Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

D Thierry, French Corrosion Institute, Brest, France


D Persson, Swerea KIMAB, Kista, Sweden
N Le Bozec, French Corrosion Institute, Brest, France
© 2018 Elsevier Inc. All rights reserved.

Introduction 55
The Role of Water in Atmospheric Corrosion: Formation of Surface Electrolyte Films 55
The Effect of Atmospheric Gases 56
Atmospheric Particles and Aerosols 58
Atmospheric Corrosion of Zinc and Zinc Coated Steel 59
The effect of carbon dioxide CO2 59
The influence of temperature 60
The effect of sulfur dioxide SO2 62
The initiation of atmospheric corrosion by particles and aerosols 62
The effect of salt composition 63
The influence of the wet/dry pattern 64
The effect of the substrate 64
The rate of atmospheric corrosion during field exposures 65
The influence of climatic parameters on the atmospheric corrosion during field exposures 66
Dose–response models for atmospheric corrosion of zinc 66
Morphology of the corrosion attack and corrosion products on zinc and zinc coated steel under field conditions 68
Corrosion products composition 70
Atmospheric Corrosion of Zinc Alloyed Coated Steel 71
Zn-Mg-Al and Zn–Mg coated steel 71
Zn–Al coated steel 74
References 76
Further Reading 78

Glossary
Crystallization Relative Humidity (CRH) The humidity at which a salt particle crystallizes.
Dose response Statistical model that relates the corrosion rate to the dose of pollutants in the atmosphere.
Deliquescence Relative Humidity (DRH) The relative humidity at which a salt particle forms an electrolyte.
Eutectic The temperature on a phase diagram at which a eutectic mixture forms.
Time of wetness Time during which the temperature is above 0 C and the relative humidity in air is above 80%.

Introduction
The Role of Water in Atmospheric Corrosion: Formation of Surface Electrolyte Films
A prerequisite for atmospheric corrosion to occur is the presence of water on the metal surfaces. Water can be in form of thin
adsorbed water films or bulk-like water films as a result of precipitation or condensation of mist or fog aerosol water droplets
on the surfaces.
The film thickness of an absorbed water film is determined by the nature of the substrate, the air relative humidity (or water
vapor pressure) and the temperature of the system. Fig. 1 shows water adsorption on a metal oxide surface (MgO at 10 C).1 Already
at a relative humidity of 30%–40% the MgO surface is covered by 2–3 monolayers of water, while multilayer adsorption occurs at
a relative humidity larger than approximately 80%. The amount of adsorbed water decreases strongly with the temperature as seen
in Fig. 2 which shows the amount of physically adsorbed water at different temperatures at 93% relative humidity.2
The time of wetness (TOW) is defined as the total time at a temperature higher than 0 C with a relative humidity higher than
80%.3 However, such a calculated TOW value does not necessarily correspond to the actual TOW, because the latter is determined
also by other factors, such the presence of salt particles on the surface.
Soluble particles, such as salt particles, play an important role in atmospheric corrosion since they absorb water above a critical
relative humidity, the deliquescence relative humidity (DRH), and form an electrolyte solution at higher relative humidity. Thus,
when salt particles are present on a metal a thin highly conductive surface electrolyte film is formed at the DRH. The DRH varies

55
56 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

Fig. 1 Adsorption isotherm for water on MgO (0 0 1) at 10 C. Data taken on ascending and descending pressures are given by closed and open
circles, respectively.1

Fig. 2 The amount of physically adsorbed water on the surface of zinc at 93% relative humidity and different temperatures.2

strongly for different salt particles as seen in Table 1. As the relative humidity is increased, more water is absorbed in the electrolyte
layer and the volume of the electrolyte becomes larger. At the same time, the solution is diluted to the point where water activity in the
salt solution corresponds to the relative humidity of the atmosphere. The reverse process of deliquescence, when the salt solution
forms salt particles from a salt solution, upon the decrease of the relative humidity is called efflorescence. The crystallization relative
humidity (CRH) is often not as well defined and occurs often at lower relative humidity compared to the DRH, due to supersaturation
of the salt solution. A schematic illustration of deliquescence and efflorescence process of a soluble particle is given in Fig. 3. Table 1
presents experimental results for the deliquescence and efflorescence processes of pure NaCl and MgCl2. However, a mixture of salt
such as NaCl and MgCl2 salts can maintain an aqueous state over a larger humidity range compared to pure NaCl.6 The marine sea
salt or sea-spray aerosols (SSAs) are composed largely of NaCl with MgCl2 as the second most abundant constituent in SSA.

The Effect of Atmospheric Gases


Atmospheric gases were early recognized as important factors for atmospheric corrosion of metallic materials. An aqueous layer on
a metal surface, in the form of an adsorbed water layer or thicker layers formed, for instance, by precipitation or marine aerosols will
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 57

Table 1 Deliquescence relative humidity and crystallization relative humidity for different salts

Salt Deliquescence relative humidity (DRH) % Crystallization relative humidity (CRH) %

NaCl 75.28a 47a


Na2SO4 84.2a 57–59
NaHSO4 52.0a <5a
NH4Cl 77.1a 47a
(NH4)2SO4 79,97a 37–40
NaNO3 74.5a <5–30a
NH4NO3 61.83a 25–32a
MgCl2 33b, 15.2c 5.5c
ZnCl2 <1
a
Ref. 4.
b
Ref. 5.
c
Ref. 6.

Fig. 3 Schematic illustration of deliquescence and efflorescence process of a soluble particle. CRH, Crystallization relative humidity; DRH, Deliques-
cence relative humidity.

contain gaseous species in equilibrium with the atmosphere. The same is true also for atmospheric droplets and aerosols prior to the
deposition on metal surfaces. The equilibrium between the concentration of a gas (p) in the atmosphere and water in contact with
the atmosphere is described by Henry’s law:
ca ðaqÞ ¼ Hcp pa (1)
3
where ca is the aqueous phase concentration of species a (in mol m ), pa is the partial pressure in the gas phase (in Pa) of the
same species, and Hcp is the Henry’s law coefficient (in mol m 3 Pa 1). A selection of Henry’s law coefficients from the compilation
by Sander7 for gases of interest for atmospheric corrosion is given in Table 2. The values of the Henry’s law coefficient are for water
as a solvent but the solubility of gases in aqueous solution depends on the composition of the solution, that is, on its ionic strength
and on dissolved organics. In general, the solubility of a gas decreases with increasing salt content in the solution. As seen in Table 2,
the solubility of different gases varies strongly and gases such as O2, CO2, and SO2 are relatively insoluble, while others such as
HNO3 and HCCOH are highly soluble. However, the absorption of gases on a metals surface is also strongly dependent on
secondary processes, such as dissociation and redox reactions which can decrease the dissolved species and allow further absorption
in the aqueous surface film. The overall reaction in neutral and alkaline conditions can be written according to Eq. (2):
O2 þ 2H2 O þ 4e /4OH (2)
It is well known that the oxygen reduction reaction proceeds by two main parallel and competitive pathways, including direct
reduction in Eq. (2) and a two-step pathway with hydrogen peroxide as intermediate species according to Eqs. (3) and (4)
O2 þ H2 O þ 2e /HO2  þ OH (3)

HO2  þ H2 O þ 2e /3OH (4)


Thus, the reduction of oxygen molecules at the cathodic sites leads to the formation of hydroxide ions and in many cases also
small amounts of intermediates and reactive oxygen species such as H2O2 and hydroxyl radicals. Hydroxide ions increase the local
pH in the aqueous surface layer and are incorporated in the dominating corrosion products formed on zinc-based materials. Since
electrolyte layers are often thin during atmospheric corrosion, the oxygen cathodic reaction is less subjected to mass transport
restrictions compared to bulk electrolytes for which diffusion of oxygen can be rate determining.
58 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

Table 2 Henry’s law coefficients for common gases (in water at


ambient temperature)7

Species Hcp (mol m 3 Pa 1)

O2 1.3  10 5
O3 1.1  10 4
H2O2 8.3  102
HO2 6.8
HO 3.8  10 1
CO2 3.3  10 4
H2S 1.0  10 3
SO2 1.3  10 2
OCS 2.0  10 4
HCl 1.5  101
Cl2 9.2  10 4
NH3 5.9$10 1
NO2 1.2  10 4
HNO3 2.1  103
HCHO 3.2  101
HCOOH 8.8  101
CH3COOH 4.0  101

Although the Henry’s law coefficient is low, carbon dioxide is present in relatively high amounts (about 400 ppm) in earth atmo-
sphere and has a strong influence on the corrosion processes of zinc. Following the absorption of CO2(g) in an aqueous solution, it
dissociates to form bicarbonate and carbonate ions according to:
CO2 ðgÞ4CO2 ðaqÞ (5)

CO2 ðaqÞ þ H2 O4HCO3 2 þ Hþ ðpKa1 ¼ 6:35Þ (6)

HCO3 2 4Hþ þ CO3 2 ðpKa2 ¼ 10:33Þ (7)


In alkaline conditions, CO2 (aq) dissociates to carbonate and hydrogen ions, allowing continued absorption of CO2. This can
occur at cathodic areas on the surface where the basic conditions are maintained by the cathodic reaction. Carbonate ions are incor-
porated in corrosion products with zinc hydroxycarbonate which is an important component in corrosion products formed on both
field exposed samples and during laboratory investigations.
Another important gas is SO2 which is present in the atmosphere largely as a result of anthropogenic emissions but also due to
natural sources, such as volcanic activity. When SO2 is dissolved in the aqueous surface film it dissociates according to the following
reactions:
SO2 ðgÞ4SO2 ðaqÞ (8)

SO2 ðaqÞ þ H2 O4HSO3 2 þ Hþ (9)

HSO3 2 4Hþ þ SO3 2 (10)


8
The oxidation of S(IV) to S(VI) in aqueous solution can be written as Eq. (11) :

Hn OSO2 ð2nÞ þ 0:5O2 /Hn SO4 ð2nÞ (11)


where n varies with pH. This oxidation reaction has been studied extensively due to its importance in atmospheric chemistry. In
the absence of oxidants or catalysts, the oxidation rate is rather slow. The oxidation of the sulfite species is homogenously catalyzed
by transition metal ions, such as Fe(II) and Mn (II) and their soluble complexes8 and oxide surfaces, such as Fe2O3 and ZnO, can
work as heterogeneous catalysts for the oxidation. Several gaseous species work as oxidants or catalysts such as NO2, O3, and H2O2.
An important promoter of sulfite oxidation is hydrogen peroxide, which leads to high reaction rates particularly at low and near
neutral pH. Hydrogen peroxide and reactive oxygen species, such as the superoxide radical can be formed in the cathodic oxygen
reaction as shown earlier and can promote the oxidation of S(IV) to S(VI).

