10
Electrochemical impedance spectroscopy and linear polarization applied to evaluation of porosity of phosphate conversion coatings on electrogalvanized steels Vanessa de Freitas Cunha Lins a, * , Geraldo Francisco de Andrade Reis b,1 , Carlos Roberto de Araujo a , Tulio Matencio b,1 a Corrosion and Surface Engineering Laboratory, Chemical Engineering Department, Federal University of Minas Gerais, Rua Espirito Santo 35, 6th floor, Belo Horizonte 30160-030, Brazil b Chemical Department, Federal University of Minas Gerais, AV. Antonio Carlos, 6627-Pampulha, Belo Horizonte 31270-901, Brazil Received 22 August 2005; received in revised form 3 June 2006; accepted 3 June 2006 Available online 11 July 2006 Abstract In this work, electrochemical impedance spectroscopy and linear polarization are used in determining porosity of zinc phosphates, and of nickel and manganese modified zinc phosphates on electrogalvanized steel. The porosity of the phosphate layers ranges from 0.1% for the manganese- modified hopeite to 8% for hopeite, using the linear polarization and the electrochemical impedance spectroscopy techniques. The porosity values measured using the two techniques were in agreement. All impedance spectra of coated samples showed two steps, with two capacitive arcs and an inductive looping. # 2006 Elsevier B.V. All rights reserved. Keywords: Phosphate conversion coatings; Electrochemical impedance spectroscopy; Porosity; Linear polarization 1. Introduction Phosphate conversion coatings are used by industries as a surface pre-treatment of metals before painting, promoting the paint adherence and protecting the substrate against corrosion. Conversion treatments are intended to interpose an inorganic layer providing both a strong bond to the metal surface and an adhesive base for the polymeric film and to slow down by long- term inhibition the destruction of the bonding by general under- film corrosion due to oxygen diffusing through the coating and by localized galvanic effects at any flaws in the organic film [1]. The extensive reviews on various aspects of phosphate coatings are given by Lorin [2], Machu [3], Rausch [4] and Freeman [5]. The phosphate layer on carbonsteel shows two phases: hopeite [Zn 3 (PO 4 ) 2 4H 2 O] and phosphophillite [Zn 2 Fe(PO 4 ) 2 4H 2 O]. As the ratio phosphophillite/hopeite increases the adhesion and corrosion resistance of the phosphate layer increases [6]. The literature usually deals with phosphate layers on steel [7,8] so the development of new methods and models for phosphate layers on zinc is of great interest. The phosphate coatings on galvanized steels present only hopeite that has low resistance in alkaline and acid environ- ments [9]. The addition of cations, such as manganese and nickel, to the hopeite is an alternative to improve the resistance of the phosphate coatings in alkaline and acid environments. There is scarce literature data on physical and chemical properties of these modified zinc phosphates [9,10]. One problem in the use of phosphate coatings in aggressive media is the existence of open porosity and pinholes [2,11,12]. These local defects form direct paths between the corrosive environment and the substrate [12]. As the corrosion reactions are initiated at the coating–substrate interface, determination of porosity is important to estimate the overall corrosion resistance of the coated materials [12]. Notter and Gabe [11] clearly describe the nature and effect of porosity on the corrosion resistance of electrodeposited coatings. They point www.elsevier.com/locate/apsusc Applied Surface Science 253 (2006) 2875–2884 * Corresponding author. Tel.: +55 31 32381773; fax: +55 31 32381789. E-mail addresses: [email protected] (V. de Freitas Cunha Lins), [email protected] (T. Matencio). 1 Tel.: +55 31 34995758; fax: +55 31 34995700. 0169-4332/$ – see front matter # 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2006.06.030

Electrochemical impedance spectroscopy and linear polarization applied to evaluation of porosity of phosphate conversion coatings on electrogalvanized steels

  • Upload
    ufmg

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

www.elsevier.com/locate/apsusc

Applied Surface Science 253 (2006) 2875–2884

Electrochemical impedance spectroscopy and linear polarization

applied to evaluation of porosity of phosphate conversion

coatings on electrogalvanized steels

Vanessa de Freitas Cunha Lins a,*, Geraldo Francisco de Andrade Reis b,1,Carlos Roberto de Araujo a, Tulio Matencio b,1

a Corrosion and Surface Engineering Laboratory, Chemical Engineering Department, Federal University of Minas Gerais,

Rua Espirito Santo 35, 6th floor, Belo Horizonte 30160-030, Brazilb Chemical Department, Federal University of Minas Gerais, AV. Antonio Carlos, 6627-Pampulha, Belo Horizonte 31270-901, Brazil

Received 22 August 2005; received in revised form 3 June 2006; accepted 3 June 2006

Available online 11 July 2006

Abstract

In this work, electrochemical impedance spectroscopy and linear polarization are used in determining porosity of zinc phosphates, and of nickel

and manganese modified zinc phosphates on electrogalvanized steel. The porosity of the phosphate layers ranges from 0.1% for the manganese-

modified hopeite to 8% for hopeite, using the linear polarization and the electrochemical impedance spectroscopy techniques. The porosity values

measured using the two techniques were in agreement. All impedance spectra of coated samples showed two steps, with two capacitive arcs and an

inductive looping.

# 2006 Elsevier B.V. All rights reserved.

Keywords: Phosphate conversion coatings; Electrochemical impedance spectroscopy; Porosity; Linear polarization

1. Introduction

Phosphate conversion coatings are used by industries as a

surface pre-treatment of metals before painting, promoting the

paint adherence and protecting the substrate against corrosion.

Conversion treatments are intended to interpose an inorganic

layer providing both a strong bond to the metal surface and an

adhesive base for the polymeric film and to slow down by long-

term inhibition the destruction of the bonding by general under-

film corrosion due to oxygen diffusing through the coating and

by localized galvanic effects at any flaws in the organic film [1].

The extensive reviews on various aspects of phosphate coatings

are given by Lorin [2], Machu [3], Rausch [4] and Freeman [5].

Thephosphate layeroncarbonsteel shows twophases: hopeite

[Zn3(PO4)2�4H2O] and phosphophillite [Zn2Fe(PO4)2�4H2O]. As

* Corresponding author. Tel.: +55 31 32381773; fax: +55 31 32381789.

E-mail addresses: [email protected] (V. de Freitas Cunha Lins),

[email protected] (T. Matencio).1 Tel.: +55 31 34995758; fax: +55 31 34995700.

