have been reported to provide corrosion protection similar to that of chromate-based conversion treatments [7]. The coatings are considered to form by precipitation of an outer chromium- and zirconium-rich layer due to an increase of pH at sites of the usual cathodic reactions: 2H þ þ 2e H 2 ð1Þ O 2 þ 2H 2 O þ 4e 4OH : ð2Þ The pH at the coating surface can increase from 3.9 to 8.5 according to measurements using a tungsten microelectrode [8]. At the same time, aluminium is oxidized, forming an inner alumina lm across the aluminium surface: 2Al þ 3H 2 OAl 2 O 3 þ 6H þ þ 6e: ð3Þ The alumina lm dissolves by reaction with hydrouoric acid [9]: Al 2 O 3 þ xF þ 6H þ 2AlF x 3x ð Þ þ 3H 2 O: ð4Þ The dissolution is balanced by formation of fresh alumina at the aluminium/alumina interface. A high electric eld is maintained across the alumina layer, which enables the outward migration of Al 3+ and inward migration of O 2ions within the alumina and continued oxida- tion of the aluminium [10]. The relatively thin residual alumina lm thickness permits electron tunnelling to support the cathodic reac- tions [11]. The coatings have been reported to contain hydrated chromi- um and zirconium oxides [12,13], which may play a role as a hydroxide ion conductor with ligand exchange with uorine [14]. Despite the absence of Cr(VI) species in the TCC coating bath, recent research has reported that Cr(VI) species are present in the coatings after ageing in air or following a corrosion test in sodium chloride solution at ambient temperature [5,13,15,16]. The Cr(VI) species in coatings formed on zinc were suggested to result from oxidation of Cr(III) species by oxygen [15]. Other work considered the possibility of oxidation of Cr(III) species by hydrogen peroxide generated by the reduction of oxygen at copper-rich particles in an AA 2024 alloy in a sodium chloride solution [16]. In the present work, the formation of a TCC coating on high purity aluminium is investigated. The substrate was selected as a relatively simple system compared with the AA 2024 aluminium alloy that has Surface & Coatings Technology 280 (2015) 317329 Corresponding author. E-mail address: [email protected] (P. Skeldon). http://dx.doi.org/10.1016/j.surfcoat.2015.09.024 0257-8972/© 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). Contents lists available at ScienceDirect Surface & Coatings Technology journal homepage: www.elsevier.com/locate/surfcoat

Upload: independent

Post on 23-Nov-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Surface & Coatings Technology 280 (2015) 317–329

Contents lists available at ScienceDirect

Surface & Coatings Technology

j ourna l homepage: www.e lsev ie r .com/ locate /sur fcoat

Trivalent chromium conversion coating formation on aluminium

J.-T. Qi, T. Hashimoto, J.R. Walton, X. Zhou, P. Skeldon ⁎, G.E. ThompsonSchool of Materials, The University of Manchester, Manchester M13 9PL, England, UK

⁎ Corresponding author.E-mail address: [email protected] (P. Skeld

http://dx.doi.org/10.1016/j.surfcoat.2015.09.0240257-8972/© 2015 The Authors. Published by Elsevier B.V

a b s t r a c t

a r t i c l e i n f o

Article history:Received 4 July 2015Revised 11 September 2015Accepted in revised form 12 September 2015Available online 16 September 2015

Keywords:AluminiumTEMRBSXPSRaman spectroscopyTrivalent chromium coating

The formation of a trivalent conversion coating on aluminium has been investigated using analytical electronmicroscopy, atomic force microscopy, ion beam analysis, glow discharge optical emission spectroscopy, Ramanspectroscopy and X-ray photoelectron spectroscopy. The coating is shown to comprise a chromium- andzirconium-rich outer layer and an aluminium-rich inner layer. Zirconium and chromium are present in chemicalstates consistent with ZrO2, ZrF4, Cr(OH)3, Cr2(SO4)3, CrF3 and CrO3 or CrO4

2−. However, negligible amounts ofCr(VI) species occurred in coatings formed in de-aerated solution. Electrochemical impedance spectroscopyrevealed that the inner layer provides the main corrosion protection during short-term tests in 0.1 M sodiumsulphate solution at room temperature.

© 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license(http://creativecommons.org/licenses/by/4.0/).

1. Introduction

Chromate conversion coatings are widely used as protectivecoatings for high-strength aluminium alloys, such as AA 2024-T3,which are susceptible to pitting corrosion [1,2]. However, due tothe toxicity and high disposal costs of Cr(VI) compounds, alterna-tive treatments are being sought, amongst which trivalent chromi-um conversion (TCC) coatings are considered promising [2–4].A TCC coating bath generally contains ZrF62−, Cr3+ and SO4

2−

constituents in an aqueous solution with a pH of 3.8–4.0 andcoating is undertaken at a temperature of 40 °C [3,5]. The bath issimilar to those used for the formation of zirconium-based conver-sion coatings, but with an addition of a trivalent chromium salt [6].The resultant coatings on aluminium alloys have been reported toprovide corrosion protection similar to that of chromate-basedconversion treatments [7].

The coatings are considered to form by precipitation of an outerchromium- and zirconium-rich layer due to an increase of pH at sitesof the usual cathodic reactions:

2Hþ þ 2e→H2 ð1Þ

O2 þ 2H2Oþ 4e→ 4OH‐: ð2Þ

The pH at the coating surface can increase from 3.9 to 8.5 accordingtomeasurements using a tungstenmicroelectrode [8]. At the same time,

on).

. This is an open access article under

aluminium is oxidized, forming an inner alumina film across thealuminium surface:

2Alþ 3H2O→Al2O3 þ 6Hþ þ 6e: ð3Þ

The alumina film dissolves by reaction with hydrofluoric acid [9]:

Al2O3 þ xF‐ þ 6Hþ→2AlFx 3‐xð Þ þ 3H2O: ð4Þ

The dissolution is balanced by formation of fresh alumina at thealuminium/alumina interface. A high electric field is maintained acrossthe alumina layer, which enables the outward migration of Al3+ andinward migration of O2− ions within the alumina and continued oxida-tion of the aluminium [10]. The relatively thin residual alumina filmthickness permits electron tunnelling to support the cathodic reac-tions [11]. The coatings have been reported to contain hydrated chromi-um and zirconium oxides [12,13], whichmay play a role as a hydroxideion conductor with ligand exchange with fluorine [14].

Despite the absence of Cr(VI) species in the TCC coating bath, recentresearch has reported that Cr(VI) species are present in the coatingsafter ageing in air or following a corrosion test in sodium chloridesolution at ambient temperature [5,13,15,16]. The Cr(VI) species incoatings formed on zinc were suggested to result from oxidation ofCr(III) species by oxygen [15]. Other work considered the possibility ofoxidation of Cr(III) species by hydrogen peroxide generated by thereduction of oxygen at copper-rich particles in an AA 2024 alloy in asodium chloride solution [16].

