Skip to content
Publicly Available Published by De Gruyter October 7, 2015

Omics-based approaches and their use in the assessment of microbial-influenced corrosion of metals

  • David J. Beale

    David J. Beale is a research scientist in the Land and Water Flagship at the Commonwealth Scientific and Industrial Research Organisation (CSIRO). He has more than 10 years of experience in the delivery of R&D projects, which include a portfolio of projects relating to sustainability, water management, and quality within the water and wastewater sector. He is also a technical expert and project leader on developing environmental metabolomic techniques for the assessment of pathogens and biofilms within water systems. He holds a bachelor’s degree with first class honours in environmental science and doctorate in analytical chemistry from RMIT University.

    EMAIL logo
    , Avinash V. Karpe

    Avinash V. Karpe has a PhD from the Faculty of Science, Engineering, and Technology, at Swinburne University of Technology and is currently a visiting scientist with CSIRO Land and Water. His primary research involves enhancing fungal bioprocessing for biofuel and medicinal metabolite production. Additionally, he is also involved in numerous metabolomic studies of fungal and bacterial communities in riverine, wastewater, and food processing systems.

    , Snehal Jadhav

    Snehal Jadhav is a postdoctoral research fellow in Swinburne University of Technology, Melbourne, working in the area of microbial proteomics and metabolomics. Currently, her research is centred on the development of strategies based on MALDI-TOF MS, GC-MS, and LC-MS to identify and characterise bacteria obtained from food, clinical, and environmental sources. Previously, her research has also focussed on the effect of natural products against bacterial biofilms formed on abiotic surfaces.

    , Tim H. Muster

    Tim H. Muster is a senior research scientist in the Cities Program in CSIRO Land & Water. He has over 20 years of research experience in the scientific disciplines of colloid, surface, and electrochemistry, with over 65 refereed journal publications. In 2007, Tim was the recipient of CSIRO Young Scientist John Philip Award and has twice won the Marshall Fordham Best Research Paper of the Australasian Corrosion Association (2003 and 2005). More recently, Tim was the recipient of a CSIRO Julius Career Award for nutrient recovery from wastewater and leads research focussed on the effective management of urban and food production waste streams for the productive recovery of water, energy, and nutrients.

    and Enzo A. Palombo

    Enzo A. Palombo is chair at the Department of Chemistry and Biotechnology at Swinburne University of Technology in Australia. He has over 25 years of experience as a microbiologist and combines his academic teaching of microbiology and environmental biology with research interests in environmental microbiology, food microbiology, diagnostic microbiology, gastrointestinal microbiota, and bioactive compound discovery.

From the journal Corrosion Reviews

Abstract

Microbial-influenced corrosion (MIC) has been known to have economic, environmental, and social implications to offshore oil and gas pipelines, concrete structures, and piped water assets. While corrosion itself is a relatively simple process, the localised manner of corrosion makes in situ assessments difficult. Furthermore, corrosion assessments tend to be measured as part of a forensic investigation. Compounding the issue further is the impact of microbiological/biofilm processes, where corrosion is influenced by the complex processes of different microorganisms performing different electrochemical reactions and secreting proteins and metabolites that can have secondary effects. While traditional microbiological culture-dependent techniques and electrochemical/physical assessments provide some insight into corrosion activity, the identity and role of microbial communities that are related to corrosion and corrosion inhibition in different materials and in different environments are scarce. One avenue to explore MIC and MIC inhibition is through the application of omics-based techniques, where insight into the bacterial population in terms of diversification and their metabolism can be further understood. As such, this paper discusses the recent progresses made in a number of fields that have used omics-based applications to improve the fundamental understanding of biofilms and MIC processes.

1 Introduction

In recent decades, there has been a considerable amount of research published on the role of microorganisms in promoting corrosion. The majority of this work has been undertaken to address the problem of microbial-influenced corrosion (MIC) in offshore oil and gas pipelines and concrete structures, with some preliminary research on microbial/metal surface interactions in water pipes. As such, a number of extensive reviews have been compiled on MIC mechanisms over the past 20 years (Beech & Gaylarde, 1999; Edyvean & Videla, 1991; Flemming, 1994; Videla, 2003; Videla & Herrera, 2009) and, most recently, the review by Beech, Sztyler, Gaylarde, Smith, and Sunner (2014), to name a few.

It has been highlighted in these reviews that there is a growing number of researchers applying omics-based techniques for the characterisation and understanding of MIC, where omics-based techniques comprise: (meta)genomics (e.g. the analysis of genetic material recovered from an organism or environmental samples), transcriptomics (e.g. the analysis of ribonucleic acid (RNA) molecules, including messenger RNA (mRNA), ribosomal ribonucleic acid (rRNA), transfer RNA, and other noncoding RNA, produced by an organism or a population), proteomics (e.g. the analysis of proteins produced by an organism or population, and their function), and metabolomics (e.g. the analysis of the small chemical compounds produced and consumed by an organism or a population), or a combination of approaches. The application of these techniques has been demonstrated to be beneficial in MIC research due to the presence of causative biofilms that are heterogeneous in nature and involve multiple organisms comprising complex and variable structures; thus, their interaction with the metal surface to which they have adhered will be similarly complex and likely to vary within the biofilm. As such, the application of omics-based techniques, specifically metagenomic techniques (also known as shotgun genomics), enables researchers to identify and characterise the entire microbial population present and, in combination with complementary techniques, identify what that population is doing in terms of gene expression (i.e. transcriptomics), protein production (i.e. proteomics), and metabolism (i.e. metabolomics) (Figure 1).

Figure 1: Application of omics-based techniques to understand and characterise biofilms and MIC.
Figure 1:

Application of omics-based techniques to understand and characterise biofilms and MIC.

2 Microbial-influenced corrosion (MIC)

Corrosion itself is a relatively simple electrochemical process that involves two matched electrochemical processes – the oxidation of the metal at the anode and a parallel chemical reduction process at the cathode. The electrochemical circuit is completed by a transfer of electrons from the anode to the cathode. Corrosion may be accelerated or retarded by encouraging or restricting either the anodic or cathodic reactions.

However, in understanding corrosion, a difficulty arises in that most corrosion takes place in a localised manner, meaning that the separation of anodes and cathodes is on a submillimetre scale, and secondly, unlike galvanic corrosion, electron transfer occurs within the conducting metal and cannot be measured. The only measurements that can be made are through physical evaluation of metal loss, oxide accumulation, or localised measurement of electrolyte composition and gradients (i.e. pH, solution potential, cation/anion concentration). Hence, where biological activity interacts with the localised electrolyte near the metal interface, it becomes increasingly difficult to isolate the (i) causal factors leading to electrochemical corrosion, (ii) casual factors leading to biofilm growth, (iii) the organic and inorganic films resulting from localised corrosion, and (iv) the chemical and localised potential gradients resulting from biofilm growth (see Figure 2).

Figure 2: What comes first, the “chicken or the egg?” Causal and resultant factors relating to “electrochemical” corrosion and microbiological changes.
Figure 2:

What comes first, the “chicken or the egg?” Causal and resultant factors relating to “electrochemical” corrosion and microbiological changes.

As a result of the overlapping consumption and production of chemical species, gaining a precise understanding of the mechanisms of MIC requires more sophisticated methods for being able to isolate microbial processes from physical electrochemical processes, hence the suitability of omics for such studies, as will be stressed throughout this paper.

Biofilm formation has been demonstrated to develop within relatively short periods of time through extracellular polymeric substance (EPS) excreted by the populating microbes (Characklis, 1981). Keevil (2004) described biofilms as a complex structure that consists of arrays of microcolonies encased in EPS but with significant channels between these stacks of microcolonies that facilitate diffusion. They are heavily hydrated gelatinous-like films that have been demonstrated to physically affect corrosion processes (Videla & Herrera, 2009) and, if heterogeneously distributed over the surface, can result in the establishment of permanent anodic and cathodic regions across the metal substrate. However, biofilms are highly reactive to the environment, and survival of the biofilm requires nutrient transfer into the interior, and thus, the film cannot behave as a complete barrier. This open structure means that biofilms can also incorporate oxide-based corrosion products and may also contain specific zones where the diffusion of particular chemical species is restricted. For instance, activity by one group of microbes may hinder or foster the development of others (Zuo, 2007). Thus, there can be a competition between sulphate-reducing bacteria (SRB) and iron-reducing bacteria (IRB) so that IRB may decrease SRB activity (Potekhina et al., 1999). In contrast, acid-producing bacteria, as well as being corrosive in themselves, may provide nutrients and environmental conditions conducive for SRB growth (Jack, Rogoz, Bramhill, & Roberge, 1994).

The major impact of biofilms and microbes on corrosion occurs through changing the rates of either anodic or cathodic reactions. Before moving on to detail the impacts and potential impact of omics, it is worth reviewing some of the current understandings of MIC and how biofilms and microbes can influence anodic and cathodic reaction rates.

2.1 Diffusion

If the reactants and/or products of anodic or cathodic processes cannot move freely to and from the metal surface, corrosion rates can be either diffusion limited or concentration limited. Videla and Herrera (2009) emphasised how the respiratory processes of microorganisms may modify the oxygen concentration at the metal interface. If the biofilm covers the whole surface, this may lead to a reduction in the cathodic reaction rate and decreased corrosion or, alternatively, if the coating is patchy, is likely to create a potential gradient that encourages localised corrosion. Edyvean and Videla (1991) illustrated the latter scenario where biofilms were shown to accelerate corrosion by forming oxygen cells in association with tubercles in potable water pipework. In addition, it is reported that aside from oxygen gradients, chemical and pH gradients can also be created within biofilms and can drive corrosion processes. For instance, pH differences within microalgal/bacteria films have been demonstrated to exceed nine units (Terry & Edyvean, 1986).

