Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Studies in Surface Science and Catalysis 151

R. Vazquez-Duhalt and R. Quintero-Ramirez (Editors)


© 2004 Published by Elsevier B.V. 193

Chapter 7

Biocorrosion

H.A. Videla ~ and L.K. Herrera b

aDepartment of Chemistry. College of Pure Sciences, INIFTA, University of


La Plata, Argentina

bFaculty of Engineering, University of Antioquia, Medellin, Colombia

1. INTRODUCTION

Microorganisms influence corrosion by changing the electrochemical


conditions at the metal/solution interface. These changes may have different
effects, ranging from the induction of localized corrosion, through a change in
the rate of general corrosion, to corrosion inhibition [1 ]. Any biological effect
that either facilitates or impedes one of the anodic or cathodic reactions of the
corrosion process, or permanently separates anodic and cathodic sites, will
increase corrosion. For instance, stimulation of the anodic reaction by acidic
metabolites or the cathodic reaction by microbial production of a cathodic
reactant like hydrogen sulfide, the breakdown of protective films or the increase
in conductivity of the liquid environment will enhance corrosion.
Although the electrochemical nature of corrosion remains valid for
microbial corrosion, the participation of the microorganisms in the process
induces several unique features, mainly the modification of the metal/solution
interface by biofilm formation. Biofilms affect the interaction between metal
surfaces and the environment, not only in biodeterioration processes like
corrosion, but also in several biotechnological processes applied to materials
recovery and handling [2]. Thus, the key to the alteration of conditions at a
metal surface, and hence the enhancement or inhibition of corrosion is the
formation of a biofilm [3]. This can be considered as a gel containing 95% or
more water, made of a matrix of exopolysaccharidic substances (EPS) in which
microbial cells, and inorganic detritus are suspended [4].
Biofilm formation is the result of an accumulation process -not
necessarily uniform in time or space [5]- that starts immediately after metal
immersion in the aqueous environment. A thin film (approximately 20-80 nm
194

thick), due to the deposition of inorganic ions and high relative molecular mass
organic compounds, is formed in a first stage. This initial film is able to alter the
electrostatic charges and wettability of the metal surface facilitating its further
colonization by bacteria. In a short time, (minutes or hours according to the
aqueous environment where the metal is immersed), microbial growth and EPS
production results in the development of a biofilm. This biofilm is a dynamic
system and the different transport processes and chemical reactions occurring at
the biofouled interface will take place from now on, through the biofilm
thickness [6] (Fig. 1).
Microbial colonization of metal surfaces drastically changes the classical
concept of the electrical interface commonly used in inorganic corrosion.
Important changes in the type and concentration of ions, pH values and
oxidation-reduction potential are induced by the biofilm, altering the passive or
active behavior of the metallic substratum and its corrosion products, as well as
the electrochemical parameters used for assessing corrosion rates [7].
Simultaneously with the biological changes that lead to biofilm
accumulation, a sequence of inorganic changes takes place at the metal surface
immediately after its immersion in an aggressive aqueous medium. This
sequence involves the process of metal dissolution and corrosion product
formation.

SUBSTRATE _ -.~

.... o o..

_.. /I--! / ..=SPERSEDCE,LSlp~TABO"TES


MASS
TRANSFER
BOUNDARY LAYER

._ "..'r~ ........... l.._! ................... L.IY. .....................


..... bit ~ OROWTH,~~Jl

DEATHI
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \

Fig. 1. Diagrammatic representation for biofilm development (from Ref. [8]).


195

Both biological and inorganic processes occur within the same time
period, but following opposite directions at the metal/solution interface.
Whereas corrosion and corrosion product accumulation occur from the metal
surface towards the solution, biofilm formation is the result of accumulation
processes directed from the bulk towards the metal surface (Fig. 2) [8]).

Fig 2. Sequence of biological and inorganic processes at a biologically and electrochemically


conditioned metal/solution interface (from Ref. [8]).

Thus, a very active interaction between the corrosion product layers and
the biofilms can be expected. The consequent corrosion behavior of the metal
will vary according to the degree of this reciprocal interaction and a concept of a
new biologically conditioned interface must be kept in mind [9]. The approach
for a sound interpretation of any microbial corrosion case must then be
interdisciplinary, and include a thorough process analysis combined with well
defined microbiological and electrochemical methodologies.
Biocorrosion has been focusing increasing attention from different
research areas in the last two decades as an answer to the demand of a wide
variety of industries. Fortunately an increasing intellectual and technical cross-
fertilization of ideas between researchers from different disciplines like
microbiology, electrochemistry and materials science has allowed a considerable
improvement in the understanding of biocorrosion to be reached.

2. SRB INDUCED CORROSION OF STEEL

Biocorrosion of carbon steel in anaerobic environments involving the


presence of SRB have been the focus of most biocorrosion research. Starting
196

with the cathodic depolarization theory (CDT) of von Wolzogen Kuhr & Van
der Vlugt [10] (Fig. 3), a copious list of papers and reviews on the anaerobic
corrosion of iron has been published [11-13]. Bacterial biofilms may develop
anaerobic regions, even in aerobic bulk water environments [2], thus allowing
SRB a very favorable environment for growth. The final result of these
processes within biofilms is to produce a wide variety of sites on the metal
surface that are markedly different from neighboring sites from a
physicochemical standpoint, thus facilitating the initiation of localized corrosion
processes.

Overall reaction 4 Fe + $ 0 4 "2-F 4 H20 3 Fe (OH) 2-k- FeS +2 OH


Metal 4 Fe ~ 4 F e z + 8 e- Anode "~Electrochemical
8 H+ + 8 e- ---~8 Had Cathode ) cell

Solution 8 H 2 0 ~ 8 H + + 8 OH- Electrolyte


8 0 4 "2 + 8 H ~ S -2 -k- 4 H 2 0
[Microbial
Microorganism ) depolarization
Corrosion products Fe -F S-2---)FeS 3 F e z +6 OH--, 3 Fe (OH)

Fig. 3. Sequence of reactions of the Cathodic Depolarization Theory. The three elements of
biocorrosion (metal/solution/microorganisms) are involved in different reactions of the whole
process.

