!ASM International 1059-9495/$19.00 JMEPEG DOI: 10.1007/s11665-013-0627-7 Microbial Corrosion in Linepipe Steel Under the Influence of a Sulfate-Reducing Consortium Isolated from an Oil Field Faisal M. AlAbbas, Charles Williamson, Shaily M. Bhola, John R. Spear, David L Olson, Brajendra Mishra, and Anthony E. Kakpovbia (Submitted January 13, 2013; in revised form June 4, 2013) This work investigates microbiologically influenced corrosion of API 5L X52 linepipe steel by a sulfatereducing bacteria (SRB) consortium. The SRB consortium used in this study was cultivated from a sour oil well in Louisiana, USA. 16S rRNA gene sequence analysis indicated that the mixed bacterial consortium contained three phylotypes: members of Proteobacteria (Desulfomicrobium sp.), Firmicutes (Clostridium sp.), and Bacteroidetes (Anaerophaga sp.). The biofilm and the pits that developed with time were characterized using field emission scanning electron microscopy (FE-SEM). In addition, electrochemical impedance spectroscopy (EIS), linear polarization resistance (LPR) and open circuit potential (OCP) were used to analyze the corrosion behavior. Through circuit modeling, EIS results were used to interpret the physicoelectric interactions between the electrode, biofilm and solution interfaces. The results confirmed that extensive localized corrosion activity of SRB is due to a formed biofilm in conjunction with a porous iron sulfide layer on the metal surface. X-ray diffraction (XRD) revealed semiconductive corrosion products predominantly composed of a mixture of siderite (FeCO3), iron sulfide (FexSy), and iron (III) oxidehydroxide (FeOOH) constituents in the corrosion products for the system exposed to the SRB consortium. Keywords biofilm, carbon steel API 5L X52, microbiologically influenced corrosion, pipeline, sulfate-reducing bacteria 1. Introduction Microbiologically influenced corrosion (MIC), or biocorrosion, is of considerable concern to the oil and gas industry. MIC is induced by indigenous microorganisms that naturally reside in hydrocarbon and secondary water injection systems (Ref 1-3). Microbial metabolic products, for example, sulfide and organic acids such as acetic acid, may alter interface chemistry, resulting in increased corrosion rates. Furthermore, biofilms consisting of complex communities of microbes and extracellular polymeric substances (EPS) that have developed on the metal surface create gradients of pH and dissolved oxygen, leading to localized forms of corrosion, such as pitting and crevice formation (Ref 1-4). Since the advent of modern oil and gas production, scientists and engineers have faced problems caused by microorganisms (Ref 5). Hydrocarbons are an excellent carbon (food) source for a wide variety of microbes in all three domains of life—the Faisal M. AlAbbas, Shaily M. Bhola, David L Olson, and Brajendra Mishra, Department of Metallurgical and Materials Engineering, Colorado School of Mines, Golden, CO 80401; Charles Williamson and John R. Spear, Department of Civil and Environmental Engineering, Colorado School of Mines, Golden, CO 80401; and Anthony E. Kakpovbia, Inspection Department, Saudi Aramco, Dhahran 31311, Saudi Arabia. Contact e-mail: falabbas@mines.edu. Journal of Materials Engineering and Performance Bacteria, Archaea and Eucarya, and microbial representatives of all the three domains likely play roles in MIC (Ref 2). Bacterial activity (e.g., sulfate reduction) is believed to be responsible for greater than 75% of the corrosion in production oil and gas wells, and it is likely that greater than 50% of failures of buried pipelines and cables are due to the metabolic activities of microbes (Ref 5). The main types of bacteria associated with metals in pipeline systems are sulfate-reducing bacteria, iron reducing bacteria, iron and manganese oxidizing bacteria, and acid producing bacteria (Ref 1, 3, 5). For several reasons, SRB have been recognized as the major contributor to MIC in pipeline systems primarily due to their predominantly anaerobic lifestyle and continuous production of corrosive hydrogen sulfide (Ref 3, 5). However, in practical situations, MIC results from synergistic interactions of different microbial consortia, which coexist in the environment and are able to affect the electrochemical processes through co-operative metabolisms (Ref 4). As early as 1926, traditional microbiologists detected the presence of SRB in oil environments (Ref 5). Studies have shown that most kinds of cultivated SRB are responsible for the production of hydrogen sulfide, which is a toxic and corrosive gas in both aerial and aqueous form (Ref 4, 5). At a minimum, hydrogen sulfide is thought to be responsible for a variety of environmental effects which include reservoir souring, contamination of natural gas and oil with H2S, corrosion of metal surfaces, the ‘‘plugging’’ of reservoirs due to the formation of extensive pore-space microbial biofilms and the precipitation of metal sulfides (Ref 5). SRB has received much attention in the oil & gas industry and MIC investigations, as these microorganisms have several detrimental metabolic activities including the ability to: (1) oxidize hydrogen, (2) use O2 and Fe3+, (3) utilize aliphatic and aromatic hydrocarbons, (4) couple sulfate reduction to the intracellular production of magnetite and (5) compete with nitrate-reducing/sulfur-oxidizing bacteria (NRB-SOB) (since they may have a nitrite reducing activity) (Ref 5, 6). In total, all of these activities can significantly degrade the quality of oil and gas in a reservoir to be economically extractable. Moreover, SRB adversely impacts the integrity of oil and gas installations. This study investigates the impact of environmental SRB (cultivated from oil field samples rather than obtained from a culture collection) on the corrosion behavior of API 5L X52 linepipe steel. The SRB consortium used in this study was cultivated from an oil well in Louisiana, USA (Ref 7). 16S rRNA gene sequence analysis was performed to identify the SRB species. The nature and kinetics of chemical and electrochemical reactions introduced by SRB activities on API 5L X52 carbon steel coupons were characterized using electrochemical impedance spectroscopy, linear polarization resistance (LPR) and open circuit potential (OCP). The biofilm and corrosion morphology were investigated using field emission scanning electron microscopy coupled with energy dispersive spectroscopy (EDS). The composition of corrosion products was evaluated using x-ray diffraction analysis. 