In the present work, the formation of a TCC coating on high purityaluminium is investigated. The substrate was selected as a relativelysimple system compared with the AA 2024 aluminium alloy that has

the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Fig. 1. The dependence of the open circuit potential of electropolished aluminium on thetime of immersion in a dilute SurTec 650 bath at 40 °C under the usual naturally-aeratedcondition (labelled O2) and under the de-oxygenated condition (labelled N2).

318 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

usually been used in the previous studies. High-resolution analyticalelectron microscopy is combined with glow-discharge opticalemission spectroscopy (GDOES), atomic force microscopy (AFM),Rutherford backscattering spectroscopy (RBS), nuclear reaction analysis(NRA), X-ray photoelectron spectroscopy (XPS) and Raman spectrosco-py. The study focuses on the composition and morphology of theresultant coatings, and provides evidence for the formation of Cr(VI)species in the absence of copper-rich, second phase particles in thesubstrate. The corrosion protection afforded by the coating is investigatedby electrochemical impedance spectroscopy.

Fig. 2. Scanning electron micrographs (3 kV SE2 signal) of TCC coatings formed on electropolis(c) and (d) 600 s.

2. Materials and methods

2.1. Materials and reagents

Specimens, with dimensions of either 30 × 24 mm or 30 × 12 mm,were cut from 99.99% aluminium sheet of ≈0.3 mm thicknessand rinsed with acetone, ethanol and deionized water, each for 5 s.Individual specimens were then electropolished for 240 s in a20% (v/v) perchloric acid (60 wt.%) and 80%(v/v) ethanol mixturebelow 10 °C, using a potential of 20 V applied between the specimenand an aluminium counter electrode, followed by rinsing in deionizedwater and drying in a stream of cool air. They were then immersed innaturally-aerated SurTec 650 solution (1:4 v/v deionized water, pH =3.9 (adjusted with 1% NaOH or 5% H2SO4)) at 40 °C for times from 15to 1200 s. The solution includes both Cr(III) and zirconium species.After removal from the coating bath, the specimens were immersed indeionized water at 40 °C for 120 s and then rinsed in deionized waterat room temperature, dried in a cool air stream, and finally aged in theambient laboratory atmosphere at about 20 °C, typically for 24 h, beforebeing subjected to analysis or corrosion testing. The exposure to theambient atmosphere is expected to result in further drying of the coat-ing. Coatingswere also formed in a solution that was de-oxygenated, bybubbling nitrogen for 1 h, in order to determine the influence ofdissolved oxygen on the formation of Cr(VI) species.

2.2. Examination techniques

The open circuit potential (OCP) of the aluminium was measuredduring coating formation for 1200 s using a Solarton electrochemicalworkstation with a Modulab software controller. The exposed areaof the coated surface was ≈2 cm2. Electrochemical impedancemeasurements were carried out on specimens coated for 60, 300

hed aluminium in a dilute SurTec 650 bath at 40 °C for different times: (a) 15 s; (b) 120 s;

Fig. 3.Atomic forcemicroscopy topography images of electropolished aluminium (a) andTCC coatings formedon electropolished aluminium in a dilute SurTec 650 bath at 40 °C for (b)15 sand (c, d) 300 s.

319J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

and 600 s from 105 to 10−2 Hz in 0.1 M Na2SO4 solution at roomtemperature (≈20 °C). The sulphate solution was selected to enablethe electrochemical properties of the coated aluminium to be deter-mined without the occurrence of pitting. The specimens were im-mersed in the solution for 30 min, before applying a sinusoidalpotential waveform with an amplitude of 10 mV about the OCP.A three-electrode cell was employed, with a saturated calomel refer-ence electrode (SCE) and a platinumwire counter electrode. EIS datawere processed by ZView software (version 3.1, Scribner Associates,Inc.) to achieve the best fitting results.

The coatings were examined by scanning electron microscopy(SEM), using a Zeiss Ultra 55 instrument with an energy-dispersiveX-ray spectroscopy (EDS) facility, operated at accelerating voltages of3 and 15 kV. AFM was employed to record the coating topography,using a Dimension 3100 microscope with Nanoscope 3a controller(Bruker, Santa Barbara, USA) and TESPA tapping mode probes. Thedata were analysed by Nanoscope Analysis software (Version 1.5).

Transmission electron microscopy (TEM) was undertaken in a JEOL2000 FX II instrument with an accelerating voltage of 120 kV. Electrontransparent cross-sections of specimens were prepared using a LEICAEM UC6 ultramicrotome with a diamond knife [17]. The sections, ofnominal thickness of 15nm,were collected on nickel grids. Additionally,≈30 to 40 nm thick sections were examined in a FEI Titan G2 80-200

ChemiStem instrument, equipped with four EDS detectors, operated at200 kV.

Elemental depth profiles through the coating thickness weredetermined using GDOES employing a GD-Profiler 2 instrument(Horiba Jobin Yvon), with a copper anode of 4 mm diameter, anargon pressure of 635 Pa, a power of 35 W, a flush time 30 s, rf of13.56 MHz and a sampling interval of 0.01 s. The emission linesused were 396.157 nm for Al, 324.759 nm for Cu, 383.834 nm forMg, 339.203 nm for Zr, 425.439 nm for Cr, 589.600 nm for Na,180.738 nm for S and 130.223 nm for O. A plasma pre-treatmentwas used to remove contaminants on the anode [18]. Three analyseswere made for each specimen in order to ensure the reproducibilityof the data.

Specimens were also analysed by RBS and NRA, using ion beamsprovided by the Van de Graaff generator at the University of Namur,Belgium. RBS employed 2.0 MeV 4He+ ions, with detection of scatteredions at 165° to the direction of the incident beam. Datawere interpretedusing the RUMP program. NRA employed the 16O(d,p1)17O reaction,using 850 keV 2H+ ions with the detection of protons at 150° to thedirection of the incident beam. The 16O contents of the specimenswere determined by comparison of the proton yields with that from areference specimen of anodized tantalum. The areas of the RBS andNRA analyses were≈1 mm2.

Fig. 4. Bright field transmission electron micrographs of ultramicrotomed cross-sections of TCC coatings formed on electropolished aluminium in a dilute SurTec 650 bath at 40 °C fordifferent times: (a) 15 s; (b) 60 s; (c) 120 s and (d) 300 s.

320 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

XPS spectra were acquired using a Kratos Axis Ultra DLD spec-trometer, with monochromatic Al Kα radiation (hν = 1486.6 eV)operated at 150 W with a base pressure of 1.0 × 10−8 mbar. Chargeneutralisation was achieved by low energy electrons produced bypassing a current of 0.18 A through a linear filament located at thelower end of the electrostatic lens. Wide scan spectra were acquiredfor binding energies between 1200 and 0 eV at 80 eV pass energyand 0.5 eV step size. The data were quantified by measuring peakareas and using theoretical sensitivity factors [19] modified for theinstrument geometry and the inelastic mean free path energy

Fig. 5. Variation of the thickness of TCC coating on electropolished aluminium with timeof immersion in a dilute SurTec 650 bath at 40 °C. The thicknesses of the coatings weredetermined from TEM observations of ultramicrotomed sections.

dependency [20]. The intensity/energy response of the instrumentwas determined following acquisition of spectra from copper, silverand gold [21]. High energy resolution spectra at 20 eV pass energyand 0.1 eV step size were acquired for chemical state determination,and peak fitting was employed to determine the position and intensityof overlapping photoelectron peaks. Charge referencing was relative tothe hydrocarbon peak at 285.0 eV. All data processing was carried outusing CasaXPS version 2.3.17 (Casa Software, Teignmouth, UK). Thearea of the XPS analysis was 700 × 300 μm.