As the localized corrosion and breakdown process is strongly dependent


of several experimental factors such as the type and concentration of aggressive
anions present in the medium and the passivating film characteristics, the effect
of sulfur anions has been studied in a series of laboratory experiments using
alkaline [14] and neutral buffered [15] solutions as well as SRB cultures in
saline media [16] under well defined experimental conditions. Taking into
account the results of these studies, a bioelectrochemical interpretation of the
biocorrosion process of carbon steel in anaerobic environments may be
summarized as follows [17]:
i) biogenic sulfides effects on carbon steel passivity breakdown is similar
to that of abiotic sulfides. The characteristics and intensity of sulfide effects on
the corrosion behavior of carbon steel is narrowly related with the nature of the
passive film already present on the metal surface (Fig. 4); ii) in neutral media
sulfide ions lead to the formation of a poorly protective film of mackinawite; iii)
the anodic breakdown of passivity would be the first stage of the corrosion
process.
197

Fig. 4. Scheme of the bioelectrochemical interpretation of the biocorrosion process of carbon


steel in anaerobic environments (from Ref. [17]).

Thus, the role of SRB may be indirect through the production of


aggressive species either as final (sulfides, bisulfides or hydrogen sulfide) or
intermediate metabolic compounds (thiosulfates, polythionates). Physico-
chemical characteristiscs of the liquid environment (pH, ionic composition,
oxygen levels) can modify the SRB effects which could eventually change from
corrosive to passivating; iv) cathodic depolarization effects attributed to SRB
hydrogenase activity or to iron sulfide films would be developed later than
passivity breakdown while corrosion process is in progress; v) the action of
biogenic sulfides can be enhanced by other aggressive anions already present in
the environment (e.g. chlorides) [18] or through microbial consortia within
adhered biofilms on the metal surface [19].
Further studies developed in our laboratory [20-23] allowed us to
conclude that SRB influenced corrosion of steel is markedly affected by the
nature and the structure of the sulfide films produced during the metal
dissolution. The environmental characteristics of the metal/biofilm/medium
interface and its surroundings (pH, ionic composition, oxygen levels, EPS
198

distribution) control the chemical and physical nature of corrosion product


layers and may change their effects on the metal behavior from corrosive to
protective. In marine environments, the impact of sulfur compounds on
corrosion is enhanced by other aggressive anions, such as the widely distributed
chlorides, already present in the medium.
The entrance of oxygen into the anaerobic environment accelerates
corrosion rate, mainly through a change in the chemical nature of iron sulfides
and elemental sulfur production. Both chemical species can provide additional
cathodic reactants to the corrosion process, acting as electron carriers between
the metal and the oxic interface within the biofilm.

2.1. Differences between biotic and abiotic media


Recently, several surface analysis techniques, electrochemical
experiments and microscopy observations were employed to clarify the role of
biotic and abiotic sulfide films in the corrosion behaviour of steel in saline
media [21, 23]. Microbiological experiments were performed under controlled
laboratory conditions using a strain of Desulfovibrio alaskensis (D. alaskensis),
known for its ability to produce EPS, isolated from a soured oil reservoir in
Alaska [24]. Atomic force microscopy (AFM) was applied to image the SRB
biofilms (Fig. 5), current transients were measured to determine the
electrochemical behaviour of steel, and SEM observations coupled with EDAX,
as well as XPS, XRD and electron microprobe analysis (EPMA) were carried
out to examine the structure and composition of biotic and abiotic sulfide films
on the surfaces of steel specimens.
The main conclusions obtained from these studies have been: i) both
abiotic and biotic iron sulfide films are related to the formation of tubercles on
the steel surface. However, for biotic solutions FeS (mackinawite) predominates,
whereas in abiotic media FeS2 (pyrite) is the major iron sulfide present (Fig. 6);
ii) the structure of the outer crust of iron sulfide acts as a barrier for the diffusion
of ions towards and from the pit cavity; iii) biotic films are more adherent to the
metal surface, while abiotic films are flaky and loosely attached; iv) the previous
history of the sulfide film may play a relevant role in the corrosion behaviour of
steel. Depending on the sulfide concentration in the medium and on the presence
or absence of biofilms and EPS, the protective characteristics of the corrosion
product may change. Biogenic layers of corrosion products can offer enhanced
protection by improving the adherence of the film to the metal, but can also
increase corrosion, inducing the presence of heterogeneities at the metal surface.
199

Fig. 5. AFM image of a SRB biofilm on AISI 316 SS (from Ref. [21]).

The differences between biotic and abiotic media containing identical


levels of corrosive compounds (i.e. iron sulfides) may be mainly attributed to the
presence of extracellular polymers and to the heterogeneities created at the metal
surface by the formation of a biofilm. In addition, the physicochemical
parameters of the corrosion product layers may change in the absence or a
presence of a biofilm, rendering these layers either protective or corrosive. This
feature can explain the observed differences in corrosion behavior of steel
between abiotic and biotic environments with respect not only to localized
pitting, but also to the hydrogen attack developed as embrittlement and crack
growth.

2.2. Hydrogen effects and SRB


Although it is widely accepted that SRB can oxidize molecular hydrogen
using their hydrogenase system, the involvement of this enzyme in the corrosion
process remains unclear. The present state of knowledge does not support the
hypothesis that the bacterial uptake of hydrogen is the rate-controlling step for
SRB-influenced corrosion of mild steel [25].
200

Fig. 6. XPS spectra of a biogenic sulfide film (top) and of an abiotic sulfide film (bottom)
(from Ref. [23]).

Hydrogen can have a considerable effect on susceptible alloys when it enters


into the atomic structure of the metal and causes embrittlement. This
phenomenon, including the acceleration of crack formation and growth, is
frequently found in industrial systems (i.e. off-shore oil platforms). In addition,
hydrogen, generated from cathodic protection can be a considerable problem in
areas of stress corrosion or corrosion-fatigue [26]. The amount of hydrogen
available to enter the steel and the ease with which it moves into the metal, is
influenced by the nature of the environment in which the steel is placed (Fig. 7).
201

r "" 1.6
E
o
Plates imbedded

-~ 1.2

9-- 1.0

o 0.8
o

'~
o 0.6

o
0.4
o
ct0
o 0.2

0
0 100 200 300 400 500 600

Time (Days)

Fig. 7. Hydrogen permeation through 50D steel, cathodically protected or unprotected and
coated or uncoated, exposed to open seawater and embedded in marine mud. 1 = Grit blasted
plus cathodic protection. 2 = CTE coating plus cathodic protection. 3 -- anti-fouling paint. 4 -
grit basted only. 5 - As received (with mill scale, uncoated) (from Ref. [29]).