2. Materials and Methods 2.1 Microorganisms and Testing Medium The SRB consortium used in this study was cultivated from water samples obtained from an oil well located in Louisiana, USA. The water samples were collected and bottled at the well head from an approximate depth of 2200 ft. as described under the NACE Standard TM0194 (2004) (Ref 8). The SRB were cultivated in a modified Baar!s medium (ATCC medium 1250). Baar!s medium is reported to be suitable when studying the influence of mixed bacterial communities on steel corrosion, though many different media recipes exist (e.g., the Postgate series of media) and any number of them likely work (Ref 9). This growth medium was composed of magnesium sulfate (2.0 g), sodium citrate (5.0 g), calcium sulfate di-hydrate (1.0 g), ammonium chloride (1.0 g), sodium chloride (25.0 g), di-potassium hydrogen orthophosphate (0.5 g), sodium lactate 60% syrup (3.5 g), and yeast extract (1.0 g). All components were per liter of distilled water. The pH of the medium was adjusted to 7.5 using 5 M sodium hydroxide and sterilized in an autoclave at 121 "C for 20 min. The SRB species were cultured in the growth medium with filter-sterilized 5% w/w ferrous ammonium sulfate added to the medium at a ratio of 0.1-5.0 mL respectively. The bacteria were incubated for 72 h at 37 "C under an oxygen-free nitrogen headspace. 2.2 Sulfate-Reducing Consortium Identifications Genomic DNA was extracted from the bacterial consortium using the MoBio Powersoil DNA extraction kit (MoBio, Carlsbad, CA) with the 10-min vortexing step replaced by 1 min of bead beating. Primers 515F (5¢-GTGCCAGCMG CCGCGGTAA-3¢) and 1391R (5¢-GACGGGCGGTGWGT RCA-3¢) (Ref 7, 10) were used to amplify 16S rRNA genes. Polymerase chain reaction (PCR), cloning and transformation were conducted as described by Sahl et al. (Ref 11). Unique restriction fragment length polymorphisms (RFLP) were sequenced on an ABI 3730 DNA sequencer at Davis Sequencing, Inc. (Davis, CA), and Sanger sequence reads were called with PHRED (Ref 12, 13) via Xplorseq (Ref 14). Sequences were aligned with the SINA aligner (Ref 15) and added (parsimony insertion) to the Silva SSURef111_NR guide tree (Ref 16) with the ARB software package (Ref 17). Sequences were also compared to the Genbank database (Ref 18) via the Basic Local Alignment Search Tool (BLAST) (Ref 19). Phylogenetic trees were created with RAxML (Ref 20). Relevant sequences were obtained from the Genbank database and aligned and masked with ssu-align (Ref 21). Phylogenetic trees were created using the GTR substitution model and the gamma distribution of rate heterogeneity. The number of bootstrap replicates was determined using the RAxML frequency-based criterion (Ref 22). 16S rRNA gene sequences produced during this research have been deposited in the Genbank with accession numbers KC756849-KC756851. 2.3 Sample Preparation Pipeline steel (API 5L X52) coupons, provided by a local energy company (Saudi ARAMCO), were used for this study, and their chemical compositions are shown in Table 1. Metallographic specimens from the received materials were prepared with standard methods for optical microscopy (1 lm final polish and 2% nital etch). Representative microstructure contains a mixture of polygonal ferrite and pearlite structures. For corrosion evaluations, the coupons were machined to a size of 10 mm 9 10 mm 9 5 mm and embedded in a mold of non-conducting epoxy resin, leaving an exposed surface area of 100 mm2. For electrical connection, a copper wire was soldered at the rear of the coupons. The coupons were polished with a progressively finer sand grinding paper until a final grit size of 600 lm was obtained. After polishing, the coupons were rinsed with distilled water, ultrasonically degreased in acetone and sterilized by exposing to pure ethanol for 24 h. 2.4 Electrochemical Tests The electrochemical measurements were made in a conventional three electrode ASTM cell coupled with a potentiostat and a high frequency impedance analyser (Gamry-600). The electrochemical cells were composed of a test coupon as a working electrode (WE), a graphite electrode as an auxiliary electrode and a saturated calomel electrode (SCE) as a reference electrode. The glasswares were autoclaved at 121 "C for 20 min at 20 psi pressure and then air dried. Graphite electrodes, purging tubes, rubber stoppers and needles were sterilized by immersing in 70 vol.% ethanol for 24 h followed by exposure to a UV lamp for 20 min. Two solutions were used in this experiment. Under a sterilized condition (in a sterilized laminar flow hood), the first cell was prepared with 600 mL of sterilized modified Baar!s growth media (described above) and the second cell was prepared with 600 mL sterilized Baar!s media inoculated with 5 mL of the SRB consortium at 106 cell/mL. The electrochemical cells were purged for 1 h with pure nitrogen gas to establish the anaerobic environment. Electrochemical tests were performed at a flow rate of 0.9 m/s (100 rpm) under atmospheric pressure conditions at 30 "C, to simulate the flow conditions of Saudi Aramco oil production pipeline. Open circuit potential values of the systems were monitored with time during the immersion period followed by periodic readings up to 336 h. Impedance measurements were performed on the system at the OCP for various time intervals from immersion up to 288 h. The frequency sweep was applied from 105 to 10!2 Hz with an AC amplitude of 10 mV. Journal of Materials Engineering and Performance Table 1 The chemical composition of API-5L X52 carbon steel coupons, wt.% Fe C Si Cr Ni Mn Cu Mo Nb Ti Al V S P Bal. 0.070 0.195 0.03 0.02 1.05 0.05 0.004 0.021 0.001 0.029 0.003 0.008 0.008 During the LPR technique, the polarization resistance (Rp) was measured on the system at a scanning amplitude of ±10 mV with reference to the OCP for various time intervals from immersion up to 336 h. The LPR plots were then fitted using the Gamry-600 electrochemical software to obtain the goodness of fit for corrosion rate values. The values were then compared using the mathematical calculation as determined by the equation given below (Ref 23); ! " 1 ba bc ðEq 1Þ icor ¼ Rp 2:3ðba þ bc Þ where ba and bc is the anodic and cathodic slope, respectively. 