0169-4332/$ – see front matter # 2006 Elsevier B.V. All rights reserved.

doi:10.1016/j.apsusc.2006.06.030

the ratio phosphophillite/hopeite increases the adhesion and

corrosion resistance of the phosphate layer increases [6]. The

literature usually deals with phosphate layers on steel [7,8] so the

development of new methods and models for phosphate layers on

zinc is of great interest.

The phosphate coatings on galvanized steels present only

hopeite that has low resistance in alkaline and acid environ-

ments [9]. The addition of cations, such as manganese and

nickel, to the hopeite is an alternative to improve the resistance

of the phosphate coatings in alkaline and acid environments.

There is scarce literature data on physical and chemical

properties of these modified zinc phosphates [9,10].

One problem in the use of phosphate coatings in aggressive

media is the existence of open porosity and pinholes [2,11,12].

These local defects form direct paths between the corrosive

environment and the substrate [12]. As the corrosion reactions

are initiated at the coating–substrate interface, determination of

porosity is important to estimate the overall corrosion

resistance of the coated materials [12]. Notter and Gabe [11]

clearly describe the nature and effect of porosity on the

corrosion resistance of electrodeposited coatings. They point

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–28842876

out different methods developed for detecting porosity of

coatings into three main categories: physical methods using

microscopy, chemical methods using different test reagents and

electrochemical methods [11]. The traditional porosity

evaluation techniques are based on active porous observation

and they are generally qualitative [13]. These techniques are not

appropriate to nanometric scale defects evaluation [14]. Typical

porosity tests for phosphate coatings analysis are the

ferrocyanide test and the copper cementation in the pores

[15]; however, electrochemical tests insure better control of the

test conditions [16], such as the capacitance measurements

developed in order to evaluate the free metal fraction [17].

Electrochemical techniques can be used for porosity

measurements and performance evaluation of coating/

substrate systems [8,18,19]. The small potential difference

between the substrate and the coating allows that the substrate

current density be used in the porosity measurement [9].

Application of electrochemical impedance spectroscopy in

the study of substrate/coating system results in a new

information about the material degradation in corrosive media

[20,21].

The objective of this work is to evaluate the porosity, using

electrochemical impedance spectroscopy, Tafel analysis and

linear polarization of the zinc phosphate and the modified zinc

phosphate coatings, with manganese and nickel additions,

deposited on electrogalvanized steels.

2. Methodology

The electrogalvanized steel sheet was produced industrially.

The samples were taken after the hardening rolling, and the

substrate was mild steel (Table 1). The steel used for the

deposition process showed a homogeneous surface in relation

to crystallographic texture [22], carbon residue and roughness.

The carbon residue in the steel surface varied between the

values 2.8 and 4.7 mg m�2, the average value was 3.6 mg m�2.

The average roughness was 0.59 mm.

The main characteristic of this process is the manner in

which high deposition rates are obtained through application of

high current density and high flow electrolyte.

The industrial electroplating line is composed of an

entrance section where the bobbins are welded, and a

chemical process where electrolytic cleaning and pickling

with H2SO4 is used in order to remove oxide layers. The

deposition section contains 20 vertical cells. Each cell has

four boxes containing one insoluble anode. This system

making connections facilitates withdrawing the anode box as

well as making possible to deposit the zinc layer on one or on

both steel faces. The electrolyte that is zinc sulfate flows

Table 1

Chemical composition of the steel

Steel Chemical composition (%, w/w)

C S Mn P Si

Carbon steel 0.060 0.014 0.26 0.020 0.009

from the anode box top and drains away between the anode

and the steel sample. The distance between the steel sample

and the anode is 6.5 mm. The anode is titanium coated with

iridium oxide. The flow rate of the electrolyte in the lower

region reaches 4 m s�1. The electrolyte drains away through

holes in the anode box in order to distribute the pressure.

Gravitational force is used to accelerate the electrolyte in

cells. The application of high current densities

(13,000 A m�2 or higher) is possible because of the high

flow rate. The zinc concentration is maintained through two

dissolution stations, and the zinc electrolyte is fed into the

stations through a by-pass system. The concentration control

is done through the flux control in the dissolution stations.

The electrolyte pH is a control variable and is measured

continuously. The electrolyte temperature was maintained at

55 8C. After the deposition, the samples were immersed in a

sulfuric solution for dissolution of the zinc sulfate that was

deposited in a pure form. The samples were then cleaned and

dried.

The phosphating of the samples was performed at the

laboratory of Henkel Surface Technologies Brasil Ltda., by

immersion in phosphate baths with the compositions presented

in Table 2. The samples were degreased with a 3 wt.% solution

of a commercial product at 60 8C and rinsed with water at room

temperature. The samples were then immerged in a bath

containing 1 g L�1 of a titanium-based refiner for 30 s at the

temperature of 60 8C. The refiner is used to prepare the metallic

surface for the nucleation and growth of the phosphate crystals.

In the next step the samples were immerged in the phosphating

solution for 1 min at the temperature of 60 8C and then rinsed

with running water and dried at 100 8C for 2 min. This

procedure was used for each of the phosphating baths shown in

Table 2.

Accelerator such as nitrate and nitrite was not used.

The mass loss method was used for the determination of

the mass of the phosphate layer. The samples were weighed

at an analytical scale with a precision of 0.1 mg. The

phosphate layer was removed by immerging each sample in a

solution of CrO3 (0.5%, wt./vol. at 70 8C) for 5 min. The

samples were then rinsed with distilled water, dried and

weighed again.

The solution used for the removal of the phosphate coating

on each sample was calibrated and then analyzed by atomic

absorption spectroscopy for the determination of the elements

present in the coating. The equipment used was an Inductive

Coupled Plasma Spectrometer-Spectroflame Modula.

The morphology of the phosphate coating was characterized

by scanning electron microscopy using a Stereoscan 360/

Cambridge microscope.

