You are on page 1of 24

Emirates Journal for Engineering Research, 11 (1), 1-24 (2006) (Review Paper)

ATMOSPHERIC CORROSION OF MATERIALS


S. SYED
Corrosion Research Group, Atomic Energy Research Institute, King Abdulaziz City for Science and Technology, P.O. Box 6086, Riyadh-11442, Saudi Arabia. Email: sabirsyed2k@yahoo.com (Received January 2006 and accepted May 2006)

. . .
In recent years, atmospheric corrosion of materials have attracted materials community for it accounts for more failures on both a tonnage basis and cost basis than any other type of environmental corrosion. Tremendous amounts of materials in industries, automobiles, bridges and buildings are exposed to the atmosphere and attacked by pollutant and water. In this review paper attention is paid to the atmospheric corrosion of various materials, influence of exposure parameters and basics theory of atmospheric corrosion. Keywords: Materials, atmospheric corrosion, corrosion processes

1. INTRODUCTION
The word corrosion is derived from the latin corrosus which means eaten away or consumed by degrees; an unpleasant word for an unpleasant process[1]. Corrosion is defined as the destruction of materials caused by chemical or electrochemical action of the surrounding environment. This phenomenon is experienced in day to day living. The most common examples of corrosion include rusting, discoloration and tarnishing[2]. Corrosion is an ever occurring material disease. It can only be reduced it cannot be prevented because thermodynamically it is a spontaneous phenomena. In fact, economy of any country would be drastically changed if there were no corrosion. For example, automobiles, ships, underground pipelines and house-hold appliances would not require coatings. The stainless steel industry would disappear and copper would be used for electrical applications. Although corrosion is inevitable, its cost could be reduced. Corrosion can be fast or slow. Sensitized 18-8 stainless steel is badly attacked in hours by polythionic acid. Railroad tracks usually show slight rusting not sufficient to affect their performance over many years. The famous iron Delhi Pillar in India was made almost 2000 years ago and is almost as good as new. Its height is 32 feet and dia 2 feet. It should be noted however, that it has been exposed mostly to arid conditions [3].

1.1. Classification of Corrosion Process Corrosion process can be conveniently classified as follows [4]:
Corrosion process

Chemical corrosion
Direct oxidation, corrosion by liquid metals, fused halides, non-aqueous solution, etc.

Electrochemical corrosion
Immersion Underground corrosion corrosion Atmospheric corrosion

Reaction of metals with dry air or oxygen is considered as a chemical corrosion. High temperature oxidation of metals and tarnishing of metals like copper, silver etc. fall in this category. Of late this is also considered to be an electrochemical process with the diffusion of oxygen (inwards) and metal ions (outwards) through the oxide layer, the electromotive force at metal-oxide interface being the driving force. Electrochemical corrosion occurs in the presence of electrolyte. The reaction is considered to take place at the metal-solution interface with the creation of local cathodic and anodic sides on the metal surface [5].

S. Syed

1.2. Atmospheric Corrosion The term atmospheric corrosion comprises the attack on metal exposed to the air as opposed to metal immersed in a liquid. Atmospheric corrosion is the most prevalent type of corrosion for common metals[6]. Atmospheric corrosion is a subject of global concern because of its importance to the service life of equipment and durability of the structural materials. While there is a general agreement on the possible types of parameters that may lead to corrosion, these studies suffer severely from the lack of generality in the sense that their predictive capability is extremely poor. Conventional atmospheric parameters that may lead to metal corrosion comprise of weathering factors such as temperature, moisture, rainfall, solar radiation, wind velocity, etc. Air pollutants such as sulphur dioxide, hydrogen sulphide, oxides of nitrogen, chlorides have also been found to contribute to atmospheric corrosion[7]. The complexity and diverse nature of the atmospheric pollutants make the prediction of the atmospheric corrosion difficult. The synergistic interaction of the variables must also be considered in the model for arriving at a definite solution. A direct approach to the problem is to measure the observed corrosion rates and the participating atmospheric parameters and correlate them. The correlation equations, thus derived, are known as damage functions and they have been found to be extremely useful, though in a restricted manner, as the results are not easily transferable from one place to another [8]. Predictability of atmospheric corrosion, in principle, should be based upon the complete understanding of the corrosion process and interdependence of the contributing parameters. Extensive data have been collected all over the world on atmospheric corrosion of metals exposed at different locations. Empirical and semi-empirical relationships have been developed to generalize these observations. Most prominent of these relationships have been the linear and exponential dependence of corrosion rate with relative humidity, pollutant levels and temperature [9]. Grossman [10] has investigated the atmospheric factors which determined the time of wetness of the outdoor structures. A thermodynamic perspective of copper tarnishing by SO2 in the presence of moisture was reported by Chawla and Payer[11]. Walters [12-13] carried out some exhaustive studies on the laboratory simulation of atmospheric corrosion by SO2 detailing the apparatus, electrochemical techniques and example results. The effect of pollutants such as SO2, NaCl, dust, etc., on the critical humidity for the rusting to occur was well documented by Vassie [14]. A statistical evaluation of the atmospheric corrosion of stainless steel was undertaken by Blank and Lherbier[15]. The atmospheric corrosion rates of mild steel and low alloy cast steels were studied by Thomas and Alderson; Briggs [16,17],

respectively. The damage function describing the atmospheric deterioration of materials due to acidic deposition was studied in detail by Lipfert [18]. Stiles and Edney studied the potential damage to galvanized steel by the dissolution of zinc into thin aqueous films as a function of residence time, acidic species and pH[19]. The corrosion product Zn2+ correlated linearly with incident H+ concentration. Some developments in the atmospheric corrosion testing were carried out by Pourbaix and Pourbaix[20] and the assessed the corrosion behaviour of different types of steel in both natural and laboratory simulated conditions. Atmospheric corrosion can further be conveniently classified into dry, damp and wet categories. Dry oxidation takes place in the atmosphere with all metals that have a negative free energy of oxide formation. The damp moisture films are created at a certain critical humidity level (largely by the adsorption of water molecules), while the wet films are associated with dew, ocean spray, rainwater, and other forms of water splashing. By its very nature, atmospheric corrosion has been reported to account for more failures in terms of cost and tonnage than any other form of corrosion. The atmospheric environments are classified as rural, urban, industrial, marine, or combinations of these. These types of atmospheres have been described as follows [21-23]. Rural: Rural environments are usually free of aggressive agents (deposition rate of SO2 and NaCl lower than 15 mg m-2 day-1). Their principal corrosives consist of moisture, relatively small amounts of sulfur oxides (SOX), and carbon dioxide (CO2) from various combustion products. Ammonia (NH3), resulting from the decomposition of farm fertilizers, may also be present. Rusting becomes pronounced when the relative humidity exceeds a certain value. For clean air this value is about 70 percent. Rural environments generally are not aggressive towards metals. This type of atmosphere is generally the least corrosive and normally does not contain chemical pollutants, but does contain organic and inorganic particulates. Arid and tropical types are especially extreme cases in the rural category. Urban: Urban atmosphere is similar to rural atmosphere where there is little industrial activity, characterized by pollution composed mainly of SOx and NOx variety, from motor vehicles and domestic fuel emissions which, with the addition of dew or fog, generate a highly corrosive wet acid film on exposed surfaces (deposition rate of SO2 higher than 15 mg m-2 day-1 and that of NaCl lower than this value). Industrial: The most potent causes of corrosion in industrial environments are the sulfur oxides (SOX) and nitrogen oxides (NOX) produced by the burning of automotive fuels and fossil fuels in power stations. The critical relative humidity, above which metals corrode, drops to about 60 percent when these airborne pollutants are deposited on the metal surface.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

These atmospheres are also associated with concentrations of chlorides, phosphates, hydrogen sulphate, ammonia and its salts. Marine: The corrosiveness of a marine environment depends on the topography of the shore, wave action at the surf line, prevailing winds and relative humidity. While the corrosiveness decreases rapidly with increasing distance from the shore, severe storms can carry salt spray inland as much as 15 km. Salt is deposited on steel surfaces by marine fog and windblown spray droplets(deposition rate of NaCl higher than 15 mg m-2 day 1). This contamination induces severe corrosion at relative humidities exceeding about 55%. This environment is characterized by proximity to the ocean and salt laden air that can produce very severe corrosion damage on many structural materials, enhance galvanic corrosion, and accelerate deterioration of protective coating systems. Marine atmospheres are usually highly corrosive. The principal culprit in marine atmospheres is the chloride (CL-) ion derived from sodium chloride.

other crevice corrosion conditions. The reduction of atmospheric oxygen is one of the most important reactions in which electrons are consumed. In the presence of gaseous air pollutants, other reduction reactions involving ozone and sulfur and nitrogen species have to be considered [24]. For atmospheric corrosion in near-neutral electrolyte solution, the oxygen reduction reaction is applicable (Eq.) O2 + 2H2O + 4e- 4O HTwo reaction steps may actually be involved, with hydrogen peroxide as an intermediate, in accordance with (Eqs). O2 + 2H2O + 2e- H2O2 + 2 H2O2 + 2e- 2 If oxygen from the atmosphere diffuses through the electrolyte film to the metal surface, a diffusionlimited current density should apply. It has been shown that a diffusion transport mechanism for oxygen is applicable only to an electrolyte-layer thickness of approximately 30 m and under strictly isothermal conditions [25]. The predicted theoretical limiting current density of oxygen reduction in an electrolyte-layer thickness of 30 m significantly exceeds practical observations of atmospheric corrosion rates. It can be argued, therefore, that the over-all rates of atmospheric corrosion are likely to be controlled not by the cathodic oxygen reduction process, but rather by the anodic reaction(s). The anodic process: Equation represents the generalized anodic reaction that corresponds to the rate-determining step of atmospheric corrosion. M Mn+ + neThe formation of corrosion products, the solubility of corrosion products in the surface electrolyte, and the formation of passive films affect the overall rate of the anodic metal dissolution process and cause deviations from sample rate equations. Passive films distinguish themselves from corrosion products, in the sense that these films tend to be more tightly adherent, are of lower thickness, and provide a higher degree of protection from corrosive attack. Atmospheric corrosive attack on a surface protected by a passive film tends to be of a localized nature. Surface pitting and stress corrosion cracking in aluminum and stainless alloys are examples of such attack. Relatively complex reaction sequences have been proposed for the corrosion product formation and breakdown processes to explain observed atmospheric corrosion rates for different classes of metals. Fundamentally, kinetic modeling rather than equilibrium assessments appears to be appropriate for the dynamic conditions of alternate wetting and drying of surfaces corroding in the atmosphere. A framework for treating atmospheric corrosion phenomena on a theoretical basis, based on six different regimes, has been presented by Graedel [26]. The regimes in this so-

2. BASIC OF ATMOSPHERIC CORROSION


2.1. Theory of Atmospheric Corrosion A fundamental requirement for electrochemical corrosion processes is the presence of an electrolyte. Thin-film invisible electrolytes tend to form on metallic surfaces under atmospheric exposure conditions after a certain critical humidity level is reached. It has been shown that for iron, the critical humidity is 60 percent in an atmosphere free of sulfur dioxide. The critical humidity level is not constant and depends on the corroding material, the tendency of corrosion products, surface deposits to absorb moisture and the presence of atmospheric pollutants. In the presence of thin-film electrolytes, atmospheric corrosion proceeds by balancing anodic and cathodic reactions. The anodic oxidation reaction involves the dissolution of the metal, while the cathodic reaction is often assumed to be the oxygen reduction reaction. It should be noted that corrosive contaminant concentrations can reach relatively high values in the thin electrolyte films, especially under conditions of alternate wetting and drying. Oxygen from the atmosphere is also readily supplied to the electrolyte under thin-film corrosion conditions. The cathodic process: If it is assumed that the surface electrolyte in extremely thin layers is neutral or even slightly acidic, then the hydrogen production reaction (Eq.) can be ignored for atmospheric corrosion of most metals and alloys. 2H+ + 2e- H2 Exceptions to this assumption would include corrosive attack under coatings, when the production of hydrogen can cause blistering of the coating, and

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

S. Syed

called GILDES-type model are the gaseous region (G), the gas-to-liquid interface (I), the surface liquid (L), the deposition layer (D), the electrodic layer (E), and the corroding solid (S). For the gaseous-layer effects, such as entrainment and detrainment of species across the liquid interface, chemical transformations in the gas phase, the effects of solar radiation on photosensitive atmospheric reactions, and temperature effects on the gas phase, reaction kinetics are important. In the interface regime, the transfer of molecules into the liquid layer prior to their chemical interaction in the liquid layer is studied. Not only does the liquid regime receive species from the gas phase, but species from the liquid are also volatilized into the gas phase. Important variables in the liquid regime include the aqueous film thickness and its effect on the concentration of species, chemical transformations in the liquid, and reactions involving metal ions originating from the electrochemical corrosion reactions. In the deposition zone, corrosion products will accumulate, following their nucleation on the substrate. The corrosion products formed under thin film atmospheric conditions are closely related to the formation of naturally occurring minerals. Over long periods of time, the most thermodynamically stable species will tend to dominate. The nature of corrosion products found on different metals exposed to the atmosphere is shown in Table 1. The solution known as the inner electrolyte can be trapped inside or under the corrosion products formed. The deposited corrosion product layers can thus be viewed as membranes, with varying degrees of resistance to ionic transport. Passivating films tend to represent strong barriers to ionic transport.
Table 1. Nature of corrosion products formed on four metals [27]. Common species Al Fe Cu Zn Al(OH)3 Al2 O3, Al2 O3, 3H2O Fe2O3, FeOOH, FeSO4, 4H2O Cu2O1,Cu4 SO4 (OH)6, Cu4 SO4 (OH)6, 2H2O, Cu3 SO4 (OH)4 ZnO, Zn5 (OH)6 (CO3)2, ZnCO3 Rarer species AlOOH, Alx (OH)Y (SO4)Z, AlCl(OH)2. 4H2O Fex (OH)Y Clz, FeCO3 Cu2 Cl (OH)3, Cu2 CO3 (OH)2, Cu2 NO3 (OH)3 Zn(OH)2, ZnSO4, Zn5 Cl2 (OH)8. H2O