2.5 Surface Analysis and Corrosion Product Compositions of the Coupons Exposed to SRB At the conclusion of each test, the WEs were carefully removed from the system for fixation. To fix the grown biofilm to the steel surface, the coupons were immersed for 1 h in a 2% glutaraldehyde solution, dehydrated with 4 ethanol solutions (15 min each): 25, 50, 75, and 100% successively, air dried overnight and then gold sputtered (Ref 24). After fixation, the coupons were examined using field emission scanning electron microscopy coupled with EDS to evaluate the morphology and chemical composition of the biofilm. Corrosion product composition was obtained using the x-ray diffraction method with a Philips PW 3040/60 spectrometer using Cu Ka radiation source. The coupons were then cleaned using a standard protocol described under the ASTM-GI-03 (Ref 25), and the pit morphology and density on the exposed coupons were examined using FE-SEM. 3. Results and Discussion 3.1 Identification of the Sulfate-Reducing Consortium 16S rRNA gene sequence analysis indicated that the mixed bacterial culture consortium contained three phylotypes that are close to members of the Proteobacteria (Desulfomicrobium sp.), Firmicutes (Clostridium sp.), and Bacteroidetes (Anaerophaga sp.) (Ref 7). The phylogenetic tree representative of the bacterial consortium has been shown in Fig. 1. Desulfomicrobium, including mesophiles and thermophiles, are known to be commonly associated with oil reservoirs. They have been isolated previously from different oil fields in the North Sea (Ref 26) and in 5 of 6 different Alberta oil fields in Canada as observed by Voordouw and colleagues (Ref 27). Leu et al. (1999) reported the presence of different strains of Desulfomicrobium obtained from various samples in distant oil fields such as formation water and drilling mud (Ref 26). Desulfomicrobium sp. are anaerobic, Gram-negative, rod-shaped, sulfate-reducing bacteria that grow on different carbon source substrates that include lactate, pyruvate, glycerol, and ethanol with optimal growth temperatures between 25 and 35 "C Journal of Materials Engineering and Performance (Ref 27). They are capable of using sulfate, thiosulfate or sulfite as a terminal electron acceptor (Ref 6, 28). Several researchers have further reported that in the presence of 0.5% NaCl, Desulfomicrobium metabolize lactate and sulfate to produce acetate, CO2 and H2S as major end products (Ref 26-28). These microorganisms can play a significant role in oil field reservoir souring by the reductive generation of hydrogen sulfide from these sulfur containing electron acceptors (Ref 28). Clostridium sp. are anaerobic microbes that are Gram-positive, sporeforming bacteria. This type of bacteria is capable of surviving high temperatures due to the formation of heat-resistant endospores (Ref 1, 6, 28). Anaerophaga sp. have also been previously identified in samples from a produced water obtained from the high-temperature Troll oil formation in the North Sea (Ref 28). The cultivation of this consortium was previously reported (Ref 7). 3.2 Surface Morphology and Element Analysis At the conclusion of the tests, the visual inspection of the coupon exposed to the biotic system revealed dense, thick and black products covered on the surface. There is a significant difference in the appearance, structure, and morphology of the corrosion products developed on the steel coupons exposed to the biotic system compared to that exposed to the abiotic system as shown in Fig. 2 and 3. Both the morphological observations and EDS elemental analysis of corrosion products of API X52 steel immersed in the biotic system are shown in Fig. 2(a)-(c). As shown in Fig. 2(a), there are two distinctive areas: A1 and A2. Quantitative EDS analysis shows A2 is composed of a higher amounts of sulfide, sodium salt, phosphate and carbon (Fig. 2(b) and Table 2). The light region (A2) is considered the outer layer where the sulfur is the predominant constituent, Fig. 2(b). On the other hand, the dark region, A1, is considered to be the inner layer in which the iron, oxygen, and carbon, in addition to sulfur and phosphorous, are predominant, as shown in Fig. 2(c) and Table 2. The outer and inner areas were chosen based on our previous research experience, and no cross section view was performed on the sample (Ref 7). The spherical characteristic structures shown in Fig. 2(a) might represent siderite (FeCO3) surrounded by a mixture of iron sulfide (FeS) and iron oxides. Similar features have been reported by Hendrik Venzlaff et al. for steel surfaces exposed to a SRB strain, Desulfopila corrodens. The fact that Desulfomicrobium have the capacity to convert the carbon source (lactate) through pyruvate to acetate along with the production of carbonate supports the formation of FeCO3 (Ref 6, 29). Furthermore, the presence of di-potassium hydrogen orthophosphate and sodium chloride in the growth media might lead to the precipitation of phosphorous-based compounds and sodium chloride on the surface as suggested by the EDS spectra (Ref 30). The morphological observations and elemental analysis of surface deposits and corrosion products on API X52 carbon steel coupons exposed to the abiotic system are shown in Fig. 3(a) and (b). There is one coherent homogenous layer of corrosion product with salt crystal deposits on the surface. As Fig. 1 The phylogenetic characterization of SRB consortium Fig. 2 FESEM and EDS analysis for the system under biotic conditions. (a) FESEM Image of carbon steel exposed to Barr!s medium inoculated with SRB, at 1009. Arrow points to a magnified image (2009) of iron carbonate layer. EDS analysis corresponding to the FESEM outer and inner region shown in Fig. 2(b) and (c), respectively Journal of Materials Engineering and Performance Fig. 3 FESEM and EDS analysis on the API X52 exposed to sterilized Baar!s medium Table 2 Comparative of EDS analysis corresponding to the abiotic and the biotic systems, respectively Element, wt.