Atmospheric Particles and Aerosols


Atmospheric particles and aerosols or particles can be either emitted from a source directly as particles (primary aerosols), such as
soil dust, sea salt, and volcanic dust or from anthropogenic sources such as industrial soot and dust produced in combustion
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 59

processes. Particles can also be formed in the atmosphere by conversion of gaseous species to particle (secondary aerosols). There
has recently been an increasing interest on the role of marine SSAs (particularly coarse aerosols in coastal environments) in the
atmospheric corrosion of metallic materials.9 SSAs are produced from waves through bubble bursting by breaking waves at the
ocean or shore. The SSAs constitute the second most abundant source of natural atmospheric aerosols (25%–50% of the aerosol
mass) after mineral dust particles.10 A study of transport of marine aerosols by Cole et al.11 indicated that coarse aerosols produced
by the surf will generally be transported 1–2 km before being scavenged or falling to the ground, while finer ocean produced aero-
sols may travel hundreds of kilometers from the coast. The residence time of the aerosol particles, therefore, varies from a few
minutes to days. NaCl constitutes about 80% of sea salts by mass and with SO42 , Mg2 þ, Ca2 þ, and Kþ as other important compo-
nents. The initially formed SSAs can react with gaseous species and other airborne particles in the atmosphere within few minute to
hours of their residence in air which can affect the chemistry of the aerosol particles strongly. Chloride containing particles and aero-
sols are not only important in coastal regions but also in road environments where road deicing salts are used. NaCl is the most
commonly used salt but salt mixtures with sodium and magnesium or calcium chloride are sometimes employed.

Atmospheric Corrosion of Zinc and Zinc Coated Steel


The effect of carbon dioxide CO2
Zinc hydroxycarbonate is a common corrosion product which is found on zinc exposed in the atmosphere under various condi-
tions. This indicates that carbon dioxide plays an important role in the atmospheric corrosion of zinc and zinc coated steel. Labo-
ratory investigations performed with controlled CO2 content have shown that the corrosion rate and the compositions of the
corrosion products are dependent on the CO2 concentration in the air. Falk et al. found that the corrosion rate of zinc increased
with the CO2 concentration in the humid air during laboratory studies in pure humid air with controlled CO2 content.12 This is
illustrated in Fig. 4. The corresponding calculated mass loss was after 4 weeks exposure in the range from 0.02 mg cm 2 for
1 ppm CO2 to 0.07 mg cm 2 for 40,000 ppm CO2 in humid air. This relatively moderate increase in corrosion with the CO2
content was attributed to the weak acidification of the surface electrolyte by the dissociation reactions of CO2 in Eqs. (3) and
(4). In the presence of CO2 the main corrosion product consisted of zinc hydroxycarbonate, Zn4(OH)6CO3$H2O and
Zn5(OH)6(CO3)2 at higher concentrations. At 1 ppm CO2, ZnO was formed.
For zinc surfaces predeposited with 70 mg cm 2 of NaCl particles, the effect was different compared the uncontaminated zinc
surfaces. The highest corrosion was observed for 1 ppm CO2 and decreased when the CO2 content increased, as seen in Fig. 5.13
The corrosion in the presence of NaCl particles was localized and led to pitting attacks on the surface, in particular for low
(< 1 ppm) CO2 concentration. The amount of soluble chloride was high at low CO2 concentration (< 1 ppm) which shows that
less chloride was bonded in insoluble corrosion products.
The effect CO2 can be explained by the pH lowering effect of the surface electrolyte by the CO2 according to reactions (6) and (7).
This promotes the formation of zinc hydroxychloride, Simonkolleite according to the following Eq. (12):
5ZnOðsÞ þ 2Cl þ 6H2 O4Zn5 ðOHÞ8 Cl2 $H2 O þ 2OH (12)

As shown in the stability diagram in Fig. 6, simonkolleite is stable at intermediate pH and high chloride concentration and
should be present only close to the anodic sites at low CO2 concentrations. At the cathodic sites, the pH is high and the formation
of ZnO will occur at these areas. When the CO2 concentration is increased, the surface electrolyte is neutralized and more

Fig. 4 Mass gain of pure zinc exposed in humid at 95% RH and 22 C with different CO2 concentrations.12
60 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

Fig. 5 Mass gain of zinc with 70 mg/cm2 of NaCl exposed in humid at 95% RH with different CO2 concentrations.13

Fig. 6 Stability diagram for Zn corrosion products on chloride solution, [Zn2 þ] ¼ 0.1 M, T ¼ 25 C.13

Simonkolleite can be formed together with zinc hydroxycarbonate. The decrease of the corrosion rate after 100 h exposure can be
explained by the conversion of soluble chloride to zinc hydroxychloride. Similar results were obtained also for hot dipped galva-
nized steel with higher corrosion rate and a more localized corrosion attack and large cathodic areas at low CO2 concentration
(< 1 ppm) compared to ambient concentrations.14 Formation of ZnO and Simonkolleite are dominating at low CO2 concentration,
while zinc hydroxycarbonate and larger amount of Simonkolleite is formed at CO2 ambient concentrations. The relative amount of
ZnO and zinc hydroxycarbonate (such as Zn5(OH)6(CO3)2) is governed by the following reaction:
5ZnO þ 2CO3 2 þ 5H2 O/Zn5 ðOHÞ6 ðCO3 Þ2 þ 4OH (13)
15
The zinc hydroxyl carbonate is formed at moderately alkaline and intermediate pH in carbonate containing solutions.

The influence of temperature


The atmospheric corrosion of zinc in presence of NaCl particles at the surface at low concentration of CO2 (< 1 ppm) and 350 ppm
CO2 and at different temperatures was studied by Lindström et al.16 The results are shown in Fig. 7. The metal loss was independent
on the temperature for 350 ppm CO2 and the effect of NaCl surface concentration was small. In contrast, when the CO2 concen-
tration was low a strong dependence on the temperature and the NaCl surface concentration was observed. In both cases, the corro-
sion was localized with pitting attacks on the zinc surface. However, heavy pitting was only observed for < 1 ppm CO2. These effects
are attributed to the absorption of CO2 which strongly affects the pH of the surface electrolyte leading to a decrease in pH at the
cathodic areas and the formation of more insoluble zinc hydroxychloride and less ZnO. Thus, the more rapid formation of insol-
uble zinc hydroxychloride at higher temperature compensate for the increase in corrosion by the temperature. The obvious effect is
that the atmospheric corrosion of zinc in presence of NaCl and CO2 is approximately independent on the temperature.
The temperature dependence of the atmospheric corrosion of zinc in the range from  4 C to 22 C has been studied. As seen in
Fig. 8, the corrosion of NaCl contaminated Zn at 95% RH was independent on the temperature above a critical temperature
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 61

Fig. 7 Metal loss as a function of temperature :, <1 ppm CO2 and 70 mg/cm2 NaCl; n, < 1 ppm CO2 and 14 mg/cm2; O, 350 ppm CO2 and
70 mg/cm2 NaCl; and ,, 350 ppm CO2 and 14 mg/cm2 NaCl.16

Fig. 8 Corrosion of zinc as a function of time and temperature. (A) Metal loss at 200 h < t < 1000 h at 22 and  4 C (in the presence of
70 mg/cm2 NaCl); (B) metal loss after 1000 h at  4, 2, 4, 10, and 22 C at two levels of NaCl; and (c) metal loss of Zn after 1000 h at 0.5,
0.0,  0.5, 1.0, and 1.5 C in the presence of 70 mg/cm2 NaCl. All experiments were carried out in 95% RH air containing 400 ppm CO2.17

of  0.5 C.17 The unusual temperature dependence was explained by a change in corrosion mechanisms between  1 C
and  0.5 C. The mode of corrosion at  0.5 C and higher is characterized by the formation of a poorly protective ZnO layer
on the surface where the O2 reduction can occur in pores in the ZnO layer, while the anodic dissolution of Zn occurs below the
ZnO layer. In this corrosion mode, zinc hydroxycarbonate, hydroxylchloride, and ZnO were the main corrosion products.
At  1.0 C and below, Zn(OH)2 (Sweetite) was formed and the zinc surface was covered by a sodium zinc carbonate
(Na2Zn3(CO3)4 3H2O)/Zn(OH)2 layer. These compounds are wide band gap insulators and this layer should not be cathodically
active. However, while this layer is permeable to chloride ions and water, it allows a separation of the anode below the layer and the
cathode on other parts of the metal surface.
62 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

The effect of sulfur dioxide SO2


SO2-induced atmospheric corrosion became a major concern due to the increase of the SO2 levels in the atmosphere by the use of
fossil fuels in connection with the industrialization. The strong effect of sulfur dioxide on the atmospheric corrosion was recognized
very early in the work by Vernon.18,19 Sulfur dioxide has a strong corrosion stimulating effect and it was early suggested that the
oxidation of O2 to form sulfate leads to a pH decrease, a dissolution of surface oxides and an increase of the corrosion. Although
the SO2 levels in the air have decreased in Europe and North America, the effect of SO2 on the atmospheric corrosion of several
metals is still significant. Furthermore, sites with very high pollution levels are today found in Asia together with high temperatures
and TOW.
Under controlled laboratory exposures, it has found that the deposition of SO2 is humidity dependent and that S(IV) sulfite is
quite stable on zinc surfaces.20–23 Combining SO2 and other gaseous species that promote the oxidation of sulfite to sulfate such as
NO2 and O3 lead to the acceleration of the corrosion and the formation of sulfate containing corrosion products.24–26 This is shown
in Fig. 9. Both NO2 and O3 give rise to a synergistic effect in combination with SO2 with an acceleration of the corrosion and
enhanced oxidation to sulfate. On the other hand, the corrosion of zinc surfaces with predeposited NaCl was lower in humid
air with SO2 compared to NaCl free surfaces which was attributed to the formation of corrosion products binding both chloride,
sulfate and blocking physically the surface.26 It was shown that SO2 oxidized rapidly in areas contaminated with NaCl particles on
zinc and copper. This was attributed to the separation of cathodic and anodic processes in the electrolyte droplet, which leads to
conditions that favor oxidation of SO2.