2.2. Corrosion rate Expressions Mostly the rates of corrosion of metals are expressed as mpy or mmpy. The relative scale for corrosion of metal is given as [28,3]. Safe: Less than 5mpy or 0.125 mmpy Moderate: 5 mpy to 50 mpy or 0.125 mmpy to 1.25mmpy. Severe: Greater than 50 mpy or 1.25 mmpy. The rate of corrosion of metal is usually measured either by gravimetric method or by electrochemical methods. The conversion factors for the two methods are. Gravimetric method: 87.6 x weight loss (mg) Corrosion rate (mmpy ) = -------------------------------Area (cm2) x time (hrs) x Density Electrochemical method: Eq.wt Corrosion rate (mmpy) = 3.2 x 1corr (mA/cm2) x -----Density Some times the corrosion rate is also given 1.44 = mmd x ---------- = mpy Density where mdd is mg per squre decimeter per day, mpy is mils per year Also the formula for calculating the corrosion rate is given as: 534W ---------DAT where W=weight loss, mg; D=density of specimen, g/cm3; A= Area of specimen. Sq.in and T= exposure time, hr mpy =

3. ATMOSPHERIC CORROSION IN VARIOUS MATERIALS


Carbon steel (hot and cold rolled) is the most widely used metal for outdoor applications although large quantities of galvanized steel, stainless steel 304, aluminum, brass and copper are also used. Metals customarily used for outdoor installations are discussed. 3.1. Carbon Steel (hot and cold rolled) Outstanding in its drawability and weldability, hot rolled carbon steel is extensively used in automobile frames, wheels, special vehicles, building structures, bridges, general structures and ships. Cold rolled carbon steel is excellent in low temperature toughness, hydrogen-induced crack resistance, fracture resistance, high weldability and formability. This steel is used in cold rolled products such as CR, GI, colour plates, structural pipes, general pipes, special pipes, pipes for machines, high pressure gas cylinders and oil well pipes [29]. Its corrosion studies are of immense interest

Any corroding surface has a complex charge distribution, producing in the adjacent electrolyte a microscopic layer with chemical and physical properties that differ from those of the nominal electrolyte. This electrodic regime influences the overall reaction kinetics in atmospheric corrosion processes. In the solid regime, the detailed mechanistic steps (sequences) in the dissolution of the solid and their kinetic characteristics are relevant.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

due to its wide spread use. The atmospheric corrosion of carbon steel has always been of prime interest for electrochemist, and corrosion engineers because it is one of the most widely used materials. In polluted atmospheres, chlorides and SO2 are the common pollutants influencing metallic corrosion. Though chlorides come from natural airborne salinity, they are considered to be a significant pollutant as a consequence of their strong action on metals during atmospheric exposure. Relationships between chloride concentration in corrosion products, or atmospheric salinity, and corrosion rates have been reported by Corvo and Morcillo et al., [30-31]. The concentration of SO2 in the atmosphere also plays an important role in determining the magnitude of atmospheric steel corrosion. The adsorbed SO2 is postulated to acidify the moisture layer at the metal interface and produce FeSO4 which undergoes hydrolysis reactions to form oxyferric hydroxides and probably regenerate H2SO4 again and further attack carbon steel. It has also been suggested [32] that once some FeSO4 and rust have been formed, the conditions become favourable to an electrochemical cycle, which operates much faster than the acid regeneration cycle. However, the existence of competitive adsorption processes between chloride and sulphur compounds in atmospheres has also been reported by several authors [31, 33-34, 8- 9]. Other pollutants, such as NOx and O3, as well as climatological and geographical parameters can also affect the atmospheric corrosion of carbon steel [9,11,35]. Organic acids, such as acetic and formic acids, also play an important role in the atmospheric corrosion of carbon steel, even when in small concentrations. The origin of these organic acids is mainly wood, plastics and paints and they cause the degradation of carbon steel near by. This is a common occurrence, for example, in products that are stocked or packed temporarily in places where those substances exist. The presence of acetic acid and formic acid has been also detected in the rain by Galloway and Likens, Graedel et al.[36-37], where they increase the acidity. Atmospheres containing 0.5 and 10 ppm acetic acid at a relative humidity (RH) of 100 % will show corrosion behavior on carbon steel [38]. Particulate matter present in the atmosphere also plays a vital role in undermining materials resistance to atmospheric corrosion. Aggressive anions such as Cl- and SO42- are renowned culprits for inducing localized attack. Sea-salt is a further important atmospheric contaminant, especially for the corrosion of carbon steel structures. The primary sources of seasalt in the atmosphere are the oceans. Normal sea wind may carry an average of 10 to 100 lb (4.5-45 kg) of sea-salt per cubic mile (4.17 km3) of air. Brierly [39] showed that sea-salt fallout may range from an extremely high level of 3000-4000 lb acre-1 y-1 (0.30.45 kg m-2 y-1) on oceanic islands and coastal areas to 3-5 lb acre-1 y-1 (3.4 x 10-4-5.6 x 10-4 kg m-2 y-1) in arid areas. Evans[40] demonstrated that the presence of

hygroscopic magnesium chloride in sea-salt or sea mist enables corrosion to take place on carbon steel at much lower relative humidity than if only sodium chloride is present. Ericsson [41] showed that sodium chloride particles on a carbon steel surface can cause corrosion at relative humidities which have been considered too low to start SO2 induced corrosion. He reported that the synergistic effect of sodium chloride and SO2 at 90% RH increased the corrosion rate of carbon steel by about 14 times than caused by sodium chloride alone. Among the climatic factors, time of wetness (TOW), temperature and rainfall are widely reported by Evans, Brown and Masters, Morcillo et al. and Feliu et al., [34,9,35,42]. Less information is available about the influence of wind [34,36] or the height of exposure sites[43]. Among the factors affecting the type and amount of atmospheric corrosion products, the main role is played by the reactivity of carbon steel. This characteristic depends on several properties such as their chemical composition (which depends on manufacturing procedures and finishing treatments) and the design and types of structures and joints. The corrosion rate also depends on atmospheric aggressiveness, which is a function of meteorological and pollution parameters. The most common characteristic of metallic atmospheric corrosion is the localized character of its nucleation. Preferential nucleation sites depend on the metal structure and are associated to the presence of different phases or environmental pollutants on discrete areas of the metallic surface. In the case of carbon steel, nucleation starts with the formation of small protuberances of corrosion products at isolated points on the metallic surface, followed by the formation of a large number of corrosion product nuclei that can cover the entire surface after relatively short exposure times. The corrosion rate after 1 year of exposure supplies critical information about the interaction between the metallic surface and environmental parameters. The wide range of values obtained is a consequence of the exposure of the base metal in atmospheres presenting different combinations of corrosive agents, depending on the location of the site and the time of year. The weight losses of metals after longer exposure times provide information about the protective character of the carbon steel corrosion products layers (SCPLs), which once sufficiently developed, attenuate the effects of meteorological and pollution variables. The barrier effect of SCPLs depends on their thickness, uniformity, porosity, solubility, adhesion and other characteristics, and can significantly affect the carbon steel corrosion rate and attack morphology through different mechanisms, as has been shown for other metals. The lowest atmospheric corrosion rates and attack nucleation densities can be found in rural atmospheres, where corrosion products also usually present

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

S. Syed

relatively smaller structures. However, as atmospheric corrosion nucleation depends on the time of wetness (TOW) of metallic surfaces, soil pollution and background atmospheric pollution, corrosion rates in rural atmospheres present a relatively wide range of values for each metal [44]. 3.2. Galvanized Steel Galvanized steel is used in telecommunication industry, power transmission lines, thermal power plant, automotive industry, roofing, siding and fencing[45-47]. The corrosion behaviour of galvanized steel in a wide variety of environments has been thoroughly investigated by Mardar and Goodwin [4849] . In areas where SO2 is present in any appreciable quantity galvanized surface will be attacked. Particulate species in the atmospheres can accelerate corrosion of galvanized steel by increasing the conductivity of the surface layer after dissolution of soluble ions from the particulate. It has been reported on the effect of particles of (NH4)2 SO4 on the corrosion of galvanized steel [50]. Atmospheric corrosion of galvanized steel structures begins through an electrochemical process when the Zn surface becomes wet with rain, mist, or dew [51-52]. A film of basic zinc carbonate, 2ZnCO33Zn (OH)2,forms on galvanized steel structures in long exposures, defined as exposures exceeding one year [53-54] . This film tends to inhibit further Zn corrosion; however, there are environmental conditions in which film removal processes compete with film formation. Haynie [55-56], described Zn corrosion in steadystate, long-exposure conditions as linear functions of these processes. C = A/B + Bt where C = total Zn corrosion, m t = exposure time, y; B = dissolution rate of the film, m Zn/y; and A = diffusivity of corrosive species through the film, m2/y. The total Zn corrosion (C) is the sum of the Zn contained in the film (A/B) and the Zn dissolved from the film by precipitation (B). Diffusivity (A) can be affected by atmospheric species combining with the film and forming various mineral phases. The dissolution rate (B) is determined by the delivery rate of acidic pollutants to the film and by the mineral phases present. Dry and wet deposition delivery mechanisms provide a means for describing the rate controlling mass transfer processes affecting corrosion of the Zn coating. Spence, et al. [46] developed a model for predicting the corrosion of galvanized steel structures based on two competing mechanisms: the formation and dissolution of the basic zinc carbonate film that forms on zinc surfaces. The model consists of a diffusivity term that describes film growth and a dissolution term that describes the rate of film removal. Dissolution

becomes the rate-determining process for predicting the long-term corrosion behaviour of galvanized steel structures. Components of the dissolution term were evaluated with data collected from field exposure experiments that were designed to separate the effects of wet and dry acidic deposition from the effects of normal weathering of galvanized steel specimens. The models dissolution term predicted the long-term corrosion of galvanized steel with reasonable accuracy. For further evaluation, the dissolution model was applied to historical, long-term corrosion data of galvanized steel products, taking into account their sizes and shapes. The field data used in this evaluation was consistent with corrosion rates predicted by the model, within the limits of uncertainty of the environmental data. Thus, the model can be used with reasonable confidence to predict corrosion behavior of different structures if environmental conditions can be properly described. 3.3. Stainless Steel 304 Stainless steels 304 was first introduced into commercial use about 70 years ago [57]. It is used in dairy equipment, fruit juice industry, food processing applications such as in mills, bakeries , slaughter and packing houses, dye tanks, pipeline buckets, dippers, railroad cars, fermentation vats and hauling equipment. A considerable amount of information has accumulated about their atmospheric corrosion behavior. 3.4. Aluminum Aluminum is one of the most abundant elements in nature. Its low density, high elastic modulus, thermal and electrical conductivity, its corrosion resistance, and its capacity to form alloys with many elements makes it one of the most useful materials of construction [58]. Aluminum and its alloys undergo black to grey staining when exposed to humid atmosphere due to condensation of moisture or rain on the surfaces [59-61]. The degree of staining does not depend on the composition of water and the staining rate is mainly controlled by the rate of diffusion of oxygen into the thin film of water condensed [62]. The stained area is mostly bayerite (Al2O3.3H2O) with the thickness of 2500-5000 [59]. Of all the aluminum alloys aluminum-magnesium alloys are highly susceptible to water staining due to formation of magnesium oxide film [63]. The corrosion of aluminum in the atmosphere has mainly been investigated through field studies [64-67]. Few laboratory investigations in controlled environments have been published. Besides a strong humidity dependence it is generally agreed that deposition of SO2 and chlorides and the pH in rain are major factors that determine the corrosion rate of aluminum. In the presence of SO2, oxidizing agents such as O3 and H2O2 may also play a role in the atmospheric corrosion of aluminum [64].

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

Compared to other metals, aluminum corrodes rather slowly under atmospheric conditions, because it forms an insulating amorphous oxide film of low solubility in air and aqueous solutions over the pH range from 4 to 8.6. However, enhanced corrosion occurs in marine environments and in some urban areas [68]. Common corrosion products found under these conditions are basic aluminum sulfates and amorphous aluminum sulfate hydrate [68-69]. Indoors, sulfate is the most abundant anion found on aluminum surfaces [70]. The presence of SO42- on these surfaces may be due to SO2 induced corrosion or sulfates associated with particle deposition. The accumulation of inorganic ionic substances is primarily due to particle deposition [71]. Ammonium and sulfate ions are the most abundant ions in fine dust particles commonly found in urban environments and may, therefore, play a dominant role in the corrosion process. This is especially true for atmospheres containing solid, ionic contaminants like ammonium sulfate. Salt spray tests, which are a common test method for industrial applications, are not suitable for studying the effect of dry deposits, since the initial period, when the particles may start to absorb water vapor from the atmosphere, cannot be simulated. Patterson and Wilkinson [73] investigated the effect of NaCl and NH4Cl particles on the corrosion of aluminum at 293 K and 80% relative humidity (RH). The weight increase due to NaCl particles was linear with time and was small compared to the weight increase due to NH4Cl particles. With NH4Cl, Al showed an initial phase of slow weight increase, and after 5 days, a sudden rapid weight gain. Also the appearance of the specimen changed after 5 days from the initial uniform deposit to a granular uneven product. The corrosion products were not identified. In a later study, Wilkinson and Patterson [74] investigated the effect of RH (70 to 90%) on the corrosion at 293 K using NH4Cl, the weight gain of Al after 90 days was seven times greater at 80% RH than at 70%. The corrosion products observed at 80% RH were much bulkier than at 70% RH which appeared after a few days. At 90% RH the bulky corrosion products appeared on the first day and continued growing. Sanyal and Bhadwar [75] investigated the effect of NaCl, Na2SO4, NH4Cl, and (NH4)2 SO4 particles on the atmospheric corrosion of aluminum at 313 K. Corrosion rates increased with increasing humidity, temperature and period of exposure, and were generally much higher than under immersed conditions in the corresponding electrolytes. The estimated minimum NO2 concentration inducing aluminum corrosion is believed to be 30 g/m3 [76]. For a urban-industrial Cuban atmosphere it was obtained diary NO2 concentrations over 31 g/m3 only in a 10% of the measurements[77]. Rural and coastal atmospheres do not report values over 30 g/m3. Therefore, the influence of this pollutant on the atmospheric corrosion may be negligible.