Alsol Alinsol N O Ni, Cu and Cr

0.052 0.002 0.0027 0.0035 0.055

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–2884 2877

Table 3

Mass and chemical composition of the phosphate coating applied on the

electrogalvanized steel

Sample Phosphate coating

Coating weight (g m�2) Chemical composition (wt.%)

Zn P Mn Ni

E-F1 3.2 42.2 9.4 3.9 1.4

E-F2 2.1 54.6 12.7 <0.1 <0.1

E-F3 1.7 37.2 9.6 4.0 <0.1

E-F4 2.9 47.0 8.3 <0.1 1.3

E-F5 2.1 64.8 12.9 <0.1 2.7

Table 2

Chemical composition of phosphatization baths

Samples Phosphate layer Chemical composition (g L�1) Acidness (mL NaOH 0.1 mol L�1)

Zn2+ Mn2+ Ni2+ Free Total

E-F1 Nickel, manganese modified zinc phosphate 1.18 0.73 0.68 0.50 18.0

E-F2 Hopeite 1.23 – – 0.50 17.9

E-F3 Manganese modified zinc phosphate 1.15 0.80 – 0.60 17.9

E-F4 Nickel modified zinc phosphate (low content) 1.16 – 0.43 0.50 17.8

E-F5 Nickel modified zinc phosphate (high content) 1.16 – 0.86 0.60 17.9

The coating porosity can be evaluated using Eq. (1) [12,23].

P ¼ Rps

Rp

� 10�ðDEcorr=baÞ (1)

where P is the coating porosity, Rps the polarization resistance

of substrate, Rp the polarization resistance of coated substrate,

DEcorr the potential difference between the substrate and the

coated substrate and ba is the anodic Tafel coefficient of

substrate.

The electrochemical parameters were obtained by electro-

chemical impedance spectroscopy (EIS), using an Autolab 30

potentiostat, and FRA for Windows v. 2.3, Eco Chemie software,

ZSimpWin Version 2.0, EChem Software and EQUIVCRT

software. The counter electrode was a platinum wire. All the

potentials were measured in relation to a saturated calomel

electrode (SCE) and a three-electrode cell was used in the

experiments. The electrolyte was 0.001 mol L�1 NaCl solution.

The selection of the electrolyte was done in order to minimize the

phosphate dissolution in the alkaline and acid media [9], and in

neutral NaCl solution no passivation takes place on the substrate

[8]. For measuring the porosity of a coating, according to Notter

and Gabe [11] and Weng et al. [8] the linear polarization method

is valid only if the polarization current of the coating is negligible

compared to the substrate and in test solution no passivation

phenomena take place on the substrate. Weng et al. [8] measured

the porosity of zinc and manganese phosphates in 0.5 mol L�1

NaCl. The first electrolyte studied was 0.5 mol L�1 NaCl, but the

dissolution of phosphates occurred, and lower concentrations of

the NaCl solution were tested. The concentration of

0.001 mol L�1 NaCl was selected.

When the corrosion potential remained stable (a variation of

less than 5 mV in 5 min), a sinusoidal ac perturbation of 5 mV

amplitude coupled with the corrosion potential is then applied

to the electrode over the frequency range 5 mHz to 100 kHz.

Linear polarization was also performed to measure the

polarization resistance of the electrogalvanized steel and the

modified zinc phosphate-coated steels. Measurements of

polarization resistance are performed between �10 mV around

the free corrosion potential with a scan rate of 0.167 mV s�1,

after an immersion of 1 h in 0.001 mol L�1 NaCl solution.

Tafel analysis was performed according the ASTM G5

(1994) Standard, using the GPES—General Purpose Electro-

chemical System V. 4.5, Eco Chemie software. The polarization

potential range was �250 mV with respect to the open circuit

potential, and the potential rate was 0.2 mV s�1. The electrolyte

used in the Tafel analysis was 0.001 mol L�1 NaCl.

The working electrodes were electrogalvanized steels and

electrogalvanized steels coated with phosphate layers.

Five measurements of electrochemical parameters using

Tafel analysis, electrochemical impedance spectroscopy and

linear polarization were performed for each sample. The

average value of polarization resistance was registered.

3. Results

3.1. Characterization of zinc and phosphate coating

The zinc coat weight of the electrogalvanized steel was

52.5 g m�2. The result of the zinc coat chemical analysis did

not show the presence of foreign elements in this layer. It was

identified the crystalline phase eta (h—Zn pure) formed by

small randomly oriented hexagonal crystals.

Table 3 shows the mass and chemical composition of the

phosphate coatings from averaging two panels for each

phosphating condition.

The results in Table 3 show that the mass of the phosphate

coating varied according to the composition of the phosphating

bath. Sample E-F1, a tricationic phosphate layer on electro-

galvanized steel, showed the highest phosphate coating mass

while the lowest mass corresponded to the Zn/Mn phosphate

layer on zinc-coated steel. Manganese has a lower molecular

mass than the zinc and nickel, the other elements that constitute

the phosphate layer. This variation was attributed to the

chemical reactivity of each phosphating bath. The increase of

the layer weight with the nickel concentration that was reported

in the literature [24] did not be observed in this work. The

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–28842878

structure of inorganic phosphate coatings depends on their film

weight [8]. Normally, light-weight phosphate coatings (0.2–

1.4 g m�2) show an amorphous structure. Middle-weight (1.4–

7.5 g m�2) as well as heavy-weight (7.5–30 g m�2) phosphate

film demonstrates a crystal structure [8]. The phosphate

coatings studied can be classified as middle-weight according

to Table 3.

According to Fig. 1 the phosphate crystals were fine,

compact and had the shape of platelets, uniformly covering the

surface of the samples. The literature [9] reported the cubic

habit of the crystals in the Zn/Mn–phosphate layer and the plate

shaped habit in the Zn/Ni–phosphate layer. The habit on the

tricationic layer was a mixture of the different influences of Mn

and Ni on the morphology [9]. No significant differences in the

morphology were observed between the phosphate coatings for

the samples investigated in this work.

