Next Article in Journal
Influence of Pre-Corrosion in NaCl Solution on Cavitation Resistance of Steel Samples (42CrMo4)
Previous Article in Journal
Recovery of Cu and Fe from a Sphalerite Concentrate by the MnO2–KI Leaching Oxidation System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparative Evaluation of Microbiologically Induced Corrosion Behaviors of 316L Austenitic and 2205 Duplex Stainless Steels Inoculated in Desulfovibrio vulgaris

1
Institute of Advanced Materials and Technology, University of Science and Technology Beijing, Beijing 100083, China
2
Beijing Institute of Machinery and Equipment, Beijing 100854, China
3
Southwest Research Institute of Technology and Engineering, Chongqing 400039, China
4
Nanjing Iron and Steel Co., Ltd., Nanjing 210035, China
5
School of Materials and Engineering, University of Science and Technology Beijing, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Metals 2025, 15(9), 1040; https://doi.org/10.3390/met15091040
Submission received: 24 August 2025 / Revised: 9 September 2025 / Accepted: 17 September 2025 / Published: 19 September 2025
(This article belongs to the Section Corrosion and Protection)

Abstract

Selecting appropriate materials is crucial for mitigating the severe economic and safety challenges posed by microbiologically induced corrosion (MIC) in marine and industrial settings. This study focuses on the MIC behavior of 316L austenitic stainless steel and 2205 duplex stainless steel that is caused by the metabolic activities of D. vulgaris during a life span of 7 days. Cell counts, weight loss, electrochemical measurements, and surface characterization were employed to evaluate the materials’ resistance to MIC. Specifically, 2205 DSS exhibited a 60% lower weight loss (0.02 vs. 0.05 mg/cm2), a 42% lower maximum pit depth (2.11 vs. 3.64 μm), and an orders-of-magnitude lower corrosion current density (0.094 vs. 2.0 μA cm−2) compared to 316L SS, demonstrating its superior resistance to D. vulgaris MIC. XRD and XPS analyses revealed that although FeS formed on both materials, FeS2—a thermodynamically stable deep-sulfidation product—was only present on 316L, indicating a more advanced corrosion stage. The absence of FeS2 on 2205 suggests limited sulfide corrosion progression. These findings confirm the advantage of duplex stainless steel in mitigating D. vulgaris-induced corrosion and provide insights into the selection of materials for MIC-prone environments.

1. Introduction

MIC originates from the metabolic activities of microorganisms that form biofilms on metal surfaces. MIC is a corrosion behavior that is induced by corrosive microbes [1,2,3,4]. It occurs in the marine industry, the oil and gas industry, water utility systems, etc. [5,6,7,8]. Corrosion is one of the key factors limiting the service life and structural reliability of metallic materials, especially in complex service environments such as marine, energy, chemical, and underground facilities. According to relevant statistics, the global economic losses caused by corrosion account for 3–4% of GDP annually, with a significant proportion related to invisible and uncontrollable MIC [9]. As a form of corrosion primarily driven by biological activity, MIC has gained widespread attention from researchers in the interdisciplinary fields of materials, environment, and microbiology due to its characteristics such as strong concealment, rapid development, and high remediation costs.
The essence of MIC is an electrochemical process triggered by the metabolic products of microorganisms in the biofilm formed on the metal surface, with corrosion mechanisms that are highly diversified and systemically coupled. The known microorganisms that induce MIC include sulfate-reducing bacteria (SRB) [10], iron-oxidizing bacteria (IOB) [11], nitrate-reducing bacteria (NRB) [12], and acid-producing bacteria (APB) [13], among others. Among them, SRB is widely considered one of the most corrosion-active microbial groups. In anaerobic environments, SRB uses sulfate as an electron acceptor, reducing it to H2S, which reacts with metal ions to form corrosion products such as FeS, leading to severe localized corrosion and even perforation failure [14].
SRB are anaerobic organisms capable of utilizing sulfate as the final electron acceptor during respiration. In contrast to soluble organic carbon, elemental iron discharges electrons extracellularly, which are subsequently employed in sulfate reduction within the SRB cytoplasm through bio-catalysis, necessitating extracellular electron transfer (EET). Consequently, this form of MIC is referred to as “EET-MIC”, arising from the energy requirements of sessile cells in biofilms that are capable of conducting EET.
In recent years, with the advancement of electrochemical technologies and microbial detection methods, there has been a deeper understanding of the mechanisms of SRB-induced corrosion. Among these, cathodic depolarization theory, the extracellular electron transfer (EET) mechanism, and the localized acidic environment formed by microbial biofilms are considered key factors driving the MIC process [15,16,17]. Furthermore, the biofilm formed by SRB can induce electrochemical microenvironments and oxygen concentration cell effects, exacerbating localized corrosion. Existing studies have shown that SRB-related MIC has become one of the main causes of stainless steel failure in subsea cables, offshore platforms, crude oil transportation systems, and wastewater treatment facilities.
To address the material degradation caused by SRB, researchers have proposed various protective strategies, including the use of efficient biocides, controlling environmental pH and redox potential, and developing antimicrobial coatings [18,19,20,21]. However, traditional biocides are often inefficient against biofilms, have high operational costs, and cause adverse environmental impacts due to the release of disinfectant by-products. Additionally, after treatment, high-speed cleaning agents are required for cleaning and rinsing to remove organic debris [22]. Coatings can slow down the corrosion rate to some extent, but the corrosion problem still persists. Moreover, if the protective layer has defects such as peeling, the corrosion rate of underground pipelines will accelerate further [23]. Therefore, improving the resistance of metal materials to SRB-induced corrosion by focusing on the alloy composition and microstructure of the materials themselves is still considered a solution with higher intrinsic safety and longer service life [24]. In this context, stainless steel materials have attracted attention in the field of MIC protection due to the presence of their passive film, high chromium/nickel content, and good resistance to pitting corrosion.
The 316L austenitic stainless steel, as a traditional general-purpose stainless steel, is widely used in industries such as chemical, marine, and medical equipment due to its excellent machinability and weldability. However, in microbiologically enriched environments, the surface of 316L is prone to being covered by SRB biofilms, which promote localized pitting corrosion and even stress corrosion cracking, revealing its performance limitations in MIC environments [25]. In contrast, 2205 duplex stainless steel, with both austenitic and ferritic structures, not only outperforms 316L in mechanical properties but also exhibits stronger resistance to localized corrosion due to its higher content of Cr, Mo, and N. It has become an important candidate material for various extreme service conditions [26]. A direct, systematic comparison of their MIC resistance is therefore of significant practical importance. It provides engineers and asset managers with a clear, data-driven basis for material selection, balancing performance requirements with economic considerations in projects where failure due to MIC poses serious risks.
Even though previous studies have investigated the corrosion characteristics of these two stainless steels in specific microbial circumstances, there is a lack of systematic comparisons to better understand the MIC behaviors and mechanisms, biofilm characteristics, and surface morphology changes of 316L and 2205 stainless steel under SRB exposure conditions. Based on the above background, this study designed a standard anaerobic circumstance to cultivate the D. vulgaris, weight loss, electrochemical testing (open circuit potential, linear polarization, electrochemical impedance spectroscopy, potential polarization), cell count, confocal laser scanning microscope (CLSM), scanning electron microscope (SEM), and X-ray diffraction (XRD) were used to analyze the corrosion behavior of 316L stainless steel and 2205 duplex stainless steel. This research presents the relationship between alloy content and MIC resistance properties, providing theoretical support for MIC resistance. A key innovation of this work is the systematic comparison of MIC behavior between 316L and 2205 stainless steels through several characterization methods, which enables comprehensive elucidation differences in MIC behaviors. Furthermore, this study offers theoretical support for optimizing material selection in MIC environments, such as marine engineering and oil and gas transportation infrastructures.

