Você está na página 1de 25

Corrosion Science,

Vol. 39,
0

No. 4, pp. 655-679, 1997 1997 ElsevierScienceLtd

Printed in Great Britain. All rights reserved 001&938X/97 $17.00 + 0.00

PII: s0010_938x(%)0015&3

A SURVEY OF ARGENTINEAN II-COPPER


J. R. VILCHE,*

ATMOSPHERIC SAMPLES

CORROSION:

F. E. VARELA,* E. N. CODARO,* B. M. ROSALES,+ G. MORIENA+ and A. FERNANDEZ+

* Instituto de Investigaciones Fisicoquimicas Teoricas y Aplicadas (INIFTA), Facultad de Ciencias Exactas, Universidad National de La Plata, Sucursal4, C.C. 16, (1900), La Plata, Argentina +CITEFA-CONICET, Centro de Investigaciones en Corrosion (CEICOR), Zufriategui 4380, (1603). Villa Martelli, Argentina Abstract-Copper samples were exposed at six sites with known ambient parameters in Argentina and the atmospheric corrosion was investigated after different outdoor exposition periods. Weight-loss measurements used to determine corrosion damage were complemented with both DC and AC electrochemical techniques, performed in 0.1 M Na#Od solution employing the exposed face of the test samples, in order to characterize the protectivenesses of the surface layers generated on copper in the distinct environments. While the ambient aggressiveness could be well evaluated from meteorological and pollution parameters and from weight-loss data, the product protective characteristics estimated through SEM and EDAX observations and electrochemical methods yielded valuable information to understand mechanistic aspects concerning the effects of physical properties, structure and contaminant content of surface corrosion products. 0 1997 Elsevier Science Ltd. All rights reserved Keywords: A. copper, B. weight-loss, B. EIS, B. SEM, B. potentiostatic, C. atmospheric corrosion.

INTRODUCTION Since atmospheric corrosion is a slow process, natural tests for the direct measurement of corrosion require very long periods of time. For this reason, new methods must be developed to replace the classical gravimetric procedures in the fight against atmospheric corrosion. Since atmospheric corrosion has been shown to be an electrochemical type phenomenon, it is possible to use electrochemical techniques for studying it as an alternative approach to traditional tests. Copper is an important functional material used in industrial, commercial and home environments. Found in its native state approximately 6000 years ago, it is considered a noble metal. The corrosion of copper, a common problem, can be aesthetically beautiful and frustratingly complex. The uniform corrosion, passivation and pitting corrosion processes of copper have been studied in different solutions since more than three decades ago. The literature on the subject until 1980 is reviewed in references.2 The corrosion of copper and its alloys depends to a great extent on the make-up of the electrolyte in contact with the metal. The reaction mechanism involves copper dissolution at local anode sites and electrochemical reduction of some species such as oxygen at cathodic areas. A given surface area may alternate from being anode and cathode to give uniform corrosion. Pitting in
Manuscript received 28 August 1996. 655

656

J. R. Vilche et al.

copper is the result of some spots remaining anodic for relatively long periods. Natural fresh water generally promotes the formation of protective coatings. From detailed examination by electrochemical and surface analytical methods, this passive layer appears to consist of a simple 0.1~0 or a duplex CuzO/CuO, Cu(OH)z film,3 the entire layer structure undergoing short and long time range phase transitions4 These layers are important in different regards. They are useful for corrosion protection and they influence electrochemical processes at copper electrodes. Actual rates of corrosion are usually very low; consequently copper is widely used in water lines, water tanks and heat exchangers. The protective properties of the anodic layer produced on copper in alkaline solutions vary with the nature of anions in the solution. Thus, in aqueous solutions containing either a carbonate or a phosphate salt, the protective properties of the anodic layer could be improved. Occasionally, insoluble copper-containing salts as malachite (CuCO,.Cu(OH),) or azurite (XuC03.Cu(OH)z) become a part of the anodic layer. Whether the anodic layer produced on copper behaves as a protective or a non-protective layer depends not only on its chemical composition but also on its compactness and adhesion to the copper substrate. The existence of soluble species, aqueous copper carbonate complexes, has been postulated,5 as corroborated by the high solubility of the following complex anions: CUCO~(,~, CU(CO&~- and CUCO~(OH)?-.~ In contact with very soft water containing large amounts of CO2 and 02, the protective films do not form due to carbonic acid effects. The rate of copper corrosion may become excessive. Atmospheric corrosion of copper has been studied extensively. Reviews by Leidheiser and Rozenfeld lo provide the reader with an historical perspective. The majority of these studies focused attention on outdoor urban corrosion. The atmospheric corrosion process is more complex than high temperature oxidation, where theories exist that explain reasonably well the observed rates. No such theory exists for atmospheric corrosion of copper, although attempts have been made to correlate observed outdoor rates to meteorological and pollution parameters.2-4 The corrosion rate of copper in laboratory tests has shown to be sensitive function of relative humidity, sulphur dioxide, nitrogen dioxide, hydrogen sulphide, ozone, hydrogen chloride and chlorine concentrations.4-8 It was observed that indoor corrosion rates obey log normal statistics over the field population of each study, although copper corrodes significantly faster outdoors (indoor rates are about 1% of outdoors values).15 Combined quartz crystal microbalance measurements with ion chromatography used to analyse the evolution of the absorbed electrolyte layer during the exposure of copper revealed that the adsorption of water on to the metal surfaces from the atmosphere reaches a steady state within 30 min at a constant relative humidity, and that the water adsorption is not the rate-limiting step in the establishment of the absorbed electrolyte and the initiation of corrosion.6 The analysis of comparative effects of SOZ, NO2 and O3 studied on laboratory and field-exposed copper7,8 suggested that synergistic effects detected for sulphur dioxide and ozone can explain the unexpectedly high corrosion rates of copper found at rural sites, which are characterized by low sulphur dioxide and nitrogen dioxide, but high ozone concentrations. Although sulphate and nitrate ions were the dominating surface species, additional constituents detected in the laboratory, but not in the field, were sulphite and nitrite. In addition, Cu20 was identified as an important compound at some sites. Both chloride and ammonia were detected as surface constituents on all field samples. 7x8 The use of multilamellar electrochemical cells, known as electrochemical atmospheric corrosion monitors, to evaluate the atmospheric corrosion rate of copper has been also reported. I9