3.2. Electrochemical techniques

3.2.1. Chemical behavior of the phosphatized

electrogalvanized steel in NaCl 0.001 mol L�1

Corrosion of phosphate-coated electrogalvanized mild steel

is usually localized at the pores. Phosphate coating is an

insulating film with pores [8]. These pores induce the

electrochemical corrosion of the zinc substrate in aqueous

solution. The assumption that the mechanism of electroche-

mical/chemical reaction in pores of the phosphate coating and

bare metal is the same is often not verified [25]. The free surface

at the bottom of the pores was found to be modified in pre-

treatment and post-treatment operations and in some cases was

found to be more active than the clean bare metal surface. The

reference sample should be at least treated in diluted

phosphoric acid that has the same free and total acidity as

the phosphating bath [25]. In the present work, the saline

solution is able to penetrate to the substrate, which is zinc-

coated steel. Therefore, the metal at the bottom of the pores is

more active than the iron of the bare steel. According to the

potential–pH equilibrium diagram for the zinc–water system at

25 8C [26], in the electrolyte solution, pH near to 7, in the

potential range applied, the anodic reaction occurs:

Zn ! Zn2þ þ 2e

Fig. 1. Morphology of (a) the E-F1 phosphate c

The cathodic reactions in neutral medium was the hydrogen

evolution and alkalization [26]:

2Hþ þ 2e ! H2 (2)

2H2O þ O2þ e ! 4OH� (3)

Another possible reactions in the Zn–H2O system at

equilibrium were:

3Hþ þHZnO2� ! 2H2O þ Zn2þ (4)

Hþ þZnðOHÞþ ! H2O þ Zn2þ (5)

2Hþ þHZnO2� ! 2H2O þ ZnðOHÞþ (6)

4Hþ þZnO22� ! 2H2O þ Zn2þ (7)

Hþ þZnO22� ! HZnO2

� (8)

2Hþ þZnðOHÞ2 ! 2H2O þ Zn2þ (9)

Hþ þHZnO2� ! ZnðOHÞ2 (10)

2Hþ þZnO22� ! ZnðOHÞ2 (11)

The alkalization may contribute to the formation of Zn(OH)2

[24]. The alkaline dissolution of zinc is a possible reaction [24]:

Zn þ 2OH� ! ZnðOHÞ2þ 2e ðE� ¼ �1:25 VÞSHE (12)

Both anodic and cathodic reactions take place at the bottom

of the pores within the coating. Because of the existence of

diffusion resistance between the inside and the outside of the

pores, the concentration of corrosion product is obviously

higher in the occluded zone than in the body solution. The local

alkalization at the bottom of the pores may produce the alkaline

dissolution of the phosphate, especially of the hopeite.

The exposition of the manganese modified zinc phosphate

layer to the electrolyte produced a formation of Mn-hydroxide

that can be formed simply by ion exchange of the phosphate

anion in the crystal lattice with OH� in the electrolyte without

any oxidation reaction necessary [9].

Tomandl and co-workers [9] proposed that the formation of

an Mn-rich hydroxide layer serves as protection for the

remaining phosphate crystals.

The effect of nickel is probably due to its cementation as

metallic nickel on the zinc surface surrounding the crystals,

oating and (b) the E-F5 phosphate coating.

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–2884 2879

which remove the zinc surface from contact with the

electrolyte. The dissolution of zinc in the electrolyte occurs

with the hydrogen evolution as the cathodic reaction. In the

potential range studied, there was an increase of the

thermodynamic driving force for cathodic hydrogen evolution,

and there was a potential range of stability of nickel, according

to the potential–pH equilibrium diagram for the nickel–water

system at 25 8C considering the hydrated oxide forms [26]. In

the potential range applied to the sample of zinc/high nickel

phosphate on zinc-coated steel, this reaction could be possible

[24]:

Ni þ 2OH� ! NiðOHÞ2þ 2e ðE� ¼ �0:72 VÞSHE (13)

The metallic nickel present on the zinc surface may support

the cathodic reaction by acting as the cathodic area on the

surface [9]. Zimmermann et al. [24] reported that during the

deposition of the phosphate layer, the cementation of Ni on the

initially corroding Zn gives rise to the formation of surface Ni–

Zn alloys, providing cathodic centers where the reduction of H+

is favored. The electrochemical behavior of Zn/Ni phosphate

layer on zinc-coated steel was less noble than the behavior of

the hopeite layer on zinc-coated steel. These results may be

explained using the literature data [24].

Zimmermann et al. [24] reported that in the absence of

accelerator, the corrosion rate seems not be high enough to get a

sufficient surface Ni enrichment. Because of this, the potential

remains in a value near that of pure zinc.

Tomandl and co-workers [9] also proposed the production of

Zn(OH)2 during the exposition of Zn/Ni phosphate layer on

zinc-coated steel in dilute hydroxide solution.

3.2.2. Tafel analysis

Fig. 2 shows the polarization curve for the electrogalvanized

steel, obtained using Tafel analysis. The electrochemical

parameters obtained were the open circuit potential (Ecorr) of

�828 mVsce (standard deviation of 4 mV), and the anodic Tafel

coefficient ba = 57 mV decade�1, with a standard deviation of

4 mV decade�1. The Tafel coefficient, b, is related to the

kinetics of the rate-determining step of the reaction, and

typically has a value of the form RT/anF, where a is a constant

Fig. 2. Electrogalvanized steel polarization curve, obtained using Tafel analysis

in 0.001 mol L�1 NaCl solution.

related to the symmetry of the effects of a change in potential on

the forward and reverse reaction rates, and is typically in the

region of 0.5; R is the gas constant; T is the absolute

temperature; n is the number of electrons involved in the

reaction; F is Faraday’s constant. RT/F is approximately

59 mV at room temperature, so values of Tafel coefficients are

typically in the region of 60–120 mV per decade, although

higher or lower values are possible [26].

3.2.3. Electrochemical impedance spectroscopy analysis

Fig. 3a shows Nyquist plot of the electrogalvanized steel in

0.001 mol L�1 NaCl solution (room temperature), and Fig. 3b,

Nyquist plot of the manganese modified zinc phosphate coating

on electrogalvanized steel in 0.001 mol L�1 NaCl solution

(room temperature).

Nyquist plot of the manganese and nickel modified zinc

phosphate coating and Nyquist plot of the hopeite coating on

electrogalvanized steel in 0.001 mol L�1 NaCl solution (room

temperature) are shown in Fig. 4.

Fig. 5 presents Nyquist plot of the high nickel modified zinc

phosphate coating and Nyquist plot of the low nickel modified

zinc phosphate coating on electrogalvanized steel in

0.001 mol L�1 NaCl solution (room temperature).

Nyquist diagrams of the phosphate-coated electrogalvanized

steels show two capacitive arcs and one inductive loop in the

low frequency region.

Fig. 3. (a) Nyquist plot of the electrogalvanized steel in 0.001 mol L�1 NaCl

solution (room temperature) and (b) Nyquist plot of the manganese modified

zinc phosphate coating on electrogalvanized steel in 0.001 mol L�1 NaCl

solution (room temperature).