% Abiotic system Whole region Biotic system Outer region (A2) Inner region (A1) Fig. 4 C O Na Si Fe S Cl Mn Ca P Total 0.01 18.73 4.54 … 68.38 1.54 4.61 2.11 0.08 0.00 100 14.99 10.35 4.20 23.58 8.66 6.96 1.69 0.89 10.06 26.98 55.09 22.43 0.00 0.00 0.00 0.00 0.00 0.00 6.84 8.72 100 100 FESEM image for the biofilm developed on the API X52 exposed to the SRB-containing medium 10009 and 20009 shown in Fig. 3(b) and Table 2, quantitative EDS analysis shows that this layer might be composed of iron oxides mixed with sodium chlorides, calcium and carbon-based compounds that accumulated from the growth medium (Ref 30). As shown in Table 2 and Fig. 2 and 3, the biotic and abiotic systems displayed significant differences in distribution and composition of corrosion products. In the presence of the SRB consortium, there is significant accumulation of sulfur-based compounds and FeCO3, and a substantial amount of corrosion Journal of Materials Engineering and Performance products growing upward is observed, and the interface does not have a rigid appearance, most likely due to the biofilm matrix, nature of which is polysaccharidic and viscoelastic (Fig. 2 and 4), while the corrosion products for the abiotic system exhibits a completely different thin-flat layer with a hard texture (Fig. 3) (Ref 1-4). The biofilm developed in the presence of the SRB consortium together with the produced corrosion products have a heterogeneous morphology and thickness (Fig. 4). The ο 14000 ∇ 12000 ο • ⊗ Intensity (CPS) 10000 8000 6000 4000 2000 ⊗ ∇ • ο ∇ ⊗ + ο ο 0 20 40 60 80 100 2θ (degree) Fig. 5 XRD spectra for the system under biotic conditions for 14 days exposure with SEM image of the examined surface FESEM micrograph (Fig. 4) reveals the presence of the corrosion products, cells, spores, and EPS fibers distributed over the coupon. At the conclusion of the experiment, the steel substrate was hardly visible. It was covered with a porous black layer. A jelly-like substance could be observed among the corrosion products, which was speculated to be biofilm produced EPS (Ref 30, 31). Rod-shaped and round shaped bacteria occupied a significant volume fraction qualitatively, (no quantitative measurements were performed); however, EPS and corrosion products have been reported to occupy 75-95% of biofilm volume, while 5-25% is occupied by the metabolizing cells (Ref 1, 30). Higher concentrations of hydrogen sulfide, phosphate-based compounds, and other potentially biotically-generated, corrosion-influencing compounds are likely to be promoted by SRB metabolism and biofilm formation (Ref 1, 3, 30, 31). Conversely, the nature of corrosion film for the abiotic system exhibits a completely different thin-flat layer with a hard texture (Fig. 3). 3.3 XRD Results The XRD spectra of API X52 steel coupons exposed to the biotic system are displayed in Fig. 5. The XRD pattern confirmed the formation of a significant amount of iron sulfide compounds that include mackinawite (Fe1+xS), other biogenetic iron sulfide (FeS), siderite (FeCO3), and iron (III) oxidehydroxide (FeOOH) (Ref 29, 32, 33). The formation of pyrite and mackinawite films in the presence of SRB is expected as a consequence of their production of sulfide. Due to the very high reactivity of H2S with iron, the mackinawite layer has been shown to form, which is much faster than what would have been expected from the typical kinetics for a precipitation process. The formed mackinawite is not stable and may dissolve depending on the solution saturation level. For the pH ranging between 4 to 7, which is a typical SRB-containing environment, the solution is often supersaturated with respect to iron sulfide, and the layer does not dissolve (Ref 34). The iron sulfide films could protect or deteriorate the surface. Thin protective films are associated with low ferrous ion concentration rates while active films are believed to form in the presence of high ferrous ion concentrations. These protective films have been proposed to fail due to disruption by microbial action, bulky growth, and oxidation (Ref 34). It has been proposed that the corrosion rate for steel can be defined by the rates of iron sulfide layer formation and breakdown (Ref 7). Possibly, the capacity of the bacterial consortium (Desulfomicrobium & Clostridium) to convert the carbon source (lactate) through pyruvate to acetate and produce carbonate resulted in the formation of FeCO3 (Ref 29). Some bacteria of the Clostridium genus (present in the consortium) are capable of fermentative utilization of lactate depending on the pH medium and produced acetate, bicarbonate or propionic acid that may be relevant to our system; however, due to our limitations, such measurements have not been made (Ref 6). It might be possible that the bacterial consortium work in a synergistic way and affect the corrosion of the alloy steel. EDS and XRD results suggested that the following reactions occurred as a result of metabolic activities of the SRB consortium and corrosion behavior of the steel surface. The general chemical and electrochemical reactions can vary according to the surface content and EDS analyses, but generally include a cathodic reaction under natural deaerated conditions (Ref 1, 3, 30, 31, 35): 2H2 O þ 2e! ! H2 þ 2OH! ðEq 2Þ The SRB Desulfomicrobium sp. may use cathodic hydrogen to reduce sulfate to sulfide as follows (Ref 1-4, 31) þ ! ! SO2! 4 þ 9H þ 8e ! HS þ 4H2 O ðEq 3Þ Reaction [3] normally happens at a slow rate without biocatalysis from bacteria, however; in systems that contain microbes, the reaction can be rapid, being enzymatically catalyzed by the hydrogenase enzyme system of Desulfomicrobium sp. Some hydrogen sulfide ions will convert to hydrogen sulfide especially at acidic pH as described (Ref 3, 30, 33) HS! þ Hþ ! H2 S ðEq 4Þ Reaction [4] is rapidly facilitated by the presence of Desulfomicrobium sp. at a pH between 1 and 7. The production of hydrogen sulfide and the oxidation of iron (anodic reaction) lead to the formation of different types of iron sulfide as follows (Ref 30, 31): Fe0 ! Fe2þ þ 2e! ðEq 5Þ Fe2þ þ H2 S ! Fey Sx þ 2Hþ ðEq 6Þ According to Liu et al. (Ref 36), the ferrous ions (Fe2+) produced by the dissolution process react with sulfide metabolized by the bacteria with the subsequent production of different forms of iron sulfide (FeSx). Therefore, it is expected that the structure of the biofilm changes according to the different growth phases of the SRB (Ref 30, 31, 35, 36). The rate determining step (RDS) for the SRB metabolic activities, Reaction [3], is the proton reduction to form hydrogen. The kinetics of this hydrogen reduction are expected to be slow due to the fact that the concentration of protons is extremely low under natural de-aerated conditions. It has been reported that in a de-aerated water environment, the formation Journal of Materials Engineering and Performance Fig. 6 FESEM analysis for the API X52 coupon surface after cleaning for the system under biotic conditions at 2009, 20009 and 40009 of hydrogen on the surface is extremely low because of limited protons (Ref 37). Hþ þ 1e! ! Hads ðEq 7Þ Therefore, when there is not enough hydrogen to drive the metabolic reaction (Reaction [3]) Desulfomicrobium sp. may switch to convert the carbon source (lactate) through pyruvate to acetate with the production of carbonate (Ref 6, 29, 30). When SRB utilize lactate as an electron donor, they produce extra ATPs by the proton motive force and the oxidation of lactate to acetate plus carbon dioxide (Ref 6). The produced hydrogen crosses the cytoplasmic membrane and is oxidized by periplasmic hydrogenase to initiate a proton motive force, which in turn bio-catalyzes Reaction [3] (Ref 4, 6). The prime reactions of this process are as follows: 2CH3 CHOHCOO! þ 2SO2! 