A survey of Argentinean atmospheric corrosion: II

657

Kinetics approaches rather than thermodynamics are suitable for studying processes involved during atmospheric corrosion, in which chemical changes occur on time scales much too short for multiphase equilibria to establish themselves. Among other in situ techniques, infrared reflection absorption spectroscopy (IRAS) and combined X-ray diffraction, infrared spectroscopy, scanning Kelvin probe measurements, and pH measurements2 have been applied to the monitoring of corrosion process in real time. Thus, copper samples were exposed in flowing corrosive air and the initial formation of a film of Cu20 was followed by in situ IRAS; the kinetics of the oxidation was found to obey a logarithmic rate law.20 In experiences of copper corrosion in sulphur dioxide-containing air, the formation of sulphite species on the surface was detected by IRAS. In urban environments, where ammonium and sulphate are the most abundant ions in fine dust particles, the corrosion mechanism was explained as dissolution of Cu followed by formation of Cu20, oxidation of Cu(1) ions to Cu(I1) ions, and precipitation of antlerite [Cus(SO,)(OH)& brochantite [CU~(SO~)(OH)~]or posnjakite [CU~(SO~)(OH)~.H~O].~This process is markedly temperature dependent. At 300 K the corrosion products formed are basic copper sulphates, while at 373 K a higher fraction of Cu20 is formed.21 Complementing previous work on aluminium and zinc samples,22 data obtained from electrochemical and microscopy techniques are discussed in this work in order to investigate the characteristics and properties of corrosion products formed on Cu samples after exposure in six test stations covering different environmental conditions in Argentina, within the frame of the MICAT (Iberoamerican Map of Atmospheric Corrosion) Project.3,23 EXPERIMENTAL METHOD

Samples of Cu (99.97%) were exposed at the following atmospheres: dry (San Juan) and subtropical (Iguazu) rural; temperate (Camet) and polar (Jubany) marine; and urbanindustrial (La Plata and Villa Martelli). Environmental data monitored in these outdoor stations are presented in Table 1, according to ambient characterization given in ISO/DP 9225.22 The corrosion products formed after 1, 2, 3 and 4 y exposure times were characterized using DC and AC electrochemical techniques, SEM observations, surface chemical analysis (EDX) and weight-loss measurements. For the weight-loss measurements, surface corrosion products were removed by pickling of corroded samples as indicated in ISO/DP 9226. The electrochemical experiments were performed at 20C in conventional threecompartment double wall glass cells containing 0.1 M Na2S04 solution, which was prepared from analytical grade (p.a. Merck) reagents and triply distilled water purified in a Mill&Q reagent grade system. Potentials were measured (and referred to in the text), against a SCE reference electrode, properly shielded. A large-area Pt plate was used as the counter electrode. The experimental setup employed for DC measurements included various Tacussel potentiostats, Servovit 13 potential scanners and EPL 2 recorders. Electrochemical impedance spectroscopy (EIS) measurements were performed employing working electrodes of geometrical area 1.33 cm2 and an activated platinum probe coupled to the reference electrode through a 10 uF capacitor to reduce phase shift errors at high frequencies. Detailed descriptions of both hardware arrangement and data processing have been given elsewhere.22*24,25EIS measurements were carried out at the

658 Table I. Environmental

J. R. Vilche ef al. data of the six test stations in Argentina Contamination Total rain (mm) 817
805

Meteorological Station (ambient condition) Camet (marine) Mean temp. (C) 14.1 13.9 14.5 13.5 16.7 Mean rel.hum. (%) 79.2 78.8 go.3 78.7 15.3 Total TOW (h) 5974 6202 6448 5834 5063

Period (y)

SOz(mg mm2 day-)

Cl-(mg

m - day-) 30 40 70 48

2 3 4 I

Villa Martelli (urban-ind.)