2. Materials and Methods

2.1. Materials Preparation

In this study, 316L austenitic stainless steel and 2205 duplex stainless steel (Nanjing Iron and Steel Co., Ltd., Nanjing, China) were utilized for all experiments. The two types of stainless steels were processed into square coupons of 1 cm × 1 cm × 0.3 cm. Table 1 lists the compositions of 316L austenitic stainless steel and 2205 duplex stainless steel.
The square coupons were sequentially polished using 180#, 400#, and 800# sandpapers to achieve consistent surfaces. The deionized water was used for cleaning and then air-dried in a clean environment or oil-free cold air. The epoxy resin was applied to seal the edges and back of the square coupons and electrochemical coupons, which were required to be dried and conserved in a clean environment. After cleaning with 100% isopropanol, the coupons were exposed to ultraviolet light for 20 min to eliminate contamination and ensure sterile conditions.

2.2. Cultural Medium

The ATCC 1249 culture medium shown in Table 2 was used to cultivate the D. vulgaris. The pH of the medium was adjusted to 7.0 ± 0.1, and 100 ppm of L-cysteine was added as a reducing agent. After sterilizing the medium at 121 °C for 20 min in an autoclave (Panasonic MLS-3751L, Kadoma, Japan), it was deoxygenated under continuous nitrogen gas flow (N2) for 45 min to ensure an anaerobic broth.
The broth was inoculated into 200 mL of culture medium (with a total bottle volume of 450 mL), leaving 250 mL of headspace with 2 mL of seeds to maintain a ratio of 1:100. The broth was incubated at 37 °C for 7 days. The incubation time was set to 7 days, a common duration in SRB-MIC studies that is sufficient for stable biofilm formation and initial corrosion product development [3]. All inoculation and transfer procedures were conducted in an anaerobic chamber.

2.3. Cell Counts

After 7 days of immersion in ATCC 1249 broth, 1 mL of broth was extracted, diluted 10 times, and shaken for planktonic cell count. At the same time, the attached cells were scraped from the material surface, diluted 100 times with phosphate-buffered saline (PBS), mixed with a Kylin-Bell VORTEX-5 vortex mixer (Haimen Qilin Bell Instrument Manufacturing Co., Ltd., Nantong, China), and then counted by a hemocytometer, the ZEISS optical microscope (Oberkochen, Germany), at 50 × 20 magnification for sessile cell count.

2.4. Biofilm Conditions

Extensive research has established that the MIC behavior of SRB is strongly related to biofilm development, which acts as a critical determinant factor of the MIC process. Therefore, the biofilms of 316L austenitic stainless steel and 2205 duplex stainless steel were observed by CLSM (Zeiss LSM780, Carl Zeiss, Oberkochen, Germany) and SEM (FEI Quanta 250, Hillsboro, OR, USA).
For CLSM observation, the coupons were rinsed with PBS at pH 7.4 to remove the medium and loosely attached planktonic cells. The cells were stained using the LIVE/DEADTM BacLightTM bacterial viability stain kit (Life Technologies, Grand Island, NY, USA) after the 7-day incubation period. The SYTO-9 dye in the kit stains live cells and dead cells green under fluorescence at excitation wavelengths of 488 nm and 559 nm. After approximately 1 h of incubation, the samples were placed under CLSM to observe the number of cells on the biofilm formed on the sample.
For SEM observation, the coupons were rinsed three times with PBS (15 s each) and then fixed with 2.5% (w/w) glutaraldehyde solution at 10 °C for 8 h. The samples were dehydrated using ethanol at concentrations of 50%, 70%, 80%, 90%, and 95% for 10 min each, and finally, 100% ethanol was used to dehydrate for 30 min after a 7-day immersion. The coupons were then gold-coated for conductivity and observed under SEM.

2.5. Weight Loss

The coupons were accurately weighed before and after the immersion period. After 7 days, MIC products were removed by a Clark solution and dried at 27 °C for 24 h. The dried coupons were measured for weight loss.
The corrosion rate ( V c o r r )   was calculated by Equation (1):
V c o r r = Δ m ρ     A     t
where Δm is the weight loss, ρ is the density (7.98 g/cm3 for 316L austenitic stainless steel and 7.80 g/cm3 for 2205 duplex stainless steel), A is the exposed surface area (1 cm2), and t is the corrosion time. This method quantitatively assesses the corrosion performance of different stainless steels [27].

