Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and...

7
International Journal of Offshore and Polar Engineering (ISSN 1053-5381) http://www.isope.org/publications Copyright © by The International Society of Offshore and Polar Engineers Vol. 22, No. 4, December 2012, pp. 257–263 Jin S. Chung Award Lecture, 2012 Modeling and Prediction of Long-Term Corrosion of Steel in Marine Environments Robert E. Melchers Centre for Infrastructure Performance and Reliability, The University of Newcastle Newcastle, NSW, Australia This paper reviews recent developments in the prediction of the likely future corrosion losses and of the maximum pit depth for steels exposed to marine environments. A robust mathematical model based on corrosion science principles and calibrated for immersion conditions to an extensive range of literature and new data is described. The model has provided explanations for the effects of steel composition, water velocity, depth of immersion and seawater salinity and also has facilitated new interpretations of data for long-term pitting corrosion. This paper briefly overviews these developments and refers to some typical applications, including marine corrosion of ship ballast tanks, corrosion of sheet piling in harbours and corrosion of offshore platform mooring chains. INTRODUCTION Physical infrastructure plays a major role in the most mod- ern societies. So-called whole-of-life assessments increasingly are being used for decision processes. Such algorithms require mod- els of sufficient rigor and robustness to represent (a) the demands or loadings expected to be placed on the system; (b) the ways in which the system may respond; and (c) prediction of likely future response, including deterioration and effectiveness of repairs. Consistent with modern decision theory, the models required for (a) and (b) are probabilistic (Melchers, 1998). Until recently, mod- els for (c) were largely ignored. Most infrastructure has expected lives of several decades. As argued previously (Melchers, 2005), the only way such pre- dictions can be made is to invoke a combination of scientific understanding of deterioration processes and sound mathematical modeling. The present paper is concerned with the development of corrosion models, particularly for longer-term exposures. Despite good maintenance regimes, and the availability of protective coat- ings and of various forms of cathodic protection, field evidence shows that existing infrastructure often shows signs of corrosion, particularly in severe environments, such as for offshore facilities, along marine coastlines and in harbors. Below some general comments are made about the basic re- quirements for models with prediction capabilities. This is fol- lowed by a a description of the development of probabilistic/ mathematical models for corrosion loss and for maximum pit depth. The main characteristics of the model are presented, and its theoretical underpinnings reviewed. Attention is then turned to pitting corrosion and the representation of uncertainty about max- imum pit depth using Extreme Value distributions. Some ongoing application-based research projects are then referenced, followed by concluding remarks. Jin S. Chung Award Lecture, 2012. Received October 1; revised manuscript received by the editors October 26, 2012. The original version (prior to the final revised manuscript) was presented at the 22nd International Offshore and Polar Engineering Conference (ISOPE- 2012), Rhodes, Greece, June 17–22, 2012. KEY WORDS: Corrosion, steel, modeling, marine, microbiology. MODELS AND MODELING Models may be considered as mathematical constructs that, at some level of abstraction, represent the phenomenon of interest. Typically models are tailored to their application and used for spe- cific purposes. They also need to provide meaningful answers. A statement such as obtained from electrochemical potentials, “Cor- rosion is highly likely,” may be of some general interest but is of limited technical value for estimating the rate of corrosion. Various approaches to model development exist. For some, a model is the outcome of an attempt to establish correlation between available data for corrosion loss or pit depth and data for the various factors believed to be of influence. Typically it pro- vides a best fit curve (or surface) through data points. There is a long history of this approach in the atmospheric corrosion lit- erature. However, any objective evaluation repeatedly shows high levels of uncertainty, low confidence levels or poor correlation coefficients (e.g. Dean and Reiser, 2002). The reason is that these models are empirical and lack reference to fundamental under- standing of the processes involved. Also, extrapolation is ques- tionable since there is no theoretical basis for it. Potentially more powerful and therefore more interesting are models based on fundamental principles and the calibrating of these models to actual (particularly field) data. The tenuous link between corrosion measured in the laboratory and in the real world has been commented upon many times, over many years (e.g. US-ONR, 2007). Theory-based models are developed from reasonable hypothe- ses of the processes likely to be involved, including physical, chemical, electrochemical and, if appropriate, microbiological influences. Calibrating a theory-based model implies that each set of data is treated as a sample set of all possible observations. In essence this approach asks: Can the data be interpreted as consistent with the model? It is implicit that some degree of uncertainty always is associated with each data point. (It is also possible for some data points to be called wrong owing to errors in observation or data processing.) As in science generally, the test is whether it is possible to disprove the model; if so, model refinement (or aban- donment) is required. And then the process can be repeated. This is a completely different way of proceeding than the traditional idea that the data are somehow sacrosanct, and that the data points

Transcript of Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and...