2.2 Polarisation past the pitting or breakdown potential

Where either the anodic or cathodic reaction rates are altered consistently over time, the potential of the underlying metal may be polarised. Polarisation can have a large impact on metals that are prone to pitting and rely upon oxide passivation to avoid corrosion. For instance, Dexter, Duquette, Siebert, and Videla (1991) suggested that the presence of a biofilm was able to shift the corrosion potential of from the passive zone to the pitting region.

2.3 Cathodic reaction rates

The cathodic reaction may be affected in a number of ways, as already mentioned; the oxygen concentration may be increased or decreased both by direct consumption (Zuo, 2007) or production of oxygen or due to the formation of barriers (by bacteria) to oxygen diffusion. Additionally, the production of acidic metabolites can increase cathodic processes through both the dissolution of protective films and through the supply of oxidising agents (i.e. protons) that encourage cathodic processes (Edyvean & Videla, 1991). Alternatively, Little, Ray, and Pope (2000) proposed that SRB may enhance cathodic reaction rates by removing atomic hydrogen from the cathode and thus avoid concentration polarisation limitations. Herrera and Videla (2009) found that the preexposure of steel to Vibrio alginolytics had a profound effect on the free corrosion potential, shifting the free corrosion potential to a more negative value. This suggests a decreased cathodic reaction rate and enabled the creation of an anaerobic environment that allowed SRB growth and an even more significant negative shift in potential.

If metal oxides are conductive, they may support cathodic reactions, or, in the case of iron in anaerobic conditions, the redox reaction of Fe(III) to Fe(II) may act as the cathodic reaction (Stratmann, 1990). In environments with a high sulphur content or containing SRB, a protective layer of FeS may form but is, like most inorganic protective layers, highly dependent upon pH (Shoesmith, Taylor, Bailey, & Owen, 1980). Other reports have suggested that FeS can act as a cathode and support the anodic dissolution of iron that is not protected by corrosion products (Edyvean & Videla, 1991). For the case of Fe(III) to Fe(II), microbes such as Shewanella oneidensis that enhance this process can have two contradictory effects: it could reduce the overall cathodic current by decreasing the iron reduction pathway, or it may disrupt the passive layer formation of Fe(III) oxides and therefore increase corrosion (Herrera & Videla, 2009).

2.4 Anodic reaction rates

Microbes may affect the rate of the anodic reaction in a number of ways. Acid metabolites can dramatically lower the local pH. Terry and Edyvean (1981) indicated that the pH differences can be up to nine units. Furthermore, Edyvean and Videla (1991) demonstrated that light/dark cycles in photosynthesis and respiration can produce daily cycles of up to 3.5 units. Sulphur-oxidising bacteria (SOB) (genus Thiobacillus) can produce up to 10% sulphuric acid with pH <2 (Edyvean & Videla, 1991). Nitrogen-utilising bacteria, microfungi, and microalgae can also produce acids (Edyvean & Videla, 1991). Such local acid regions can dissolve any protective films and accelerate the anodic reaction.

The slime produced by iron reducers may absorb iron, and such absorption can have diverse effects depending on the state of the oxide/biofilm. It could limit oxide formation, which would promote corrosion. On the other hand, it could also limit the acidification caused by the hydration of metal ions, which can lead to local acidification and pitting.

2.5 The role for omics

Corrosion is the result of a series of chemical, physical, and (micro)biological processes leading to the deterioration of materials. The mechanisms of MIC and MIC inhibition are not completely understood because they cannot be linked to a single biochemical reaction or specific microbial species or group (Kip & van Veen, 2015). Corrosion is influenced by the complex processes of different microorganisms performing different electrochemical reactions and secreting proteins and metabolites that can have secondary effects. Information on the identity and role of microbial communities that are related to corrosion and corrosion inhibition in different materials and in different environments is scarce. As some microorganisms are able to both cause and inhibit corrosion, we pay particular interest to their potential role as corrosion-controlling agents. One avenue to explore MIC and MIC inhibition is through the application of omics, where insight into the bacterial population in terms of diversification and their metabolism can be understood. Table 1 provides a summary of omics-based techniques applied to the investigation of MIC, proceeded by a discussion of their use and benefits. It should be noted, it is not the intention of the authors to provide a review of the different analytical platforms applied in omics-based research. The focus of the review presented is to highlight the application and benefits of omics-based techniques in the assessment of MIC of metals.

Table 1

Summary of omics-based techniques applied to the investigation of MIC.

Industry/ContextMaterialOmics techniqueSummaryReferences
WastewaterConcrete sewerMetagenomics

Transciptomics
Taxonomic and functional characterisation.

Abundance of SRB and SOB bacteria. Enrichment of genes associated with metal resistance.
Gomez-Alvarez, Revetta, and Domingo (2012)
MetagenomicsCharacterisation of diverse population comprising archaea, fungi, and several bacterial groups.Vincke, Boon, and Verstraete (2001)
Electricity transmission towerSteelMetagenomics (and culture dependent techniques)Culture-dependent techniques resulted in 31 isolates.

Metagenomic technique resulted in 160 clone sequences comprising filamentous fungi, yeast and bacteria.
Sette et al. (2010)
Marine environmentsCarbon steelMetagenomicsCharacterisation of diverse population comprising bacteria and archaea.

Dominant group was methanogenic archaea.
Usher, Kaksonen, and MacLeod (2014)
Oil production facilityCarbon SteelMetagenomics

Metabolomics
Characterisation of diverse population comprising thermophilic sulfidogenic anaerobes and mesophiles.

Hydrocarbon metabolites detected.
Lenhart et al. (2014)
Extra heavy crude oilCarbon steelMetagenomicsMolecular identification, characterization.

Phylogenesis of novel ascomycetes (Wickerhamia spp., Cladosporium spp. BM-103) and basidiomycetes (Rhodotorula spp.)

Degradation of crude oil resulting in to further degradation of carrier pipelines and storage tanks.
Naranjo et al. (2015)
Food industryAD0-aluminium alloysIonomicsFungal mediated corrosion of AD0 alloys within 60-day incubation.

Alternaria alternata fungus caused maximum corrosion, mediated by several exometabolites.
Smirnov, Belov, Sokolova, Kuzina, and Kartashov (2008)
WoodChromated-copper arsenate (CCA) preservative-treated woodMetabolomicsDisposal of CCA-treated wooden poles due to toxic preservatives such as hexavalent chromium.

Brown rot fungi (Antrodia vaillantii, Poria placenta) caused an oxalic acid-mediated leaching of chromium and copper up to 52.4% and 15.6%, respectively.
Sierra-Alvarez (2007)
Aerospace industryAA2024-T3 aluminium alloyIonomicsConsortium of A. niger and bacteria caused heavy leaching of Cr6+ preservative coating on aluminium alloy to form nontoxic, elemental Cr.Lee, Little, Ray, and Stropki (2012)
VariousCarbon steel/aluminiumMetabolomicsFungal-mediated degradation of metals.

Mitosporic fungi such as Chrysosporium merdarium, Penicillium cyclopium, Arthrinium phaeospermum, Cladosporium herbarum, and A. niger caused mild to severe deterioration in carbon steel and aluminium alloys in 30-day incubation period. The likely mediators were indicated as organic acids, indole alkaloids, trichothecenes, amides, etc.
Lugauskas et al. (2009)
Cooling water system of nuclear reactorCarbon steelMetagenomicsCharacterisation of SIR such as Desulfovibrio, Desulfomicrobium, and Desulfosarcina.Balamurugan, Hiren Joshi, and Rao (2011)
Household water pipesCopperMetagenomicsCharacterisation of diverse population comprising bacteria.

Mainly Alpha, Beta, and Gammaproteobacter species were isolated, including Acinetobacter, Sphingomonas, Chryseobacterium, and Methylobacterium species.
Pavissich, Vargas, González, Pastén, and Pizarro (2010)
Energy emission towersCarbon steel/zincMetagenomicsBacterial and fungal communities were identified.

Mainly Acinetobacter and Stenotrophomonas bacterial species and fungal genera such as Capnobotryella Coniosporium and Fellomyces were found.
Oliveira et al. (2011)

3 Application of omics-based techniques to characterise MIC and biofilms

Metal corrosion is one of the main causes of damage to petroleum pipeline systems, resulting in great economic losses to the industry. Around 40% of the internal pipeline corrosion in the gas industry has been attributed to MIC (Chakraborty, DasGupta, and Bhadury, 2014). Through the use of metagenomics and transcriptomics, MIC biofilm communities can be studied at both their compositional and functional levels. A number of traditional techniques, such as clone libraries and genetic fingerprinting, along with more recent metagenomics and transcriptomics, are being used to characterise and understand MIC biofilms (Chakraborty et al., 2014). Genetic fingerprinting techniques share a common trait of generating profiles of microbial communities based on direct analysis of polymerase chain reaction products amplified from environmental DNA (Muyzer, 1999). Over the years, microbial ecologists have used a number of techniques that produce community fingerprints based on either sequence polymorphisms or length polymorphisms. The community profiles from different samples can be compared using computer-assisted cluster analysis (i.e. chemometrics). In general, genetic fingerprinting techniques are rapid and allow simultaneous analyses of multiple samples. Since fingerprinting approaches have been devised to demonstrate effects on or differences between microbial communities and do not necessarily provide direct taxonomic identities, these approaches are specifically significant in the context of assessing the impacts of abiotic and biotic factors on microbial communities in contaminated sites (Chakraborty et al., 2014).