4 ! þ ! 4CH3 COO! þ HCO! 3 þ H2 S þ HS þ H þ CO2 ðEq 8Þ Carbonate ions will react with ferrous ions and form insoluble siderite (FeCO3) in addition to iron sulfide (FeS). The net reaction will be as follows (Ref 29): Journal of Materials Engineering and Performance ! 4Fe þ SO2! 4 þ 3HCO3 þ H2 O ! FeS ðs) þ 3FeCO3 ðs) þ 5OH! ; DG& ¼ !86:1 kJ/mol Fe ðEq 9Þ Closer inspection of the API X52 carbon steel electrode in Fig. 6 shows variations in surface roughness. Generalized corrosion was observed in an irregular pattern on the coupon as in Fig. 6(a). The images at higher magnifications in Fig. 6(b) also show preferential etching of selected pearlite grains in a regular, rectangular, and repetitive manner. The attack appears to be in an un-preferred orientation. An inter-granular attack that produces microstructural relief and etching of pearlite microstructures can be observed in Fig. 6(c) and (d). The rectangular attack, found predominantly in the more corroded areas on the surface, may be the result of the galvanic coupling between the semiconductive iron sulfide, which acts as a cathode, and the steel surface, or anode. Galvanic corrosion is either a chemical or an electrochemical corrosion phenomenon. The latter is due to a potential difference between two different surfaces connected through a circuit that allows for current flow to occur from more active metal (anode), to less active surface (cathode). In our case, the iron sulfide is considered the cathode where the calculated equilibrium FESEM analysis for the API X52 coupon surface after cleaning for the system under abiotic conditions exposure at 1009 and 5009 potential of sulfate reduction (SO42!/FeS) is !0.25 V, and the steel surface is the anode with calculated equilibrium potential of !0.44 V (Ref 29). Anodic corrosion of iron by galvanic coupling with iron sulfide has been proposed previously (Ref 37, 38). Galvanic corrosion attacks the iron matrix adjacent to iron sulfide deposits. Furthermore, the conductivity of the iron sulfide augments this coupling between the solution and underlying surface that, in turn, catalyzes the corrosion process. The conductivity of heterogeneous corrosion products composed of FeS and FeCO3 in SRB cultures has been reported to be around 50 Sm!1, which is higher than that of many typical semiconductors such as silicon. This conductivity is mainly due to the presence of FeS as FeCO3, which is considered an insulating mineral (Ref 29). The preferred attacks shown in the pearlite microstructures (Fig. 6c) are attributed to the nature of heterogeneous structures of pearlite. Pearlite consists of plates of cementite (Fe3C) in a matrix of ferrite (Ref 33). The structural and compositional inhomogeneity within the pearlite structure invites the possibility of this preferred attack and enhances the localized corrosion (Ref 39). When pearlite is exposed to localized corrosion, as in our case, galvanic microcells between ferrite (cathode) and cementite (anode) are generated. Consequently, the surface will exhibit deeper and larger pits. However, in a real system where oxygen might exist, the corrosion mechanism becomes quite complicated, and not only the sulfate and sulfide but also several other ionic species (e.g. chlorides) may be involved. On the other hand, minimal corrosion damage was observed in the abiotic system where the polishing marks are still visible (Fig. 7). 3.4 Open Circuit Potential/Polarization Resistance The OCP variations for the biotic and abiotic systems are shown in Fig. 8. The Ecorr as a function of time data revealed that in the biotic system, a substantial shift of Ecorr towards noble values (-600 mV/SCE) occurred for the first 100 h and then remained more or less stable throughout the period of exposure. This potential shift was attributed to the growth of the SRB species, their metabolic activities, and subsequent accumulation of iron sulfide on the surface. SRB attached to the Biotic System Abiotic System -600 OCP (mV) /SCE Fig. 7 -650 -700 -750 0 50 100 150 200 250 300 350 Time (hrs) Fig. 8 OCP variations under biotic and abiotic conditions coupon surface, colonized and reproduced to form a biofilm, and the activity of microbes in this biofilm subsequently altered the electrochemical processes taking place at the steel surface. These alterations include pH changes, H2S production, iron sulfide formation and even EPS production. These factors collectively enhance the reduction capacity of the system and accelerate anodic dissolution (Ref 30, 31, 35, 36). In stark contrast, the abiotic system had a notable increase of the Ecorr, which then remained more or less steady at approximately !690 mV/SCE. This potential shift might be attributed to the accumulation of the growth medium constituents such as organic compounds, potassium, sodium chloride, and phosphorous on the coupon surface (Ref 31). There is a difference in noble direction of approximately 90 mV/SCE between the biotic and abiotic systems. This positive shift in Ecorr is known as ennoblement. Ennoblement has been reported for different alloys exposed to microbes. It is probably the most Journal of Materials Engineering and Performance 3000 70 Biotic System Abiotic System 60 Corrosion Rate (MPY) Rp (Ω.cm2) 2500 2000 1500 1000 500 50 40 30 20 10 0 0 0 (a) Fig. 9 Biotic System Abiotic System 50 100 150 200 250 300 350 Time (hrs) (b) 0 50 100 150 200 250 300 350 Time (hrs) (a) Polarization resistance (Rp) and (b) corrosion rate variations under biotic and abiotic conditions notable phenomenon in many MIC investigations (Ref 1-3). The ennoblement has been attributed to microbial colonization, biofilm formation and the deposition of sulfide, which collectively result in organometallic catalysis and acidification of the electrode surface (Ref 1). The accumulation of corrosion products such as iron sulfide will result in galvanic coupling with the steel surface and may contribute to this ennoblement phenomenon. Ennoblement promotes pitting corrosion, which is more critical for passive alloys (Ref 1-3). The polarization resistance variations for biotic and abiotic systems are shown in Fig. 9(a). The Rp as a function of time data revealed that in the biotic system, a substantial decrease of Rp to 500 X cm2 was followed by another decrease to approximately 250 X cm2 at 350 h. There was a noticeable increase in the