Page 1: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

International Journal of Offshore and Polar Engineering (ISSN 1053-5381) http://www.isope.org/publicationsCopyright © by The International Society of Offshore and Polar EngineersVol. 22, No. 4, December 2012, pp. 257–263

Jin S. Chung Award Lecture, 2012

Modeling and Prediction of Long-Term Corrosion of Steel in Marine Environments

Robert E. MelchersCentre for Infrastructure Performance and Reliability, The University of Newcastle

Newcastle, NSW, Australia

This paper reviews recent developments in the prediction of the likely future corrosion losses and of the maximum pitdepth for steels exposed to marine environments. A robust mathematical model based on corrosion science principles andcalibrated for immersion conditions to an extensive range of literature and new data is described. The model has providedexplanations for the effects of steel composition, water velocity, depth of immersion and seawater salinity and also hasfacilitated new interpretations of data for long-term pitting corrosion. This paper briefly overviews these developments andrefers to some typical applications, including marine corrosion of ship ballast tanks, corrosion of sheet piling in harboursand corrosion of offshore platform mooring chains.

INTRODUCTION

Physical infrastructure plays a major role in the most mod-ern societies. So-called whole-of-life assessments increasingly arebeing used for decision processes. Such algorithms require mod-els of sufficient rigor and robustness to represent (a) the demandsor loadings expected to be placed on the system; (b) the ways inwhich the system may respond; and (c) prediction of likely futureresponse, including deterioration and effectiveness of repairs.Consistent with modern decision theory, the models required for(a) and (b) are probabilistic (Melchers, 1998). Until recently, mod-els for (c) were largely ignored.

Most infrastructure has expected lives of several decades. Asargued previously (Melchers, 2005), the only way such pre-dictions can be made is to invoke a combination of scientificunderstanding of deterioration processes and sound mathematicalmodeling. The present paper is concerned with the development ofcorrosion models, particularly for longer-term exposures. Despitegood maintenance regimes, and the availability of protective coat-ings and of various forms of cathodic protection, field evidenceshows that existing infrastructure often shows signs of corrosion,particularly in severe environments, such as for offshore facilities,along marine coastlines and in harbors.

Below some general comments are made about the basic re-quirements for models with prediction capabilities. This is fol-lowed by a a description of the development of probabilistic/mathematical models for corrosion loss and for maximum pitdepth. The main characteristics of the model are presented, andits theoretical underpinnings reviewed. Attention is then turned topitting corrosion and the representation of uncertainty about max-imum pit depth using Extreme Value distributions. Some ongoingapplication-based research projects are then referenced, followedby concluding remarks.

Jin S. Chung Award Lecture, 2012. Received October 1; revisedmanuscript received by the editors October 26, 2012. The originalversion (prior to the final revised manuscript) was presented at the22nd International Offshore and Polar Engineering Conference (ISOPE-2012), Rhodes, Greece, June 17–22, 2012.

KEY WORDS: Corrosion, steel, modeling, marine, microbiology.

MODELS AND MODELING

Models may be considered as mathematical constructs that, atsome level of abstraction, represent the phenomenon of interest.Typically models are tailored to their application and used for spe-cific purposes. They also need to provide meaningful answers. Astatement such as obtained from electrochemical potentials, “Cor-rosion is highly likely,” may be of some general interest but is oflimited technical value for estimating the rate of corrosion.

Various approaches to model development exist. For some,a model is the outcome of an attempt to establish correlationbetween available data for corrosion loss or pit depth and data forthe various factors believed to be of influence. Typically it pro-vides a best fit curve (or surface) through data points. There isa long history of this approach in the atmospheric corrosion lit-erature. However, any objective evaluation repeatedly shows highlevels of uncertainty, low confidence levels or poor correlationcoefficients (e.g. Dean and Reiser, 2002). The reason is that thesemodels are empirical and lack reference to fundamental under-standing of the processes involved. Also, extrapolation is ques-tionable since there is no theoretical basis for it.

Potentially more powerful and therefore more interesting aremodels based on fundamental principles and the calibrating ofthese models to actual (particularly field) data. The tenuous linkbetween corrosion measured in the laboratory and in the realworld has been commented upon many times, over many years(e.g. US-ONR, 2007).

Theory-based models are developed from reasonable hypothe-ses of the processes likely to be involved, including physical,chemical, electrochemical and, if appropriate, microbiologicalinfluences.

Calibrating a theory-based model implies that each set of datais treated as a sample set of all possible observations. In essencethis approach asks: Can the data be interpreted as consistent withthe model? It is implicit that some degree of uncertainty alwaysis associated with each data point. (It is also possible for somedata points to be called wrong owing to errors in observation ordata processing.) As in science generally, the test is whether it ispossible to disprove the model; if so, model refinement (or aban-donment) is required. And then the process can be repeated. Thisis a completely different way of proceeding than the traditionalidea that the data are somehow sacrosanct, and that the data points

Page 2: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

258 Modeling and Prediction of Long-Term Corrosion of Steel in Marine Environments

are the only pieces of information from which to construct a so-called model.