3.5. Brass Brasses are widely used engineering materials and have found their applications in electrical, air conditioning, marine, construction and fabrication industries. 70/30 brass has phase structure and is resistant to many organic and inorganic reagents. Anodic reactions for the corrosion of brass is the dissolution of both Cu and Zn, but in some corrodents there is a preferential dissolution of zinc leading to dezincification of the alloy and results in the loss of tensile strength of the alloy. In acidic medium the cathodic reactions are the discharge of hydrogen ion to form nascent hydrogen and reduction of dissolved oxygen to form water which may further give nitrous acid which may combine with H+ to form nitrosonium ion. Reduction of nitrosonium ion may result in the formation of nitrous acid. The formation of nitrosyl ion and nitrous acid [78] represent the limiting stage in the overall cathodic reaction of the reduction of nitrous acid. Nitrous acid is an active species for the corrosion of brass and its concentration controls the dissolution of 70/30 brass in HNO3 acid [79-80]. 3.6. Copper Copper is on one hand, a natural component in most ecosystems and on the other, a metal that always has found many applications in old and modern societies. Todays society relies on the electrical conductivity of a small number of metals to convey power and signal information to electrical and electronic equipment, architectural and artistic. One of the most important metals in this respect is copper. Due to the widespread use of copper in many applications besides conductors, e.g., as roofing material, for decorative structures, and statues, there is a vast knowledge of the long time behavior of the atmospheric corrosion of unsheltered copper outdoors. For instance, copper patina formation, its stages, and most important constituents, together with mechanistic deliberations, have been thoroughly examined [81-83]. The general sequence of atmospheric corrosion of copper is well known. Initially oxygen and water react with a fresh copper surface forming a sequential structure consisting of Cu2O/CuO/[Cu(OH)2 or CuO. xH2O], the main component being Cu2O, cuprite [82] .This is later followed by reaction with pollutants present as gases (e.g., SO2, NO2, O3, Cl2, HCl, and H2S) as ionic constituents of aerosol particles or as ions in precipitation. Eventually a patina of several different compounds forms on top of the initially formed cuprite layer. Important copper compounds found as patina constituents are Cu2Cl(OH)3, atacamite, Cu4SO4 (OH)6. H2O, posnjakite, and Cu4SO4 (OH)6, brochantitie, in rural and urban areas. In urban areas Cu2SO4 (OH)4, antlerite, and Cu2CO3(OH)2, malachite, are also found [82]. When copper is used for electronic signal transmission, very small defects on contact surfaces due to atmospheric corrosion can cause partial or complete failure of the

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

S. Syed

device. Even a relatively benign environment or short exposures in more polluted environments can cause damage. Thus, it is very important to investigate minor attacks preceding actual patina formation. In the literature, information on corrosion processes on copper can be found in studies conducted in the laboratory, in indoor field exposures, and in sheltered and unsheltered outdoor exposures. Since the pioneering work by Vernon [84], the effect of different combinations of gaseous pollutants has been investigated in several laboratory studies [85-89]. Vernon also realized the importance of aerosol particles as corrosion accelerators [84, 90]. Current concentrations of aerosol particles and their impact on reliability of electronics have been studied by Sinclair, Sinclair and Psotakelty and Sinclair et al. [91-93,72] in several field exposures. These investigations later inspired multianalytical laboratory studies of the influence of ammonium sulfate on the corrosion of copper [94]. Field studies combining pollutant monitoring and copper corrosion measurements have been pursued both indoors [85,95-96] and in sheltered outdoor environments [97-99]. For atmospheric corrosion to occur, the presence of water on the metal surface is also essential [84-85, 94]. The amount of water sorbed is dependent on relative humidity and on surface deposits [100].

4. INFLUENCE OF EXPOSURE PARAMETERS


4.1. Humidity Atmospheric air is a mixture of dry air and water vapor. In industrial and sea shore places, gases such as SO2, Cl2 and H2S and particulates of NaCl and other salts are present. The air humidity is characterized by the indices RH, absolute humidity, moisture content and specific air humidity [4]. 4.2. Critical Relative Humidity The primary value of the critical relative humidity denotes that humidity below which no corrosion of the metal in question takes place. However, it is important to know whether this refers to a clean metal surface or one covered with corrosion products. In the latter case a secondary critical humidity is usually found at which the rate of corrosion increases markedly [84]. This is attributed to the hygroscopic nature of the corrosion product. In the case of iron and steel it appears that there may even be a tertiary critical humidity [101]. Thus at about 60% RH rusting commences at a very slow rate (primary value) [102] at 75-80% RH there is a sharp increase in corrosion rate probably attributable to capillary condensation of moisture within the rust[84,103]. At 90% RH there is a further increase in rusting rate [101], corresponding to the vapour pressure of saturated ferrous sulphate solution[104], ferrous

sulphate being identifiable in rust as crystalline agglomerates[105]. The primary critical RH for uncorroded metal surfaces seems to be virtually the same for all metals, but the secondary values vary quite widely. It has been found, that at high relative humidities, aluminum and iron show no SO2 + NO2 synergism [106] , and that for steel in negligible [107-108]. It was reported that a thick layer of water on the metal surface seems to act as a sink for SO2, but as a barrier for NO2 [108]. For metals with a protecting oxide film, NO2 may even act as an inhibitor; otherwise, there seems to be synergistic effects [106]. It has been showed by several authors that the SO2 + NO2 synergism on copper corrosion is only active at high relative humidity (90%) [109], and steel corrosion at low relative humidity [110-111, 108,106]. Kucera et al. [128] reported that the synergistic effect of SO2 + O3 can be both stronger than SO2 + NO2, as for copper, and weaker, as for nickel. In other works no synergistic effects of simultaneous interaction of SO2 and NO2 with either nickel or copper have been observed [112]. This dry deposition is in most cases dominating and SO2 exerts the strongest corrosive effect [106]. The role of NO2 has not yet been clarified and its strong synergistic effect with SO2, shown for many materials in different laboratory studies, has not been observed in the field exposure and may be due to the strong correlation between SO2, NO2 and O3 concentrations [106] . 4.3. Specific AtmospHeric Corrodents (Pollutants) The electrolyte film on the surface will contain various materials deposited from the atmosphere or originating from the corroding metal. The composition of the electrolyte is often the factor which determines the rate of corrosion. Joel and Karl [186] reports the combined effects of fire, corrosive smoke, and particulate as fire corrosivity. While the effects of fire corrosivity are well known, little quantitative information is available concerning the mechanisms involved and the degree to which materials, particularly metals, are susceptible. In this investigation, the deleterious effects of combustion products generated during fires on normal construction materials are documented. Table 2 summarizes the depth of corrosion data corresponding to post-test exposure in a humidity chamber. Karlsson, et al. [147] reported the serious corrosive damage to carbon steel can be caused by fume gas containing 50-500 ppm SO2, 1-2 g m_3 of alkali chloride dust and 8-15% oxygen at temperatures of 3500oC. It is confirmed by experiments, that a liquid phase may exist in a system containing Fe2O3, alkali chlorides, SO2 and O2, and that the melting point is as low as 310oC. The liquid phase is very corrosive and attacks stainless steel 304 and low alloyed steel at approximately the same rate.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

Table 2. Depth of corrosion data corresponding to post-test exposure in a humidity chamber following combustion product exposure [186]. Material UL Identification Post-test exposure 27oC (80oF) 75% RH corrosion value (micro inch) 3 4 1 day 2 days days days 35.5 31.0 33.0 0.0 53.5 39.0 53.5 0.0 70.5 46.5 66.5 0.0 79.0 52.0 73.0 0.0

4.3.1. Sulfur dioxide Sulfur dioxide, a product of the combustion of sulfurcontaining fossil fuels, plays an important role in atmospheric corrosion in urban and industrial atmospheres. It is adsorbed on metal surfaces, has a high solubility in water, and tends to form sulfuric acid in the presence of surface moisture films. Sulfate ions are formed in the surface moisture layer by the oxidation of sulfur dioxide in accordance with Eq. SO2 + O2 + 2e- SO42The required electrons are thought to originate from the anodic dissolution reaction and from the oxidation of ferrous to ferric ions. It is the formation of sulfate ions that is considered to be the main corrosion accelerating effect from sulfur dioxide. For iron and steel, the presence of these sulfate ions ultimately leads to the formation of iron sulfate (FeSO4). Iron sulfate is known to be a corrosion product component in industrial atmospheres and is mainly found in layers at the metal surface. The iron sulfate is hydrolyzed by the reaction expressed by Eq. FeSO4 +2H2O FeOOH + SO42- + 3H+ + eThe corrosion-stimulating sulfate ions are liberated by this reaction, leading to an auto-catalytic type of attack on iron [25,26,113]. The acidification of the electrolyte could arguably also lead to accelerated corrosion rates, but this effect is likely to be of secondary importance because of the buffering effects of hydroxide and oxide corrosion products. In nonferrous materials such as zinc, sulfate ions also stimulate corrosion, but the auto-catalytic corrosion mechanism is not easily established. Corroding zinc tends to be covered by stable zinc oxides and hydroxides, and this protective covering is only gradually destroyed at its interface with the atmosphere. In moderately corrosive atmospheres, sulfates present in zinc corrosion products tend to be bound relatively stronger, with limited water solubility. At very high levels of sulfur dioxide, dissolution of protective layers and the formation of more soluble corrosion products is associated with higher corrosion rates. It is well known that SO2 pollutant substantially enhances the corrosion rates of metals exposed in the atmosphere. Rozenfeld [114] has suggested that, because of its greater solubility (SO2 is about 2600 times more soluble than oxygen), it might be reduced at cathodic sites more rapidly than oxygen, consequently increasing anodic dissolution rates. In solution, electro-chemical reduction of SO2-3 competes with its oxidative conversion to SO2-4. However, Seinfeld, [115] states that, in the absence of catalysts, solution phase oxidation of SO2-3 by dissolved oxygen is slow. Under these circumstances SO2 may persist for a sufficient length of time to act as a cathodic depolarizer in the manner as suggested by Rozenfeld.

Stainless steel (304) Stainless steel (304) Stainless steel (304) Stainless steel (304)

SS1 SS2 SS3 SS4

Allam, et al. [187] have used the energy dispersive X-ray mico-analysis, X-ray diffraction and fluorescence, Auger, X-ray photo-electron spectroscopy and Fourier transform infrared spectroscopy to characterize corrosion products on carbon steel after atmospheric exposure for periods up to 12 months in an industrial environment near the west coast of the Arabian Gulf. The results indicate that atmospheric corrosion starts by the formation of small blisters at discrete locations on the metal surface, presumably the anodic sites. The blister covers are very rich in iron chlorides and contain iron oxyhydroxides, oxides, sulphates and possibly hydroxide. The formation of iron chlorides as the primary corrosion product is only limited to the early stages of blister formation due to the aggressive nature of chloride ions. Chloride formation during later stages may be partially impaired since it requires the inward transport of fresh chloride ions through the then thick rust layer. In contrast, the formation of iron sulphates at the rust-metal interface continues by the acid regeneration mechanism (which leads to the electrochemical mechanism); therefore it is less dependent on the supply of fresh sulphate ions from the surface electrolyte through the growing rust layer. Chawla and Payer [188] studied the early stage of corrosion of copper by moist air containing 0.5% sulfur dioxide (75% relative humidity at 25oC). Scanning electron microscopy, Auger electron spectroscopy, and transmission electron microscopy have been used to examine the surface topography, surface chemistry and microstructure of copper foils before and after exposure to the corrosive atmosphere. The analysis shows that in the initial stage of corrosion a mixture of copper oxide and copper sulfide forms on the surface. Reduction of sulfur dioxide to sulfide on the metal surface indicates that sulfur dioxide is a cathodic depolarizer in the early stages of corrosion. The primary contaminants in the air that lead to atmospheric corrosion are SOx, NOx, chlorides, carbon dioxide, hydrogen per oxide, hydrogen chloride, ozone, oxygen, hydrogen sulphide, organic acids and saline particles etc.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

S. Syed

4.3.2. Nitrogen compounds Nitrogen oxide emissions originate from combustion processes other than those emitting SOx. Road traffic and energy production are the primary sources. Most of the nitrogen oxides are emitted as NO in combustion processes. In the atmosphere oxidation to NO2 takes place successfully according to 2NO + O2 2NO2 As the pollutant moves further from the source it is further oxidized by the influence of ozone: NO + O3 NO2 + O2 Near the emission source nitrogen dioxide is considered to be the primary pollutant. The NO2/NO ratio in the atmosphere varies with time and distance from the source. Allowed enough time the NOx may be further oxidized according to the reaction 2NO + H2O + 3/2 O2 2HNO3 Since this reaction occurs at a very slow rate, the amounts of HNO3 and nitrates in the vicinity of the source are very low. In urban atmosphere NO2 levels show a less promising trend. This has led to some interest in the corrosive effect of NO2 on copper. Thus Eriksson and Johansson [116] exposed copper to humid air containing NO2. Simon, et al., [117] studied the corrosion products formed on copper exposed to humid air containing NO2 identified cuprite (Cu2O) and basic copper nitrate (Cu2(OH)3 NO3) on the surface using XPS. The general conclusion of these studies is that NO2 in the ppm range has very slight corrosive effects on copper. 4.3.3. Chlorides Atmospheric salinity distinctly increases atmospheric corrosion rates. Apart from the enhanced surface electrolyte formation by hygroscopic salts such as NaCl and MgCl2, direct participation of chloride ions in the electrochemical corrosion reactions is also likely. In ferrous metals, chloride anions are known to compete with hydroxyl ions to combine with ferrous cations produced in the anodic reaction. In the case of hydroxyl ions, stable passivating species tend to be produced. In contrast, iron-chloride complexes tend to be unstable (soluble), resulting in further stimulation of corrosive attack. On this basis, metals such as zinc and copper, whose chloride salts tend to be less soluble than those of iron, should be less prone to chloride-induced corrosion damage [25] and this was consistent with practical experience. Chloride ions is the most common and important atmospheric corrosive agent, as has been reported by different authors all over the world [42,118,107,119]. However, different works concerning the influence of chloride ion on metallic corrosion have been reported[42,107].