The initiation of atmospheric corrosion by particles and aerosols


Given the importance of SSAs in coastal regions and the influence of chloride containing particles in road environments, the initi-
ation of the atmospheric corrosion by salt particles have been studied in several investigations.
It was shown first by Neufeld et al.27 that NaCl particles induced the initiation of corrosion on zinc surfaces that lead to
secondary spreading of alkaline electrolyte following the initial droplet formation, see Fig. 10. This effect is a consequence of
the establishment of anodic regions in the center of the droplet and cathodic regions at the drop edge. The secondary spreading
is a general effect observed also for metals as iron and copper.28,29 The effect is probably related to instabilities of the droplet driven

Fig. 9 Mass gain of zinc samples exposed in humid air at 95% RH with SO2 (225 ppb), NO2 (400 ppb), and O3 (400 ppb).25

Fig. 10 Secondary spreading of a NaCl droplet (A) 10 min, (B) 20 min, and (C) 40 min after placement of the droplet on Zn surface.9
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 63

by surface tension changes at the edge and the secondary spreading effect will increase the effective cathodic area.30 The extent of
secondary spreading of electrolyte outside the droplet is smaller in ambient CO2 content than at low concentration27,31 and is
inhibited in the presence of SO2. Oxidation of sulfite to sulfate ions occurs rapidly in the cathodic areas at droplet edge and
secondary spreading electrolyte layer.31,32
On zinc, the corrosion products formed in the original droplet area are Simonkolleite, Zn5(OH)8Cl2$H2O and zinc hydroxycar-
bonate at ambient CO2 concentration while the formation of zinc hydroxycarbonate is suppressed at low CO2 content.22,26 When
SO2 is present in the air zinc sulfate and basic zinc sulfate are formed. Schematic descriptions of the corrosion initiation on zinc
surface at low and ambient CO2 concentration are shown in Fig. 11.
Cole et al. studied the initiation of the atmospheric corrosion on zinc surfaces by fine seawater droplets and it was observed that
soon after the droplets were placed on the surface, a thin secondary spreading film formed outside the original droplet area.33,34
Analysis indicated the presence of NaZn4Cl(OH)60 4SO4$6H2O (gordaite) and Zn5(OH)8Cl2$H2O (simonkolleite) in the center
of the drop when the droplet volume was maintained. However, during droplet evaporation, a wider range of species was detected,
including (Zn,Mg)SO4$4H2O (boyleite), zinc chlorate hydrate, and sodium zinc chloride hydrate. In the secondary spread region,
an amorphous phase with a composition and morphology consistent with zinc hydroxycarbonate was present. The phase analysis is
supported by electrochemical data and by analysis of changes in droplet chemistry with time.

The effect of salt composition


The composition of the salt has a strong effect on the atmospheric corrosion under laboratory conditions. Exposures of zinc with
chlorides and sulfate salts of sodium, magnesium, ammonium, and zinc to humid air at constant humidity and temperature
showed considerable variation on metal loss for different salts.35 The metal loss was the highest for samples with sodium chloride
(NaCl) deposits while the presence of Zn2 þ and Mg2 þ led to lower mass loss due to the formation of precipitates in the form of
hydroxide, oxide, or hydroxy salts at the cathodic areas. Similar results were obtained for sulfate salts, that is, the highest corrosion
rate was registered for the samples treated with sodium sulfate (Na2SO4). It was actually found that the corrosion rate of zinc in the
atmosphere was related to the amount of sodium present on the surface, rather than to the amount of chloride or sulfate. The effect
of cations on the atmospheric corrosion of zinc covered by chloride deposits was studied by Prosek et al.36 The cations strongly
affected the corrosion rate of zinc and the corrosivity of cations of chloride salts for zinc increased in order of Mg2 þ < Ca2 þ < Naþ,
see Fig. 12. It was observed that with calcium and magnesium chloride, the formation of hydrozincite decreased during zinc expo-
sure in wet air. This is due to the decrease in pH of the surface electrolyte, caused by the hydrolysis of the cations and the subsequent
enhanced formation of simonkolleite.

Fig. 11 Schematic illustration of the NaCl-induced atmospheric corrosion of zinc in humid air with different CO2 concentration (<5 and
350 ppm).31
64 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

Fig. 12 Effect of cations and their mixtures on weight loss of zinc with 1400 mg/m2 of chloride; numbers indicate the molar ratio of cations in salt
used from. 35

Barton et al. determined the corrosion rate of steel and zinc in model coastal atmospheres with different chloride deposition.32
The corrosion rate was lower for zinc treated with artificial seawater than for zinc with pure NaCl deposition. The authors explained
that the principal difference between the two test solutions lied in the content of calcium and magnesium of artificial seawater. The
hydrolysis of these ions decreases the pH at the metal surface.

The influence of the wet/dry pattern


Atmospheric corrosion seldom takes place under stationary conditions with constant humidity and temperature. Instead, metal
surfaces are exposed in the atmosphere subjected to periodic wetting and drying events as well as periodic variations in humidity
and temperature. The effect of the wet/dry pattern has been studied in several investigations. Yadav et al. investigated the effect of
wet/dry cycling and TOW on zinc and galvanized steel by electrochemical impedance spectroscopy measurements.37,38 The corro-
sion attack was heterogeneous and localized corrosion attacks were observed under wet–dry cyclic conditions in the presence of
chloride ions. The progress of corrosion was similar for both substrates until the anodic localized attack reached the ZnFe layer close
to the steel substrate which became cathodic.
The effect of the TOW is shown in Fig. 13 which gives the reciprocal of the polarization resistance during the wet/dry cycling.
High corrosion rates of the Zn coating were obtained when the drying periods in each cycle was shorter since the surface was wet for
longer periods. On the other hand, long drying periods in each cycle lead to shorter times for the appearance of red rust, although
the total amount of corrosion was smaller. This indicates that the galvanic protection ability of zinc coating layer was not sufficient
under a very thin electrolyte layer.
Zhu et al. studied the effect of drying on open and confined zinc surfaces during exposure to periodic wetting and drying.39 The
wet/dry pattern had a strong influence on the composition of the corrosion products on both open and confined surfaces on zinc. It
found that at longer drying times ( 23 h), the corrosion products were mainly composed of Simonkolleite, Zn5(OH)8Cl2$H2O for
the open surface while Hydrozincite, Zn5(OH)6(CO3)2 was dominating on the confined surfaces. The open surface dries up rapidly
compared to the confined surface which has different drying characteristic with longer effective drying. The corrosion rate and extent
of localization of the corrosion was higher for the confined surfaces and Zn5(OH)6(CO3)2 and ZnO were formed at the cathodic
areas on surface. For shorter drying times, the corrosion rate increases for open surfaces and the composition of the corrosion prod-
ucts was similar with a mixture of simonkolleite, hydrozincite, and ZnO. The results for open surface of electrogalvanized steel40 was
similar as for zinc for the same duration of exposure while the corrosion products on confined surface was dominated by ZnO due
to the strong alkalization of the crevice due to the oxygen reduction which took place on the bare steel substrate following localized
corrosion penetration of the zinc coating.

The effect of the substrate


Although most investigations on the atmospheric corrosion are performed using pure zinc, galvanized steel and electroplated steel
are of much larger technological interest due to its extensive use for corrosion protection of steel. In most cases, zinc coated steel is
produced by the hot dipping in melted zinc or by electrodeposition. The latter process produces an essentially pure zinc layer on the
steel surface while hot dipped galvanizing results in the formation of Zn–Fe phases with increasing amount of Fe toward the steel
below a zone of pure Zn. In the continuous hot dipped galvanized process, Al (in the order of 0.2%) is added to the zinc in order to
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 65

Fig. 13 Corrosion monitoring results of galvanized steel during wet and dry cycle of. immersion in 0.5 M NaCl at 298 K for 1 h and drying at 298 K
60% RH for (A) 11 h, (B) 7 h and (C) 3 h on a horizontal specimen surface, and (D) 3 h on an inclined specimen surface. The appearance of red rust
is indicated by the arrows.38

form a Fe–Al phase on the steel surface which suppresses the formation of Zn–Fe phases. Galvanized steel sheets used for building
applications, appliances and automotive applications have thin zinc coatings produced by the continuous hot dipped galvanized
process or electroplated steel. The continuous hot dipped galvanized zinc layer contains a small amount of Al and it is well known
that the surface has a thin layer of aluminium oxide.41
Until no exposed free steel surface or Zn–Fe phases appears, the atmospheric corrosion rate of hot-dip galvanized steel behaves
very similarly to that of pure zinc and variations in mass losses of zinc and hot-dip galvanized steel of the same order of magnitude
as that between two simultaneously exposed samples of the same metal.42 However, the presence of Al in the continuous hot dip-
ped galvanized zinc layer seems to have an effect on the very initial corrosion product formation.43 The Zn–Fe phases present in
piece wise hot dipped steel are often found to have a better corrosion resistance compared to pure zinc layers. It is known that
the life of hot dipped galvanized steel can be prolonged by annealing. However, corrosion products of Zn–Fe alloys are yellowish
brown in color reflecting the iron content in the layer.

