Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Surface & Coatings Technology 236 (2013) 252–261

Contents lists available at ScienceDirect

Surface & Coatings Technology


journal homepage: www.elsevier.com/locate/surfcoat

Corrosion resistance of chromium-free conversion coatings deposited on


electrogalvanized steel from potassium hexafluorotitanate(IV)
containing bath
J. Winiarski ⁎, J. Masalski, B. Szczygieł
Division of Surface Engineering, Catalysis and Corrosion, Wroclaw University of Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland

a r t i c l e i n f o a b s t r a c t

Article history: Chromium-free conversion coatings were deposited from a bath containing potassium hexafluorotitanate(IV).
Received 29 April 2013 The substrate was electrogalvanized carbon steel. The obtained coatings had a microspheroidal structure and
Accepted in revised form 29 September 2013 were homogenous. An XPS analysis showed that the coatings were composed of: oxides, hydroxides, phosphates,
Available online 5 October 2013
carbonates and fluorocompounds. As part of this study, the influence of coating deposition time on the surface
morphology, chemical composition and physicomechanical properties of the samples was examined. DC polari-
Keywords:
Electrogalvanized steel
zation measurements, electrochemical impedance spectroscopy and neutral salt spray tests showed that the
Conversion coating investigated Cr-free coatings markedly increase the corrosion resistance of electrogalvanized steel in a chloride
Corrosion resistance environment.
XPS © 2013 Elsevier B.V. All rights reserved.
EIS

1. Introduction amounts of phosphates and organic compounds codeposited in the


course of conversion layer formation [8]. The manganese(II) phosphate
The harmfulness of the Cr(VI) compounds used in the chemical present in the bath increases the coating deposition rate while the or-
treatment of the surface of: zinc, aluminum alloys and magnesium is ganic components incorporated into the coating increase its resistance
commonly known. The current legal norms force manufacturers to to the aggressive action of chloride ions [10]. Polarization test results
abandon the deposition of coatings from baths based on Cr(VI) com- showed that the protective action of coatings containing titanium
pounds, for processes using substances less harmful to the natural envi- consists in inhibiting the cathodic reaction. This may be indicative of
ronment. Usually chromate treatment in baths based on Cr(III) constrained oxygen reduction kinetics in the cathodic sites of the coat-
compounds is used in the industry today. This solution, however, is ing [13].
not fully satisfactory and there is wide research aimed at obtaining XPS showed that the coatings obtained from a bath based on
totally Cr-free replacements for chromate coatings. hexafluorotitanate(IV) compounds containing an organic addition are
Tetraoxomolybdates(VI) [1–3], rare earth metal salts [4] or silicon composed of mainly: C, O, N, P, F, Zn, Ti and Mn [8,13]. The depth profile
compounds [5] seem to offer a chance of eliminating chromium com- measurements performed for the coatings deposited from this bath, but
pounds from the baths used for depositing conversion coatings. Studies with no organic addition, showed that the coating's top layer contains
have shown that the coatings deposited from such baths reduce the rate mainly: O, Ti and P, whereas Mn appears only in the inner layer [8].
of corrosion of the zinc substrate, but their protective action has a typi- Thus one can infer that the sparingly soluble manganese compounds
cal barrier-like character and is of relatively short duration. precipitate earlier than the titanium and phosphorous compounds.
Decidedly the highest hopes of obtaining good quality Cr-free There is no zinc in the oxidized form (Zn2+) on the surface of the coat-
conversion coatings lie in baths based on titanium or zirconium com- ing. In a measurable quantity such zinc occurs only at the conversion
pounds. Titanium based coatings can be deposited from baths in coating/zinc substrate interface.
which the source of Ti is hexafluorotitanate(IV) acid, its salts [6–13] or The aim of this study was to obtain titanium containing conversion
titanium chlorides (TiCl4, TiCl3) [14]. In the literature one can find coatings from a bath with a simple chemical composition. Such coatings
descriptions of titanium containing coatings deposited on hot-dip gal- were deposited on the surface of electrogalvanized steel. The protective
vanized steel from a bath with a pH of 2.0–5.0 and T = 25–40 °C [8]. properties of the coatings should be similar to those of chromate coat-
The coatings are homogenous, devoid of cracks and much thinner (by ings and should considerably improve the corrosion resistance of galva-
a few tens of nm) than chromate coatings. They contain considerable nized steel. A Cr-free bath composed of popular and easily available
reagents: K2TiF6, H3PO4, Mn(NO3)2 and H2O2 was proposed. The aim
⁎ Corresponding author. Tel.: +48 71 320 3193; fax: +48 71 328 04 25. of the study was to determine the effect of deposition time on the pro-
E-mail address: juliusz.winiarski@pwr.wroc.pl (J. Winiarski). tective performance of the conversion coatings and to characterize their

0257-8972/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.surfcoat.2013.09.056
J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261 253