4.3.4. Carbon dioxide The concentration of carbon dioxide in the atmosphere is about 350 ppm [120]. The effect of CO2 on the atmospheric corrosion of zinc was investigated by falk, et al. and lindstrom, et al. [121-122]. They reported that ambient concentrations of CO2 inhibit the NaClinduced corrosion of zinc. This effect is important for understanding zinc corrosion in cases where the supply of CO2 is limited, eg., in crevices and under paint films. In the case of zinc and copper, carbonatecontaining corrosion products are often reported from the field [68,123]. In contrast, aluminum carbonate has not been identified as a corrosion product in the atmosphere [64]. A laboratory study of the effect of CO2 on the atmospheric corrosion of aluminum is reported. The samples were exposed to pure air with 95% relative humidity and the concentration of CO2 was <1 and 350 ppm, respectively. Atmospheric corrosion of aluminum is about 10-20 times faster in CO2 free humid air compared to air containing ambient levels of CO2. In the absence of CO2, bayerite, Al (OH)3, forms. Only minute amounts of carbonate were found on the surface after exposure to CO2 containing air [124]. 4.3.5. Hydrogen peroxide In recent years H2O2 has been found to be a common atmospheric constituent. Typical midsummer concentrations can be as high as 10-30 ppb (gas phase[125] and 10-100 m precipitation[126]). Such concentrations are often greater than those of SO2 [127] or SO2-4 [128].The importance of the presence of hydrogen peroxide is that metallurgical research has demonstrated the propensity for H2O2 and its related radicals to be reactive toward iron and iron alloys. The general result of such interactions is that contact of the metal surface with a solution containing H2O2 results in decomposition of the H2O2 to produce HOX. radicals and the subsequent formation of passive layers on steel [129,130]. Once a passive layer has formed, the decomposition rate of the H2O2 becomes very low [131]. Mayne and Burkill [132] have suggested that the passive layer is formed by interaction with peroxy compounds, and Marsh, et al. [133] have demonstrated such an effect by experiments in which the HOx. radicals are generated by -radiation. The later researchers found that the oxidizing radiolysis products can have an inhibitive effect on the initiation of localized corrosion at potentials less than approximately +0.5 V vs. SHE, as are present in aqueous solutions exposed to the atmosphere [134]. Their work was done in connection with the corrosion of iron in industrial applications, but, as seen above, precipitation and surface deposition from the atmosphere readily supply H2O2 to outdoor steel surfaces. In addition to aiding in the formation of passive oxyhydroxide layers, H2O2 can interfere with the reduction by bisulfite ion of the relatively unreactive

10

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

iron (III) ions to the more reactive iron (II) form, since the reaction. HSO-3 + H2O2 HSO-4 + H2O Thus, in the summer months when atmospheric chemistry is vigorous and H2O2 concentrations are high[135] it is generally true that in water films on steel surfaces (H2O2) > [Fe3+], rust and dissolution of the steel is inhibited. In the winter months, or in very high sulfur environments, [Fe3+] > [H2O2], the reduction of iron (III) to iron (II) is promoted while surface passivation is inhibited, and steel dissolution proceeds rapidly. In recent years the sulfur emissions to the atmosphere in some parts of the world, particularly North America, has decreased or remained relatively constant, while increased emissions of oxides of nitrogen and hydrocarbons have resulted in increase in atmospheric H2O2. The consequence of these changes would be an expected decrease in the corrosion rate of steel, especially in the summer; unfortunately, the historical data needed to examine this conjecture is not available. 4.3.6. Hydrogen chloride Very little work has been reported on the effect of HCl on the degradation of materials in the environment. This is probably because HCl, which is present outdoors in markedly reduced concentrations when compared with SO2, has not been considered to contribute to significant degradation of materials. The major natural source of HCl is from volcanic fumaroles; however, atmospheric transformations of chlorinated hydrocarbons also contribute to outdoor levels, [136] as do emissions from waste incineration especially of PVC and chlorinated solvents[137]. Although the concentration of chloride species in rainwater has been measured frequently, it is generally difficult to distinguish between chloride derived from combustion with that derived from sea-salt aerosol or, in specific locations, road dering salt. The first major study of atmospheric degradation of metals by HCl was carried out by Feitnecht [138] who exposed zinc, iron and copper to HCl vapours at varying humidities between 50% and 95% RH. He found that HCl reacted with metals only when a critical relative humidity was exceeded, the value of which was approximately that of the vapor pressure over a saturated solution of the metal chloride formed during corrosion. Barton and Bartonova [139] carried out a much more extensive investigation of the corrosive effect of HCl gas at concentrations between 7 and 10 ppm on zinc, mild steel and copper at temperatures between 20 and 50oC and at relative humiditys of 70 and 95%. Two distinct stages were seen in the behavior. The first stage was characterized by a non-linear increase in weight loss with time, which Barton and Bartonova termed the indication period for steady-state corrosion. The second stage, after about 16 days exposure,

showed steady-state corrosion with a linear increase in weight-loss with time. The primary corrosion products found on iron were FeO(OH), Fe3O4 and FeCl2, whilst those found on zinc were 4Zn (OH)2. ZnCl2, Zn (OH)2 and ZnO. The amount of chloride in the corrosion product tended to decrease slowly with time, during the initial period of exposure. After the steady state corrosion region had been reached, the composition of the corrosion product remained unchanged. The corrosion rate was measured at different temperatures in the steady state region. For zinc, the corrosion rate decreased as the temperature increased; for iron, the corrosion rate increased with temperatures up to 40oC, but decreased at 50oC. Barton and Bartonova proposed the following mechanisms to explain their observations. Iron: Initially, a uniform film of FeO (OH) and FeCl2 forms, presumably by reaction with the air-formed surface film on iron. Subsequently, the FeCl2 reacts as follows. FeCl2 + H2O + O2 = FeO (OH) The metal then reacts with HCl released: Fe + 2HCl + O2 = FeCl2 + H2O (2) The steady-state can be represented by a combination of these two reactions. The rate of the reactions did not appear to depend on the diffusion of HCl to the surface since the corrosion rate was similar in flowing and stationery atmospheres. The implication is that the corrosion rate is dependent on chemical reaction rate. Zinc: During the initial period, the following reactions may occur: Zn + 2HCl + O2 = ZnCl2 + H2O (3) (1)

ZnCl2 + 4H2O + 2O2 + 4Zn = ZnCl2. 4Zn (OH)2 (4) In the steady state, reactions (3) and (4) are significant; a reaction involving the destruction of ZnCl2. 4Zn (OH)2 is also important : ZnCl2.4Zn(OH)2+2HCl =2ZnCl2+3Zn (OH)2+2H2O(5) The steady-state is represented by a combination of reactions (3), (4) and (5). The kinetics of corrosion is controlled by the transfer of HCl to the corrosion product atmosphere interface, its adsorption and the subsequent production of soluble ZnCl2. The corrosion rate also depends on the hydroxide / chloride ratio in the corrosion product as the hydroxides are more protective than the chlorides. No literature is available on the combined corrosive effects of gaseous SO2 and HCl pollutant. Corrosion of iron and zinc in HCl has been carried out under a range of conditions simulating natural atmospheres. The corrosion rate of zinc did not significantly increase upon exposure to HCl at presentation rates typical of the highest found in an urban area (2.5 x 10-6 mg cm-2 s-1). This is explained by the formation of protective basic zinc chloride by

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

11

S. Syed

reaction with pre-existing zinc hydroxide. At higher levels of pollutant (representative of episodic industrial atmospheres) reaction of HCl with zinc hydroxide proceeds further to form soluble zinc chloride. Consequently, the protective ability of the corrosion product is lost. Under these conditions, corrosion rate of zinc is controlled by the availability of HCl at the metal surface. In contrast, the corrosion rate of mild steel samples exposed to HCl presentation rates typical of the highest encountered in an urban atmosphere was 18 times than found in an unpolluted atmosphere. Increasing the HCl presentation rate further did not result in a significant increase in corrosion rate. These results are explained by the reaction of HCl at the metal surface and the subsequent formation of FeCl2, which is oxidized to FeO(OH), liberating HCl which can initiate further corrosion. The reaction sequence forms a cycle and is, therefore, apparently independent of incoming HCl [140]. 4.3.7. Ozone In Urban environment where O3 sometimes reaches extreme values, sulphate and oxide formation were accelerated more than expected [99]. A few laboratory exposures have recognized O3 as a potential agent for atmospheric corrosion of copper. The general oxidative power of O3, increases the oxidation of H2S to sulphur and SO2 to SO3 [141]. Later Graedel et al. [142] unambiguously showed that O3 enhances the atmospheric sulphadation of copper. Laboratory experiments involving SO2 and O3 as gaseous pollutants are scarce, Svensson and Johansson[143] showed that O3 increases the corrosion attack on galvanized steel by oxidizing SO2 much more efficiently than NO2. Similar evidence was provided for copper by Eriksson [144], who found a basic copper sulphate, Cu2..5 SO4 (OH)3, 2H2O, after one day of exposure to SO2+O3 at 90% relative humidity. Copper samples are exposed at 75% relative humidity to explore the possible influence of ozone on the atmospheric corrosion rate of copper, various combinations of the gaseous pollutants such as sulphur dioxide, nitrogen dioxide and ozone were added. Ozone promotes the oxidation of SO2 to sulphate more efficiently than NO2 does. A synergism between SO2 and O3 is suggested. This synergism includes both the oxidation of Sulphur dioxide by ozone and the capability of ozone to form oxides, hydroxides or other oxygen containing reaction products in the presence of smaller amounts of SO2. The synergistic effect possibly can explain the unexpectedly high corrosion rates of copper found at rural sites. The rural sites are characterized by low SO2 and NO2 concentrations, and by high ozone concentrations [145]. It has been reported that the presence of ozone in the atmosphere may lead to an increase in the sulfur dioxide deposition rate. While the accelerating effect of ozone on zinc corrosion appears to be very limited,

both aluminum and copper have been noted to undergo distinctly accelerated attack in its presence [24] . Role of ozone on the SO2 induced atmospheric corrosion of copper increases the formation rate of both Cu2O and CuSO4. x H2O all over the surface [146]. 4.3.8. Oxygen Oxygen is a natural constituent of air and is readily absorbed from the air into the water film on the metal surface, which may be considered saturated, thus promoting any oxidation reactions. Serious corrosive damage to iron alloys can be caused by 8-15% of oxygen at temperature of 3-500oC. It is confirmed by experiments [147]. 4.3.9. Hydrogen sulphide Trace amount of hydrogen sulphide is present in some contaminated atmospheres. Hydrogen sulphide is known to be extremely corrosive to most metals and alloys. The tarnishing of copper in a test atmosphere consisting of air and hydrogen sulphide has been described. Initially, linear growth is observed and the tarnished film is made up of cuprous oxide. In a later stage the film growth becomes parabolic and cuprous sulphide is formed [148]. Cuprous oxide protects copper from further attack by sulphur compounds in a dry atmosphere. In the presence of water, present as an adsorbed film, cuprous oxide reacts with hydrogen sulphide and a sandwiched reaction layer (Copper/ cuprous oxide/cuprous sulphide) is formed [149]. In humid air (no hydrogen sulphide present) thin copper films is oxidized to cuprous oxide [150,151]. According to Leest [152] and Abbott [153] when copper is exposed to H2S atmosphere the corrosion products formed was Cu2S. Lenglet, et al. [154] reports that when copper exposed to H2S atmosphere the corrosion products are Cu2S and sulphate, most likely as posjnakite, Cu4SO4 (OH)6. H2O. 4.3.10. Organic acids Organic acids, such as acetic and formic acids, also play an important role in the atmospheric corrosion of metals, even when in small concentrations. The origin of these organic acids is mainly from woods, plastics, certain paints, rubbers, resins, and other materials likely to be found alongside packaged metal items [159166] , and they cause the degradation of metals nearby. This is a common occurrence, for example, in products that are stocked or packed temporarily in places where those substances exist. The presence of acetic acid and formic acids has been detected in the rain [36,37], where they increase the acidity. The influence of organic acid vapors on copper structures after prolonged exposure in a town atmosphere was studied by Vernon [155]. The presence of these acids in outdoor atmospheres is a source of free acidity in precipitation in industrial areas [156158,128,194,96] . Acetic vapor is also present in industrial atmospheres, e.g., from vinegar in the food processing

12

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

industry and from the decomposition of raw materials in the paper industry. Organic acid anions constitute about 0.1 to 1% of the total ion concentration in the corrosion-products (patina) on copper exposed to the outdoor atmosphere for extended periods [37]. An important aspect of acetic acid vapor is that it causes metal corrosion at very low concentrations [167-170]. Atmospheres containing 0.5 and 10 ppm acetic acid at a relative humidity (RH) will show corrosion behavior on mild steel [38]. A study was made for the copper corrosion rate and corrosion products originated by the action of acetic acid vapor at 100% relative humidity. Copper plates were exposed to an acetic acid contaminated atmosphere for a period of 21 days. Five acetic vapor concentration levels were used. The copper corrosion rate was in the range of 1 to 23 mg/dm2 day. Some of the compounds identified were cuprite (Cu2O), copper acetate hydrate [Cu(CH3COO)2. 2H2O] and copper hydroxide acetate [Cu4(OH)(CH3COO)7. 2H2O]. This last compound was also characterized. The thickness of the patina layers was 4 to 8 nm for amorphous cuprite, 11 to 48 nm for cuprite, and 225 nm for copper acetate. The patina, in which the cementation process of different corrosion-product layers plays an important role, is formed by the reaction of acetic vapor with copper through porous cuprite paths [171]. 4.3.11. Saline particles These are of two main types. The first is ammonium sulphate, the effect of submicron-sized particles of (NH4)2SO4 on the corrosion of copper in air at 373 K. This study was undertaken because ammonium and sulfate ions are usually the most abundant ions in fine dust particles commonly found in urban environments. These particles can lead to accelerated corrosion of electronic materials, e.g., copper, which is used in microelectronic devices, circuit boards and connectors [172-173, 72]. It was found that copper reacts with (NH4)2 SO4 particles only if the critical relative humidity (CRH) of (NH4)2 SO4 is exceeded. The CRH is that humidity at which the salt starts to absorb significant amounts of water. At relative humidities (RH) above the CRH, copper was heavily corroded leading to a thick Cu2O-layer overgrown by crystals of basic copper sulfates. A corrosion mechanism that explains these results was proposed [94]. The corrosion mechanism of copper at 373 and 300 K in the presence of submicron (NH4)2 SO4 particle deposits has been investigated. Several in situ techniques have been used to monitor the corrosion process in real time. At and above the critical relative humidity of (NH4)2 SO4, dissolution of Cu is followed by formation of Cu2O, oxidation of Cu(I) ions to Cu(II) ions and precipitation of antlerite [Cu3(SO4)(OH)4], brochantite [Cu4(SO4)(OH)6], or posnjakite [Cu4(SO4)(OH)6. H2O]. The amount of corrosion product formed increases with the amount of