The rate of atmospheric corrosion during field exposures


It is often observed that the long-term kinetics of the atmospheric corrosion of zinc follow a simple power law:
M ¼ atn or=log M ¼ log a þ n log t (14)
where M is the mass loss after t years exposure and a and n are constants. Corrosion data for zinc at different test sites reasonably
fits a linear log–log plot with correlation coefficients close to unity.44
The equation can be derived for diffusion controlled reactions of a film covered metal surface and for a pure diffusion control n
will have a value of ½. If the diffusion process is accelerated by removal of the layer by erosion, dissolution, cracking, or other similar
processes which is probably the case for atmospheric corrosion, the exponent n will have a value higher than 0.5 approaching
a limiting value of 1.0.44 It has been found that the corrosion rate of skyward-facing surfaces of galvanized steel is linear with
time, while corrosion on the groundward side is parabolic in shape,45 which is consistent with the diffusion model, since the sky-
wark side is subjected to precipitation and washing events which periodically remove a part of the corrosion product layer. Fig. 14
shows examples of the corrosion of zinc samples after long-term exposures at different exposure sites in Spain.46 The general trend
was linear for the different sites irrespective of the type of atmosphere (rural, urban, industrial, and marine), see Table 3. However,
concerning the behavior of the two marine sites, Alicante and Barcelona the corrosion rates deviates from the linear behavior during
the first 10 years exposure. This was attributed to corrosion product formation-compaction-detachment processes.
66 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

Fig. 14 Zinc corrosion after long-term exposures in different exposure sites in Spain. Small diagrams show corrosion rates.46

Table 3 Characteristics of atmospheric test stations in Spain in Fig. 14,47

Deposition rate (mg/m2 d) Zinc


Test site Time of First year ISO 19 corrosivity
Atmosphere (code) wetness (h/year) SO2 Cl corrosion (mm) category

Rural El Escorial 3900 8 Negligible 0.9 C3


Urban Madrid 2100 56 Negligible 1.4 C3
Industrial Bilbao 3000 81 41 5.2 C5
Marine
Light Barcelona 3200 52 27 2.2 C4
Severe Alicante 4300 126 219 5.8 C5

The influence of climatic parameters on the atmospheric corrosion during field exposures
There exists a large body of work in which the atmospheric corrosion of zinc and galvanized steel has been studied in different
climates with the influence of corrosion-stimulating species, such as SO2 and chloride. The SO2 content in the atmosphere has
been of large interest due to the strong corrosion-stimulating effect. However, there is a trend of decreasing SO2 concentrations
which is reflected in a decrease in the zinc corrosion as seen in the data from the long-term ICP materials test program,48
Figs. 15 and 16. The ICP materials test program has 52 test sites all over Europe as well a few sites in the United States, Canada,
and Israel.
The effect of chloride deposition is exemplified in Fig. 17 for marine test sites in the MICAT program in Latin-America.47 The
effect of chloride between the sites leads to a variation in corrosivity class from C1 to CX according to the ISO 9223.3

Dose–response models for atmospheric corrosion of zinc


A significant body of work has been made with the aim to establish numerical relationships between atmospheric corrosion rate and
climatic parameters. Statistical models of atmospheric corrosion have been elaborated since the 1950s. The models relate the corro-
sion of metallic materials to pollutants dose and climatic parameters are denoted dose–response models. These are based on data
from the results of atmospheric test programs and environmental data and should reflect the physicochemical nature of the corro-
sion process. Dose–response functions for the atmospheric corrosion of zinc surfaces in atmospheres with significant contributions
of SO2 and Cl deposition to the corrosion are as follows:49

CZn ¼ 0:0053½SO2 0:43 ½Tow0:53 expffZn g þ 0:00071½Cl0:68 ½Tow0:30 expf0:11T g (15)


Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 67

Fig. 15 Average SO2, NO2, and O3 levels relative year 1987 for sites in the ICP materials test program.47

Fig. 16 Average relative zinc corrosion for the sites in the ICP materials test program for traditional zinc and blasted zinc. The zinc corrosion of
traditional zinc at 1987 was taken as reference.48

Fig. 17 Zinc corrosion (1 year) versus chloride deposition for marine test sites with insignificant effect of SO2. Adapted from Reference Almeida, E.,
Morcillo, M., Rosales B. (2000). Atmospheric Corrosion of Zinc: Part 2: Marine Atmospheres. Br. Corros. J. 35, 289–296.
68 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel


fZnðT Þ ¼ 0 for T  10 C;


fZnðT Þ ¼ –0:032ðT–10Þ for T > 10 C;

where CZn is the corrosion rate of zinc and [SO2] and [Cl] are the deposition rates of SO2 and chloride, respectively. The corro-
sion rates for zinc calculated by Eq. (12) versus temperature for different deposition rates of SO2 and chloride are shown in Fig. 18.
The mean experimental data of relative humidity and TOW are given in Fig. 19 which were obtained by processing climate data
from several exposure programs.50

Morphology of the corrosion attack and corrosion products on zinc and zinc coated steel under field conditions
The corrosion attacks on pure zinc and zinc coatings are often regarded as uniform which also is the assumption when the extent of
the corrosion is determined from the mass loss of the metal samples. However, in studies where the morphology of the corrosion
attack has been considered more explicitly, the corrosion attack of zinc is often quite uneven and sometimes localized. Pitting-like
attack is observed, particularly in marine environments. Cole et al. observed pitting of pure zinc samples after exposure in marine
and other sites in tropical exposure countries with relatively low SOx levels.51 No pitting, or pits with very low aspect ratio (depth/
width) was seen for specimens exposed at sites with high SOx levels. Cross sections of specimens exposed to marine conditions
exhibited thick but very irregular corrosion layers containing both pedestal formed features, which were associated to the corrosion
pits. A cross section of a corrosion pit on zinc from the tropical exposure is shown in Fig. 20.
For zinc exposed up to 4 years in a marine tropical atmosphere in China, Cui et al. observed the atmospheric corrosion of zinc
initiated with localized corrosion, which is characterized with some pits, grooves, and intergranular corrosion.47 As the exposure
time increased, more pits were generated and coalesced with the neighbor pits resulting in a more uniform corrosion attack. The
occluded pit in in the marine sites in these studies were enriched in Cl and S, while the Zn, O, and C content was high on the outer
parts of the corrosion product scale.
The corrosion morphology for zinc surfaces after long-term exposures up to 16 years exposure in Spain was investigated by De la
Fuente et al.46 Pitting was observed on zinc after long-term exposures in rural, urban atmospheres, while a corrosion attack profile
with peaks and valleys was observed in the industrial atmosphere. The corrosion product layers were thin in these environments in

Fig. 18 The temperature dependence of the annual corrosion rate of zinc, mm/year, calculated from the dose–response function (6) for nine combi-
nations of the depositions of SO2 and Cl, mg/(m2 day). The group depositions of Cl are indicated on the graph for each set of curves. The
following depositions of SO2 correspond to the numbers of curves (1, 4, 7)d10; (2, 5, 8)d35; (3, 6, 9)d80. The value of Tow changes with T
according to the graph in Fig. 19 for all the curves.49

Fig. 19 Mean experimental data of Rh and Tow as a function of temperature.


Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 69

Fig. 20 Cross section of corrosion products underlying corrosion pit on a zinc sample exposed in marine tropical site (Cowley beach).51

contrast to the corrosion product layers formed in marine atmospheres. The latter showed the existence of two clearly differentiated
subzones: a cracked inner sublayer and an outer sublayer, with numerous microcracks and fine microscopic channels. Pitting-like
corrosion attack of zinc has also been reported in older studies.52
The morphology of the corrosion attack and the corrosion product scale for zinc coated steel was investigated after a worldwide
exposure in different environments.53 The corrosion attack was localized for continuous hot dipped galvanized steel after exposures
in marine exposure sites as well as for sites classified as marine/industrial and industrial/urban. Fig. 21 shows a cross section of

Fig. 21 Micrograph and SEM-EDS maps of O, S, and Cl for a cross section of HDG exposed 2 years at the site in Wanning.53
70 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

a hot-dip galvanized steel coating after 2 years exposure in a marine exposure site in Wanning in China.50 The zinc coating is
completely penetrated to the steel surface by the pitting attack. The SEM-EDS analysis shows that the S and Cl is enriched in the
inner parts and bottom of the pits, which is similar to the observations made by Cole and Cui for pure zinc.51,52
The amount of corrosion products on zinc surfaces is dependent on many factors, where the amount of precipitation is partic-
ularly important. The effect of rain washing was indicated already in the older work by Schikorr which found that the amount of
corrosion products that remains on the surface is similar to the amount washed away.54 The amount of corrosion products is often
larger on sheltered samples although the corrosion rate is usually lower on them compared to unsheltered conditions.55 The corro-
sion and run off of zinc are influenced by different physiochemical mechanisms. The latter is dependent on electrochemical reac-
tions at the metal/corrosion product interfaces while the latter is governed by chemical reactions at the corrosion product/
atmosphere interface.56

Corrosion products composition


The composition of the corrosion products has been the subject of investigation in numerous investigations over the years.55–58 The
dominating corrosion products found in field exposures on zinc and zinc coated steel are basic salts of zinc containing anions, such
as carbonate, chloride, and sulfate. Zinc oxide is frequently found and zinc hydroxide has been reported as well. Nonbasic salt of
zinc, such as ZnCO3 and ZnSO4$nH2O have also been reported, see Table 4.
Odnevall and Leygraf have described reaction sequences how different corrosion products are formed in environments consid-
ering differences the relative amount of chloride and sulfur pollution.5,59 Considering the crystal structural resemblances between
zinc hydroxycarbonate, hydroxychloride, hydroxyl sulfate, and the hydroxyl chloro sulfates, it is plausible that the transformation
between these is facilitated under the proper environmental conditions. For instance, the proposed formation sequences under shel-
tered conditions in ambient atmosphere dominated by sulfur pollution and lower pollutions are described in Fig. 22.
However, the formation of different corrosion products should be strongly connected with the electrochemical processes
which occur at the metal/corrosion product interface. As discussed by Cole,51 the formation of corrosion products on zinc by
the influence of acidic and moderately acidic aerosol particles directly related to the electrochemical processes taking place during
initiation and propagation of corrosion processes.