2.6. Electrochemical Testing

Electrochemical tests were conducted using a three-electrode system: the working electrode was 316L austenitic stainless steel and 2205 duplex stainless steel. The reference electrode was a saturated calomel electrode (SCE, filled with saturated KCl solution). The counter electrode was a platinum plate with a surface area of 1 cm2. The experiments were conducted in 450 mL anaerobic bottles, with each bottle containing 200 mL of deoxygenated D. vulgaris culture medium with 2 mL of seeds.
The electrochemical tests were performed by the electrochemical station (Reference 600 Plus, Gamry, Warminster, PA, USA). Open circuit potential (OCP) was measured until the OCP value became stable. Linear polarization resistance (LPR) was scanned at a rate of 0.1667 mV s−1 in the range of −10 mV to +10 mV vs. OCP each day. EIS was performed at OCP by applying a sinusoidal signal of 10 mV (amplitude) at a frequency ranging from 104 to 10−2 Hz each day. And we used ZView2 software to fit and analyze the obtained data. Potentiodynamic polarization curves were scanned from OCP to −200 mV vs. OCP and from OCP to +200 mV vs. OCP at a rate of 0.1667 mV s−1 once at the end of the 7 d incubation. The corrosion potential (Ecorr), corrosion current density (icorr), and anodic and (absolute) cathodic Tafel slopes (βa and βc) were determined from a Tafel analysis of the polarization curves [8].

2.7. Corrosion Pits Observation

After 7 days of D. vulgaris incubation, the 316L austenitic stainless steel and 2205 duplex stainless steel coupons were retrieved. The SEM was used to observe the corrosion pits on the square coupon surfaces. The coupons were cleaned using a fresh Clarke’s solution to remove biofilms and corrosion products before pit image analysis under SEM. The chemical composition of Clarke’s solution consisted of 50 g of stannous chloride and 20 g of antimony chloride in 1 L of hydrochloric acid solution (specific gravity: 1.19).
The MIC pit depth was measured by CLSM. The profile with the deepest pit was chosen as the representative profile to indicate the severity of the MIC pitting corrosion on 316L austenitic stainless steel and 2205 duplex stainless steel.

2.8. Corrosion Product Analysis

After the 7-day culture experiment, the samples were removed from the anaerobic bottles, dried under nitrogen for about 10 min, and then subjected to X-ray diffraction (XRD; Bruker D8 Discovery model, Bruker AXS GmbH, Karlsruhe, Germany) and X-ray photoelectron spectroscopy (XPS; AXIS Supra, Shimadzu, Kyoto, Japan) to analyze the chemical composition of the corrosion products, with the latter providing more information. The top biofilm of 316L austenitic stainless steel and 2205 duplex stainless steel coupons was removed by a cotton swab to reveal corrosion products underneath. It was then dried under N2 for approximately 10 min before being placed under XRD and XPS.

3. Results and Discussion

3.1. Cell Counting

The planktonic cell in the culture medium also showed slight differences between the two types of stainless steel. In Figure 1, the planktonic cell of the 2205 duplex stainless steel group was lower than that in the 316L stainless steel, with counts of 3.2 × 107 ± 1.2 × 106 cells/mL and 4 × 107 ± 2.3 × 106 cells/mL, respectively (error bars stand for the standard deviations from three independent coupons). A lower cell count typically leads to less corrosion damage, while a higher cell count may indicate more severe corrosion issues for the material.
The sessile cell count (Figure 2) on the 2205 duplex stainless steel was relatively lower (2.5 × 106 ± 5.4 × 105 cells/cm2) than the 316L stainless steel, which showed a higher attachment of D. vulgaris cells (4.0 × 106 ± 8.9 × 105 cells/cm2), which corresponds to the weight loss trends. Error bars stand for the standard deviations from three independent samples. Higher sessile cell count requires more free electrons as nutrients for D. vulgaris metabolism, thereby promoting the rapid growth of D. vulgaris on the 316L surface.

3.2. CLSM

CLSM was applied to evaluate the distribution of live cells on the coupons. From the fluorescence images (Figure 3), it can be observed that the green fluorescent spots on the surface of 2205 stainless steel are sparse, with a discrete distribution and low brightness, indicating limited attachment of D. vulgaris cells on its surface and weak biofilm formation capability. In contrast, the surface of 316L stainless steel exhibits dense fluorescent spots with enhanced brightness, suggesting that the surface of 316L is more prone to D. vulgaris adhesion and proliferation, and the biofilm development is more extensive. Thus, it can be inferred that 2205 stainless steel has an advantage in inhibiting D. vulgaris biofilm formation.

3.3. Weight Loss Analysis

The weight loss results of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days are shown in Figure 4. It can be observed that the corrosion loss of 2205 duplex stainless steel is relatively low, which is related to its higher content of Cr and Mo. These two compositions typically enhance the traditional corrosion resistance property of iron [28,29], which also enhances the MIC resistance property. The results show that the weight loss of 2205 duplex stainless steel is 0.02 mg/cm2, while that of 316L stainless steel is 0.05 ± 0.01 mg/cm2. This indicates that the MIC induced by D. vulgaris on 316L stainless steel is more severe than that on 2205 duplex stainless steel. The trend of weight loss is consistent with sessile cell counts, which is followed by EET-MIC.

3.4. Pit Depths

Morphologies of MIC pits on the 316L stainless steel and 2205 duplex stainless steel after 7 days of incubation with biofilms and corrosion products removed were examined under CLSM, with pit depths of 3.635 μm and 2.113 μm (in Figure 5), respectively. The pit depth 2205 duplex stainless steel was relatively smooth with fewer MIC products, showing better MIC resistance than 316L austenitic stainless steel. This was attributed to higher Cr and Mo content, which forms a relatively stable passive layer to protect the rare material. Furthermore, both austenite and ferrite phases also contribute to the corrosion resistance of the 2205 duplex [30,31].

