ArticlePDF AvailableLiterature Review

Microbially mediated metal corrosion

Authors:

Abstract

A wide diversity of microorganisms, typically growing as biofilms, has been implicated in corrosion, a multi-trillion dollar a year problem. Aerobic microorganisms establish conditions that promote metal corrosion, but most corrosion has been attributed to anaerobes. Microbially produced organic acids, sulfide and extracellular hydrogenases can accelerate metallic iron (Fe0) oxidation coupled to hydrogen (H2) production, as can respiratory anaerobes consuming H2 as an electron donor. Some bacteria and archaea directly accept electrons from Fe0 to support anaerobic respiration, often with c-type cytochromes as the apparent outer-surface electrical contact with the metal. Functional genetic studies are beginning to define corrosion mechanisms more rigorously. Omics studies are revealing which microorganisms are associated with corrosion, but new strategies for recovering corrosive microorganisms in culture are required to evaluate corrosive capabilities and mechanisms. Interdisciplinary studies of the interactions among microorganisms and between microorganisms and metals in corrosive biofilms show promise for developing new technologies to detect and prevent corrosion. In this Review, we explore the role of microorganisms in metal corrosion and discuss potential ways to mitigate it.
Nature Reviews Microbiology
nature reviews microbiology https://doi.org/10.1038/s41579-023-00920-3
Review article Check for updates
Microbially mediated
metal corrosion
Dake Xu 1,2, Tingyue Gu 3,4,5,6 & Derek R. Lovley 1,7
Abstract
A wide diversity of microorganisms, typically growing as biolms,
has been implicated in corrosion, a multi-trillion dollar a year problem.
Aerobic microorganisms establish conditions that promote metal
corrosion, but most corrosion has been attributed to anaerobes.
Microbially produced organic acids, sulde and extracellular
hydrogenases can accelerate metallic iron (Fe0) oxidation coupled
to hydrogen (H2) production, as can respiratory anaerobes consuming
H2 as an electron donor. Some bacteria and archaea directly accept
electrons from Fe0 to support anaerobic respiration, often with c-type
cytochromes as the apparent outer-surface electrical contact with the
metal. Functional genetic studies are beginning to dene corrosion
mechanisms more rigorously. Omics studies are revealing which
microorganisms are associated with corrosion, but new strategies for
recovering corrosive microorganisms in culture are required to evaluate
corrosive capabilities and mechanisms. Interdisciplinary studies of the
interactions among microorganisms and between microorganisms
and metals in corrosive biolms show promise for developing new
technologies to detect and prevent corrosion. In this Review, we explore
the role of microorganisms in metal corrosion and discuss potential
ways to mitigate it.
Sections
Introduction
Key reactions in microbial
metal corrosion
Microbial corrosion of metals
under aerobic conditions
Microbial corrosion of metals
under anaerobic conditions
Microbial diversity in metal
corrosion
Mitigating microbial corrosion
Outlook
1Electrobiomaterials Institute, Key Laboratory for Anisotropy and Texture of Materials (Ministry of Education),
Northeastern University, Shenyang, China. 2Shenyang National Laboratory for Materials Science, Northeastern
University, Shenyang, China. 3Department of Chemical & Biomolecular Engineering, Ohio University, Athens,
OH, USA. 4Department of Biological Sciences, Ohio University, Athens, OH, USA. 5Institute for Corrosion and
Multiphase Technology, Ohio University, Athens, OH, USA. 6Institute for Sustainable Energy and the Environment,
Ohio University, Athens, OH, USA. 7Department of Microbiology, University of Massachusetts, Amherst, MA, USA.
e-mail: gu@ohio.edu
Nature Reviews Microbiology
Review article
in the complex biofilms associated with corrosion outside the labora
-
tory. The understanding of microbial corrosion of non-ferrous metals
is even more poorly understood. The goal of this Review is to critically
review the previous research on microbial metal corrosion and its
prevention and to highlight exciting opportunities for future research.
Key reactions in microbial metal corrosion
In the air, Fe0 is abiotically oxidized with the reduction of oxygen (O2)
(Fig.2a, reaction 1), and the Fe
2+
generated is further abiotically oxi-
dized to Fe(III) oxides, generating the familiar ‘rust’ associated with
corroding iron. Aqueous environments foster the growth of chemically
and physically heterogeneous biofilms on metal surfaces populated
with diverse microorganisms. O
2
is often unavailable as an Fe
0
oxidant
within these biofilms. Potential Fe0 oxidants under anaerobic condi-
tions include protons (H
+
), nitrate, Fe(III), sulfate, carbon dioxide (CO
2
)
and hydrogen sulfide (H
2
S). Only H
+
(Fig.2a, reaction 2) and H
2
S (Fig.2a,
reaction 3) react abiotically with Fe
0
, generating H
2
(refs. 9,1214). Fe
0
oxidation with other potential electron acceptors requires micro-
bial catalysis (Fig.2a, right), in which a diversity of microorganisms
(Table 1 and Fig.2b) accept electrons from Fe
0
to support anaerobic
respiration, either directly or indirectly via mechanisms detailed later.
Fe
2+
generated from these various reactions can react with sulfide,
carbonate, phosphate or additional iron ions to produce a wide range
of minerals (Fig.2a, reactions 4–9), which typically accumulate on
corroding iron, becoming incorporated in biofilms, and are visual
signs of intense corrosion. Key to understanding the factors control-
ling the rate and extent of iron corrosion in specific environments
is discerning which of these many possible reactions is taking place
within corrosive biofilms.
Similar to Fe0, metallic nickel (Ni0), metallicaluminium (Al0),
metalliczinc (Zn
0
) and metallictitanium (Ti
0
) readily react with O
2
,
generating metal oxides of the metals that can form a protective film,
reducing additional O
2
access to the underlying metal and trapping
metal ions resulting from the metal oxidation at the surface, making
further metal oxidation thermodynamically unfavourable (Fig.2c).
However, if the passive film becomes damaged, exposing the under-
lying metals to water under anaerobic conditions, the metals will
reduce H
+
to H
2
(refs. 15,16). By contrast, the oxidation of metallic
copper (Cu0) is too electropositive to directly generate H2 (ref. 17),
but in sulfidogenic environments H
2
S can react with Cu
0
to release H
2
(refs. 17,18) (Fig.2d).
Microbial corrosion of metals under aerobic
conditions
Biofilms of Fe2+-oxidizing bacteria along with the Fe(III) oxides that
are generated from the Fe
2+
oxidation (Fig.2a) are commonly associ-
ated with corrosion of ferrous metals. In aqueous environments, Fe2+-
oxidizing bacteria accelerate the oxidation of Fe
2+
to ferric ion (Fe
3+
)
coupled to O
2
reduction over the abiotic reaction rate
19
. These include
lithoautotrophic bacteria such as Mariprofundus, Gallionella and Sider-
oxydans spp., which conserve energy from Fe2+ oxidation that is then
used to fix CO2, and heterotrophs such as Leptothrix and Sphaerotilus
spp., which grow heterotrophically and catalyse Fe2+ oxidation without
energetic benefit19. Rapid removal of Fe2+ should promote abiotic Fe0
oxidation coupled to O2 reduction (Fig.2a, reaction 1) by reducing
Fe
2+
accumulation. In low pH environments, such as those inhabited
by Acidithiobacillus ferrooxidans, soluble Fe3+, which has a high redox
potential, is also an Fe0 oxidant (Fig.2a, reaction 9), generating more
Fe
2+
to further support A. ferrooxidans respiration
20
. Thus, corrosion of
Introduction
Of the many important impacts of microbial biofilms, corrosion is the
most economically damaging, accounting for more than two-thirds of
all expenses attributed to biofilm growth and activity
1
. The more than
US$2.7trillion annual cost of microbial corrosion dwarfs the biofilm
costs associated with human health, food and agriculture, and energy
and the environment combined
1
. Well-known examples of the cata-
strophic impact of microbial corrosion include the notorious 2006
Alaskan pipeline leak, which in addition to environmental damage
disrupted global oil markets2. In the United States alone, water leaks
associated with corrosion of water distribution pipes, much of it attrib-
uted to microbial activity, will require a trillion-dollar investment in
new infrastructure over the next 25years
3
. Microbial corrosion impacts
not only pipelines but also a diversity of other structures and devices
including nuclear waste storage facilities, heat exchangers, reinforced
concrete, oil and gas infrastructure, water utilities, fuel systems, power
plants, underground storage tanks, marine platforms, offshore wind
turbines and dental devices (Fig.1).
Humans have been producing metallic iron (Fe
0
) for only a few
thousand years. Thus, corrosion processes associated with micro-
organisms probably evolved for reasons other than utilizing Fe0 as an
energy source. For example, the microbial hydrogen (H2) oxidation that
is key to some forms of microbial corrosion is a likely reflection of the
early evolution of H2 metabolism and the continued importance of H2
as a central intermediate in many modern anaerobic environments
4
.
Microorganisms that can corrode by directly extracting electrons
from metals are likely to use mechanisms previously evolved for micro-
bial electron extracellular exchange with minerals and other microbial
species5,6.
Few microbiologists have studied microbial corrosion, despite
its economic and environmental impacts. The study of microbial cor-
rosion shares many core concepts and analytical methods with other
microbiology subdisciplines (Box1), and corrosion investigations
seem to be entering a renaissance era in which new, interdisciplinary
approaches are beginning to provide fresh insights into the ways in
which microbial activity can cause metal corrosion. Better understand-
ing of corrosion mechanisms is expected to lead to improved strategies
for detecting corrosion before it causes severe damage, and to new
approaches for corrosion mitigation.
Microorganisms can deteriorate the quality of many metals includ-
ing aluminium, copper, nickel and titanium, but most microbiological
research has focused on the role of microorganisms in the corrosion
of iron-containing ferrous metals, such as carbon and stainless steels
7
.
Corroding iron metals are typically covered with biofilms (Fig.1) of
diverse structure and function that initially promote deterioration
of the metal by oxidizing Fe0 to ferrous ion (Fe2+):
Fe →Fe+2e (1)
02+−
For this reaction to proceed, the electrons (e) released from Fe0
must be transferred to one of a diversity of potential electron-accepting
molecules (Fig.2a).The question that has been asked for more than a
century8 is which of the myriad possible abiotic and microbiological
reactions involving Fe
0
(Fig.2a) are the most important. The popularity
of different mechanisms has waxed and waned over time and there is
still little consensus on how microbial corrosion takes place. Detailed
histories of microbial corrosion research are available elsewhere
911
.
Only a few proposed mechanisms have been rigorously evaluated under
well-defined laboratory conditions and none have been demonstrated
Nature Reviews Microbiology
Review article
iron-containing steel was threefold to sixfold faster in the presence of
A. ferrooxidans than in sterile controls, even when A. ferrooxidans was
not in direct contact with the iron20.
At circumneutral pH, Fe2+-oxidizing isolates have shown mixed
results for whether pure cultures accelerate iron corrosion19,2124. Fe3+
combines with hydroxyl ions to form Fe(III) oxides (Fig.2a, reaction 4),
a
Microbial corrosion
Various industries
Power plants Fuel systems Water utilities Oil and gas Rebar and
concrete
Heat
exchangers
Nuclear waste
storage
Oshore
assets
Ships and
port facilities
Dental
devices
Manned
spacecraft
Marine environments Medical Space
Other
Alcaligenes
Zoogloea
Diaphorobacter
Azospirillum
Uncultured
Anaerospora
Paludibacter
Desulfovibrio
Desulfomicrobium
Petrimonas
Opitutus
Erysipelothrix
Brevundimonas
Xanthobacter
Sporomusa
Magnetospirillum
Pseudomonas
Hydrogenophaga
Relative abundance of bacteria (%)
10
20
30
40
50
60
70
80
90
100
0
Halomonas
Methanosphaerula
Methanoregula
Nitrosopumilus
Unclassiied
Mathanobacterium
Methanomassiliicoccus
Methanolinea
Aquisphaera
Desulfitibacter
Methanospirillum
Methanosphaera
Methanothermobacter
Methanomethylovorans
Nitrososphaera
Methanoculleus
Methanothrix
Methanosarcina
Relative abundance of archaea (%)
10
20
30
40
50
60
70
80
90
100
0
100.00 µm
50.00 µm
194.30 µm
150.00 µm
0.00 µm
Width: 636.4 µm Height: 636.4 µm Depth: 194.3 µm
30 µm 0
50
100
y (µm)
z (µm)
x (µm)
Largest pit
depth = 26 µm
150
200
250
300
300
34
0
050 100 150 200 250
25 µm
20 µm
15 µm
10 µm
5 µm
0 µm
b
e f
c d
Fig. 1 | Examples of microbial corrosion. a, Diversity of materials that are
susceptible to microbial corrosion. Areas affected by corrosion include various
industries, such as oil and gas, water utilities and power plants, as well as marine
environments, and space and medical sectors. b, Severe microbial corrosion of a
carbon steel pipeline used in shale gas production. c, Biofilm associated with the
corroding carbon steel surface imaged with confocal scanning laser microscopy.
d, Confocal scanning laser microscopy image of carbon steel, after removal of
biofilm and corrosion products, revealing the morphology of pits generated
by microbial corrosion. e,f, Heatmaps of the diversity of bacteria (part e) and
archaea (part f) on the surface of a corroding carbon steel pipeline. Parts b,c
and d unpublished data from Dake Xu laboratory. Parts e and f replotted with
data from ref. 87.
Nature Reviews Microbiology
Review article
which precipitate on cells and the metal surface
19,23
. Depending on con-
ditions, Fe(III) oxides may or may not abiotically oxidize Fe0 (ref. 25).
Mn(IV) oxides produced by Mn2+-oxidizing bacteria are strong Fe0
oxidants. Fe0 oxidation with Mn(IV) regenerates Mn2+ (refs. 2527).