surface morphology, chemical composition and adhesion to the zinc repeated several times until a constant carbon content on the surface
substrate and the wear resistance of the samples in friction conditions. was obtained. The AVANTAGE ver. 4.75 (Thermoelectric) software was
used to register and process the spectra. The background line was deter-
2. Experimental details mined by the Shirley method. The position of the spectra was calibrated
relative to C 1 s 284.6 eV. The NIST X-ray Photoelectron Spectroscopy
2.1. Preparation of materials and conversion coatings Database [15] and the Handbook of X-ray Photoelectron Spectroscopy
[16] were used to identify the spectra.
Chromium-free conversion coatings were deposited on disks with DC polarization measurements were performed in aerated
diameter d = 14.8 mm and thickness h = 1.8 mm, made of carbon 0.5 mol dm−3 NaCl solution with pH = 6.5 using a (Schlumberger)
steel. Prior to deposition the surfaces of the samples had been prepared potentiostat equipped with a classic three-electrode cell. The counter
according to the procedure: grinding with abrasive paper with a grit of electrode was a platinum disk and the reference electrode was a saturat-
360–2000, rinsing with distilled water, degreasing in acetone in an ul- ed calomel electrode (SCE) equipped with a Luggin capillary. The work-
trasonic wash (T = 25 ± 1 °C, t = 1 min), drying and again rinsing ing electrode area exposed to the NaCl solution amounted to 1.0cm2. The
with water. Zinc coatings were electrolytically deposited from a slightly potential measured after 10min since the immersion of the sample in the
acid chloride bath SurTec®758 composed (g dm−3) of: ZnCl2 — 50, NaCl solution was adopted as the corrosion potential (Ecorr). After this
KCl — 250, H3BO3 — 5 and brightening additions, at pH = 5.4–5.7, T = time, the variation in the potential was smaller than 1 mV min−1. After
25 ± 1 °C and current density j = 1.5 A dm−2, while the bath was fixing Ecorr, polarization curves were recorded in the potentiodynamic
being stirred with a magnetic stirrer. The obtained zinc coatings were mode, in a potential range from −100 mV to +150 mV relative to
fine-grained and showed good luster. After 25 min of deposition their Ecorr at a potential change rate of 0.5 mV s−1.
thickness amounted to 12 μm. After zinc plating, the samples were EIS measurements were carried out using a Gamry Reference 600
rinsed with distilled water, activated in a 0.5 wt.% HNO3 solution at potentiostat. The counter electrode was a graphite rod. The working
25 ± 1 °C for 10s and again rinsed with water. Conversion coatings con- electrode area amounted to 1.0 cm2. The samples were immersed in
taining titanium were obtained by immersing the samples in a solution the corrosive medium (0.5 mol dm−3 NaCl) for 24 h. During immersion
composed (mmol dm−3) of: K2TiF6 — 8, H2O2 — 26, H3PO4 — 50 and the impedance spectra were recorded at the corrosion potential and at
Mn(NO3)2 — 30. All the reagents were of analytical grade. The bath's sinusoidal voltage excitation with an amplitude of 10 mV in a frequency
temperature amounted to respectively 25, 40 and 60 ± 1 °C and deposi- range of 100 kHz–10 mHz. The impedance data were displayed as
tion time was changed from 20 s to 300 s. On the basis of the previous Nyquist and Bode plots. The acquired data were curve fitted and ana-
XPS studies of similar chromium-free conversion coatings containing ti- lyzed using the ZView 3.2c (Scribner Associates) software.
tanium [11,14] it can be assumed that the approximate thickness of the The salt spray test was carried out in a SK 400M-TR (Liebisch) cham-
coatings amounted to 50–80 nm. The bath's pH was fixed at 3.0 by ber. Electroplated steel samples with conversion coatings were tested.
adding an appropriate amount of 10wt.% Na2CO3 solution. After conver- Samples had the form of 40 × 50 × 0.5 mm plates. The test conditions
sion coating the samples were rinsed with water and dried at room conformed to EN-ISO 9227:2006. The temperature in the chamber
temperature (25 °C) for 20 min. The bath used for depositing the con- was T = 35 °C and the salt concentration amounted to 50 g dm−3.
version coatings was stable and even after a few weeks of storage no in- The adhesion of the conversion coatings to the zinc coating was test-
soluble deposit would precipitate from it. Cr(III)-based conversion ed by the scratch method in a CSEM Micro-Combi-Tester using Rockwell
coatings were used as the reference for evaluating the properties of diamond indenter with a radius of 200 μm. The scratch track length was
the (Cr-free) coatings containing titanium. The Cr(III)-based treatment set to 5 mm. Maximum load Pmax was 20 N. The load was changed line-
was performed in a commercial SurTec®667 solution at room temper- arly and over the whole scratch track it ranged from 0.01 N to Pmax. The
ature (25 °C). The electrogalvanized samples were immersed in this indenter travel rate was set at 5 mm min−1. The measurements were
treatment bath for 30s (pH =1.9). The samples were then rinsed in dis- performed three times for each sample. The tests and the analysis
tilled water and dried at room temperature (25 °C) for 20 min. were carried out in accordance with PN-EN 1071-3. Dry friction wear
tests were carried out using a ball-on-plane type tribotester in rotational
2.2. Measurement and characterization motion in accordance with ISO 20808:2004. Three samples of each type
were tested. Wear resistance, as an indicator of the volumetric wear
The morphology of the samples coated with the conversion coatings (Wv (mm3 N−1 m−1)) of the samples, was calculated from the measure-
was examined using a VEGA II SBH (TESCAN) scanning electron micro- ments of the cross section of the groove which appeared as a result of
scope. Prior to examination the surface of the samples had been sputter- the friction of a polished Φ 6 mm ball made of Si3N4 under: normal
coated with a thin layer of gold. load Fn = 0.25 N, rotational speed n = 60 rpm−1, number of cycles N =
Thickness of the zinc coatings was measured using the magnetic 5000 and friction track radius R = 5 mm. The groove's geometry was
induction method and a DUALSCOPE FMP40 (Fischer) thickness gauge measured in four places (at every 90°) on its circumference by the con-
equipped with an FGAB1.3 probe. tact method using a Micro-Combi-Tester device.
The chemical composition of the conversion coatings was studied
using X-ray photoelectron spectroscopy (XPS). Photoelectron spectra 3. Results and discussion
were recorded by means of an ESCALAB-210 (VG Scientific) spec-
trometer using a nonmonochromatized X-radiation Al-Kα = 1486.6 eV 3.1. SEM and XPS characterization
(14.5 kV, 20 mA). The measurements were conducted in the analyzer's
chamber in a vacuum of below 5·10−9 mbar. A survey scan was All the samples with the Cr-free coatings deposited at a temperature
recorded for an energy range of 1350 eV to 0 eV with a pitch of 0.75 eV of 60 °C were characterized by luster, light blue–yellow color and an in-
at constant analyzer energy CAE = 100 eV. High resolution spectra delible milky tarnish. Regardless of deposition time, the coatings were
were recorded with a pitch of 0.1 eV at CAE = 40 eV at a counting time homogenous and devoid of cracks.
of 50 ms. The sample would be placed perpendicularly to the analyzer The compositions of the coatings deposited for 300 s from the bath
axis and the X-ray source would be set at an angle of 60°. The first anal- based on K2TiF6 were determined using X-ray photoelectron spectros-
ysis of the surface was carried out in order to preliminarily determine copy (XPS). The high resolution spectra recorded after the surface had
the carbon content. Then the surface of the sample was cleaned with been cleaned with a beam of Ar+ ions (E = 2 keV, i = 100 μA) revealed
an Ar+ (2keV, 100μA) ion beam using an AG-21 argon gun, maintaining that the coating contained (% at.): C — 10.58, O — 55.18, Ti — 4.31,
argon pressure at a level of 3.5·10−6–4.0·10−6 mbar. Cleaning was Mn — 3.09, P — 13.48 and Zn — 8.59. Moreover, small quantities of
254 J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261

Fig. 1. High-resolution spectra of photoelectrons C 1s (a) and O 1s (b) for conversion coating deposited from bath based on K2TiF6 for 300 s (T = 60 °C, pH = 3.0), recorded after surface
cleaning with Ar+ ions.

(% at.): K — 1.8, N — 1.31 and F — 1.66 were determined. Fig. 1 shows the occurs at 642.58 eV (Fig. 2b), which may be indicative of the presence
spectra of photoelectrons: C 1s, K 2p and O 1s. Peak C 1s can be of Mn(IV) compounds in the coating [23], e.g. MnO2 [20] or MnO2·xH2O
decomposed into three component peaks. The energy (284.6 eV) of [26]. However, due to the relatively large Full Width at Half Maximum
the first peak is characteristic of the carbon–carbon or carbon–hydrogen value of Mn 2p3/2 peak (FWHM = 3.77 eV) it can be concluded, that
bond (Fig. 1a) [13,17,18]. This subpeak can be related to adsorbed or- the existence of small amounts of Mn(III) compounds, e.g. Mn2O3 [15]
ganic impurities due to the atmospheric exposure of the sample [17]. in the coating is possible. The small components of Mn 2p spectra
The second peak with an energy of 285.99 eV is characteristic of carbon visible respectively at 646.2 eV and 660.3 eV are the weak satellites of
bond to oxygen (Fig. 1a) [13,17]. The component peak with the highest Mn 2p3/2 and Mn 2p1/2 peaks. The existence of Mn(II) compounds in
energy (289.09 eV) indicates the presence of carbonates [13,17]. Hence the coating is rather dubious, due to the absence of a satellite peak at
it can be concluded that potassium occurs in the form of carbonates in about +5 eV from the Mn 2p3/2 [23,27]. Peak N 1s is symmetric and
the coating. Peak O 1s can be decomposed into three component its maximum occurs at 400.24 eV (Fig. 3a). The peak is difficult to
peaks (Fig. 1b). The energy (530.39 eV) of the first peak indicates the deconvolute. One can only surmise that the analyzed coating does not
presence of metal oxides (M–O in Fig. 1b) [13,19–21]. The second include nitrates since nitrogen bonding energy in nitrates is ~407 eV
peak (531.83 eV), dominating in the O 1s spectrum, is characteristic of [16]. The maximum of peak Zn 2p3/2 occurs at 1022.9 eV (Fig. 3b)
hydroxyl groups in metal hydroxides [19,21,22], phosphates [13] and which evidences, that the coating is composed mainly of Zn(II) com-
also of CO2−3 groups [23] (M–OH and P–O in Fig. 1b). The peak at pounds. On the basis of a single spectrum of Zn 2p3/2 one cannot
533.21 eV corresponds to the oxygen–hydrogen bond [13,15] (H2O in unequivocally determine the form in which zinc occurs in the investi-
Fig. 1b), which is characteristic of adsorbed water [21,23,24]. Fig. 2 gated coating due to the very small difference between the metallic
shows the spectra of Ti 2p and Mn 2p. The maximum of peak Ti 2p3/2 zinc bonding energy and the bonding energy of zinc in zinc oxide or
corresponds to 459.4 eV (Fig. 2a). The maximum of peak Ti 2p3/2 indi- zinc hydroxide [13,19]. The maximum of peak Zn 2p3/2 occurs at
cates the presence of titanium (IV) oxide in the analyzed coating 1022.9 eV (Fig. 3b) which evidences, that the coating is composed
[15,25]. This supposition seems to be confirmed by the difference in mainly of Zn(II) compounds. Nevertheless, peak Zn 2p3/2 is not fully-
BE (~5.5 eV) between peak Ti 2p1/2 and peak Ti 2p3/2, which is similar symmetric, as evidenced by the inflection point on the side of lower en-
as for the clean spectrum of TiO2 [25]. The maximum of peak Mn 2p3/2 ergies. This asymmetry indicates the presence of small amounts of

Fig. 2. High-resolution spectra of photoelectrons Ti 2p (a) and Mn 2p (b) for conversion coating deposited from bath based on K2TiF6 for 300 s (T = 60 °C, pH = 3.0), recorded after surface
cleaning with Ar+ ions.
J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261 255

Fig. 3. High-resolution spectra of photoelectrons: N 1s (a), Zn 2p (b), P 2p (c) and F 1s (d) for conversion coating deposited from bath based on K2TiF6 for 300 s (T = 60 °C, pH = 3.0),
recorded after surface cleaning with Ar+ ions.

metallic zinc on the analyzed surface, which may be due to the reduc- 1 μA cm−2. As deposition time is increased, the rate of corrosion of the
tion of Zn2+ to Zn0 under X-ray radiation or due to the nonuniformities samples decreases. The lowest values of jcorr were measured in the
in the coating thickness. Peak P 2p is symmetric and its maximum oc- case of the samples with the coating deposited for respectively 180 s
curs at ~134 eV (Fig. 3c). This value may indicate the presence of the and 300 s (jcorr ~ 0.3 μA cm−2). This value is nearly twenty times
HPO2−4 groups [24] or K3PO4 [24,28]. BE= 685.09 eV, to which the max- lower than the corrosion current for the sample with only zinc coating
imum of peak F 1s corresponds (Fig. 3d), indicates that fluorine occurs (jcorr = 5.8 μA cm−2, Ecorr = −1.04 V) and it is close to jcorr calculated
in the deposited conversion coating, probably in the form of fluorides for the electrogalvanized steel with Cr(III)-based conversion coating
[17] or other fluorocompounds, e.g. K2TiF6 (BE = 685.0 eV) [15]. (0.62 μA cm−2, Ecorr = −1.13 V).
The results of the XPS analyses showed that the obtained conversion Fig. 4 shows polarization curves for the samples with the Cr-free
coatings contain mainly: C, O, Ti, Mn, P and Zn. However, on the basis of coatings and for comparison, for a sample with Cr(III)-based conversion
the recorded high resolution XPS spectra it was not possible to unequiv- coatings and a sample with only a zinc coating. All the conversion coat-
ocally determine the chemical compounds present in the coating. Metal ings deposited on electrogalvanized steel significantly lower corrosion
oxides and hydroxides, phosphates, carbonates and small amounts of current density. The polarization curves of the two samples with the
fluorocompounds predominate. Cr-free coatings have a similar shape. No differences in the slopes of
either the anodic or cathodic branches are observed. Only the Ecorr of
3.2. DC polarization measurements

DC polarization measurements were carried out in a non-deaerated Table 1


0.5 mol dm−3 NaCl solution at pH = 6.5 and T = 25 °C. After 10 min of Results of polarization measurements for samples with conversion coatings, estimated
from potentiodynamic polarization curves. Measurements were performed in aerated
immersion the open circuit potential was relatively stable and the cor-
0.5 mol dm−3 NaCl solution at pH = 6.5.
rosion resistance of the samples could be preliminarily compared on
the basis of DC measurements. Table 1 shows the corrosion potentials Deposition time T = 25 °C T = 40 °C T = 60 °C
t (s)
(Ecorr) and corrosion current densities (jcorr) measured for the samples Ecorr jcorr Ecorr jcorr Ecorr jcorr
with Cr-free coatings, which were deposited from the bath based on (V) (μA cm−2) (V) (μA cm−2) (V) (μA cm−2)
K2TiF6. It was noticed that the higher the temperature of the bath, the 20 −1.01 1.5 −1.09 2.0 −1.17 0.5
lower the jcorr values of the coatings. One can suppose that higher 60 −1.05 3.9 −1.08 1.0 −1.11 0.5
bath temperature favors the deposition of more compact coatings 180 −1.09 2.1 −1.07 1.4 −1.11 0.3
300 −1.07 2.0 −1.12 6.6 −1.10 0.3
with better barrier properties. At 60 °C the corrosion current is below
256 J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261

3.3. Impedance measurements

Impedance spectra (each time preceded by corrosion potential


measurement) were recorded after 3h and 24h of immersion in aerated
0.5moldm−3 NaCl solution at room temperature. They are shown as the
Nyquist impedance plots in Fig. 5 (Fig. 5a — for electrogalvanized steel;
Fig. 5b — for Cr(III)-based conversion coating on zinc; Fig. 5c and d — for
electrogalvanized steel with a titanium containing conversion coating
deposited for respectively 60 s and 300 s at T = 60 °C and pH = 3.0).
Fig. 6 shows Bode plots with fitting lines for the investigated samples.
The distinct semicircle in the frequency range of 104 Hz–0.5 Hz in
the zinc coating spectrum (Fig. 5a) indicates that zinc oxidation is con-
trolled mainly by the charge transfer reaction on the metal-solution
phase boundary. The small fragment of the flattened semicircle in
the range of the lowest frequencies indicates some diffusion con-
straints imposed on mass transport by the layer of corrosion products
present on the surface or in the defects in the zinc coating [29]. The ob-
tained experimental data are in quite good agreement with the equiv-
alent circuit shown in Fig. 7a. In this equivalent circuit, R1 represents
Fig. 4. Polarization curves recorded for electrogalvanized steel, Cr(III)-based conversion the electrolyte resistance, CPE1 and R2 represent respectively the ca-
coating and Ti-containing conversion coatings deposited for 60 s and 300 s from bath
based on K2TiF6. All curves determined in non-deaerated 0.5 mol dm−3 NaCl solution at
pacitive and resistive responses of the coating, CPE2 corresponds to
pH = 6.5. the double layer capacitance and R3 represents the charge transfer re-
sistance. In order to take the diffusion effect into account, well known
diffusional element Ws (Finite Length Warburg — Short Circuit Termi-
the sample with the coating deposited for 60s is slightly shifted towards nus [30]) connected in series with R3 was implemented in the model.
negative values relative to the sample with the coating deposited for The results obtained from the fitting procedure are shown in Table 2.
300 s. Small current density fluctuations, noticeable particularly in the The low values of parameter χ2 prove that thanks to the implementa-
cathodic–anodic transition area, are conspicuous. They may be indica- tion of this quite simple equivalent circuit a good fit to the experimen-
tive of considerable surface development of the deposited coatings tal data was achieved (Fig. 6). Since all the recorded spectra do not
and their inadequate density. The cathodic branch of the polarization have a purely capacitive character, a constant phase element (CPE)
curve for the sample with the Cr(III)-based conversion coating has a was used instead of the capacitor in order to fit the model better to
small slope which indicates that the rate of the cathodic process is the experimental data. The impedance of the CPE is defined by the
somewhat constrained. The anodic branch has a steep slope and its formula: ZCPE = T−1 (jω)−P, where: T — a time constant parameter
passive range is about 100 mV wide. The anodic process is markedly (s−P Ω−1 cm−2), ω — the angular frequency of the AC signal (rad−1)
constrained and thereby limits the corrosion rate (Fig. 4). The wide pas- and P — the CPE exponent. The capacitance of the double layer was
sive range is indicative of good protective properties of the trivalent calculated using the equation Cdl = T(ω*)P − 1, where ω* = (RT)−1/P
chromium conversion coating. The shape of the polarization curve (R is the charge transfer resistance connected in parallel with CPE)
obtained for the sample with the zinc coating (without a conversion [31]. According to the above equations, in the 3–24 h exposure range
coating) shows that the reaction of the anodic oxidation of the zinc pro- the double layer capacitance slightly decreases from 61.2 (μF cm−2) to
ceeds without any major constraints. The rate of corrosion is clearly 59.6 (μFcm−2). Also charge transfer resistance decreases, whereas coat-
constrained by the cathodic process, i.e. probably the reduction of the ing resistance increases nearly tenfold (Table 2). The decrease in the
oxygen dissolved in aerated 0.5 mol dm−3 NaCl solution. value of CPE2-P combined with an increase in the value of R2 can be

Fig. 5. Nyquist plots of EIS data obtained from: electrogalvanized steel (a), Cr(III)-based conversion coating (b), Cr-free conversion coatings deposited for 60 s (c) and 300 s (d) at different
times of exposure to 0.5 mol dm−3 NaCl solution.
J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261 257

Fig. 6. Bode plots of EIS data obtained after 3 h of immersion in 0.5 mol dm−3 NaCl solution for: electrogalvanized steel (a), Cr(III)-based conversion coating (b), Cr-free conversion coatings
deposited for 60 s (c) and 300 s (d) and after 24 of immersion for: zinc coated steel (e), Cr(III)-based conversion coating (f) and Cr-free conversion coatings deposited for 60 s (g) and 300 s (h).

caused by the accumulation of zinc coating corrosion products, while (Table 3). The resistance (R2) of the Cr(III)-based coating seems to be
the slightly decreasing charge transfer resistance (R3) is evidence of constant during the 24 h of immersion, amounting to 1 kΩ cm2. The
increasing electrolyte/steel contact area. Generally, the investigated values of CPE2-T and CPE2-P also seem to be constant, which confirms
system can be well described by the same equivalent electric circuit the finding that such properties of the investigated system as: the distri-
for 3–24 h of immersion of the sample in NaCl solution. bution of the reaction rates, the surface roughness and the heterogene-
In the case of the trivalent chromium conversion coating on zinc, ity of the conversion coating do not change. The resistance of the charge
two time constants can be seen in the Bode plot (Fig. 6b and f). The transfer reaction decreases from 28.7 kΩ cm2 to 21.0 kΩ cm2, which is
first of them, being in the high frequency range, is associated with the probably the consequence of the slow degradation of the coating. Nev-
properties of the conversion coating while the second one, in the ertheless, after 24 h of exposure the impedance modulus for this sample
range of lower frequencies, is associated with the corrosion process. still exceeded 20 kΩ cm2, which is evidence of the very good protective
The equivalent circuit shown in Fig. 7b was used for such coatings. In properties of the Cr(III)-based conversion coating.
this equivalent circuit, R1 represents the electrolyte resistance, CPE1 The Nyquist impedance spectra for the samples with Cr-free conver-
and R2 represent respectively the capacitive and resistive responses of sion coatings contain three capacitive loops, but only the second and the
the coating defects, CPE2 corresponds to the double layer capacitance third one are visible (Fig. 5c and d). The first capacitive loop at the
and R3 represents the charge transfer resistance. The increasing value highest frequencies is not visible since its time constant is much smaller
of the imaginary part (Z″) of the impedance in the lowest frequency than those of the other loops. In order for the capacitive loops to be
range (10−1–10−2 Hz) is evidence of diffusion constraints in the corro- clearly separated in the Nyquist system the following conditions apply-
sion process (Fig. 5b). In order to take this phenomenon into account, a ing to resistance ratio (1) and time constants (2) should be satisfied:
constant phase element (CPE3) connected in series with R3 was added 0.2 b (R1/R2) b 5 (1) and (R1C1)/(R2C2) N 20 (2), where: R1 — a resis-
to the equivalent circuit. A good fit to the experimental data (Fig. 6b tance connected in parallel with capacitance C1; R2 — a resistance
and f) was obtained as indicated by the low values of parameter χ2 connected in parallel with capacitance C2.
258 J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261