Table 4 Zinc compounds reported in corrosion products on zinc and zinc coated
steel after field exposure and laboratory investigations

Compound Field exposures/atmosphere

ZnO Rural, urban, marine, industrial


Zn(OH)2 Rural, urban, marine, industrial
ZnCO3 Rural, urban, marine
Zn4(OH)6CO3$H2O Marine
Zn5(OH)6(CO3)2 Rural, urban, marine, industrial
Zn5(OH)8Cl2$H2O Marine, industrial
ZnSO4 Industrial
Zn4(OH)6SO4$nH2O Rural, urban, marine, industrial
NaZn4Cl(OH)60 4SO4$6H2O Urban, marine
Zn4Cl2(OH)6SO4$5H2O Rural, urban, marine, industrial

Adapted from References Zhang, X. G. (1996). Corrosion and Electrochemistry of Zinc. Plenum
Press: New York; Graedel T. E. (1989). Corrosion mechanisms for Zinc Exposed in the Atmosphere.
J. Electrochem. Soc. 136, 193C–203C.

Fig. 22 Schematic reaction sequence for the formation of zinc corrosion products on zinc or galvanized steel.57
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 71

The composition of the corrosion products varied with the environmental conditions as illustrated in Table 4 for sites within a world-
wide exposure zinc coated steel.53 In sites which could be classified as marine, marine/urban, and marine/industrial, Zn5(OH)6(CO3)2
was the dominating corrosion product but Zn5(OH)8Cl2$H2O, (Zn(OH)2)3$ZnSO4$nH2O, and NaZn4(SO4)(OH)6Cl$6HyO were
also present on most of these sites. The corrosion was localized as seen in Fig. 21 with corrosion pits enriched with sulfate and chloride
containing corrosion products close to the anodic area in the bottom of the pit while zinc hydroxycarbonate was present in the outer
parts of the corrosion products and outside the pits where cathodic processes occurs. The relative content of sulfate versus carbonate
containing corrosion products for the pure marine sites was relatively low and increased for sites classified as marine/urban and
marine/industrial. Sites classified as industrial/urban sites have the highest sulfate to carbonate ratio. The sulfate/carbonate ratio reflect
variations in the composition of the corrosion products which are related to the environmental conditions at the different exposure sites.
The formation of corrosion products on hot dipped galvanized steel during the localized corrosion attack in marine and marine/
industrial atmospheres is described schematically in Fig. 23. Corrosion initiation is assumed following the deposition of aerosol
particles and droplets and the establishment of cathodic regions at the edge of the droplet and anodic attack of the zinc coating
in the center of the droplet. Zn4Cl2(OH)6SO4$5H2O and NaZn4Cl(OH)60 4SO4$6H2O can be formed shortly after initiation in
accordance with laboratory studies of the interaction of SSA droplets with zinc surfaces.33 Zinc hydroxycarbonate, Zn5(OH)6(CO3)2
which is the dominating phase at the pure marine sites and is present mainly in the outer parts of the corrosion product layers and
outside the pits. Propagation of the corrosion pit can occur beneath the corrosion product scale and leads to the enrichment of S and
Cl containing corrosion products in the pit. This requires that chloride and sulfate ions are transported through the corrosion
product scale, either through defects and cracks in the corrosion product scale or by ionic transport through corrosion product
membranes.60 It should be emphasized that the formation of corrosion products in zinc and galvanized steel in an atmospheric
environment is a complex and continuously changing process. The composition of the corrosion scale on the field exposed samples
is not only controlled by the electrochemical processes at the metal/corrosion product interfacial region but also with the interaction
of the corrosion product with the atmosphere. Run-off of components from the corrosion products can alter the composition of
outer layers depending on the solubility of different phases in the patina.

Atmospheric Corrosion of Zinc Alloyed Coated Steel


Zn-Mg-Al and Zn–Mg coated steel
As already mentioned in previous articles, hot dip galvanized steel (HDG) is widely used for applications in the automotive,
building, and appliance industry. This is linked to the corrosion properties of HDG providing both a barrier and a galvanic protec-
tion of the steel substrate. In the automotive industry, there is a need to move to new coatings with improved corrosion resistance in

Fig. 23 Schematic model of initiation and propagation of atmospheric corrosion of HDG marine environments and marine/urban, marine/industrial
sites.54
72 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

order to (1) reduce the thickness of the coating and (2) eliminate the need of additional corrosion protection in confined areas such
as hem flanges.
In the last 2–3 decades, a large body of literature has been published on further enhancements of the corrosion properties of
HDG by additions of magnesium, aluminium, or a combination of both.61–67 New coating alloys containing aluminium and
magnesium have been developed for continuous galvanized steel sheet.61–69 The zinc magnesium aluminium coated steel will
be refered as ZM in this review. These coatings normally contain 0.2%–11% of magnesium and 0.1%–3% of aluminium which
are added to zinc. The addition of magnesium and aluminium results in a complex microstructure containing primary zinc,
MgZn2, binary eutectic MgZn2–Zn, binary eutectic Zn–Al, and a ternary eutectic depending upon the composition of the coating.
This is illustrated in Fig. 24 for a coating containing 2% of magnesium and 2% of aluminium. In addition to this, there are recent
reports on the corrosion of zinc’magnesium coatings produced by physical vapor deposition (PVD).70,71 It should be mentioned
that the microstructure may differ according to the exact composition of the coating. However, the phases mentioned earlier are
generally included. In this case, the coating is mainly formed by an inner layer of zinc and an outer layer of MgZn2. Alternatively,
Zn and Mg are codeposited simultaneously and coatings containing from about 10 to 16 wt% Mg are formed. As PVD and the hot-
dip galvanizing process result in materials with very different structures and surface concentrations of magnesium and aluminium, it
is often difficult to compare the results.
The corrosion performance of these novel coatings have been studied both under laboratory and field conditions. Fig. 25 shows
the results obtained after 6 weeks exposure in different cyclic corrosion tests. For more details on the different cyclic corrosion tests
use reference.72 The relative performance of the ZM coated steel is dependent upon testing conditions and it is highest for tests that
involve a high chloride load (e.g., neutral sat spray at 5% and VDA 621-415). It should be noticed that this generally corresponds to
conditions for which very high corrosion on HDG and zinc are observed. Similar results have been observed in the literature when
comparing the time to red rust of zinc’magnesium coated steel produced by PVD and submitted to different accelerated tests.64

Fig. 24 SEM micrograph of a cross section of the ZnAl2%Mg2% coated steel coating including EDX-analyses.43

Fig. 25 Weight loss of Zn-Mg-Al coated steel versus HDG as a function of testing conditions.
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 73

Although, the corrosion performance of ZM coated steel is always superior to HDG in chloride environments, it exists few reports
with an opposite situation. LeBozec et al.14 and Persson et al.43 reported that ZM coated steel showed poor corrosion behavior in
CO2 free and SO2 environments, respectively. One possible explanation provided by the authors was that under these conditions
the surface pH became very alkaline (e.g., CO2 free environments) or acidic (SO2 containing environments), resulting in a much
higher dissolution of the zinc and aluminium rich phases.
Long-term data on the corrosion behavior of ZM coated steel compared to HDG are very scarse. Salgueiro and coworkers
compared the corrosion behavior of ZM to HDG in different environmental conditions including marine, rural, and urban loca-
tions.73 A ratio of improvement between 2.4 and 3.3 was found after 2 years of exposure. Similar results were found by Tomandl
et al. after 1 year exposure in marine and marine tropical field sites74 and by Thierry et al. after 4 years exposure at 15 different field
station worldwide.75 It should be noted that a good performance of ZM coated steel compared to HDG was also measured in indus-
trial environments that contain moderate amount of SO2. It should be noticed that as already mentioned for HDG, the corrosion of
ZM materials was highly localized after field exposure with a preferential attack of the binary or ternary eutectics.75
Despite the large body of literature on the subject, there is still no clear single explanation of the role of magnesium and
aluminium in the enhanced corrosion properties of Zn–Mg and Zn-Al-Mg coated steel compared to HDG. However, several mech-
anisms for the improvement are suggested in the literature.
It is rather well accepted that the binary and ternary eutectics containing magnesium and magnesium and aluminium, respec-
tively, are dissolved preferentially. Some authors have attributed the better behavior of these coatings to the formation of Mg con-
taining corrosion products.76–78 Possible hypothesizes on the effect have been an enhanced insulating character of the zinc
corrosion products due to magnesium and a limited charge transfer reaction at the grain boundaries.79,80 The role of magnesium
has also been explained through a reduction of the surface alkalinity by the formation of Mg(OH)2 which favors the formation of
simonkolleite and zinc hydroxy sulfate.76,81–83 The results observed in Refs. 76–80 imply that magnesium is present in the inner
layer of the corrosion products formed on Zn-Al-Mg coated steel, although no corrosion products associated with magnesium
has been clearly identified. The presence of aluminium in zinc-aluminium’magnesium coated steel seems to result in different
mechanisms with the formation of layer double hydroxides (LDH). These compounds have the general formula Zn(or
Mg)xAly(A)m(OH)n$zH2O (where A is an anion and in atmospheric corrosion often carbonate). Although the formation of LDH
has been well documented (see for instance Refs. 43, 81,82), the protective character of these compounds is still a subject for further
investigations. The mechanisms of formation of LDH and ZnO have been studied by Salgueiro in different electrolytes.84 This is
schematically illustrated in Fig. 26. It was found that the formation of LDH is delayed in NH4þ and HCO3– environments due