The Mn
2+
-mediated electron shuttling between Fe
0
and Mn
2+
-oxidizing
bacteria corrodes Fe
0
faster than direct abiotic Fe
0
oxidation with O
2
(Fig.3a). In a similar manner, elemental iodine (I
2
) oxidizes Fe
0
, pro-
ducing the iodine ion (I). Then, I-oxidizing bacteria regenerate I2 to
further promote Fe0 oxidation28. In this way, corrosion is microbially
accelerated in pipelines carrying water with high concentrations of I2
(ref. 28).
Fe2+-oxidizing and Mn2+-oxidizing bacteria are early biofilm
colonizers, consuming O
2
, generating oxide coatings on the metal
surface and creating a low O2 environment near the biofilm–
metal interface
19,22,27
. Heterotrophic microorganisms contribute to
O2 removal within corrosion biofilms because organics are avail-
able to microorganisms in many environments in which corrosion
eventually develops29. As O2 is depleted deeper within biofilms,
microorganisms capable of fermentative metabolism and anaerobic
respiration become established near the metal–biofilm interface19,30,
where anaerobic microbial activity promotes corrosion via multiple
mechanisms (Fig.3a).
Box 1
The interconnection of corrosion biolms research with other
disciplines
The study of microbial corrosion can advance other microbiology
subdisciplines, which share common concepts and experimental
approaches, and vice versa. Electrotrophy, the direct uptake
of electrons from extracellular electron donors, is a rapidly
evolving ield not only in corrosion but also in microbial ecology,
bioengineering and biogeochemistry6. For example, the question
of the relative role of direct metal-to-microorganism electron transfer
or hydrogen (H2)-mediated electron transfer in metal corrosion
parallels inquiries into the relative importance of direct interspecies
electron transfer and interspecies H2 transfer feeding methanogens
in anaerobic digesters, soils and sediments130. Identiication of
microorganisms capable of serving as the electron-accepting partner
for direct interspecies electron transfer is revealing microorganisms
that are also eective in corrosion via direct metal-to-microorganism
electron transfer35. Challenging questions key for optimizing
microbial electrosynthesis — the bioelectrochemical production
of organic commodities from carbon dioxide (CO2)131134 — are
similar to those for corrosion: do microorganisms primarily accept
electrons from electrodes via H2, electron shuttles or direct
electrode-to-microorganism electron transfer; and which electron
transfer method is most practical for scaling. Corroding metal
surfaces, which select for microorganisms that are highly eective
in extracting electrons from electroactive surfaces, might be an
excellent environment for recruiting microorganisms well adapted
for electron uptake from cathodes in electrosynthesis. Elucidation
of mechanisms for electron uptake from cathodes yields insights into
how microorganisms might accept electrons from Fe0 (refs. 62,135).
The study of bioelectrochemical technologies and microbial
corrosion already relies on similar electrochemical techniques to
document the rate, extent and type of electron exchange (Box2).
Adhesion to metal and bioilm development is a key feature of
corrosion136. Thus, advancements in bioilm theory90,137,138 have clear
application to corrosion. In turn, future investigation of corrosion
bioilm development, composition and diversity has the potential
to contribute general knowledge to bioilm ecology and evolution
in complex environments.
Fig. 2 | Key reactions and diversity of microorganisms involved in metal
corrosion. a, Key reactions associated with microbial corrosion of ferrous
metals. Each of these reactions has been shown to be thermodynamically
favourable under conditions related to corrosion. Numbers 1–9 refer to abiotic
reactions associated with corrosion. b, Diversity and spatial heterogeneity
of microorganisms within a biofilm involved in corrosion of ferrous metals.
Hydrogen (H2)released from metallic iron (Fe0)and produced by fermentative
bacteria (representative genera in grey box, left) is consumed by denitrifiers
(green box), sulfate reducers (yellow box), methanogens (light blue box) and
acetogens (red box). Heterotrophs (magenta box), sulfide-oxidizing consortia
(yellow box) and iron oxidizers (grey box, right) consume O2 promoting
anaerobic corrosion reactions. Electroactive methanogens (magenta box) and
Fe(III) reducers (orange boxes) directly extract electrons from Fe0 generating
ferrous ion (Fe2+)that iron oxidizers convert to Fe(III) oxides that serve as an
additional electron acceptor for Fe(III) reducers. The diffusion of organics
and electron acceptors (oxygen (O2), nitrate (NO3), sulfate (SO42–)) from the
surrounding environment into the biofilm and their preferential consumption
with the biofilm (O2, then NO3, then SO42–), as well as the availability of Fe0 as
an electron donor at the base of the biofilm, results in vertical heterogeneity
of diverse functional populations. Lightning bolts designate direct electron
transfer from Fe0 to cells. c, Key reactions associated with the corrosion of
aluminium, zinc and titanium. Films of metal oxides, hydroxides and/or sulfides
block O2 and proton (H+)contact with the underlying metal and the outward
diffusion of metal ions. Microbial or chemical damage to the films permits O2
and H+ to access the metal and allows the release of metal ions generated from
the metal oxidation. d, Key reactions associated with copper corrosion. Under
aerobic conditions, metallic copper (Cu0)is abiotically oxidized and the cupric
ion (Cu2+)generated combines with microbially produced oxalic acid, forming
insoluble copper oxalate precipitates. Under anaerobic conditions, sulfate
reducers produce hydrogen sulfide (H2S) that reacts with Cu0 to generate H2
and copper sulfide (Cu2S) precipitates. Al0, metallic aluminium; CH4, methane;
CH3COOH, acetic acid; CO2, carbon dioxide; CO32–, carbonate; e, electron;
Fe3+, ferric ion; FeCO3, ferrous carbonate; Fe3O4, magnetite; Fe(OH)3, iron
hydroxide; FePO4, ferric phosphate; FeS, iron sulfide; H2O, water; N2, nitrogen;
NH4+, ammonium; OH, hydroxide; PO43–, phosphate; S2–, sulfide; SO42–, sulfate;
Ti0, metallic titanium; Zn0, metallic zinc.
Nature Reviews Microbiology
Review article
O2H2O
O2
Iron oxidizers
H2O
OH
Fe(OH)3FeS
5–8
Fe2+
Fe0Fe2+
Fe
2+
Fe
0
Fe2+
S2–
CO3
2–
PO4
3–
Fe3+ H+
Fe
2+
Fe
0
Fe
2+
Fe
0
H2
Respiratory
anaerobes
H2
Fe3+
CH4, acetate
S2–
Fe2+
N2, NH4
+
Succinate
CO2
SO4
2–
Fe(III)
NO3
Fumarate
Ferrous minerals:
sulide, carbonate,
phosphate, magnetite
Electrons transferred to microorganisms
via numerous mechanisms (Fig. 3)
Fe0
H2S + Fe0
e
Al3+, Zn2+, Tin
Outward diusion
blocked inhibiting
further metal oxidation
Microbially or chemically
damaged passive ilm
Undamaged passive
ilm of metal oxides,
hydroxides, sulides
O2, H+ penetration
blocked
Al0
Zn0
Ti0
Al3+
Zn2+
Tin+
e
1 2 Fe0 + 4 H+ + O2
2 Fe0 + 2 H+
3 Fe0 + H2S
4 Fe3+ + 3 OH
5 Fe2+ + S2–
2 Fe2+ + 2 H2O
Fe2+ + H2
FeS + H2
Fe(OH)3
FeS
6 Fe2+ + CO3
2–
7 3 Fe2+ + 2 PO4
3–
8 Fe2+ + 2 Fe3+ + 8 OH
9 Fe0 + 2 Fe3+
FeCO3
Fe3(PO4)2
Fe3O4 + 4 H2O
3 Fe2+
Abiotic reactions
CO2 + 8 H+ + 8 e
2 CO2 + 8 H+ + 8 e
SO4
2– + 8 H+ + 8 e
Fe(III) + e
2 NO3
+ 12 H+ + 10 e
NO3
+ 10 H+ + 8 e
Fumarate + 2 H+ + 2 e
CH4 + 2 H2O
CH3COOH + 2 H2O
S= + 4 H2O
Fe2+
N2 + 6 H2O
NH4
+ + 3 H2O
Succinate
Anaerobic respiration
4 Fe2+ + 4 H+ + O24 Fe3+ + 2 H2O
Aerobic respiration
H2H+
H2O
H2O
H2O
Fe(III)
Fe(III)
Fe(III)
Fe0
Fe2+
H2H+
Fe0Fe2+
Fe0Fe2+
Fe
2+
Fe
0
Fe0
Al0, Zn0 or Ti0
Acetobacterium
Sporomusa
Methanobacterium
Methanococcus
Methanosarcina
Methanosaeta
NO3
Fe2+
NH4
+
Fe2+
Fe2+
Acetate CO
2
H2
H2
H+
CO
2
O
2
O
2
CO
2
CO
2
O
2
H2H+
CO
2
CH
4
CH
4
SO
4
2
SO
4
2
SO
4
2
S
2
S
2
H2H+
H
2
H
+
NO3
NO3
O2
N2
H2
Organics
Organic acids
Clostridia
Bacillus
Enterobacter
Anaerolinea
Mariprofundus
Dechloromonas
Sideroxydans
Pseudomonas
Marinobacter
Thalassospira
Desulfovibrio
Desulfoferrobacter
Desulfotignum
Desulfobacula
Desulfobulbus
Sulfur-oxidizing
consortia
Precipitation
Vibrio
Pseudoalteromonas
Flexibacter
Marinobacter
Fe2+
Geobacter
Shewanella
Geobacter
Geothrix
Rhodoferax
Organic acids
Organics
Biofilm
a
b
c
O2, H+H2OH2S H2
Microbially produced oxalic acid Microbially
produced sulide
Microbial
consumption
Copper oxalate
precipitate
Cu0
Aerobic Anaerobic
d
Cu0Cu2+
Cu2+
C C
O
O
O
OCu2S precipitate
Cu0
12
4
93
Organics
Microbial consumption
O2H2OH2H+
Nature Reviews Microbiology
Review article
The role of aerobic bacteria in the corrosion of non-ferrous metals
does not seem to have been intensively investigated, but microorgan-
isms can accelerate corrosion under aerobic conditions, presumably
by releasing metabolites that destroy the protective metal oxide films
occluding the underlying metal31,32. The fungus Aspergillus niger pro-
duced organic acids, lowering the pH to promote the oxidation of
Cu0 (Fig.2d). Oxalic acid, the most abundant acid produced, further
accelerated corrosion by chelating the cupric ion (Cu2+) generated from
Cu0 oxidation and forming copper oxalate precipitates33.
Microbial corrosion of metals under anaerobic
conditions
Unlike O2, potential oxidants for Fe0 in anoxic environments, such
as nitrate, sulfate and CO2, do not spontaneously abiotically react
with Fe0 to oxidize and corrode it. Microbial catalysis is required.
Hence, microorganisms have their greatest impact on corrosion of
ferrous metals in anaerobic environments. The activity of anaerobic
microorganisms promotes the oxidation of Fe
0
through multiple mech-
anisms (Fig.3b,c) including the production of metabolites that enhance
the oxidation of Fe
0
to Fe
2+
with the reduction of H
+
to H
2
; H
2
-mediated
electron transfer between the metal and the microorganism; direct
metal-to-microorganism electron transfer; and redox-active organic
molecules shuttling electrons between Fe0 and microorganisms. These
corrosion mechanisms have been associated with phylogenetically
and physiologically diverse microorganisms (Table1 and Figs.1,2).
In some instances, corrosion mechanisms are considered rigor-
ously defined, with genetic studies or other approaches that defini-
tively rule out alternatives. In other studies, the mechanisms have only
been inferred from indirect evidence (Table1). Until recently, micro-
organisms and iron sources were typically added to poorly defined,
Table 1 | Examples of the diversity of microorganisms participating in Fe0 oxidation under anaerobic conditions and
proposed corrosion mechanisms
Microorganism Relevant physiology and
electron acceptor Fe0 as sole
donor Fe0 forms Proposed mechanism
Acetobacterium spp. Acetogen; CO2+13,124 Pure Fe0 (refs. 13,124) H2 intermediate (V)13,124
Archaeoglobus Hyperthermophile, acetogen;
CO2 heterotroph; SO42– +125; –125,126 Carbon steela,125; carbon steel126 Sulide (I)125; direct (I) but H2
utilizer125,126
Bacillus licheniformis Fermentative dissimilatory nitrate
reduction Carbon steela,97; X80 steelb,98;
stainless steelc,72 Direct (I)97,98; ribolavin (I)72
Clostridium celerecrescens Fermentative API XL 52 steeldOrganic acids54
Desulfoferrobacter sulitae H2-oxidizing sulfate reducer +Pure Fe0H2 intermediate and direct (I)127
Desulfopila corrodens (strain IS4) H2-oxidizing and lactate-oxidizing
sulfate reducer +Pure Fe0 (refs. 36,61); mild
steele,51 H2 intermediate (I); direct (I)51,61
Desulfovibrio ferrophilus (strain IS5) H2-oxidizing and lactate-oxidizing
sulfate reducer +Pure Fe0 (refs. 36,61); mild steel51 H2 intermediate (I)36; direct (I)51,61
Desulfovibrio vulgaris H2-oxidizing and lactate-oxidizing
sulfate reducer +13,38; –60 Pure Fe0 (refs. 13,38); carbon
steel60; carbon steel17 H2 intermediate(V)13,38; direct (I)60;
lavin shuttle (I)17
Dethiosulfovibrio peptidovorans Fermentative, thiosulfate reducer Mild steel Unspeciied128
Enterobacter roggenkampii Fermentative Carbon steel Organic acids44
Geobacter spp. Electroactive with fumarate,
nitrate and Fe(III) as electron
acceptors
+14,34,36 Pure Fe0 (ref. 14); stainless
steel34,36 Direct (V)14,34
Methanospirillum hungateiMethanogen +13 Pure Fe0 (ref. 13) H2 intermediate (V)13
Methanobacterium spp. Methanogen +12; +71 Pure Fe0 (ref. 12); carbon steel71 H2 intermediate (V)12; H2
intermediate and direct (I)71
Methanococcus spp. Methanogen +12 Pure Fe0 (refs. 12,59) H2 intermediate (V)12,57,59,95
Methanosarcina barkeri Methanogen +Carbon steel Electron shuttle (I)96
Methanosarcina acetivorans Methanogen +Fe0; stainless steel Direct (V)35
Prolixibactersp. Nitrate reducer Fe0Direct (I)70 or H2 (I)7
Pseudomonas aeruginosa Nitrate reducer Carbon steel Electron shuttle (I)99
Shewanella oneidensis Electroactive, fumarate, nitrate +37; –29 Carbon steel37; stainless steel29 H2 intermediate (V)37; direct (V)29,37
Sporomusa spp. Acetogen +Fe0H2 intermediate (I)124,129
Sulfurimonas sp. Sulide-oxidizing nitrate reducer Carbon steel Nitrite and sulfur intermediates100
Wolinella succinogenes Nitrate reducer +13 Pure Fe0 (ref. 13) H2 intermediate (V)13
CO2, carbon dioxide; Fe0, metallic iron; H2, hydrogen; I, mechanism inferred from indirect evidence; SO42–, sulfate; V, mechanism veriied. aCarbon steels composed of approximately 99% Fe0
combined with small amounts of carbon, manganese, phosphate, sulfur and sometimes silica. bX80 steel composition: 0.093wt% carbon, 0.069wt% silicon, 1.85wt% manganese, 0.013wt%
phosphate, 0.0068wt% sulfur, 0.3wt% chromium, 0.15wt% nickel, 0.19wt% copper, 0.038wt% aluminium, 0.011wt% titanium and balance Fe0. c316L stainless steel: 0.019wt% carbon, 0.43wt%
silicon, 1.18wt% manganese, 0.032wt% phosphate, 0.0006wt% sulfur, 10.5wt% nickel, 16.78wt% chromium, 2.09wt% molybdenum and balance Fe0. dAPI XL 52 steel: 0.11wt% carbon, 0.955wt%
manganese, 0.175wt% silicon; 0.005wt% phosphate, 0.022wt% sulfur, 0.037wt% chromium, 0.293wt% copper, 0.013wt% nickel and balance Fe0. eMild steel >99.37% iron.