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–28842880

Fig. 4. Nyquist plot of the manganese and nickel modified zinc phosphate

coating (a) and Nyquist plot of the hopeite coating (b) on electrogalvanized steel

in 0.001 mol L�1 NaCl solution (room temperature).

Fig. 5. Nyquist plot of the high nickel modified zinc phosphate coating (a) and

Nyquist plot of the low nickel modified zinc phosphate coating (b) on electro-

galvanized steel in 0.001 mol L�1 NaCl solution (room temperature).

Fig. 6. Equivalent circuit proposed to simulated the impedance behavior of all

samples. Signification of parameter was done in the text.

ZSimpWin Version 2.0 and EQUIVCRT software allow

non-ideal electrochemical behavior to be modeled, using

constant phase element (CPE) to replace pure capacitance.

Fig. 6 shows the equivalent circuit proposed to simulate the

impedance behavior of all samples using Zview software. R1 is

the Ohmic or uncompensated resistance of the solution between

the working and reference electrodes. This electrolyte

resistance depends on the sample area and the solution

concentration. R2 is the electrolytic resistance through the pores

(pinholes), which is in parallel with C1, coating capacitance. R3

is a charge transfer resistance, in parallel with CPE1, which

represents the double layer capacitance at the coating/metal

interface. Constant phase element showed the impedance:

ZCPE ¼1

½TðjwÞ#�(14)

T is a constant in F cm�2 sw � 1, w is related to the angle of

rotation of a purely capacitive line on the complex plane plots:

A ¼ 90�ð1� ’Þ (15)

Only when w = 1, T = Cdl, and purely capacitive behavior is

obtained.

R4 is the resistance associated to the inductive processes, and

the L1 is the pseudo-inductance. The electrochemical behavior

of the phosphate layers on the zinc-coated steel obtained using

EIS are according to the literature data [8,27]. The EIS results in

this work indicate an electrochemical process with two steps,

and an inductive loop in the low frequency region. The first

capacitive arc is related to the surface coating with phosphate.

The second arc is related to the corrosive process by charge

transference, where the value of substrate resistance was

calculated. Weng et al. [8] obtained two semi-circles in the

Nyquist diagram of the Mn phosphate layer on zinc-coated

steel, and proposed a similar equivalent circuit with Warburg

impedance for the impedance spectra of phosphate coatings in

aqueous solution. Weng et al. [8] reported that the diffusion

resistance plays an important role in the protection of phosphate

coatings, and the influence of diffusion grows with the increase

of solution pH. In an acid medium, no diffusion can be found

during the corrosion of phosphate coatings on tool steel. In a

neutral medium of 0.5 mol L�1 NaCl, the impedance spectrum

reveals a characteristic of finite-length diffusion [8]. Nyquist

plots obtained in this work have the form presented in literature

for the model, which ignores the Warburg impedance [28]. If

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–2884 2881

Table 4

Electrochemical parameters obtained using the EIS technique

Sample Ecorr (mVSCE) R1 (V cm2) R2 (V cm2) C1 (mF cm�2) t(R2C1) (s) R3 (V cm2) t(R3CPE1) (s)

Substrate S828 4882 6308 36.8�824 2639 4161 0.23�830 4953 8142

Hopeite (Zn) S867 2303 6485 23.2 24800 1.04�860 1880 6236 0.15 58397

�865 2543 6523 29240

Zn–high Ni S975 1628 5613 45.6 8350 0.54�953 838 2002 0.25 4461

�925 1300 1500 10430

Zn–low Ni S914 906 4544 122.8 9000 0.73�936 2003 5645 0.56 7142

�922 1569 4789 12870

Zn–Mn S1012 1457 1952 75.7 3800 0.37�984 913 7122 0.19 12045

�1006 1489 2543 5019

Zn–Mn–Ni S890 1518 9590 16.2 35550 1.94�893 1567 7333 0.15 37650

�854 2367 9633 35000

Table 5

Porosity of the phosphate coating on electrogalvanized steels

Sample Porosity (%)

Electrochemical

impedance technique

Linear polarization

echnique

Hopeite (Zn) 6.9 8.2

Zn–high Ni 0.2 0.9

Zn–low Ni 2.4 1.8

Zn–Mn 0.1 0.1

Zn–Mn–Ni 1.9 1.1

the Warburg diffusion impedance is included, the curve shape

depends on the two competitive controlling mechanisms of

corrosion rate: charge transfer and diffusion.

Pye [28] reported that for real coated metal systems two

semicircles in the Nyquist plot could be distinguished only

when their time constant values are not too close. The criteria

followed to observe two semicircles are 0.2 < R3/R2 < 5,

which is observed in this work, and t(R3CPE1)/t(R2C1) > 20.

The ratio R3/R2 ranges from 1.48 to 3.82, but the ratio and

t(R3CPE1)/t(R2C1) ranges from 1.30 to 12.94. But in this work,

two time constants are identified using the Bode plot.

The inductive loop can be due to the coating porosity and

superficial adsorption [29,30]. The literature [24,31] cited the

adsorption of species such as OH� and NiOH on the surface of

phosphate layers on zinc-coated steels. Adsorption of products

of the reactions in the Zn–H2O system such as Zn(OH)+ and

HZnO2� is also possible. The presence of Hads is commonly

accepted in the postulated reaction mechanisms and was also

identified on nickel deposits by impedance measurements as a

low frequency capacitive loop [31]. Munoz and Schultze [31]

related the high absorbability of H on Ni and the Volmer-Tafel

mechanism where the desorption step is the controlling one:

Hþ þ e ! Hads (16)

2Hads ! H2 (17)

The less noble corrosion potentials observed for the Zn/Ni

phosphate on zinc-coated steel in relation to the corrosion

potentials of the Zn phosphate on electrogalvanized steel may

be due to the absence of accelerators in the phosphating bath.

Close to the corrosion potential, the preferential adsorption of

NO2� of the accelerator against NiOH decreases the rate of Ni

dissolution, shifting the onset of anodic currents to more

positive potentials. A similar effect is also given by the

adsorption of hydrogen [31]. Zimmerman et al. [24] also

reported that the preferential adsorption of NO2� against OH�

hinders both the active dissolution of Ni and the formation of a

passivating hydroxide/oxide film.

Table 4 shows electrochemical parameters values obtained

from fitting and corrosion potential, Ecorr. Five spectra were

obtained for each sample, and Table 4 shows three of them.