Raman spectra were obtained using a Renishaw 2000 Ramaninstrument with an argon laser (514 nm, 12.5 mW power excitation).Before carrying out the examination, the laser was calibrated usingthe silicon peak at 520 cm−1. The integration time was 30 s, with10 times accumulation to avoid the effect of stray light noise. Theanalysed area was≈1 μm in diameter.

3. Results and discussion

3.1. OCP measurements

The OCPs of aluminium during immersion for 1200 s in thenaturally-aerated and de-aerated SurTec 650 solutions are shownin Fig. 1. The potential for the former solution displays an initialrapid fall, a subsequent small peak at 85 s and a gradual rise to a rel-atively steady value. The minimum and final potentials are −1.54and −1.50 V (SCE) respectively. Repeated measurements gavesimilar results. The initial fall in potential is due to thinning of thepre-existing oxide film on the aluminium surface, which enablesthe cathodic and anodic reactions to proceed. The initial decreasein potential of 0.35 V indicates thinning of the film by about 0.4 nm[11,22]. The potential for the de-aerated solution displays a similar gen-eral trend to that of the naturally-aerated solution, but with slightlyhigher minimum and final potentials of −1.51 and −1.45 V (SCE).

Fig. 6.GDOES elemental depth profiles of electropolished aluminium following immersionfor 300 s in a dilute SurTec 650 bath at 40 °C. (a) Zr, Cr, C, and (b) S, O, Cu. The signal from Siin (b) results from the procedure used to clean the anode. The profiles in (a) and (b) aresmoothed by 10 neighboured points. The dashed line marks the aluminium/coatinginterface. The insert in (a) shows the Al signal.

321J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

The OCP of the coated specimens after several seconds of immersion inthe coating bath is far below the Cr(III)/Cr(VI) equilibrium potential(≈−1 VSCE) [23,24], indicating that direct oxidation of Cr(III) ions onthe aluminium substrate should not occur.

3.2. Morphology of the coatings and growth kinetics

Fig. 2(a–c) shows scanning electron micrographs of the surfaceof the TCC-coated aluminium after treatments of 15, 120 and 600 srespectively, with subsequent post-treatment consisting of immersionin deionized water at 40 °C for 120 s, rinsing in deionized water andstorage in ambient air for 24 h. After 15 s,most of the surface is relativelyfeatureless, although occasional areas reveal deposits of material with anodular appearance (Fig. 2(a)). EDS analysis of the deposits indicates anatomic ratio of chromium to zirconium of ≈3, which compares with aratio of ≈0.46 ± 0.03 in the adjacent coating material determined bylater RBS analysis. A deposit was observed at the bottom of the vesselcontaining the coating solution after treating several specimens.However, SEM examination of the dried deposit revealed particlesmuch larger than those on the coating surface, with an atomic ratio ofchromium to zirconium of ≈0.1. A previous study of the role ofhexafluorozirconate ions on the formation of CCCs on aluminium alloysalso revealed the presence of nodules at the coating surface with a sim-ilar morphology to those of Fig. 2(a), although the reason for their for-mation was not identified [25]. It has been noted in earlier work thatzirconium species in solution can develop large agglomerates at

pH 4.5 from a stable colloidal dispersion [26]. After treatment for120 s, in addition to the nodular features, cracking and detachment ofthe coating are evident (Fig. 2(b)). The detachmentwasmore extensiveafter treatment for 600 s, with the surface revealing both small and largeareas of coating loss (Fig. 2(c)). The increased magnification image ofFig. 2(d) shows the details of a region of local detachment. About 2%of the aluminium surface was affected by coating detachment fortimes up to 300 s; by 600 s, the coating had detached from ≈25% ofthe surface. However, other specimens coated for 600 s that had notbeen rinsed in deionized water following formation of the coating re-vealed only localized detachment of the coating, similar in appearanceto the small regions in Fig. 2(c, d). Thus, the detachment from largeareas is possibly associated with the rinsing procedure.

The topographies of the electropolished aluminium and the TCCcoatings formed for 15 and 300 s are displayed in the AFM images ofFig. 3(a–d). The specimen treated for 15 s received a post-treatmentconsisting of immersion in deionized water at 40 °C for 120 s, rinsingin deionized water and storage in ambient air for 24 h. However, thespecimen treated for 300 s was examined immediately after formationof the coating, i.e. without subsequent immersion and rinsing in deion-ized water. The aluminium substrate in the non-coated condition re-veals a furrowed surface due to the electropolishing pre-treatment(Fig. 3(a)), consistent with the earlier reports on the morphology of(110) surfaces [27,28]. Following treatments for 15 and 300 s, a thincoating has formed,with the furrows of the original surface being faintlyevident (Fig. 3(b, c)). Nodular depositswith similar dimensions and fea-tures to those revealed by SEM observation (Fig. 2(a)) occur locally asindividual or lines of particles. The linear features are possibly formeddue to cracking of the coating followed by enhanced deposition of coat-ingmaterial at the exposed substrate. After 300 s, furrows were still ev-ident (Fig. 3(c)) and cracks were also observed at some locations in thecoating (see arrow in Fig. 3(d)). The cracks were formed either duringthe growth of the coating and the subsequent drying of the specimenin air or in the brief interval between forming the coating and the exam-ination by AFM.

Bright field transmission electron micrographs of cross-sections ofthe coated aluminium following treatments in the TCC bath for 15, 60,120 and 300 s and the standard rinsing and immersion treatment,followed by storage in ambient air for 24 h, are shown in Fig. 4(a–d) respectively. At each of these times, the aluminium surface iscovered by a coating of relatively uniform thickness. The main part ofthe coatingmaterial appears relatively dark due mainly to the presenceof elements of high atomic number, later shown to be chromium andzirconium, compared with the aluminium substrate. The absence ofdiffracting regions suggests that the coatings have amorphous struc-tures. The thickness of the coating increaseswith increase of the immer-sion time. The coating has a fine texture and contains crack or fissures,which are especially evident for the coating formed for the longesttime. These features may result from the mechanical damage by thediamond knife during sectioning of the coating and dehydration of thecoating in the microscope. However, their spacing within the coatingwas on a much finer scale than the cracks in the coating that wererevealed by SEMandAFM, suggesting that they resultmainly fromdam-age during ultramicrotomy. A ≈2 nm-thick region of different appear-ance is present at the coating base, which is most easily observed inthe specimens treated for 120 and 300 s. Later analyses indicate thatthis region consists of an aluminium-rich film, with negligible amountsof chromium or zirconium. Thus, the results of the TEM examinationindicate that the coatings consist of two layers, with an inner layerthat remains only a few nanometres thick and an outer layer thatgrows significantly in thickness as the time of immersion in the coatingbath increases.