3.5. Electrochemical Performance Testing

In Figure 6, the OCP of both stainless steels fluctuates over time (the error bars represent the average values calculated from two parallel working electrodes). The overall OCP value of 2205 duplex stainless steel is lower than that of 316L stainless steel, indirectly proving that 2205 duplex stainless steel has a weaker thermodynamic tendency on the MIC process [21].
For MIC, an activated system, open circuit potential (OCP) as a thermodynamic tendency parameter often lacks high regularity and reliability. The kinetic parameters such as linear polarization resistance (LPR) offer higher reliability. In Figure 7, the Rp value of 2205 duplex stainless steel is higher than that of 316L stainless steel (the error bars represent the average values calculated from two parallel working electrodes). The Rp results are completely consistent with other experimental results, establishing a consistent mechanism to illustrate the outstanding MIC resistance of 2205 duplex stainless steel. The lower sessile cell count on 2205 duplex stainless steel (Figure 2) resulted in lower electron consumption to maintain the metabolism of the D. vulgaris requirement (EET-MIC), which directly induced a lower weight loss (Figure 4), higher Rp, and higher charge transfer (Rct) (Table 3). and lower MIC current density (icorr, Table 4).
Electrochemical impedance spectroscopy (EIS) is generally a non-destructive testing method for samples and provides more electrochemical information on the MIC rate [32]. The Nyquist and Bode plots for 2205 duplex stainless steel and 316L stainless steel for 1 day, 3 days, and 7 days are shown in Figure 8.
The equivalent circuit model Rs(Qf(Rf(QdlRct))) was selected based on the physical characteristics of the system. It is a well-established model for MIC studies where a biofilm/corrosion product layer forms on the metal surface. In Table 3, Rs represents the solution resistance; Rf represents the biofilm resistance; Rct represents the charge transfer resistance. The sum of biofilm resistance and charge transfer resistance is used to evaluate the corrosion rate in MIC. In Table 3, CPE represents the constant phase element, which simulates a non-ideal double-layer capacitance. A non-ideal capacitor differs from an ideal capacitor in that it has a time constant with a dispersion effect due to surface inhomogeneity. The impedance of the constant phase element is related to frequency as follows:
Z   =   1 Y 0 ( j ω ) n
where Y0 is the modulus, j is the imaginary unit, ω is the angular frequency, and n is the element exponent.
The EIS fitting circuit model requires inclusion of the biofilm/corrosion product film resistance and charge transfer resistance, as shown in Figure 9. The charge transfer resistance (Rct) for both steels is high (>105 Ω cm2), indicative of their inherent passivity. Rct values in the range of 105–106 Ω·cm2 are well-documented in the literature for stainless steels in various near-neutral environments, even under MIC conditions. For example, Ebeagwu et al. reported Rct values of 2 × 105 Ω·cm2 for 2205 DSS in a simulated marine environment [33]. Li et al. found that the Rct value of 316 SS in SRB medium was 105 Ω·cm2 [34]. However, the Rct value for 2205 DSS is consistently higher than that for 316L SS at each time point, signifying a more effective barrier against charge transfer and thus a lower corrosion rate. Concurrently, the biofilm resistance (Rf) for 2205 increases more significantly over time (from 33 to 173 Ω cm2) compared to 316L (from 34 to 162 Ω cm2). The constant phase element exponents (n1, n2) are less than 1, reflecting the non-ideal capacitive behavior due to surface inhomogeneity caused by biofilm coverage and initial corrosion. The trends are consistent with weight loss, cell counts, open circuit potential, and linear polarization resistance.
Figure 8. Variation in EIS for 2205 duplex stainless steel and 316L stainless steel in 1, 3, and 7 days of incubation in 450 mL anaerobic bottles.
Figure 8. Variation in EIS for 2205 duplex stainless steel and 316L stainless steel in 1, 3, and 7 days of incubation in 450 mL anaerobic bottles.
Metals 15 01040 g008
Figure 9. Equivalent circuit diagram for Figure 8.
Figure 9. Equivalent circuit diagram for Figure 8.
Metals 15 01040 g009
Table 3. Equivalent circuit fitting parameters for Figure 9.
Table 3. Equivalent circuit fitting parameters for Figure 9.
T   d R s
( Ω   c m 2 )
Q f
( Ω 1   c m 2   s n )
n 1 R f
( Ω   c m 2 )
Q d l
( Ω 1   c m 2   s n )
n 2 R c t
( Ω   c m 2 )
2205125.365.39 × 10−50.87336.21 × 10−50.854.23 × 105
322.46.96 × 10−50.83828.35 × 10−50.833.67 × 105
724.667.19 × 10−50.811736.10 × 10−50.824.04 × 105
316L125.815.72 × 10−50.86347.02 × 10−50.833.76 × 105
323.327.63 × 10−50.811076.83 × 10−50.813.16 × 105
723.737.6 × 10−50.801625.69 × 10−50.803.50 × 105
Figure 10 shows the PDP curves at the end of the 7-day incubation. Table 4 lists the electrochemical parameters obtained from Tafel curve fitting, including the corrosion current density (icorr), corrosion potential (Ecorr), and the cathodic (βc) and anodic (βa) Tafel slopes. The icorr values of 2205 duplex stainless steel and 316L stainless steel were 0.094 μA cm−2 and 2.0 μA cm−2, respectively, indicating that the 316L stainless steel had a higher MIC rate, corroborating the weight loss, cell counts, pit depth, OCP, and LPR data.
Table 4. Tafel curve fitting results.
Table 4. Tafel curve fitting results.
icorr (μA cm−2)Ecorr (V) vs. SCEβa (mV/dec)βc (mV/dec)
2205 SS9.4 × 10−2–0.37121–241
316L SS2.0–0.3196–709
Figure 10. Potentiodynamic polarization curves of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Figure 10. Potentiodynamic polarization curves of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Metals 15 01040 g010

3.6. Surface and Cross-Sectional Morphology

SEM observed the biofilm formation on the surface and cross-section of the D. vulgaris specimens, as shown in Figure 11. The surfaces of both samples are covered with loosely structured corrosion products, and their structures are similar, indicating that the two types of steel undergo the same corrosion mechanism.
The 2205 stainless steel shows only a thin layer of surface corrosion, with a clear interface between the corrosion layer and the substrate. No significant delamination or cracking is observed, suggesting that the corrosion has not progressed deeply, and the overall material structure remains intact. In contrast, 316L stainless steel shows significant thickening of the corrosion layer, with some areas exhibiting corrosion cracks and delamination. This proves that 2205 duplex stainless steel has higher reliability in engineering applications under microbiologically corrosive environments.