The parameters marked in black are of the spectrum that

represents an average of all results. The values appearing in

bold font in Table 4 are of the spectra, which represent an

average of the results.

Electrogalvanized mild steel is an electrochemically active

material. Anodic dissolution occurs if it is exposed to

0.001 mol L�1 NaCl solution, resulting in uniform macro-

scale removal of zinc from the exposed areas.

The electrolyte resistance, R1, varies according to the

exposed area of the samples and the solution concentration. The

value of the R1 was the highest for the zinc-coated steel, which

presented a higher exposed area of zinc than the other samples.

Among the samples with phosphate layer on zinc-coated steel,

the Zn phosphate layer on zinc-coated steel showed the highest

value of electrolyte resistance. This sample also showed the

highest porosity (Table 5) and the highest exposed zinc area

among the samples studied. According to Weng et al. [8], the

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–28842882

porosity is directly responsible for the exposed area of substrate

at the pores within the coating.

The corrosion potential of zinc-coated steel and hopeite

layer on zinc-coated steel (Table 4) was �0.828 V (SCE) and

�0.867 V (SCE). The anodic polarized Zn surface Zn/

Zn2+showed E8 = �0.81 V with [Zn2+] = 0.081 [24]. The

values obtained for the corrosion potentials of zinc-coated

steel and hopeite layer on zinc-coated steel were according to

Zimmermann et al. [24].

The proposed crystal structure of the manganese modified

zinc phosphate on zinc substrate is Mn2Zn(PO4)2�4H2O and

Zn3(PO4)2�4H2O [8]. Manganese occupies both octahedral and

tetrahedral positions in the hopeite crystal lattice [9]. The

exposition of the manganese modified zinc phosphate layer to the

electrolyte produced a formation of Mn-hydroxide that can be

formed simply by ion exchange of the phosphate anion in the

crystal lattice with OH� in the electrolyte without any oxidation

reaction necessary [9]. This ion exchange can have facilitated the

oxidation of zinc, generating the less noble corrosion potential

among the samples. The anodic reaction of zinc oxidation, and

the cathodic reaction of hydrogen evolution probably occurred

[9,26]. The alkalization may produce Zn(OH)2. Table 4 shows

that the values of the electrolytic resistance through the pores, R2,

and the charge transfer resistance, R3, were the lowest values

among the samples studied. The corrosion potential of the Zn/Mn

phosphate layer on zinc-coated steel was the less noble potential,

according to the resistance values.

The Zn/Ni–phosphate layer on zinc-coated steel showed

values of R2, electrolytic resistance through the pores, lower than

the pore resistance of the hopeite (Table 4). The open circuit

potential of the Zn/Ni phosphate layer on galvanized steel was

lower than the Ecorr of the Zn phosphate (Table 4). The less noble

behavior of the Zn–Ni phosphate layer on galvanized steel was

observed. The effect of nickel is probably due to its cementation

as metallic nickel on the zinc surface surrounding the crystals.

The dissolution of zinc in the electrolyte occurs with the

hydrogen evolution as the cathodic reaction. In the potential

range studied, there was an increase of the thermodynamic

driving force for cathodic hydrogen evolution, and there was a

potential range of stability of nickel, according to the potential–

pH equilibrium diagram for the nickel–water system at 25 8Cconsidering the hydrated oxide forms [26]. In the potential range

applied to the sample of zinc/high nickel phosphate on zinc-

coated steel, this reaction could be possible [24]:

Ni þ 2OH� ! NiðOHÞ2þ 2e ðE� ¼ �0:72 VÞ (18)

The metallic nickel present on the zinc surface may support

the cathodic reaction by acting as the cathodic area on the

surface [9]. Zimmermann et al. [24] reported that during the

deposition of the phosphate layer, the cementation of Ni on the

initially corroding Zn gives rise to the formation of surface Ni–

Zn alloys, providing cathodic centers where the reduction of H+

is favored. This hypothesis that nickel metallic on the Zn

surface and Ni–Zn alloys are cathodic centers could explain the

less noble electrochemical behavior of Zn/Ni phosphate layer

on zinc-coated steel in relation to the behavior of the hopeite

layer on zinc-coated steel.

The Zn/Mn/Ni phosphate layer on zinc-coated steel

presented the highest pore resistance (R2) than the samples

studied, and the corrosion potential of the trication phosphate

on zinc-coated steel was higher than the Ecorr of the Zn/Mn and

Zn/Ni phosphate layers on galvanized steel.

The porosity evaluation was done using the Eq. (1) and the

electrochemical parameters obtained using the electrochemical

impedance spectroscopy (Table 4) and the Tafel analysis

(Fig. 2). The Tafel coefficient was ba = 57 mV decade�1. The

polarization resistance values were obtained using the EIS

technique, and are the sum of the pore resistance (R2), the

electrolyte resistance and the charge transfer resistance. Liu

et al. [32] used the same equation for the estimation of porosity

of a PVD ceramic-coated mild steel. In Nyquist diagram, the

interception of the capacitive arc with the real axis in the high

frequency region provides the electrolyte resistance (R1), the

sum of electrolyte resistance and the pore resistance (R1 + R2),

and the interception of the capacitive arc with the real axis in

the low frequency region provides the sum of the electrolyte

resistance, pore resistance and charge transfer resistance

(R1 + R2 + R3) [28]. In the presence of an inductive loop in

the low frequency region, the charge transfer resistance was

calculated according literature [33,34]. The interception

relative to the sum of the electrolyte resistance, pore resistance

and charge transfer resistance is the interception of the loop

inductive with the real axis in the low frequency region.

Table 5 shows the values of porosity of the phosphate

coating on galvanized steels, using electrochemical impedance

spectroscopy. The porosity ranges from 0.1% for the

manganese-modified hopeite to 6.9% for hopeite. According

to the literature data, porosities from 0.2% to 2% are obtained

for zinc phosphates [2]. Weng et al. [8] reported that the mean

porosity of both Zn and Mn phosphates on tool steel was below

1.5%.

Hopeite presented the highest porosity (6.9%) among the

phosphate coatings studied. This result is in agreement with the

corrosion behavior of the hopeite in alkaline and acid

environments, in which it presents low resistance and

dissolution. Tomandl and co-workers [9] reported that the

manganese and nickel addition to the hopeite decreased the rate

of phosphate leaching in proportion to the percent incorporation

of these species. The high porosity of the zinc phosphate

observed in the NaCl solution may be due to the leaching

process. The dissolution of the phosphate layer was observed in

all samples after immersion during 12 h in 3% (w/w) NaCl

solution. The porosity of hopeite could be overestimated

because upon immersion the pore size and number may

increase and lead to overestimating all results, especially the

porosity of hopeite.