Fig. 5 reveals that the coating thickness, determined frommeasurements on the ultramicrotomed coating cross-sections thatwere observed by TEM, increases with immersion time, reaching≈93 nm after 600 s. The thickness measurements were made at

Fig. 7. High angular annular dark field (HAADF) transmission electron micrograph and EDS maps of the coating formed on electropolished aluminium for 300 s in a dilute SurTec 650bath at 40 °C.

322 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

several points along the length of coating cross-section, with the av-erage values and the standard deviations shown in the figure.The rate of thickening increases in the initial ≈60 s, and thenprogressively decreases at longer times. The average rate of growthis ≈0.27 nm s−1 up to 300 s and ≈0.08 nm s−1 between 300 and600 s. Dardona et al. [29] reported a growth rate before 300 sof 0.4 nm s−1, which is reasonably similar to the present valueof ≈0.3 nm s−1. A reduced growth rate after 300 s may arise fromthe impeded oxidation of aluminium due to reduced transportof species across the thickening coating. However, changes in thethickness and composition of the alumina layer at the base of thecoating may also occur that reduce the rate of electron tunnellingand, hence, the rate of the cathodic reaction and the deposition ofthe coating material.

3.3. Composition of the coatings

GDOES elemental depth profiles of the coated aluminium after a TCCtreatment for 300 s and the standard post-treatment are shown inFig. 6(a, b). The coating contains chromium, zirconium, oxygen, and car-bon and sulphur species. The aluminium/coating interface, determinedby the half height of the trailing edge of the oxygen signal, is markedby a dashed line. Whether or not aluminium is present in the coatingis unclear because of the low intensity of the aluminium signal andthe limited depth resolution of the analysis. Chromium and zirconiumappear to be distributed throughout the coating thickness. The varia-tions in the intensities of the chromium and zirconium signals withinthe coating may be due to changes in the coating composition and/orof the sputtering rate across the coating thickness. Carbon and sulphur

Fig. 8. Experimental and simulated (solid line) RBS spectra for the trivalent chromiumcoatings formed on electropolished aluminium for 300 s in a dilute SurTec 650 bathat 40 °C.

Table 2Elemental concentrations (g cm−3) from RBS and NRA analyses of coatings formed onaluminium for 60, 120 and 300 s.

Zr Cr S F Al ONRA CNRA Density

60 s 0.76 0.20 0.01 0.08 0.54 0.76 0.16 2.3120 s 1.00 0.27 0.01 0.28 0.39 0.72 0.09 2.6300 s 1.10 0.28 0.01 0.29 0.12 0.71 0.06 2.6

323J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

also appear to be present at all regions of the coating, with an enhance-ment in the optical emissions near the substrate. However, later EDSmapping did not reveal an increase in the concentrations of theseelements, suggesting that the intensity variations may arise fromchanges in the sputtering rate. A peak in the copper signal near thealloy/coating interface indicates an enrichment of the copper impurityin the aluminium substrate immediately beneath the coating.

EDS elemental maps of aluminium, oxygen, zirconium, chromiumand fluorine in an ultramicrotomed cross-section of the previousspecimen that was examined in the TEM are displayed in Fig. 7. Mostof the coating thickness appears to contain relatively uniform distribu-tions of the elements, although the signal for aluminium is muchlower than the signals from the other elements, indicating a relativelylow concentration. However, a thin, fluorine-rich region, which alsocontains aluminium and oxygen, is present adjacent to the substrate.

Fig. 8 presents the experimental and simulated RBS spectra for thealuminium after treatment in the TCC bath for 300 s followed by thestandard post-treatment and further storage in laboratory air for1 week before the analysis. The spectrum for the coated specimenshows peaks for oxygen, fluorine, sulphur, chromium, zirconium andhafnium in the coating; concerning the last, hafnium is an impurity inthe zirconium reagent used in the bath formulation. The detection ofsulphur in the coating is probably due to the presence of sulphate inthe coating solution. Similar spectra were obtained for treatmenttimes of 60 and 120 s apart from reductions in the magnitude of thepeaks due to the presence of thinner coatings. The formation of the coat-ing results in a displacement of the aluminium leading edge to a lowerenergy comparedwith the non-treated aluminium. The results of fittingof the spectra are given in Table 1 for specimens treated for times of 60,120 and 300 s. The fitting used a two-layer model. The inner layerconsisted of aluminium, oxygen and fluorine and the outer layerconsisted of oxygen, fluorine, sulphur, chromium and zirconium,

Table 1Elemental concentrations (×1015 atomic/cm2) from RBS and NRA analyses of coatingsformed on aluminium for 60, 120 and 300 s.

Zr Cr ORBS S F Al O NRA C NRA Cr/Zr

60 s 7.9 3.7 46.5 0.3 4.0 19.0 45.1 12.6 0.47120 s 25.0 11.7 122.9 1.0 33.3 33.3 103.4 16.7 0.47300 s 60.4 26.9 222.2 2.1 77.0 22.2 221.8 25.6 0.45

in agreement with the elemental distributions indicated by the EDSmaps. The table also includes the results of the analyses of oxygenand carbon by NRA. The analysis of oxygen by NRA is more accuratethan by RBS. Carbon and also hydrogen are too light to be detected byRBS. The coatings contain chromium and zirconium with an atomicratio that did not change significantly for all treatment times, indicatingthat the species are added to the coating in relatively constant propor-tions as the thickness of the coating increases. The average valueof the Cr:Zr ratio was ≈ 0.46 ± 0.03. In comparison, a TCC coatingformed on AA 2024 alloy under cathodic polarization, using a differentsolution composition to the present one,was reported to consist of a hy-drous oxide of composition Cr2O3∙ iH2O ∙x(ZrO2∙ jH2O) (i = 2.10 ± 0.55,j=1.60±0.45, and x=0.85±0.14)with a Cr/Zr ratio of 2.42 [30]. Thecomposition was attributed to the fast deposition of positively-chargedchromium (III) ions on the negatively-charged aluminium surface,while the hydrolysis of ZrF62− ions limited the deposition of the zirconi-um species. The coating compositions, determined by RBS and NRA, andthe coating thicknesses, determined by TEM, indicate that the averagedensity of the coatings was ≈2.5 g cm−3 (Table 2). The hydrolysis ofchromium sulphate, which is a probable component of the coatingsolution is shown in Fig. 9, based on the previously proposed hydrolysisof chromium chloride [3,31].

3.4. Chemical states of coating species

XPS analyses were carried out on a specimen treated in the TCC bathfor 120 s followed by the standard post-treatment. The depth of analysiswas ≈5 nm. The surface elemental concentrations are presented inTable 3. The values in the second roware corrected for the carbonaceousover layer [32]. Fig. 10 shows thehigh resolutionXPS spectra from theAl2p and the Zr 3d photoelectron regions. Five peaks were used to fit theintensity in the region of 73.0 to 84.0 eV, including three peaks for Al 2pand two peaks for Cr 3s at 77.1 eV and 80.9 eV. The presence of alumin-ium oxide, oxyfluoride and fluoride is indicated by the Al 2p peaks at74.7, 75.3 and 76.2 eV respectively. The Zr 3d region contains two spinorbit split doublets separated by 2.4 eV. With the intensity ratio of theZr 3d5/2 to Zr 3d3/2 components fixed at 1.5:1, the peak fitting of Zr3d5/2 peak indicated chemical states consistent with ZrO2 and ZrF4,with 95% of the zirconium being in the form of oxide [33].