1226 1153 1729

2 3 4 I 2 3 4 I 2 3 4

Iguazu (rural)

17.1 17.0 15.8 20.6 20.9 22.1 2!.1 18.0 20.0 18.3 IX.3 2.0 m-3.I -2.9 PO.6 17.0 16.7 13.6 17.1

71.5 73.6 70.1 77.8 73.8 14.7 78.6 50.6 49.3 50.8 51.8 83.8 84.0 84.5 84.3 77.5 76.8 78.0 17.9

4227 4509 3478 5831 5530 5547 5167 1002 847 865 972 2693 2425 2588 3299 5198 4955 553 I 4869

983 1420 1422 2158 2624 1720 2020 35


Ill

IO 9 IO

San Juan (rural)

I
2 3 4

Jubany (marine)

93 67 *

6 6
30

* *
1830 1178 126.1 361 1088 6.2 8.2 6. I 6.5

24

I 2 3
4 * Not available

La Plata (urbanind.)

corrosion potential in the frequency range 1 mHz <.f< 50 kHz, using a Solartron 1250 FRA and a 1186 EI integrated with a PC system. SEM observations were made using a Philips 515 microscope, which was coupled to a 9100 EDAX for surface analysis. Test samples were metallized with Ni previous to SEM observation. SEM and EDAX data were used to characterize the morphology and chemical heterogeneities of corrosion products. Analysis of the saline pollutants was made by immersion of the samples in boiling distilled water for 1 h followed by measurement of the amounts of Cl- and SOd2- ions in the remaining solution at 25C. The MICAT Project was established to examine the specific influence of SO2 and Cl- ions. The results expressed as mg of anion per area unit were

A survey of Argentinean atmospheric corrosion: II

659

compared with the data obtained in the corresponding station.422 EXPERIMENTAL


Weight-loss studies

candle (IS0 9224) and outdoor

RESULTS

AND DISCUSSION

Annual corrosion rates of Cu were determined over both 100 x 150 mm areas exposed to the ambient facing to the sky and to the ground, on the basis of mass losses, density and exposure time. The edges of the samples were considered negligible compared to those areas. Figure 1 shows the corrosion rates, expressed in pm y - , for different exposure times. Data obtained in marine, urban and rural environments follow clearly different trends. The ambient variables with most influence on the experimental results are time of wetness (TOW), temperature (2) and pollutant concentration.14 The TOW is determining of the difference in annual corrosion rates observed between the two rural stations San Juan (dry) and Iguazu (subtropical), and also between the urban ones Villa Martelli and La Plata. On the other hand, T and pollutant concentration account for the distinct magnitudes observed between the marine stations Camet and Jubany. The similarity of values obtained for the first annual corrosion rates of the rural ambient of Iguazu and the urban site of Villa Martelli can be explained as due to the compensation of effects between pollutant concentrations and TOW, T and RH (Relative Humidity) values. After the first year the effect of pollutants is almost negligible.
Potentiostatic step polarization data

The electrochemical character of atmospheric corrosion suggested the possibility of using DC anodic and cathodic steps of 20 mV amplitude to evaluate the protectiveness of the corrosion products formed after different exposure times in the six distinct environments. Pieces of 1 cm2 cut from the outdoor test samples were immersed in the support electrolyte 0.1 M Na2S04 for 1, 24 and 48 h before each potentiostatic pulse and

Co) Camet
[I{ (-1 SanJuan (+I Jubany ( l 1 La Plata

~~Gllartel: Ii

-I

2 3 t I years

Fig. 1. Corrosion rates of copper from weight-loss measurements.

660

J. R. Vilche et al

then the attained stationary current densities were measured. These data can be associated, in principle, with corrosion rates detected in natural exposures during rain periods whose duration should be similar to the above mentioned immersion times. The effect of both weathering and immersion time, the latter being proportional to the TOW of the test site, can be analysed separately. It was expected that an aggressive environment would produce results in agreement with high corrosion rates while mild ambient conditions would be reflected in low current densities, corresponding to the low mass losses measured. Neither high current densities were obtained for corrosion films grown in aggressive ambient nor low figures were determined for corrosion products formed in mild environments (Table 2). On the contrary, an opposite trend for the corrosion rates in the distinct ambient was observed after pulses, and that can be probably attributed to the increase in the pollutant content as a function of time (Table 3). Moreover, the main consistent result was the loss in protectiveness, corresponding to an increase of the current density, due to the partial dissolution of the products in the support electrolyte (Table 2) also observed in EIS data for different immersion times.

Table

2. Current densities attained in anodic (A) and cathodic (C) potentiostatic time. Copper samples weathered from ([4 - &,, = 20 mV) after 1, 24 and 48 h immersion Argentinean stations Immersion Station San Juan Period (y) j PA cme2 1 2 3 4 A time (h) 48 jpAcme2 C 1 0.02 0.06 0.02 0.11 0.02 0.08 0.01 0.01 0.03 0.03 0.03 0.12 0.11 0.04 0.06 Immersion 24 0.05 0.05 0.03 0.25 0.04 0.08 0.04 0.01 0.07 0.02 0.07 0.13 0.13 0.06 0.07

polarizations

1 to 4 y in the

time (h) 48 0.22 0.14 0.05 0.46 0.05 0.05 0.04 0.02 0.13 0.14 0.09 0.09 0.19 0.09 0.11
__