The introduction of probability allows both for the uncertaintyassociated with the level of abstraction used for constructing themodel, and for the inevitable uncertainties in data (Melchers,1998). An appropriate probabilistic formulation for corrosion lossc4t5 as a function of exposure period t is:

c4t5= b4t5 · f 4t5+ �4t5 (1)

where f 4t5 is the mean-value function for corrosion loss, �4t5 isa zero-mean error function, and b4t5 is a bias function with unitmean, all as functions of t. If the mean value function f 4t5 is apoor choice, the relationship to real data will be poor, and therewill be a large error term �4t5. Conversely, a high-quality modelshould produce a good fit to data and leave only a small errorterm. It follows that model development should focus primarilyon a sufficiently good-quality mean value function. This approachhas been explored intensely in recent years for marine corrosion ofsteel (Melchers, 2003a, b). In the following, attention is confinedmainly to the mean value function.

MODELS FOR CORROSION

For engineers and for asset managers interested in the predic-tion of the future rate of corrosion, the problem may be illustratedusing the hypothetical but distinctly nonlinear corrosion loss curveshown in Fig. 1a. At time ti, the estimated corrosion loss is c4ti5shown at A. Of interest is the likely rate of corrosion r4ti5 fort > ti, i.e. projection AC. Simple linear extrapolation from theorigin O through the current observation A (i.e. the line OAB)usually will overestimate the rate. Knowledge of the underlyingmodel for the corrosion process, i.e. OAC, allows the likely futurecorrosion loss rate r4ti5 (i.e. AC) to be estimated. For this, knowl-edge of the model (OA) is essential.

From Fig. 1a and from Eq. 1, it is clear that in developing a(probabilistic) corrosion or pit depth model (see also Fig. 1b) thefirst priority should be the mean value function f 4t5. The exist-ing applied corrosion literature, reflected in many well-knowntexts, suggests various corrosion rates, apparently constant in time(Jones, 1996). In engineering and naval architecture this is still thecommon understanding although various other models have beenproposed. A review of these is available (Melchers 2008a). How-ever, for most of these the actual corrosion loss measurements,such as for ships plating, for example, have very high uncertain-ties (Guedes Soares et al., 2006). In the atmospheric corrosionliterature, the almost universally accepted model is the so-calledpower-law model, which has corrosion loss c4t5 as a function ofexposure time t represented by Feliu et al. (1993):

c4t5=AtB (2)

where A and B are constants obtained from fitting Eq. 2 to theavailable data. Despite its long history, it has been shown only

Fig. 1 (a) Corrosion model 0AC and prediction of corrosion rates(b) showing estimation of uncertainty in model from various data

recently that the derivation of this model from diffusion princi-ples involves a number of major assumptions (Melchers, 2003c).Good-quality atmospheric corrosion data deviate from Eq. 2 in aconsistent manner, and A and B are functions of the length of thedata record (Melchers, 2009). This is inconsistent with a good-quality model.

A MORE ROBUST MODEL

To build a model based on corrosion theory it is necessary tohave some understanding of the processes involved. The corrosionof steel in seawater or in coastal marine conditions is corrosionin wet environments. Corrosion can occur only in the presenceof water and with an electron acceptor such as oxygen. Corro-sion is an electro-chemical reaction, distinguished from a purelychemical reaction, as it involves the physical transport of electronsfrom 1 site (anode) to another (cathode), with a correspondingpart of the chemical section occurring at these sites. Of course thesites may be very close to each other (nanometers) or much fur-ther apart, provided there is electrical conductivity between them.Water provides such an electrolyte but steel has much greaterconductivity. The initiation of corrosion invariably is by pittingat a micro-scale, set up by very small differences in the surfacetopography and grain structure of a steel surface. This permitsvery small (electrical) potential differences to develop and even-tually lead to pit initiation (Jones, 1996; Burstein et al., 2004).The kinetic reactions involved typically are very fast, much fasterthan the transport of various species through processes such asdiffusion and effusion. In most cases these latter processes controlthe rate of the corrosion reaction(s). In real seawater invariablythere also will be a colonization of the steel surface by biofilms(Little and Lee, 2007). Usually these provide environments suit-able for colonization by microorganisms such as bacteria presentin seawater.