Rp to '3000 X cm2 at 100 h. This increase is attributed to the formation of a film of siderite (FeCO3) and mackinawite that provide short-term protection of the steel surface. It has been reported that the mackinawite layer partially protects the steel from corrosion (Ref 7). The substantial drop in the Rp is attributed to the formation of a stable biofilm and conductive iron sulfide mixed layers on the surface. SRB metabolic activities produce hydrogen sulfide and form organic compounds such as an EPS with acids at the metal/biofilm interface (Ref 1, 3, 30). These factors, along with galvanic coupling between the iron sulfide and underlying surface, create an aggressive environment that leads to this substantial decrease of polarization resistance. In contrast, in the abiotic system, Rp decreased to about 1000 X cm2 at the first 50 h followed by an increase to about 1200 X cm2, which then remained more or less steady throughout the experiment. The initial drop in the Rp was due to the corrosion effects of the deposited nutrients (i.e., sulfide and sodium chloride) on the surface. However, when stable deposits and a corrosion film were formed on the surface, it provided protection as indicated by a steady resistance afterwards (Ref 40). The polarization resistance is inversely proportional to the corrosion rate, implying a higher corrosion rate at lower resistance. The corrosion rate plots over time for the biotic and abiotic systems are shown in Fig. 9(b). Journal of Materials Engineering and Performance The corrosion rate for the biotic system increased significantly after 100 h, which is in agreement with the shift of OCP. The corrosion rate reached a maximum value of 45 mpy after 250 h, whereas the corrosion rate for the abiotic system for the same interval is approximately 15 mpy. It should be noted that the utilization of the LPR method in this study should not replace or underestimate the weight loss (WL) method in determining the corrosion rate. The WL is absolutely more reliable as it provides direct proof of corrosion rather than indirect electrochemical measurements (e.g., LPR). Several general statements could be made to explain the high corrosion rates observed under the biotic conditions. The microenvironment at the metal-liquid layer could have been altered by biofilms via attachment of bacteria and deposition of organic by-products such as: EPS, hydrogen sulfide, and iron sulfide (Reaction [3] and [8] and Fig. 4). These changes enhance the kinetics of the corrosion processes at the metal surface, which was reflected in an increase in Ecorr accompanied by a decrease in Rp. Moreover, the accumulation of iron sulfide (Reaction [5] and [8]) on the carbon steel coupons forms a galvanic cell with the adjacent steel surface, resulting in further enhancement of the corrosion process (Ref 1, 3). The observed rectangular pits and preferred pearlite microstructures attacks (Fig. 6) are due to these galvanic couplings. However, it was unclear what portion of an increase of corrosion resulted from the changes induced by the bacterial metabolic activities and what portion was due to other redox couples promoted by the iron sulfide galvanic coupling. 3.5 Electrical Impedance Spectroscopy Results Figure 10(a) displays the Nyquist plots for a carbon steel coupon exposed to the abiotic system. At low frequencies (LF) (Fig. 10a), the magnitude of the capacitive loop represented by the semicircle diameter increased with time. These LF magnitudes represent the change in the charge transfer resistance (Rct) that describes the evolution of the anodic reaction controlled by charge transfer processes (Ref 30, 31, 35, 41). The resistance increased to values around 4000 X cm2. The diagram obtained at 192 and 288 h exhibited a bigger loop diameter, which 6000 Phase Angle (degree) 96 h 4000 -Z" (Ω.cm2) 0h -80 0 hrs 96 hrs 192 hrs 288 hrs 2000 192 h -60 288 h -40 -20 0 0 (a) Fig. 10 0 1500 3000 4500 6000 (b) 2 Z' (Ω.cm ) 10-1 100 101 102 103 104 105 Frequency (Hz) EIS date abiotic system; (a) Nyquist plots and (b) phase angle plots Fig. 11 Circuits model used to fit the EIS data for the abiotic system indicates an increase in charge transfer resistance. Under these conditions, it was found that the corrosion process occurred in the first 50 h, and a steady corrosion rate was observed as the exposure time increased (Fig. 9b). In the abiotic system, the anodic reaction is represented by reaction (5), and the cathodic reaction is shown by reaction (2). The steady corrosion rate is possibly due to the protective effect of a mixed layer of deposits and corrosion products that are possibly composed of sodium chloride, sulfide, potassium, and carbon-based compounds on the electrode surface (Ref 40). The formation of a capacitive layer on the steel surface is confirmed by phase angle spectra (Fig. 10b) that displayed one time constant or one peak at 10 Hz. The equivalent electrical circuit based on the minimum deviation between the measured and fitted data for the abiotic condition is shown in Fig. 11. The circuit includes: • • 10-2 A resistance Rs considered as the solution resistance. Parallel connection of a charge transfer resistance (Rct) for the steel surface and constant phase element (CPE) associated with a double layer capacitance due to the formation of a heterogeneous layer composed of corrosion products along with other compounds deposited from the growth media. In general, a CPE is used instead of capacitor to compensate for the deviation from ideal behavior. The impedance of CPE is defined by the following equation (Ref 29, 41-43): ZCPE ¼ ðCPEÞ!1 ðjxÞ!a ðEq 10Þ When the carbon steel was exposed to the biotic system, the EIS spectra varied significantly with exposure time as shown in Fig. 12(a). The LF magnitude, represented by the semicircle diameter, significantly decreased with time, indicating a decrease in charge transfer resistance (Rct) and a subsequent increase in the corrosion rate as supported by Fig. 9(b). The SRB consortium impacts the corrosion rate by at least three mechanisms, via biofilm formation, reduction of sulfates and production of hydrogen sulfide and subsequent formation of iron sulfide (Ref 30, 31, 35). For the first 48 h, the medium frequency (MF) response presented in the phase diagram in Fig. 12(c) shows one time constant that indicates an activation control process, which is represented by the circuit diagram for the abiotic system shown in Fig. 11. This behavior is attributed to the formation of an unstable conditioning layer based on a mixture of inorganic / organic