I
0.03 0.05 0.03 0.99 0.19 0.14 0.03 0.03 0.02 0.05 0.04 0.43 0.14 0.06 0.27

24 0.05 0.06 0.04 1.79 0.20 0.04 0.05 0.03 0.20 0.03 0.09 0.26 0.17 0.06 0.13

I .66 0.09
o.ib

lguazu

1 2 3
4

.A

V. Martelli

1 2 3
4

La Plata

I 2
3
4

6.33 0.15 0.07 0.05 0.05 0.14 0.14 0.17 0.21 0.24 0.10 0.32

Camet

1
2 3 4 1 2 3 4

A 0.02 0.1 I 0.01 0.05 0.10 0.17 0.10 0.05 0.17 0.07 0.03 0.14 0.11 0.06 0.16 0.02 0.08 0.09 0.23 0.12

Jubany

0.01 0.13 0.01 0.07 0.16 0.20 0.06

0.05 0.18 0.07 0.10 0.19 0.11

0.55 0.21 0.05 0.10 0.11 0.20 0.11

A survey of Argentinean atmospheric corrosion: II Table 3.

661

Pollutant content in the corrosion products of copper weathered from 1 to 3y exposure in the Argentinean test stations Concentration

Station (ambient condition) Iguazu (rural) V. Martelh (urban-indust.) Camet (marine) Jubany (marine) 1Y 29.8 86.0 491.6 1031.2

Cl-(mg me2) 2Y 14.4 53.3 474.9 3Y 52.8 40.8 464.0 1221.3 JY 83.9 107.5 6.8

SOd2-(mg me2) 2Y 42.1 130.7 75.1 70.9 3Y 101.3 83.2 102.7 111.7

SEM and EDAX data

SEM analysis on atmospheric exposed samples showed that the attack on Cu is lower than that previously found on Zn,22 whereas the thickness of the passive layers seems to be more uniform. The plan and cross-section observations of corrosion products revealed good agreement with the corrosion rates estimated from weight-loss measurements. Using EDAX, the presence of Cl and S proceeding from both the marine and industrial atmospheric pollutants as well as the soil were determined. Distribution mapping can be obtained when their weight-fractions are higher than 2%, and then, interesting correlation with results of the other techniques can be established. EDAX analysis on corrosion products with several morphology and size showed that the flat areas, in which uniform Cu corrosion was observed, contain lower quantity of environmental pollutants (Figs 2-7). The large structures present higher pollutant contents and are associated with localized attacks, with pitting under the nuclei. The larger the nuclei size is, the wider and deeper the pit becomes, as it can be seen in the cross-section corresponding to the Figs 2-7. The reason for this is that both Cl- and SO2 stimulate the localized attack and then constitute partly the insoluble corrosion products on the metal. The depth of the pits, which can be seen in the cross-section SEM micrographs, increases with the time of exposure in the test sites with high contents in Cl- and SO2 pollutants (Figs 6 and 7(b)). Comparing morphology and EDAX surface analysis of the Cu samples exposed from 1 to 4 y, an increase in the density of globular corrosion products is observed, but maintaining the size of each unit which is determined by the pollutant level. Comparable sizes of the structures shown in Figs 2-7 resulted from similar total amounts of the different pollutant in the stations, according to the data shown in the corresponding EDAX. Results for 2,3 and 4 y showed an increase in the density of particles with the exposure time. In the cross-section micrographs, few pits of small depth and poor cohesion were observed, mainly in Iguazu (Fig. 3b). A correlation between the size of the grown product nuclei and those of the metallic grains beneath them was also found. As in the case of the Zn samples,22 the pollutants are included in the layers of corrosion products on Cu, forming chlorides and basic sulphates, but since the corrosion products on Cu are insoluble, the mapping allows determination of their distribution along the thickness of the corrosion product layers. Figure 7 shows a change in the shape of the surface layer

WT % 2.54 5.68 S CL CA cu -----looon 0.38 1.60 0.68 89.11

Fig. 2.

MEB-EDX

of Cu samples after

, 1y (a,b) and 4 y (c,d) in San Juan station. (a,c) In plan, (b,d) cross section.

formed on Cu, turning to be rather flat than globular. This non-typical morphology, which is clearly observed in Fig. 7a, is similar to that found on Zn in the same Antarctic station,22 and can be related to its development under an ice layer, which flattens the structures making them lose the globular shape characteristic of the corrosion products formed on Cu in all the explored continental environments.4 The corresponding cross-section micrographs reveal the localized character of the attack, which originates the globular products adjacent to areas with uniform attack. These two regions correspond to the brighter and darker areas, respectively, in the plan micrographs of Figs 2-6. Both the mapping of S and Cl and the EDAX diagrams exhibited their heterogeneous distributions,

A survey of Argentinean atmospheric corrosion: II

663

(4

c
wr %
0.52 ____--.--- 99.49 Kmo

f < rc

Fig. 2.