Nature Reviews Microbiology
Review article
organic-rich media for mechanistic studies7,9. Elucidating the routes
for corrosion in such systems is difficult because multiple types of
microbial metabolism can take place simultaneously. Understanding
corrosion dynamics under complex conditions such as those found
in real-world situations is the ultimate goal. However, studies with Fe0
as the only electron donor available to support respiration, coupled
with quantitative monitoring of respiration and appropriate controls
to account for the possible role of electron shuttles, can yield more
readily interpretable results
7,9,14,3437
. Functional genetic studies
14,34,35,37,38
or other approaches that can associate a loss of protein function with
an inhibition of corrosion in model corrosion isolates, or possibly even
within mixed-species corrosion communities
39
, are among the most
powerful tools to reveal corrosion mechanisms. In the future, compara
-
tive molecular analysis of zones of high and low corrosion rates
4049
may reveal important microorganisms, corrosion mechanisms and
diagnostic molecular signals for corrosion without the need to first
culture isolates. With further study, other parameters, such as isotope
fractionation or type of mineral formation, might be developed into
diagnoses for different mechanisms of iron corrosion21,50,51.
H2 as an intermediary electron carrier
Electron transfer between Fe0 and microorganisms via a H2 intermediary
has been central to discussions of microbial Fe0 corrosion, as this was
first proposed as the mechanism by which sulfate reducers promote
corrosion52. Even at circumneutral pH, the protons available in water can
abiotically accept electrons from Fe
0
to generate H
2
(refs. 1214) (Fig.2a,
reaction 2). Lower pH supplies more H
+
, favouring H
2
production. Thus,
the release of CO2, acetic acid (CH3COOH) and other short-chain fatty
acids during microbial metabolism accelerates corrosion
53,54
. Micro-
bial oxidation of the H
2
at the metal surface recycles H
+
, avoiding net
H+ consumption during Fe0 oxidation and making H+ available at the
surface for additional Fe0 oxidation. Localized zones of lower pH, and
thus higher corrosion, may form within heterogeneous biofilms, lead-
ing to pitting of the metal
53,55
. Hydrogenases released from cell lysis
5658
,
or that are actively secreted outside the cell
59
, catalyse H
2
production.
H2S produced during microbial sulfate reduction may also react with
Fe
0
to enhance H
2
production, either through a direct reaction with Fe
0
(Fig.2a, reaction 3) or iron sulfide (FeS) deposits providing a conductive
surface to facilitate electron transfer from Fe0 to H+ (ref. 25).
Anaerobes aggressively compete for H2 in anaerobic environ-
ments4. Thus, rapid microbial H2 uptake is likely to maintain low H2
concentrations during corrosion, thermodynamically favouring fur-
ther reduction of H
+
to H
2
(ref. 13). When microorganisms capable of
H
2
utilization are deprived of organic electron donors
60
, more aggres-
sive corrosion may be a metabolic adaption to gain more energy from
Fe0 oxidation via H2. Diverse microorganisms reducing either nitrate,
fumarate, Fe(III), sulfate or CO2 have all been shown to anaerobically
oxidize H
2
produced from Fe
0
(Table1). Faster corrosion of Al
0
and
Zn0 in the presence of H2-utilizing sulfate-reducing bacteria suggests
that microbial H2 uptake also promotes corrosion of these metals15,16.
Several different experimental strategies have been used to dem-
onstrate the importance of H2 as the electron carrier between Fe0 and
microorganisms. Early studies physically separated Fe
0
and micro-
organisms, permitting only gas exchange12,13. In an elegant study, the
effectiveness of Methanococcus maripaludis strains in Fe
0
corrosion
was linked to the presence of genes for an extracellular hydrogenase
59
.
Studies with a mutant of Desulfovibrio vulgaris unable to use H2 dem-
onstrated that a H2 intermediary electron carrier was essential for
Fe0 to serve as an electron donor for sulfate reduction38. By contrast,
deleting the uptake hydrogenases of Shewanella oneidensis only partly
prevented corrosion, demonstrating that H
2
uptake was just one of the
routes for Fe0 oxidation37.
Methods for genetic manipulation are not yet available for many
corrosive microorganisms. An alternative approach to evaluate
whether microorganisms rely on H
2
as an electron carrier between Fe
0
and cells is to determine their ability to use 316L-grade stainless steel,
which does not generate H2, as an electron donor34,36. For example, it
was proposed that the sulfate reducers Desulfovibrio ferrophilus and
Desulfopila corrodens were capable of direct metal-to-microorganism
electron transfer because they oxidized Fe
0
much faster than other
H
2
-utilizing isolates
61
. However, such comparisons are not a defini-
tive criterion for direct electron transfer, as evidenced from the
M. maripaludis extracellular hydrogenase studies59 described above,
in which the strains that corroded Fe0 the fastest employed an extra-
cellular hydrogenase. D. ferrophilus and D. corrodens could use pure
Fe0 as an electron donor, but not 316L stainless steel, suggesting that
they relied on H2 as an electron carrier to oxidize Fe0 (ref. 36). Digest-
ing outer-surface proteins with protease did not inhibit D. ferrophilus
sulfate reduction with Fe
0
as the electron donor
62
, also suggesting
dependence on an electron carrier, such as H2, that is oxidized within
the cell. The mechanisms of Fe
0
corrosion by D. ferrophilus are of great
interest because its rate of corrosion above abiotic controls (1.6mm
per year uniform corrosion; 1.5cm per year pitting corrosion, which is
a localized form of corrosion where cavities are formed in the material)
exceeds severe corrosion rates in pipelines
63,64
. Therefore, now that
methods for making gene deletions in D. ferrophilus are available
65
, the
role of H2 as an intermediary electron carrier for D. ferrophilus corrosion
should be further evaluated with hydrogenase gene deletion studies
similar to those recently demonstrated for D. vulgaris38.
Direct metal-to-microorganism electron transfer
Direct metal-to-microorganism electron transfer, known more simply
as electrobiocorrosion, refers to corrosion when the initial electron
acceptor for electrons derived from Fe
0
is an outer-surface electrical
contact on the cell surface. Electrobiocorrosion has been observed for
several microorganisms in which the possibility for H2 uptake could
be rigorously ruled out14,29,34,36,37. In each of these studies, an outer-
surface, multi-haem c-type cytochrome was implicated as an important
electrical contact between Fe0 and the corroding microorganism. For
example, a strain of Geobacter sulfurreducens genetically modified
to prevent H
2
uptake, as well as wild-type Geobacter metallireducens,
which cannot use H2, reduced either fumarate, nitrate or Fe(III) oxide,
with pure Fe0 or stainless steel as the electron donor14,34,36. Deletion of
genes for outer-surface, multi-haem c-type cytochromes known to
be involved in other forms of extracellular electron exchange inhib-
ited electron uptake from Fe0 in both Geobacter strains. Fe0 served as
the electron donor for the reduction of CO
2
to methane by Methano-
sarcina acetivorans, which is unable to use H2 as an electron donor35.
Deletion of the gene for MmcA, a multi-haem membrane-bound c-type
cytochrome, previously shown to be involved in electron transport to
extracellular electron donors
66
and electron uptake from other cells
67
,
inhibited methane production with Fe0 as the electron donor35. All of
the cells of the Geobacter and Methanosarcina spp. seemed to be in
direct contact with the metal
14,34,67
, suggesting that potential strategies
for long-range electron transport over multiple cell lengths, such as
electrically conductive pili
5
, were probably not necessary for corro-
sion. However, the ability of some electroactive microorganisms68 and
mixed microbial communities
69
to form conductive biofilms suggests
Nature Reviews Microbiology
Review article
that, under some circumstances, microorganisms at a distance from
the metal surface could contribute to direct electron uptake from iron.
Simultaneous electron uptake via a H
2
intermediate and outer-
surface c-type cytochromes is possible
37
. Deletion of the genes for
one or more of the outer-membrane multi-haem c-type cytochrome
conduits responsible for other forms of extracellular electron exchange
in S. oneidensis partially inhibited Fe
0
oxidation
29,37
, but, as noted above,
so did deletion of uptake hydrogenases37.
Several other claims for direct metal-to-microorganism elec-
tron transfer have not been rigorously evaluated. For example,
direct electron uptake was proposed for a nitrate-reducing Prolixi-
bacter sp.
70
, but the apparent consumption of H
2
during Fe
0
corro-
sion suggests that H2 served as an intermediary electron carrier7.
Similar concerns exist for claims that Methanobacterium strains
IM1 (ref. 61) and TO1 (ref. 71) are capable of direct electron uptake
because they can utilize H2, and the possibility that their high corrosion
Initial colonization by Fe2+
oxidizers and heterotrophs
Mechanism 1:
Producing an O2-consuming bioilm to
provide a habitat for corrosive anaerobes
Mechanism 2:
Microbial recycling of Fe0 oxidants
Passivation layer of Fe(III) oxide blocks
key microbial corrosion mechanisms
Corrosion enabled
Sulfate reducer
Respiration depletes
O2 in bioilm
Corrosive anaerobes grow
in anaerobic zones
Fe2+ oxidizer
Fe0
Fe0
I2
Fe
2+
Fe
0
I
Fe3+
Mn4+ Mn2+
Fe2+
Fe2+, Mn2+ or I oxidizer
O
2
H
2
O
Anaerobe
Outer surface cytochromes
Potentially corrosive
microorganisms
Conductive minerals
X
reduced
X
oxidized
Fe
2+
Fe
0
Fe
2+
Fe
2+
Fe
2+
Fe
0
X
reduced
X
oxidized
Fe
0
Fe
0
Fe
2+
Fe
0
Fe0
H
+
, CO
2
2H
+
2H
+
H
2
H
+
H
2
, organics
Fe
2+
Fe
0
Fe
2+
Fe
0
Fe
2+
X
reduced
X
oxidized
Flavin
oxidized
Flavin
reduced
Fe
0
Fe0
Sulide
minerals
facilitate H2
production
Organic acids
neutralized and
diusion restricted
Restricted diusion
of H+ in and H2 out
limiting H2 production
Fe(III) reducers
remove Fe(III) oxide
CH3COOH
CH3COOH
CH3COO
Hydrogenase
catalyses H2
production
SO
4
2
S
2
X
reduced
X
oxidized
Lower pH favours
Fe0 oxidation
Organic acids–COOH
Fermentable substrate Xreduced = CH4, S2–, N2, Fe(II), succinateXoxidized = CO2, SO4
2–, NO3
, Fe(III), fumarate
Organic acids–COO
Fe
2+
X
reduced
X
oxidized
Fe
0
2H
+
H
2
Fe(III) oxide layer
Insulating Fe(III) oxide
blocks electron transfer
to potentially corrosive
microorganisms
Direct electron transfer to
corrosive microorganisms
H
+
H
2
H
+
H
+
H
2
H
2
H
2
Fe
2+
Fe
0
Fe
2+
Fe
0
H
+
Fe
2+
Fe
0
CO
2
CH
4
SO
4
2
S
2
Heterotroph
a
b
c
d
Nature Reviews Microbiology
Review article
rates are due to the release of hydrogenases or other factors cannot
be discarded.