The net result is that, usually within hours of first exposure,very small pits initiate on a steel surface (Burstein et al., 2004).These pits grow quickly in depth to some 100 microns (Butler etal., 1972) within days of exposure and later they grow in width.Many early pits stop growing soon after formation and are essen-tially overtaken by others. It follows that microscopic examina-tion of a corroded steel surface invariably reveals a complex mixof larger and smaller pits as well as unaffected regions (cath-odes), at least for some time. Fig. 2 shows a schematic view ofthe development of pitting with increased exposure time. In par-ticular it shows that the initial major pits stop growing in depth

Fig. 2 Development of pitting as a function of time, showing ini-tial pitting broadening out to form a rough plateau on which newpits then form (Jeffrey and Melchers, 2007)

Fig. 3 Successive views of pitted surface of steel coupons

Page 3: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

International Journal of Offshore and Polar Engineering, Vol. 22, No. 4, December 2012, pp. 257–263 259

Fig. 4 Data and trends for corrosion of mild steel in 4 differenttropical exposure environments

but amalgamate to form shallow depressions, and that later newpitting develops on the depression surfaces. The result is the for-mation of a series of depressions and a range of pit depths andsizes. This shows that the growth of pit depth is not a contin-uous process, at least for longer exposures. Fig. 3 shows somemicroscope photographs (at the same scale) of the progression ofpitting (Jeffrey and Melchers, 2007). This pattern of behavior forpit growth and development contrasts with the conventional wis-dom, which assumes a continuous single functional process forpit depth development.

It follows immediately from these longer-term field observa-tions that so-called uniform or general corrosion is an erroneousbut still convenient practical concept. In practice it usually isobtained from the changes in mass loss of nominally identicalcoupons exposed for different periods. Fig. 4 shows typical curvesfor the development of mass loss with time, for 4 different marineexposure environments (Southwell et al., 1958; Forgeson et al.,1960). Because of its practical importance for structural strengthloss and because of its central position in conventional corrosionresearch and testing, the uniform corrosion model is used here formodel development.

The environmental factors E that influence even the apparentlysimplest case and the corrosion of fully immersed steel in sea-water are summarized in Table 1, together with their importance(Schumacher, 1979). Also of potential importance are: (a) steelcomposition, (b) size and orientation of the steel item, (c) sur-face roughness of the steel, and (d) shielding of the object byadjacent structures. The model development described below doesnot consider all these factors but is confined to a specific andpractically useful situation: mild steel with normal or acid pick-led finish, exposed in coastal seawater at various locations, andwithout obvious water pollution. This eliminates, or very muchequalizes, most factors in Table 1, such as water quality, watervelocity and wave action, oxygenation—since coastal seawater isnear saturated, and pH—since pH varies very little between sea-waters. The major variables left are (a) water temperature and (b)microbiological influences. The former is simplified to the annual

Factor Importance Factor Importance

bacteria high pollutants variesbiomass low temperature highoxygen supply short term pressure nocarbon dioxide little suspended solids nosalinity not by itself wave action highpH high water velocity highcarbonate solubility low

Table 1 Selection of factors E known to influence corrosion

Fig. 5 Model for corrosion showing (a) sequential phases, modelparameters (r0, ca, ta, cs , rs5 and early and later influence of nutri-ent availability levels; and (b) effect of increasing water tempera-ture and reducing dissolved oxygen concentration (Melchers andJeffrey, 2008)

average seawater temperature, and the latter is considered only interms of nutrients necessary for microbiological activity; if theseare absent or severely limited, microbiological activity cannot beimportant.

The principal features of the model are summarized in Fig. 5.The underlying model (base model) is based on the idealizationthat the corrosion process changes as rust layers build up on themetal surface, and that these are not removed to any major extent.This results in a model composed of a sequence of phases (sub-processes) labeled 0–4. The first phase (0) is quite short (days)and includes both bacterial and electrochemical processes. Scien-tifically of great interest, it is of limited engineering importance asit has little influence over longer-term corrosion. In phase 1, the(instantaneous or tangent) rate of corrosion is limited by so-calledconcentration control, in which the rate of transport of oxygenfrom the surrounding (sea)water or moisture layers is controlledby the rate at which it diffuses under the concentration gradientestablished next to the metal surface (Jones, 1996). In phase 2,the rate of corrosion is controlled by the slowly decreasing rate ofdiffusion of oxygen through the increasing thickness of the rustlayers. This is analogous to the model first proposed by Tammann(1923) and can be represented mathematically as a diffusion pro-cess in a simplified form (Melchers, 2003c).

As is well-known, under wet corrosion conditions the anodicreaction is the dissociation of ferrous metal from ferrous ionsin which metallic iron oxidizes (i.e. becomes an electron donor)(Jones, 1996). The overall reaction is:

2Fe + O2 + 2H2O → 2Fe(OH)2 (3)

Note that the reaction does not include the chlorides in the sea-water. Even for chlorid-containing environments, the rate-limitingstep is the rate of diffusion of oxygen to the corroding surface(Evans, 1960; Brasher, 1967; Melchers, 2006). However, the pres-ence of chlorides in seawater does permit some other reactionsinvolving oxygen as the electron acceptor. One is the produc-tion of ferrous chloride, summarized by the equilibrium reaction(Sharland and Tasker, 1988):

Fe2+ + 2Cl− ⇔ FeCl2 (4)

When exposed to air or oxygen FeCl2 oxidizes quickly to FeCl3

and then to the chloride containing �-FeOOH and producing also�-FeOOH (Waseda and Suzuki, 2006). As a result, FeCl2 is sel-dom detected in rusts exposed to the atmosphere.