One coating used in order to improve the corrosion

resistance of the phosphate coating on galvanized steel is

the tricationic phosphate, which shows an intermediary value

of porosity of 1.9%. A low value of porosity is beneficial to the

corrosion resistance of the coating that acts as a physical

barrier against corrosive agents. In general, a lower porosity

means a smaller corrosion rate of substrate [8]. On the contrary,

considering the adhesion performance, the porosity of a

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–2884 2883

Table 6

Electrochemical parameters obtained using the linear polarization technique

Sample Ecorr (mVSCE) Rp (V cm2)

Electrogalvanized steel �820 12725

�824 13580

�828 10349

Hopeite (Zn) �867 25174

�860 34875

�868 24800

Zn–high Ni �975 9809

�953 4661

�925 9524

Zn–low Ni �914 8104

�936 20557

�916 9694

Zn–Mn �1012 14285

�984 9672

�1010 10567

Zn–Mn–Ni �980 35501

�893 34799

�854 37022

coating may have some advantages. If a phosphate coating is

used as a pre-treatment layer between the substrate and a top

layer, the pores in the coating may absorb oil, wax or paint very

well, so that the bonding ability between substrate and top

coating is improved. Therefore, the porosity of the tricationic

phosphate could be favorable to the both applications.

The coating pore resistance represents the resistance of areas

on the coating with more rapid solution uptake. The nickel

addition to the zinc phosphate promotes an effect of pore

resistance and corrosion potential reduction. The porosity

decreases with the nickel addition to the zinc phosphate

(Table 5). Mizuno et al. [35] reported that nickel is deposited on

the less noble substrate in the phosphate bath. The nuclei of the

phosphate crystals are formed on the sites, which are not

covered with nickel because of polarization. Subsequently the

crystals grow on the nuclei incorporating nickel. The porosity

values of the nickel/zinc phosphates on the electrogalvanized

steel may be underestimated because of the measurements

taken into account the pores with active zinc at the bottom. This

result is according to the decreasing of porosity with the nickel

addition in the phosphate bath.

The manganese introduction to the zinc phosphate reduces the

pore resistance and the corrosion potential. However, the

porosity decreases strongly. The porosity obtained for the

manganese-modified hopeite (0.1%) is according the literature

data [8]. Weng et al. [8] studied the porosity of zinc and

manganese phosphate coatings, using the polarization resistance

technique. They obtained that the porosity of Mn phosphate on

steel was smaller than the porosity of the Zn phosphate layer on

steel, indicating that manganese phosphate is more compact than

Zn phosphate. Weng et al. [8] obtained an average porosity of less

than 1.5% for zinc and manganese phosphate coatings. The

porosity is a function of a 10th power that varies with the

difference between the potentials of the substrate and of the steel

coated with the phosphate layer (Eq. (1)). So, the higher

polarization resistance did not signify a smaller porosity.

Pore resistance of the steel coated with nickel and

manganese-modified hopeite was higher than the steel coated

with hopeite, and the corrosion potential was similar. The

porosity of the steel coated with Ni, Mn and Zn phosphate was

lower than the galvanized steel coated with hopeite. This result

is in agreement with the literature [2] that reported that the Ni,

Mn, Zn phosphate coatings has a higher corrosion resistance

than the hopeite in aqueous media, alkaline or acid, and that the

grain size of the Ni, Mn, Zn phosphate coatings is smaller than

the grains of hopeite.

The porosity range obtained in this work is in accordance

with the results of Ponte et al. [36] that reported values of

porosity from 1.99% up to 5.38% for the tricationic phosphate

coatings of Fe/Zn/Mn. According Kwiatkowski [25], the value

of a phosphate coating porosity may vary significantly

depending on process details and the method of determination.

The polarization resistance was also measured using linear

polarization, and the porosity of phosphate coatings was

calculated using Eq. (1). Creus et al. [12] used linear

polarization to evaluate porosity of ceramic coatings (TiN,

CrN) on steel, using the Eq. (1). Weng et al. [8] used linear

polarization to measure the porosity of the zinc and manganese

phosphates on tool steel, in the 0.5 mol L�1 NaCl solution. Liu

et al. [32] also used Eq. (1) to measure the porosity of PVD

coated steel in 0.5 mol L�1 NaCl solution.

The porosity values obtained using this technique is also

shown in Table 5. The porosity values measured using the two

techniques were in agreement. Table 6 shows the values of

polarization resistance obtained using the linear polarization

technique.

The introduction of the phosphate layer on the electro-

galvanized steel reduced the corrosion potential of the

electrogalvanized steel, which is �828 mV (SCE). Corrosion

potentials in the range from �1012 mV up to �854 mV (SCE)

were measured for the phosphate-coated electrogalvanized

steels. Corrosion potential values from �968 mV (SCE) to

�835 mV of the phosphate-coated galvanized steels are

obtained in a previous work [37]. Probably the porous

phosphate layer induces a localized corrosion of the zinc,

and the zinc surface without coating presents a generalized

corrosion, that occurs in a less extent. Chemical analysis of the

electrolyte after electrochemical tests is in agreement with this

hypothesis. The electrolyte after the EIS tests of the

electrogalvanized samples presents a zinc concentration of

1.5 mg L�1. The results of the electrolyte analysis after EIS test

with the phosphate coating on electrogalvanized steels show

zinc contents from 2.51 mg L�1 (nickel manganese zinc

phosphate) to 9.29 mg L�1 (nickel zinc phosphate).

4. Conclusions

The porosity ranges from 0.1% for the manganese-modified

hopeite to 6.9% for hopeite, using the electrochemical

impedance spectroscopy technique.

The electrochemical reaction showed two steps, with two

capacitive arcs and an inductive looping.

V. de Freitas Cunha Lins et al. / Applied Surface Science 253 (2006) 2875–28842884

The electrogalvanized steel without phosphate coatings

showed a higher corrosion potential than the steel coated with

phosphates.

Hopeite presented the highest porosity among the phosphate

coatings studied. This result is in agreement with the corrosion

behavior of the hopeite in aqueous media, alkaline or acid, in

which it presents low resistance and dissolution.