Fig. 26 Schematic illustration of the mechanism of formation of ZnO and LDH in NaCl and rain water electrolytes on HDG and ZM coatings
according to.84
74 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

to their buffering effects.84 It should be noticed that as demonstrated by LeBozec and coworkers, LDH are not formed in CO2 free
environments resulting in poor performance of these coatings.14
Finally, it should be noticed that although these coatings present a very complex microstructure, no systematic investigation has
been performed on the effect of microstructure on the corrosion behavior of ZnAlMg coated steel. However, the results obtained by
Prosek et al. on model alloys having composition close to that used for coated steel have shown that microstructure can play an
important role. Indeed, it was shown that alloys with finer microstructure led to better corrosion performance probably due to
a blocking of the cathodic sites.85

Zn–Al coated steel


Zn–Al coated steel are widely used in the construction industry as an alternative to replace HDG due to their superior corrosion
properties. Basically, two main coatings are commonly used, Galfan that contains 5% aluminium and Galvalume that contains
55% Al and 1.6% Si. The 5% aluminium alloy coating has a relatively complex microstructure. During solidification, the b’zinc
(proeutectic) phase, containing about 1% aluminium, crystallizes first, followed by the crystallization of a eutectic of zinc

Fig. 27 SEM image on Galfan coating according to Ref. 86

Fig. 28 SEM micrographs of Galvalume coatings alloyed with different contents of magnesium: (A) Galvalume coating, (B) G–0.5 Mg coating, (C)
G–1.5 Mg coating, and (D) G–2.5 Mg coating.88
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 75

Fig. 29 Sequence of formation of corrosion products on Zn, Galfan, and pure Aluminium in marine environments according to reference.86

Fig. 30 Mechanisms of formation of corrosion products formed in marine environments on galvalume coated steel according to [95].

containing about 5% aluminium. The microstructure of Galfan is thus characterized by a two-phase structure, a zinc-rich eta (h)
proeutectoid phase surrounded by a eutectic type phase consisting of beta (b) aluminium and eta (h) zinc lamellae (see
Fig. 27). However, the microstructure can be varied depending upon the cooling rate.87 The microstructure of Zn 55%Al (Galva-
lume) shows aluminium-rich dendrites, zinc-rich interdendritic areas interspersed with occasional silicon particles and a thin inter-
metallic layer which binds the coating to the steel (see Fig. 28). It should be noticed that addition of magnesium in Galvalume leads
to some well-crystallized Mg-containing dendritics (see Fig. 28C and D).
In the case of Zn5Al (Galfan), the corrosion rate upon long-term exposure is generally 1.5–2 times lower than that of HDG. This
has been associated to the formation of different corrosion products as displayed in Fig. 29. As explained before for ZM materials,
the role of hydrotalcites is probably very important in the performance of Zn5Al coated steel.
In the case of Zn55Al (Galvalume), it is believed that the corrosion starts in the interdendritic phases of the coating, mainly
affecting the zinc in the alloy.86,89 Ramus Moreira and coworkers identified two different phases in the dendritic and interdendritic
areas and concluded that the atmospheric corrosion process begins in an aluminium rich phase present in the interdendritic areas.90
On the other hand, reports on the identification of corrosion products on these materials are rather scarce. Seré and coworkers iden-
tified basic zinc aluminium carbonate (Zn6Al2(OH)16CO3 4H2O) together with minor quantities of Bayerite (Al(OH)3) in the
corrosion products formed upon exposure to salt spray test.91 Similar results were also found by other authors on Zn55Al under
wet storage conditions.92 Li found Zn5(OH)8Cl2, Zn4(CO3)2(OH)6 and Zn6Al2(OH)16CO3 after immersion in aerated seawater.93
In a recent investigation by Persson et al. and based on exposure at a marine site, it was found that the corrosion rate of Zn55Al was
about four times lower than that of HDG at the same location. These values are in good agreement to those reported elsewhere by
Palma and coworkers86 in different locations in Spain and they are consistent with the observations that Zn55Al coated steel has
a considerably lower corrosion rate in severe marine and polluted industrial environments.94 A mechanism of corrosion protection
was also proposed as illustrated in Fig. 30. From this, it was concluded that hydrotalcites play an important role in the corrosion
protection and that their formation in the interdendritic areas blocks the cathodic reaction because of their inability to act as cathode
and the overall corrosion rate should decrease. Finally as reported by Li et al., addition of magnesium in galvalume at about 1.5%
further increased the corrosion properties of the coating due a better compactness of the corrosion products and enhanced barrier
properties.88

See also: Corrosion of Steel in Concrete: New Challenges; Kinetics of Oxide Growth of Passive Films on Transition Metals; Wet Corrosion of Stainless
Steels and Other Chromium-Bearing Alloys.
76 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