compounds; it is essentially the time of biofilm formation (Ref 30, 31, 35). However, after biofilm formation in conjunction with the development of iron sulfide and siderite film and when a steady state is reached at 96 h, another time constant shows up in the phase angle spectra in Fig. 12(c). Figure 13(a) shows the circuit model used to fit the EIS data at 96 and 192 h, where the compact biofilm (bf) formed constitutes an additional parallel combination of resistance and capacitance. At 288 h, the compact biofilm layer and iron sulfide layers start developing pores and the corresponding EIS data fit the circuit model shown in Fig. 13(b). The enhancement of the dissolution kinetics of the metallic surface is evidenced by the decrease of the magnitude of charge transfer resistance (Rct) with time as shown in Fig. 12(a) and 13(b). EIS spectra suggest that the formation of an adherent biofilm along with a mixed layer of iron sulfide and siderite established electrochemical cells on the steel surface and subsequently enhanced the corrosion process. EIS results draw general statements about how the corrosion proceeds in the biotic system: • Before 96 h, the MF response showed one time constant (Fig. 12c) that was attributed to the formation of an unstable layer of a mixture of corrosion products, mainly ferric hydroxide and organic compounds. At this stage, the SRB bacteria attached to the surface, assimilated lactate and reduced sulfate to sulfide ions. Biogenic hydrogen sulfide and a subsequent mixed layer of iron sulfide and siderite along with EPS are formed by the precipitation of ferrous ions with sulfide and carbonate ions. Journal of Materials Engineering and Performance 3000 200 0h 96 h 288 hrs 366 hrs 150 288 h -Z'' (Ω*cm2) -Z" (Ω*cm2) 192 h 2000 1000 100 50 0 (a) 0 1000 2000 0 3000 0 Z' (Ω*cm ) 100 150 200 2 Z'(Ω*cm ) -80 Phase Angle (degree) 50 (b) 2 0h 96 h 192 h 288 h -60 -40 -20 0 (c) Fig. 12 10-2 10-1 100 102 103 104 105 Frequency (Hz) EIS date for biotic system; (a), (b) Nyquist plots, and (c) phase angle plots Fig. 13 Circuit models used to fit the EIS data for the biotic system (a) at 96 and 192 h (b) at 288 and 366 h • 101 At 96 h, the mixed layers of EPS and semiconductive corrosion products were stable as evidenced by the phase angle spectra that reveal two time constants (Fig. 12c). At this stage, the microenvironment changes induced by the bacterial metabolic activities in the biofilm and the galvanic coupling between the iron sulfide with the underlying surface increased the corrosion rate significantly, as shown in Fig. 9(b). The galvanic coupling attacks were Journal of Materials Engineering and Performance more pronounced with pearlite microstructures, as represented in Fig. 6, due to their structure and compositional heterogeneity that induced electrochemical potential gradients. • At the final stage, 288 h, with the proliferation of the SRB, production of excess hydrogen sulfide and accumulation of excess corrosion products, the biofilm and iron sulfide film decomposed, cracked and became loose and porous due to the production of polysulfide products and the induced intrinsic physical growth stresses (Ref 7, 32, 35, 44). Subsequently, the steel surface was exposed to the aggressive medium again, which accelerated the corrosion rate significantly ('45 mpy), as shown in Fig. 9(b). 4. Conclusions In this study, MIC of API 5L X52 carbon steel coupons was investigated by exposure of the coupons to a SRB consortium cultivated from produced water from a production sour oil well. 16S rRNA gene sequence analysis indicated that the mixed bacterial culture consortium contained three phylotypes that are close to members of the Proteobacteria (Desulfomicrobium sp.), Firmicutes (Clostridium sp.) and Bacteroidetes (Anaerophaga sp.); all described previously as species associated with MIC. In the presence of a SRB-biofilm, substantial levels of sulfide were detected. XRD revealed the presence of different phases, siderite (FeCO3), iron sulfide (FexSy), and iron oxide (FeOOH) constituents in the corrosion products for the system exposed to the SRB consortium. The microenvironment at the metal-liquid layer became highly altered by the bacterial biofilms via attachment of bacteria, deposition of organic byproducts, (e.g., EPS and the production of hydrogen sulfide) and subsequent development of semiconductive iron sulfide. The metabolic activities resulted in accumulation of a substantial amount of iron sulfide that promoted galvanic coupling with the underlying surface and resulted in general corrosion. Thereafter, the corrosion rate increased dramatically. The corrosion of the steel coupons was significantly more severe in the biotic conditions compared to the abiotic control. The corrosion rate was ' 45 mpy in the biotic system, while it was '15 mpy for the abiotic system. The nature of the corrosion was generalized with preferential etching of select pearlite microstructures in a regular, rectangular, and repeating manner. The galvanic coupling attacks were more marked with pearlite microstructures due to high electrochemical potential gradients. Moreover, the biofilm, as well as iron sulfide altered the kinetic behavior of the system by inducing an extra time constant in the circuit model. Acknowledgments The authors acknowledge and appreciate the Saudi Aramco and Inspection Department Management for their continual support for this project. References 1. B.J. Little and J.S. Lee, Microbiologically Influenced Corrosion, Wiley, Hoboken, NJ, 2007 2. R. Bhola, S.M. Bhola, B. Mishra, and D.L. Olson, Microbiologically Influenced Corrosion and Its Mitigation: A Review, Mater. Sci. Res. India, 2010, 7(2), p 407–412 3. R. Javaherdashti, Microbiologically Influenced Corrosion: An Engineering Insight, Springer, London, 2008 4. W.A. Hamilton, Sulfate-Reducing Bacteria and Anaerobic Corrosion, Annu. Rev. Microbiol., 1985, 39, p 195–217 5. M. Bethencourt, F. Botana, and M. Cano, Biocorrosion of Carbon Steel Alloys by an Hydrogenotrophic Sulfate-Reducing Bacterium Desulfovibrio Capillatus Isolated from a Mexican Oil Field Separator, Corros. Sci., 2006, 48, p 2417–2431 6. M. Madigan, Brock Biology of Microorganisms, 12th ed., Pearson Benjamin Cummings, San Francisco, 2009 7. F.M. AlAbbas, C. Williamson, S.M. Bhola, J.R. Spear, D.L. Olson, B. Mishra, and A.E. Kakpovbia, Influence of Sulfate Reducing Bacterial Biofilm on Corrosion Behavior of Low-Alloy, High-Strength Steel (API-5L X80), Int. Biodeterior. Biodegrad., 2013, 78, p 34–42 8. NACE Standard TM0194-2004, Field Monitoring of Bacterial Growth in