(NH4)2 SO4, particles, relative humidity (RH) and temperature [94]. The effect of (NH4)2 SO4 particles on the atmospheric corrosion of aluminum has been investigated at 300 and 373 K at various relative humidities (RH) levels. Aluminum reacts with (NH4)2 SO4 particles only at or above the critical relative humidity (CRH) at either temperature. The corrosion rate increases with increasing RH and temperature. Above the CRH of (NH4)2 SO4, droplets are formed on the particles at both temperatures, making electrochemical reactions possible. The (NH4)2 SO4 decomposes and ammonia evaporates from the droplets. At 373 K mixed ammonium-metal-sulfates are formed, followed by basic metal sulfates; oxide formation is enhanced at 373 K compared to 300 K. At 300 K no solid corrosion products containing Al are found, but it was shown that Al dissolves in the droplets. A corrosion mechanism has been proposed that explains the experimental observations, including pH and corrosion potential changes with time [174]. The second is marine salt, mainly sodium chloride but quite appreciable quantities of potassium, magnesium and calcium ions are analyzed in rainfall [175]. 4.3.12. Other atmospheric contaminant and airborne particles The corrosive effects of gaseous chlorine in the presence of moisture tend to be stronger than those of chloride salt anions because of the acidic character of the former species [25]. Airborne particles are divisible into two groups: aThe inert non-absorbent particles, usually siliceous, which can only affect corrosion by facilitating differential aeration processes at points of contact, and b- The absorbent particles such as charcoal and soot are intrinsically inert but have surfaces or infrastructures that adsorb SO2 by either co-adsorption of water vapor or condensation of water within the structure and catalyse the formation of a corrosive acid electrolyte solution. Dirt with soot assists the formation of patinae on copper and its alloys by retaining soluble corrosion products long enough for them to be converted to protective, insoluble basic salts [176]. 4.4. Temperature The effect of temperature on atmospheric corrosion rates is also quite complex. An increase in temperature will tend to stimulate corrosive attack by increasing the rate of electrochemical reactions and diffusion processes. For a constant humidity level, an increase in temperature would lead to a higher corrosion rate. Raising the temperature will, however, generally lead to a decrease in relative humidity and more rapid evaporation of surface electrolyte. When the time of wetness is reduced in this manner, the overall corrosion rate tends to diminish [27].

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

13

S. Syed

For closed air spaces, such as indoor atmospheres, it has been pointed out that the increase in relative humidity associated with a drop in temperature has an overriding effect on corrosion rate [177]. This implies that simple air conditioning that decreases the temperature without additional dehumidification will accelerate atmospheric corrosion damage. At temperatures below freezing, where the electrolyte film solidifies, electrochemical corrosion activity will drop to negligible levels. The very low atmospheric corrosion rates reported in extremely cold climates are consistent with this effect.

Table 3. Corrosion (m) and durability factors (DF) of aluminum [178]. Area
South

Station
1 2 3 4 1 2 3 4 1 2 3 4

Period 1982-83
m 0.37 0.41 0.33 0.10 0.81 0.49 0.31 0.33 DF 51 39 34 144 45 76 75 79 -

Period 1983-84
m 0.24 0.32 0.29 0.12 0.63 0.5 0.17 0.19 0.18 0.11 0.13 0.18 DF 43 41 93 52 77 125 134 66 98 82 53

North West

Central

5. RELATED PREVIOUS EXPERIENCE AND WORK DONE WORLDWIDE


Natesan, et al. [178] reported that the loss due to corrosion is often compared with that of other calamities such as earthquake and cyclone. In fact similar to earthquake and cyclone, corrosion is a natural process, only difference being that its impact is invariably indirect. In the case of earthquake, mapping of seismic zones is already in practice. In the case of cyclone also, weather prediction is available on a global level. Different countries are independently preparing their own corrosivity maps confined to the region of their interest. With the on going liberalization and globalization of the industries, there is an urgent need for preparing corrosion maps on a global level. The different organizations which are involved are to be brought together and a centralized data bank created. The data generated are presented and discussed in the light of the global data. Collection of data on the corrosivity of atmospheres in different locations has been going on in different countries. Certain interesting observations are highlighted here. Spain: Researchers from Centro Nacional de Investigaciones Metalurgicas, Ciudad Universitaria, Madrid have taken into consideration three clearly differentiated meteorological areas in Spain: the central, north western and southern areas. Table 3 shows the data on corrosion (m) of aluminium and their respective durability factors [DF] obtained at three different locations. It can be seen that the durability factor of aluminium varies not only from location to location but also from station to station. Durability factor for aluminium varies from 34 to 144. USA: Laque center for corrosion is a pioneering institution involved in carrying out atmospheric corrosion studies. This center had ranked the corrosivity of a number of sites in Canada as well as USA using the mass loss technique. The data had indicated that short term mass loss data can exhibit wide variations because of uncontrolled environmental factors in natural atmospheric environments and seasonal effects. Thus longer exposure (e.g. 1-2 years) is intended to average out the influence of large fluctuations in short term (e.g. 1 month) environmental variables [178].

Table 4. Durability factors for sites at which carbon steel, galvanized steel and Aluminium exposed [178]. Sites Paraparaumu Flock House Levin Kairaga Tiwai point Invercargill Gore Tapanui Carbon steel 1 1 1 1 1 1 1 1 Galvanized steel 26 38 49 75 22 63 53 43 Aluminium 400 69 323 677 341 843 121 73

New Zealand: Is a group of islands in the southwest pacific. The climate is warm and humid with prevailing westerly winds depositing large amounts of sea salt far inland, condition thought to pose severe atmospheric corrosion hazard. Results of an atmospheric corrosivity survey of New Zealand are reported. Carbon steel, aluminium and galvanized steel are exposed for one year at 168 sites located throughout New Zealand. One year corrosion rates ranged between 18 4800 gm-2 per year for carbon steel and 0.7 1417 gm-2 per year for galvanized steel. Results for aluminium were significantly greater than zero at a number of severe marine sites. Maximum corrosion rate of 2.6 gm-2 per year was found. And at one industrial site a rate of 1.3 gm-2 per year was recorded. A clear correspondence between corrosion rates and proximity to the coast is evident in these results, inferring that atmospheric corrosion rate in New Zealand are related to levels of chloride deposition. Table 4 shows the durability factors for carbon steel, galvanized steel and aluminium. The durability factors for galvanized steel and aluminium varies from place to place. Durability factor for galvanized steel and aluminium varies from 22 to 75 and 69 to 843 respectively. Germany: The results of long time examinations of the corrosivity of the atmosphere demonstrated that in the period from 1979 to 1989 on the territory of the formerly GDR no changes could be established.

14

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

Table 5. Typical durability factor based on relative corrosion rates for galvanized steel and aluminium (one year data) [178].
mg/sq.dm.

25 20 15 10 5 0 Rainy Summer Season SEASON Winter

Location SVRECT, Surat MPT, Mormugoa NIO, Goa NMPT, Mangalore IOC, Mumbai INS Naval Base, Kochi Kayamkulam CECRI Unit, Tuticorin Mandapam Camp Nagapattinam Cuddalore INS Naval Chennai Near Nellore CECRI Unit Kochi Mettupalayam MPL, Manali Tirupur LPSC Mahendragiri Coimbatore Portblair

Galvanized steel 10.75 2.46 13.27 16.16 18.89 24.29 3.43 8.69 89.75 12.21 44.94 31.64 25.22 3.86 1.4

Aluminium 6.95 41.26 120.96 21.25 27.2 180 84 3.09 155.79 V.High 44.15 60.93 551.72 502.78 33.33 81.56 6 2.52 4 9.5

Figure 1. Average seasonal corrosion rate (in mg/sq.dm.) of different months.

The comparison of these results with those obtained at the same territory in the period 1989-1994 shows a significant decrease of corrosivity against metallic materials which is caused by a lower deposition rate of the corrosion pollutant SO2. This positive development can be explained by an improvement of the situation which occurs due to sequence of changes of the industrial structure as well as by active measure of environmental protection in the new countries of FRG after the political changes in the year 1989 [178]. India: It is almost 32 years since the first corrosion map of India was brought out. Over these years, lot of environmental changes has occurred due to industrialization, population growth and enormous vehicle population. Durability data clearly indicated that non ferrous viz galvanized steel and aluminium have better durability factors. However the factors vary from location to location. If durability factor and cost factor are taken together it can be clearly seen that aluminium has appreciable cost benefit ratio. At certain locations galvanized steel may prove to be a cost effective candidate material. Table 5 shows typical durability factor based on relative corrosion rates for galvanized steel and aluminium (one year data). Vashi, et al. [179] reports the corrosion rate of aluminium as well as the sulphation rate was determined under outdoor exposure at Ankleswar, South Gujarath representing an industrial atmosphere. Monthly corrosion rate of aluminium vary from 4 to 30 (1 to 5 m/y) mg/sq.dm. the values for yearly rates being 65 to 126 (1 to 15 m/y) mg/sq.dm. Aluminium or aluminium coated sheets would, therefore, give better performance in comparison with mild steel or zinc.

The corrosion rate of aluminium in rainy months (22.7 mg/sq.dm.) was higher than the rate inwinter months (6.5 mg/sq.dm.) and summer months (10.5 mg/sq. dm) (Fig. 1) [179]. Odnevall, et al. [180] summarized the results from an extensive field exposure program implemented to study possible seasonal dependencies of copper corrosion rates and runoff rates. Two year exposures in one urban and one rural environment were performed at four different starting seasons. An extensive multi-analytical approach was undertaken of all exposed samples. Seasonal differences in corrosion product formation was observed during the first month of exposure and attributed mainly to differences in relative humidity conditions. Seasonal differences in corrosion rate at the rural site could be discerned throughout the whole two-year exposure, again, mainly attributed to differences in relative humidity. No seasonal effect could be observed at the urban site indicating that other parameters influenced the corrosion kinetics at this site. While corrosion rates exhibit a continuous decrease with exposure time, the yearly runoff rates are independent of time. Depending on starting months the yearly copper runoff rates ranged from 1.1 to 1.7 gm 2y 1 for the urban site, and from 0.6 to 1.0 gm 2y1 for the rural site. These seasonal variations were primarily attributed to differences in precipitation quantity and environmental characterist-ics. Runoff rates are significantly lower than corrosion rates as long as the adhering copper patina is growing with exposure time. Almeida, et al. [44] summarizes the results obtained in the MICAT project for carbon steel specimens exposed for 1 to 4 years in 22 rural and urban atmospheres in the Ibero-American region. Test site characterization, chemical and morphological determination of the steel corrosion product layers (SCPLs) contributed to understanding the corrosion phenomena involved. It was observed how some climatological factors could affect steel corrosion rates and SCPL properties. Although the studied atmospheres were classified into different ISO groups, steel corrosion rates did not differ significantly

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

15

S. Syed

Table 6. MICAT S0P0,S0P1 and S0P2 site characteristics [44].


Carbonsteel corrosion rate (m year_1) 1year 4 years Name Carbon steel corrosion products

Deposition rate (mg m-2 d-1)

Time of wetness (TOW)

ISO Class (**)

Cl-

Cuzco Arties San Juan Iguazu Trinidad Riobamba Granada Brasflia Cuernavaca Pucallpa V. Martelli Arequipa San Pedro Belem Cotove Tortosa Leon Guayaquil La Plata Mexico S.L.Potosi Sao Paulo

(*) 1.7 (*) (*) 1.5 1.1 (*) (*) (*) (*) (*) (*) (*) (*) (*) (*) (*) 1.5 (*) (*) (*) (*)

(*) 9.0 (*) (*) 0.7 1.2 6.2 (*) 7.9 (*) 9.0 (*) 0.6 (*) 0.3 5.3 (*) 3.0 6.9 13.6 18.9 57.8

4 3 3 5 4 4 3 4 3 5 4 2 5 5 4 4 3 4 4 3 3 5

C2-C3 C3 C3 C3 C2-C3 C2-C3 -

1.4 3.9 4.9 5.7 6.7 8.4 8.5 12.9 13.4 14.3 14.7 15.4 17.0 19.4 19.6 20.2 20.8 22.6 28.1 9.7 31.1

0.8 3.3 1.9 2.8 5.2 5.2 8.7 6.6 8.0 7.7 7.8 6.2 6.1 10.7 12.9 15.3 6.6 20.3

C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C2 C3 C2 C3 C2

L.G L L L.G L L.G L.G L L L.G L.G L.G L.G L.G L L.G

L.G L L.G L.G L.G L.G L.G

L.G L.G L.G L.G L.G

66 101 96 38 35 91 65 87 52 308 53 57 61 41 46 249 34 84 49 270

20.6 8.3

(*) Apparently unpolluted atmosphere ( < 3 mg Cl- m-2 d-1 and < 10 mg SO2 m-2 d-1 (**) Based on climatological and pollution parameters; (*** ) Based on 1st year corrosion rate; - Not available ; C2 - Low; C3 - Medium; L lepidocrocite ( -FeOOH); G goethite ( - FeOOH);

Table 7. Copper and aluminium corrosion rates (g/m2 a + standard deviation) [181].
Exposure time (months) 6 12 18 Aluminium 6 0.3 + 0.04 12 0.3 + 0.02 18 0.2 + 0.01 Outdoor 27.4 + 1.6 19.4 + 0.4 14.3 + 0.6 Sheltered 9.8 + 0.8 11.5 + 0.5 8.6 + 0.7 1.1+ 0.14 1.2+ 0.14 0.8 + 0.11 Rural Ventilated shed 3.5 + 0.2 1.9 + 0.3 1.9 + 0.3 0.2 + 0.03 0.3 + 0.01 0.2 + 0.00 Closed space 0.2 + 0.04 0.2 + 0.02 2.0 + 1.2 0.4 + 0.08 0.3 + 0.10 0.3 + 0.04 Outdoor 5.3 + 0.7 34.6 + 0.04 30.3 + 2.0 4.3 + 0.2 3.2 + 0.03 2.2 + 0.4 Coastal Sheltered 44.5 + 0.5 54.1 + 2.6 72.4 + 3.5 4.8 + 0.1 4.0 + 0.4 3.6 + 0.2 Ventilated shed 19.8 + 4.5 47.2 + 3.7 33.1 + 0.5 2.2 + 0.1 1.6 + 0.3 1.4 + 0.1 Urban - industrial Outdoor Sheltered 23.9 + 1.8 6.8 + 1.5 19.8 + 0.04 8.5 + 0.4 15.0 + 0.5 7.5 + 0.7 0.7 + 0.05 0.6 + 0.05 0.4 + 0.00 1.1 + 0.04 1.7 + 0.13 1.8 + 0.1