3.7. Corrosion Products

According to the XRD analysis (Figure 12), the results show that after being immersed in the D. vulgaris environment for 7 days, both 2205 and 316L stainless steels underwent sulfide reactions, generating FeS corrosion products. The 316L stainless steel (red curve) exhibits significant FeS diffraction peaks, indicating more severe sulfide corrosion in the D. vulgaris environment. In contrast, the FeS peaks of 2205 stainless steel (black curve) are weaker, suggesting stronger corrosion resistance and no significant sulfide corrosion. These results demonstrate that the XRD spectra of both stainless steels are similar, indicating that the chemical nature of MIC and the corrosion products generated are not affected by the different steel compositions.
Figure 13 and Table 5 present the high-resolution Fe 2p3/2 XPS spectra and the relative composition distribution of corrosion products of 2205 duplex stainless steel and 316L austenitic stainless steel after 7 days of immersion in the D. vulgaris environment. The fitting results show that the surface of the samples primarily formed corrosion products such as FeS, FeO, Fe3O4, and Fe2O3, with FeS2 further detected on the 316L surface.
FeS in 2205 accounted for 19.88%, higher than 10.65% in 316L, indicating that both stainless steels were affected by D. vulgaris and underwent sulfidation. Li et al. observed a similar shift (713.6 eV) in iron sulfide produced by MIC [34]. Notably, FeS2 was only detected on the 316L surface, with an atomic percentage of 4.28%. FeS2 is a thermodynamically more stable deep-sulfidation product [35,36], typically formed from FeS under sustained anaerobic conditions, and its presence indicates that the corrosion process has entered a later, irreversible stage.
Regarding oxide products, both materials showed the presence of FeO, Fe3O4, and Fe2O3. These oxides represent different stages of the corrosion process: FeO is the primary oxidation product of divalent iron, with a loose, unstable structure, commonly seen in the early stages of corrosion; Fe3O4 is a mixed-valence oxide, formed in the intermediate corrosion stage, with certain electrochemical activity that may accelerate corrosion; Fe2O3, the final oxidation product of trivalent iron, is the most thermodynamically stable [37,38], and typically appears after the complete failure of the passive film, signaling the deepening of corrosion and the failure of surface protection mechanisms. Therefore, an increase in Fe2O3 can be considered an indicator of intensified oxidative corrosion.
Combining the quantitative results, the 2205 sample primarily contains Fe2O3 (42.82%), with relatively low amounts of FeO and Fe3O4. The corrosion reaction mainly remains in the early-to-intermediate stages. In contrast, 316L shows a significant increase in Fe3O4 (38.47%) alongside Fe2O3 (46.60%), indicating its inferior corrosion resistance.

4. Conclusions

This study investigates the corrosion behavior of 2205 duplex stainless steel and 316L austenitic stainless steel in the D. vulgaris environment over a 7-day immersion period. The results demonstrate that 2205 duplex stainless steel exhibits superior resistance to MIC compared to 316L austenitic stainless steel. Specifically, 2205 DSS showed a 60% lower weight loss (0.02 mg/cm2 vs. 0.05 mg/cm2), 42% shallower maximum pit depth (2.11 μm vs. 3.64 μm), and an orders-of-magnitude lower corrosion current density (0.094 μA cm−2 vs. 2.0 μA cm−2) than 316L SS.
The electrochemical and morphological analysis revealed that 2205 stainless steel has not only a lower corrosion rate but also significantly fewer D. vulgaris cells attached to its surface (2.5 × 106 cells/cm2 vs. 4.0 × 106 cells/cm2 for 316L SS) and less significant corrosion product accumulation. In contrast, 316L stainless steel showed higher D. vulgaris cell attachment, more pronounced pitting corrosion, and a higher corrosion rate, indicating its higher susceptibility to MIC.
XRD analysis confirmed the formation of FeS on both samples, with 316L showing more pronounced sulfide corrosion. In addition, XPS results revealed the presence of FeS2 on 316L but not on 2205. As FeS2 is a thermodynamically stable, deep-sulfidation product, its detection on 316L suggests that the corrosion process progressed to a more advanced and irreversible stage, whereas 2205 remained in an early-to-intermediate corrosion phase. This result reinforces that 2205 not only limits microbial adhesion and maintains structural integrity but also suppresses the formation of deep corrosion products under D. vulgaris influence.
Future work will involve long-term immersion experiments (e.g., 30 days or longer) to monitor corrosion evolution. Techniques such as Localized Electrochemical Impedance Spectroscopy (LEIS) and critical pitting temperature (CPT)/potential (CPP) measurements will be employed to gain deeper insights into the localized corrosion initiation and propagation mechanisms under MIC conditions.