(Continued)

being associated with larger and higher roughness of the globular products. These results were found in all of the environments with high pollutant level in Latin America, though lower levels in Argentinas stations mask the results.14 Exfoliation was still found in rural environments such as those of San Juan (Fig. 2 corresponding to 3 and 4 y) and Iguazu (Fig. 3b, 1 y). It can also be observed in Villa Martelli (Fig. 4b, 3 y), Camet (Fig. 6b, 4 y) and Jubany (Fig. 7b, 1 y). Analysis of pollutants Table 3 shows the results of the pollutants extracted from the corrosion products layers grown in the different environments of Argentinas six stations within the Corrosion

664

J. R. Vilche ef al.

Fig. 3

MEB-EDX

of Cu samples after 1 y (a,b) and 3 y (c,d) in Iguazu station. cross section.

(a,c) In plan, (b,d)

Mapping Project4322 with the object of comparing the influence of the corrosion product composition and the pollutant solubility during the wetting-periods. Thus, pollutant contents were obtained for the whole thickness of the surface layer, and since the included copper micrographs point out their detailed distributions, the calculation of the percentage accompanied by the corresponding EDAX diagrams allows comparison with the total values obtained through this technique. The best correlation was obtained when comparing the results of both techniques on samples of the same series, because the corresponding analysis were made on different pieces

A survey of Argentinean atmospheric corrosion: II

665

c
f
I

(4

WT % 0.21 _________99.19 IW.00

Fig. 3.

(Continued)

of the same sample. As the soluble compounds can be lixiviated during rain just before the collection of the respective series, the comparison with the data of other series can show, in general, quite different values. However, the results for pollutant content in corrosion layers grown on Cu samples are in agreement for both techniques, because the pollutants are incorporated as basic compounds consisting of insoluble Cl and S containing species, which are not reached by the rain. Since the Cl and S containing compounds present in the corrosion products formed on Cu samples are insoluble at ambient temperature, they can not be cleaned by the rain action during the outdoor exposition and can be easily detectable with the EDAX technique. Though, they are lixiviated by hydrolysis during the boiling in distilled water. In this way, the results of both techniques are in agreement only for the case of Cu samples.**

ELkM AL SI S FE CU

WC x 081 0 73 034 1.06 _____~~o_q 10000

r)

-.^-._.____._-1._.___

---._

171g. 4.

MEB-EDX

of Cu samples after 2 y (a,b) and 3 y (c,d) years in Villa Martelli station. plan. (b-d) cross section.

(a,c) In

Electrochemical

impedance spectroscopy

i EIS)

data

The impedance measurements of the copper samples were performed to determine the transfer function of the metal/corrosion products/electrolyte system. This was made in order to establish a relationship between the environmental data monitored in the outdoor test stations and the characteristics of the passive layers formed on the exposed samples. Typical impedance responses at the corrosion potential, which lies at ca. - 0.40 f 0.05 V, are shown as Nyquist (Fig. 8) and Bode plots (Fig. 9) for copper samples exposed in the test sites during different times. The shape of the corresponding Nyquist diagrams exhibits two slightly distorted capacitive semicircles and a third contribution at very low frequencies, which is related to slow processes on the corrosion layer such as diffusion. It is important to

A survey of Argentinean atmospheric corrosion: II

667

wr %
0.36 0.48 99.16 -1oooo

Fig. 4.

(Continued)

mention that the partial dissolution of the corrosion products in the support electrolyte yields a decrease in the impedance values, as it can be seen in Fig. 8 for different immersion times. These results indicate clearly that EIS is a very sensitive technique for characterizing surface layers. Since the passive films formed on copper are usually described as bi-layered structures, the two time constants at higher frequencies can be associated with a compact Cut0 inner layer and a porous outer layer, composed by cupric sulphate, nitrate or chloride depending on the atmospheric pollutants. 17** The experimental data can be discussed taking into account the impedance of the twolayer structure of the surface corrosion products in parallel with the cathodic reaction impedance, Rcath and an additional time constant for the mass-transport contribution (Z,),

J. R. Vilche et al.

ELEM SI

WT % 2.31

(b)

Fig. 5. MEB-EDX of Cu samples after 1 y (a,b) and 2 y (c,d) in La Plats station. (a,c) In plan, (b,d)

cross section.

and in series with the uncompensated transfer function:

ohmic

resistance

(RJ,

according

to the following
-I

20~):

RQ+

JOC (.

in + R,;')-'+(jwC,,

R,-:)-]-+(&ah +L-j

(1)

where Ci, and Ri, are the capacitance and resistance of the C&O inner layer, respectively, and C,, and R,, those related to the outer porous layer contribution. For each set of experimental impedance data the parameters involved in the transfer function (1) were evaluated using non-linear least-square fit procedures. The results of the

A survey of Argentinean atmospheric corrosion: II

669

WI- % 2.65 5.91 3.35 1.28 0.73 86.07

----iCo%-

Fig. 5.

(Continued)

computer fit to measured impedance spectra are shown in Figs 10-12. The excellent agreement between the experimental results and optimum fit data indicates that the interface is well represented by the proposed model. The fit parameters obtained for each sample at its corresponding corrosion potential are summarized in Table 4. The ohmic resistance Rn was close to 77.5 Q cm for the whole set of samples. The parameters Ci,, Ri,, C,, and R,, obtained at the corrosion potential of each sample are in agreement with the data resulting from the previous techniques, although they were found to be only weakly exposure time-dependent. Provided there are not dielectric relaxations in the measurable frequency range, Ci, can be expressed by

670

J. R. Vilche et al.

3 75
CL K ,l FE UJ -----loooo9 74 14.81 0 49 I65 69.51

-.._._I.