16

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Soluble salts (mg m-2) SO42Cl47 213 341 104 133 411 522 94 245 811 113 0 128 259 487 487 294 518 128 2 102 0

ISO Class (***)

SO2

2 years

3 years

1 year

Atmospheric Corrosion of Materials

between them. The only common characteristic of these atmospheres was an increase in SCPLs protectiveness with exposure time. Table 6 summarizes the experimental results corresponding to carbon steel surfaces exposed in rural and urban atmospheres during the MICAT project. Antonio, et al., [181] reports the atmospheric corrosion of copper and aluminium exposed at three test sites indoors and outdoors (coastal, urbanindustrial and rural) under different exposure conditions up to 18 months based on the influence of environmental parameters and main pollutants (SO2 and chlorides) on the atmospheric corrosion of metals at Cuba. The interaction between the chloride deposition rate with the time of rainfall (outdoors) and with the time of wetness at temperature between 5oC and 25oC (indoors) were found to be the most significant variables influencing the corrosion of the two metals investigated; although other variables appeared to be important in the corrosion process depending on the metal nature. A classification of the atmospheric corrosion aggressivity of the test sites based both on environmental data and corrosion rate measurements was made according to ISO 9223. The corrosion aggressivity prognostic of this standard is not always in agreement with the results obtained in Cuban atmospheric conditions. Table 7 reported the average corrosion values for copper and aluminium, which corroded uniformly. Indira, et al.[182] studied the atmospheric corrosion behavior of galvanized steel and aluminium in marine environment at Kochi, India for a period of one year (from August `98 to July `99). Monthly corrosion rates of these metals were determined by mass loss method. The pollutants like chloride and sulphur dioxide present in atmosphere have been estimated periodically. These values were correlated with corrosion rate values. The rate of galvanized steel and aluminium are in the range of 0.0025 to 0.0314 mpy and 0 to 0.0014 mpy respectively. The durability values clearly indicated that non ferrous metals viz., galvanized steel and aluminium have better durability. Table 8 gives the monthly corrosion rates of galvanized steel and aluminium. Lopez-Delgado et al.[171] studied the copper corrosion rate and corrosion products originated by the action of acetic acid vapor at 100% relative humidity. Copper plates were exposed to an acetic acid contaminated atmosphere for a period of 21 days at Madrid, Spain. Five acetic vapor concentration levels were used. The copper corrosion rate was in the range of 1 to 23 mg/dm2 day. Some of the compounds identified were cuprite (Cu2O), copper acetate hydrate (Cu(CH3COO)2 2H2O), and copper hydroxide acetate (Cu4 OH(CH3COO)7. 2H2O). This last compound was also characterized. The thickness of the patina layers was 4 to 8 nm for amorphous cuprite, 11 to 48 nm for cuprite, and 225 nm for copper acetate.

1 0 8 6 4 2 0 Winter Summer Rainy Winter

SEASON 1 994-1 995 1 995-96

Figure 2. Average seasonal corrosion rate (mg/dm2) of aluminium. Table 8. Atmospheric corrosion rates of galvanized steel and aluminum at Kochi marine site [182]. S. No. 1 2 3 4 5 6 7 8 9 10 11 12 Corrosion rate is mpy Month Galvanized steel January 0.00466 February 0.0061 March 0.0098 April 0.0314 May 0.0229 June 0.0233 July 0.0291 August 0.007362 September 0.00821 October 0.00827 November 0.0044 December 0.0025 Aluminium Negligible -do0.00058 0.0014 0.00132 0.00121 0.0012 Nil Nil Nil Nil Nil

The patina, in which the cementation process of different corrosion-product layers plays an important role, is formed by the reaction of acetic vapor with copper through porous cuprite paths. Vashi and Patel[65] studied and determined the corrosion rate of aluminium as well as the sulphation rate under outdoor exposure at Baroda, India (Petrochemical Complex Area, Central Gujarat) representing an industrial atmosphere. Monthly corrosion rate of aluminium vary from 0.6 to 4.9 mg/sq.dm (0.3 to 2.2 m/y). The values for yearly rates being 25.5 to 37.6 mg/sq.dm (0.9 to 1.4m/y). Aluminium or aluminium coated sheets would, therefore, give better performance. Figure 2 shows average seasonal corrosion rate of aluminium. Mohan, et al.,[183] studied the proper corrosion preventive methods. Design engineers require corrosivity data of an area and the pollutants present in atmosphere. The intensity of corrosion of various metals and the pollution data of different locations in the southern region is reported. Table 9 shows the corrosion rates of aluminium, copper and stainless steel 304 for 3 months, 6 months, 9 months and one year. Corrosion rate at Site 1, 2 and 3 which are on eastern coast of India are higher than the sites 4 and 5, which are on western coast of India. Uniformly corroding metal copper shows a slight decrease in corrosion rate during prolonged exposure whereas in the case of aluminium and stainless steel 304 the initial corrosion rate is very low and increase rapidly with increase in time with the formation of pits.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

17

S. Syed

Table 9. Corrosion rate of metals at various locations [183]. Aluminium 3 months 6 months 9 months One year 3 months 6 months 9 months One year 3 months 6 months 9 months One year 3 months 6 months 9 months One year 3 months 6 months 9 months One year 0.0007 0.007 0.0026 0.0029 0.0110 0.0092 0.0017 0.0020 0.0001 0.0004 0.0005 0.0002 0.0001 0.0004 0.0014
Corrosion Loss in mg/sq.dm/month

14 12 10 8 6 4 2 0 Rainy Winter Summer

Period exposure

Corrosion rate in mm/y Copper Stainless steel 304 Site I 0.03000 0.00150 0.02000 0.00500 0.01800 0.00550 0.01500 0.00500 Site II Site III 0.25500 0.00060 0.01460 0.00060 0.00512 0.00027 0.00870 0.00040 Site IV Site V -

SEASON

Figure 3. Average seasonal corrosion rate (mg/sq.dm./month) of aluminium.

Table 10. Corrosion rate of aluminium from weight loss measurements [23]. Test station San juan Period (years) 1 2 3 4 1 2 3 4 1 2 3 4 ( m / year) 0.015 0.057 0.015 0.012 0.005 0.0015 0.14 0.21 0.29 0.13 0.08 0.15 1.54 1.07 1.38 1.19 0.43 0.67 Test station Iguazu ( m / year) 0.03 0.08 0.10 0.06 0.03 0.03 0.05 0.03 0.06 0.02 0.03 0.04 0.05 0.08 0.06 0.05 0.02 -

Camet

Villa Martelli

Vashi and Patel [184] reports the corrosion rates of aluminium as well as the sulphation rate have been determined under outdoor conditions of exposure at Surat (South Gujarat) representing an industrial atmosphere. The monthly rates of outdoor corrosion of aluminium vary from 1 to 40 mg/sq.dm (0.44 to 17.8 m/y), the values for yearly rates being 24 to 55 mg/dq.dm (0.87 to 2.01 m/y). Aluminium or aluminium coates sheets would, therefore, give better performance. Figure 3 shows average seasonal corrosion rate of aluminium. Vilche, et al. [23] reports the atmospheric corrosion of aluminum.The samples exposed at six test sites were investigated with known ambient parameters at Argentina after different outdoor exposition periods. Corrosion damage was determined by weight-loss measurements. Both D.C. and A.C. electrochemical techniques, performed in 0.1 M Na2SO4 solution employing the exposed face of the test samples, were used to characterize the protectivenesses of the surface layers generated on the metals at different exposure times during the atmospheric corrosion process in the distinct environments. The time dependence of corrosion rates as compared with the weight-loss data as well as with information from SEM and EDAX observations of the rusts concerning the structure and morphology of surface corrosion products and the presence of pollutants. The corrosion rates of Al were calculated by considering the total affected area, faces to the sky and to the ground, metal density and exposure time. Corrosion rates expressed in m year-1 are shown in Table 10.

Jubany

La Plata

Julve and Gustems[185] studied the outdoor corrosion resistance of zinc-electroplated steel (15 m thickness) for a period of four years at 11 outdoor exposure sites in the metropolitan area of Barcelona. The results obtained were evaluated by means of the mass loss method and by the ASTM `rating number method. The corrosion observed in the exposed samples (zinc- electroplated steel) was discussed, as well as the influence of relative humidity, chlorides, sulphur dioxide and particles in suspension. The corrosion increased in marine environment of zinc electroplated steel. Table 11 summarizes the corrosion rate of zinc -electroplated steel. Eriksson et al. [109] reported the reaction of SO2 and NO2 in sub-ppm concentrations in humid air with a copper surface. The deposition of SO2 was studied in a time-resolved manner and the formation of HNO2 was measured. Corrosion products were analyzed using XRD, FTIR, FTIRAS, and ion chromatography. SIV-O species on the copper surface was analyzed quantitatively using a sensitivity of 1011 mol.

18

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

Table 11. Corrosion rate (kg m-2 y-1) of zinc - electroplated steel 185]. Outdoor exposure site Molina Poble Nou Hospitalet Sant Adria Badalona Montcada Sant Vicenc Viladecans Cornella Prat Port Type of atmosphere Urban Urban-industrial Urban Urban-industrial Urban-industrial Urban-industrial Rural-industrial Rural Urban Rural-Industrial Marine Zinc-electroplated steel 0.0055 0.0089 0.0059 0.0088 0.0092 0.0071 0.0056 0.0052 0.0049 0.0056 0.0300

Table 13. Rating scale for examination [189]. Rating 0 1 2 3 Surface appearance of plate No corrosion or staining Staining and/or spot rusting (0-25%) Discoloration and/or increased corrosion (25-75%) Corrosion over entire surface (75-100%) Weld appearance Unaffected Minor corrosion Crevice attack and/or pitting corrosion visible Severe corrosion; cracking

Table 14. Corrosion rating after 3 years exposure according to rating scheme in Table 13* [189]. Panel 1A 1B 1C 2A 2B 2C 3A 3B 3C 4A 4B 4C Site 1 Panel 1 2 1 1 2 0 0 2 0 0 1-2 0 Weld 0-1 0 0-1 0 0 0 0 0 0-1 0 0 0 Site 2 Panel 2 3 1 1 2 1 1 3 1 0 1 0 Weld 1 0 0-1 0 0-1 0 1 1-2 1 0 0-1 0 Site 3 Panel 0 1 0-1 0 1 0 0 1 0 0 0-1 0 Weld 0 0-1 0 0 0 0 0 0-1 0 0 0 0

Table 12. Mass gain and metal loss of copper samples exposed to SO2 and NO2 at 90% RH for 4 weeks [109]. Atmosphere Pure air SO2 NO2 NO2 - SO2 Mass gain of dried samples (mg/cm2) 0.002 0.043 0.008 0.18 Metal loss (mg/cm2) 0.012 0.050 0.021 0.270 Mass of corrosion products / Mass of Cu corroded 1.2 1.86 1.4 1.68

Corrosion rates and corrosion products are reported. The corrosion of copper in humid SO2 polluted air is accelerated strongly by the presence of NO2. Table 12 summarizes the metal loss and mass gain data from all experiments performed at 90% RH. Boulton, et al. [189] reports the results of a 3 years study of the corrosion performance of welded panels of stainless steel 304 in three atmospheric environments. The alloys chosen were special grades used in process industries for fabrication of plant for exterior applications such as tankage, pipework and processing equipment. The panels were welded with austenitic filler metals and three surface finishes were employed : ground, polished and pickled .The three environments chosen for the exposure sites were severe marine, urban-marine and a geothermal atmosphere. These examinations were carried out in natural light and the extent of any corrosion attack or discoloration was marked according to the rating scale given in Table 13. The trends for the urban-marine site followed the same gradation as that found at the severe marine site. A summary of the degree of corrosion after 3 years, for all panels, is given in Table 14. Mohan, et al. [190] reported that the rate of corrosion of copper is affected by relative humidity, temperature and pollutants like sulphur dioxide and chloride etc. Determination of corrosion rate in the atmosphere is time consuming. Hence laboratory tests like Damp Heat Alternating Atmosphere test containing sulphur dioxide is designed to simulate the conditions in the industrial atmosphere.