Fig. 11 presents the high resolution spectra for (a) the C 1s and(b) the O 1s photoelectron regions. The fitting of the C 1s peak wasachieved with a C–H or C–C group at 285 eV, a silicone group at285.7 eV, a C–O group at 288.5 eV, a Si–O–C group at 287.6 eV, a C=OorO–C–O at 288.6 eV and a COOHgroup at 589.4 eV. Carbon contamina-tion has previously been regarded to result from contaminants inthe deionized water [34,35]. Five peaks were employed to fit the O 1sphotoelectron region. These peaks are expected for oxides, hydroxidesand carbonaceous contamination. Biesinger et al. reported peaks at530.1, 531.6 and 533.2 eV for the oxide, hydroxide and hydrated speciesof chromium [36]. Oxygen in ZrO2 and Al2O3 is expected to displaypeaks at 530.5 and 531.5 eV respectively.

The fitting of the Cr 2p3/2 peak is shown in Fig. 12. The fitting inFig. 12(a) includes the presence of chromium fluoride [37,38], whichhas not been considered in previously reported XPS data [12,13,39].The overall chromium species include Cr(OH)3 (577.28 eV), Cr2(SO4)3(578.49 eV), CrF3 (580.03 eV) with 59.7, 28.6, and 10.4% of the totalchromium being associated with the respective species. Small peaks

Fig. 9. Possible hydrolysis mechanism of chromium sulphate.

324 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

are also included for Cr(VI) species (579.58 eV—consistent with eitherCrO3 or CrO4

2−) [38]. These peaks are close to the detection limits ofthe analysis and fitting procedure, and indicate an upper limit on theCr(VI) of 1.3% of the total chromium. This amount of Cr(VI) would rep-resent≈0.03 at.% of all the atoms in the analysed region of the coating.Fig. 12(b) shows the fitting of the Cr 2p3/2 peak without consideringCrF3. The amount of Cr(VI) is now increased to 9.8% of the total chromi-um, although the fitting is not as good as previously. The measurementof the amount of Cr(VI) was not significantly affected by reduction ofCr(VI) to Cr(III) during the analysis. This was determined from therate of reduction of CrO3 powder over a period of 60 h of analysis,which indicated that only ≈1% of the Cr(VI) would be reduced duringthe usual analysis time for a coating of 2 h.

Table 4 compares the atomic concentrations determined by XPS inthe outer ≈5 nm of the coating thickness and by RBS and NRA acrossthe total coating thickness. Less fluorine and aluminium are detectedby XPS than by RBS, which is consistent with the concentration offluorine and aluminium being enhanced near the base of the coating.The surface elemental concentration of fluorine determined from the F2s peak (31.4 eV binding energy) is ≈10.6 times greater than thatfrom the F 1s peak (684.9 eV binding energy), indicating enrichmentof fluoride in the sub-surface region. The XPS yields an increased con-centration of zirconium relative to chromium compared with the ratiodetermined by RBS, which is possibly related to the difference in thedepths of analysis of the two techniques. An increased amount ofzirconium at the coating surface may result from zirconium depositingpreferentially with the pH increase [26,30] or from leaching of chromi-um species from the coating [6]. The latter process may have occurredduring the post-treatment of the coated aluminium,which involved im-mersion in deionized water at 40 °C for 120 s and subsequent rinsing indeionized water at room temperature.

Raman spectra for electropolished aluminium and a TCC coatingformed for 1200 s in the SurTec 650 solutions under naturally-aerated and de-oxygenated conditions are shown in Fig. 13(a). Theresults of the Raman analysis apply to the whole thickness of thecoating. Analysis of the coatings by RBS indicated that there was nosignificant difference in the atomic ratio of chromium to zirconiumin the coatings. Furthermore, the coatings were of similar thickness.Fig. 13(b) shows the fittings of the Raman spectra in the range 700 and

Table 3Elemental concentrations (at.%) determined from XPS analyses of a coating formed onaluminium for 120 s, before and following correction for the carbonaceous overlayer.

Zr 3d Cr 2p O 1s C 1s Al 2p S 2p F 1s

Before correction 9.7 2.3 34.8 42.7 2.1 0.4 4.5After correction 22.0 5.9 57.1 - 3.1 0.7 8.6

1050 cm−1 for TCC coatings formed under the naturally-aerated con-dition. The specimens were analysed immediately after the 1200 simmersion treatment in the conversion coating bath followed bydrying in cool air. The 438, 804 and 945 cm−1 peaks in the spectrumof electropolished aluminium are attributed to the air-formed aluminafilm [40–42]. The peaks for the coating formed in the naturally-aerated solution reveal the presence of zirconium oxide (ZrO2, 233and 470 cm−1) [43], chromium hydroxide (Cr(OH)3, 536 cm−1)[16], a mixture of aluminium hydroxide and oxide (438, 804 and945 cm−1) [41], Cr(VI) species (866 cm−1) [16,44] and the S–Obond in a SO4

2−-containing compound, such as Cr2(SO4)3 (995,1050 and 1155 cm−1) [45]. Notably, the peak centred at 866 cm−1,indicative of the presence of Cr(VI) species, is consistent in positionand shape with either a mixed oxide that contains aluminium oxide,Cr(III) oxide and Cr(VI) oxide or CrO4

2− [40,41]. Since the coating wasonly exposed to air for no more than 1 h before the measurementsweremade, the Cr(VI) is probably formed during formation of the coat-ing. The Cr(VI) was not detectable when the solution was de-aerated,indicating that the formation of Cr(VI) is associated with reductionof oxygen during the growth of the coating. The reduction of oxygengenerates hydrogen peroxide according to the reaction:

O2 þ 2H2Oþ 2e→H2O2 þ 2OH‐: ð5Þ

The hydrogen peroxide may subsequently oxidize Cr(III) species inthe coating, for instance according to the reaction:

2Cr OHð Þ3 þ 3H2O2 → 2Cr OHð Þ6: ð6Þ

The quantitative XPS analysis of the Cr2p and Al2p photoelectronregions in a coating formed for 1200 s under the de-aerated conditionis shown in Fig. 14(a, b) respectively. This coatingwas analysedwithoutimmersion or rinsing in deionizedwater. The fitting of the Cr 2p data in-cluded the species used for the naturally-aerated solution, with the ad-dition of Cr2O4

2− species; the Cr(VI) species represent 2.6% of the totalchromium in the analysed region. The amount of Cr(VI) is not regardedas significant, being at the limit of detection set by the accuracy of thefitting procedure. The relatively high signal from aluminium metal inthe Al2p photoelectron region probably originates from exposure ofthe metal due to cracking and detachment of the coating.