Organic molecules as electron carriers between metals and
microorganisms
Redox-active organic molecules are possible alternatives to H2 as
intermediary electron carriers between Fe
0
and anaerobic respira-
tory microorganisms during corrosion (Fig.3c). Organic electron
shuttles important in other forms of extracellular electron exchange,
such as mineral reduction and interspecies electron exchange, include
humic substances and a diversity of microbial metabolites, such as
flavins, phenazines and pyocyanins5. Faster corrosion in cultures of
sulfate-reducing D. vulgaris or D. ferrophilus was attributed to ribofla-
vin functioning as an electron shuttle
63,7274
. However, it has yet to be
demonstrated that Fe0 reduces riboflavin, or that reduced riboflavin
can serve as the electron donor for sulfate reduction, the two reactions
necessary for riboflavin to function as a shuttle. Riboflavin amend-
ments did not stimulate sulfate reduction in cultures of D. vulgaris with
Fe
0
as the sole electron donor
38
and the midpoint potential of flavins
seems to be too positive for flavins to function as an electron shuttle
for microbial reduction of sulfate to sulfide
25,38
. Flavins play multiple
roles in microbial physiology, with functioning as extracellular elec-
tron shuttles being one of the least common. Strategies designed to
specifically detect electron shuttling, such as separating the corrod-
ing metal and the microorganisms by incorporating the metal within
microporous beads
75
or genetic modifications that prevent extracel-
lular flavin release76, are potential approaches to further evaluate the
role of organic electron shuttles in Fe0 corrosion.
Pseudomonas aeruginosa accelerated Ti0 corrosion via a mecha-
nism that involved secretion of phenazine-1-carboxylate77. One possi-
bility is that the phenazine is an electron shuttle, accepting electrons
from Ti0, with reduced phenazine serving as an electron donor for O2
respiration. However, key reactions in this model, such as Ti0 reduction
of phenazine and respiration with reduced phenazine as the electron
donor, have yet to be verified.
Another less well-studied impact of anaerobes on corrosion is the
removal or partial damage of protective surfaces known as passivation
layers. For example, the accumulation of Fe(III) oxides, produced either
by Fe2+-oxidizing bacteria or abiotic oxidation (Fig.2a), may obstruct
corrosive microorganisms from attacking the metal. Microbial Fe(III)
reduction can remove the protective layer (Fig.3d), accelerating cor-
rosion
10,7881
. Extracellular production of hydrogen peroxide associ-
ated with P. aeruginosa release of reduced phenazine oxidized Cr(III)
oxide — an important component of the passivation layer on stainless
steel — to Cr6+, making the stainless steel more vulnerable to corrosion82.
Acidic metabolites may also attack minerals in passivation layers
83
.
Passive film damage is particularly important for very active metals
such as Ti0, Al0 and Zn0, because they readily react with water without
a passive film.
Microbial diversity in metal corrosion
In addition to the diversity of microorganisms shown to enhance
corrosion of metals in pure culture (Table1), molecular analyses of
microbial communities associated with corroding metal surfaces have
described a wide diversity of bacteria and archaea enriched within
corrosion biofilms
30,4145,47,49,8490
. Physiological diversity (Table1 and
Fig.2b) determined from studies with defined cultures includes aerobes
(heterotrophs, iron-oxidizing and sulfur-oxidizing bacteria), facul-
tative microorganisms (heterotrophs, nitrate reducers) and diverse
anaerobes including fermenters, dissimilatory nitrate reducers, Fe(III)
and Mn(IV) reducers, sulfate reducers, methanogens and acetogens.
In molecular studies of mixed communities, phylogenetic characteri-
zation, which has often been at the level of family or above, provides
insufficient physiological information to infer a substantial role in
corrosion because corrosion rates can vary substantially even between
species of the same genus
63
. Furthermore, some microorganisms within
corrosion biofilms may not significantly contribute to corrosion but
simply utilize the metal surface as a support for growth on nutrients
in the surrounding environment and/or nutrients released from the
microorganisms actively engaged in corrosion43,47,48,86,90,91.
Although progress is being made towards identifying genes that
could indicate a high potential for microbial corrosion within a micro-
bial community46, it is not yet possible to predict corrosion mechanisms
from the presence of specific genes. For example, although gene dele-
tion studies have indicated that in some microorganisms outer-surface,
multi-haem cytochromes are important electrical contacts for direct
electron uptake from Fe
0
(refs.
14,34,35,37
), the presence of similar genes in
other microorganisms may not be indicative of a direct electron uptake
capability
36
. Many microorganisms that are capable of exchanging
electrons with other microbial species, minerals or electrodes lack
outer-surface c-type cytochromes5, and might also electrically interact
with metals during corrosion. The presence of hydrogenase genes is
not necessarily predictive of the ability of microorganisms to use H2 as
an intermediary electron carrier for corrosion35.
Therefore, it is still necessary to study representative isolates to
determine whether microorganisms within corrosion biofilms are likely
to have the possibility to enhance corrosion and to evaluate their cor-
rosive mechanism(s). Corrosion is often associated with a high abun-
dance and metabolic activity of sulfate reducers
41,45,48,9193
, most notably
Desulfovibrio spp., which have served as model microorganisms for
Fig. 3 | Details of key mechanisms for microbial corrosion of ferrous metals.
a, Under aerobic conditions, iron-oxidizing microorganisms colonize the
metal surface and develop biofilms that provide zones of low oxygen (O2)that
enable the growth of corrosive anaerobes, such as sulfate reducers. In another
mechanism, iron-oxidizing, manganese-oxidizing and iodine-oxidizing
microorganisms generate oxidants for abiotic metallic iron (Fe0)oxidation.
b, Under anaerobic conditions, sulfide minerals formed as the result of microbial
sulfate reduction, and protons released from organic acid fermentation products
promote the oxidation of Fe0 coupled with the reduction of proton (H+)to
hydrogen (H2), as do extracellular hydrogenases released from microorganisms.
c, Anaerobic respiration with Fe0 as the electron donor can occur with
direct metal-to-microorganism electron transfer facilitated by multi-haem
outer-surface c-type cytochromesorconductive minerals. Alternatively, soluble
redox-active molecules, such as flavin, function as an electron shuttle between
Fe0 and the microorganism. Lightning bolts designate direct electron transfer
from Fe0 to cells. d, Fe(III) oxides protect Fe0 from corrosion because they are an
insulating layer preventing direct electron transfer from Fe0 to microorganisms
and restricting access of organic acids and other sources of H+, thus limiting H+
availability for Fe0 oxidation coupled to H+ reduction to H2. When Fe(III) reducers
remove Fe(III) oxides, Fe0 oxidation coupled either to direct electron transfer
or H+ reduction is possible. H2 oxidizers, such as sulfate reducers, oxidize H2,
resupplying H+ for additional Fe0 oxidation. CH4, methane; CH3COOH, acetic
acid; CO2, carbon dioxide; Fe2+, ferrous ion; Fe3+, ferric ion; I, iodine ion;
I2, elemental iodine; NO3, nitrate; S2–, sulfide; SO42–, sulfate.
Nature Reviews Microbiology
Review article
corrosion studies since the earliest investigations on microbial cor-
rosion10,11,52. More than 50 corrosion studies have focused on just one
species, D. vulgaris
11
, which as noted above relies on H
2
as an inter-
mediary electron carrier for Fe0 oxidation38. For all sulfate reducers
there is the possibility that production of H2S may directly lead to a
non-enzymatic corrosive reaction of sulfide with Fe
0
, especially at low
pH10 (Fig.2a, reaction 3).
Methanogens can also have an important role in corro-
sion12,13,30,35,46,50,57,59,61,71,9496. Many methanogens grow with the H2
released from Fe0 as the sole electron donor, as can the corrosive ace-
togens studied to date (Table1). Direct electron uptake has only been
rigorously demonstrated for M. acetivorans35.
As noted above, the Fe(III)-reducing microorganisms G. sulfurredu-
cens, G. metallireducens and S. oneidensis are capable of direct electron
uptake from Fe0 (refs. 14,29,34,37), but both S. oneidensis37 and G. sul-
furreducens14 can also accept electrons from Fe0 via H2. G. metalliredu-
cens and S. oneidensis also directly accept electrons from Fe
0
with nitrate
as the electron acceptor. Amongst other corrosive nitrate reducers,
Wolinella succinogenes grows with H
2
evolved from Fe
0
(ref. 13) whereas
it has been inferred that Bacillus licheniformis72,97,98 and P. aeruginosa99
may take up electrons via direct metal-to-microorganism electron
transfer or an organic electron shuttle (Table1). Nitrite generated from
nitrate reduction may also act as an abiotic oxidant for Fe0 (ref. 100).
Corrosion has primarily been studied with isolates recovered
under culture conditions in which the ability to interact with metals
was not a selective pressure. A better understanding of the diversity of
microorganisms contributing to corrosion will require a rethinking
of isolation strategies to isolate microorganisms representative of
those that predominate in corrosive biofilms7,9. Due to the complex
-
ity of microbial communities involved in corrosion, it seems unlikely
that any one isolate will be able to replicate the full range of corrosive
reactions likely to take place in mature, highly corrosive biofilms
21,86,101
.
Therefore, in order to approximate the microbial corrosive activity in
the real world with laboratory studies it will be necessary to mimic the
diverse communities with biofilms built from appropriate combina-
tions of diverse physiological microbial types
21,22,84,102104
. For example,
the combination of an Fe(II)-oxidizing isolate with an Fe(III) reducer21
or a sulfate reducer
22
accelerated corrosion beyond that achieved with
individual isolates. Iron was corroded faster when an Fe(III)-reducing
strain of Geobacillus was added to a consortium of two fermentative
Bacillus spp. and the sulfate reducer Desulfotomaculum sp. SRB-M
79
.
In a study involving two species of sulfate reducers, a syntroph and
a methanogen, different combinations of co-cultures and culture
conditions demonstrated the importance of both sulfide and acetate
production in promoting corrosion102.
Mitigating microbial corrosion
Strategies for mitigating microbial corrosion typically rely on physical
and chemical methods for removing biofilms, but biological-based
techniques are evolving (Fig.4). Periodic scrapping to remove biofilms
Ag+
HO
HO
OH
OH
N N
OO
O
O
Pipeline pig Biocide cocktail
Flow
Physical scrubbing (pigging)
with biocide
Cu
Cu
Ag
Ag
Ce Ce
Cu2+
Ce3+
aNovel metal alloy
b
Live cell Dead cell Polymer coating
Biologically produced chemicals
Chelator Surfactant
Metal
Dispersing
Nature-mimicking
peptide
Quorum-sensing
inhibitor
Phage
Replication
D-Amino acid
Coatings
cAntibioilm agents
d
Bioilm dispersal
ePhage therapy
fBifunctional corrosion inhibitor
gProtective bioilm
h
Smart coating
Antimicrobial
Metal
Agent to heal
coating damage
HO
H2N
OH
O
Live cell
Dead cell Antimicrobial
molecule
Corrosion barrier
EPS
Corrosive
microbe barrier
Biomineralization layer
O2 barrier
O2 corrosion
Cell lysis
Fig. 4 | Various methods for disrupting biofilms to prevent or mitigate
microbial metal corrosion. a, Surface scrubbing of a biofilm using a pipeline
pig tool (pigging) coupled with biocide application. b, Novel metal alloys that
incorporate silver(Ag), copper (Cu)and cerium(Ce) release metal ions that
are toxic to corroding microorganisms. c, Plain polymer coating or smart
coating with self-healing ability to repair damage to the coating and release
antimicrobials, thus protecting the metal from corrosion. d, Antibiofilm agents
that inhibit biofilm growth, such as EDTA or surfactant molecules, weaken
microbial activity in biofilms or eliminate biofilms. e, Diverse chemical agents
can disperse biofilms. f, Phage treatment to lyse biofilm cells. g, Bifunctional
corrosion inhibitor molecules can both protect the metal surface and have
antimicrobial activity. h, Protective biofilm can keep corrosive microorganisms
and abiotic corrosive agents, such as oxygen (O2), away from the metal surface.
Cu2+, cupric ion; EPS, exopolymeric substances. Nature-mimicking peptide in
part e adapted with permissionfrom ref. 108, Elsevier.
Nature Reviews Microbiology
Review article
as well as periodic chemical treatments that broadly kill the micro
-
organisms within biofilms is the most common approach105. To limit
corrosion, iron can be alloyed with other metals that either slowly
release biocidal metal ions, such as copper, or with chromium, which
forms a protective oxide layer on the iron surface
7
. Various coatings
that are toxic to microorganisms and/or prevent attachment have
also been developed
7
. However, biological concepts that may be more
environmentally friendly and cost-effective are emerging.
For example, biologically produced chemicals such as -limonene,
a component of citrus peels, may be effective antimicrobials106.
-Amino acids
107
or peptides
108
that either kill microorganisms or serve
as a signal for biofilm dispersal can enhance the effectiveness of tradi
-
tional biocides7,109. Deploying microorganisms that can produce killing
or dispersal agents offers the possibility of continuous in situ genera-
tion of biofilm inhibitors at low cost110. Predatory Bdellovibrio111 or quo-
rum sensing inhibitors112 reduced steel corrosion by sulfate reducers
in laboratory incubations. However, these approaches, or applying
‘cocktails’ of bacteriophages to lyse corrosive bacterial biofilms
109
,
remain aspirational corrosion mitigation goals.
Encouraging the growth of self-sustaining biofilms that can block
corrosive agents is an attractive concept also in need of further devel-
opment
7
. In aerobic environments, biofilms that cover metal surfaces
with an O2-consuming layer limit O2 access to the metal, thus prevent-
ing abiotic oxidation
109
. Reduced minerals formed within biofilms can
also contribute to O
2
removal. Extracellular polymeric substances
can block O
2
, metabolites or microorganisms from accessing metal
surfaces113. Microbially produced carbonate minerals can protect
metal surfaces114. Natural communities contributing to carbonate
formation likely protected the iron sheet piles that reinforce waterways
and dykes in the Netherlands from corrosion over a 50-year deploy-
ment115,116. A highly diverse microbial community, enriched in organic
acid-oxidizing bacteria and methanogens, was enriched near the piles
and was associated with the accumulation of carbonates. An improved
understanding of the mechanisms by which microbial communities in
soils and sediments can promote carbonate precipitation near metal
surfaces could lead to important new strategies for sustainable and
inexpensive corrosion prevention.