*Upper significant surface only;

Table 15. Corrosion of copper (mg/dm2) at 80% RH [190]. Cycle s 1 2 3 4 5 1 2 3 4 5 Copper 0.1 l SO2 35oC 1.2 1.3 1.5 1.6 1.7 40oC 0.8 1.1 0.9 1.1 1.6 0.53 0.67 0.93 1.47 2.13 0.53 0.67 1.47 2.0 2.7 0.2 l SO2

The rate of corrosion of copper are determined in the Kesternich Type Apparatus for five cycles, each cycle consisting of 8 hours `ON and 16 hours `OFF. Experiments performed with different humidities from 60 to 100% RH at different temperatures 35oC and 40oC and different concentration of sulphur dioxide 0.1 to 0.3 litres shows that the corrosion rate increases in all the cases. Table 15 shows the corrosion rate of copper at 80% RH at 35oC and 40oC with 0.1 l and 0.2 l sulphur dioxide. Table 16 indicates the corrosion rates of copper at 35oC and 40oC at 100% RH with sulphur dioxide of different concentrations. Table 17 shows the corrosion rates of copper at different relative humidities and temperatures with 0.1 l of sulphur dioxide.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

19

S. Syed

Table 16. Corrosion of copper (mg/dm2) at 100% RH [190]. Cycles 0.1 l SO2 1 2 3 4 5 1 2 3 4 5 0.5 1.2 1.6 1.7 1.5 1.9 2.6 2.7 2.7 Copper 0.2 l SO2 35oC 0.27 0.87 0.80 1.2 40oC 0.93 2.17 4.0 4.0 6.2 (mg/dm2) 0.27 0.9 1.73 2.27 2.27 at 0.1 litre of sulphur 100% RH 35oC 0.5 1.2 1.6 1.7 40oC 1.5 1.9 2.6 2.7 2.6 0.3 l SO2

Table 19. Weight losses after 1 and 2 years, exposure and salt spray test [47]. Exposure site Penghu Keelung Peitou Kaoshiung Pingtung Taitung Salt spray Salty spray Salt spray Exposure period (years) 1 2 1 2 1 2 1 2 1 2 1 2 0.25 2 1.3 Wt. Loss (g.m-2) Class A Class B Class C 83.5 29.4 4.7 66.8 4.7 20.5 9.5 85.1 4.3 85.1 0.9 98.4 225 855 33.5 0 0 0 0 0 8.2 41.1 42.3 0 0 0 0 0 7.6 36.2

Table 17. Corrosion rates of copper dioxide [190]. Cycle s 1 2 3 4 5 60% RH 35oC 1.0 1.9 2.1 2.4 3.1 40oC 1.73 1.9 2.5 2.9 3.7

80% RH 35oC 1.2 1.3 1.5 1.6 1.7 40oC 0.8 1.1 0.9 1.1 1.6

Table 20. Corrosion rate of metals at site No.1 [191]. Period of exposure Quarterly Jan-Mar Apl-Jun Jul-Sep Oct Dec Half-yearly Jan-June Jul-Dec Nine months Jul-Mar Yearly Jul-Jun Corrosion rate (mmd) Aluminium Copper 0.028 0.214 0.818 0.093 0.151 0.686 0.131 0.150 0.981 5.670 6.233 3.114 3.155 3.563 1.251 2.146 Stainless steel 304 0.007 0.078 0.140 0.012 0.060 0.140 0.060 0.088

Table 18. Corrosion rates of copper (mg/dm2) at 0.2 litre of sulphur dioxide [190]. Cycle s 1 2 3 4 5 60% RH 35oC 40oC 1.5 2.9 5.3 7.7 8.7 80% RH 35oC 0.53 0.67 1.47 2.00 2.70 40oC 0.53 0.67 0.93 1.47 2.13 100% RH 35oC 0.27 0.87 0.80 1.20 40oC 0.93 2.17 4.00 4.00 6.20

Table 18 presents the corrosion rates of copper at different relative humidities and temperatures with 0.2 l of sulphur dioxide. Horng, et al. [47] initiated the exposure tests of 3 classes (A,B,C ASTM 475-78) of galvanized steel wire strand in 1984 in rural, marine, and industrial areas of Taiwan. They studied the atmospheric corrosion behaviors of galvanized steel wire strand from the results of two years exposure test. All the results were compared with the laboratory results derived from salt spray test. The role of salinity and other weathering parameters were analyzed, and amazingly, the salinity and the corrosivity at Penghu is exceptionally high and class A galvanized steel wire strand started to rust after three months exposure. The corrosion products of testing specimens were also analyzed with X-ray diffraction and SEM-EDS and the possible mechanisms of atmospheric corrosion of galvanized steel wire strand were determined.

Table 19 shows the weight loss results. Class A of galvanized steel wire strand at Penghu (Pescadores) has the highest corrosion rate of all the exposure sites after 2 years exposure. The average corrosion rate of class A galvanized steel wire strand at Penghu is from three times to fourteen times greater than those at the other exposure sites. The corrosion rates of class B and class C are smaller than class As at Penghu after one year exposure. Table 20 gives the corrosion rates of aluminium, copper and stainless steel 304 for quarterly, half-yearly, 9 months and one year at the 45m site (No.1). Table 21 shows the yearly corrosion rates of aluminium, copper and stainless steel 304 by monthwise at 45 m (Site No.1). Sundaram, et al. [191] reported that the corrosion of metals in the atmosphere is greatly influenced by man made and natural environments. The presence of moisture is a must for the corrosion reaction to take place in the atmosphere. The prevalance of high humidity along with high temperature and intensive solar radiation in tropical climate is dangerous as it

20

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

Table 21. Yearly corrosion rate of metals at site No.1 [191]. Period of exposure Jan-Dec Feb-Jan Mar-Feb Apl-Mar May-Apl Jun-May Jul-Jun Aug-Jul Sep-Aug Oct-Sep Nov-Oct Dec-Nov Corrosion rate ( mmd) Aluminium Copper 0.061 0.082 0.062 0.070 0.117 0.150 0.098 0.069 0.076 0.044 0.099 2.121 1.987 1.929 2.274 2.325 2.146 2.718 3.004 3.232 3.055 3.485 Stainless steel 304 0.050 0.044 0.039 0.043 0.047 0.088 0.027 0.041 0.027 0.032 0.038

Table 22. Corrosion rate of metals at site No.2 [191]. Period of exposure Quarterly Jan-Mar Apl-Jun Jul-Sep Oct-Dec Half yearly Jan-Jun Jul-Dec Nine months Jul-Mar Yearly Jan-Dec Jul-Jun Corrosion rate ( mdd) Aluminium Copper 0.0234 0.0685 0.0697 0.0278 0.053 0.057 0.047 0.0403 0.0425 2.331 4.700 5.352 1.097 2.986 3.027 2.183 2.187 1.874 Stainless steel 304 0.0265 0.0494 0.0873 0.0431 0.0545 0.0492 0.033 0.033 0.0926

atmosphere of Mandapam Camp. Weight-loss measurement was employed in computing the corrosion rates. The monthly corrosion rate of brass and the quarterly corrosion rate of aluminium were in line with the aggressiveness of the environment decided by the salinity value, temperature and humidity. The corrosion products formed on all the three alloys tested were compact and adherent. The direct attack of pollutants on the metals for further corrosion was hindered by the presence of these layers, and their protective nature was explicit from the exponential decrease of corrosion rates with time. In the high chloride containing atmosphere, the protection offered by the corrosion product on aluminium was not so effective as that on the copper base alloys. Table 23 shows the corrosion rate of the metals for 2 years exposure in comparison with a few other studies. Leest[152] exposed the sputtered and electrodeposited copper specimens to the laboratory atmosphere for 18 months, after which the corrosion products were analysed by electrochemical methods and reflectance spectra. The formation of corrosion products takes place in three stages: initially cuprous oxide is formed, followed by the growth of cuprous sulphide, and finally curprous oxide is converted to cuprous sulphide. The presence of water is essential for these corrosion reactions to occur.

6. CONCLUSIONS
This paper provides a comprehensive state of the art review of the atmospheric corrosion of materials in various atmospheres such as rural, urban, industrial, marine, or combinations of these. It also provides basics of atmospheric corrosion, influence of exposure parameters namely critical relative humidity, temperature, specific atmospheric corrodants (pollutants) and other atmospheric contaminant and airborne particles. Finally it provides the related work done worldwide on atmospheric corrosion of various materials such as hot and cold carbon steel, galvanized steel, stainless steel 304, copper, brass and aluminium. Therefore, material scientists have an important role to play in selection of materials because of atmospheric corrosion accounts for more failures on both a tonnage basis and cost basis than any other type of environmental corrosion. REFERENCES
[1] [2] [3] Trimgham T.C.E, 1958. Causes and Prevention of Corrosion in AIR Craft, SIR ISAAC, Pitman & Sons, Ltd. Speller F.N., 1951. Corrosion (Causes and Prevention) McGraw-Hill Book Company Inc., New York. Fontana M.G, 1987. Corrosion Engineering, 3rd Ed. McGraw-Hill Book Company Inc. New York.

Table 23. Corrosion rate, g/m2/day [193]. Metal Copper Brass Aluminium Present study 2 years 0.050 India 2 years 0.127 0.0585 0.00698 Sweden 2 years 0.00245 0.00200 Panama 1 year 0.10130 0.03014 0.00274

causes rapid deterioration of materials. Investigations made at Mandapam Camp, a tropical marine location on the south-east coast of India. Table 22 shows the results of aluminium, copper and stainless steel 304 at site No.2. Espada, et al., [192] reported the three years of exposure for field tests of atmospheric corrosion of low carbon steel, and copper at twenty-four sites located along the Galicia Coast (North-Western of Spain. The test sites were chosen in a large variety of marine environments, at different distances from the sea and altitudes. A statistical analysis of data lets to obtain equations for to predict the corrosion rate at greater time. They present a correlation coefficient of 0.967 for low-carbon steel, and 0.905 for copper. Ananth, et al. [193] studied the corrosion rates of commercially available aluminium and brass were determined by exposure in the tropical marine

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

21

S. Syed

[4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41]

Natesan M, 1995. PhD Thesis, Anna University, Chennai, India. Evans U.R and Hoar T.P, 1932. Proc. Roy. Soc.(A), 137-343. Naixin X, Zhao L, Ding C, Zhang C, Li R and Zhong Q, 2002. Corros. Sci., 44-163. Brown P.W and Masters L.W, 1982. Atmospheric Corrosion, Wiley, New York. Scully J.C, 1990. The Fundamentals of Corrosion 3rd Edn., Pergamon Press, New York. Brown P.W and Masters L.W, 1982. Atmospheric Corrosion, Wiley, New York. Grossman P.R, 1987. Atmospheric Factors Affecting Engineering Metals, ASTM STP, 646. Chawla S.K and Payer J.H, 1990. Corros. 46:860. Walters G.W, 1991. Corros. Sci., 32: 1331. Walters G.W, 1991. Corros. Sci., 32:1353. Vassie P.R, 1987. Br. Corros. J., 22:37. Black H.L and Lherbier L.W, 1968. Metal Corrosion in the Atmosphere, ASTM STP, 435:3. Thomas H.E and Alderson H.N,1968. Metal Corrosion in the Atmosphere, ASTM STP; 435:83. Briggs C.W, 1968. Metal Corrosion in the Atmosphere, ASTM STP; 435:271. Lipfert F.W, 1987. Mater. Perf., 26:12. Stiles D.C and Edney E.O, 1989. Corros, 45:896. Pourbaix M and Pourbaix A, 1989. Corros., 45:71. Naeemi A.H and Albrecht P, 1984. Int. Cong. Metallic Corros., Toronto, Canada, 418. Money K. L, 1987. Metals Handbook Corrosion, Metals Park, Ohio, ASM International, 204. Vilche J. R, Varela F.E, Acuna G, Codaro E.N, Rosales B.M, 1995. Fernandez A and Moriena G, Corros. Sci., 37:941. Oesch S and Faller M, 1997. Corros. Sci., 39:1505. Barton K, 1976. Protection against Atmospheric Corrosion, John Wiley and Sons, London, Graedel T.E,1996. Corros. Sci., 38:2153. Roberge P.R, 1999. Handbook of Corrosion Engineering, McGraw-Hill, USA. 58. Uhligh H.H, 1985. Corrosion and Corrosion Control, John Wiley and Sons, New York. Sastry T.P and Rao V.V, 1981. J. Electrochem. Soc. India, 30:289. Corvo F, 1984. Corros, 40: 4-15. Morcillo M, Chico B, Otero E and Mariaca L, 1999. Mater. Perf., 24:72. Evans U.R, 1981. An Introduction to Metallic Corrosion, Arnold, U.K. Morcillo M, 1995. Atmospheric Corrosion , ASTM STP 1239, ASTM, Philadelphia, 257. Evans U. R. and Taylor C.A 1972. .J, Corros. Sci., 12:227. Morcillo M, Chico B, Mariaca L and Otero E, 1999. Corros. Sci., 41:91. Galloway J.N and Likens G.E,1976. Water Air Soil Poll., 6:241. Graedel T.E, McGrory-Joy C and Franey J.P, 1986. J. Electrochem. Soc., 133:452. Alves R.C.D and Ferreira M.G.S, 1992. J. Electrical Chem., 340:137. Brierly W.B, 1965. J. Environ. Sci., 15:5. Evans U.R, 1976. The Corossion and Oxidation of Metals: Second Supplementary Vol. 8, Edward Arnold, London. Ericsson R, 1978. Werks. Korros., 29:400.

[42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54]

[55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75]

Feliu S, Morcillo M and Feliu Jr. S, 1993. Corros. Sci., 34:403. Sawant S.S and Wagh A. B, 1991. Corros. Prev. Cont., 6:75. Almeida E, Morcillo M, Rosales B and Marrocos M, 2000. Mater. Corros. 51:859. Balasubramanium R, Mungole M.N and Ramesh K.A.V, 2001. J.Electrochem. Soc.India, 50:107. Spence J.W, Haynie F.H, Lipfert F.W, Cramer S.D and McDonald L.G, 1992. Corros., 48:1009. Horng, Y.T, Yiang I and Chang T.C, 1987. 10th Int. Cong. Metallic Corros., Madras, India. Mardar A.R, 1997. ASM Handbook, ASM, Metals Park, Vol. 20, 470. Goodwin F.E, 1990. Metallurgy and Performance, TMS, Warrendale, 183. Lobnig R.E, 1996. J. Electrochem. Soc., 143:1539. Nriagu, S.O, 1978. Sulfur in the Environment. Part II. Ecological Effects, John Wiley, New York. Anderson F.A, 1956. ASTM STP 175 Philadelphia, PA. Barton K, 1972. Protection Against Atmospheric Corrosion, John Wiley, New York. Flinn D.R, Cramer S.D, Carter J.P, Hurwitz D.M and Linstrom P.J, 1986. Materials Degradation Caused by Acid Rain, American Chemical Society, Washington, D.C. Haynie F.H, 1988. ASTM STP 965 Philadelphia, PA:ASTM, 282. Haynie F.H, 1980. ASTM STP 691 Philadelphia, PA:ASTM, 157. Evans T.E, 1972. Proc. 4th Int. Cong. Metallic Corros., Texas, 408. Hatch J.E, 1984. Aluminium: Properties and Physical Metallurgy, Metral Park,OH, ASM,Int., 200. Godard H.P, Jepson W.B and Bothwell M.R, 1967. The Corrosion of Light Metals, John Wiley & Sons Inc., New York, 170. Hatch and John E, 1984. Aluminium: Properties and Physical Metallurgy, ASM OHIO, 256. Caruthers W.H, 1986. The Theory and Practice, NACE, Houston, Texas, USA.. Pearson E.C, 1952. Huff H.J and Hay R.H, Can. J. Technol.,30:311. Cohen H.M, 1988. Aluminium, 35:197. Graedel T.E, 1989. J. Electrochem. Soc., 136:193. Vashi R.T and Patel H.G, 1997. Bull. Electrochem., 13:343. Mattson E, 1985. Chemtech, 15:234. Johansson E, 1984. KI report, Swedish Corrosion Institute. Graedel T.E, J. 1989. Electrochem. Soc., 136:204C. Barton K, 1976. Protection Against Atmospheric Corrosion, John Wiley& Sons, New York. Sinclair J.D and Psota-Kelty L.A, 1984. Proc. 9th Int. Cong. Metallic Corros., Ottawa, Canada, 8:296. Sinclair J.D, 1988. J. Electrochem. Soc., 135:89C. Sinclair J.D and Psota-Kelty L.A, 1990. Atmos.Environ., 24A:627. Patterson W.S and Wilkinson J.H, 1938. J.Soc. Chem. Ind., 57:445. Wilkinson J.H and Patterson W.S, 1941. ibid,60:42. Sanyal B and Bhadwar D.V, 1962. J. Sci. Ind. Res., 21D: 243.