4 MIC biofilm genome and gene expression

Metagenomics was developed in the first decade of the 21st century and laid the foundation of all the omics-based techniques. Metagenomics revolutionised microbial ecology research, where the collective microbial genome collected from a sample is analysed and reconfigured to provide knowledge of the community’s diversification (Handelsman, 2004). Transcriptomic tools are used to gain functional insights into the activities of microbial communities by studying their mRNA transcriptional profiles (Pascault et al., 2015).

Advances in high-throughput next-generation sequencing (NGS) (Jiang et al., 2008) technology for direct sequencing of environmental DNA (i.e. shotgun metagenomics) are transforming the field of microbiology (Ahmed et al., accepted for publication; Gomez-Alvarez, 2014). NGS technologies are now regularly being applied in comparative metagenomic and transcriptomic studies, which provide the data for functional annotations, taxonomic comparisons, community profile, and metabolic reconstructions.

Gomez-Alvarez et al. (2012) analysed the whole metagenome to determine microbial composition and functional genes associated with biomass harvested from sections of a corroded wastewater pipe. Taxonomic and functional analysis demonstrated that approximately 90% of the total diversity was associated with the phyla Actinobacteria, Bacteroidetes, Firmicutes, and Proteobacteria. Furthermore, the biofilm was found to have an abundance of SOB and SRB. Gomez-Alvarez et al. also demonstrated, combined with transcriptomics, an enrichment of genes associated with heavy metal resistance, virulence (protein secretion systems), and stress response in the biofilm analysed.

Vincke et al. (2001) investigated the genetic fingerprint of the microbiota on corroded concrete sewer pipes using denaturing gradient gel electrophoresis (DGGE) of 16S rRNA gene fragments. The DGGE profiles of the bacterial communities present on the concrete surface changed as observed by shifts occurring at the level of the dominance of bands from noncorroded places to the most severely corroded places. By means of statistical tools, it was possible to distinguish two different groups, corresponding to the microbial communities on corroded and noncorroded surfaces, respectively. Characterisation of the microbial communities indicated that the sequences of typical bands showed the highest level of identity to sequences from the bacterial strains Thiobacillus thiooxidans, Acidithiobacillus spp., Mycobacterium spp., and different heterotrophs belonging to the α-, β- and γ-Proteobacteria, Acidobacteria, and Actinobacteria (Vincke et al., 2001). In addition, Santo Domingo et al. (2011) investigated the microbial composition of concrete biofilms within wastewater collection systems using metagenomics molecular assays. They found that α-, β-, γ-, and δ-Proteobacteria represented 15%, 22%, 11%, and 4% of the clones, respectively, in the 2457 sequences analysed. Bacteria implicated in concrete corrosion were found in the clone libraries while archaea, fungi, and several bacterial groups were also detected using group-specific assays, which highlights that concrete sewer biofilms are more diverse than previously reported (Santo Domingo et al., 2011). However, this diversity was not the case for corrosive biofilms in oil pipelines analysed with DGGE (Neria-González, Wang, Ramírez, Romero, and Hernández-Rodríguez, 2006).

Sette et al. (2010) profiled the fungal community structure found in Brazilian energy transmission towers that displayed signs of corrosion and/or biofilm formation using both culture-dependent and independent methods. In total, 31 isolates comprising 10 filamentous fungi and 21 yeasts were recovered from enrichment cultures, comprising Aspergillus fumigatus, A. niger, Candida albicans, C. pseudointermedia, C. tropicalis, Cryptococcus laurentii, Debaryomyces nepalensis, Exophiala dermatitidis, Fusarium spp., F. solani, Paecilomyces lilacinus, Trichoderma citrinoviride, T. longibrachiatum, and Pichia guilliermondii. Metagenomic analyses based on 160 clone sequences revealed 30 operational taxonomic units (OTUs) comprising 20 OTUs of filamentous fungi and 10 OTUs of yeasts. The majority of OTUs were related to the genera Capnobotryella, Cryptococcus, Devriesia, Fellomyces, Fusarium, Kockovaella, Panaeolus, Rhodotorula, Sirobasidium, Sporobolomyces, Strelitziana, and Teratosphaeria. While the fungi identified are ubiquitous fungi commonly found in other environments, some have been related to MIC of metals. Furthermore, Sette et al. (2010) emphasised the need for complementary approaches to assess the microbial assemblage of unusual environments, stating that not one single approach was sufficient to characterise a community effectively.

Usher et al. (2014) investigated marine rust on carbon steel in order to determine the importance of various microorganisms in corrosion. Microorganisms were imaged (Figure 3), identified, and enumerated by pyrosequencing. The analysis demonstrated the presence of a diverse population of bacteria and archaea. However, the dominant group was methanogenic archaea, representing 53.5% of all sequences. One methanogenic species, Methanococcus maripaludis, comprised 31% of sequences and can significantly increase corrosion rates by extracting electrons directly from steel (Usher et al., 2014). Furthermore, Mohan, Bibby, Lipus, Hammack, and Gregory (2014) analysed hydraulic fracturing source water and wastewater produced during fracking using metagenomic and metabolomic techniques. While not related to corrosion, the metabolic profile revealed a relative increase in genes responsible for carbohydrate metabolism, respiration, sporulation and dormancy, iron acquisition and metabolism, stress response, and sulphur metabolism in the produced water samples from a diverse microbial population (Mohan et al., 2014). These results suggest that microbial communities in potable water have an increased genetic ability to handle stress, which has significant implications for infrastructure management, such as biofilm control and combating MIC.

Figure 3: Scanning electron microscopy (SEM) images of microorganisms in marine rust tubercles sampled from carbon steel. After Usher et al. (2014). Note: the white arrows highlight microorganisms in the marine rust tubercles sampled from carbon steel. Extracellular polymeric substances (EPSs) are indicated with a red arrow. Panel A is a micrograph of carbon steel pipe (control), while panels B and C are micrographs of marine tubercules.
Figure 3:

Scanning electron microscopy (SEM) images of microorganisms in marine rust tubercles sampled from carbon steel. After Usher et al. (2014). Note: the white arrows highlight microorganisms in the marine rust tubercles sampled from carbon steel. Extracellular polymeric substances (EPSs) are indicated with a red arrow. Panel A is a micrograph of carbon steel pipe (control), while panels B and C are micrographs of marine tubercules.

With sequencing costs decreasing, NGS is enabling an increasing number of laboratories to taxonomically (and functionally) classify a wide range of the organisms that are present, and the extension of these techniques to assess microbiological communities that influence corrosions is a logical step (Douterelo et al., 2014).

5 Proteomics and the emergence of matrix-assisted laser desorption/ionisation-time of flight

Proteomics is a discipline focused on the identification of proteins, and metaproteomics is defined as the characterization of the entire protein complement of a microbial community (Douterelo et al., 2014). Proteomics has been successfully used to investigate microbial community functions in marine environments, freshwater ecosystems, and biofilms from an acid mine drainage (Douterelo et al., 2014). Wikieł, Datsenko, Vera, and Sand (2014) studied the metabolic activity and biofilm development of Desulfovibrio alaskensis correlated to electrochemical response of carbon steel surfaces. While carbon steel was protected from the microbial attachment and exposed to metabolic products, only one spike in corrosion potential was recorded and coincided with increased sulphide concentrations during biofilm formation and ultimately resulted in corrosion rates that did not exceed 0.05 mm/year. Furthermore, 150 proteins were detected in the EPS matrix responsible for biofilm formation and maturation (Wikieł et al., 2014).

As previously stated, MIC is greatly influenced by the resident microbial community at any one site. Sites experiencing MIC are generally colonised by multispecies biofilms rather than a single species of bacteria (Li, Whitfield, and Van Vliet Krystyn, 2013). Thus, investigation of the microbes involved in the corrosion process becomes a crucial parameter for adopting a suitable control strategy. Identification of microbes responsible for MIC is generally performed using conventional phenotypic methods that utilise culture-specific growth media and biochemical tests or by molecular methods such as 16S rRNA gene sequencing and the more recent metagenomics (Gomez-Alvarez, 2014; Gomez-Alvarez et al., 2012) and metabolomic approaches (Beale, Dunn, & Marney, 2010; Beale, Morrison, Key, & Palombo, 2014). Phenotypic methods are not always considered reliable and often require interpretation by skilled personnel. On the other hand, molecular methods of identification are much more reliable but can be expensive, time consuming, and difficult to interpret and not readily incorporated into routine laboratory microbiological analysis (Jadhav, Sevior, Bhave, & Palombo, 2014; Van Veen, Claas, & Kuijper, 2010).

A recent proteomics-based microbial identification approach that has gained increased popularity for its rapidity and cost-effectiveness is matrix-assisted laser desorption/ionisation-time of flight (MALDI-TOF) mass spectrometry (MS). Although MALDI-TOF MS was developed in the late 1980s, it is only in the last decade that it has emerged as a revolutionary technology in microbial diagnostics for the identification of different aerobic and anaerobic bacteria, yeasts and fungi (Anderson et al., 2012; Böhme et al., 2011; Griffin et al., 2012; Reich, Bosshard, Stark, Beyser, & Borgmann, 2013; Wieser, Schneider, Jung, & Schubert, 2012). Microbial detection using MALDI-TOF MS is also referred to as “whole cell” MS since the entire cell is analysed with minimal or no preextraction. In this technique, a single bacterial colony from solid culture media is spotted onto a target plate and overlaid with a matrix solution (containing solvents such as ethanol, acetonitrile, and trifluroacetic acid). The solvents in the matrix penetrate the cell wall and make the inner proteins available for crystallisation. A nitrogen laser is then utilised to ionise the microbial proteins (mainly ribosomal proteins); these ions subsequently travel through the TOF tube under vacuum until they reach the detector to generate characteristic spectra (Croxatto, Prod’hom, & Greub, 2012; Welker & Moore, 2011). The test spectra are then compared to commercially available reference databases such as MALDI Biotyper (Bruker Daltonics, Bremen, Germany) and VITEK MS (bioMérieux, Marcy l’Etoile, France), which include spectra obtained from known reference strains (Tani et al., 2012) (Figure 4). The main advantages of using MALDI-TOF MS in routine diagnostics is that is it a simple, cost-effective, rapid, and reliable technique that can be automated and incorporated into high-throughput diagnostic laboratories. More importantly, it is a single system that can be used for detecting multiple microbes.

Figure 4: Schematic representation of MALDI-TOF MS mechanism. Note: The analyte-matrix mixture is bombarded with a nitrogen laser beam, which ionises the analyte (red and orange circles) and the matrix ions (blue circles). These ions travel to the detector based on their relative mass to charge (m/z) ratio.
Figure 4:

Schematic representation of MALDI-TOF MS mechanism. Note: The analyte-matrix mixture is bombarded with a nitrogen laser beam, which ionises the analyte (red and orange circles) and the matrix ions (blue circles). These ions travel to the detector based on their relative mass to charge (m/z) ratio.

MALDI-TOF MS has been extensively used for the identification and subtyping of different clinically significant microbes (Murray, 2010). Other than routine identification from solid culture media, this technique has been used to detect bacteria from more complex samples such as urine (Ferreira et al., 2010), positive blood cultures (Drancourt, 2010), stool samples (Sparbier, Weller, Boogen, & Kostrzewa, 2012), and, more recently, from different food matrices (Jadhav et al., 2014). Some of the other applications of MALDI-TOF MS also include determination of antibiotic resistance in bacteria (Knox et al., 2014), analysis of specific protein biomarkers, and the determination of secondary metabolites (Vávrová, Matoulková, Žová, & Šedo, 2014). Whilst there have been numerous studies highlighting MALDI-TOF MS as a powerful tool for identification of clinical pathogens, few studies have investigated its application in the detection of environmentally important microbes, such as the ones involved in MIC.

Some of the microbial groups linked to MIC include SRB, methanogens, sulphur-oxidising bacteria, nitrate-reducing bacteria, and iron-oxidising bacteria (Li et al., 2013). Genera representing the above-mentioned groups that have been studied using MALDI-TOF MS include Pseudomonas (Anderson et al., 2012; Seng et al., 2013; Vithanage, Yeager, Jadhav, Palombo, & Datta, 2014), Desulfovibrio (Barreau, Pagnier, & La Scola, 2013; Justesen et al., 2012; Seng et al., 2013), Desulfitobacterium (Krader & Emerson, 2004), Methanothermobacter and Methanococcus (Krader & Emerson, 2004), and Sphingomonas (Seng et al., 2013). Since the commercial databases were originally developed for the identification of clinically significant pathogens, they have incomplete representation of environmental microorganisms (Kopcakova et al., 2014). Thus, it is possible that this technology, in its current format, may show limited success for identification of microbes implicated in MIC.

In a study conducted by Vávrová et al. (2014), MALDI-TOF MS was used to identify anaerobic bacteria obtained from biofilms in a brewery environment. Initially, 47% of isolates were able to be identified to the species level. Following creation of an in-house database for relevant bacteria that were absent in the commercial database, the speciation rate increased to about 85%. A similar approach was used by Ziegler et al. (2015) for identification of root nodule soil bacteria. After the development of a rhizobia-specific in-house database, identification of the bacteria was found to be in complete congruence with phylogenetic analysis. The above-mentioned studies highlight one of the main advantages of this technology – enabling the user to update the database with microbes more relevant to the type of analysis. Such in-house databases for bacteria linked to industry-specific MIC can be created and could also be exchanged between laboratories. Addition of more reference spectra to existing databases will result in more reliable identification of microbes responsible for or implicated in MIC.

When investigating MIC, it is important to understand all aspects of microbial biofilms. Cells in biofilms have been found to secrete certain extracellular polysaccharides (Belinky, Flikshtein, Lechenko, Gepstein, & Dosoretz, 2003), which serve as a binding component to provide a structural framework to the biofilm. EPS is known to vary with the type of microbial communities found within the biofilms. In an interesting study, Hasan, Gopal, and Wu (2011) used MALDI-TOF MS for the direct analysis of EPS produced by a biofilm formed on an aluminium surface exposed to sea water. The MIC was attributed predominantly to a fungus after characterising a specific polysaccharide “pullulan” in the biofilm. Thus, not only can this technology be used for microbial identification, but it can also be used for the direct analysis of biofilms involved in MIC. This can greatly reduce the time of analysis required for detecting the type of MIC at a particular site.

6 MIC biofilm metabolome and metabolomics

Metabolomics is the identification of the metabolites that are produced and consumed via either physical and/or chemical metamorphosis in biological samples (Beale et al., 2010). Metabolomics has been applied to understand metal corrosion and metals in the environment (Beale et al., 2010; Jones et al., 2015), food contamination (Beale, Morrison, & Palombo, 2014), general fungal and yeast metabolism (Brown et al., 2013; Christensen & Nielsen, 2000; Larsson, Snoep, Norbeck, & Albers, 2011; Niklas, Schneider, & Heinzle, 2010; Si, Luo, Xiao, & Zhao, 2014), and fungal based biomass degradation (Karpe, Beale, Harding, & Palombo, 2015). Furthermore, quantitative and qualitative metabolic profiling and flux analysis methods have been used to indicate the consumption of various substrates and relevant product generation. These approaches would considerably improve the understanding of fungal/bacterial-mediated corrosion of various metals and their minerals/alloys/metalloids. Metabolomic approaches also provide an insight to important metabolic pathways in bacterial (Cortés-Tolalpa et al., 2014) and fungal (Si et al., 2014) systems. These approaches have potential to improve fungal processing of metals and their related metalloids, alloys, and minerals.

Relatively few metabolomic-MIC studies have been reported. The role of corrosion products on MIC of carbon steel has been investigated by gas chromatography MS (GC/MS) (Liu, Xu, & Zeng, 2000), where S2- and organic acids were found to destroy the protective layer and promote hydrogen permeation. Furthermore, GC/MS-based metabolomics has also been explored as a potential tool in monitoring MIC in copper pipes in water distribution systems (Beale et al., 2010; Beale, Dunn, Morrison, Porter, & Marlow, 2012). It was found that the biofilm metabolites of bacteria-causing Cu pipe MIC comprised a combination of organic acids, amino acids, and lipids. These are common in biological metabolic processes, specifically those relating to soluble monomers and SRB substrates. In addition, the range of carboxylic acids released from the isolates (Bosea, Methylobacterium, Paenibacillus, Sphingomonas, and Variovorax) suggests that the corrosion potential of these biofilms varies, which would account for localised pitting corrosion commonly observed in metallic pipes (Beale et al., 2014). Kouremenos, Beale, Antti, and Palombo (2014) investigated the metabolic profile of Pseudomonas putida in potable water exposed to high and low doses of soluble and insoluble iron using liquid chromatography (LC) TOF-MS and identified metabolites that included nucleobases, nucleosides, dipeptides, tripeptides, amino acids, fatty acids, sugars, and phospholipids, significant as a response to exposure. While not directly related to MIC, the approach by Kouremenos et al. (2014) and the preliminary work of Beale et al. (2010, 2012) demonstrate the feasibility of GC- and LC-based metabolomics techniques to assess microbial populations exposed to metals or isolated from biofilms known to cause MIC in potable water networks. Furthermore, additional research on biofilms and biofilms known to cause MIC using other metabolomic based platforms, such as direct-injection MS, capillary electrophoresis-MS, Fourier transform infrared spectroscopy, and nuclear magnetic resonance spectroscopy, has been described (Dunn & Ellis, 2005).

In a study by Booth et al. (2011), the difference in response to metal stress between sessile and planktonic bacterial populations was characterised. The planktonic population had an oxidative stress response to copper ion exposure, whereas the same species within a biofilm environment responded by shifts in exopolysaccharide-related metabolism. This finding suggests that microbial responses pertinent to corrosion are different between sessile and planktonic populations, and through more research using metabolomic-based techniques, linkages between the metabolite activity of both sessile and planktonic populations and their release of extracellular metabolites from a biofilm can be achieved. While biofilms, in general, and their influence on corrosion, in particular, have been subject to extensive research, there has been limited experimental work performed on the in situ characterization of the organic compounds within biofilms (Beech et al., 2014; Graeber, Boehm, & Kuever, 2014).

Fungi, especially the spore-generating species, have the ability to survive under variable environmental stresses. Similar to bacteria, fungi can metabolise a number of carbon sources for survival, thus displaying a versatile nature of nutrient uptake. Fungi have been reported to grow in environments rich in heavy metals, such as mines and contaminated mine sites (Fogarty & Tobin, 1996; Gadd, 2007; Gadd et al., 2014).

A vast number of buildings, dams, and other constructions are based upon posttensioned concrete utilization. Stainless steel is heavily used to provide flexibility and support to this posttensioned concrete. However, to protect this steel from environmental factors such as humidity- and temperature-based corrosions, it is first coated with grease, which acts as a lubricant and corrosion protector. However, it has been recently reported that a number of fungi have the ability to utilise this grease as a source of nutrition, thus causing considerable contamination of this lubricant (Little, Staehle, & Davis, 2001; Lugauskas et al., 2009). The resultant fungal contamination causes the production of several organic acids. These acids cause considerable pitting in stainless steel over a long period, as evident from Figure 5 (De Leo, Campanella, Proverbio, & Urzì, 2013; Ehrlich, 1997).

Figure 5: SEM images of fungal mediated corrosion of lubrication grease and stainless steel tendons used in building construction. Sourced from De Leo et al. (2013), with permission from Elsevier. Note: Photographs (A), (C), (E) and (G) show growth of Aspergillus falvus, P. commune, F. solani, and A. acidophilus, respectively. Photographs (B), (D), (F), and (H) show the pit formation on stainless steel by Aspergillus falvus (Ø 400 μm), P. commune (Ø 350 μm), F. solani (Ø 1.2 mm), and A. acidophilus (Ø 1.2 mm), respectively.
Figure 5:

SEM images of fungal mediated corrosion of lubrication grease and stainless steel tendons used in building construction. Sourced from De Leo et al. (2013), with permission from Elsevier. Note: Photographs (A), (C), (E) and (G) show growth of Aspergillus falvus, P. commune, F. solani, and A. acidophilus, respectively. Photographs (B), (D), (F), and (H) show the pit formation on stainless steel by Aspergillus falvus (Ø 400 μm), P. commune (Ø 350 μm), F. solani (Ø 1.2 mm), and A. acidophilus (Ø 1.2 mm), respectively.

It has been recently reported that Aspergillus flavus, Penicillium commune, F. solani, and Acidomyces acidophilus displayed the ability to metabolise OVOLINE 71C grease and produce numerous organic acids (De Leo et al., 2013). The acids were observed to cause pitting in stainless steel within a period of 10 months. The level of corrosion was equated with that caused by H2SO4 at pH 3. Although the authors reported that 10 months of fungal-mediated corrosion was not considerable, longer periods would be able to cause significant damage to the steel. Similar results were observed by earlier workers investigating metal corrosion by different microbes (Emde, Smith, & Facey, 1992). In this instance, the authors observed that water distribution system pipes had numerous fungi present in the corrosion pockets. A considerable proportion of those organisms were reported as the yeast Rhodotorula, which was considered as one of the causal organisms for pipe corrosion (Emde et al., 1992).

Vanadium is a metal that has found increased usage in various industrial applications such as stainless steel and sulphuric acid manufacturing (Moskalyk & Alfantazi, 2003). In addition, an estimated 110,000 metric tonnes/year of vanadium is released in the air by burning petroleum products and mining (Anke, 2005). It has been observed (Ceci et al., 2014) that fungi such as A. niger, Aspergillus ustus, Clonostachys rosea, and P. lilacinus were able to colonise vanadate [Pb5(VO4)3Cl], causing pits and surface degradation. The degradation, especially by A. niger, caused precipitation of lead oxalate, which resulted in an increase in fungal tolerance towards toxic heavy metals such as lead, vanadium, and their salts. According to the authors, this fungal degradation of vanadium minerals has numerous applications in stabilising and controlling various industrial (including nuclear) wastes.

The major metabolites generated by fungi to counter metal toxicities are organic acids, especially numerous metal oxalates. A large number of fungi excrete oxalic acid, which reacts with the surrounding metals, metalloids, or metal minerals to convert them to metal oxalates. This not only decreases the toxicity of these metals (allowing the development of tolerance in fungi) but also causes structural damages, which may become prominent after a considerable time (Gadd et al., 2014). While not desirable for industry trying to combat fungal mediated corrosion, this property has immense potential for the treatment of heavy metal contaminated sites. The experiments conducted by several groups highlighted that the fungi such as Aspergillus terrus, Trichoderma viride (Joshi, Swarup, Maheshwari, Kumar, & Singh, 2011), and Paecilomyces javanicus (Rhee, Hillier, Pendlowski, & Gadd, 2014) have the ability to convert as much as 60 mg/g of lead (Pb) to nonhazardous minerals, such as pyromorphite. A similar process, termed biomachining, was developed and reported recently for tin (Sn), copper (Cu), aluminium (Al), and nickel (Ni) using A. niger (Jadhav & Hocheng, 2014).

Contrary to the above-mentioned experiments where various fungi have been used for metal or metal-based mineral corrosion, some experiments have also shown the utilization of fungal corrosive properties in restoration of metal and alloys. It has been proposed that fungi such as A. niger corrode copper to produce copper oxalate (Gharieb, Ali, & El-Shoura, 2004; Sayer & Gadd, 1997). Based on this principle, it was later reported that treatment of bronze coupons coated with Beauveria bassiana for 2 weeks at room temperature generated a coating of copper oxalate. This resulted in the prevention of further copper corrosion by environmental factors such as pollutants and humidity (Joseph, Simon, Mazzeo, Job, & Wörle, 2012). This aspect of fungal-based metal mineralization was also reported recently for the purpose of gold nanoparticle generation. It was observed that Rhizopus oryzae was able to reduce up to 130 μm HAuCl4 to Au+ and later to Au0 in an intracellular mechanism within an incubation period of 12 h at pH 4. Gold nanoparticles generated during the fungal-mediated reduction process were calculated to be of about 15 nm diameter, thus indicating their application as biomedical nanoparticles (Das, Liang, Schmidt, Laffir, & Marsili, 2012).

7 Conclusion

The study of functional genes involved in metabolic pathways is essential when attempting to link microbial diversity with specific ecological functions. In the context of this review, better knowledge of the role that microorganisms play in MIC and MIC processes such as biofilm formation and corrosion is required and through the application of omics-based approaches is being realised. Enhanced knowledge through the application of omics-based techniques provides a holistic opportunity to measure respiratory processes that influence oxygen concentrations, metabolite (acidic) reaction kinetics, mechanisms, pathways, and characterising biofilm population diversity and dynamics. Furthermore, the application of multiomic approaches provides further potential of accelerated understanding of these complex populations and processes through the combined assimilation of microorganism populations that are present and their activity, both localised and throughout a system.

The multiomic technology in MIC is novel as compared to various clinical applications. Although it is currently considered to be a nascent field, various potential applications resulting in considerable implications are either currently underway or are expected to appear on the scientific horizon in the near future. Until recently, a multiomic approach has not been applied in MIC and MIC-based research (Graeber et al., 2014). However, the research on concrete pipelines demonstrated the power of multiomics to identify key metabolic networks and gene functions within biofilms from corroded surfaces. It is expected that the exponential development in the overall omics knowledge, coupled with an rapidly developing technology in this field, will assist in exploration of various applications and related discoveries in the field of MIC and biofilm related research.


Corresponding author: David J. Beale, Land and Water Flagship, Commonwealth Scientific and Industrial Research Organisation (CSIRO), PO Box 2583, Brisbane 4001, Queensland, Australia, e-mail:

About the authors

David J. Beale

David J. Beale is a research scientist in the Land and Water Flagship at the Commonwealth Scientific and Industrial Research Organisation (CSIRO). He has more than 10 years of experience in the delivery of R&D projects, which include a portfolio of projects relating to sustainability, water management, and quality within the water and wastewater sector. He is also a technical expert and project leader on developing environmental metabolomic techniques for the assessment of pathogens and biofilms within water systems. He holds a bachelor’s degree with first class honours in environmental science and doctorate in analytical chemistry from RMIT University.

Avinash V. Karpe

Avinash V. Karpe has a PhD from the Faculty of Science, Engineering, and Technology, at Swinburne University of Technology and is currently a visiting scientist with CSIRO Land and Water. His primary research involves enhancing fungal bioprocessing for biofuel and medicinal metabolite production. Additionally, he is also involved in numerous metabolomic studies of fungal and bacterial communities in riverine, wastewater, and food processing systems.

Snehal Jadhav

Snehal Jadhav is a postdoctoral research fellow in Swinburne University of Technology, Melbourne, working in the area of microbial proteomics and metabolomics. Currently, her research is centred on the development of strategies based on MALDI-TOF MS, GC-MS, and LC-MS to identify and characterise bacteria obtained from food, clinical, and environmental sources. Previously, her research has also focussed on the effect of natural products against bacterial biofilms formed on abiotic surfaces.

Tim H. Muster

Tim H. Muster is a senior research scientist in the Cities Program in CSIRO Land & Water. He has over 20 years of research experience in the scientific disciplines of colloid, surface, and electrochemistry, with over 65 refereed journal publications. In 2007, Tim was the recipient of CSIRO Young Scientist John Philip Award and has twice won the Marshall Fordham Best Research Paper of the Australasian Corrosion Association (2003 and 2005). More recently, Tim was the recipient of a CSIRO Julius Career Award for nutrient recovery from wastewater and leads research focussed on the effective management of urban and food production waste streams for the productive recovery of water, energy, and nutrients.

Enzo A. Palombo

Enzo A. Palombo is chair at the Department of Chemistry and Biotechnology at Swinburne University of Technology in Australia. He has over 25 years of experience as a microbiologist and combines his academic teaching of microbiology and environmental biology with research interests in environmental microbiology, food microbiology, diagnostic microbiology, gastrointestinal microbiota, and bioactive compound discovery.

Acknowledgments

The authors gratefully acknowledge the help, support, and useful discussions on the manuscript provided by colleagues from the Land and Water Flagship at the Commonwealth Scientific and Industrial Research Organisation (CSIRO) and Swinburne University of Technology. Furthermore, the authors acknowledge the facilities, scientific, and technical assistance of the Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy, Characterization & Analysis, the University of Western Australia, a facility funded by the University, State and Commonwealth Governments.

References

Ahmed W, Staley C, Sadowsky MJ, Gyawali P, Sidhu JPS, Palmer A, Beale D, Toze S. Toolbox approaches using molecular markers and 16S rDNA amplicon datasets for the identification of fecal pollution in surface water. Appl Environ Microbiol 2015. doi: 10.1128/AEM.02032-15.10.1128/AEM.02032-15Search in Google Scholar PubMed PubMed Central

Anderson NW, Buchan BW, Riebe KM, Parsons, LN, Gnacinski S, Ledeboer NA. Effects of solid-medium type on routine identification of bacterial isolates by use of matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol 2012; 50: 1008–1013.10.1128/JCM.05209-11Search in Google Scholar PubMed PubMed Central

Anke M. Vanadium – An element both essential and toxic to plants, animals and humans? Anal Real Acad Nac Farm 2004; 70: 961–999.Search in Google Scholar

Balamurugan P, Hiren Joshi M, Rao TS. Microbial fouling community analysis of the cooling water system of a nuclear test reactor with emphasis on sulphate reducing bacteria. Biofouling 2011; 27: 967–978.10.1080/08927014.2011.618636Search in Google Scholar PubMed

Barreau M, Pagnier I, La Scola B. Improving the identification of anaerobes in the clinical microbiology laboratory through MALDI-TOF mass spectrometry. Anaerobe 2013; 22: 123–125.10.1016/j.anaerobe.2013.04.011Search in Google Scholar PubMed

Beale DJ, Dunn MS, Marney D. Application of GC-MS metabolic profiling to ‘blue-green water’ from microbial influenced corrosion in copper pipes. Corros Sci 2010; 52: 3140–3145.10.1016/j.corsci.2010.04.039Search in Google Scholar

Beale DJ, Dunn MS, Morrison PD, Porter NA, Marlow DR. Characterisation of bulk water samples from copper pipes undergoing microbially influenced corrosion by diagnostic metabolomic profiling. Corros Sci 2012; 55: 272–279.10.1016/j.corsci.2011.10.026Search in Google Scholar

Beale DJ, Morrison PD, Palombo, EA. Detection of Listeria in milk using non-targeted metabolic profiling of Listeria monocytogenes: a proof-of-concept application. Food Control 2014; 42: 343–346.10.1016/j.foodcont.2014.01.022Search in Google Scholar

Beale DJ, Morrison PD, Key C, Palombo EA. Metabolic profiling of biofilm bacteria known to cause microbial influenced corrosion. Water Sci Technol 2014; 69: 1–8.10.2166/wst.2013.425Search in Google Scholar PubMed

Beech IB, Gaylarde CC. Recent advances in the study of biocorrosion – an overview. Rev Microbiol 1999; 30: 177–190.Search in Google Scholar

Beech IB, Sztyler M, Gaylarde CC, Smith WL, Sunner J. 2 – Biofilms and biocorrosion. In: Liengen T, Féron D, Basséguy R, Beech IB, editors. Understanding Biocorrosion. Oxford, UK: Woodhead Publishing, 2014: 33–56.Search in Google Scholar

Belinky PA, Flikshtein N, Lechenko S, Gepstein S, Dosoretz CG. Reactive oxygen species and induction of lignin peroxidase in Phanerochaete chrysosporium. Appl Environ Microbiol 2003; 69: 6500–6506.10.1128/AEM.69.11.6500-6506.2003Search in Google Scholar PubMed PubMed Central

Böhme K, Fernández-No IC, Barros-Velázquez J, Gallardo JM, Cañas B, Calo-Mata P. Rapid species identification of seafood spoilage and pathogenic Gram-positive bacteria by MALDI-TOF mass fingerprinting. Electrophoresis 2011; 32: 2951–2965.10.1002/elps.201100217Search in Google Scholar PubMed

Booth SC, Workentine ML, Wen J, Shaykhutdinov R, Vogel HJ, Ceri H, Turner RJ, Weljie AM. Differences in metabolism between the biofilm and planktonic response to metal stress. J Proteome Res 2011; 10: 3190–3199.10.1021/pr2002353Search in Google Scholar

Brown S, Bashkirova L, Berka R, Chandler T, Doty T, McCall K, McCulloch M, McFarland S, Thompson S, Yaver D, Berry A. Metabolic engineering of Aspergillus oryzae NRRL 3488 for increased production of l-malic acid. Appl Microbiol Biotechnol 2013; 97: 8903–8912.10.1007/s00253-013-5132-2Search in Google Scholar

Ceci A, Rhee YJ, Kierans M, Hillier S, Pendlowski H, Gray N, Persiani AM, Gadd GM. Transformation of vanadinite [Pb5(VO4)3Cl] by fungi. Environ Microbiol 2015; 17: 2018–2034.10.1111/1462-2920.12612Search in Google Scholar

Chakraborty A, DasGupta CK, Bhadury P. 4 – Application of molecular techniques for the assessment of microbial communities in contaminated sites. In: Das S, editor. Microbial biodegradation and bioremediation. Oxford, UK: Elsevier, 2014: 85–113.Search in Google Scholar

Characklis WG. Bioengineering report: fouling biofilm development: a process analysis. Biotechnol Bioeng 1981; 23: 1923–1960.10.1002/bit.260230902Search in Google Scholar

Christensen B, Nielsen J. Metabolic network analysis of penicillium chrysogenumusing13c-labeled glucose. Biotechnol Bioeng 2000; 68: 652–659.10.1002/(SICI)1097-0290(20000620)68:6<652::AID-BIT8>3.0.CO;2-JSearch in Google Scholar

Cortés-Tolalpa L, Gutiérrez-Ríos RM, Martínez LM, de Anda R, Gosset G, Bolívar F, Escalante A. Global transcriptomic analysis of an engineered Escherichia coli strain lacking the phosphoenolpyruvate: carbohydrate phosphotransferase system during shikimic acid production in rich culture medium. Microb Cell Fact 2014; 13: 28.10.1186/1475-2859-13-28Search in Google Scholar

Croxatto A, Prod’hom G, Greub G. Applications of MALDI-TOF mass spectrometry in clinical diagnostic microbiology. FEMS Microbiol Rev 2012; 36: 380–407.10.1111/j.1574-6976.2011.00298.xSearch in Google Scholar

Das SK, Liang J, Schmidt M, Laffir F, Marsili E. Biomineralization mechanism of gold by zygomycete fungi Rhizopous oryzae. ACS Nano 2012; 6: 6165–6173.10.1021/nn301502sSearch in Google Scholar

De Leo F, Campanella G, Proverbio E, Urzì C. Laboratory tests of fungal biocorrosion of unbonded lubricated post-tensioned tendons. Constr Build Mater 2013; 49: 821–827.10.1016/j.conbuildmat.2013.08.071Search in Google Scholar

Dexter SC, Duquette DJ, Siebert OW, Videla HA. Use and limitations of electrochemical techniques for investigating microbial corrosion. Corrosion 1991; 47: 308–318.10.5006/1.3585258Search in Google Scholar

Douterelo I, Boxall JB, Deines P, Sekar R, Fish KE, Biggs CA. Methodological approaches for studying the microbial ecology of drinking water distribution systems. Water Res 2014; 65: 134–156.10.1016/j.watres.2014.07.008Search in Google Scholar

Drancourt M. Detection of microorganisms in blood specimens using matrix-assisted laser desorption ionization time-of-flight mass spectrometry: a review. Clin Microbiol Infect 2010; 16: 1620–1625.10.1111/j.1469-0691.2010.03290.xSearch in Google Scholar

Dunn WB, Ellis DI. Metabolomics: current analytical platforms and methodologies. TrAC Trends Anal Chem 2005; 24: 285–294.10.1016/j.trac.2004.11.021Search in Google Scholar

Edyvean RGJ, Videla HA. Biological corrosion. Interdiscip Sci Rev 1991; 16: 267–282.10.1179/isr.1991.16.3.267Search in Google Scholar

Ehrlich HL. Microbes and metals. Appl Microbiol Biotechnol 1997; 48: 687–692.10.1007/s002530051116Search in Google Scholar

Emde KME, Smith DW, Facey R. Initial investigation of microbially influenced corrosion –MIC-- in a low temperature water distribution system. Water Res 1992; 26: 169–175.10.1016/0043-1354(92)90216-QSearch in Google Scholar

Ferreira L, Sánchez-Juanes F, González-Ávila M, Cembrero-Fuciños D, Herrero-Hernández A, González-Buitrago JM, Muñoz-Bellido JL. Direct identification of urinary tract pathogens from urine samples by matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol 2010; 48: 2110–2115.10.1128/JCM.02215-09Search in Google Scholar

Flemming HC. Microbial deterioration of materials-fundamentals-economical and technical overview. Werkst Korros 1994; 45: 5–9.10.1002/maco.19940450105Search in Google Scholar

Fogarty RV, Tobin JM. Fungal melanins and their interactions with metals. Enzyme Microb Technol 1996; 19: 311–317.10.1016/0141-0229(96)00002-6Search in Google Scholar

Gadd GM. Geomycology: biogeochemical transformations of rocks, minerals, metals and radionuclides by fungi, bioweathering and bioremediation. Mycol Res 2007; 111: 3–49.10.1016/j.mycres.2006.12.001Search in Google Scholar PubMed

Gadd GM, Bahri-Esfahani J, Li Q, Rhee YJ, Wei Z, Fomina M, Liang X. Oxalate production by fungi: significance in geomycology, biodeterioration and bioremediation. Fungal Biol Rev 2014; 28: 36–55.10.1016/j.fbr.2014.05.001Search in Google Scholar

Gharieb M, Ali M, El-Shoura A. Transformation of copper oxychloride fungicide into copper oxalate by tolerant fungi and the effect of nitrogen source on tolerance. Biodegradation 2004; 15: 49–57.10.1023/B:BIOD.0000009962.48723.dfSearch in Google Scholar