It is suggested that the formation of Cr(VI) species under naturally-aerated conditions is associated with generation of H2O2 by the reduc-tion of dissolved oxygen as O2 + 2H++2e− → H2O2 (+0.94 VSCE).The reduction reaction is proposed to take place at sufficientlythin regions of the alumina layer at the base of the coating, wheretunnelling of electrons can occur, or at flaw sites in the layer, associ-ated with the influences on the film composition and morphology of

Fig. 10. High resolution XPS spectrum for (a) Al 2p and (b) Zr 3d photoelectron regionsand curve fitting for a coating formed on electropolished aluminium for 120 s in a diluteSurTec 650 bath at 40 °C. Fig. 11.High resolution XPS spectrum for (a) C 1s and (b) O 1s photoelectron regions and

curvefitting for a coating formed on electropolished aluminium for 120 s in a dilute SurTec650 bath at 40 °C.

325J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

impurities in the aluminium substrate. The H2O2 oxidizes the Cr(III)species that have been transported across the permeable outer coat-ing region, forming Cr(VI) species which can then diffuse toward thecoating surface, as shown in Fig. 15. Cr(VI) species may then be lostto the coating bath.

An accurate value for the amount of Cr(VI) species that remainwithin the coating cannot be given from the present results. TheRaman results are qualitative, while the XPS analyses apply onlyto the outer region of the coating, which has a total thickness

of ≈80 nm after a coating treatment of 300 s. Furthermore, thepresence of CrF3 interferes significantly with detection of Cr(VI)species. If the CrF3 is included in the fitting of the Cr 2p peak, the ne-cessity for consideration of Cr(VI) is marginal. If the CrF3 is excludedfrom the fitting, an upper limit of ≈1 wt.% of Cr(VI) in the coating isestimated, since chromium constitutes ≈9.8 wt.% of the coating onthe basis of the RBS results of Table 2, while XPS indicates that a

Fig. 12.High resolution XPS spectrum and curve fitting for the Cr 2p photoelectron regionfor a coating formed on electropolished aluminium for 120 s in a dilute SurTec 650 bath at40 °C. (a) Fitting with CrF3 and (b) fitting without CrF3.

Table 4Comparison of the elemental concentrations (at.%) from RBS, NRA and XPS analyses of acoating formed on aluminium for 120 s.

ONRA F S Cr Zr Al

RBS & NRA 46.1 14.8 0.4 5.2 11.1 14.8XPS 57.1 8.6 0.7 5.9 22.0 3.1

Fig. 13. (a) Raman spectra for electropolished aluminium and the TCC coatings formed for1200 s in a dilute SurTec 650 bath at 40 °C under either naturally aerated or de-aeratedconditions. All data were smoothed over 10 neighbouring points. (b) Fitting of spectrumin the range from 700 to 1050 cm−1 for the naturally-aerated condition.

326 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

maximum of 9.8% of the chromium may be in the form of Cr(VI).Inclusion of CrF3 suggested an upper limit on the amount of Cr(VI)of ≈0.1 wt.%. These estimates assume that the Cr(VI) is distributeduniformly through the coating thickness. However, this may not bethe case if Cr(VI) is formed in the inner coating regions and diffusesoutward, in which case a greater amount of Cr(VI) would be expectedto be present in the coatings.

3.5. Electrochemical impedance spectroscopy

Fig. 16(a, b) shows the dependence of the impedance on frequencyand phase angle-frequency for the aluminium in the non-coated condi-tion and after coating for 60, 300 and 600 s, under naturally-aeratedcondition, followed by immersion for 120 s in deionized water at40 °C and rinsing in deionized water at room temperature. The EISdata were obtained in 0.1 M Na2SO4 solution. The measurementscommenced after a period of 30 min immersion in the solution toallow the OCP to stabilize. The OCPs of the uncoated aluminium andthe coatings formed for 60, 300 and 600 s were similar, with valuesof≈−1.2 V vs SCE. The EIS datawere fitted using the equivalent circuitsof Fig. 16(c, d). The measurements were repeated on three specimensfor each condition of the alloy. The average values and standard

Fig. 14.High resolution of (a) Cr2p and (b) Al2p and photoelectron regions in the coatingsformed on electropolished aluminium for 1200 s in a dilute SurTec 650 bath at 40 °C underthe de-aerated condition.

Fig. 15. Schematic diagram of the TCC coating formation and oxidation of Cr(III) to Cr(VI)by hydrogen peroxide. (a) Cathodic reactions proceed at local regions, with deposition ofcoatingmaterial in response to local increase in pH. (b) Transport of species throughporesand oxidation of Cr(III) by hydrogen peroxide generated by reduction of oxygen.

327J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

deviations obtained for the circuit elements, which reveal a good repro-ducibility of the results, are given in Table 5.

The data for the non-coated aluminium were fitted using a sim-ple circuit comprising the solution resistance, i.e. the resistancebetween the reference electrode and the specimen (Re), and a par-allel combination of a resistance (Rct) representing the charge transferresistance and a constant phase element (Qdl), with an impedancegiven by Z = (jω)-nQ−1. The exponent was close to 1 indicating that

the impedance wasmainly due to the double layer capacitance. The cir-cuit for the coated specimens included an additional resistance (Rcoat)and constant phase element (Qcoat) that represent the effects of thecoating; n has been associatedwith the heterogeneity of the TCC coatingin previouswork [46]. Table 5 reveals that the charge transfer resistancedecreases with increase of the time of the coating treatment, with a rel-atively large decrease occurring after treatment of 300 s. The cause ofthe decrease is possibly associated with the accumulation of fluorideions at the base of the coating, which thin the alumina film present onthe aluminium substrate. In contrast, the coating resistance increaseswith increase of the treatment time,whichmay be due to a combinationof an increase in the coating thickness and change in the coating mor-phology, for instance a reduction in pore size. The values of the coatingresistance are much lower than those of the charge transfer resistanceand are of the order of the solution resistance. Hence, the corrosionprotection is mainly provided by the aluminium-rich, inner layer ofthe coating. However, the low frequency impedance of the aluminiumcoated for 300 and 600 s was lower than that of the uncoated alumini-um and the aluminium coated for 60 s. The reduced impedance of theformer specimens is possibly related to the accumulation of fluorideions near the base of the coating, which was evident by analyticalTEM in a coating formed for 300 s. The presence of the fluoride ionsthen reduced the protection provided by the aluminium-rich layer.

Fig. 16.Dependence of themodulus of the impedance on (a) frequency and (b) phase angle for the bare aluminium in the non-coated condition and following coating for 60, 300 and 600 s.The specimens were post-treated by immersion for 120 s in deionized water at 40 °C. The impedance measurements were carried out in 0.1 M Na2SO4 solution. The equivalent circuitmodels for uncoated and coated aluminium are shown in (c) and (d) respectively.