An effective field-scale manipulation of the microbial commu-
nity to diminish corrosion is the addition of nitrate, which effectively
promotes the growth of nitrate reducers and inhibits the growth and
activity of more corrosive sulfate reducers
117
. Other microbially based
mitigation strategies that have been based on laboratory studies, typi-
cally with a limited diversity of potentially corrosive microorganisms,
will require much more extensive testing and development before
they will be field applicable7,109.
Outlook
Research to date has conclusively demonstrated that microbial corro-
sion is a serious economic problem that threatens industry, health and
the environment. However, the process is still so poorly understood
that there are inadequate strategies to definitively detect and sustain-
ably mitigate this damaging activity. Improved understanding of which
microorganisms are responsible for corrosion and how they corrode
is essential. Molecular analysis of corrosion communities4049 will be
an important approach, but more detailed mechanistic studies are
required before it will be possible to meaningfully develop a ‘molecu-
lar or sequence-based Koch’s postulates’118,119 approach to diagnose
microbial corrosion and monitor the success of mitigation strategies.
Needed are attempts to apply Koch’s original culture-based postu-
lates to microbial corrosion with skilful isolation of microorganisms
from corroding materials and evaluation of their corrosive capabili-
ties
7,9
. Comparative genomics and transcriptomics of strains that are
closely related phylogenetically and metabolically, but differ greatly
in their corrosion capabilities, is expected to be a productive avenue
for elucidating corrosion mechanisms.
The development of genetic methods to rigorously evaluate possi-
ble corrosion mechanisms in novel microorganisms will also be needed.
Beyond providing new diagnostic molecular tools for corrosion,
Box 2
Diagnosis and monitoring of microbial corrosion
Diagnosing and preventing microbial corrosion requires methods
not only to estimate the rate and extent of corrosion but, ideally, also
to discriminate between abiotic corrosion and corrosion in which
microorganisms are the causative agents, because the mitigation
of each type of corrosion may be dierent. At present, there is
insuicient information to predict corrosion rates or mechanisms,
or even to deinitely ascertain whether corrosion is taking place,
based on an assay of any microbial genes or proteins7,25. In laboratory
studies, rates of corrosion are commonly determined from
measurements of metal weight loss and assessments of metal pitting
with confocal, atomic force or scanning electron microscopy29,34,37.
A diversity of electrochemical techniques in which the corroding
metal of interest serves as the working electrode are available
and can provide online monitoring with near real-time estimates
of microbial corrosion rates without disrupting bioilm activity7,139.
Scanning electrochemical microscopy is an exciting technology for
simultaneous high-resolution topographical and chemical species
mapping within liquid environments140. Fine-scale electrochemical
variations on the surface of corroding metal can be mapped,
associating microbial cells with microbially catalysed redox reactions
within heterogeneous environments141. Microsensors can potentially
elucidate the availability of electron acceptors and nutrients, as
well as pH and metabolites, at a ine scale within corrosion bioilms
to provide better insights into environmental factors inluencing
microbial activity and reactions at the metal–bioilm interface7,139.
X-ray photo-electron spectroscopy, X-ray diraction and Fourier-
transform infrared spectroscopy can aid in identifying passive ilms
and the mineral products of corrosion7, potentially linking speciic
minerals with mechanisms of corrosion. However, most of the
corrosion monitoring approaches described here are not translatable
to monitoring of corrosion in open, dynamic environments, outside
the laboratory. We need new tools for continuous or periodic
monitoring of microbial abundance, community composition and
activity in order to track microbial corrosion in real time.
Nature Reviews Microbiology
Review article
mechanistic studies will likely yield insights into novel mitigation
approaches. For example, if direct electron uptake from Fe0 via multi-
haem c-type cytochromes and/or H2-mediated metal-to-microorganism
electron uptake are found to be the most important mechanisms for
corrosion, then a search for treatments that specifically target the
cytochromes and hydrogenases will be warranted.
As physiological properties and contributions of individual
components of corrosion biofilms are better understood, there will
be a need to develop an understanding of the web of interactions
between individual populations within biofilms. Most microbial cor-
rosion studies have been conducted with static incubations, over short
times, often in the presence of high concentrations of readily degraded
organic substrates. More realistic conditions of flow, longer times
and organic concentrations and types are likely to yield results more
relevant to actual corrosion environments. Methods are also needed
to specifically evaluate the activity of the ‘bottom feeders’ that are
localized at the metal–biofilm interface and likely to be most active in
corrosion processes even though they may not be the most numerous
microorganisms within the biofilm.
An eventual goal is to build genome-scale metabolic models that
describe the activity of microorganisms in corrosion biofilms and their
interactions with other biofilm components
93
, analogous to similar
applications in host microbiomes, bioremediation, wastewater treat-
ment and biogeochemical cycling120123. Advances in electrochemical
and imaging methods (Box2) will make crucial contributions to this
complex, interdisciplinary problem25.
There are great opportunities for basic research in microbial metal
corrosion that not only can aid in solving this economically important
problem but also can expand insights into biofilm ecology beyond
those already being developed primarily from the study of medically
relevant biofilms. Thus, the investigation of metal corrosion is highly
recommended to upcoming generations of microbial ecologists,
physiologists and environmental engineers.
Published online: xx xx xxxx
References
1. Camara, M. et al. Economic signiicance of bioilms: a multidisciplinary and cross-
sectoral challenge. npj Bioilms Microbiomes 8, 42 (2022).
This perspective article discusses the substantial economic costs of microbial
corrosive activity.
2. Jacobson, G. A. Corrosion at Prudhoe Bay — a lesson on the line. Mater. Perform. 46,
26–34 (2007).
3. Spark, A., Wang, K., Cole, I., Law, D. & Ward, L. Microbiologically inluenced corrosion:
a review of the studies conducted on buried pipelines. Corros. Rev. 38, 231–262
(2020).
4. Morita, R. Y. Is H2 the universal energy source for long-term survival? Microb. Ecol. 38,
307–320 (2000).
This study discusses the central role of H2 in diverse microbial communities.
5. Lovley, D. R. & Holmes, D. E. Electromicrobiology: the ecophysiology of phylogenetically
diverse electroactive microorganisms. Nat. Rev. Microbiol. 20, 5–19 (2022).
6. Lovley, D. R. Electrotrophy: other microbial species, iron, and electrodes as electron
donors for microbial respirations. Bioresour. Technol. 345, 126553 (2022).
7. Lekbach, Y. et al. Microbial corrosion of metals — the corrosion microbiome. Adv. Microb.
Physiol. 78, 317–390 (2021).
8. Gaines, R. Bacterial activity as a corrosive inluence in the soil. J. Ind. Eng. Chem. 2,
128–130 (1910).
This study provides one of the earliest fact-documented evidences of corrosive
activity of microorganisms on metal.
9. Enning, D. & Garrelfs, J. Corrosion of iron by sulfate-reducing bacteria: new views of an
old problem. Appl. Environ. Microbiol. 80, 1226–1236 (2014).
10. Hamilton, W. A. Sulphate-reducing bacteria and anaerobic corrosion. Annu. Rev.
Microbiol. 39, 195–217 (1985).
This study presents a comprehensive historical account of early studies on the role
of sulfate-reducing bacteria in corrosion.
11. Ueki, T. & Lovley, D. R. Desulfovibrio vulgaris as a model microbe for the study of
corrosion under sulfatereducing conditions. mLife 1, 13–20 (2022).
12. Daniels, L., Belay, N., Rajagopal, B. S. & Weimer, P. J. Bacterial methanogensis and growth
from CO2 with elemental iron as the sole source of electrons. Science 237, 509–511 (1987).
13. Rajagopal, B. S. & LeGall, J. Utilization of cathodic hydrogen by hydrogen-oxidizing
bacteria. Appl. Microbiol. Biotechnol. 31, 406–412 (1989).
14. Tang, H.-Y., Holmes, D. E., Ueki, T., Palacios, P. A. & Lovley, D. R. Iron corrosion via direct
metal-microbe electron transfer. mBio 10, e00303–e00319 (2019).
To our knowledge, this study is the irst to rigorously demonstrate direct
metal-to-microorganism electron transfer with genetic analyses.
15. Guan, F. et al. Interaction between sulfate−reducing bacteria and aluminum alloys —
corrosion mechanisms of 5052 and Al–Zn–In–Cd aluminum alloys. J. Mater. Sci. Technol.
36, 55–64 (2020).
16. Wang, D. et al. Sulfate reducing bacterium Desulfovibrio vulgaris caused severe
microbiologically inluenced corrosion of zinc and galvanized steel. Int. Biodeterior.
Biodegrad. 157, 105160 (2021).
17. Wang, D. et al. Distinguishing two dierent microbiologically inluenced corrosion (MIC)
mechanisms using an electron mediator and hydrogen evolution detection. Corros. Sci.
177, 108993 (2020).
18. Chen, Z., Dou, W., Chen, S., Pu, Y. & Xu, Z. Inluence of nutrition on Cu corrosion by
Desulfovibrio vulgaris in anaerobic environment. Bioelectrochemistry 144, 108040
(2022).
19. Emerson, D. The role of iron-oxidizing bacteria in biocorrosion: a review. Biofouling 34,
989–1000 (2019).
This review presents important insights into how iron-oxidizing bacteria can
contribute to metal corrosion.
20. Wang, H., Ju, L.-K., Castandea, H., Cheng, G. & Newby, B. Z. Corrosion of carbon
steel C1010 in the presence of iron oxidizing bacteria Acidithiobacillus ferrooxidans.
Corros. Sci. 89, 250–257 (2014).
21. Lee, J. S., McBeth, J. M., Ray, R. I., Little, B. J. & Emerson, D. Iron cycling at corroding
carbon steel surfaces. Biofouling 29, 1243–1252 (2013).
22. Liu, H. et al. Corrosion behavior of carbon steel in the presence of sulfate reducing
bacteria and iron oxidizing bacteria cultured in oilield produced water. Corros. Sci.
100, 484–495 (2015).
23. Chen, S., Deng, H., Liu, G. & Zhang, D. Corrosioin of Q235 carbon steel in seawater
containing Mariprofundus ferrooxydans and Thalassospira sp. Front. Microbiol. 10,
936 (2020).
24. Yue, Y., Lv, M. & Du, M. The corrosion behavior and mechanism of X65 steel induced by
ironoxidizing bacteria in the seawater environment. Mater. Corros. 70, 1852–1861 (2019).
25. Little, B., Hinks, J. & Blackwood, D. J. Microbially inluenced corrosion: towards an
interdisciplinary perspective on mechanisms. Int. Biodeterior. Biodegrad. 154, 105062
(2020).
This article presents an interdisciplinary analysis and synthesis of a wide range of
abiotic and biotic corrosion studies.
26. Hamilton, W. A. Microbially inluenced corrosion as a model system for the study of
metal microbe interactions: a unifying electron transfer hypothesis. Biofouling 19, 65–76
(2003).
27. Dexter, S. C., Xu, K. & Luther, G. L. Mn cycling in marine bioilms: eect on the rate
of localized corrosion. Biofouling 19, 139–149 (2003).
28. Wakai, S. et al. Corrosion of iron by iodide-oxidizing bacteria isolated from brine in an
iodine production facility. Microb. Ecol. 68, 519–527 (2014).
This study shows that a diversity of redox-active components can serve as electron
shuttles to support corrosion.
29. Zhou, E. et al. Direct microbial electron uptake as a mechanism for stainless steel
corrosion in aerobic environments. Water Res. 219, 118553 (2022).
This study shows that aerobic respiration generates anaerobic conditions within
bioilms promoting corrosive microbial electron uptake.
30. Usher, K. M., Kaksonen, A. H. & MacLeod, I. D. Marine rust tubercles harbour iron
corroding archaea and sulphate reducing bacteria. Corros. Sci. 83, 189–197 (2014).
31. Wang, J. et al. Corrosion behavior of Aspergillus niger on 7075 aluminum alloy and the
inhibition eect of zinc pyrithione biocide. J. Electrochem. Soc. 166, G39–G46 (2019).
32. Zhang, T., Wang, J., Zhang, G. & Liu, H. The corrosion promoting mechanism of
Aspergillus niger on 5083 aluminum alloy and inhibition performance of miconazole
nitrate. Corros. Sci. 176, 108930 (2020).
33. Zhao, J., Csetenyi, L. & Gadd, G. M. Biocorrosion of copper metal by Aspergillus niger.
Int. Biodeterior. Biodegrad. 154, 105081 (2020).
34. Tang, H.-Y. et al. Stainless steel corrosion via direct iron-to-microbe electron transfer
by Geobacter species. ISME J. 15, 3084–3093 (2021).
This study demonstrates the lack of H2 generation from stainless steel as a tool for
distinguishing the role of H2 in corrosion.
35. Holmes, D. E. et al. Cytochrome-mediated direct electron uptake from metallic iron
by Methanosarcina acetivorans. mLife 1, 443–447 (2022).
To our knowledge, this study provides the irst genetic evidence for direct electron
uptake from Fe0 by a methanogen.
36. Liang, D. et al. Extracellular electron exchange capabilities of Desulfovibrio ferrophilus
and Desulfopila corrodens. Environ. Sci. Technol. 55, 16195–16203 (2021).
37. Hernandez-Santana, A., Sulita, J. M. & Nanny, M. A. Shewanella oneidensis MR-1
accelerates the corrosion of carbon steel using multiple electron transfer mechanisms.
Int. Biodeterior. Biodegrad. 173, 105439 (2022).
This study presents a genetic approach to elucidate multiple iron corrosion
mechanisms within one microorganism.
Nature Reviews Microbiology
Review article
38. Woodard, T. L., Ueki, T. & Lovley, D. R. H2 is a major intermediate in Desulfovibrio vulgaris
corrosion of Iron. mBio 14, e00076-23 (2023).