Provided the data sets available are sufficiently rich, eventu-ally the corrosion loss (and maximum pit depth) trends show an

Page 4: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

260 Modeling and Prediction of Long-Term Corrosion of Steel in Marine Environments

upswing in the instantaneous corrosion rate. This is the early partof phase 3 in Fig 5a. The latter part of phase 3 has a decliningcorrosion rate, eventually leading to an almost constant corrosionrate, represented by phase 4.

Phases 3 and 4 have been associated with biologically influ-enced corrosion (Lee et al., 1995; Jeffrey and Melchers, 2003).This arises from the development of anoxic conditions withinmicro-niches in the rust layers, particularly those close to the cor-roding metal. These provide conditions suitable for colonization,for example, by the obligatory anaerobic sulfate-reducing bacte-ria (SRB). In field conditions, as distinct from laboratory studies,bacterial colonization and activity invariably involve broad com-munities of microorganisms (Hamilton, 1985). This cohabitationprovides mutual support for energy transfer and nutrient avail-ability, although typically the community requires nutrients andenergy inputs from the external environment. It follows that therate of nutrient supply may play a rate-controlling role in corro-sion resulting from bacteria such as SRB.

Bacteria also may play a role in corrosion during phases 0 and1 in the anoxic niches that can develop adjacent to the corrodingsteel surface (Cragnolino and Tuovinen, 1984). This is a transientphenomenon (Melchers, 2010), as indicated in Fig. 5a. Duringthese early phases there is little to impede the diffusion of nutri-ents to the bacteria from the (supposedly nutrient-rich) externalenvironment. This is not the case in phases 3 and 4, when a con-siderable rust layer is present between the bacteria and the exter-nal environment.

Anoxic conditions also may permit wet corrosion under thecathodic hydrogen evolution reaction. The mechanism withinwhich these reactions can occur at the corroding interface for theearly part of phase 3 has been postulated as locally similar tothose that cause crevice and pitting corrosion, both of which canproceed very quickly in time (Melchers, 2011). Under these con-ditions the corrosion rate is controlled, initially at least, by therate of (outward) diffusion of molecular hydrogen. For a giventhickness and diffusivity of the rust layers present at the end ofphase 2, this is much faster than the inward diffusion of oxygenbecause of the much smaller size of molecular hydrogen. The rel-ative rates are shown schematically as ra and rb respectively inFig 5a. The likely long-term corrosion rate limiting mechanismsare the inward diffusion of water or the outward movement offerrous ions. The details remain to be determined.

The hydrogen-reduction reaction (Fig. 7) produces (black) mag-netite. Usually it is found at the base of older rusts, highly adher-ent to the underlying metal although it may be quite weak ifhydrated. It has a characteristic smell very different from theblack corrosion product hydrogen sulfide FeS (or similar sulfideproducts) produced by microbiological activity. FeS results fromthe metabolism of hydrogen and the sulfates abundantly presentin seawater. Typically it is soft and slimy and easily removed to

Fig. 6 Data and calibrated trend curves for model parameters r0

and ta (Fig. 5a)

Fig. 7 Schematic representation of effects of small amounts ofalloying materials on corrosion loss; effects are different for vari-ous parts (phases) of model. Also, higher carbon content appearsto increase both cs and rs (Melchers, 2003d).

Factor Summary influence

Water velocity Increases corrosion in the first few weeks as rustsbuild up and thereafter little additional effect.

Salinity Important, but effect also influenced by waterhardness. Some soft fresh waters are highlycorrosive.

Timing Steel exposed, say in autumn or winter, has ini-tially a slower rate of corrosion. This movesmodel (to the right) in time by about 6 months.

Pressure No evidence that it has any noticeable effect.Depth Depth effects result mainly from temperature,

DO and nutrient levels.Size/Area Precise shape or area involved in corrosion not

important for moderate-sized objects, i.e. withdistances insufficient for the water propertiesto change.

Table 2 Some factors influencing seawater corrosion of steel

reveal bright metal underneath. FeS is not always observed butmagnetite is common (Waseda and Suzuki, 2006).

Fig. 5 summarizes, schematically, the influence of the mainfactors. Fig. 6 shows the calibration curves for just 2 of the param-eters r0 and ta. As expected of field data collected under uncon-trolled conditions, there is variability in the data (Lee et al., 2010)but the trends are quite clear.