328 J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

The relatively low resistance of the coating indicates its permeabilitythat allows access of the electrolyte to the substrate throughpores and cracks. The most inhomogeneous coating, as indicated by arelatively low value of n, is that formed for 600 s, which is consistentwith the earlier findings from SEM of increased cracking and detach-ment of the coating. The effective capacitances per unit area of the coat-ings formed for 60, 300 and 600 s are given by Ceff = Q1/nR(1 − n)/n,where R represents the ohmic resistance of the electrolyte [46–48].The maximum effective capacitance per unit area of the inner doublelayer was measured as ≈3.2 μF/cm2 for the specimen coated for by atreatment of 600 s; the value is in good agreement with that reportedby Guo and Frankel [6].

4. Conclusions

1. The trivalent chromium conversion coating formed on the highpurity aluminium consists of two main layers. The outer layer,which constitutes most of the coating thickness, consists of AlF3,Al2O3, AlOxF, Cr(OH)3, CrF3, Cr2(SO4)3, ZrO2 and ZrF4 species.The inner layer is aluminium-rich, with the presence of oxideand fluoride species. The inner coating layer provides the maincorrosion protection.

2. The average rate of coating growth for treatment times up to 300 s is≈0.27 nms−1. Thereafter, the growth rate slows significantly, whichis due to hindered transport of reactant and product species by

Table 5Parameters of TCC coatings/Al systems in 0.1M sodium sulphate solution obtained from EISaluminium respectively.

Re Rcoat Rct Qcoat

Ω cm2 (×10−6

Bare Al 37 ± 2 – 37.0 ± 1.8 × 105 –60 s-TCC 23 ± 2 25 ± 5 32.5 ± 3.1 × 105 0.6 ± 0300 s-TCC 26 ± 1 45 ± 3 1.7 ± 0.1 × 105 1.8 ± 0600 s-TCC 24 ± 1 85 ± 3 1.1 ± 0.1 × 105 3.6 ± 0

the thickening coating layers or changes in the composition andthickness of the inner layer that affect the rate of electron tunnelling.

3. The coatings may partly detach from the aluminium substrateduring rinsing with water shortly after their formation. Further-more, cracks develop in coatings formed after ≈60 to 300 s. Thecracks are probably initiated by stress in the coating related todrying of the coating.

4. Raman spectroscopy revealed qualitatively the presence of Cr(VI)species in the coating and provided evidence of the role of oxygenin Cr(VI) formation. The presence of CrF3 interferes with the analysisof Cr(VI) species by XPS. With the assumption of a uniform distribu-tion of Cr(VI) in the coating, the amount of Cr(VI) was estimated as≈0.1 to 1% of the coating weight.

Acknowledgements

The authors acknowledge Engineering and Physical ScienceResearch Council (U.K.) for support of the LATEST 2 ProgrammeGrant and the China Council Scholarship for financial support of J-T.Qi. The technical support from X.L. Zhong and H. Liu is also greatlyappreciated. They also thank the European Community for financialassistance within the Integrating Activity “Support of Public and In-dustrial Research Using Ion Beam Technology (SPIRIT)”, under ECcontract no. 227012.

data using the equivalent circuits of Fig. 16(c) and (d) for bare aluminium and coated

Qdl ncoat ndl Ceff-dl

sn/(Ω cm2)) μF/cm2

3.8 ± 0.5 – 1 3.8 ± 0.5.4 4.2 ± 0.2 0.94 ± 0.01 0.96 ± 0.01 3.3 ± 0.3.2 4.3 ± 0.3 0.91 ± 0.01 0.96 ± 0.01 3.0 ± 0.1.1 3.6 ± 0.1 0.80 ± 0.01 0.99 ± 0.01 3.2 ± 0.4

329J.-T. Qi et al. / Surface & Coatings Technology 280 (2015) 317–329

References

[1] G.S. Chen, M. Gao, R.P. Wei, Microconstituent-induced pitting corrosion in alumi-num alloy 2024-T3, Corrosion 52 (1996) 8–15.

[2] M.W. Kendig, R.G. Buchheit, Corrosion inhibition of aluminum and aluminum alloysby soluble chromates, chromate coatings, and chromate-free coatings, Corrosion 59(2003) 379–400.

[3] F. Pearistein, V.S. Agarwala, Trivalent chromium solutions for applying chemicalconversion coatings to aluminium alloys or for sealing anodied aluminium, Plat.Surf. Finish. 81 (1994) 50–55.

[4] Occupational Exposure to Hexavalent Chromium. , United States Department ofLabor, US, 2006.

[5] A. Iyer, W. Willis, S. Frueh, A. Fowler, J. Barnes, L. Hagos, J. Escarega, J.L. Scala, S.L.Suib, Characterization of NAVAIR trivalent chromium process (TCP) coatings and so-lutions, Plat. Surf. Finish. 5 (2010) 32–42.

[6] Y. Guo, G.S. Frankel, Active corrosion inhibition of AA2024-T3 by trivalent chromeprocess treatment, Corrosion 68 (2012) 1–10.

[7] C.A. Matzdorf, W.C. Nickerson, E. Lipnickas, Evaluation of modified zirconium/triva-lent chromium conversion coatings by accelerated corrosion and electrochemicaltechniques, Proc. 2005 Tri-Service Corrosion Conference, Orlando, Florida, U.S.A.,2005.

[8] L. Li, A.L. Desouza, G.M. Swain, In situ pH measurement during the formation ofconversion coatings on an aluminum alloy (AA2024), Analyst 138 (2013)4398–4402.

[9] J. Cerezo, P. Taheri, I. Vandendael, R. Posner, K. Lill, J.H.W. de Wit, J.M.C. Mol, H.Terryn, Influence of surface hydroxyls on the formation of Zr-based conversion coat-ings on AA6014 aluminum alloy, Surf. Coat. Technol. 254 (2014) 277–283.

[10] G.E. Thompson, R.C. Furneaux, G.C.Wood, J.A. Richardson, J.S. Goode, Nucleation andgrowth of porous anodic films on aluminium, Nature 272 (1978) 433–435.

[11] K. Shimizu, G.M. Brown, K. Kobayashi, G.E. Thompson, G.C. Wood, The role of elec-tron tunnelling in the development of chemical conversion coatings on high purityaluminium, Corros. Sci. 34 (1993) 1853–1857.

[12] Y. Guo, G.S. Frankel, Characterization of trivalent chromium process coating onAA2024-T3, Surf. Coat. Technol. 206 (2012) 3895–3902.

[13] L.L. Li, G.P. Swain, A. Howell, D. Woodbury, G.M. Swain, The formation, structure,electrochemical properties and stability of trivalent chrome process (TCP) coat-ings on AA2024, J. Electrochem. Soc. 158 (2011) C274–C283.

[14] S. Licht, X. Yu, D. Zheng, Cathodic chemistry of high performance Zr coated alkalinematerials, Chem. Commun. (2006) 4341–4343.

[15] T. Rochester, Z.W. Kennedy, Unexpected results from corrosion testing of trivalentpassivates, Plat. Surf. Finish. 94 (2007) 14–18.

[16] L.L. Li, D.Y. Kim, G.M. Swain, Transient formation of chromate in trivalent chromiumprocess (TCP) coatings on AA2024 as probed by Raman spectroscopy, J.Electrochem. Soc. 159 (2012) C326–C333.