22

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

Atmospheric Corrosion of Materials

[76] [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] [96]

[97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] [111]

Arroyave C and Morcillo M, 1995. Corros. Sci., 37:293. Cuesta O, 1993. Ph. D. Thesis, Royal Institute of Technology, Stockholm, Sweden. Schmid G.H, Z. 1959. Electrochem, 63:1183. Donchenko M.I and Saenko T.V, 1979. Metall., 15:96. Divers E and Mellor J.W, 1952. Comprehensive Treatise on Inorganic and Theoretical Chemistry, Longmanns, London, 3:94. Graedel T.E, Nassau K. and Franey J.P, 1987. Corros. Sci., 27:639. Graedel T.E,1987. ibid., 27:721. Graedel T.E,1987. ibid., 27:741. Vernon W.H.J, 1935. Trans. Far. Soc., 31:1668. Rice D.W, Peterson P, Rigby E.B, Phipps P.B, Cappell R.J and Tremoureus R, 1981. J. Electrochem. Soc., 128:275. Graedel T.E, Franey J.P, Gualtieri G.J, Kammlott G.W and Malm D. L, 1985. Corros. Sci., 25:1163. Zakipour S, Leygraf C and Portnoff G, 1986. J. Electrochem. Soc., 135:873. Dante J.F and Kelly R.G, 1993. ibid., 140:1890. Eriksson P, Johansson L.G and Strandberg H, 1993. ibid., 140:53. Vernon W.H.J, 1931. Trans. Far. Soc., 27:255. Sinclair J.D, 1993. The Electrochem. Soc. Proce. Series, Pennington, New Jersey, 93:325. Sinclair J.D, Psota-Kelty L.A, Weschler C.J and Shields H.C, J. 1990. Electrochem. Soc., 137:1200. Sinclair J.D, Psota-Kelty L.A and Peins G.A, 1992. Atmos. Environ., 26A:871. Lobnig, R.E., Frankenthal R.P, Siconolfi J.D, Sinclair J.D and Stratmann M, 1994. J. Electrochem. Soc., 141:2935. Sharma S.P, 1978. ibid., 125:2005. Fukuda Y, Fukushima T, Sulaiman A, Musalam I, Yap L.C,Chotimongkol L, Judabong S, Potjanart A, Keowkangwal O, Yoshihara K and Tosa M, 1991. J. Electrochem. Soc., 138:1238. Odnevall I and Leygraf C, 1995. ibid, 142:3682. Forslund M and Leygraf C, 1993. The Electrochem. Soc. Proc. Series, Pennington, New Jersey, 486. Tidblad J and Leygraf C,1995. J. Electrochem. Soc., 142:749. Forslund M and Leygraf C, 1996. ibid., 143:105. Schikorr G, 1964. Werkst. Korros., 15:457. Skorchelletti V.V and Tukachinsky S.E, 1955. J.Appl. Chem., 28:615. Barton K and Bartonova Z, 1970. Werkst. Korros., 21:85. Schikorr G, 1967. Werkst. Korros., 18:514. Schwarz H, 1965. ibid., 16:93. Kucera V, et al., 1996. 13th Int. Corros. Cong., Melbourne, Australia. Arroyave C, Lopez F.A and Morcillo M, 1995. Corros. Sci., 37:1751. Arroyave C and Morcillo M, 1996. 13th Int. Corros. Cong., Melbourne, Australia. Eriksson P, Johansson L.G and Strandberg H, 1993. J. Electrochem Soc., 140: 53. Baboian R, 1991. Corrosion 91. Paper No.371, NACE. Graedel T.E and Frankenthal R.P, 1990. J. Electochem. Soc., 8:137.

[112] Tidblad J, Leygraf C and Kucera V, 1991. J. Electrochem. Soc., 138:3592. [113] Fyfe D, 1994. Corrosion, Oxford, ButterworthHeinemann, London. [114] Rozenfeld I.L, 1961. Proc. 1st Int. Cong. Metallic Corros., Butterworth, London, 243. [115] Seinfeld J.H, 1986. The Atmospheric Chemistry and Physics of Air Pollution, Wiley, New York. [116] Ericsson P and Johansson L.G, 1986. Proc. 10th Scandinavian Corros. Cong., Stockholm. [117] Simon D, Mollimard D, Perrin C and Bardolle J, 1988. Proc. 14th Int. Conf. Electric Contacts, Paris. [118] Feliu S, Morcillo M and Feliu Jr. S, 1993. Corros. Sci., 34:415. [119] Corvo F and Leon I, 1988. Rev Iberoamericana de Corrosion y Proteccion, 5:19. [120] Butcher S.S, Charlson R.J,Orians G.H and Wolfe G.V, 1992. Global Biogeochemical Cycles, Academic press, Inc., London. [121] Falk T, Svensson J.E and Johansson L.G, 1988. J.Electrochem. Soc., 145:39. [122] Lindstrom R ,Svensson J.E and Johansson L.G, 2000. J.Electrochem. Soc.,147;1751. [123] Graedel T.E, 1987. J. Electrochem. Soc., 134:109. [124] Blucher B D, Lindstrom R, Svensson J.E and Johansson L.G, 2001. J. Electrochem. Soc., 148: B127. [125] Kok, G.L, Darnall K.R, Winer A.M, Pitts J.N, Jr and Gay B W, 1978. Environ. Sci. Technol., 12:1077. [126] Zika R, Saltzman E., Chameides W.L and Davis D.D, 1982. J. Geophys. Res., 87:5015. [127] Raynor G.S, Smith M.E and Singer I.A, 1974. J. Air Poll Contr. Assoc., 24:586. [128] Graedel T.E, 1986. J. Electrochem. Soc., 133:2476. [129] Petrova T.I, Kharitonova N.L and Popova A.I, 1982. Tr. Mon Energ. Inst., 575:86. [130] Moskivin, L.N, Efimov A.A, Teterin V.F and Tomilov S.S, 1982. Teploenergetika (Moscow), 9:12. [131] Calvo E.J and Schiffrin D.J, 1984. J. Electrochem. Soc., 10:257. [132] Mayne J.E.O, and Burkill J.A, 1986. Br. Corros. J., 21:20. [133] Marsh G.P, Taylor K.J, Bryan G and Worthingus S.E, 1986. Corros. Sci., 26:971. [134] Garrels R.M, 1954. Geochim. Cosmochim. Acta, 5 :153. [135] Van V, C.C, Luria M, Ray J.D, Gunter R.L, Wellman D.L and Boatman J.F, 1987. EOS-Trans. AGU, 68:1212. [136] Seinfeld J,, 1986. Atmospheric Chemistry and Physics of Air Pollution, Wiley, New York. [137] Manning M, 1988. Air pollution, acid rain and the environment. Report by Watt Committee on Energy HMSO, U.K. [138] Feitnecht W, 1952. Chimia, 6:3. [139] Barton K and Bartonova Z, 1969. Proc. 3rd Int. Congr. Metallic Corros., 4:403. [140] Askey A, Lyon S.B, Thompson G.E, Johnson J.B, Wood G.C, Cooke M and Sage P, 1933. Corros. Soc., 34:233. [141] Rice D.W, Peterson P, Rigby E.B, Phipps P.B.P, Cappell R.J and Tremoureux R, 1981. ibid, 128:275.

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

23

S. Syed

[142] Graedel T.E, Franey J.P and Kammlott G.W, 1984. Corros. Sci., 224:599. [143] Svensson J.E and L.G. Johansson, 1993. J. Electrochem. Soc., 140:2210. [144] Eriksson P, PhD Thesis. 1992. Chalmers University of Technology, Goteborg, Sweden. [145] Zakipour S, Tidblad J and Leygraf C, 1995. J. Electrochem. Soc., 142:757. [146] Aastrup T, Wadsak M, Schreiner M and Leygraf C, 2000. J. Electrochem. Soc., 147:2543. [147] Karlsson A, Moller P. J and Vagn J, 1990. Corros. Sci., 30:153. [148] Franey J.P, Graedel T.E and Kammlott G.W, 1982. Atmospheric Corrosion, J.Wiley & Sons, Inc. [149] Sharma S.P, 1980. J. Electrochem. Soc., 127:21. [150] Beck L and Gunter J.R, 1984. Thin Solid Films, 117:131. [151] Santucci S and Picozzi P, 1984. Thin Solid Films, 113:243. [152] Leest R.E.v.d, 1986. Werkst. Korros., 37:629. [153] Abbott W.H, 1992. Proc. 16th Int. Conf, Eectrical Contacts, Loughborough, England. [154] Lenglet M. Lopitaux J, Leygraf C, Odnevall I, Carballeira M, Noualhaguet J.C, Guinement J, Gautier J and Boissel J, 1995. J. Electrochem. Soc., 142:3690. [155] Vernon W.H.J, 1934. J.Chem. Soc., 1853. [156] Keene W.C, Galloway J.N and Holden J.D, Jr., 1983. J. Geophys. Res., 88:5122. [157] Kawamura K and Kaplan I.R, 1984. Anal. Chem., 56:1616. [158] Keene W.C and Galloway J.N, 1984. Atmos Environ., 18:2491. [159] Clarke S.G. and Longhurst E.E, 1961. J. Appl. Chem., 11:435. [160] Evans U.R, 1951. Chem. Ind., 706. [161] Schikorr G, 1961. Werkst. Korros., 12:1. [162] Arni P.C, Cochrane G.C and Gray J.D, 1965. J. Appl. Chem., 15:305. [163] Donovan P.D, 1986. Protection of Metals from Corrosion in Storage and Transit, Ellis Horwood, Chichester, England, 78. [164] Donovan P.D and Stringer J, 1971. Br. Corros. J., 6:132. [165] Donovan P.D and Stringer J, 1972. Proc. 4th Int. Cong. Metallic Corros., NACE, Houston, Texas, 537. [166] Cermakova D and Vlchkova Y, 1966. Proc. 3rd Int. Cong. Metallic Corros., Moscow, 497. [167] Bastidas J.M and Mora E.M, 1998. Can. Metall. Quart., 37:57. [168] Parkinson A.R, 1980. Anti-Corros., Methods. Mater., 37:11. [169] Knotkova-Cermakova D and Vlchkova Y, 1971. Br. Corros. J., 6:17.

[170] Notoya T, 1991. J. Mater. Sci. Lett., 10:389. [171] Lopez-Delgado A, Cano, E, Bastidas J.M and Lopez F.A, 1998. J. Electrochem Soc., 145:4140. [172] Lobnig R.E, Frankenthal R.P, Siconolfi D,J and Sinclair J.D, 1993. J. Electrochem. Soc., 140;1902. [173] Sinclair J.D, 1988. ibid., 135:89C. [174] Lobnig, R.E, Siconolfi D.J, Maisano J, Grundmeier G, Streckel H, Frankenthal R.P., 1996. Stratmann M and Sinclair J.D, J. Electrochem. Soc., 143:1175. [175] Stevenson, C.M and Roy Q.J, 1968. Mateorol. Soc., 94:56. [176] Shreir, L.L, Jarman R.A and Burstein G.T, 1994. Corrosion Vol. 1, Butterworth-Heinemann, Ltd. Great Britain, 2:35. [177] Sharp S, 1990. Mater. Perf., 29:43. [178] Natesan M, Palaniswamy N, Rengaswamy N.S, Rajesh Kuamr S and Raghavan M, 2002. 7th. Int. Symp. Advance in Electrochem. Sci. & Technol, Chennai, India. [179] Vashi R. T, Malek G. M, Champaneri V. A and Patel R.N, 2002. Bull.Electrochem., 18:91. [180] Odnevall W.I and Leygraf C, 2001. Corros. Sci., 43:2379. [181] Antonio R. M and Corvo F, 2000. Corros. Sci., 42:1123. [182] Indira C.J, Prasad K.V, Syamala Kumari V.S, Namboodiri P.N.N, Natesan M, Palaniswamy N and Raghavan M, 2000. 10th Nat. Cong. Corros, Cont., Madurai, India, 148. [183] Mohan P.S, Natesan M, Sundaram M and Balakrishnan K, 1996. Bull. Electrochem., 12:91. [184] Vashi R.T and Patel R. N, 1996. Bull. Electrochem., 12:477. [185] Julve E. and Gustems L.L, 1993. Corros. Sci., 35:1273. [186] Joel P. S. and Karl W, 1992. Mater. Perf., 31:46. [187] Allam I.M, Arlow J.S and Saricimen H, 1991. Corros.Sci., 32: 417. [188] Chawla S.K and Payer J.H, 1990. J. Electrochem. Soc., 137:60. [189] Boulton L.H, Miller N.A and Sanders M.C, 1988. Br. Corros. J., 23:117. [190] Mohan P.S, Sundaram M and Guruviah S, 1987. 10th Int. Cong. Metallic Corros., Madras, India, 179. [191] Sundaram M, Mohan P.S and Ananth V, 1987. 10th Int. Cong. Metallic Corros., Madras, India, 143. [192] Espada L, Merino P, Gonzalez A and Sanchez A, 1987. 10th Int. Cong. Metallic Corros., Madras, India, 3. [193] Ananth V, Subramanian G, Mohan P.S and Palraj S, 1986. Bull. Electrochem., 2:541. [194] Graedel T.E, 1994. J. Electrochem. Soc., 141:922.

24

Emirates Journal for Engineering Research, Vol. 11, No.1, 2006

You might also like