Author Contributions

Conceptualization, Z.L., Y.C., and Q.G.; Methodology, Z.L., Y.C., and Q.G.; Software, X.Z., Y.L., and Y.F.; Validation, X.Z., X.L., and J.C.; Formal Analysis, Q.G., Y.L., and J.Y.; Investigation, J.C.; Resources, Z.L. and Y.F.; Data Curation, Q.G., J.C., and J.Y.; Writing—Original Draft, Z.L., Y.L., Y.F., and J.Y.; Writing—Review and Editing, Z.L., Y.C., J.C., and Y.F.; Visualization, X.L., and J.C.; Supervision, Z.L.; Project Administration, Z.L.; Funding Acquisition, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the HDH Cooperation Foundation of Southwest Research Institute of Technology and Engineering (HDHDW59A010301), the Natural Science Foundation of Chongqing, China (No. CSTB2024NSCQ-MSX0976), and the National Natural Science Foundation of China (grant number 52401074).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Authors Jiaxing Cai and Yi Fan were employed by the company Nanjing Iron and Steel Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Muyzer, G.; Stams, A.J.M. The ecology and biotechnology of sulphate-reducing bacteria. Nat. Rev. Microbiol. 2008, 6, 441–454. [Google Scholar] [CrossRef]
  2. Li, Z.; Sun, W.; Zhou, H.; Zhang, M.X.; Fan, Y.Q.; Gu, T.Y.; Wang, F.H.; Xu, D.K. Advanced microbial technologies for in-depth studies of microbiologically influenced corrosion and its mitigation. Corros. Sci. 2025, 256, 113211. [Google Scholar] [CrossRef]
  3. Dong, Y.Z.; Yang, L.L.; Fan, Y.Q.; Liu, D.; Zhou, E.Z.; He, J.J.; Cui, M.M.; Li, Y.X.; Wang, F.H.; Gu, T.Y.; et al. Size effects of conductive nanoparticles in microbiologically influenced corrosion. Corros. Sci. 2025, 255, 113130. [Google Scholar] [CrossRef]
  4. Xu, D.; Jia, R.; Li, Y.; Gu, T. Advances in the treatment of problematic industrial biofilms. World J. Microbiol. Biotechnol. 2017, 33, 97. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Y.F.; Ning, C.Y. Latest research progress of marine microbiological corrosion and bio-fouling, and new approaches of marine anti-corrosion and anti-fouling. Bioact. Mater. 2019, 4, 189–195. [Google Scholar] [CrossRef]
  6. Yazdi, M.; Khan, F.; Abbassi, R.; Quddus, N.; Castaneda-Lopez, H. A review of risk-based decision-making models for MIC in offshore pipelines. Reliab. Eng. Syst. Saf. 2022, 223, 108474. [Google Scholar] [CrossRef]
  7. Jia, R.; Yang, D.Q.; Abd Rahman, H.B.; Gu, T.Y. Laboratory testing of enhanced biocide mitigation of an oilfield biofilm and its microbiologically influenced corrosion of carbon steel in the presence of oilfield chemicals. Int. Biodeterior. Biodegrad. 2017, 125, 116–124. [Google Scholar] [CrossRef]
  8. Jia, R.; Yang, D.Q.; Xu, J.; Xu, D.K.; Gu, T.Y. Microbiologically influenced corrosion of C1018 carbon steel by nitrate reducing Pseudomonas aeruginosa biofilm under organic carbon starvation. Corros. Sci. 2017, 127, 1–9. [Google Scholar] [CrossRef]
  9. Chen, S.Q.; Li, Y.; Cheng, Y.F. Nanopatterning of steel by one-step anodization for anti-adhesion of bacteria. Sci. Rep. 2017, 7, 5326. [Google Scholar] [CrossRef]
  10. Yu, L.; Duan, J.Z.; Zhao, W.; Huang, Y.L.; Hou, B.R. Characteristics of hydrogen evolution and oxidation catalyzed by Desulfovibrio caledoniensis biofilm on pyrolytic graphite electrode. Electrochim. Acta 2011, 56, 9041–9047. [Google Scholar] [CrossRef]
  11. Liu, H.W.; Gu, T.Y.; Zhang, G.A.; Cheng, Y.F.; Wang, H.T.; Liu, H.F. The effect of magneticfield on biomineralization and corrosion behavior of carbon steel induced by iron-oxidizing bacteria. Corros. Sci. 2016, 102, 93–102. [Google Scholar] [CrossRef]
  12. Liu, B.; Fan, E.D.; Jia, J.H.; Du, C.W.; Liu, Z.Y.; Li, X.G. Corrosion mechanism of nitrate reducing bacteria on X80 steel correlated to its intermediate metabolite nitrite. Constr. Build. Mater. 2021, 303, 124454. [Google Scholar] [CrossRef]
  13. Sowards, J.W.; Mansfield, E. Corrosion of copper and steel alloys in a simulated underground storage-tank sump environment containing acid-producing bacteria. Corros. Sci. 2014, 87, 460–471. [Google Scholar] [CrossRef]
  14. Liu, X.Z.; Wang, Y.H.; Song, Y.W.; Liu, W.F.; Zhang, J.; Li, N.N.; Dong, K.H.; Cai, Y.; Han, E.-H. The respective roles of sulfate-reducing bacteria (SRB) and iron-oxidizing bacteria (IOB) in the mixed microbial corrosion process of carbon steel pipelines. Corros. Sci. 2024, 240, 112479. [Google Scholar] [CrossRef]
  15. Gu, T.Y.; Jia, R.; Unsal, T.; Xu, D.K. Toward a better understanding of microbiologically influenced corrosion caused by sulfate reducing bacteria. J. Mater. Sci. Technol. 2019, 35, 631–636. [Google Scholar] [CrossRef]
  16. Sun, D.X.; Wu, M.; Xie, F. Effect of sulfate-reducing bacteria and cathodic potential on stress corrosion cracking of X70 steel in sea-mud simulated solution. Mater. Sci. Eng. A 2018, 721, 135–144. [Google Scholar] [CrossRef]
  17. Enning, D.; Garrelfs, J. Corrosion of iron by sulfate-reducing bacteria: New views of an old problem. Appl. Environ. Microbiol. 2014, 80, 1226–1236. [Google Scholar] [CrossRef]
  18. Rasheed, P.A.; Jabbar, K.A.; Rasool, K.; Pandey, R.P.; Sliem, M.H.; Helal, M.; Samara, A.; Abdullah, A.M.; Mahmoud, K.A. Controlling the biocorrosion of sulfate-reducing bacteria (SRB) on carbon steel using ZnO/chitosan nanocomposite as an eco-friendly biocide. Corros. Sci. 2019, 148, 397–406. [Google Scholar] [CrossRef]
  19. Xue, Y.; Voordouw, G. Control of microbial sulfide production with biocides and nitrate in oil reservoir simulating bioreactors. Front. Microbiol. 2015, 6, 1387. [Google Scholar] [CrossRef]
  20. Gutierrez, O.; Park, D.; Sharma, K.R.; Yuan, Z.G. Effects of long-term pH elevation on the sulfate-reducing and methanogenic activities of anaerobic sewer biofilms. Water Res. 2009, 43, 2549–2557. [Google Scholar] [CrossRef]
  21. Zhang, L.M.; Yan, M.C.; Zhang, S.D.; Zhu, L.Y.; Umoh, A.J.; Ma, A.L.; Zhneg, Y.G.; Wang, J.Q. Significantly enhanced resistance to SRB corrosion via Fe-based amorphous coating designed with high dose corrosion-resistant and antibacterial elements. Corros. Sci. 2020, 164, 108305. [Google Scholar] [CrossRef]
  22. Nguyen, T.; Roddick, F.A.; Fan, L. Biofouling of water treatment membranes: A review of the underlying causes, monitoring techniques and control measures. Membranes 2012, 2, 804–840. [Google Scholar] [CrossRef] [PubMed]
  23. Cui, Y.Y.; Qin, Y.X.; Ding, Q.M.; Gao, Y.N. Study on corrosion behavior of X80 steel under stripping coating by sulfate reducing bacteria. Bmc Biotechnol. 2021, 21, 5. [Google Scholar] [CrossRef] [PubMed]
  24. Song, D.Z.; Zou, J.T.; Sun, L.L.; Zhang, Y.P.; Zhang, J.Y.; Liang, X.H.; Zhang, S.Q.; Li, Y.S.; Li, H.J.; Xi, B.; et al. Enhanced the SRB corrosion resistance of 316L stainless steel via adjusting the addition of Cu and Ce elements. Vacuum 2024, 224, 113183. [Google Scholar] [CrossRef]
  25. Chen, S.; Cheng, Y.F.; Voordouw, G. A comparative study of corrosion of 316L stainless steel in biotic and abiotic sulfide environments. Int. Biodeterior. Biodegrad. 2017, 120, 91–96. [Google Scholar] [CrossRef]
  26. Shen, T.W.; Feng, L.; Shao, Y.X.; Chen, Y.X.; Xia, X.J.; Chen, Y.L.; Zhao, W. Relationship between phase transformation and corrosion performance of 2205 duplex stainless steel fabricated by laser powder bed fusion. Mater. Today Commun. 2025, 44, 111919. [Google Scholar] [CrossRef]
  27. Grengg, C.; Mittermayr, F.; Ukrainczyk, N.; Koraimann, G.; Kienesberger, S.; Dietzel, M. Advances in concrete materials for sewer systems affected by microbial induced concrete corrosion: A review. Water Res. 2018, 134, 341–352. [Google Scholar] [CrossRef]