and h), as evidenced by the low values of parameter χ2 (Table 4). The
first time constant can be associated with the diffusion processes caused
by the presence of a layer of corrosion products (CPE1 and R2) on the
conversion coating. The second time constant is representative of the
conversion coating and is described by the coating capacitance (CPE2)
and the coating resistance (R3). The third time constant is determined
by the charge transfer resistance (R4) and by the double layer capaci-
tance (represented by CPE3). As one can see in Fig. 5c and d, the imped-
ance spectra for the titanium containing coatings deposited for both 60 s
and 300 s show a similar shape. According to the results presented in
Table 4, after 3h of exposure to the 0.5moldm−3 NaCl solution the resis-
tance of the layer of corrosion products on the surface of the conversion
coating (R2) deposited for 300 s (R2 = 176 Ω cm2) is lower than that of
the coating deposited for 60s (R2= 287Ωcm2). Both the charge transfer
resistance (R4) and the resistance associated with the conversion coating
(R3) assume much higher values (N 6 kΩ cm2) for the sample with the
coating deposited for 300 s. After 24 h of exposure to the 0.5 mol dm−3
NaCl solution the three phase angle maxima are still visible in the Bode
plots for the two samples (Fig. 6d and h) and the impedance modulus
reaches lower values than after 3 h of exposure. The samples with the
coating deposited over the longer time (300s) show much better protec-
tive properties after both the times of exposure to the chloride solution.
A comparison of the behavior of this coating after 3 and 24 h of immer-
sion in the 0.5 mol dm−3 NaCl solution shows that the coating's resis-
tance (R3) decreases, reaching 1740 Ω cm2. The resistance (R2) of the
layer of corrosion products also decreases (Table 4). This trend indicates
the gradual degradation of the conversion coating. The charge transfer
resistance (R4) decreases threefold, while the relatively high double
layer capacitance increases over tenfold (from 4.23·10−4 F cm−2 to
48.1·10−4 F cm−2). Such a large increase in double layer capacitance
may be caused by electrolyte penetration under the conversion coating
in places where the latter is damaged. This would also explain the trend
in R4 values. Nevertheless, the EIS measurement results indicate that
the titanium containing Cr-free conversion coatings increase the corro-
sion resistance of zinc coatings in a chloride solution.

3.4. Neutral salt spray test

Through DC polarization measurements it is possible to determine


the instantaneous rate of corrosion of samples and to evaluate the bar-
rier properties of the oxidation products forming during exposure. In
the case of unstable systems, to which the investigated conversion coat-
ings belong, one should take into account that the obtained results may
significantly differ from the corrosion resistance in real conditions. For
this reason, the samples coated with the titanium containing coatings
were subjected to tests in a salt spray chamber. The degree of rusting
of the samples' surface after each successive 24 h exposure to neutral
salt spray atmosphere was evaluated. Fig. 8 shows the condition of the
surface of the samples before exposure and after 96 h of testing in the
salt spray chamber. After the first 24 h of testing no traces of corrosion
were found on the surface of the tested samples. Tarnish composed of
zinc corrosion products formed on only the steel sample coated with
the zinc coating, but without the conversion coating. The appearance
of the surface of the plates changed after 48 h of exposure. The sample
with Cr-free conversion coating deposited for 60 s underwent consider-
Fig. 7. Models of equivalent circuits proposed for curve fitting of EIS data for: electro-
able corrosion whereas the sample with the coating deposited for 300 s
galvanized steel (a), Cr(III)-conversion coating on zinc (b) and Cr-free conversion coating (c).
showed only a few spots of corrosion of the zinc substrate, situated
mainly on the edges. The sample with the zinc coating was covered
On the Bode plots for these samples one can see clearly three phase with zinc corrosion products on more than 50% of its surface. The
angle maxima (Fig. 6c, d, g and h). The first maximum can be observed Cr(III)-based conversion coating continued to provide total protection
at high frequencies (10 kHz–2 kHz), the second one at middle fre- to the substrate.
quencies (100Hz–0.5Hz) and the third one appears at lower frequencies After 96h of exposure, zinc corrosion products occurred on 5% of the
(0.5 Hz–0.01 Hz). In this case, a two-layer model (equivalent circuit on surface of the sample with the Cr-free coating deposited for 300 s. In the
Fig. 7c) has been proposed, taking into account the effect of the corrosion case of the sample with a coating deposited for 60 s, the degree of
products present on the conversion coating's surface and in its hollows. rusting was substantially greater. In the case of the steel coated with
The circuit ensured a good fit to the experimental data (Fig. 6c, d, g only a zinc coating, serious damage to the coating, as evidenced by the
J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261 259

Table 2
Fitting results for EIS data acquired for zinc coated steel in aerated 0.5 mol dm−3 NaCl solution at pH = 6.5.

Time R1 CPE1-T CPE1-P R2 CPE2-T CPE2-P R3 Ws1-R Ws1-T Ws1-P χ2 10−4


(h) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (Ω cm2) (Ω cm2) (sP)

3 5.8 135 0.91 10.6 99.5 0.87 390 319 56.7 0.40 6.8
24 6.0 333 0.80 101 66.9 0.97 355 247 35.3 0.44 4.8

Table 3
Fitting results for EIS data acquired for Cr(III)-based conversion coating on zinc in aerated 0.5 mol dm−3 NaCl solution at pH = 6.5.

Time R1 CPE1-T CPE1-P R2 CPE2-T CPE2-P R3 CPE3-T CPE3-P χ2 10−4


(h) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (kΩ cm2) (10−6 Ω−1 cm−2 s−P)

3 3.3 12.9 0.90 923 36.2 0.62 28.7 220 0.37 2.4
24 3.6 15.2 0.88 1003 37.2 0.65 21.0 356 0.66 3.3

Table 4
Fitting results for EIS data acquired for Cr-free conversion coatings in aerated 0.5 mol dm−3 NaCl solution at pH = 6.5 versus deposition time (t).