In seawater, where structures are cathodically protected from corrosion,


hydrogen is available at the metal surface, particularly in anaerobic
environments. Biological activity at the surface, especially bacterial activity, can
enhance the entry of hydrogen into the metal. This action can be direct, by
producing metabolites like sulfides which encourage the entry of hydrogen into
the steel, and indirect, by both disrupting any surface films and due to the need
of increasing cathodic protection when bacteria are present [27]. SRB are the
main bacteria responsible for the enhancement of hydrogen entry as they are
active in the areas where hydrogen will be generated under cathodic protection,
i.e. anaerobic environments such as marine muds (Fig. 7).
It is indisputable that a biologically active environment is able to induce
conditions conductive to the enhancement of corrosion-fatigue crack growth and
hydrogen embrittlement by the activity of SRB [28-30]. A "biologically active
environment" involves not just the presence of microorganisms such as SRB or
a single metabolite i.e. sulfide, but the effect of all the activities of the bacteria
and their interactions with other components of the environment (e.g.
degradation by large fouling organisms, production of EPS, interactions of
202

bacteria and the metal). Indeed, the local environment surrounding the metal
surface is very different from that without bacteria, even if the same levels of
sulfide are detectable.
These areas are particularly important, as pipelines and the bottoms of the
legs of offshore platforms can be buried in marine muds where anaerobic
conditions are predominant and SRB activity is intense (Fig. 8). Moreover, sour
environments, such as those frequently found in oil production activities, are
particularly aggressive due to high levels of hydrogen available at the metal
surface or in a crack, as a consequence of sulfide poisoning of the recombination
reaction at the cathode [31 ]. In such habitats hydrogen effects can be altered by
the presence of organic molecules on the metal surface and the existence of a
biofilm with its EPS matrix. This feature can explain the differences between
general embrittlement effects (as measured by hydrogen flux) and crak tip
effects (as measured by crack growth). Embrittlement results in the general
lowering of strength of the material causing it to fail in a catastrophic way at a
lower load as it has been reported in high sulfide biological environments more
than in low sulfide abiotic environments, even though the rest of the crack
growth curve is very similar (Fig. 9) [28].
EPS and other organic molecules related to biofilms hinder dissolution
and dissociation reactions and adsorption processes in the crack. Even under low
frequency cyclic loading the crack tip opens and closes rapidly. Thus, any effect
of the environment must occur fast and organic material dragged into the crack
could have a blocking impact.
The results referred here serve to illustrate the complex nature of the
interactions between SRB biofilm and the steel. In many cases, bacterial
metabolism within the biofilm generates sulfides, and consequently, this is the
main cause of the corrosiveness of the environment. However, microbial
metabolic activity is also responsible for the release of EPS which may have a
blocking effect on hydrogen entry into the metal.

3. B I O C O R R O S I O N OF ALUMINUM ALLOYS IN F U E L / W A T E R
SYSTEMS

Microbial contamination of hydrocarbon fuels is the main cause of serious


problems concerning the quality of maintenance of the product, as well as the
corrosion of metals used during the processes of extraction, production,
distribution, and storage of the fuel.
203

Fig. 8. Composite diagram of an offshore structure showing the main sites ofbiodeterioration
problems: 1. marine fouling; 2. drill cuttings around legs; 3. oil storage and transport; 4.
water-filled legs; 5. production system; 6. seawater injection system; 7. downhole pipework;
8. reservoir problems (from Ref. [31]).

Microoganisms usually present in droplets or thin films are sufficient to


allow microbial growth and the development of biocorrosion. Because of the
electrochemical nature ofbiocorrosion an aqueous environment is also required.
Fungal and bacterial growth can also occur on side walls not necessarily
adjacent to large waterbottoms. Hyphal growth of fungi usually ramify over
fuel/water interface becoming a site for further water entrapment. As a result of
this microbial propagation action, penetration and breakdown of fuel tank
coatings and passive films occurs on the metal, leading to the onset of localized
corrosion processes generally of the type of pitting attack.
Due to its economic and technological importance, the biodeterioration of
jet fuel and subsequent biocorrosion of aluminum alloys used in the aircraft
industry has been extensively studied [32-35]. Since the end of the 70's several
of our publications have been devoted to elucidate the mechanisms affecting the
whole biocorrosion process [36-39].
204

Fig. 9. Crack growth rates of a RQT 501 steel in biologically active H2S seawater
environments; solid line: crack growth rate in seawater; dotted line: crack growth rate in 520
ppm HzS in abiotic natural seawater (from Ref. [28]).

Different types of microorganisms usually present in soils and natural


waters utilize paraffinic hydrocarbons in the range C10-C~8 (kerosene fraction)
more easily than in the range C5-C9 (gasoline fraction). In a water-flee fuel the
microbial detrimental activity is nil. Conversely, as soon as water becomes
available, the growth of microorganisms is possible by utilizing the
hydrocarbons as a source of carbon for their metabolic activity. Chemical
contaminants present in the fuel as well as in bilge water provide some kind of
nitrogen source and the necessary trace elements for growth. Therefore, the
major requirement for microbial activity is the presence of water in the storage
tank. On this respect, during the processing of hydrocarbon products is
practically impossible to avoid the existence of water. Although good
housekeeping procedures could minimize the amount of water accompanying
the fuel, the microorganisms are able to generate their own water phase for
further proliferation [40]. Microbial metabolism generate in this way, small
ecosystems which retain water in certain areas of the storage tank such as
tubercles, slime deposits or even biofilms on the metal surface. The microbial
ecology of aircraft fuel tanks and storage systems correspond to a wide diversity
205