  28. Lü, X.H.; Zhang, X.X.; Zhao, K.F.; Chen, Z.M.; Li, J.; Wang, C. Study on corrosion resistance mechanism of Cr-containing steel in simulated oilfield produced fluid. Int. J. Electrochem. Sci. 2023, 18, 100154. [Google Scholar] [CrossRef]
  29. Tang, Y.H.; Li, B.; Shi, H.Y.; Guo, Y.X.; Zhang, S.Z.; Zhang, J.S.; Zhang, X.Y.; Liu, R.P. Simultaneous improvement of corrosion and wear resistance of Fe–Mn–Al–C lightweight steels: The role of Cr/Mo. Mater. Charact. 2023, 205, 113274. [Google Scholar] [CrossRef]
  30. Suo, D.H.; Dai, W.; Liu, Y.Y.; Zhang, B.; Zheng, K.K.; Tu, W.R.; Jiang, Y.M.; Li, J.; Sun, Y.T. Optimizing annealing temperature for Duplex Stainless Steel 2205 in acidic NaCl environments according to corrosion resistance. Corros. Sci. 2023, 222, 111374. [Google Scholar] [CrossRef]
  31. Song, X.; Hu, Y.; Yan, Z.J.; Qi, H.P. Corrosion resistance improvement in 6Cr13 martensitic stainless steel via quenching-tempering and partitioning. Mater. Corros. 2023, 74, 544–550. [Google Scholar] [CrossRef]
  32. Cai, X.; Chang, S.K.; Yang, M.M.; Li, S.J.; Wang, C.; Qiao, Y.X.; Zhou, J.; Xue, F. From cast to wire-arc DED: An investigation on NAB alloy MIC resistance. Corros. Sci. 2025, 251, 112910. [Google Scholar] [CrossRef]
  33. Ebeagwu, M.C.; Wei, B.; Cai, Z.; Udoh, I.I.; Xu, J.; Sun, C. Insights into temperature-dependent microbiologically influenced corrosion of 2205 duplex stainless steel induced by Desulfovibrio vulgaris. Electrochim. Acta 2025, 538, 146967. [Google Scholar] [CrossRef]
  34. Li, J.L.; Chen, L.J.; Wei, B.; Wei, B.X.; Xu, J.; Wu, T.Q.; Sun, C. Corrosion characterization in SLM-manufactured and wrought 316 L Stainless Steel induced by Desulfovibrio desulfuricans. Constr. Build. Mater. 2025, 470, 140602. [Google Scholar] [CrossRef]
  35. Son, S.; Hyun, S.P.; Charlet, L.D.; Kwon, K. Thermodynamic stability reversal of iron sulfides at the nanoscale: Insights into the iron sulfide formation in low-temperature aqueous solution. Geochim. Cosmochim. Acta 2022, 338, 220–228. [Google Scholar] [CrossRef]
  36. Thiel, J.; Byrne, J.M.; Kappler, A.; Pester, M. Pyrite formation from FeS and H2S is mediated through microbial redox activity. Proc. Natl. Acad. Sci. USA 2019, 116, 6897–6902. [Google Scholar] [CrossRef] [PubMed]
  37. Kim, Y.S.; Kim, J.G. Corrosion behavior of pipeline carbon steel under different iron oxide deposits in the district heating system. Metals 2017, 7, 182. [Google Scholar] [CrossRef]
  38. Avdeev, Y.G.; Kuznetsov, Y.I. Iron oxide and oxyhydroxide phases formed on steel surfaces and their dissolution in acidic media. Review. Int. J. Corros. Scale Inhib. 2023, 12, 366–409. [Google Scholar]
Figure 1. Planktonic cell count results of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Figure 1. Planktonic cell count results of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Metals 15 01040 g001
Figure 2. Sessile cell count results of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Figure 2. Sessile cell count results of 2205 duplex stainless steel and 316L stainless steel after being immersed in D. vulgaris standard culture medium for 7 days.
Metals 15 01040 g002
Figure 3. Confocal laser microscopy images of 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days: (a) 2205; (b) 316L.
Figure 3. Confocal laser microscopy images of 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days: (a) 2205; (b) 316L.
Metals 15 01040 g003
Figure 4. Weight losses of 2205 duplex stainless steel and 316L stainless steel after 7 days of incubation in 450 mL anaerobic bottles (each error bar represents the range of readings from 3 coupons in the same anaerobic bottle).
Figure 4. Weight losses of 2205 duplex stainless steel and 316L stainless steel after 7 days of incubation in 450 mL anaerobic bottles (each error bar represents the range of readings from 3 coupons in the same anaerobic bottle).
Metals 15 01040 g004
Figure 5. Maximum pit depths on 2205 duplex stainless steel (a) and 316L stainless steel (b) after 7 days of incubation in 450 mL anaerobic bottles.
Figure 5. Maximum pit depths on 2205 duplex stainless steel (a) and 316L stainless steel (b) after 7 days of incubation in 450 mL anaerobic bottles.
Metals 15 01040 g005
Figure 6. Variation in OCP vs. time for 2205 duplex stainless steel and 316L stainless steel during 7 days of incubation in 450 mL anaerobic bottles.
Figure 6. Variation in OCP vs. time for 2205 duplex stainless steel and 316L stainless steel during 7 days of incubation in 450 mL anaerobic bottles.
Metals 15 01040 g006
Figure 7. Variation of Rp with days for 2205 duplex stainless steel and 316L stainless steel for 7 days of incubation in 450 mL anaerobic bottles.
Figure 7. Variation of Rp with days for 2205 duplex stainless steel and 316L stainless steel for 7 days of incubation in 450 mL anaerobic bottles.
Metals 15 01040 g007
Figure 11. Surface morphology and cross-sectional morphology of 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days: (a1,a2) 2205; (b1,b2) 316L.
Figure 11. Surface morphology and cross-sectional morphology of 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days: (a1,a2) 2205; (b1,b2) 316L.
Metals 15 01040 g011
Figure 12. XRD spectra of corrosion products on 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days.
Figure 12. XRD spectra of corrosion products on 2205 and 316L flat samples after being immersed in the D. vulgaris environment for 7 days.
Metals 15 01040 g012
Figure 13. The Fe 2p3/2 XPS spectra of 2205 and 316L after 7 days of immersion in the D. vulgaris environment.
Figure 13. The Fe 2p3/2 XPS spectra of 2205 and 316L after 7 days of immersion in the D. vulgaris environment.
Metals 15 01040 g013
Table 1. Chemical composition of 316L stainless steel and 2205 duplex stainless steel (wt%). (PREN* = 1%Cr + 3.3%Mo + 16%N).
Table 1. Chemical composition of 316L stainless steel and 2205 duplex stainless steel (wt%). (PREN* = 1%Cr + 3.3%Mo + 16%N).
MaterialClSiMnSPCrNiMoNFePREN*
316L0.020.51.40.0030.03517.4512.112.30.03Bal.25.52
22050.030.61.50.0010.02622.305.803.10.16Bal.35.09
Table 2. Composition of ATCC 1249 culture medium.
Table 2. Composition of ATCC 1249 culture medium.
ComponentChemicalAmount
Component IMgSO4∙7H2O4.1 g/L
Sodium citrate5.0 g/L
CaSO4∙2H2O1.0 g/L
NH4Cl1.0 g/L
Distilled water400 mL/L
Component IIK2HPO40.5 g/L
Distilled water200 mL/L
Component IIISodium lactate4.5 mL/L
Yeast extract1.0 g/L
Distilled water400 mL/L
Component IVFe(NH4)2(SO4)21.0 g/L
Table 5. XPS analysis results of the relative composition of key corrosion products in 2205 and 316L after 7 days of immersion in the D. vulgaris environment.
Table 5. XPS analysis results of the relative composition of key corrosion products in 2205 and 316L after 7 days of immersion in the D. vulgaris environment.
Corrosion Product2205316L
Binding Energy (eV)Atom %Binding Energy (eV)Atom %
FeS713.619.88713.610.65
Fe2O3711.642.82711.646.60
Fe3O4708.15.59710.238.47
FeO710.731.71~~
FeS2~~707.34.28
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Z.; Chen, Y.; Guo, Q.; Zhang, X.; Li, X.; Li, Y.; Cai, J.; Fan, Y.; Yang, J. A Comparative Evaluation of Microbiologically Induced Corrosion Behaviors of 316L Austenitic and 2205 Duplex Stainless Steels Inoculated in Desulfovibrio vulgaris. Metals 2025, 15, 1040. https://doi.org/10.3390/met15091040