References

1. Foster, M.; D’Agostino, M.; Passno, D. Water on MgO(100)dAn Infrared Study at Ambient Temperatures. Surf. Sci. 2005, 590, 31–41.
2. Strekalov, P. V.; Agafonov, V. V.; Mikailovkii, Y. N. Effect of temperature on the adsorption of moisture and the corrosion rate of zinc under atmospheric conditions. Prot. Met.
1972, 8, 521.
3. ISO 9223 Standard. Corrosion of Metals and AlloysdCorrosivity of Atmospheres: Classification, 2012.
4. Jacobsson, M. Z. Fundamentals of Atmospheric Modelling, Cambridge University Press, 2015.
5. Lide, D. R. In Handbook of Chemistry and Physics Eighty; Lide, D. R., Ed., 3rd ed.; CRC Press: Boca Raton, FL, 2002.
6. Gupta, D.; Eom, H.-J.; Cho, H.-R.; Ro, C.-U. Hygroscopic Behavior of NaCl–MgCl2 Mixture Particles as Nascent Sea-Spray Aerosol Surrogates and Observation of Efflo-
rescence During Humidification. Atmos. Chem. Phys. 2015, 15, 11273–11290.
7. Sander, R. Compilation of Henry’s Law Constants (version 4.0) for Water as Solvent. Atmos. Chem. Phys. 2015, 15 (8), 4399–4981.
8. Martin, L. R. In SO2, NO and NO2 oxidation mechanisms. In Acid Precipitation Series, Vol. 3, Calvert, Ed., Vol. 3; Butterworth Publ., 1984; pp 63–100.
9. Cole, Y. S.; Azmat, N. S.; Kanta, A.; Venkatraman, M. What Really Controls the Atmospheric Corrosion of Zinc? Effect of Marine Aerosols on Atmospheric Corrosion of Zinc.
Int. Mater. Rev. 2009, 54 (3), 117–133.
10. Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the Upper and Lower Atmosphere: Theory, Experiments, and Applications, Academic Press: San Diego, 2000. xxii, 969 pp.
11. Cole, I. S.; Paterson, D. A.; Ganther, W. D. Holistic Model for Atmospheric Corrosion Part 1-Theoretical Framework for Production, Transportation and Deposition of Marine
Salts. Corros. Eng. Sci. Technol. 2003, 38 (2), 129–134.
12. Falk, T.; Svensson, J.-E.; Johansson, L.-G. The Role of Carbon Dioxide in the Atmospheric Corrosion of Zinc. J. Electrochem. Soc. 1998, 145, 39–44.
13. Falk, T.; Svensson, J.-E.; Johansson, L.-G. The Influence of CO2 and NaCl on the Atmospheric Corrosion of Zinc. J. Electrochem. Soc. 1998, 145, 2993–3000.
14. LeBozec, N.; Thierry, D.; Rohwerder, M.; Persson, D.; Luckeneder, G.; Luxem, L. The Effect of Carbon Dioxide on the Atmospheric Corrosion of Zn-Al-Mg Coated Steel.
Corros. Sci. 2013, 74, 379–386.
15. Kannangara, C.; Conway, B. E. Zinc Oxidation and Redeposition Processes in Aqueous Alkali and Carbonate Solutions. J. Electrochem. Soc. 1987, 134, 521–523.
16. Lindström, R.; Svensson, J.-E.; Johansson, L.-G. The Atmospheric Corrosion of Zinc in the Presence of NaCldThe Influence of Carbon Dioxyde and Temperature.
J. Electrochem. Soc. 2000, 147, 1751–1757.
17. Esmaily, M.; Mortazavi, N.; Svensson, J. E.; Johansson, L.-G. Evidence for an Unusual Temperature Dependence of the Atmospheric Corrosion of Zinc. J. Electrochem. Soc.
2016, 163, C864–C872.
18. Vernon, W. H. J.; Whitby, L. The Open-Air Corrosion of Copper. A Chemical Study of the Surface Patina. J. Inst. Metals 1929, 42, 181–202.
19. Vernon, W. H. J. A Laboratory Study of the Atmospheric Corrosion of Metals Part IdThe Corrosion of Copper in Certain Synthetic Atmospheres, with Particular Reference to
the Influence of Sulphur Dioxide in Air of Various Relative Humidities. Trans. Faraday Soc. 1931, 27, 255–277.
20. Sydberger, T.; Vannerberg, N. G. The Influence of the Relative Humidity and Corrosion Products on the Adsorption of Sulfur Dioxide on Metal Surfaces. Corros. Sci. 1972,
12, 775.
21. Duncan, J. R.; Spedding, D. J. The Effect of Relative Humidity on Adsorption of Sulphur Dioxide on to Metal Surfaces. Corros. Sci. 1973, 13, 993.
22. Duncan, J. R.; Spedding, D. J. Initial Reactions of SO2 after Adsorption on to Metals. Corros. Sci. 1973, 13, 881.
23. Persson, D.; Leygraf, C. Initial Interaction of Sulfur Dioxide with Water Covered Metal Surfaces: An In Situ IRAS Study. J. Electrochem. Soc. 1995, 142, 1459–1468.
24. Svensson, J.-E.; Johansson, L.-G. In A Laboratory Study of the Atmospheric Corrosion of Zinc: Synergistic Effects of Trace Amounts of SO2, NO2 and NaClA Laboratory Study
of the Atmospheric Corrosion of Zinc: Synergistic Effects of Trace Amounts of SO2, NO2 and NaCl; 1989In 11th Scandinavian Corrosion Congress, Stavanger, Norway, F47.
25. Svensson, J.-E.; Johansson, L.-G. A Laboratory Study of the Effect of Ozone, Nitrogen Dioxide and Sulfur Dioxide on the Atmospheric Corrosion of Zinc. J. Electrochem. Soc.
1993, 140, 2210–2216.
26. Svensson, J.-E.; Johansson, L.-G. A Laboratory Study of the Initial Stages of the Atmospheric Corrosion of Zinc in the Presence of NaCl; Influence of SO2 and NO2. Corros.
Sci. 1993, 34, 721–740.
27. Neufeld, A. K.; Cole, I. S.; Bond, A. M.; Furman, S. A. The Initiation Mechanism of Corrosion of Zinc by Sodium Chloride Particle Deposition. Corros. Sci. 2002, 44, 555–572.
28. Tsuru, T.; Tamiya, K.-I.; Nishikata, A. Formation and Growth of Micro-Droplets During the Initial Stage of Atmospheric Corrosion. Electrochim. Acta 2004, 49, 2709–2715.
29. Chen, Z. Y.; Persson, D.; Nazarov, A.; Zakipour, S.; Thierry, D.; Leygraf, C. In Situ Studies of the Effect of CO2 on the Initial NaCl-Induced Atmospheric Corrosion of Copper.
J. Electrochem. Soc. 2005, 152, B342–B351.
30. Nazarov, A.; Thierry, D. Rate Determining Reactions of Atmospheric Corrosion. Electrochim. Acta 2004, 49, 2717–2724.
31. Chen, Z. Y.; Persson, D.; Leygraf, C. Initial NaCl-Particle Induced Atmospheric Corrosion of ZincdEffect of CO2 and SO2. Corros. Sci. 2008, 50 (1), 111–123.
32. Barton, K.; Cuc, T. D.; Bartonova, S. Eifluss der Kationenzusammensetzung von Chloridlösungen beim Modellieren von Korrosionseinflussen der Kustenatmospharen. Mater.
Corros. 1977, 28, 17–19.
33. Cole, I. S.; Muster, T. H.; Furman, S. A.; Wright, N.; Bradbury, A. Products Formed during the Interaction of Seawater Droplets with Zinc Surfaces: I. Results from 1- and 2.5-
Day Exposures. J. Electrochem. Soc. 2008, 155, C244–C255.
34. Azmat, N. S.; Ralston, K. D.; Muddle, B. C.; Cole, I. S. Corrosion of Zn under Acidified Marine Droplets. Corros. Sci. 2011, 53, 1604–1615.
35. Lindstrom, R.; Svensson, J.-E.; Johansson, L.-G. The Influence of Salt Deposits on the Atmospheric Corrosion of Zinc. J. Electrochem. Soc. 2002, 149, B57–B64.
36. Prosek, T.; Thierry, D.; Taxén, C.; Maixner, J. Effect of Cations on Corrosion of Zinc and Carbon Steel Covered With Chloride Deposits Under Atmospheric Conditions. Corros.
Sci. 2007, 49, 2676–2693.
37. Yadav, A. P.; Nishikata, A.; Tsuru, T. Degradation Mechanism of Galvanized Steel in Wet–Dry Cyclic Environment Containing Chloride Ions. Corros. Sci. 2004, 46, 361–376.
38. Yadav, A. P.; Nishikata, A.; Tsuru, T. Electrochemical Impedance Study on Galvanized Steel Corrosion Under Cyclic Wet–Dry ConditionsdInfluence of Time of Wetness.
Corros. Sci. 2004, 46, 169–181.
39. Zhu, F.; Persson, D.; Thierry, D.; Taxen, C. Formation of Corrosion Products on Open and Confined Zinc Surfaces Exposed to Periodic Wet/Dry Conditions. Corrosion 2000,
56, 1256–1265.
40. Zhu, F.; Persson, D.; Thierry, D. Formation of Corrosion Products on Open and Confined Metal Surfaces Exposed to Periodic Wet/Dry ConditionsdA Comparison Between Zinc
and Electrogalvanized Steel. Corrosion 2001, 57, 582–590.
41. Feliu, S.; Barranco, V. XPS Study of the Surface Chemistry of Conventional Hot-Dip Galvanised Pure Zn, Galvanneal and Zn–Al Alloy Coatings on Steel. Acta Mater. 2003, 51,
5413–5424.
42. Atteraas, L.; Haagenrud, S. Atmospheric Corrosion in Norway. In Atmospheric Corrosion; Ailor, W. H., Ed., J. Wiley: New York, 1982; pp 873–891.
43. Persson, D.; Thierry, D.; Lebozec, N.; Prosek, T. In Situ Infrared Reflection Spectroscopy Studies of the Initial Atmospheric Corrosion of Zn–Al–Mg coated Steel. Corros. Sci.
2013, 72, 54–63.
44. Benarie, M.; Lipfert, F. L. A General Corrosion Function in Terms of Atmospheric Pollutant Concentrations and Rain pH. Atmos. Environ. 1986, 20, 1947–1958.
45. Legault, R. A.; Pearson, V. P. Atmospheric Corrosion in Marine Environments. Corrosion 1978, 34, 433–437.
46. De la Fuente, D.; Castaño, J. G.; Morcillo, M. Long-Term Atmospheric Corrosion of Zinc. Corros. Sci. 2007, 49 (3), 1420–1436.
47. Almeida, E.; Morcillo, M.; Rosales, B. Atmospheric Corrosion of Zinc. Part 2: Marine Atmospheres. Br. Corros. J. 2000, 35, 289–296.
48. Tidblad, J.; Kucera, V.; Ferm, M.; et al. Effects of Air Pollution on Materials and Cultural Heritage: ICP Materials Celebrates 25 Years of Research. Int. J. Corros. 2012, 2012,
496321, 16 p.
Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel 77