of microorganisms, although in almost all cases the microbial contaminants


reported were fungi and bacteria [41]. The fungus Hormoconis resinae
(formerly Cladosporium resinae), and some species of the genera Aspergillus,
Penicillium and Fusarium have been isolated in jet fuels and fuel storage
systems [42]. H. resinae has also been reported as a contaminant of fuel oils for
marine and terrestrial turbine engines [43].
Due to the high rate of consumption of kerosene fuel by jet engines, even
small amounts of microbial sludges in the fuel becomes hazardous. The
combination of microorganisms and their associated water creates major
problems of fuel contamination, filter plugging, fuel gauge malfunction and fuel
tank corrosion. The corrosion of the tank wall and subsequent leakage of
hydrocarbon fuel can cause important economic losses, as well as different
troubles related with soil and underground water contamination [44]. The
corrosion attack is generally located at the tank bottom where there is an active
microbial population associated with free water.
To allow microbial growth, the environment must provide the basic
elements (carbon, oxygen, hydrogen, nitrogen, phosphorous) and some other
elements needed in much smaller amounts, but nevertheless essential for a
normal metabolism. Although carbon and hydrogen are abundantly supplied by
the hydrocarbon chains of the fuel (usually 95% of the product), an essential
element for their biodegradation is oxygen. At low concentrations of oxygen, the
rate of hydrocarbon oxidation is lower, although in aircraft fuel tanks there is a
periodically replenishment of oxygen during tank refuelling. Usually, the
nitrogen and phosphorous availability in kerosene fuel is the limiting factor for
microbial growth. These elements are generally present as nitrates and
phosphates, either in the water phase or in the additives present in the fuel.
The action of fungal contaminants of jet fuels in corrosion can be accomplished
through:
a) Local increase in the proton concentration, derived from organic acidic
metabolites,
b) Greater oxidizing characteristics of the medium favoring pitting attack,
c) Metabolite production decreasing the surface energy of the interface
passive film/electrolyte
d) Microbial adhesion processes enhancing metal dissolution, and
e) Microbial uptake of corrosion inhibitors (mainly nitrates and
phosphates).
A simplified scheme to illustrate the mechanisms influencing biocorrosion
of aluminum alloys in fuel / water systems is shown in Figure 10 [45].
A protective action of some contaminating bacteria (e.g.S. marcescens),
can be developed when an active degradation of hydrocarbon chains is not
accompanied by important reduction in pH levels. In those circumstances, the
organic anions enhance the passive behavior of aluminum [ 1].
206

Fig. 10. Simplified scheme of the initiation of pitting attack on aluminum alloys in fuel/water
systems (from Ref. [45]).

4. AEROBIC CORROSION OF IRON

Metal-oxidizing microorganisms create environments for the accumulation of


chloride ions, forming acidic ferric chloride and manganic chloride, which are
corrosive to steel. However, the main mechanism to explain the action of these
microorganisms in biocorrosion is the formation of differential aeration cells
associated with tubercles. A typical aeration cell of this kind is formed when the
tubercle structure at its outer area acts as a cathode due to the continuous oxygen
supply from the water. Oppositely, an anode (anaerobic area ) is formed at the
bottom of the formation, where access of oxygen is limited.
In all cases corrosion acceleration is mainly due to the presence of SRB in
the inner area of the tubercles where anaerobic conditions are created. Thus, it is
highly feasible that the fastidious iron and manganese depositing bacteria
require other organisms to create suitable conditions for their growth. One
important feature of this case of biocorrosion is that once the tubercle (and the
electrochemical cell) is formed, even the dead of the microorganisms does not
207

extinguish the differential aeration mechanism, since a substantial bamer to the


intake of oxygen has been established.
Thus, biocorrosion of cast iron in low flow or stagnant regions of water
distribution pipelines is mainly due to tubercles formed by iron-oxidizing and
manganese-oxidizing bacteria. These bacteria are able to oxidize soluble ferrous
compounds to less soluble ferric compounds (e.g. ferric hydroxide) which
precipitate at the inner wall of the pipelines under the form of tubercles (discrete
hemispherical mounds).
Tubercles create environments suitable for the growth of other hazardous
microorganisms like the SRB which are generally found at the inner (anaerobic)
regions of the deposits. Near the tubercles other aerobic microorganisms like
slime-forming bacteria (e.g. Pseudomonas) are frequently found. In addition,
tubercles are able to impede the penetration of biocides or corrosion inhibitors
diminishing their effectiveness.
Finally, iron-oxidizing bacteria are able to concentrate chlorides and manganese
ions forming acidic ferric chloride and manganic chloride which are highly
corrosive.

5. M I C R O B I A L INHIBITION OF CORROSION

Corrosion inhibition is the slow down of the corrosion reaction usually


performed by substances which, when added in small amounts to an
environment, decrease the rate of attack by this environment on a metal
(corrosion inhibitors). Microorganisms are able to drastically change the
electrochemical conditions at the metal/solution interface. These changes can
range from the induction or acceleration of corrosion to corrosion inhibition.
There are several examples of microbial effects that could enhance corrosion: i)
stimulation of the anodic reaction by acidic metabolites or the cathodic reaction
by microbial production of a new alternative cathodic reactant (e.g. H2S) ; ii) the
microbial breakdown of protective films or iii) the increase in conductivity of
the liquid environment. However, microbial effects causing corrosion inhibition
has been seldomly mentioned in the literature [46].
Microorganisms can aid to achieve corrosion inhibition according to some
of the following mechanisms: i) neutralizing the action of corrosive substances
present in the environment; ii) forming protective films or stabilizing a pre-
existing protective film on a metal; iii) inducing a decrease in the medium
corrosiveness.
As general key features of microbial inhibition of corrosion the following
points can be summarized [47]:
- Biocorrosion and its counter-process, microbial inhibition of corrosion, are
rarely linked to a single mechanism of to a single species of microorganisms.
208

- Either the corrosive or the inhibitory action of bacteria are developed at


biofilmed metal surfaces where complex biofilm/protective films interactions
occur. Biological activity leads to important changes in the type and
concentrations of ions, pH and oxygen levels inducing significant variations in
the physical and chemical characteristics of the environment as well as in the
electrochemical parameters used to measure the corrosion rate.
- The main mechanisms of corrosion inhibition by bacteria are always linked to
a marked modification of the environmental conditions at the metal/solution
interface by biological activity.
- Microbial corrosion inhibition is frequently accomplished through: i) a
decrease in the cathodic rate by microbial consumption of a cathodic reactant
(e.g. oxygen consumption by respiratory activity); ii) decreasing the medium
aggressiveness in restricted areas of the metal solution interface (e.g. by
neutralizing acidity); iii) providing or stabilizing protective films on the metal
(e.g. biofilm exopolymers with metal binding capacity).
- It must be stressed that in practical situations, the inhibitory action of bacteria
can be reversed to a corrosive action within bacterial consortia structured in
biofilm thickness.
A proper understanding of the identity and role of microbial contaminants
in the specific environment of a metal surface may be used to induce corrosion
inhibition by bacteria as a useful tool to prevent frequent biodeterioration effects
encountered in practice.