Deposition Time R1 CPE1-T CPE1-P R2 CPE2-T CPE2-P R3 CPE3-T CPE3-P R4 χ2 10−4


time (h) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (Ω cm2) (10−6 Ω−1 cm−2 s−P) (kΩ cm2) (10−4 Ω−1 cm−2 s−P) (kΩ cm2)
(t)

t = 60 s 3 2.9 16.8 0.71 287 41.6 0.67 4.44 7.29 0.95 3.24 8.9
24 6.3 63.0 0.69 47.8 71.4 0.76 0.78 150 0.83 1.06 5.9
t = 300 s 3 4.7 26.5 0.69 176 29.8 0.74 6.10 4.07 0.96 6.31 7.8
24 3.9 25.8 0.67 125 36.2 0.75 1.74 39.2 0.91 2.01 6.0

presence of the products of corrosion of the steel substrate, took place. better performance of the samples in the salt spray test. This observa-
The trivalent chromium conversion coating continued to perform best. tion is consistent with the results of the DC polarization and impedance
Only small white patches, without any signs of corrosion of the zinc sub- measurements.
strate, could be noticed on this sample.
The results of the salt spray tests indicate that the Cr(III)-based 3.5. Adhesion and wear resistance
conversion coatings have better protective properties than the titanium
containing conversion coatings. The latter when deposited on a zinc The degradation of the Cr-free conversion coatings deposited from
coating markedly improve the corrosion resistance of the system. The the bath based on K2TiF6 observed in the course of the scratch test (t =
extension of Cr-free coating deposition time from 60 to 300 s results in 20–300 s, T = 60 °C, pH = 3.0) was similar for all the tested samples.

Fig. 8. Condition of surface of: electrogalvanized steel (a), sample with Cr(III)-based conversion coating (b), sample with Cr-free coating deposited for 60 s (c) and 300 s (d), initially and
after 96 h of exposure to neutral salt spray.
260 J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261

Fig. 9. Appearance of surface of sample with Cr-free coating deposited for 20 s (T = 60 °C, pH = 3.0) during scratch test after force: LC1 (a), LC2 (b) and LC3 (c) was reached; magnification
200×.

For instance, in the case of the coating deposited for 20 s, the conversion increase in the volumetric wear index, which in the case of the sample
coating was found to initially undergo cohesive cracking ahead of the with the coating deposited for 300 s reached 1661·10−6 mm3 N−1 m−1.
moving indenter (Fig. 9) as a result of the plastic deformation of the This dependence indicates that as the thickness of the Cr-free conver-
substrate and the large tensile stresses in the conversion coating sion coatings increases, the wear resistance decreases. Moreover, an
under load LC1 of 1.2 N. An increase in the load to LC2 (4.5 N) would analysis of the obtained results shows that the wear resistance of all
lead to the first coating abrasions on the scratch track. Load LC3 (8.1 N) the samples with the Cr-free coatings is relatively low and at least
resulted in the complete removal of the coating from the scratch twice lower than that of the reference samples covered with a Cr(III)-
track. Regardless of deposition time, the adhesion of the Cr-free conver- based conversion coatings.
sion coatings to the zinc substrate was relatively constant (Table 5).
Load LC1 increases slightly with deposition time, reaching a maximum
for the coating deposited for 60s (1.6N). Loads LC2 and LC3 assume max- 4. Conclusions
imum values (respectively 7.7 N and 10.7 N) for the coating deposited
for 60 s. In the case of the reference Cr(III)-based conversion coating,
1). From a bath based on K2TiF6 (T = 60 °C) one can deposit homoge-
the measured loads were greater (amounting to respectively 8.3 N
nous Cr-free coatings. The bath is characterized by relatively high
and 14.6 N). Of all the tested samples, the trivalent chromium conver-
stability.
sion coatings are characterized by the best adhesion.
2). The obtained Cr-free conversion coatings contain the following ele-
Wear resistance of the samples was determined from measurements
ments: C, O, Ti, Mn, P, Zn and small quantities of: K, N and F. The
of the profile of the groove formed as a result of rubbing with a ball (the
main compounds forming the coating are probably: oxides, hydrox-
ball, 6 mm in diameter, was made of Si3N4 and had a polished surface).
ides, phosphates and carbonates.
The groove's width (w), depth (d), cross sectional area (Ap) and the
3). The results of the DC polarization and EIS measurements showed
wear index (Wv) after friction for the tested samples are shown in
that regardless of deposition time length, the obtained conversion
Table 6. From among the tested samples with Cr-free coatings the sam-
coatings (Ti-containing) increase the corrosion resistance of electro-
ple with the coating deposited for 20s (Wv = 926·10−6 mm3 N−1 m−1)
galvanized steel from ten to twentyfold. The lowest value of jcorr =
showed the greatest wear resistance while the wear resistance of the
0.3 μA cm−2 was measured for the coatings deposited for 300 s at
sample with the coating deposited for 60 s was only slightly worse.
pH = 3.0 and T = 60 °C. This value is close to that for the Cr(III)-
The extension of coating deposition time results in nearly a twofold
based conversion coating on zinc. The results of the EIS measure-
ments and the salt spray tests show that the protective performance
of the coatings deteriorates as their time of service in an environ-
Table 5
ment containing NaCl increases.
Scratch test results for Ti containing conversion coatings (deposition conditions: T = 60 °C,
pH = 3.0) as function of deposition time, and for reference Cr(III)-based conversion 4). When the time of deposition of Cr-free conversion coatings is ex-
coating. Standard deviation values are given in brackets. tended from 20 s to 300 s, the wear resistance of the samples
decreases. The coatings deposited for 60 s show the best adhesion
Deposition time LC1 LC2 LC3
t (s) (N) (N) (N)
from all the deposited Cr-free coatings. The tested reference triva-
lent chromium conversion coatings are characterized by better
20 1.2 (0.2) 4.5 (0.5) 8.1 (0.3)
adhesion and wear resistance than the Ti-containing coatings.
60 1.6 (0.1) 7.7 (0.4) 10.7 (0.3)
180 1.5 (0.2) 6.3 (0.5) 8.4 (0.4) 5). The results of investigations indicate that titanium containing con-
300 1.5 (0.2) 6.1 (0.7) 8.1 (0.7) version coatings should be considered as potential substitutes for
Cr(III) 3.2 (0.3) 8.3 (0.8) 14.6 (1.1) the more toxic chromate coatings.

Table 6
Volumetric wear test results for samples with conversion coatings deposited from bath based on K2TiF6 (deposition conditions: T = 60 °C, pH = 3.0) and for Cr(III)-based conversion
coatings after rubbing with Si3N4 ball at Fn = 0.5 N, n = 60 rpm−1, R = 5 mm, N = 5 000 cycles. Standard deviation values are given in brackets.