Since each phase of the model describes a different principalcorrosion mechanism, each may be expected to respond differ-ently to steel composition (Fig. 7) and to various environmentalconditions, including water temperature, dissolved oxygen, salin-ity, carbonate and sulfate contents. Indeed this has been observedin careful data analyses. This also has facilitated extension of themodel to conditions other than natural, unpolluted, largely quies-cent coastal seawater. Table 2 shows the influence of some otherinfluencing factors.

PITTING CORROSION

Pitting is of interest for the perforation of pipe walls andtanks. This implies that the maximum pit depth is of inter-est. Measurement of maximum pit depths on multiple mild-steelcoupons exposed for various periods of immersion in aerated nat-ural sweater has shown, repeatedly, a distinctly nonlinear, bimodaltrend (e.g. Fig. 8). The trend curve depends on the exposure con-ditions, a little on steel composition, but mainly on water temper-ature (Melchers, 2004a). The bimodal pit depth trend curve shownin Fig. 8 is similar to the model for corrosion loss. This is notsurprising as pitting underlies also the uniform corrosion process.

Page 5: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

International Journal of Offshore and Polar Engineering, Vol. 22, No. 4, December 2012, pp. 257–263 261

Fig. 8 Maximum pit depths as a function of exposure period fornatural coastal seawater (Taylors Beach, Australia)

In applications, interest lies mainly in the maximum pit depththat may occur for larger steel surfaces, or equivalently, overlonger exposure periods. The conventional approach is to assumethat the statistics of maximum observed pit depths may be rep-resented in terms of Extreme Value (EV) statistics (Galambos,1987). In fact, pitting corrosion usually is considered one of theprime applications of EV analysis.

In brief, the maximum of independent maxima may be con-sidered distributed according to the well-known Gumbel ExtremeValue distribution (for maxima) (Galambos, 1987). Fig. 9 showsan example of a so-called Gumbel plot. It is a special probabil-ity paper in which the cumulative probability distribution, shownon the left axis, is distorted in such a way that data that are trulyGumbel distributed plot as a straight line. (This is similar in con-cept to the so-called Normal probability plot). The left axis con-ventionally is labeled using the standardized variable w, definedbelow. The equivalent cumulative probability (i.e. the probabilitythat the variable of interest is less than the value on the horizontalaxis) is shown on the right vertical axis.

For pitting, each pit depth is assumed to be an independentmaximum (or, can be so assumed in the limit as the number of pitsbecomes large). Then, on the standard Gumbel plot, Fig. 9, thestandardized variable for the maxima is given by w = 4y− u5�which is defined through the cumulative distribution function

Fig. 9 Gumbel plot with maximum pit depth data at various expo-sure periods for 18 surfaces of mild-steel plate coupons continu-ously immersed in Pacific Ocean seawater (Melchers, 2004b)

Fig. 10 Gumbel plot of Fig. 9 with each data set is off-set by 0.5mm for clarity; straight light lines through data are Gumbel trendsand comparison to data shows that data trends are not linear asassumed in Fig. 9 (Melchers, 2008b).

(CDF) for y, i.e. by FY 4y5 and its probability density function(PDF) fY 4y5 as:

FY 4y5= FW 64y− u5�7 with FW 4w5= exp4−e−w5 (5)

fY 4y5= �fW 64y− u5�7 (6)

Here u and � respectively are the mode and slope of the Gum-bel distribution and are related to the mean �Y and standard devi-ation �Y through �Y = u+ 101396/� and �Y = 0040825�/�.

A number of standard techniques exist to plot the data from adata set on a Gumbel plot (Galambos, 1987). Fig. 9 shows 6 setsof data, each one for a different period of exposure. It is seenthat straight lines can be fitted through each data set, indicatingthe conventional Gumbel trend. It is seen also that the slopes ofthe lines (cf. � in Eqs. 5 and 6) increase with longer exposure,indicating greater variability in pit depth with increasing exposureperiod. Closer examination of maximum pit depth data for longerperiods of exposure has shown that the data are not particularlyclosely linear, but that they have a distinctive characteristic trend,also seen for other data sets (Melchers, 2008b). Fig. 10 shows themodified trends.

The most important observation in Fig. 10 is that for exposures> 105 years and for the deeper pits (i.e. those above and to theright of the line AA), the extreme value distribution that best fitsthose data is the Frechet EV distribution (Galambos, 1987). It isevident that there can be very large differences between Gumbeland Frechet in the probabilities associated with a given depth ofpitting. This may have important practical implications.