6. E L E C T R O C H E M I C A L INTEPRETATION OF BIOCORROSION

The basic concepts of electrochemical corrosion are valid in biocorrosion and


could be used to interpret the acceleration of the corrosion process by
microorganisms in different aqueous media both under anaerobic or aerobic
conditions.
Biocorrosion is rarely interpreted by a single mechanism or rarely caused
by single species of microoganisms [7]. Hence, it is important to be cautious in
the interpretation of data supplied by electrochemical methods. Frequently, these
techniques have been used in complex media where the characteristics and
properties of passive films are not well understood. For instance, the presence of
complex deposits of corrosion products, metabolites and EPS may dramatically
reduce the usefulness of some electrochemical results. Moreover, it has to be
kept in mind that microbial colonization of passive metals can drastically change
their resistance to breakdown by changing locally the type and concentration of
ions, pH values, oxygen gradients and inhibitor levels. These changes should
result in important alterations in the electrochemical behavior of the metal and in
the electrochemical parameters measured in laboratory experiments. In
209

conclusion, whenever electrochemical techniques will be used for the evaluation


ofbiocorrosion, the actual condition of the metal surface must be considered [9].
In classical electrochemical studies, the interface between a metal and the
surrounding electrolyte has been characterized by a certain distribution of
electrical charges giving rise to the so-called electrical double layer [8]. The
knowledge about the structure of the double layer at the metal/solution interface
is mainly based on experimental data obtained with the dropping mercury
electrode. Thus, the behavior of the interface between mercury and several
aqueous electrolyte solutions could be considered approximately equal to that of
an ideally polarizable interface. However, at the light of the present knowledge
on biocorrosion, it can be easily inferred that this behavior is markedly different
from that corresponding to the complex metal/solution interface associated with
biocorrosion [9]. Consequently, electrochemical concepts used for inorganic
corrosion analysis will have to be adapted to the characteristics of the
biologically conditioned interface.
A detailed description of the electrochemical methods for evaluating
biocorrosion is out of the scope of this chapter. A wide variety of electrochemical
techniques such as corrosion and redox potentials measurements, Tafel and
potentiodynamic polarization, linear polarization and electrical resistance
probes, and several modem electrochemical techniques like alternating current
methods or electrochemical noise have been reviewed by several authors in
relation to their use in biocorrosion evaluation [48-50].
A careful use of several electrochemical concepts and methods should be
coupled when possible, with visual inspection, microscopy, innovative surface
analysis techniques and a sound microbiological analysis to reach an adequate
characterization of the causative microorganisms, their interactions and the
effects of their metabolic activities and products on corrosion.

7. PREVENTION AND CONTROL

One of the classic concepts for maintaining an industrial system free of


biodeteriorating effects is "to keep the system clean". Although this is a very
difficult task to accomplish in practice, there are several general methods that
can be used. These methods can be broadly classified in: i) physical and, ii)
chemical. Among the former, flushing is perhaps the most simple, although of
limited efficacy. A special case is the use of flushing supported by cleaners or
jointly with chemical agents that induce biofilm detachment. Abrasive or non-
abrasive sponge balls are frequently employed in the industry. The former could
present problems related with protective passive films that can be damaged, and
the second is not very effective with thick biofilms.
With reference to chemical methods, the most common approach for
controlling biofouling problems in industrial water systems, is the use of
210

biocides. These substances can be either oxidizing or non-oxidizing toxicants.


Chlorine, ozone and bromine are three typical oxidizing agents of industrial use.
Non-oxidizing biocides are claimed to be more effective than oxidizing biocides
for an overall control of algae, fungi, and bacteria. They have a greater
persistence and many of them are pH independent. Often a combination of
oxidizing and non-oxidizing biocides or two non-oxidizing biocides is used to
optimize microbiological control of industrial water systems. Typical biocides of
the second type are formaldehyde, glutaraldehyde, isothiazolones and quaternary
ammonium compounds.
Increasing legislative requirements and the necessity for greater
environmental acceptability, have contributed to restrict the use of some
traditional biocides and to develop either new compounds or carefully selected
blends of existing biocides.

Table 1
Biocides used in industrial water systems. Properties and usual concentrations.

Chlorine: effective against bacteria and algae; oxidizing; pH dependent; concentration range:
0.1-0.2 mg/1 (continuous treatment)
Chlorine dioxide: effective against bacteria, in a lesser extent against fungi and algae;
oxidizing; not dependent on the pH; concentration range: 0.1-1.0 mg/1
Bromine: effective against bacteria and algae; oxidizing; wide pH range; concentration
range: 0.05-0.1 mg/1
Ozone: effective against bacteria and biofilms; oxidizing; pH dependent; concentration
range: 0.2-0.5 rag/1
Methylene-bis-thiocyanate: effective against bacteria; non-oxidizing; hydrolyses at pH
higher than 8.0; concentration range: 1.5-8.0 mg/1
Isothiazolones: effective against bacteria, algae and biofilms; non-oxidizing; not dependent
on the pH; concentration range: 0.9-10 mg/1
QUATS: effective against bacteria and algae; non-oxidizing; surface activity; concentration
range: 8-35 mg/1
Glutaraldehyde: effective against bacteria, algae, fungi and biofilms; non-oxidizing; wide
pH range; concentration range: 10-70 mg/1
THPS (tetra kis-hydroximethil phosphonium): effective against bacteria, algae and fungi;
low environmental toxicity; specific action againts BRS.
211

Taking into account environmental concems, the use of ozone for


different types of industrial water systems, presents several advantages with
respect to other biocides [51 ]. The unique combination of high toxicity of ozone
during treatment, with no toxicant discharge, could make ozone the biocide of
choice for the present decade, if an appropriate balance between positive effects
and costs is reached. Several publications from our laboratory on ozone biocidal
action on sessile and planktonic bacteria, its mechanisms of disinfection and the
optimization of its use can be found in the literature [52].
Among the most promisory non-oxidizing biocides, the THPS (tetra kis-
hydroximethil phosphonium) is a new compound of wide spectrum: effective on
bacteria, fungi and algae. It is being widely used in the oil industry due to its
capacity of dissolving the ferrous sulfide. Its main advantange is its low
environmental toxicity.