One coating used in order to improve the corrosion

resistance of the phosphate coating on galvanized steel is the

tricationic phosphate, which shows an intermediary value of

porosity of 1.9%.

As the nickel content decreases from 0.83% (w/w) to 0.43%

(w/w) in the Ni, Zn phosphate, the corrosion potential increase.

However, the porosity also increases. The nickel addition to the

hopeite promotes a decrease of the polarization resistance and

corrosion potential, but again the porosity decreases.

The manganese addition to the hopeite decreases the

corrosion potential and the polarization resistance, decreasing

the corrosion resistance of the material in this medium.

However, the manganese addition to the hopeite decreases

significantly the porosity of the coating.

The addition of manganese and nickel to the hopeite was

beneficial, and increases the polarization resistance and

decreases the phosphate porosity.

Using linear polarization technique, the porosity ranges

from 0.1% for the manganese-modified hopeite to 8.2% for the

hopeite. The porosity values measured using the two

techniques, EIS and linear polarization, were in agreement.

References

[1] G. Bombara, U. Bernabai, Surf. Technol. 11 (1980) 393.

[2] G. Lorin, Phosphating of Metals-Constitution, Finishing Publications

Ltd., Middlesex, 1974.

[3] W. Machu, Die Phosphatierung, Verlag Chemie, 1950.

[4] W. Rausch, The Phosphating of Metals, Finishing Publications Ltd.,

Middlesex, 1990.

[5] D.B. Freeman, Phosphating and Metal Pre-Treatment, Wooghead-Faul-

kner, Cambridge, 1986.

[6] L. Fedrizzi, F. Deflorian, S. Rossi, L. Fambri, P.L. Bonora, Prog. Org.

Coat. 42 (2001) 65.

[7] N. Bretherton, G.E. Thompson, S. Turgoose, Corros. Sci. 35 (1993) 1297–

1303.

[8] D. Weng, P. Jokiel, A. Uebleis, H. Boehni, Surf. Coat. Technol. 88 (1996)

147–156.

[9] K. Ogle, A. Tomandl, N. Meddahi, M. Wolpers, Corros. Sci. 46 (2004)

979.

[10] N. Sato, T. Minami, H. Kono, Surf. Coat. Technol. 37 (1989) 23.

[11] I.M. Notter, D.R. Gabe, Corros. Rev. 5 (1992) 217.

[12] J. Creus, H. Mazille, H. Idrissi, Surf. Coat. Technol. 130 (2000) 224.

[13] E. Julve, Pintura y Acabados (1986) 80.

[14] J.R. Roos, J.P. Celis, F. Chonglun, J. Electrochem. Soc. 137 (04) (1990)

1096.

[15] A. Losch, E. Klusmann, J.W. Schultze, Electrochim. Acta 39 (1994)

1183.

[16] A. Losch, J.W. Schultze, H.D. Speckman, A new electrochemical method

for the determination of the free surface of phosphate layers, in: EURO-

CORR ’91 Proceedings, Budapeste, (1991), pp. 413–418.

[17] A. Losch, J.W. Schultze, J. Electroanal. Chem. 359 (1993) 39.

[18] H. Leidheiser Jr., Corrosion 39 (1983) 189.

[19] B.S. Skerry, D.A. Eden, Prog. Org. Coat. 15 (1987) 269.

[20] F. Deflorian, L. Fedrizzi, A. Locaspi, P.L. Bonora, Electrochim. Acta 38

(1993) 1945.

[21] F. Mansfeld, S.L. Jeanjaquet, M.W. Kendig, Corros. Sci. 26 (1986) 734.

[22] J.F. Silva Filho, V.F.C. Lins, Surf. Coat. Technol. 200 (9) (2006) 2892.

[23] J. Aromaa, Mater. Sci. Eng. A 140 (1991) 722.

[24] D. Zimmermann, A.G. Munoz, J.W. Schultze, Electrochim. Acta 48

(2003) 3267–3277.

[25] L. Kwiatkowski, Surf. Eng. 20 (4) (2004) 292.

[26] P.R. Roberge, Hanbook of Corrosion Engineering, The Mc-Graw-Hill

Companies, New York, 2000, p. 1139.

[27] S. Wolynec, Tecnicas Eletroquımicas em Corrosao, EDUSP—Editora da

Universidade de Sao Paulo, Sao Paulo, 2003, p. 166.

[28] W.A. Pye, Field performance of epoxy-coated reinforcing steel in Virginia

Bridge Decks, Blacksburg—Virginia Polytechnic Institute and State

University, Virginia, 1998, 199 pp., Dissertation (Doctor of Philosophy

in Civil Engineering).

[29] D.C. Silverman, Primer on the AC impedance technique, in: R. Baboian

(Ed.), Electrochemical Techniques for Corrosion Engineering, Houston–

NACE—National Association of Corrosion Engineers, Texas, 1986, p. 73.

[30] A.M.O. Brett, C.M.A. Brett, Electroquımica: Princıpios, metodos e

aplicacoes, Coimbra, Livraria Almedina, Coimbra, 1996, p. 471.

[31] A.G. Munoz, J.W. Schultze, Electrochim. Acta 49 (2004) 293–301.

[32] C. Liu, Q. Bi, A. Leyland, A. Mathews, Corros. Sci. 45 (2003) 1257.

[33] R. Baboian, Electrochemical Techniques for Corrosion Engineering,

National Association of Corrosion Engineers, Texas, 1986, p. 73.

[34] J.R. McDonald, Impedance Spectroscopy: Emphasizing Solid Materials

and Systems, John Wiley and Sons, New York, 1987, p. 346.

[35] K. Mizuno, K. Suzuki, T. Ohtsubo, M. Suzuki, Characterization of

phosphate conversion coatings, in: Proceedings of the International Con-

ference on Zinc and Zinc Alloy Coated Steel Sheet, GALVATECH, Tokyo,

The Iron and Steel Institute, Japan, 1989.

[36] H.A. Ponte, A.M. Maul, E.A. Alvarenga, Mater. Res. 5 (2002) 439.

[37] K.S. Fernandes, V.F.C. Lins, E.A. Alvarenga, M.M.A.M. Schvartzman, in:

Proceedings of the XIV Congresso Brasileiro de Engenharia Quımica,

ABEQ, Natal, 2002.