APPLICATIONS

The above principles and models have been applied to the cor-rosion inside water ballast tanks for ships (Gudze and Melchers,2008), the corrosion of steel piles in harbors by so-called acceler-ated low water corrosion (Melchers and Jeffrey, 2009), the corro-sion of mooring chains for offshore floating production and stor-age platforms (Fontaine et al., 2012), and for the internal corrosionof water injection pipelines used in the oil industry (Comãnescuet al., 2012).

CONCLUSIONS

A review of data and influencing factors for the corrosion ofsteels in marine environments has shown that corrosion loss as aprocess is best represented by the bi-model model that represents

Page 6: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

262 Modeling and Prediction of Long-Term Corrosion of Steel in Marine Environments

the changes that occur as corrosion develops, including microbi-ological effects, both on mass loss and for pitting. The represen-tation of maximum pit depth statistics requires consideration ofthese changes. The conventional Gumbel extreme value distribu-tion ignores these changes and may not be adequate for predic-tion. Several applications using these approaches are referenced.

REFERENCES

Brasher, Dora M (1967). “Stability of the Oxide Film on Metalsin Relation to Inhibition of Corrosion—I: Mild Steel in thePresence of Aggressive Ions,” Brit Corr J, Vol 2, pp 95–103.

Burstein, GT, Liu, C, Souto, RM, and Vines, SP (2004). “Originsof Pitting Corrosion,” Corr Eng, Sci and Tech, Vol 39, No 1,pp 25–30.

Butler, G, Stretton, P, and Beynon, JG (1972). “Initiation andGrowth of Pits on High-Purity Iron and Its Alloys withChromium and Copper in Neutral Chloride Solutions,” BritCorr J, Vol 7, No 7, pp 168–173.

Comanescu, I, Melchers, RE, and Taxén, C (2012). “CorrelationBetween MIC and Water Quality, Pigging Frequency and Bio-cide Dosing in Oil Field Water Injection Pipelines,” Eurocorr2012, Istanbul.

Cragnolino, G, and Tuovinen, OH (1984). “The Role of Sulphate-reducing and Sulphur-oxidizing Bacteria on the Localized Cor-rosion of Iron-Base Alloys: A Review,” Int Biodeterioration,Vol 20, No 1, pp 9–26.

Evans, UR (1960). The Corrosion and Oxidation of Metals: Sci-entific Principles and Practical Applications,” London: EdwardArnold.

Feliu, S, Morcillo, M, and Feliu, S Jr (1993). “The Predictionof Atmospheric Corrosion from Meteorological and PollutionParameters—1: Annual Corrosion;—2: Long-Term Forecasts,”Corrosion Sci, Vol 34, No 3, pp 403–422.

Fontaine, E, Potts, AE, Kai-tung, Ma, Arredondo, AL, and Melch-ers, RE (2012). “SCORCH JIP: Examination and Testing ofSeverely-corroded Mooring Chains from West Africa,” OffshoreTech Conf, Houston, Paper 23012.

Forgeson, BW, Southwell, CR, and Alexander, AL (1960). “Cor-rosion of Metals in Tropical Environments—Part 3: Underwa-ter Corrosion of 10 Structural Steels,” Corrosion, Vol 16, No 3,pp 105t–114t.

Galambos, J (1987). The Asymptotic Theory of Extreme OrderStatistics, 2nd Edition, Krieger, Malabar, Florida.

Guedes Soares C, Garbatov, Y, Zayed, A, Wang, G, Melchers, RE,Paik, JK, and Cui, W (2006). “Non-linear Corrosion Model forImmersed Steel Plates Accounting for Environmental Factors,”Trans 2005 Soc Naval Architects and Marine Eng, Vol 113,pp 306–329.

Gudze, MT, and Melchers, RE (2008) “Operational Based Corro-sion Analysis in Naval Ships,” Corrosion Sci, Vol 50, No 12,pp 3296–3307.

Hamilton, WA (1985). “Sulphate-reducing Bacteria and AnaerobicCorrosion,” Ann Rev Microbiol, Vol 39, pp 195–217.

Jeffrey, R, and Melchers, RE (2003). “Bacteriological Influence inthe Development of Iron Sulphide Species in Marine ImmersionEnvironments,” Corrosion Sci, Vol 45, No 4, pp 693–714.

Jeffrey, R, and Melchers, RE (2007). “The Changing Topographyof Corroding Mild Steel Surfaces in Seawater,” Corrosion Sci,Vol 49, pp 2270–2288.

Jones, D (1996). Principles and Prevention of Corrosion, 2nd Edi-tion, Upper Saddle River, New Jersey, Prentice-Hall.

Lee, JS, RI, Ray, and Little, BJ (2010). “The Influence of Experi-mental Conditions on the Outcome of Laboratory InvestigationsUsing Natural Coastal Seawaters,” Corrosion, Vol 66, No 1,105001, pp 1–6.