8. M O N I T O R I N G BIOCORROSION

Monitoring programs for biofouling and biocorrosion have been mainly focused
in the assessment of planktonic populations in water samples, and generalized
corrosion by using corrosion coupons or some kind of resistance or polarization
resistance probes.
The main objections to these monitoring programs are: i) the planktonic
population does not properly reflect the type and number of organisms living in
the biofilm and causing biodeterioration problems; ii) susceptibility of
planktonic microorganisms to antimicrobial agents markedly differs from that of
sessile microorganisms within the biofilm, mainly because of the protective
action of their EPS. Thus, the monitoring methods adopted must provide
information of well-established biofilms like those developed in system water.
From the corrosion side, the electrical resistance method is appropriate for
indicating a change in the general corrosion rate, but the results are difficult to
interpret in the presence of localized corrosion like pitting, the most frequent
form of attack found in biocorrosion cases [48]. If biofilms or localized
corrosion are present, the polarization resistance will reveal that something is
happening, but may not give an accurate measure of the corrosion rate. Only the
use of any of these techniques jointly with other electrochemical methods or
parameters assessing localized corrosion hazard can provide valuable data for
monitoring the deleterious effects ofbiocorrosion and biofouling.
Owing to the variables of dissimilar nature involved in biofouling and
biocorrosion, an effective monitoring program, either for the laboratory or the
field, must necessarily supply information on water quality, corrosive attack,
sessile and planktonic bacteria populations, biofilms characteristics, and
chemical composition of inorganic and biological deposits [53].
212

Fig. 11. RENAprobeVMsampler device.

Sampling devices for monitoring biocorrosion and biofilms can be


simultaneously used to assess corrosion attack after the removal of biological
and inorganic deposits, giving a wider and more useful information. Sampling
devices fall into two main types: a) directly implanted and, b) side-stream
implanted. Metal coupons, generally made with the same structural material of
the system, present a known surface area, to enable an accurate count of sessile
bacteria per square cm. after biofilm detachment. Coupons are mounted in
holding assemblies which are inserted in the pipework of the laboratory or
industrial system.
Two practical cases of monitoring biocorrosion in industrial waters
developed in our laboratory have been described in the literature [54, 55].

8.1. Monitoring biocorrosion in chemical industry cooling water systems


A monitoring program based on laboratory and field measurements for
assessing biodeterioration on mild steel and stainless steel in recirculating
cooling water systems has been reported [54]. This program was based on: i)
water quality control; ii) corrosion monitoring in the field (weight loss and linear
polarization probe); iii) laboratory corrosion tests (polarization techniques and
corrosion potential vs. time measurements), and iv) use of a multipurpose
sampling device that allows monitoring of sessile populations, biofilms,
corrosive attack, morphology and intensity, and biological and inorganic
deposits analysis. The side stream sampling device (RENAprobe TM) allows
realiable biofilm and corrosion sampling without introducing important
modifications in the system (Fig. 11).
213

8.2. Oilfield water injection system


Among the several factors related to biocorrosion hazard in secondary oil
recovery operations are: i) velocity, temperature, oxygen level and redox
potential of the injection water; ii) chemical composition, pH amount of organic
matter and depth of pumping of the injection water and; iii) effectivenes of the
biocide to gain access to the biofilms. Taking into account these factors a
multiple monitoring program for biofouling and biocorrosion was implemented
in an oild field in Argentina [55]. The goal of the program was to perform a
rapid evaluation of the bacteriological status of the system, and to implement an
effective biocide treatment for sessile bacteria.

9. NEW DEVELOPMENTS IN BIOCORROSION

Recent improvements in analytical, microbiological, electrochemical and


microscopical techniques and instrumentation have allowed the development of
new methods for laboratory and field assessment of biocorrosion in industrial
systems.
Chemical analysis within the biofilm by means of microsensors is one of
the most exciting advances in instrumentation [56]. Biofilm systems have been
considered to be diffusion limited [57]. As a consequence, chemical conditions
at the surface and within biofilms can vary dramatically over a distance of a few
micrometers. Thus, the information obtained from bulk water analysis has a
limited value and must be closely analysed before any conclusions are drawn
about the behavior of the metal/solution interface. Direct measurements inside
biofilms are restricted by: i) the small thickness of the biofilm; ii) the diffusion
limitation of concentration profiles across the biofilm; iii) the heterogeneous
nature of the biofilm. The latter aspect is specially important not only in relation
to microbial coverage of the surface but also with respect to biocorrosion. An
example of microsensor technology applied to evaluate vertical profiles of
chemical species in biofilm systems has been reported [58]. Equally important
tools for the study of biofilm structure are the fiber optic microprobe or optrode
used for finding the location of the biofilm/bulk water interface, and the
mapping of electric fields [59] by means of the scanning vibrating microprobe
(SVM).
Advanced microbiological techniques such as DNA probes have also been
applied to biocorrosion and biofouling research [60]. Although these techniques
are restricted to the laboratory at present, their joint utilization with
microbiological field measurements can be highly useful for monitoring
biocorrosion.
Recent developments in microscopy such as the environmental scanning
electron microscope (ESEM), the confocal laser microscope (CSL) and the
214

atomic force microscope (AFM) permit biofilm observation in real time and
without intoducing distortion of the samples. There is an increasing number of
references using these innovative technologies in recent biocorrosion literature
[61-63].
A combination of CSL and microelectrode techniques allowed correlation
of oxygen concentration profiles with biofilm structure [64]. CSL facilitates the
visualization of biofilm structures by eliminating the interference arising from
out of focus objects [65]. Observations performed under flow conditions and
using physiologically active biofilms, provided information to construct a new
conceptual model ofbiofilm structure.
On the corrosion side, new electrochemical test methods for the study of
localized corrosion phenomena in biocorrosion analysis and monitoring have
been reported [66]. As an example, an electrochemical sensor for monitoring
biofilms on metallic surfaces in real time has been recently presented [67]. The
system provides an immediate indication of the condition of biological activity
on probe surfaces and it is a powerful tool to optimize biocide treatment (Fig.
12).

10. CONCLUDING REMARKS

Biocorrosion is rarely linked to a single mechanism or to a single species of


microorganisms.
Biofilms mediate the interaction between metal surfaces and the liquid
environment. This interaction leads to an important modification of the
metal/solution interface drastically changing the types and concentrations of
ions, pH and oxygen levels.
Biofilm/corrosion product layers interactions at the metal/solution
interface condition the electrochemical behavior of metal surfaces in biological
media.