39. Rubin, B. E. et al. Species-and site-speciic genome editing in complex bacterial
communities. Nat. Microbiol. 7, 34–47 (2022).
40. Neria, I., Wang, E. T., Ramirez, F., Romero, J. M. & Hernandez-Rodriguez, C.
Characterization of bacterial community associated to bioilms of corroded oil pipelines
from the southeast of Mexico. Anaerobe 12, 122–133 (2006).
41. Paisse, S. et al. Sulfate-reducing bacteria inhabiting natural corrosion deposits from
marine steel structures. Appl. Microbiol. Biotechnol. 97, 749–7504 (2013).
42. Zhang, G. et al. The bacterial community signiicantly promotes cast iron corrosion in
reclaimed wastewater distribution systems. Microbiome 6, 222 (2018).
43. Mugge, R. L., Lee, J. S., Brown, T. T. & Hamdan, L. J. Marine bioilm bacterial community
response and carbon steel loss following Deepwater Horizon spill contaminant
exposure. Biofouling 35, 870–882 (2019).
44. Salgar-Chaparro, S. J., Darwin, A., Kaksonen, A. H. & Machuca, L. L. Carbon steel corrosion
by bacteria from failed seal rings at an oshore facility. Sci. Rep. 10, 12287 (2020).
45. Salgar-Chaparro, S. J., Lepkova, K., Pojtanabuntoeng, T., Darwin, A. & Machuca, L. L.
Nutrient level determines bioilm characteristics and subsequent impact on microbial
corrosion and biocide eectiveness. Appl. Environ. Microbiol. 86, e02885-19 (2020).
46. Lahme, S., Mand, J., Longwell, J., Smith, R. & Enning, D. Severe corrosion of carbon steel
in oil ield produced water can be linked to methanogenic archaea containing a special
type of [NiFe] hydrogenase. Appl. Environ. Microbiol. 87, e01819–e01820 (2021).
This study demonstrates the possibility of diagnosing microbial corrosion
mechanisms with molecular analysis.
47. Garrison, C. E. & Field, E. K. Introducing a “core steel microbiome” and community
functional analysis associated with microbially inluenced corrosion. FEMS Microb. Ecol.
97, iaa237 (2021).
48. Wakai, S. et al. Dynamics of microbial communities on the corrosion behavior of steel in
freshwater environment. npj Mater. Degrad. 6, 45 (2022).
This work shows the dynamic succession of microbial communities during the initial
stages of metal corrosion.
49. Gosi, P. et al. Prediction of long-term localized corrosion rates in a carbon steel cooling
water system is enhanced by metagenome analysis. Eng. Fail. Anal. 141, 106733 (2022).
50. Tamisier, M. et al. Iron corrosion by methanogenic archaea characterized by stable
isotope eects and crust mineralogy. Environ. Microbiol. 24, 583–595 (2022).
51. Enning, D. et al. Marine sulfate-reducing bacteria cause serious corrosion of iron under
electroconductive biogenic mineral crust. Environ. Microbiol. 14, 1772–1787 (2012).
This study shows high rates of corrosion by D. ferrophilus and D. corrodens,
highlighting the need to elucidate corrosive mechanisms in these microorganisms.
52. von Wolzogen Kűhr, C. A. H. & van der Vlugt, L. S. The graphitization of cast iron as an
electrobiochemical process in anaerobic soil. Water 18, 147–165 (1934).
53. Sulita, J. M., Phelps, T. J. & Little, B. Carbon dixoide corrosion and acetate: a hypothesis
on the inluence of microorganisms. Corrosion 64, 854–859 (2008).
54. Ramos Monroy, O. A., Ruiz Ordaz, N., Hernández Gayosso, M. J., Juárez Ramírez, C.
& Galíndez Mayer, J. The corrosion process caused by the activity of the anaerobic
sporulated bacterium Clostridium celerecrescens on API XL 52 steel. Environ. Sci. Pollut.
Res. 26, 29991–30002 (2019).
55. Kryachko, Y. & Hemmingsen, S. M. The role of localized acidity generation in microbialy
inlfuenced corrosion. Curr. Microbiol. 74, 870–876 (2017).
56. Chatelus, C. et al. Hydrogenase activity in aged, nonviable Desulfovibrio vulgaris cultures
and its signiicance in anaerobic biocorrosion. Appl. Environ. Microbiol. 53, 1708–1710
(1987).
57. Deutzmann, J. S., Sahin, M. & Spormann, A. M. Extracellular enzymes facilitate electron
uptake in biocorrosion and bioelectrosynthesis. mBio 6, e00496-15 (2015).
This study demonstrates that assumptions of direct electron uptake are inadvisable
without rigorous experimental validation.
58. Rouvre, I. & Basseguy, R. Exacerbation of the mild steel corrosion process by direct
electron transfer between [Fe–Fe]-hydrogenase and material surface. Corros. Sci. 111,
199–211 (2016).
59. Tsurumaru, H. et al. An extracellular [NiFe] hydrogenase mediating iron corrosion is
encoded in a genetically unstable genomic island in Methanococcus maripaludis. Sci.
Rep. 8, 1–10 (2018).
This study reveals an important mechanism for microbial stimulation of H2 production
from corroding iron.
60. Dou, W. et al. Electrochemical investigation of increased carbon steel corrosion via
extracellular electron transfer by a sulfate reducing bacterium under carbon source
starvation. Corros. Sci. 150, 258–267 (2019).
This study shows that the physiological state of microorganisms may be an important
determinant impacting corrosive activity.
61. Dinh, H. T. et al. Iron corrosion by novel anaerobic microorganisms. Nature 427, 829–832
(2004).
62. McCully, A. L. & Spormann, A. M. Direct cathodic electron uptake coupled to sulfate
reduction by Desulfovibrio ferrophilus IS5 bioilms. Environ. Microbiol. 22, 4794–4807
(2020).
63. Wang, D. et al. Aggressive corrosion of carbon steel by Desulfovibrio ferrophilus IS5
bioilm was further accelerated by ribolavin. Bioelectrochemistry 142, 107920 (2021).
64. Xu, L., Kijkla, P., Kumseranee, S., Punpruk, S. & Gu, T. “Corrosion-resistant” chromium steels
for oil and gas pipelines can suer from very severe pitting corrosion by a sulfate-reducing
bacterium. J. Mater. Sci. Technol. https://doi.org/10.1016/j.jmst.2023.01.008 (2023).
65. Ueki, T., Woodard, T. L. & Lovley, D. R. Genetic manipulation of Desulfovibrio ferrophilus
and evaluation of Fe(III) oxide reduction mechanisms. Microbiol. Spectr. 10, e0392222
(2022).
66. Holmes, D. E. et al. A membrane-bound cytochrome enables Methanosarcina acetivorans
to conserve energy from extracellular electron transfer. mBio 10, e00789-19 (2019).
67. Holmes, D. E., Zhou, J., Ueki, T., Woodard, T. L. & Lovley, D. R. Mechanisms for electron
uptake by Methanosarcina acetivorans during direct interspecies electron transfer. mBio
12, e02344-21 (2021).
68. Malvankar, N. S. et al. Tunable metallic-like conductivity in nanostructured bioilms
comprised of microbial nanowires. Nat. Nanotechnol. 6, 573–579 (2011).
69. Shrestha, P. M. et al. Correlation between microbial community and granule conductivity
in anaerobic bioreactors for brewery wastewater treatment. Bioresour. Tech. 174,
306–310 (2014).
70. Iino, T. et al. Iron corrosion induced by nonhydrogenotrophic nitrate-reducing
Prolixibacter sp. strain MIC1-1. Appl. Environ. Microbiol. 81, 1839–1846 (2015).
71. Hirano, S. et al. Novel Methanobacterium strain induces severe corrosion by retrieving
electrons from Fe0 under a freshwater environment. Microorganisms 10, 270 (2022).
72. Jin, Y. et al. Sharing ribolavin as an electron shuttle enhances the corrosivity of a mixed
consortium of Shewanella oneidensis and Bacillus licheniformis against 316L stainless
steel. Electrochim. Acta 316, 93–104 (2019).
73. Li, H. et al. Extracellular electron transfer is a bottleneck in the microbiologically
inluenced corrosion of C1018 carbon steel by the bioilm of sulfate-reducing bacterium
Desulfovibrio vulgaris. PLoS ONE 10, e0136183 (2015).
74. Zhang, P., Xu, D., Li, Y., Yang, K. & Gu, T. Electron mediators accelerate the
microbiologically inluenced corrosion of 304 stainless steel by the Desulfovibrio
vulgaris bioilm. Bioelectrochemistry 101, 14–21 (2015).
75. Nevin, K. P. & Lovley, D. R. Novel mechanisms for accessing insoluble Fe(III) oxide during
dissimilatory Fe(III) reduction by Geothrix fermentans. Appl. Environ. Microbiol. 68,
2294–2299 (2002).
76. Kotloski, N. J. & Gralnick, J. A. Flavin electron shuttles dominate extracellular electron
transfer by Shewanella oneidensis. mBio 4, e00553-12 (2013).
77. Liu, D. et al. Electron transfer mediator PCN secreted by aerobic marine Pseudomonas
aeruginosa accelerates microbiologically inluenced corrosion of TC4 titanium alloy.
J. Mater. Sci. Technol. 79, 101–108 (2021).
78. Herrera, L. K. & Videla, H. A. Role of iron-reducing bacteria in corrosion and protection
of carbon steel. Int. Biodeterior. Biodegrad. 63, 891–895 (2009).
79. Valencia-Cantero, E. & Pena-Cabriales, J. J. Eects of iron-reducing bacteria on carbon
steel corrosion induced by thermophilic sulfate-reducing consortia. J. Microbiol.
Biotechnol. 24, 280–286 (2014).
80. Hu, Y. et al. Microbiologically inluenced corrosion of stainless steels by Bacillus subtilis
via bidirectional extracellular electron transfer. Corros. Sci. 207, 110608 (2022).
81. Huang, L. et al. Acceleration of corrosion of 304 stainless steel by outward extracellular
electron transfer of Pseudomonas aeruginosa bioilm. Corros. Sci. 199, 110159 (2022).
82. Zhou, E. et al. Accelerated biocorrosion of stainless steel in marine water via extracellular
electron transfer encoding gene phzH of Pseudomonas aeruginosa. Water Res. 220,
118634 (2022).
83. Dong, Y. et al. Severe microbiologically inluenced corrosion of S32654 super
austenitic stainless steel by acid producing bacterium Acidithiobacillus caldus SM-1.
Bioelctrochemistry 123, 34–44 (2018).
84. Mand, J., Park, H. S., Jack, T. R. & Voordouw, G. The role of acetogens in microbially
inluenced corrosion of steel. Front. Microbiol. 5, 268 (2014).
85. Capao, A., Moreira-Filho, P., Garcia, M., Bitati, S. & Procopio, L. Marine bacterial
community analysis on 316L stainless steel coupons by illumina MiSeq sequencing.
Biotechnol. Lett. 42, 1431–1448 (2020).
86. Salgar-Chaparro, S. J., Lepkova, K., Pojtanabuntoeng, T., Darwin, A. & Machuca, L. L.
Microbiologically inluenced corrosion as a function of environmental conditions:
a laboratory study using oilield multispecies bioilms. Corros. Sci. 169, 108595
(2020).
87. Zhou, E. et al. Methanogenic archaea and sulfate reducing bacteria induce severe
corrosion of steel pipelines after hydrostatic testing. J. Mater. Sci. Technol. 48, 72–83
(2020).
88. Hirano, S., Nagaoka, T. & Matsumoto, N. Microbial community dynamics in a crust formed
on carbon steel SS400 during corrosion. Corros. Eng. Sci. Technol. 55, 685–692 (2020).
89. Huang, Y. et al. Responses of soil microbiome to steel corrosion. npj Bioilms
Microbiomes 7, 6 (2021).
90. Tuck, B., Watkin, E., Somers, A. & Machuca, L. L. A critical review of marine bioilms
on metallic materials. npj Mater. Degrad. 6, 25 (2022).
91. Vigneron, A. et al. Complementary microorganisms in highly corrosive bioilms from
an oshore oil production facility. Appl. Environ. Microbiol. 82, 2545–2554 (2016).
92. Li, X.-X. et al. Responses of microbial community composition to temperature gradient
and carbon steel corrosion in production water of petroleum reservoir. Front. Microbiol.
8, 2379 (2017).
93. Rajala, P., Cheng, D.-Q., Rice, S. A. & Lauro, F. M. Sulfate-dependant microbially induced
corrosion of mild steel in the deep sea: a 10-year microbiome study. Microbiome 10, 4
(2022).
94. Zhang, T., Fang, H. H. P. & Ko, B. C. B. Methanogen population in a marine bioilm
corrosive to mild steel. Appl. Microbiol. Biotechnol. 63, 101–106 (2003).
95. Uchiyama, T., Ito, K., Mori, K., Tsurumaru, H. & Harayama, S. Iron-corroding methanogen
isolated from a crude-oil storage tank. Appl. Environ. Microbiol. 76, 1783–1788 (2010).
Nature Reviews Microbiology
Review article
96. Tan, J. L., Goh, P. C. & Blackwood, D. J. Inluence of H2S-producing chemical species
in culture medium and energy source starvation on carbon steel corrosion caused
by methanogens. Corros. Sci. 119, 102–111 (2017).
97. Xu, D., Li, Y., Song, F. & Gu, T. Laboratory investigation of microbiologically inluenced
corrosion of C1018 carbon steel by nitrate reducing bacterium Bacillus licheniformis.
Corros. Sci. 77, 385–390 (2013).
98. Li, J., Liu, Z., Lou, Y., Du, C. & Li, X. Evidencing the uptake of electrons from X80 steel
by Bacillus licheniformis with redox probe, 5-cyano-2,3-ditolyl tetrazolium chloride.