. .._._-

--

_J_-

Fig. h

MEB-EDX

of Cu samples after 7 5 (a.h) and 4 y (c,d) in Camet station. cross section

(a,c) In plan, (b.d)

C:,, = tto/d

(2)

where to is the permitivity of the vacuum, 8.85 x lo- I4 F cm-, and t and d the dielectric constant and the thickness of the oxide film, respectively. Thus, d can be evaluated from (2) by assuming a conservative value of t = 20 for the CUE0 and experimental values of Ci,. From the C,, data shown in Table 4, the average value for the calculated thickness results 38.6 urn, though the values themselves vary between 0.2 and 300 urn. Values of CuzO

A survey of Argentinean atmospheric corrosion: II

671

KEL!ZM "AL SI

s
Ii CL K FE

WI % 7.01 15.42 0.69 12.72 0.64 2.33 61.20 _________ loO.cm

UT % 0.23 1.81 97.96 -----__-_ 100.00

Fig. 6. (Continued)

resistivity were reported to be between 10 and 50Qm,26 those corresponding to the compounds present in the outer layer such as cupric sulphate were as high as 7 x 1O*2 am.27 This is in agreement with the fact that the calculated values of Ri were smaller than those of R,, in almost all the experimental measurements (Table 4). As the structure of the corrosion product layers formed on Cu samples after exposure at the atmosphere is close to that found on Zn samples in a previous work,** i.e. a compact inner layer and a porous outer layer, the transfer functions used to interpret the experimental results were similar. It is worth noting, however, that Cu samples developed

Fig. 7.

MEB-EDX

of Cu samples after

I y (a,b) and 3 y (c,d) in Jubany station. (a,c) In plan, (b,d) cross section.

product layers with higher impedance values than those found on Zn. In this way, the impedance contribution corresponding to the oxygen reduction reaction was considered in parallel with that due to the passive film and characterized by its charge transfer resistance Rcath and a diffusional contribution ZW which accounts for the mass transport controlled oxygen electroreduction reaction.** This cathodic reaction occurs in the conditions of the experimental impedance measurement and it does not necessarily correspond to that during the atmosphere exposition. In those cases in which the impedance of the passive layers on the sample is lower than that of the oxygen reaction, the influence of this cathodic process on corrosion

A surveyof Argentinean atmospheric corrosion:II

wr%
2.03 2.62 0.95 ----___-_ 94.40 Icm.00

Fig. 7.

(Continued)

the impedance spectra becomes negligible. Thus, the parameters related to the latter process are hidden and can not be determined by the fitting procedure, as it can be seen in Table 4. The Warburg component, Zw, related to a process under diffusion control through a finite region of length 1corresponding to the thickness of the composite passive layer can be described by2 Zw = Roo(jS)-12tanh(jS) 112 (31

where the diffusion resistance R ~0 is the limit of Zw(iw) as o+O and the parameter S = 1w/ D. For high values of either w or 12/D, tanh(iS) approaches 1 and the Warburg impedance is represented by

674

J. R. Vilche et al

-80 3 G , -40 g

0 0
Fig. 8.

40

80 Re/kQcm2

120

160

Impedance diagram obtained with Cu samples weathered 1 y at La Plata test station after 1 to 72 h immersion times in the electrolyte.

ZW = R~0/1(2w/D)-~(l the corresponding

-j)

= cm- 12(1 -j)

(4)

values of the Warburg coefficient (Tare given in Table 4. From the corrosion rate data, an increasingly protective effect of the products formed with exposure time (Fig. 1) evidently depends on their pollutant content (Table 3) and on the TOW (Table 1). DC potentiostatic steps did not show any monotonous trend with exposure time. Factors such as the heterogeneous morphology associated to an irregular pollutant distribution in the corrosion products limits the generalised analysis. For this reason a separate discussion will be done for each different environment.

(0) 2 years (0) 3 years

-2

0 log

-2

-4

[f/Hz]

Fig. 9.

Impedance diagram obtained with Cu samples weathered at Jubany test station after different exposure times.

A survey of Argentinean atmospheric corrosion: II

675

7 f E N = 4

0 ()lyeaY

(0) 2 years (v) 3 years (v) 4 years

ST

80

8 ii 8

60 40 20

-2

0 log

-2

-4

[f/Hz]

Fig. 10. Impedance diagram for Cu samples after different exposition times at Iguazu test station. Full line traces correspond to results fitted by the transfer function (equation (1)).

(0) 1 year (0) 2 years (v) 3 years

L
I

-2

0
log

-2 [f/Hz]

-4

Fig. 11. Impedance diagram for Cu samples after different exposition times at San Juan test station. Full line traces correspond to results fitted by the transfer function (equation (I)).

616

J. R. Vilche et al

-2

0 log [f/Hz]

-2

-4

Fig. 12. Impedance diagram for Cu samples after different exposition times at Villa Martelli test station. Full line traces correspond to results fitted by the transfer function (equation (1)).