Deposition time Groove width Groove depth Cross sectional area Volumetric wear
t (s) w (μm) d (μm) of the groove Wv·10−6 (mm3 N−1 m−1)
Ap (μm2)

20 392 (43) 5.1 (1.0) 1157 (221) 926 (176)


60 477 (93) 5.0 (0.9) 1247 (462) 998 (370)
180 625 (103) 5.9 (0.5) 1761 (347) 1254 (246)
300 746 (199) 6.8 (0.1) 2077 (75) 1661 (60)
Cr(III) 361 (80) 3.5 (0.5) 680 (17) 544 (14)
J. Winiarski et al. / Surface & Coatings Technology 236 (2013) 252–261 261

Acknowledgments [12] M.A. Smit, J.A. Hunter, J.D.B. Sharman, G.M. Scamans, J.M. Sykes, Corros. Sci. 45
(2003) 1903–1920.
[13] S. Le Manchet, D. Verchère, J. Landoulsi, Thin Solid Films 520 (2012) 2009–2016.
This work was supported by a grant from the Ministry of Science and [14] B. Szczygieł, J. Winiarski, W. Tylus, Mater. Chem. Phys. 129 (2011) 1126–1131.
Higher Education of Poland (Grant N N209 022039). [15] C.D. Wagner, A.V. Naumkin, A. Kraut-Vass, J.W. Allison, C.J. Powell, J.R. Rumble Jr.,
NIST Standard Reference Database 20, Version 4.1, 2003.
The adhesion and wear resistance measurements were done within [16] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, in: J. Chastain, R.C. King Jr. (Eds.),
statutory activity by the Ministry of Science and Higher Education of Handbook of X-ray Photoelectron Spectroscopy, Ulvac-PHI Inc., Physical Electronics
Poland (contract no. 20140). USA, Inc., 1995
[17] P. Hammer, A.P. Rizzato, S.H. Pulcinelli, C.V. Santilli, J. Electron Spectrosc. 156–158
(2007) 128–134.
[18] L. Martinez, E. Roman, J.L. de Segovia, S. Poupard, J. Creus, F. Pedraza, Appl. Surf. Sci.
References 257 (2011) 6202–6207.
[19] B. Szczygieł, A. Laszczyńska, W. Tylus, Surf. Coat. Technol. 204 (2010) 1438–1444.
[1] Y.W. Song, D.Y. Shan, R.S. Chen, E.H. Han, Surf. Coat. Technol. 204 (2010) 3182–3187. [20] G. Wang, M. Zhang, R. Wu, Appl. Surf. Sci. 258 (2012) 2648–2654.
[2] A.A.O. Magalhães, I.C.P. Margarit, O.R. Mattos, J. Electroanal. Chem. 572 (2004) [21] X. Zhang, W.G. Sloof, A. Hovestad, E.P.M. van Westing, H. Terryn, J.H.W. de Wit, Surf.
433–440. Coat. Technol. 197 (2005) 168–176.
[3] L. Fachikov, D. Ivanova, Appl. Surf. Sci. 258 (2012) 10160–10167. [22] H. Ardelean, I. Frateur, S. Zanna, A. Atrens, P. Marcus, Corros. Sci. 51 (2009) 3030–3038.
[4] Y. Hamlaoui, C. Rémazeilles, M. Bordes, L. Tifouti, F. Pedraza, Corros. Sci. 52 (2010) [23] B. Kucharczyk, W. Tylus, Appl. Catal. A 335 (2008) 28–36.
1020–1025. [24] T. Ishizaki, R. Kudo, T. Omi, K. Teshima, T. Sonoda, I. Shigematsu, M. Sakamoto,
[5] V. Dikinis, G. Niaura, V. Rėzaitė, I. Demčenko, R. Šarmaitis, Trans. Inst. Met. Finish. 85 Electrochim. Acta 62 (2012) 19–29.
(2007) 87–93. [25] S. Hashimoto, A. Tanaka, Surf. Interface Anal. 34 (2002) 262–265.
[6] R. Berger, U. Bexell, T.M. Grehk, S.E. Hörnström, Surf. Coat. Technol. 202 (2007) [26] F.A. AL-Sagheer, M.I. Zaki, Colloids Surf. A 173 (2000) 193–204.
391–397. [27] D. Kovacheva, T. Trendafilova, K. Petrov, A. Hewat, J. Solid State Chem. 169 (2002)
[7] L. Fedrizzi, F. Deflorian, P.L. Bonora, Electrochim. Acta 42 (1997) 969–978. 44–52.
[8] B. Wilson, N. Fink, G. Grundmeier, Electrochim. Acta 51 (2006) 3066–3075. [28] R. Zeng, Z. Lan, L. Kong, Y. Huang, H. Cui, Surf. Coat. Technol. 205 (2011) 3347–3355.
[9] M.A. Smit, J.M. Sykes, J.A. Hunter, J.D.B. Sharman, G.M. Scamans, Surf. Eng. 15 (1999) [29] L. Fedrizzi, L. Ciaghi, P.L. Bonora, R. Fratesi, G. Roventi, J. Appl. Electrochem. 22
407–410. (1992) 247–254.
[10] S. Le Manchet, J. Landoulsi, C. Richard, D. Verchère, Surf. Coat. Technol. 205 (2010) [30] ZPlot for Windows, Electrochemical Impedance Software, Scribner Associates,
475–482. Southern Pines, 2001.
[11] J. Winiarski, B. Szczygieł, Ochr. Przed Koroz. 5 (2012) 224–228. [31] J. Kubisztal, J. Niedbała, A. Budniok, Surf. Interface Anal. 42 (2010) 1222–1225.

You might also like