Corros. Sci. 168, 108569 (2020).
99. Jia, R., Yang, D., Xu, D. & Gu, T. Electron transfer mediators accelerated the
microbiologically inluence corrosion against carbon steel by nitrate reducing
Pseudomonas aeruginosa bioilm. Bioelectrochemistry 118, 38–46 (2017).
100. Lahme, S. et al. Metabolites of an oil ield sulide-oxidizing, nitrate-reducing Sulfurimonas
sp. cause severe corrosion. Appl. Environ. Microbiol. 85, e01891-18 (2019).
101. Vigneron, A., Head, I. M. & Tsesmetzis, N. Damage to oshore production facilities
by corrosive microbial bioilms. Appl. Microbiol. Biotechnol. 102, 2525–2533 (2018).
102. Lyles, C. N., Le, H. M., Beasley, W. H., McInerney, M. J. & Sulita, J. M. Anaerobic
hydrocarbon and fatty acid metabolism by syntrophic bacteria and their impact on
carbon steel corrosion. Front. Microbiol. 5, 114 (2014).
This study provides an example of the use of deined multi-species microbial consortia
to investigate corrosion mechanisms.
103. Batmanghelich, F., Li, L. & Seo, Y. Inluence of multispecies bioilms of Pseudomonas
aeruginosa and Desulfovibrio vulgaris on the corrosion of cast iron. Corros. Sci. 121,
94–104 (2017).
104. Liu, H. & Chen, Y. F. Corrosion of X52 pipeline steel in a simulated soil solution with
coexistence of Desulfovibrio desulfuricans and Pseudomonas aeruginosa bacteria.
Corros. Sci. 173, 108753 (2020).
105. Jia, R., Unsal, T., Xu, D., Lekbach, Y. & Gu, T. Microbiologically inluenced corrosion and
current mitigation strategies: a state of the art review. Int. Biodeterior. Biodegrad. 137,
42–58 (2019).
106. Unsal, T. et al. Foodgrade dlimonene enhanced a green biocide in the mitigation of
carbon steel biocorrosion by a mixedculture bioilm consortium. Bioprocess. Bioeng.
45, 669–678 (2022).
107. Li, Y., Jia, R., Al-Mahamedh, H., Xu, D. & Gu, T. Enhanced biocide mititgtation of ield
bioilm consortia by a mixture of d-amino acids. Front. Microbiol. 7, 896 (2016).
108. Jia, R. et al. A sea anemone-inspired small synthetic peptide at sub-ppm concentrations
enhanced bioilm mitigation. Int. Biodeterior. Biodegrad. 139, 78–85 (2019).
109. Lou, Y. et al. Microbiologically inluenced corrosion inhibition mechanisms in corrosion
protection: a review. Bioelctrochemistry 141, 107883 (2021).
This article presents a comprehensive review of strategies for inhibiting microbial
corrosion.
110. Ornek, D., Wood, T. K., Hsu, C. H., Sun, Z. & Mansfeld, F. Pitting corrosion control of
aluminum 2024 using protective bioilms that secrete corrosion inhibitors. Corrosion
58, 761–767 (2002).
111. Qiu, L. et al. Inhibition eect of Bedellovibrio bacteriovorus on the corrosion of X70
pipeline steel induced by sulfate-reducing bacteria. Anti-Corros. Methods Mater. 63,
260–274 (2016).
112. Scarascia, G. et al. Eect of quorum sensing on the ability of Desulfovibrio vulgaris
to form bioilms and to biocorrode carbon steel in saline conditions. Appl. Environ.
Microbiol. 86, e01664-19 (2020).
113. Li, Z. et al. Marine bioilms with signiicant corrosion inhibition performance by secreting
extracellular polymeric substances. ACS Appl. Mater. Interfaces 13, 47272–47282 (2021).
114. Guo, N. et al. Marine bacteria inhibit corrosion of steel via synergistic biomineralization.
J. Mater. Sci. Technol. 66, 82–90 (2021).
115. Kip, N. et al. Methanogens predominate in natural corrosion protective layers on metal
sheet piles. Sci. Rep. 7, 11899 (2017).
116. In’t Zandt, M. H. et al. High-level abundances of Methanobacteriales and
Syntrophobacterales may help to prevent corrosion of metal sheet piles. Appl. Environ.
Microbiol. 85, e01369-19 (2019).
This work uses ield studies to show that some anaerobes may protect metal surfaces
from corrosion rather than accelerating it.
117. Thorstenson, T. et al. in Conf. Proc. Corrosion 2002 02033 (NACE International, 2002).
118. Falkow, S. Molecular Koch’s postulates applied to microbial pathogenicity. Rev. Infect.
Dis. 10, S274–S276 (1988).
119. Fredricks, D. N. & Relman, D. A. Sequence-based identiication of microbial pathogens:
a reconsideration of Koch’s postulates. Clin. Microbiol. Rev. 9, 18–33 (1996).
120. Mahadevan, R., Palsson, B. O. & Lovley, D. R. In situ to in silico and back: elucidating the
physiology and ecology of Geobacter spp. using genome-scale modelling. Nat. Rev.
Microbiol. 9, 39–50 (2011).
121. Esvap, E. & Ulgen, K. O. Advances in genome-scale metabolic modeling toward
microbial community analysis of the human microbiome. ACS Synth. Biol. 10, 2121–2137
(2021).
122. Borer, B. & Or, D. Spatiotemporal metabolic modeling of bacterial life in complex
habitats. Curr. Opin. Biotechnol. 67, 65–71 (2021).
123. Colarusso, A. V., Goodchild-Michelman, I., Rayle, M. & Zomorrodi, A. R. Computational
modeling of metabolism in microbial communities on a genome-scale. Curr. Opin. Syst. Biol.
26, 46–57 (2021).
124. Philips, J. et al. An Acetobacterium strain isolated with metallic iron as electron donor
enhances iron corrosion by a similar mechanism as Sporomusa sphaeroides. FEMS
Microbiol. Ecol. 95, iy222 (2019).
125. Ali, O. A. et al. Iron corrosion induced by the hyperthermophilic sulfate-reducing
archaeon Archaeoglobus fulgidus at 70°C. Int. Biodeterior. Biodegrad. 154, 105056
(2020).
126. Jia, R., Yang, D., Xu, D. & Gu, T. Carbon steel biocorrosion at 80°C by a thermophilic
sulfate reducing archaeon bioilm provides evidence for its utilization of elemental iron
as electron donor through extracellular electron transfer. Corros. Sci. 145, 47–54 (2018).
This study expands the known diversity of corrosive microorganisms to
hyperthermophiles.
127. Davidova, I. A., Duncan, K. E., Wiley, G. & Najar, F. Z. Desulfoferrobacter sulitae gen.
nov., sp. nov., a novel sulphate-reducing bacterium in the Deltaproteobacteria capable
of autotrophic growth with hydrogen or elemental iron. Int. J. Syst. Evol. Microbiol. 72,
005483 (2022).
128. Magot, M. et al. Dethiosulfovibrio peptidovorans gen. nov., sp. nov., a new anaerobic,
slightly halophilic, thiosulfate-reducing bacterium from corroding oshore oil wells.
Int. J. Syst. Bacteriol. 47, 818–824 (1997).
129. Kato, S., Yumoto, I. & Kamagata, Y. Isolation of acetogenic bacteria that induce
biocorrosion by utilizing metallic iron as the sole electron donor. Appl. Environ.
Microbiol. 81, 67–73 (2014).
130. Lovley, D. R. Syntrophy goes electric: direct interspecies electron transfer. Ann. Rev.
Microbiol. 71, 643–664 (2017).
131. Nevin, K. P., Woodard, T. L., Franks, A. E., Summers, Z. M. & Lovley, D. R. Microbial
electrosynthesis: feeding microbes electricity to convert carbon dioxide and water
to multicarbon extracellular organic compounds. mBio 1, e00103–e00110 (2010).
132. Lovley, D. R. & Nevin, K. P. Electrobiocommodities: powering microbial production
of fuels and commodity chemicals from carbon dioxide with electricity. Curr. Opin.
Biotechnol. 24, 385–390 (2013).
133. Tremblay, P.-L., Angenent, L. T. & Zhang, T. Extracellular electron uptake: among
autotrophs and mediated by surfaces. Trends Biotechnol. 35, 360–371 (2017).
134. Logan, B. E., Rossi, R., Ragab, A. & Saikaly, P. E. Electroactive microorganisms in
bioelectrochemical systems. Nat. Rev. Microbiol. 17, 307–319 (2019).
135. Venzla, H. et al. Accelerated cathodic reaction in microbial corrosion of iron due to
direct electron uptake by sulfate-reducing bacteria. Corros. Sci. 66, 88–96 (2013).
136. Pal, M. K. & Lavanya, M. Microbially inluenced corrosion: understanding bioadlhesion
and bioilm formation. J. Bio. Tribo. Corros. 8, 76 (2022).
137. Sauer, K. et al. The bioilm life cycle: expanding the conceptual model of bioilm
formation. Nat. Rev. Microbiol. 20, 608–620 (2022).
138. Jo, J., Price-Whelan, A. & Dietrich, L. E. P. Gradients and consequences of heterogeneity
in bioilms. Nat. Rev. Microbiol. 20, 593–607 (2022).
139. Usher, K. M., Kaksonen, A. H., Cole, I. & Marney, D. Critical review: microbially inluenced
corrosion of buried carbon steel pipes. Int. Biodeterior. Biodegrad. 93, 84–106 (2014).
140. Traxler, I., Singewald, T. D., Schimo-Aichhorn, G., Hild, S. & Valtiner, M. Scanning
electrochemical microscopy methods (SECM) and ion-selective microelectrodes for
corrosion studies. Corros. Rev. 65, 1213–1224 (2022).
This article overviews the emerging advanced methods for assessing microbial
corrosion.
141. Li, Z. et al. Adaptive bidirectional extracellular electron transfer during accelerated
microbiologically inluenced corrosion of stainless steel. Commun. Mater. 2, 67 (2021).
Acknowledgements
D.X. was inancially supported by the National Key Research and Development Program
of China (No. 2022YFB3808800) and the National Natural Science Foundation of China
(No. U2006219) while working on this Review. The authors apologize to all investigators
whose excellent work could not be cited due to space constraints.
Author contributions
The authors contributed equally to all aspects of the article.
Competing interests
The authors declare no competing interests.
Additional information
Peer review information Nature Reviews Microbiology thanks Daniel John Blackwood,
Muhammad Awais Javed and the other, anonymous, reviewer(s) for their contribution to
the peer review of this work.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional ailiations.
Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
article under a publishing agreement with the author(s) or other rightsholder(s); author
self-archiving of the accepted manuscript version of this article is solely governed by the
terms of such publishing agreement and applicable law.
© Springer Nature Limited 2023
... Metal corrosion phenomena cause significant economic losses, including the deterioration of existing infrastructures (Cámara et al., 2022;Xu et al., 2023). Microorganisms are involved in these phenomena, which are referred to as microbiologically influenced corrosion (MIC). ...
Article
Microbiologically influenced corrosion refers to the corrosion of metal materials caused or promoted by microorganisms. Although some novel iron-corrosive microorganisms have been discovered in various manmade and natural freshwater and seawater environments, microbiologically influenced corrosion in the deep sea has not been investigated in detail. In the present study, we collected slime-like precipitates composed of corrosion products and microbial communities from a geochemical reactor set on an artificial hydrothermal vent for 14.5 months, and conducted culture-dependent and -independent microbial community ana­lyses with corrosive activity measurements. After enrichment cultivation at 37, 50, and 70°C with zero-valent iron particles, some of the microbial consortia showed accelerated iron dissolution, which was approximately 10- to 50-fold higher than that of the abiotic control. In a comparative ana­lysis based on the corrosion acceleration ratio and amplicon sequencing of the 16S rRNA gene, three types of corrosion were estimated: the methanogen-induced type, methanogen–sulfate-reducing bacteria cooperative type, and sulfate-reducing Firmicutes-induced type. The methanogen-induced and methanogen–sulfate-reducing bacteria cooperative types were observed at 50°C, while the sulfate-reducing Firmicutes-induced type was noted at 37°C. The present results suggest the microbial components associated with microbiologically influenced corrosion in deep-sea hydrothermal systems, providing important insights for the development of future deep-sea resources with metal infrastructures.
... It is known that microbiologically influenced corrosion (MIC) is a serious menace to various assets in the oil and gas industry (D. Xu et al., 2023;Yazdi et al., 2021). It has been reported to be responsible for nearly 20% of the total corrosion costs (Kermani and Harrop, 1996;Skovhus et al., 2017). ...
Article
Full-text available
To combat abiotic CO 2 corrosion of pipelines, chromium steels (CrSs) are used to replace carbon steels, but CrSs can suffer very severe pitting corrosion caused by microbiologically influenced corrosion (MIC) because their passive films are not as good as those on high-grade stainless steels, and their MIC often involves (semi-)conductive corrosion product films. In this study, severe pitting corrosion (2.0 cm/a pitting corrosion rate) with a 7-day weight loss of 3.8 ± 0.5 mg/cm ² (0.26 mm/a uniform corrosion rate) was observed on 13Cr coupons incubated anaerobically with a highly corrosive pure-strain sulfate reducing bacterium (SRB) Desulfovibrio ferrophilus IS5 in 125 mL anaerobic vials filled with 50 mL enriched artificial seawater at 28°C. A popular green biocide, namely tetrakis hydroxymethyl phosphonium sulfate (THPS), was enhanced by biofilm dispersing Peptide A (a 14-mer) to mitigate SRB MIC against 13Cr. The 7-day weight losses for coupons with 50 ppm (w/w) THPS, 50 ppm THPS + 100 nM (180 ppb) Peptide A and 100 ppm THPS were reduced to 2.2 ± 0.2 mg/cm ² , 1.5 ± 0.5 mg/cm ² , and 0.3 ± 0.2 mg/cm ² , respectively. The pitting rates also decreased from 20 mm/a to 12 mm/a, 8.6 mm/a, and 1.5 mm/a, respectively based on the maximum pit depth data for the 7-day incubation. Electrochemical tests using a miniature electrochemical glass cell design supported the weight loss trend with additional transient corrosion rate information. THPS was found to be effective in mitigating severe pitting corrosion on 13Cr, and the enhancement effect of Peptide A for THPS was manifested. This work has significant implications in field applications when CrSs are considered as metal choices to replace carbon steels to combat abiotic CO 2 corrosion in pipelines. When SRB MIC is a possible threat, a mitigation plan needs to be implemented to prevent potentially very severe pitting that can lead to pinhole leaks.