[17] R.S. Timsit, J.L. Hutchison, M.C. Thornton, Preparation of metal specimens for HREMby ultramicrotomy, Ultramicroscopy 15 (1984) 371–374.

[18] I.S. Molchan, G.E. Thompson, P. Skeldon, N. Trigoulet, P. Chapon, A. Tempez, J.Malherbe, L.L. Revilla, N. Bordel, P. Belenguer, T. Nelis, A. Zahri, L. Therese, P.Guillot, M. Ganciu, J. Michler, M. Hohl, The concept of plasma cleaning in glow dis-charge spectrometry, J. Anal. At. Spectrom. 24 (2009) 734–741.

[19] J.H. Scofield, Hartree-Slater subshell photoionization cross-sections at 1254 and1487 eV, J. Electron Spectrosc. Relat. Phenom. 8 (1976) 129–137.

[20] J. Walton, N. Fairley, A traceable quantification procedure for a multi-mode X-rayphotoelectron spectrometer, J. Electron Spectrosc. Relat. Phenom. 150 (2006)15–20.

[21] M.P. Seah, A system for the intensity calibration of electron spectrometers, J. Elec-tron Spectrosc. Relat. Phenom. 71 (1995) 191–204.

[22] G.M. Brown, K. Shimizu, K. Kobayashi, G.E. Thompson, G.C. Wood, The growth ofchromate conversion coatings on high purity aluminium, Corros. Sci. 34 (1993)1045–1054.

[23] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Ssolutions, PergammonPress, London, 1996.

[24] R. Cornelius, H. Crews, J. Caruso, K.G. Heumann (Eds.), Handbook of Elemental Spe-ciation, Handbook of Elemental Speciation II: Species in the Environment, Food,Medicine and Occupational Health, John Wiley & Sons, Ltd 2005, p. 121.

[25] D. Chidambaram, C.R. Clayton, G.P. Halada, The role of hexafluorozirconate in theformation of chromate conversion coatings on aluminum alloys, Electrochim. Acta51 (2006) 2862–2871.

[26] M. Rumyantsev, A. Shauly, S.G. Yiantsios, D. Hasson, A.J. Karabelas, R. Semiat, Param-eters affecting the properties of dynamic membranes formed by Zr hydroxidecolloids, Desalination 131 (2000) 189–200.

[27] V.V. Konovalov, G. Zangari, R.M. Metzger, Highly ordered nanotopographies onelectropolished aluminum single crystals, Chem. Mater. 11 (1999) 1949–1951.

[28] G.M. Brown, K. Shimizu, K. Kobayashi, G.E. Thompson, G.C. Wood, The morphology,structure and mechanism of growth of chemical conversion coatings on aluminium,Corros. Sci. 33 (1992) 1371–1385.

[29] S. Dardona, M. Jaworowski, In situ spectroscopic ellipsometry studies of trivalentchromium coating on aluminum, Appl. Phys. Lett. 97 (2010) 18190801–18190803.

[30] X.C. Dong, P. Wang, S. Argekar, D.W. Schaefer, Structure and composition of trivalentchromium process (TCP) films on Al alloy, Langmuir 26 (2010) 10833–10841.

[31] F. Pearlstein, V. S. Agarawala, Trivalent chromium conversion coatings for alumi-num, Google Patents, US5304257 A (1994).

[32] M.P. Seah, Simple universal curve for the energy-dependent electron attenuationlength for all materials, Surf. Interface Anal. 44 (2012) 1353–1359.

[33] A.V. Naumkin, A. Kraut-Vass, S.W. Gaarenstroom, C.J. Powell, NIST X-ray photoelec-tron spectroscopy database: NIST standard reference database 20, version 4.1,http://srdata.nist.gov/xps/2012.

[34] Laboratory Water, Its Importance and Application, USA, Department of Health &Human Services, 2013.

[35] M.R. Alexander, G. Beamson, P. Bailey, T.C.Q. Noakes, P. Skeldon, G.E. Thompson, Thedistribution of hydroxyl ions at the surface of anodic alumina, Surf. Interface Anal.35 (2003) 649–657.

[36] M.C. Biesinger, L.W.M. Lau, A.R. Gerson, R.S.C. Smart, Resolving surface chemicalstates in XPS analysis of first row transition metals, oxides and hydroxides: Sc, Ti,V, Cu and Zn, Appl. Surf. Sci. 257 (2010) 887–898.

[37] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.S.C. Smart, Re-solving surface chemical states in XPS analysis of first row transition metals, ox-ides and hydroxides: Cr, Mn, Fe, Co and Ni, Appl. Surf. Sci. 257 (2011)2717–2730.

[38] M.C. Biesinger, C. Brown, J.R. Mycroft, R.D. Davidson, N.S. McIntyre, X-ray photoelec-tron spectroscopy studies of chromium compounds, Surf. Interface Anal. 36 (2004)1550–1563.

[39] W.K. Chen, J.L. Lee, C.Y. Bai, K.H. Hou, M.D. Ger, Growth and characteristics ofCr(III)-based conversion coating on aluminum alloy, J. Taiwan Inst. Chem. Eng.43 (2012) 989–995.

[40] L. Xia, R.L. McCreery, Chemistry of a chromate conversion coating on aluminumalloy AA2024-T3 probed by vibrational spectroscopy, J. Electrochem. Soc. 145(1998) 3083–3089.

[41] J.D. Ramsey, R.L. McCreery, Raman microscopy of chromate interactions with cor-roding aluminum alloy 2024-T3, Corros. Sci. 46 (2004) 1729–1739.

[42] K.A. Rodgers, Routine identification of aluminium hydroxide polymorphs with thelaser Raman microprobe, Clay Miner. 28 (1993) 85–99.

[43] A. Sarfraz, R. Posner, M.M. Lange, K. Lill, A. Erbe, Role of intermetallics and copper inthe deposition of ZrO2 conversion coatings on AA6014, J. Electrochem. Soc. 161(2014) C509–C516.

[44] E.L. Lee, I.E. Wachs, Molecular design and in situ spectroscopic investigation of multi-layered supported M1Ox/M2Ox/SiO2 catalysts, J. Phys. Chem. C 112 (2008)20418–20428.

[45] F.G. Baglin, D. Breger, Identification of the zirconium sulfate species in highly acidicaqueous solutions by Raman spectroscopy, Inorg. Nucl. Chem. Lett. 12 (1976)173–177.

[46] L. Li, K.P. Doran, G.M. Swain, Electrochemical characterization of trivalent chromiumprocess (TCP) coatings on aluminum alloys 6061 and 7075, J. Electrochem. Soc. 160(2013) C396–C401.

[47] B. Hirschorn, M.E. Orazem, B. Tribollet, V. Vivier, I. Frateur, M. Musiani, Determina-tion of effective capacitance and film thickness from constant-phase-element pa-rameters, Electrochim. Acta 55 (2010) 6218–6227.

[48] G.J. Brug, A.L.G. Vandeneeden, M. Sluytersrehbach, J.H. Sluyters, The analysis of elec-trode impedances complicated by the presence of a constant phase element, J.Electroanal. Chem. 176 (1984) 275–295.