Acknowledgement
H.A. Videla acknowledges the financial support of the Agencia de Promoci6n
Cientifica y Tecnol6gica of Argentina through the project PICT/99 6782 on
Biodeterioration of materials.
215

Fig. 12. Scheme of an electrochemical sensor for monitoring biofilms (from Ref. [67])

REFERENCES

[1] H.A. Videla, Corrosion/96, paper No. 223, NACE International, Houston, TX (1996).
[2] H.A. Videla and W.G. Characklis, Int. Biodeterior. Biodegr., 29 (1992) 195.
[3] R.G.J. Edyvean and H.A. Videla, Interdisc. Sci. Rev., 16 (1991), 267.
[4] G.G. Geesey, Am. Soc. Microbiol. News, 48, (1982) 9.
[5] W.G. Characklis and K.C. Marshall, Biofilms: A Basis for an interdisciplinary
approach. In: Biofilms, W.G. Characklis and K.C. Marshall (eds.) John Wiley and Sons
Ltd, New York, 1990, pp. 3-15.
[6] W.G. Characklis, Biotech. Bioeng., 23, (1981) 1923.
[7] H.A. Videla, Metal Dissolution/redox in Biofilms. In: Structure and Function of
Biofilms, W.G. Characklis and P.A. Wilderer (eds.) John Wiley & Sons, Chichester,
U.K, 1989, pp. 301-320.
[8] H.A. Videla, Microbially induced corrosion: An updated overview. In: Biodeterioration
and Biodegradation 8, H.W. Rossmoore (ed.) Elsevier Applied Science, London, UK,
1991, pp. 63-88.
[9] H.A. Videla, Electrochemical aspects ofbiocorrosion. In: Bioextraction and
biodeterioration of metals, C.C. Gaylarde and H.A. Videla (eds.) Cambridge University
Press, Cambridge, UK, 1995, pp. 85-127.
216

[ 10] G.A.H. von Wolzogen Ktihr and L.R. Van der Vlugt, Water (den Haag), 18, (1934) 147
(Translation in Corrosion 17, (1961) 293).
[ 11 ] J.D.A. Miller, Metals, in: Microbial biodeterioration, A.D. Rose (ed.), Academic Press,
New York, 1981, pp. 149-202.
[12] R.C. Salvarezza and H.A. Videla, Corrosion, 36 (1980) 550.
[13] G. Gragnolino and O.H. Tuovinen, International Biodeterioration, 20 (1984) 9.
[14] R.C. Salvarezza, H.A. Videla and A.J. Arvia, Corros. Sci., 22 (1982) 815.
[15] R.C. Salvarezza, H.A. Videla and A.J. Arvia, Corros. Sci., 23 (1983) 717.
[16] H.A. Videla, Corrosion of mild steel induced by sulfate-reducing bacteria-a study of
passivity breakdown by biogenic sulfides. In: Biologically Induced Corrosion, S.C.
Dexter (ed.), NACE-8, Houston, TX, 1986, pp. 162-171.
[ 17] H.A. Videla, Electrochemical interpretation of the role of microorganisms in corrosion.
In: Biodeterioration 7, D.R. Houghton, R.N. Smith and H.O.W. Eggins (eds.) Elsevier
Applied Science, London, UK, 1988, pp. 359-371.
[ 18] C.A. Acosta, R.C. Salvarezza, H.A. Videla and A.J. Arvia, Electrochemical behaviour
of mild steel in sulfide and chloride containing solutions, Passivity of Metals and
Semiconductors, M. Froment (ed.) 1984, pp. 387-392.
[ 19] J.W. Costerton and G.G. Geesey, The microbial ecology of source colonization and of
consequent corrosion. In: Biologically Induced Corrosion, S.C. Dexter (ed.) NACE-8,
Houston, TX, NACE, 1986, pp. 223-232.
[20] H.A. Videla, C.L. Swords, M.F.L. de Mele, R.G. Edyvean, P,Watkins and I.B. Beech,
Corrosion/98, paper No. 289, NACE International, Houston, TX (1998).
[21] H.A. Videla, R.G. Edyvean, C.L. Swords, M.F.L. de Mele and I.B. Beech,
Corrosion/99, paper No. 163, NACE International, Houston, TX (1999).
[22] H.A. Videla, Biofouling, 15 (2000) 37.
[23] H.A. Videla, C.L. Swords and R.G. Edyvean, Corrosion/2002, paper No. 02557, NACE
International, Houston, TX (2002).
[24] I.B. Beech, V. Zinkevich, R. Tapper and R. Avci, Journal of Microbiological Methods,
36(1999) 3.
[25] W. Lee, Z. Lewandowski, P.H. Nielsen and W.A. Hamilton, Biofouling, 8 (1995) 165.
[26] Tumbull (ed) Hydrogen transport and cracking in metals, London Institute of Materials,
UK, 1995.
[27] S.G. G6mez de Saravia, M.F.L. de Mele, H.A. Videla and R.G.J. Edyvean, Biofouling,
11 (1997) 1.
[28] C.J. Thomas, R.G.J. Edyvean and R. Brook, Biofouling, 1 (1988) 65.
[29] R.G.J. Edyvean, J. Benson, C.J. Thomas, I.B. Beech, and H.A. Videla, Corrosion/97,
paper No. 206, NACE International, Houston, TX (1997).
[30] R.G.J. Edyvean, J. Benson, I.B. Beech and H.A. Videla, Microbiological influences on
hydrogen effects on steels. In: Hydrogen Energy Progress XII, vol. 3, J.C. Bolcich and
T.N. Veziroglu (eds.) 1998, pp. 1755-1762.
[31] R.G.J. Edyvean, International Biodeterioration, 23 (1987) 199.
[32] E.C. Hill, Biodegradation of petroleum products. In: Petroleum Microbiology, R.M.
Atlas (ed.), Chap. 15, MacMillan Publishing Co., New York, 1984, p. 579-617.
[33] P. Me Kenzie, A.S. Akbar and J.D. Miller, Fungal Corrosion of Aircraft Fuel Tank
Alloys, Technical paper, The Institute of Petroleum, London, UK (1977) 37.
[34] H.G. Hedrick, Materials Protection, 9 (1970) 27.
[35] D.G. Parbery, Material und Organismen, 6 (1971) 161.
[36] M.F.L. de Mele, R.C. Salvarezza and H.A. Videla, International Biodeterioration
Bulletin, 15 (1979) 39.
217