Article
This study conducted a comparative proteomic analysis to identify potential genetic markers for the biological function of chemolithoautotrophic iron oxidation in the marine bacterium Ghiorsea bivora . To date, this is the only characterized species in the class Zetaproteobacteria that is not an obligate iron‐oxidizer, providing a unique opportunity to investigate differential protein expression to identify key genes involved in iron‐oxidation at circumneutral pH. Over 1000 proteins were identified under both iron‐ and hydrogen‐oxidizing conditions, with differentially expressed proteins found in both treatments. Notably, a gene cluster upregulated during iron oxidation was identified. This cluster contains genes encoding for cytochromes that share sequence similarity with the known iron‐oxidase, Cyc2. Interestingly, these cytochromes, conserved in both Bacteria and Archaea, do not exhibit the typical β‐barrel structure of Cyc2. This cluster potentially encodes a biological nanowire‐like transmembrane complex containing multiple redox proteins spanning the inner membrane, periplasm, outer membrane, and extracellular space. The upregulation of key genes associated with this complex during iron‐oxidizing conditions was confirmed by quantitative reverse transcription‐PCR. These findings were further supported by electromicrobiological methods, which demonstrated negative current production by G. bivora in a three‐electrode system poised at a cathodic potential. This research provides significant insights into the biological function of chemolithoautotrophic iron oxidation.
Article
Full-text available
Desulfovibrio vulgaris has been a primary pure culture sulfate reducer for developing microbial corrosion concepts. Multiple mechanisms for how it accepts electrons from Fe0 have been proposed. We investigated Fe0 oxidation with a mutant of D. vulgaris in which hydrogenase genes were deleted. The hydrogenase mutant grew as well as the parental strain with lactate as the electron donor, but unlike the parental strain, it was not able to grow on H2. The parental strain reduced sulfate with Fe0 as the sole electron donor, but the hydrogenase mutant did not. H2 accumulated over time in Fe0 cultures of the hydrogenase mutant and sterile controls but not in parental strain cultures. Sulfide stimulated H2 production in uninoculated controls apparently by both reacting with Fe0 to generate H2 and facilitating electron transfer from Fe0 to H+. Parental strain supernatants did not accelerate H2 production from Fe0, ruling out a role for extracellular hydrogenases. Previously proposed electron transfer between Fe0 and D. vulgaris via soluble electron shuttles was not evident. The hydrogenase mutant did not reduce sulfate in the presence of Fe0 and either riboflavin or anthraquinone-2,6-disulfonate, and these potential electron shuttles did not stimulate parental strain sulfate reduction with Fe0 as the electron donor. The results demonstrate that D. vulgaris primarily accepts electrons from Fe0 via H2 as an intermediary electron carrier. These findings clarify the interpretation of previous D. vulgaris corrosion studies and suggest that H2-mediated electron transfer is an important mechanism for iron corrosion under sulfate-reducing conditions. IMPORTANCE Microbial corrosion of iron in the presence of sulfate-reducing microorganisms is economically significant. There is substantial debate over how microbes accelerate iron corrosion. Tools for genetic manipulation have only been developed for a few Fe(III)-reducing and methanogenic microorganisms known to corrode iron and in each case those microbes were found to accept electrons from Fe0 via direct electron transfer. However, iron corrosion is often most intense in the presence of sulfate-reducing microbes. The finding that Desulfovibrio vulgaris relies on H2 to shuttle electrons between Fe0 and cells revives the concept, developed in some of the earliest studies on microbial corrosion, that sulfate reducers consumption of H2 is a major microbial corrosion mechanism. The results further emphasize that direct Fe0-to-microbe electron transfer has yet to be rigorously demonstrated in sulfate-reducing microbes.
Article
Full-text available
Corrosion of iron‐containing metals under sulfate‐reducing conditions is an economically important problem. Microbial strains now known as Desulfovibrio vulgaris served as the model microbes in many of the foundational studies that developed existing models for the corrosion of iron‐containing metals under sulfate‐reducing conditions. Proposed mechanisms for corrosion by D. vulgaris include: (1) H2 consumption to accelerate the oxidation of Fe0 coupled to the reduction of protons to H2; (2) production of sulfide that combines with ferrous iron to form iron sulfide coatings that promote H2 production; (3) moribund cells release hydrogenases that catalyze Fe0 oxidation with the production of H2; (4) direct electron transfer from Fe0 to cells; and (5) flavins serving as an electron shuttle for electron transfer between Fe0 and cells. The demonstrated possibility of conducting transcriptomic and proteomic analysis of cells growing on metal surfaces suggests that similar studies on D. vulgaris corrosion biofilms can aid in identifying proteins that play an important role in corrosion. Tools for making targeted gene deletions in D. vulgaris are available for functional genetic studies. These approaches, coupled with instrumentation for the detection of low concentrations of H2, and proven techniques for evaluating putative electron shuttle function, are expected to make it possible to determine which of the proposed mechanisms for D. vulgaris corrosion are most important.
Article
Full-text available
The sulfate-reducing microbe Desulfovibrio ferrophilus is of interest due to its relatively rare ability to also grow with Fe(III) oxide as an electron acceptor and its rapid corrosion of metallic iron. Previous studies have suggested multiple agents for D. ferrophilus extracellular electron exchange including a soluble electron shuttle, electrically conductive pili, and outer surface multiheme c-type cytochromes. However, the previous lack of a strategy for genetic manipulation of D. ferrophilus limited mechanistic investigations. We developed an electroporation-mediated transformation method that enabled replacement of D. ferrophilus genes of interest with an antibiotic resistance gene via double-crossover homologous recombination. Genes were identified that are essential for flagellum-based motility and the expression of the two types of D. ferrophilus pili. Disrupting flagellum-based motility or expression of either of the two pili did not inhibit Fe(III) oxide reduction, nor did deleting genes for multiheme c-type cytochromes predicted to be associated with the outer membrane. Although redundancies in cytochrome or pilus function might explain some of these phenotypes, overall, the results are consistent with D. ferrophilus primarily reducing Fe(III) oxide via an electron shuttle. The finding that D. ferrophilus is genetically tractable not only will aid in elucidating further details of its mechanisms for Fe(III) oxide reduction but also provides a new experimental approach for developing a better understanding of some of its other unique features, such as the ability to corrode metallic iron at high rates and accept electrons from negatively poised electrodes. IMPORTANCE Desulfovibrio ferrophilus is an important pure culture model for Fe(III) oxide reduction and the corrosion of iron-containing metals in anaerobic marine environments. This study demonstrates that D. ferrophilus is genetically tractable, an important advance for elucidating the mechanisms by which it interacts with extracellular electron acceptors and donors. The results demonstrate that there is not one specific outer surface multiheme D. ferrophilus c-type cytochrome that is essential for Fe(III) oxide reduction. This finding, coupled with the lack of apparent porin-cytochrome conduits encoded in the D. ferrophilus genome and the finding that deleting genes for pilus and flagellum expression did not inhibit Fe(III) oxide reduction, suggests that D. ferrophilus has adopted strategies for extracellular electron exchange that are different from those of intensively studied electroactive microbes like Shewanella and Geobacter species. Thus, the ability to genetically manipulate D. ferrophilus is likely to lead to new mechanistic concepts in electromicrobiology.
Article
Full-text available
A mesophilic sulphate-reducing micro-organism, able to grow chemolithoautotrophically with H 2 /CO 2 (20 : 80) and with elemental iron as a sole electron donor, was isolated from a consortium capable of degrading long-chain paraffins and designated strain DRH4 T . Cells were oval shaped often with bright refractile cores and occurred singly or in pairs. The cells formed pili. Strain DRH4 T could grow chemolithoautotrophically with H 2 /CO 2 or elemental iron and chemoorganotrophically utilizing a number of organic substrates, such as fatty acids from formate to octanoate (C 1 –C 8 ). Sulphate and thiosulphate served as terminal electron acceptors, but sulphite and nitrate did not. Optimal growth was observed from 37 to 40 °C and pH from 6.5 to 7.2. Strain DRH4 T did not require NaCl for growth and could proliferate under a broad range of salinities from freshwater (1 g l ⁻¹ NaCl) to seawater (27 g l ⁻¹ NaCl) conditions. The genomic DNA G+C content was 54.46 mol %. Based on 16S rRNA gene sequence analysis. strain DRH4 T was distinct from previously described Deltaproteobacteria species exhibiting the closest affiliation to Desulforhabdus amnigena ASRB1 T , Syntrophobacterium sulfatireducens TB8106 T and Desulfovirga adipica 12016 T with 93.35, 93.42 and 92.85 % similarity, respectively. Strain DRH4 T showed significant physiological differences with the aforementioned organisms. Based on physiological differences and phylogenetic comparisons, we propose to classify DRH4 T as the type strain (=DSM 113 455 T =JCM 39 248 T ) of a novel species of a new genus with the name Desulfoferrobacter suflitae gen. nov., sp. nov.
Article
Full-text available
Over the last 30 years, scanning electrochemical microscopy (SECM) has become a fundamental technique in corrosion research. With its high spatial resolution and its ability to study local electrochemistry, it contributes essentially to the understanding of corrosion processes. By using selective micro- and nano-sensors, concentration profiles of different corrosion relevant species, from protons to metal ions, can be established. This review provides a comprehensive overview about SECM based techniques and discusses various types of microsensors, including materials selection and preparation techniques, and it provides extensive tables on redox-couples for specific corrosion research applications.
Article
The corrosion of 304 stainless steel by Bacillus subtilis was investigated under different glycerol (electron donor) and nitrate (electron acceptor) concentrations. When nitrate was deficient, the passive film served as an alternative electron acceptor to complete outward extracellular electron transfer (EET) of B. subtilis, inducing more severe corrosion especially in the early stage. When glycerol concentration was deficient, the inward EET of B. subtilis was significantly enhanced, resulting in accelerated corrosion in the late stage. These results demonstrated that bidirectional EET was involved in the corrosion of stainless steel by B. subtilis and depended on the electron donor/acceptor concentrations.
Article
To predict variation of maximum localized penetration with exposure time, long-term localized corrosion was assessed in an emergency cooling water system composed of two carbon steel pipelines of 700 mm diameter transporting raw river water at flow velocities of 1 m/s and 0.1 m/s. Field tests, visual inspection, ultrasonic testing, BART testing, SEM-EDS and metagenomic analyses were performed to assess the progress of long-term corrosion and determine the influence of microbes in the corrosion process. High corrosion was linked to sulphate reducing bacteria and potentially to methanogenic archaea in the low-velocity pipeline, while moderate corrosion was linked to non-sulphate reducing bacteria in the higher velocity pipeline. Using historical and literature data available as well as our own test results, an empirical model was developed to predict Maximum Localized Penetration change over time to be applied in the ageing management of cooling water systems. Molecular Microbiological Methods in combination with traditional techniques are useful tools in the ageing management of pipelines. By applying the empirical model developed and the approach presented, unexpected through-wall leaking can be avoided, thus, saving costs and assets.
Article
Bacterial biofilms are often defined as communities of surface-attached bacteria and are typically depicted with a classic mushroom-shaped structure characteristic of Pseudomonas aeruginosa. However, it has become evident that this is not how all biofilms develop, especially in vivo, in clinical and industrial settings, and in the environment, where biofilms often are observed as non-surface-attached aggregates. In this Review, we describe the origin of the current five-step biofilm development model and why it fails to capture many aspects of bacterial biofilm physiology. We aim to present a simplistic developmental model for biofilm formation that is flexible enough to include all the diverse scenarios and microenvironments where biofilms are formed. With this new expanded, inclusive model, we hereby introduce a common platform for developing an understanding of biofilms and anti-biofilm strategies that can be tailored to the microenvironment under investigation. In this Review, Bjarnsholt and colleagues propose a revised conceptual model of the biofilm life cycle that encompasses the three major steps of biofilm formation — aggregation, growth and disaggregation — independently of surfaces, and initiation from single-cell planktonic bacteria, and thus represents a broader range of biofilm systems.
Article
Despite increasing interest in corrosion caused by direct electron transfer (DET) from steel surfaces to microbial cells, the validity of this mechanism is debated. This is often because of the inability to differentiate between the extent of corrosion due to DET and the microbial consumption of H2. Shewanella oneidensis is linked to corrosion of steel and capable of engaging in extracellular electron transfer and consuming H2. However, it is not clear how S. oneidensis corrodes steel. We tested the ability of S. oneidensis to corrode carbon steel through a DET mechanism independently from its H2-consuming capabilities by performing anaerobic corrosion experiments of carbon steel with an S. oneidensis strain incapable of consuming H2 (ΔhydAΔhyaB) and a strain unable to engage in DET (ΔMtr). We found that S. oneidensis accelerates the steel corrosion rate by up to four times relative to an abiotic control and corrosion is coupled to the reduction of fumarate to succinate. More localized corrosion and higher cell density on the steel surface is observed when cells are restricted to DET alone than when only H2 consumption is operational. Coupon weight loss and Linear Polarization Resistance measurements show that the wild-type strain preferentially used the DET mechanism to corrode steel under anoxic conditions. This study illustrates the use of mutant strains as an experimental approach to distinguish DET from H2-mediated electron transfer.