Table 4.

Fitting parameters C, (nF cm-*) 26.2 89.2 40.7 0.32 2.18 0.057 0.214 1.77 47.3 112.0
0.745

for copper samples exposed at the different test sites Ri (kR cm) 0.527 0.530 108.7 1.315 715.9 26.65 21.23 3015. 909.8 0.662 4.737 61.52 20.36 36.15 7.819 87.55 35.10
&-ah SD

Test station San Juan

Exposure time (y)

(nF>-) 1.46 8.44 5.58 10.3 0.350 0.388 0.289 0.010 0.260 0.094 1.99 8.90 10.4 7.29 2.34 3.01 2.53 5.16 0.545 1.41

(kQk2) 34.71 61.76 17.09 1890. 1799. 1686. 217.1 4124. 183.7 64.20 1390. 483.1 176.1 365.2 17.25 6.079 0.3977 27.24 230.5

(kn cm*) 13.34 6.342 14.74 142.4 109.3 212.5 2.222 33.58 84.79 1189. 147.9 49.79 46.56 0.101 41.77 187.9

(kR cm2) 13.87 69.66 98.01 247.7 3730. 466.7 45.48 8897. 663.7 480.3 16.19 114.9 314.8 343.5 24.25 66.46 4.791 56.52 217.6

Camet

Iguazu
L

3 4 Villa Martelli

I
3 4 1 2 1 2 3

La Plata Jubany

0.157 6.03 3.65 13.3 3.52 0.468

A survey of Argentinean atmospheric corrosion: II

67-l

The lowest corrosion rates (Fig. 1) were determined in the rural test station of San Juan, with results from 2 to 4y within the values obtained in the three first annual periods. Comparing Fig. 1 and Table 2 from 1 to 4y with Fig. 2b, the morphology is evidently responsible for the poor barrier effect of the corrosion products at 4 y exposure time. This morphology corresponds to the highest current values reached after anodic and cathodic pulses, which also increases with immersion time. It can also be noticed that the protectiveness, estimated though the corrosion rates, was time-independent (Fig. 1). Protectiveness in Iguazu followed the same trend, which could be attributed to an increase in pollutant content masking the effect associated with the thickening of the corrosion products layer after the third year. Otherwise, these product films would increase their protective properties with time of exposure, which correspond to a decrease in the current measured when potentiostatic pulses are applied. At the Villa Martelli urban ambient, protective effect of the products formed in the first year was observed, though no further increase was noticed with time. The morphology in cross section (Fig. 4) after 2-3 y does not justify better performance with increasing exposure time although the pollutant levels decreased (Table 3). In the urban test station La Plata no protective corrosion layers were observed, indicating low adherence, especially in the micrograph in cross section shown after the second year of exposure. A low increase in the protective character of the products can be noticed with time, both from mass losses and from pulses results. In the samples weathered in the marine stations, corrosion rate decrease is the most marked (Fig. l), in spite of the highest Cl- and SOf- contents increase with time. The current values when pulses are applied are correlative to the pollutant content in the corrosion products and at Jubany environment during the respective test period. It has been stated that chloride ions penetrate protective oxide films through pores, flaws (cracks) or other weak spots.30 It was argued that no matter how compact, a passive layer contains flaws through which the chloride ion easily penetrates. The flaws were stated to be large enough to permit the passage of large aggressive ions. This viewpoint stresses that the surface oxide layer plays an essential inert role in the pitting process. The role of halide ions centres around: (a) competitive adsorption with OH- on the available copper surface thus creating sites that are more liable for electrochemical dissolution and (b) competition with OH- attached to Cu(I1) in a soluble intermediate stage, thus enhancing film rupture (through dissolution).30 Furthermore, it was shown that the first step of the anodic dissolution of copper in chloride solution is the formation of the complex CuC12-, and that during the anodic polarization there is always an equilibrium between two compounds: a thin layer of CuCl and a dense liquid layer of dissolved CUCI~-.~ Nevertheless, the degree of the influence of aggressive anions in breaking down protective films depends on both the nature of the surface (oxide) and the added anion. Then, although cuprous chloride or basic cupric chloride (3CU(OH)2.CUC12) can be formed in the surface film,6 the values of the solubility products indicate that the formation of oxides is favored and that the basic copper carbonate is less soluble than chloride compounds.6 The insoluble CuO (or CUE) film is much less stable in the presence of I- compared with Cl-. Compared with Cl- and I- ions, the presence of phosphates and to a lesser extent sulphates, partly inactivate the copper surface and cause a pronounced anodic current decrease of Cu(I1) process.30 Solubility of all compounds increases considerably with decreasing PH.~

678

J. R. Vilche er al.

CONCLUSIONS Investigations on outdoor exposed Cu samples show that the protectiveness of corrosion products as a function of outdoor exposure time is markedly increased in marine environments, moderate in the urban sites and almost negligible in rural environments. While corrosion product morphology is the main factor responsible for the variation in the corrosion rates found among the test stations, pollutant content explains additional and singular tendencies, specifically detected in each site. Furthermore, electrochemical impedance measurements represent a powerful tool to obtain information on the properties of the passivated interface. The experimental results are interpreted in terms of a double passive layer structure model, a compact CuzO inner layer and a porous outer layer, composed by cupric sulphate, nitrate or chloride. The cathodic process, oxygen reduction reaction, present in the experimental measurement conditions was also considered as contributing to the whole electrode impedance.
Acknowledgements-This research project was financially supported by the Consejo National de Investigaciones Cientificas y T&cnicas, the Comision de Investigaciones Cientificas de la Provincia de Buenos Aires, the CYTED program of Spain in the coordination of the MICAT project, and the Fundacion Antorchas.