[37] R.C. Salvarezza, M.F.L. de Mele and H.A. Videla, Intemational Biodeterioration
Bulletin, 15 (1979) 125.
[38] R.C. Salvarezza, M.F.L. de Mele and H.A. Videla, Br. Corros. J., 16 (1981) 162.
[39] R.C. Salvarezza and H.A. Videla, Electrochemical behavior of aluminum in
Cladosporium resinae cultures, in: Biodeterioration 6, S. Barry, D.R. Houghton, G.C.
Llewellyn and C.E. O' Rear (eds.) CAB International Mycological Institute, London,
UK, 1986, p. 212-218.
[40] R.J. Watkinson, Hydrocarbon degradation, in: Microbial Problems and Corrosion in Oil
and Oil Products Storage, E.C. Hill (ed.) The Institute of Petroleum, London, UK, 1984,
pp. 50-56.
[41] R.A. Neihof and M. May, Intemational Biodeterioration Bulletin, 19 (1983) 59.
[42] H.A. Videla, P.S. Guiamet, S.M. do Valle and E.H. Reinoso, Corrosion/88, paper No.
91, NACE, Houston, TX (1988).
[43] D. Allsopp and K.J. Seal, Introduction to Biodeterioration, Edward Arnold, London,
1986.
[44] S. Holmes, Microbiology of hydrocarbon fuels. In: Proceedings 2nd International
Conference on Long-Term Storage Stabilities of Liquid Fuels, Southwest Research
Institute, San Antonio, TX, 1986, 336.
[45] H.A. Videla, Practical Cases. In: Manual of Biocorrosion, Chapter 7, Boca Raton,
Florida: CRC Lewis Publishers, 1996, pp. 179-220.
[46] H.A. Videla, Corrosion inhibition in the presence of microbial corrosion. In: Reviews
on Corrosion Inhibitor Science and Technology, Vol 2, A. Raman and P. Labine (eds.),
Chapter IX, NACE Intemational, Houston, TX, 1996, pp. 1-11.
[47] H.A. Videla, Corrosion Inhibition by Bacteria. In: Manual of Biocorrosion, Chapter 5,
Boca Raton, Florida: CRC Lewis Publishers, 1996, pp. 121-135.
[48] S.C. Dexter, D.J. Duquette, O.W. Siebert and H.A. Videla, Corrosion/89, paper No.
616, NACE International, Houston, TX (1989).
[49] D.J. Duquette, Electrochemical techniques for evaluation of Microbiologically
Influenced Corrosion processes. Advantages and disadvantages In: Argentine-USA
Workshop on Biodeterioration (CONICET-NSF), H.A. Videla (ed.) Aquatec Quimica
S.A., Sao Paulo, Brazil, 1996, pp. 15-32.
[50] F. Mansfeld and B.J. Little, Corrosion/90, paper No. 108, NACE International,
Houston, TX (1990).
[51] J.M. Brook and P.R. Puckorius, Corrosion/91, paper No. 212, NACE Intemational.
Houston. TX (1991).
[52] H.A. Videla, M.R. Viera, P.S. Guiamet and M.F.L. de Mele, Corrosion/99, paper No.
186, NACE Intemational, Houston, TX (1999).
[53] H.A. Videla, Detection, Identification and Monitoring. In: Manual of Biocorrosion,
Chapter 6, Boca Raton, Florida: CRC Lewis Publishers, 1996, pp. 137-178.
[54] H.A. Videla, F. Bianchi, M.M.S. Freitas, C.G. Canales and J.F. Wilkes, Monitoring
biocorrosion and biofilms in industrial waters: a practical approach. In: Microbiological
Influenced Corrosion Testing, J.R. Kems and B.J. Little (eds.) ASTM Publications STP
1232, American Society for Testing and Materials, Philadelphia, PA, 1994, pp. 128-
137.
[55] H.A. Videla, P.S. Guiamet, O.R. Pardini, E. Echarte, D. Trujillo and M.M.S. Freitas,
Corrosion/91, paper No. 103, NACE International, Houston, TX (1991).
[56] Z. Lewandowski, Dissolved oxygen gradients near microbially colonized surfaces. In:
Biofouling and biocorrosion in industrial water systems, G.G. Geesey, Z. Lewandowski
and H.C. Flemming (eds.) Lewis Publishers, Boca Raton, FL, 1994, pp. 175-188.
218

[57] W.G. Characklis, Biofilm processes, in: Biofilms, W.G. Characklis and K.C. Marshall
(eds.) John Wiley & Sons, New York, 1990, pp. 195-231.
[58] Z. Lewandowski, T. Funk, F. Roe and B. Little, Spatial distribution ofpH at mild steel
surfaces using an iridium oxide microelectrode. In: Microbiologically Influenced
Corrosion Testing, J.R. Kearns and B.J. Little (eds.) ASTM Publications STP 1232,
Philadelphia, PA, 1994, pp. 61-69.

[59] Z. Lewandowski, F. Roe, T Funk and D. Chen, Proc. NSF-CONICET Workshop,


Biocorrosion and Biofouling: Metal/Microbe Interactions, Buckman Laboratories
International, Inc., Memphis, TN, 1993, pp. 52-61.
[60] S. Le Borgne, J. Jan, J.M. Romero and M. Amaya, Corrosion/2002, paper No. 02461,
NACE International, Houston, TX, (2002).
[61] A. Steele, D.T. Goddard and I.B. Beech, Int. Biodet. Biodegrad. 34 (1994) 35.
[62] A. Steele, I.B. Beech and D.T. Goddard, Proc. 1995 International Conference on
Microbialy Influenced Corrosion, American Welding Society-NACE International,
Houston, TX, 1995, 73/1-13.
[63] J.W. Costerton, Structure ofbiofilms. In: Biofouling and Biocorrosion in Industrial
Water Systems, G.G. Geesey, Z. Lewandowski and H.C. Flemming (eds.) Lewis
Publishers, Boca Raton, FL, 1994, pp. 1-14.
[64] Z. Lewandowski, P. Stoodley and F. Roe, Corrosion/95, paper No. 222, NACE
International, Houston, TX (1995).
[65] D.E. Caldwell, D.R. Korber and J.R. Lawrence, Adv. Microbial Ecol., 12 (1992) 1.
[66] H.A. Videla, Fundamentals of electrochemistry. In: Manual of Biocorrosion, Chapter 4,
Boca Raton, Florida: CRC Lewis Publishers, 1996, pp. 73-120.
[67] G.J. Licina, Corrosion/2001, paper No. 01442, NACE International, Houston, TX
(1995).

You might also like