49. Mikhailov, A. A.; Tidblad, J.; Kucera, V. The Classification System of ISO 9223 Standard and the Dose-Response Functions Assessing the Corrosivity of Outdoor Atmospheres.
Prot. Met. 2004, 40 (6), 541–550.
50. Mikhailov, A. A.; Tidblad, J.; Kucera, V. Model for the Prediction of the Time of Wetness from Average Annual Data on Relative Air Humidity and Air Temperature. Prot. Met.
2000, 36, 533–540.
51. Cole, I. S.; Ganther, W. D.; Furman, S. A.; Muster, T. H.; Neufeld, A. K. Pitting of Zinc: Observations on Atmospheric Corrosion in Tropical Countries. Corros. Sci. 2010, 52 (3),
848–858.
52. Cui, Z.; Li, X.; Xiao, K.; et al. Corrosion Behavior of Field-Exposed Zinc in a Tropical Marine Atmosphere. Corrosion 2014, 70 (7), 731–748.
53. Persson, D.; Thierry, D.; Karlsson, O. Corrosion and Corrosion Products of Hot Dipped Galvanized Steel during Long Term Atmospheric Exposure at Different Sites Worldwide.
Corros. Sci. 2017.
54. Schikorr, G. Relation between the Atmospheric Corrosion of Zinc and Nickel and the Sulphur Dioxide Content of the Air (in German). Metall 1961, 15, 981–987.
55. Zhang, X. G. Corrosion and Electrochemistry of Zinc, Plenum Press: New York, 1996.
56. Odnevall Wallinder, I.; Leygraf, C. A Critical Review on Corrosion and Runoff from Zinc and Zinc-Based Alloys in Atmospheric Environments. Corrosion 2017, 73 (9),
2327–6932.
57. Graedel, T. E. Corrosion Mechanisms for Zinc Exposed in the Atmosphere. J. Electrochem. Soc. 1989, 136, 193C–203C.
58. Leygraf, C.; Graedel, T. Atmospheric Corrosion, John Wiley & Sons: New York, 2000.
59. Odnevall, I.; Leygraf, C. In Atmospheric Corrosion, ASTM STP 1239; Kirk, W. W., Lawson, H. H., Eds., American Society for Testing of Materials: Philadelphia, 1995; pp
215–229.
60. Saito, M.; Tokuda, K.; Shimoda, N. Nippon Steel & Sumitomo Metal Technical Report No. 108, March 2015, 2015.
61. LeBozec, N.; Thierry, D.; Peltola, A.; et al. Corrosion Performance of Zn-Mg-Al Coated Steel in Accelerated Corrosion Tests Used in the Automotive Industry and Field
Exposures. Mater. Corros. 2013, 64, 969–978.
62. Prosek, T.; Larché, N.; Vlot, M.; Goodwin, F.; Thierry, D. Corrosion Performance of Zn-Al-Mg Coatings in Open and Confined Zones in Conditions Simulating Automotive
Applications. Mater. Corros. 2010, 60, 412–420.
63. Vlot, M.; Bleeker, R.; Maalman, T.; Van Perlstein, E. In MagiZinc™; A new generation of hot-dip galvanised productsMagiZinc™; A new generation of hot-dip galvanised
products; 2006In Proc. of Galvanized Steel Sheet Forum 2006, Dusseldorf.
64. Koll, T.; Beier, F.; Böddeker, T.; Hinkel, M.; Fritzsche, C.; Schulz, S. In Performance of a ZM-Coating with 1.6% Mg and 1.6% Al in the Autobody Process ChainPerformance
of a ZM-Coating with 1.6% Mg and 1.6% Al in the Autobody Process ChainIn 4th International Conference on Steels in Cars and Trucks, Braunschweig, Germany, June 15–
19, Verlag Stahleisen GmbH: Dusseldorf, Germany, 2014; p 2014.
65. Goo, B. J.; Lee, B. R.; Moon, M. B.; Oh, S. J. In Corrosion Properties of Zn Alloy Coated Steel Sheet for Automotive PanelsCorrosion Properties of Zn Alloy Coated Steel
Sheet for Automotive Panels; 2014In 19th International Corrosion Congress, November 2–6, 2014, Jeju, Korea.
66. Prosek, T.; Nazarov, A.; Le Gac, A.; Thierry, D. Coil-Coated Zn–Mg and Zn–Al–Mg: Effect of Climatic Parameters on the Corrosion at Cut Edges. Prog. Org. Coat. 2015, 83,
26–35.
67. Frankel, G. S. Pitting Corrosion of Metals: A Review of the Critical Factors. J. Electrochem. Soc. 1998, 145, 2186–2198.
68. Kairy, S. K.; Rometsch, P. A.; Diao, K.; et al. Exploring the Electrochemistry of 6xxx Series Aluminium Alloys as a Function of Si to Mg Ratio, Cu Content, Ageing Conditions
and Microstructure. Electrochim. Acta 2016, 190, 92–103.
69. Zander, D.; Pieper, C.; Köster, U. In Influence of the Casting Method on Microstructure and Corrosion of AZ91 and AM50; Kainer, K. U., Ed.In Proceedings of the 7th
International Conference on Magnesium Alloys and their Applications, Wiley-VCH Verlag: Weinheim, 2007; pp 757–762.
70. Krieg, R.; Vimalanandan, A.; Rohwerder, M. Corrosion of Zinc and Zn-Mg Alloys With Varying Microstructures and Magnesium Contents. J. Electrochem. Soc. 2014, 161,
C156–C161.
71. Vimalanandan, A.; Bashir, A.; Rohwerder, M. Zn–Mg and Zn–Mg–Al Alloys for Improved Corrosion Protection of Steel: Some New Aspects. Mater. Corros. 2014, 65,
392–400.
72. LeBozec, N.; Blandin, N.; Thierry, D. Accelerated Corrosion Tests in the Automotive Industry: A Comparison of the Performance Towards Cosmetic. Mater. Corros. 2008, 59,
889–894.
73. Salgueiro Azevedo, M.; Allély, C.; Ogle, K.; Volovitch, P. Corrosion Mechanisms of Zn(Mg, Al) Coated Steel in Accelerated Tests and Natural Exposure: 1. The Role of
Electrolyte Composition in the Nature of Corrosion Products and Relative Corrosion Rate. Corros. Sci. 2015, 90, 472–481.
74. Tomandl, A.; Labrenz, E. The Behavior of ZnAlMg Alloys in Marine Environments. Mater. Corros. 2016, 67, 1286–1296.
75. Thierry D., Person D. and LeBozec N. Performance of Zn-Al-Mg coated steel upon 4 years exposure in field. Corros. Sci., in preparation.
76. Hosking N.C., Ström M., Shipway P. and Rudd C., Corrosion Resistance of Zinc Magnesium Coated Steel. Corros. Sci. 49, 3669–3695.
77. Prosek, T.; Nazarov, A.; Bexell, U.; Thierry, D.; Serak, J. Corrosion Mechanisms of Model Zinc Magnesium Alloys in Atmospheric Conditions. Corros. Sci. 2008, 50,
2216–2231.
78. Tsujimura, T.; Komatsu, A.; Andoh, A. In Influence of mg Content in Coating Layer and Coating Structure on Corrosion Resistance of Hot Dip Zn-Al-Mg Alloy Coated Steel
SheetInfluence of mg Content in Coating Layer and Coating Structure on Corrosion Resistance of Hot Dip Zn-Al-Mg Alloy Coated Steel SheetIn Proceedings Galvatech 01,
June 26–28, Brussels, Belgium; 2001; pp 145–152.
79. Diler, E.; Rioual, S.; Lescop, B.; Thierry, D.; Rouvellou, B. Chemistry of Formation of Corrosion Products of Zn and MgZn Pure Phases under Atmospheric Conditions. Corros.
Sci. 2012, 65, 178–186.
80. Li, B.; Dong, A.; Zhu, G.; et al. Investigation of the Corrosion Behavior of Continuously Hot-Dip Zn-Mg Coating. Surf. Coat. Technol. 2012, 206, 3989–3999.
81. Volovitch, P.; Vu, T.; Allely, C.; Abdel, A.; Ogle, K. Understanding Corrosion Via Corrosion Products Characterization. Corros. Sci. 2011, 53, 2437–2445.
82. Schuerz, S.; Fleischanderl, M.; Luckeneder, G.; et al. Corrosion Behavior of Zn-Al-Mg Coated Steel Sheet in Sodium Chloride Containing Environments. Corros. Sci. 2009, 51,
2355–2365.
83. Elvins, J.; Spittle, J.; Sullivan, J.; Dorsley, D. The Effect of Magnesium Additions on the Microstructure and Cut Edge Resistance of Zinc Aluminium Alloy Galvanized Steel.
Corros. Sci. 2008, 50, 1650–1658.
84. Salgueiro, M.; Allély, C.; Ogle, K.; Volovitch, P. Corrosion Mechanisms of Zn (Mg,Al) Coated Steel in Accelerated Corrosion Tests and Natural Exposure. Corros. Sci. 2015, 90,
472–481.
85. Prosek, T.; Hagström, J.; Persson, D.; et al. Effect of Microstructure of Zn-Al and Zn-Al-Mg Model Alloys on Corrosion Stability. Corros. Sci. 2016, 110, 71–81.
86. Palma, E.; Puente, J. M.; Morcillo, M. The Atmospheric Corrosion Mechanism of 55%Al-Zn Coating on Steel. Corros. Sci. 1998, 40, 61–68.
87. Elvins, J.; Spittle, J. Microstructural Changes in Zinc Aluminium Alloy Galvanizing as a Function of Processing Parameters and their Influence on Corrosion. Corros. Sci. 2005,
47, 2740–2759.
88. Li, S.; Gao, B.; Tu, G.; et al. Effect of Magnesium on the Microstructure and Corrosion Resistance of Zn-55Al-1.6Si Coatings. Constr. Build. Mater. 2014, 71, 124–131.
89. Zoccola, J. C.; Townsend, H. E.; Borzillo, A. R.; Horton, J. B. In Atmospheric Corrosion Behavior of Aluminium-Zinc Alloy-Coated Steel; Coburn, S. K., Ed., American Society
for Testing Materials: Philadelphia, 1978; pp 165–184. ASTM STP 646.
90. Ramus Moreira, A.; Panossian, Z.; Camargo, P. L.; et al. Zn/55Al Coating Microstructure and Corrosion Mechanism. Corros. Sci. 2006, 48, 564–576.
91. Seré, P. R.; Zapponi, M.; Elsner, C. I.; Di Sarli, A. R. Comparative Corrosion Behaviour of 55Aluminium–Zinc Alloy and Zinc Hot-Dip Coatings Deposited on Low Carbon Steel
Substrates. Corros. Sci. 1998, 40, 1711–1723.
92. Odnevall Wallinder, I.; He, W.; Augustsson, P. E.; Leygraf, C. Corros. Sci. 1999, 41, 2229–2249.
78 Atmospheric Corrosion of Zinc and Zinc Alloyed Coated Steel

93. Li, Y. Formation of Nano-Crystalline Corrosion Products on Zn–Al Alloy Coating Exposed to Seawater. Corros. Sci. 2001, 43, 1793–1800.
94. Persson, D.; LeBozec, N.; Thierry, D. Corrosion Product Formation on Zn55Al Coated Steel upon Exposure in a Marine Atmosphere. Corros. Sci. 2011, 53, 720–726.
95. Zhang, X.; Leygraf, C.; Wallinder, I. Atmospheric Corrosion of Galfan Coatings on Steel in Chloride Rich Environments. Corros. Sci. 2013, 73, 62–71.

Further Reading

Leygraf, C.; Odnevall Wallinder, I.; Tidblad, J.; Graedel, T. Atmospheric Corrosion, Wiley & Sons, 2016.
Marcus, P., Ed. Corrosion Mechanisms in Theory and Practice, 3rd edn.; CRC Press, 2011.
Zhang, X. G. Corrosion and Electrochemistry of Zinc, Springer Science, 2013.
Porter, F. C. Corrosion Resistance of Zinc and Zinc Alloys, Marcel Dekker Inc., 1994.
Dunbar, S. R.; Showak, W. Atmospheric Corrosion of Zinc and its Alloys. In Atmospheric Corrosion; Ailor, W. H. J., Ed., Wiley and Sons: New York, 1982; pp 529–552.

You might also like