Gomez-Alvarez V. Biofilm-growing bacteria involved in the corrosion of concrete wastewater pipes: protocols for comparative metagenomic analyses. In: Donelli G, editor. Microbial biofilms. Vol. 1147. New York: Springer, 2014: 323–340.Search in Google Scholar

Gomez-Alvarez V, Revetta R, Domingo JW. Metagenome analyses of corroded concrete wastewater pipe biofilms reveal a complex microbial system. BMC Microbiol 2012; 12: 122.10.1186/1471-2180-12-122Search in Google Scholar PubMed PubMed Central

Graeber M, Boehm S, Kuever J. 3 – Molecular methods for studying biocorrosion. In: Liengen T, Féron D, Basséguy R, Beech IB, editors. Understanding biocorrosion. Oxford, UK: Woodhead Publishing, 2014: 57–75.Search in Google Scholar

Griffin PM, Price GR, Schooneveldt JM, Schlebusch S, Tilse MH, Urbanski T, Hamilton B, Venter D. Use of matrix-assisted laser desorption ionization-time of flight mass spectrometry to identify vancomycin-resistant enterococci and investigate the epidemiology of an outbreak. J Clin Microbiol 2012; 50: 2918–2931.10.1128/JCM.01000-12Search in Google Scholar PubMed PubMed Central

Handelsman J. Metagenomics: application of genomics to uncultured microorganisms. Microbiol Mol Biol Rev 2004; 68: 669–685.10.1128/MMBR.68.4.669-685.2004Search in Google Scholar PubMed PubMed Central

Hasan N, Gopal J, Wu HF. Rapid, sensitive and direct analysis of exopolysaccharides from biofilm on aluminum surfaces exposed to sea water using MALDI-TOF MS. J Mass Spectrom 2011; 46: 1160–1167.10.1002/jms.2003Search in Google Scholar PubMed

Herrera LK, Videla HA. Role of iron-reducing bacteria in corrosion and protection of carbon steel. Int Biodeterior Biodegrad 2009; 63: 891–895.10.1016/j.ibiod.2009.06.003Search in Google Scholar

Jack TR, Rogoz E, Bramhill B, Roberge PR. The characterization of sulphate-reducing bacteria in heavy oil and waterflood operations. In: Kearns JR, Little BJ, editors. Microbiologically influenced corrosion testing. Vol. 1232. W Conshohocken, PA: American Society Testing and Materials, 1994: 108–117.Search in Google Scholar

Jadhav U, Hocheng H. Use of Aspergillus niger 34770 culture supernatant for tin metal removal. Corros Sci 2014; 82: 248–254.10.1016/j.corsci.2014.01.011Search in Google Scholar

Jadhav S, Sevior D, Bhave M, Palombo EA. Detection of Listeria monocytogenes from selective enrichment broth using MALDI-TOF mass spectrometry. J Proteomics 2014; 97: 100–106.10.1016/j.jprot.2013.09.014Search in Google Scholar PubMed

Jiang F, Kongsaeree P, Charron R, Lajoie C, Xu H, Scot G, Kelly C. Production and separation of manganese peroxidase from heme amended yeast cultures. Biotechnol Bioeng 2008; 99: 540–549.10.1002/bit.21590Search in Google Scholar PubMed

Jones OA, Dias DA, Callahan DL, Kouremenos KA, Beale DJ, Roessner U. The use of metabolomics in the study of metals in biological systems. Metallomics 2015; 7: 29–38.10.1039/C4MT00123KSearch in Google Scholar PubMed

Joseph E, Simon A, Mazzeo R, Job D, Wörle M. Spectroscopic characterization of an innovative biological treatment for corroded metal artefacts. J Raman Spectrosc 2012; 43: 1612–1616.10.1002/jrs.4164Search in Google Scholar

Joshi PK, Swarup A, Maheshwari S, Kumar R, Singh N. Bioremediation of heavy metals in liquid media through fungi isolated from contaminated sources. Ind J Microbiol 2011; 51: 482–487.10.1007/s12088-011-0110-9Search in Google Scholar PubMed PubMed Central

Justesen US, Holm A, Knudsen E, Andersen LB, Jensen TG, Kemp M, Skov MN, Gahrn-Hansen B, Møller JK. Species identification of clinical isolates of anaerobic bacteria: a comparison of two matrix-assisted laser desorption ionization-time of flight mass spectrometry systems. J Clin Microbiol 2012; 50: 542.10.1128/JCM.06542-11Search in Google Scholar

Karpe AV, Beale DJ, Harding IH, Palombo EA. Optimization of degradation of winery-derived biomass waste by Ascomycetes. J Chem Tech Biotechnol 2015; 90: 1793–1801.10.1002/jctb.4486Search in Google Scholar

Keevil CW. The physico-chemistry of biofilm-mediated pitting corrosion of copper pipe supplying potable water. Water Sci Technol 2004; 49: 91–98.10.2166/wst.2004.0096Search in Google Scholar

Kip N, van Veen JA. The dual role of microbes in corrosion. ISME J 2015; 9: 542–551.10.1038/ismej.2014.169Search in Google Scholar PubMed PubMed Central

Knox J, Jadhav S, Sevior D, Agyekum A, Whipp M, Waring L, Iredell J, Palombo E. (2014). Phenotypic detection of carbapenemase-producing Enterobacteriaceae by use of Matrix-Assisted Laser Desorption Ionization–Time of Flight Mass Spectrometry and the Carba NP test. J Clin Microbiol 2014; 52: 4075–4077.Search in Google Scholar

Kopcakova A, Stramova Z, Kvasnova S, Godany A, Perhacova Z, Pristas P. Need for database extension for reliable identification of bacteria from extreme environments using MALDI TOF mass spectrometry. Chem Pap 2014; 68: 1435–1442.10.2478/s11696-014-0612-0Search in Google Scholar

Kouremenos KA, Beale DJ, Antti H, Palombo EA. Liquid chromatography time of flight mass spectrometry based environmental metabolomics for the analysis of Pseudomonas putida bacteria in potable water. J Chromatogr 2014; 966: 179–186.10.1016/j.jchromb.2014.02.058Search in Google Scholar PubMed

Krader P, Emerson D. Identification of archaea and some extremophilic bacteria using matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry. Extremophiles 2004; 8: 259–268.10.1007/s00792-004-0382-7Search in Google Scholar PubMed

Larsson C, Snoep JL, Norbeck J, Albers E. Flux balance analysis for ethylene formation in genetically engineered Saccharomyces cerevisiae. IET Syst Biol 2011; 5: 245–251.10.1049/iet-syb.2010.0027Search in Google Scholar

Lee JS, Little BJ, Ray RI, Stropki JT. Fate of Cr+6 from a coating in an electrolyte with microorganisms. J Electrochem Soc 2012; 159: C530–C538.10.1149/2.057211jesSearch in Google Scholar

Lenhart TR, Duncan KE, Beech IB, Sunner JA, Smith W, Bonifay V, Biri B, Suflita JM. Identification and characterization of microbial biofilm communities associated with corroded oil pipeline surfaces. Biofouling 2014; 30: 823–835.10.1080/08927014.2014.931379Search in Google Scholar

Li K, Whitfield M, Van Vliet Krystyn J. Beating the bugs: roles of microbial biofilms in corrosion. Corros Rev 2013; 31: 73–84.10.1515/corrrev-2013-0019Search in Google Scholar

Little BJ, Ray RI, Pope RK. Relationship between corrosion and the biological sulfur cycle: a review. Corrosion 2000; 56: 433–443.10.5006/1.3280548Search in Google Scholar

Little B, Staehle R, Davis R. Fungal influenced corrosion of post-tensioned cables. Int Biodeterior Biodegrad 2001; 47: 71–77.10.1016/S0964-8305(01)00039-7Search in Google Scholar

Liu H, Xu L, Zeng J. Role of corrosion products in biofilms in microbiologically induced corrosion of carbon steel. Br Corros J 2000; 35: 131–135.10.1179/000705900101501155Search in Google Scholar

Lugauskas A, Prosycevas I, Ramanauskas R, Asta Griguceviciene A, Selskiene A, Pakštas V. The influence of micromycetes on the corrosion behaviour of metals (Steel, Al) under conditions of the environment polluted with organic substances. Mater Sci (Medziagotyra) 2009; 15: 224–235.Search in Google Scholar

Mohan AM, Bibby KJ, Lipus D, Hammack RW, Gregory KB. The functional potential of microbial communities in hydraulic fracturing source water and produced water from natural gas extraction characterized by metagenomic sequencing. PLoS One 2014; 9: e107682.10.1371/journal.pone.0107682Search in Google Scholar

Moskalyk RR, Alfantazi AM. Processing of vanadium: a review. Miner Eng 2003; 16: 793–805.10.1016/S0892-6875(03)00213-9Search in Google Scholar

Murray PR. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry: usefulness for taxonomy and epidemiology. Clin Microbiol Infect 2010; 16: 1626–1630.10.1111/j.1469-0691.2010.03364.xSearch in Google Scholar PubMed

Muyzer G. DGGE/TGGE a method for identifying genes from natural ecosystems. Current Opin Microbiol 1999; 2: 317–322.10.1016/S1369-5274(99)80055-1Search in Google Scholar

Naranjo L, Pernía B, Inojosa Y, Rojas D, D’Anna LS, González M, De Sisto Á. First evidence of fungal strains isolated and identified from naphtha storage tanks and transporting pipelines in Venezuelan oil facilities. Adv Microbiol 2015; 5: 143.10.4236/aim.2015.52014Search in Google Scholar