REFERENCES
1. U. Bertocci and D. Turner, Encyclopedia of Electrochemistry of fhe Elements (ed. A.J. Bard), Vol. II, Marcel Dekker, New York (1974). 2. J. Van Muylder, Comprehensive Treatise of Electrochemisrry (eds. J. OM. Bockris, B.E. Conway, E. Yeager and R.E. White), 4, pp. l-96, Plenum Press, New York (1981). 3. H.-D. Speckmann and H.-H. Strehblow, We&s?. Korros. 35, 512 (1984). 4. M. Perez S, M. Barrera, S. Gonzalez, R.M. Souto, R.C. Salvarezza and A.J. Arvia, Elecrrochim. Acta 35,
1337 (1990). 5. M. Perez S, R.M. Souto, M. Barrera, S. Gonzalez, R.C. Salvarezza and A.J. Arvia, Electrochim. Acta 38, 703 (1993). 6. M. Drogowska, L. Brossard and H. Menard, J. Elecirochem. Sot. 139, 39 (1992). 7. S.R. Ribotta, M.E. Folquer and J.R. Vilche, Corrosion 51, 682 (1995). 8. S.R. Ribotta, M.E. Folquer, L.M. Gassa and J.R. Vilche, Corrosion, in press. 9. H. Leidheiser, The Corrosion of Copper, Tin, and their Alloys, John Wiley and Sons, Inc., New York (1974). 10. I.L. Rozenfeld, Atmospheric Corrosion of Metals, NACE, Houston, Texas (1972). 11. C. Wagner and K. Grunewald, Z. Phys. Chem. B40, 455 (1938). 12. H. Guttman and P.J. Sereda, ASTM STP 435, 326 (1968). 13. S. Feliu, M. Morcillo and S. Feliu Jr, Corros. Sci. 34, 403 (1992); ibid 34, 415 (1992). 14. W.W. Kirk and H.H. Lawson (eds.), Atmospheric Corrosion STP 1239, ASTM, Philadelphia (1995). 15. D. W. Rice, P. Peterson, E.B. Rigby, P.B.P. Phipps, R.J. Cappell and R. Tremourex, J. Electrochem. Sot. 128, 275 (1981). 16. J.F. Dante and R.G. Kelly, J. Elecrrochem. Sot. 140, 1890 (1993). 17. J. Tidblad and C. Leygraf, J. Electrochem. Sot. 142, 749 (1995). 18. S. Zakipour, J. Tidblad and C. Leygraf, J. Electrochem. Sot. 142, 757 (1995). 19. J.A. Gonzalez, E. Otero and C. Cabanas, Br. Corros. J. 25, 125 (1990). 20. D. Persson and C. Leygraf, 1. Electrochem. Sot. 140, 1256 (1993). 21. R.E. Lobnig, R.P. Frankenthal, D.J. Siconolfi, J.D. Sinclair and M. Stratmann, /. Electrochem. Sot. 141, 2935 (1994). 22. J.R. Vilche, F.E. Varela, G. Acuiia, E.N. Codaro, B.M. Resales, A. Femandez and G. Moriena, Corros. Sci.

37, 941 (1995). 23. M. Morcillo, Proc. 1st Panamerican Corrosion and Protection Congress, , Mar &I Plats 1, 21 l-225 (1992). 24. E.B. Castro, S.G. Real, S.B. Saidman, J.R. Vilche and R.H. Milocco, Mater. Sci. Forum 44/45, 417 (1989). 25. R.H. Milocco, E.B. Castro, S.G. Real and J.R. Vilche, Transient Techniques in Corrosion Science and

A survey of Argentinean atmospheric corrosion: II

679

Engineering (eds W.H. Smyrl, D.D. Macdonald and W.J. Lorenz), p. 88. The Electrochemical Society, Pennington (1989). 26. R.S. Carmichael (ed.), Handbook of Physical Properties of Rocks, 1, CRC Press, Boca Raton, Florida (1982). 27. E.W. Washburn (ed.), International Critical Tables of Numerical Data: Physics, Chemistry and Technology, VI, p 153, McGraw-Hill, New York (1928). 28. K. Jiittner, K. Manandhar, U. Seifert-Kraus, W.J. Lorenz and E. Schmidt, Werkstoffe und Korrosion 37, 377 (1986). 29. J. Ross Macdonald (ed.), Impedance Spectroscopy, Interscience, New York (1987). 30. F.M. Al-Kharati and Y.A. El-Tantawy, Corros. Sci. 22, 1-12 (1982). 31. J. Crousier, L. Pardessus and J.-P. Crousier, Electrochim. Acta 33, 1039 (1988).

Você também pode gostar