Lee, W, Lewandowski, Z, Nielsen, PH, and Hamilton, WA (1995).“Role of Sulfate-reducing Bacteria in Corrosion of Mild Steel:A Review,” Biofouling, Vol 8, pp 165–194.

Little, BJ, and Lee, JS (2007). Microbiologically Induced Corro-sion, John Wiley & Sons, New York.

Melchers, RE (1998). Structural Reliability Analysis and Predic-tion, Chichester, John Wiley.

Melchers, RE (2003a). “Modeling of Marine Immersion Corro-sion for Mild and Low Alloy Steels—Part 1: PhenomenologicalModel,” Corrosion, Vol 59, No 4, pp 319–334.

Melchers, RE (2003b). “Modeling of Marine Immersion Corro-sion for Mild and Low Alloy Steels—Part 2: Uncertainty Esti-mation,” Corrosion, Vol 59, No 4, pp 335–344.

Melchers, RE (2003c). “Mathematical Modelling of the Diffu-sion Controlled Phase in Marine Immersion Corrosion of MildSteel,” Corrosion Sci, Vol 45, No 5, pp 923–940.

Melchers, RE (2003d). “Effect on Marine Immersion Corrosion ofCarbon Content of Low Alloy Steels,” Corrosion Sci, Vol 45,No 11, pp 2609–2625.

Melchers, RE (2004a). ‘Pitting Corrosion of Mild Steel in MarineImmersion Environment—1: Maximum Pit Depth,” Corrosion,Vol 60, No 9, pp 824–836.

Melchers, RE (2004b). “Pitting Corrosion of Mild Steel in MarineImmersion Environment—Part 2: Variability of Maximum PitDepth,” Corrosion, Vol 60, No 10, pp 937–944.

Melchers, RE (2005). “The Effect of Corrosion on the StructuralReliability of Steel Offshore Structures,” Corrosion Sci, Vol 47,No 10, pp 2391–2410.

Melchers, RE (2006). “Modelling Immersion Corrosion of Struc-tural Steels in Natural Fresh and Brackish Waters,” CorrosionSci, Vol 48, No 12, pp 4174–4201.

Melchers, RE (2008a). “Development of New Applied Models forSteel Corrosion in Marine Applications Including Shipping,”Ships and Offshore Struct, Vol 3, No 2, pp 135–144.

Melchers, RE (2008b). “Extreme Value Statistics and Long-TermMarine Pitting Corrosion of Steel,” Prob Eng Mech, Vol 23,pp 482–488.

Melchers, RE (2009). “Experiments, Science and Intuition in theDevelopment of Models for the Corrosion of Infrastructure,”Proc ACA Conf Corrosion & Protection, Coffs Harbour.

Melchers, RE (2010). “The Transient Early and the Longer-TermInfluence of Bacteria on the Marine Corrosion of Steel,” CorrEng Sci Tech, Vol 45, No 40, pp 257–261.

Melchers, RE (2011). “Modelling Long-Term Corrosion of MildSteel in Unpolluted Fresh and Sterile Seawaters,” Proc 18th IntCorr Congress, Perth, Aust Corr Assoc, CDRom, Paper 156.

Melchers, RE, and Jeffrey, R (2008). “The Critical Involvementof Anaerobic Bacterial Activity in Modelling the CorrosionBehaviour of Mild Steel in Marine Environments,” Elec Acta,Vol 54, pp 80–85.

Melchers, RE, and Jeffrey, R (2009). “The Influence of WaterNutrient Concentration on the Marine Corrosion of Long SteelStrips,” NACE Conf, San Antonio, Texas, TEG 187X, Paper10223.

Page 7: Modeling and Prediction of Long-Term Corrosion of Steel in ... · PDF fileModeling and Prediction of Long-Term Corrosion of Steel in Marine Environments ... where f4t5 is the mean-value

International Journal of Offshore and Polar Engineering, Vol. 22, No. 4, December 2012, pp. 257–263 263

Schumacher, M, editor (1979). Seawater Corrosion Handbook,Park Ridge, New Jersey; Noyes Data Corp.

Sharland, SM, and Tasker, PW. (1988) “A Mathematical Model ofCrevice and Pitting Corrosion—I: The Physical Model,” Cor-rosion Sci, Vol 28, No 6, pp 603–620.

Southwell, CR, Forgeson, BW, and Alexander, AL (1958). “Cor-rosion of Metals in Tropical Environments—Part 2: Atmo-

spheric Corrosion of 10 Structural Steels,” Corrosion, Vol 14,No 9, pp 53–59.

Tammann, G (1923). Lehrbuch der Metallographie, 2nd Edition,Leipzig.

US-ONR (2007). Second World Congress on Corrosion in theMilitary, Naples, US Office of Naval Research (Global).

Waseda, Y, and Suzuki, S (editors) (2006). Characterization ofCorrosion Products on Steel Surfaces, Springer, Berlin.