Neria-González I, Wang ET, Ramírez F, Romero JM, Hernández-Rodríguez C. Characterization of bacterial community associated to biofilms of corroded oil pipelines from the southeast of Mexico. Anaerobe 2006; 12: 122–133.10.1016/j.anaerobe.2006.02.001Search in Google Scholar PubMed

Niklas J, Schneider K, Heinzle E. Metabolic flux analysis in eukaryotes. Curr Opin Biotechnol 2010; 21: 63–69.10.1016/j.copbio.2010.01.011Search in Google Scholar PubMed

Oliveira VM, Lopes-Oliveira PF, Passarini MRZ, Menezes CBA, Oliveira WRC, Rocha AJ, Sette LD. Molecular analysis of microbial diversity in corrosion samples from energy transmission towers. Biofouling 2011; 27: 435–447.10.1080/08927014.2011.581751Search in Google Scholar PubMed

Pascault N, Loux V, Derozier S, Martin V, Debroas D, Maloufi S, Humbert J-F, Leloup J. Technical challenges in metatranscriptomic studies applied to the bacterial communities of freshwater ecosystems. Genetica 2015; 143: 157–167.10.1007/s10709-014-9783-4Search in Google Scholar PubMed

Pavissich JP, Vargas IT, González B, Pastén PA, Pizarro GE. Culture dependent and independent analyses of bacterial communities involved in copper plumbing corrosion. J Appl Microbiol 2010; 109: 771–782.10.1111/j.1365-2672.2010.04704.xSearch in Google Scholar PubMed

Potekhina JS, Sherisheva NG, Povetkina LP, Pospelov AP, Rakitina TA, Warnecke F, Gottschalk G. Role of microorganisms in corrosion inhibition of metals in aquatic habitats. Appl Microbiol Biotechnol 1999; 52: 639–646.10.1007/s002530051571Search in Google Scholar

Reich M, Bosshard PP, Stark M, Beyser K, Borgmann S. Species identification of bacteria and fungi from solid and liquid culture media by MALDI-TOF mass spectrometry. J Bacteriol Parasitol 2013; S5: 1–8.10.4172/2155-9597.S5-002Search in Google Scholar

Rhee YJ, Hillier S, Pendlowski H, Gadd GM. Fungal transformation of metallic lead to pyromorphite in liquid medium. Chemosphere 2014; 113: 17–21.10.1016/j.chemosphere.2014.03.085Search in Google Scholar PubMed

Santo Domingo JW, Revetta RP, Iker B, Gomez-Alvarez V, Garcia J, Sullivan J, Weast J. Molecular survey of concrete sewer biofilm microbial communities. Biofouling 2011; 27: 993–1001.10.1080/08927014.2011.618637Search in Google Scholar PubMed

Sayer JA, Gadd GM. Solubilization and transformation of insoluble inorganic metal compounds to insoluble metal oxalates by Aspergillus niger. Mycol Res 1997; 101: 653–661.10.1017/S0953756296003140Search in Google Scholar

Seng P, Abat C, Rolain JM, Colson P, Lagier JC, Gouriet F, Fournier PE, Drancourt M, Scola BL, Raoult D. Identification of rare pathogenic bacteria in a clinical microbiology laboratory: impact of matrix-assisted laser desorption ionization-time of flight mass spectrometry. J Clin Microbiol 2013; 51: 2182–2194.10.1128/JCM.00492-13Search in Google Scholar PubMed PubMed Central

Sette L, Passarini M, Rodrigues A, Leal R, Simioni K, Nobre F, de Brito B, da Rocha A, Pagnocca F. Fungal diversity associated with Brazilian energy transmission towers. Fungal Diversity 2010; 44: 53–63.10.1007/s13225-010-0048-ySearch in Google Scholar

Shoesmith DW, Taylor P, Bailey MG, Owen DG. The formation of ferrous monosulphide polymorphos during the corrosion of iron by aqueous hydrogen-sulfide at 21-degrees C. J Electrochem Soc 1980; 127: 1007–1015.10.1149/1.2129808Search in Google Scholar

Si T, Luo Y, Xiao H, Zhao H. Utilizing an endogenous pathway for 1-butanol production in Saccharomyces cerevisiae. Metab Eng 2014; 22: 60–68.10.1016/j.ymben.2014.01.002Search in Google Scholar PubMed

Sierra-Alvarez R. Fungal bioleaching of metals in preservative-treated wood. Proc Biochem 2007; 42: 798–804.10.1016/j.procbio.2007.01.019Search in Google Scholar

Smirnov VF, Belov DV, Sokolova TN, Kuzina OV, Kartashov VR. Microbiological corrosion of aluminum alloys. Appl Biochem Microbiol 2008; 44: 192–196.10.1134/S0003683808020117Search in Google Scholar

Sparbier K, Weller U, Boogen C, Kostrzewa M. Rapid detection of Salmonella sp. by means of a combination of selective enrichment broth and MALDI-TOF MS. Eur J Clin Microbiol Infect Dis 2012; 31: 767–773.10.1007/s10096-011-1373-0Search in Google Scholar PubMed

Stratmann M. The atmospheric corrosion of iron – a discussion of the physicochemical fundamentals of this omnipresent corrosion process – invited review. Ber Bunsen Ges 1990; 94: 626–639.10.1002/bbpc.19900940603Search in Google Scholar

Tani A, Sahin N, Matsuyama Y, Enomoto T, Nishimura N, Yokota A, Kimbara K. High-throughput identification and screening of novel Methylobacterium species using whole-cell MALDI-TOF/MS analysis. PLoS One 2012; 7: e40784.10.1371/journal.pone.0040784Search in Google Scholar PubMed PubMed Central

Terry LA, Edyvean RGJ. Mocroalgae and corrosion. Botanica Marina 1981; 24: 177–183.Search in Google Scholar

Terry LA, Edyvean RGJ. Recent investigations into the effects of algae on corrosion. In: Evans LV, Hoagland KD, editors. Studies in environmental science. Vol. 28. Chapter 15. Amsterdam: Elsevier, 1986: 211–229.Search in Google Scholar

Usher KM, Kaksonen AH, MacLeod ID. Marine rust tubercles harbour iron corroding archaea and sulphate reducing bacteria. Corros Sci 2014; 83: 189–197.10.1016/j.corsci.2014.02.014Search in Google Scholar

Van Veen SQ, Claas ECJ, Kuijper EJ. High-throughput identification of bacteria and yeast by matrix-assisted laser desorption ionization-time of flight mass spectrometry in conventional medical microbiology laboratories. J Clin Microbiol 2010; 48: 900–907.10.1128/JCM.02071-09Search in Google Scholar PubMed PubMed Central

Vávrová A, Matoulková D, Žová TB, Šedo O. MALDI-TOF MS analysis of anaerobic bacteria isolated from biofilm-covered surfaces in brewery bottling halls. J Am Soc Brew Chem 2014; 72: 95–101.10.1094/ASBCJ-2014-0324-01Search in Google Scholar

Videla HA. Biocorrosion and biofouling of metals and alloys of industrial usage. Present state of the art at the beginning of the new millennium. Rev Metal 2003; 5: 256–264.Search in Google Scholar

Videla HA, Herrera LK. Understanding microbial inhibition of corrosion. A comprehensive overview. Int Biodeterior Biodegrad 2009; 63: 896–900.Search in Google Scholar

Vincke E, Boon N, Verstraete W. Analysis of the microbial communities on corroded concrete sewer pipes – a case study. Appl Microbiol Biotechnol 2001; 57: 776–785.10.1007/s002530100826Search in Google Scholar PubMed

Vithanage NR, Yeager TR, Jadhav SR, Palombo EA, Datta N. Comparison of identification systems for psychrotrophic bacteria isolated from raw bovine milk. Int J Food Microbiol 2014; 189: 26–38.10.1016/j.ijfoodmicro.2014.07.023Search in Google Scholar PubMed

Welker M, Moore ERB. Applications of whole-cell matrix-assisted laser-desorption/ionization time-of-flight mass spectrometry in systematic microbiology. Syst Appl Microbiol 2011; 34: 2–11.10.1016/j.syapm.2010.11.013Search in Google Scholar PubMed

Wieser A, Schneider L, Jung J, Schubert S. MALDI-TOF MS in microbiological diagnostics – identification of microorganisms and beyond (mini review). Appl Microbiol Biotechnol 2012; 93: 965–974.10.1007/s00253-011-3783-4Search in Google Scholar PubMed

Wikieł AJ, Datsenko I, Vera M, Sand W. Impact of Desulfovibrio alaskensis biofilms on corrosion behaviour of carbon steel in marine environment. Bioelectrochemistry 2014; 97: 52–60.10.1016/j.bioelechem.2013.09.008Search in Google Scholar PubMed

Ziegler D, Pothier J, Ardley J, Fossou R, Pflüger V, de Meyer S, Vogel G, Tonolla M, Howieson J, Reeve W, Perret X. Ribosomal protein biomarkers provide root nodule bacterial identification by MALDI-TOF MS. Appl Microbiol Biotechnol 2015; 99: 1–16.10.1007/s00253-015-6515-3Search in Google Scholar PubMed

Zuo R. Biofilms: strategies for metal corrosion inhibition employing microorganisms. Appl Microbiol Biotechnol 2007; 76: 1245–1253.10.1007/s00253-007-1130-6Search in Google Scholar PubMed

Received: 2015-5-28
Accepted: 2015-8-20
Published Online: 2015-10-7
Published in Print: 2016-3-1

©2016 by De Gruyter

Downloaded on 18.4.2024 from https://www.degruyter.com/document/doi/10.1515/corrrev-2015-0046/html
Scroll to top button