Oil and Gas Systems, NACE, Houston, TX, 2004 9. P.J. Antony, R.K. Singh, R. Mohanram, K. Pradeep, and R. Raman, Influence of Thermal Aging on Sulfate-Reducing Bacteria (SRB)Influenced Corrosion Behaviour of 2205 Duplex Stainless Steel, Corros. Sci., 2008, 50, p 1858–1864 10. D.J. Lane, 16S/23S rRNA Sequencing, Nucleic Acid Techniques in Bacterial Systematics, E. Stackebrandt and M. Goodfellow, Ed., Wiley, Chichester, 1991, p 115–175 11. J.W. Sahl, N. Fairfield, J. Kirk Harris, D. Wettergreen, W.C. Stone, and J.R. Spear, Novel Microbial Diversity Retrieved by Autonomous Robotic Exploration of the World!s Deepest Vertical Phreatic Sinkhole, Astrobiology, 2010, 10(2), p 201–213 12. E. Brent and P. Green, Base-Calling of Automated Sequencer Traces Using Phred. II. Error Probabilities, Genome Res., 1998, 8(3), p 186–194 13. E. Brent, L. Hillier, M.C. Wendl, and P. Green, Base-Calling of Automated Sequencer Traces Using Phred. I. Accuracy Assessment, Genome Res., 1998, 8(3), p 175–185 14. D.N. Frank, XplorSeq: A Software Environment for Integrated Management and Phylogenetic Analysis of Metagenomic Sequence Data, BMC Bioinformatics, 2008, 9(1), p 420 15. P. Elmar, J. Peplies, and F. Oliver Glöckner, SINA: Accurate HighThroughput Multiple Sequence Alignment of Ribosomal RNA Genes, Bioinformatics, 2012, 28(14), p 1823–1829 16. P. Elmar, C. Quast, K. Knittel, B.M. Fuchs, W. Ludwig, J. Peplies, and F. Oliver Glöckner, SILVA: A Comprehensive Online Resource for Quality Checked and Aligned Ribosomal RNA Sequence Data Compatible with ARB, Nucleic Acids Res., 2007, 35(21), p 7188–7196 17. W. Ludwig, S. Oliver, W. Ralf, R. Lothar, M. Harald, Yadhukumar, B. Arno, et al. ARB: A Software Environment for Sequence Data, Nucleic Acids Res., 2004, 32(4), p 1363–1371 18. D.A. Benson, K.M. Ilene, D.J. Lipman, J. Ostell, and D.L. Wheeler, GenBank, Nucleic Acids Res., 2005, 33, p D34–D38 19. S.F. Altschul, W. Gish, W. Miller, E.W. Myers, and D.J. Lipman, Basic Local Alignment Search Tool, J. Mol. Biol., 1990, 215(3), p 403–410 20. A. Stamatakis, RAxML-VI-HPC: Maximum Likelihood-Based Phylogenetic Analyses with Thousands of Taxa and Mixed Models, Bioinformatics, 2006, 22(21), p 2688–2690 21. E.P. Nawrocki, Structural RNA Homology Search and Alignment Using Covariance Models, Washington University, Saint Louis, MO, 2009 22. N.D. Pattengale, M. Alipour, O.R.P. Bininda-Emonds, B.M.E. Moret, and A. Stamatakis, How Many Bootstrap Replicates are Necessary?, J. Comput. Biol., 2010, 17(3), p 337–354 23. D.A. Jones and P.S. Amy, A Thermodynamic Interpretation of Microbiologically Influenced, Corrosion, 2002, 8(8), p 938–945 24. C. Xua, Y. Zhanga, B. Chenga, and W. Zhub, Pitting Corrosion Behavior of 316L Stainless Steel in the Media of Sulphate-Reducing and Iron-Oxidizing Bacteria, Mater. Charact., 2008, 59(3), p 245–255 25. ASTM G1-03, Standard Practice for Preparing, Cleaning and Evaluating Corrosion Test Specimens, ASTM, Philadelphia, PA, 2009, p 17–23 26. J.Y. Leu, T.C. McGovern, A.R. Porter, and W.A. Hamilton, The Same Species of Sulphate-Reducing Desulfomicrobium Occur in Different Oil Field Environments in the North Sea, Lett. Appl. Microbiol., 1999, 29, p 246–252 27. G. Voordouw, J.K. Voordouw, and T.R. Jack, Identification of Distinct Communities of Sulfate-Reducing Bacteria in Oil Fields by Reverse Sample Genome Probing, Appl. Environ. Microbiol., 1992, 58, p 3542–3552 28. H. Dahle, F. Garshol, M. Madsen, and N. Birkeland, Microbial Community Structure Analysis of Produced Water from a HighTemperature North Sea Oil-Field, Biomed. Life Sci., 2008, 9, p 37–49 29. D. Enning, H. Venzlaff, J. Garrelfs, H.T. Dinh, V. Meyer, K. Mayrhofer, A.W. Hassel, M. Stratmann, and F. Widdel, Marine Sulfate-Reducing Bacteria Cause Serious Corrosion of Iron Under Electroconductive Biogenic Mineral Crust, Environ. Microbiol., 2012, 14(7), p 1772–1787 30. H. Castaneda and X.D. Benetton, SRB-Biofilm Influenced in Active Corrosion Sites Formed at the Steel-Electrolyte Interface When Exposed to Artificial Seawater Conditions, Corros. Sci., 2008, 50(4), p 1169–1183 31. D. Cetin and L. Aksu, Corrosion Behavior of Low-Alloy Steel in the Presence of Desulfotomaculum sp, Corro. Sci., 2009, 51, p 1584–1588 32. H. Venzlaff, D. Enning, J. Srinivasan, K.J.J. Mayrhofer, A.W. Hassel, H.F. Widdel, and M. Stratmann, Accelerated Cathodic Reaction in Microbial Corrosion of Iron Due to Direct Electron Uptake by Sulphate Reducing Bacteria, Corros. Sci., 2013, 66, p 88–96 33. F.M. AlAbbas, J.R. Spear, A. Kakpovbia, N.M. Balhareth, D.L. Olson, and B. Mishra, Bacterial Attachment to Metal Substrate and Its Effects on Microbiologically-Influenced Corrosion in Transporting Hydrocarbon Pipelines, J. Pipeline Eng., 2012, 2(1), p 63–72 Journal of Materials Engineering and Performance 34. R.G.J. Edyvean, Hydrogen Sulphide—A Corrosive Metabolite, Int. Biodeterior., 1991, 27, p 109–120 35. F. Kuanga, J. Wang, L. Yana, and D. Zhanga, Effects of SulfateReducing Bacteria on the Corrosion Behavior of Carbon Steel, Electrochim. Acta, 2007, 52, p 6084–6088 36. T. Liu, H. Liu, Y. Hu, L. Zhou, and B. Zheng, Growth Characteristics of Thermophile Sulfate-Reducing Bacteria and its Effect on Carbon Steel, Mater. Corros., 2009, 60(3), p 218–224 37. W. Lee, Z. Lewandowski, P.H. Nielsen, and W.A. Hamilton, Role of Sulfate-Reducing Bacteria in Corrosion of Mild Steel: A Review, Biofouling, 1995, 8, p 165–194 38. R.A. King and J.D. Miller, Corrosion by Sulfate Reducing Bacteria, Nature, 1971, 233, p 491–492 39. E. Robert, R. Hill, and R. Abbaschian, Physical Metallurgy Principles, 3rd ed., PWS-Kent Publication, Boston, 1992 Journal of Materials Engineering and Performance 40. ASM Handbook Volume 13 A, Corrosion: Fundamentals, Testing and Protection, ASM International, Materials Park, OH 41. O.R. Monroy, M.H. Gayosso, N. Ordaz, G. Olivares, and C.J. Ramı́rez, Corrosion of API, XL 52 Steel in Presence of Clostridium Celerecrescens, Mater. Corros., 2011, 62(9), p 878–883 42. S.M. Bhola, R. Bhola, B. Mishra, and D.L. Olson, Electrochemical Impedance Spectroscopic Characterization of the Oxide Film Formed over Low Modulus Ti-35.5Nb-7.3Zr-5.7Ta Alloy in Phosphate Buffer Saline at Various Potentials, J. Mater. Sci., 2010, 45(22), p 6179–6186 43. S.M. Bhola, R. Bhola, L. Jain, B. Mishra, and D.L. Olson, Corrosion Behavior of Mild Carbon Steel in Ethanolic Solutions, J. Mater. Eng. Perform., 2011, 20(3), p 409–416 44. J.F.D. Stott, What Progress in the Understanding of Microbially Induced Corrosion Has Been Made in the Last 25 Years? A Personal Viewpoint, Corrosion Sci., 1993, 35(1-4), p 667–673