AMA Style

Li Z, Chen Y, Guo Q, Zhang X, Li X, Li Y, Cai J, Fan Y, Yang J. A Comparative Evaluation of Microbiologically Induced Corrosion Behaviors of 316L Austenitic and 2205 Duplex Stainless Steels Inoculated in Desulfovibrio vulgaris. Metals. 2025; 15(9):1040. https://doi.org/10.3390/met15091040

Chicago/Turabian Style

Li, Zhong, Yuzhou Chen, Qiang Guo, Xiaohu Zhang, Xiaolong Li, Yong Li, Jiaxing Cai, Yi Fan, and Jike Yang. 2025. "A Comparative Evaluation of Microbiologically Induced Corrosion Behaviors of 316L Austenitic and 2205 Duplex Stainless Steels Inoculated in Desulfovibrio vulgaris" Metals 15, no. 9: 1040. https://doi.org/10.3390/met15091040

APA Style

Li, Z., Chen, Y., Guo, Q., Zhang, X., Li, X., Li, Y., Cai, J., Fan, Y., & Yang, J. (2025). A Comparative Evaluation of Microbiologically Induced Corrosion Behaviors of 316L Austenitic and 2205 Duplex Stainless Steels Inoculated in Desulfovibrio vulgaris. Metals, 15(9), 1040. https://doi.org/10.3390/met15091040

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop