Next Article in Journal
Comparison of the Physical Care Burden on Formal Caregivers between Manual Human Care Using a Paper Diaper and Robot-Aided Care in Excretion Care
Previous Article in Journal
Do Higher Transcranial Direct Current Stimulation Doses Lead to Greater Gains in Upper Limb Motor Function in Post-Stroke Patients?
Previous Article in Special Issue
Adsorption of Fluoride onto Acid-Modified Low-Cost Pyrolusite Ore: Adsorption Characteristics and Efficiencies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Manganese-Oxidizing Bacteria (MnOB): Applications, Future Concerns, and Challenges

School of Environmental and Municipal Engineering, Qingdao University of Technology, Qingdao 266520, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2023, 20(2), 1272; https://doi.org/10.3390/ijerph20021272
Submission received: 1 November 2022 / Revised: 5 January 2023 / Accepted: 9 January 2023 / Published: 10 January 2023

Abstract

:
Groundwater serving as a drinking water resource usually contains manganese ions (Mn2+) that exceed drinking standards. Based on the Mn biogeochemical cycle at the hydrosphere scale, bioprocesses consisting of aeration, biofiltration, and disinfection are well known as a cost-effective and environmentally friendly ecotechnology for removing Mn2+. The design of aeration and biofiltration units, which are critical components, is significantly influenced by coexisting iron and ammonia in groundwater; however, there is no unified standard for optimizing bioprocess operation. In addition to the groundwater purification, it was also found that manganese-oxidizing bacteria (MnOB)-derived biogenic Mn oxides (bioMnOx), a by-product, have a low crystallinity and a relatively high specific surface area; the MnOB supplied with Mn2+ can be developed for contaminated water remediation. As a result, according to previous studies, this paper summarized and provided operational suggestions for the removal of Mn2+ from groundwater. This review also anticipated challenges and future concerns, as well as opportunities for bioMnOx applications. These could improve our understanding of the MnOB group and its practical applications.

1. Introduction

The elements on Earth have their own biogeochemical cycle pathways on the biosphere scale, and they may interact with one another [1,2]. These cycles are critical to the water ecosystem [1]. For example, the effective circulation of nitrogen elements from nitrate to nitrogen gas in the hydrosphere can prevent eutrophication. This implies that water or wastewater treatment technologies can be developed using the redox cycles [3].
It is critical to understand the macroscopic manganese (Mn) cycle. The element Mn is abundant in the earth’s crust. It is ranked fifth among metal elements and second only to iron among transition metals [4]. The Mn element has 11 valence states, ranging from −3 to +7; however, only +2, +3, and +4 are naturally present in the biosphere [5]. Dissolved Mn2+ can be found in a variety of water bodies, including surface water (rivers and lakes) [6,7,8], groundwater [9,10], and oceans [11]. Because free or dissolved Mn3+ is extremely unstable, it is easily transformed into Mn2+ or Mn4+ through disproportionation [12], or an oxyhydroxide (MnOOH) is formed. Mn oxides (MnOx) are typically found in sediments and Mn ores, appearing in a complicated form of Mn4+, Mn3+, and Mn2+. The interconversion of Mn2+, Mn3+, and Mn4+, as shown in Figure 1, forms the biogeochemical cycle of Mn at the hydrosphere scale. MnOx is the most powerful natural oxidant in pathway I, with typical redox potentials of 1.23 V [13] and 1.51 V [14], respectively (Equations (1) and (2)). As a result, a redox reaction between MnOx and reducing substances (such as organic matter in sediments) might occur, resulting in the release of Mn2+ into the surrounding aquatic environment. As in pathway II, the dissolved Mn2+ will be re-oxidized into MnOx by manganese-oxidizing bacteria (MnOB) and oxygen. In comparison to oxygen, however, it appears that the MnOB group dominates pathway II due to a greater oxidation rate [15], which is several orders of magnitude higher [16].
MnO2(s) + 4H+ + 2e = Mn2+ + 2H2O     E0 = 1.23 V
MnOOH(s) + 3H+ + e = Mn2+ + 2H2O     E0 = 1.51 V
MnOB have received attention throughout the past 20 years. The number of publications retrieved with the keywords “manganese oxidizing bacteria or bioMnOx” from well-known publishers is increasing. VOSviewer software (version 1.6.18) was used to process Web of Science Core Collection data. The most common topics, as can be seen in Figure 2, are the oxidation and removal of Mn2+, the characterization and use of bioMnOx, the identification of MnOB, and multicopper oxidases. On the one hand, compared to the removal of Mn2+ from groundwater by chemical oxidants such as chlorine, potassium permanganate, and ozone, the oxidation of Mn2+ to Mn3+ and Mn4+ using MnOB without any chemicals offers a low-cost and environmentally friendly method of removing Mn2+ for drinking purposes. However, there are currently no specifications for the design or operation of bioprocesses that can be referenced as a guide, which could result in a number of operational issues, such as excessive energy use. Moreover, in addition to groundwater purification, because the cycle in Figure 1 is regulated by numerous environmental factors [17], organic compounds such as antibiotics [18] and dyes [19] can be absorbed and degraded by biogenic MnOx (bioMnOx). It possesses advantages over chemical approaches for the treatment of contaminated water since it is a natural biosorbent and oxidant. However, the release of Mn2+ has been demonstrated during the degradation of organic compounds, which can lead to secondary pollution or a decrease in reactivity.
Therefore, the properties of MnOB and bioMnOx were investigated in this paper. The bioprocess for removing Mn2+ from groundwater using functional MnOB was evaluated, and its operational suggestions were also concluded. Moreover, this paper suggests a suitable process technique in accordance with the applications of bioMnOx in bioremediation. It is possible to provide an in-depth comprehension of foundation research and engineering construction.

2. The MnOB Group

2.1. The Phylogenetic Diversity of MnOB

Microorganisms that can oxidize Mn2+ into insoluble bioMnOx are known as Mn oxidizing organisms, and these organisms include bacteria, eukaryotes, fungi, and others [20]. As can be observed, the MnOB group has received the greatest attention, most likely because of its significant contribution to Mn2+ oxidation. Although MnOB can be isolated, most of them are still impossible to distinguish through pure culture. This can be attributed to selection differences in the medium [20], spontaneous loss of Mn oxidation ability in the laboratory [21], or to the fact that some MnOB strains do not oxidize Mn2+ alone [22]. Leptothrix, Pseudomonas, Hyphomicrobium, and Bacillus have so far been recognized as typical MnOB genera [15,20,23,24]. According to Hope et al. [25], Leptothrix is a dominant MnOB in Mn removal biofilters. Burger et al. [26] found that only one of four full-scale water supply plants had Leptothrix in its biofilters, whereas the other three plants that tested positive for biological Mn removal did not. Cheng et al. [27] demonstrated that Gallionella and Crenothrix were the dominant MnOB in a pilot-scale biofilter. Yang et al. [28] indicated that in addition to Pseudomonas, Hyphomicrobium, and Bacillus, the MnOB in a biofilter were also dominated by Acinetobacter, Pedomicrobium, Hydrogenophaga, and Microbacterium. This implies that the well-known, typical MnOB may not exist in a Mn-rich environment. Table 1 lists the MnOB group identified in various Mn environments. In addition to the typical MnOB genera mentioned above, there are a number of additional MnOB found in rivers, lakes, oceans, groundwater, and Mn deposits. These bacteria belong to a variety of phyla; isolated MnOB differ from one another even in the same water environment. Furthermore, some MnOB remain unrecognized. According to [22], the majority of the MnOB members in a culture belong to the phylum Nitrospirae but are distantly related to Nitrospira and Leptospirillum.
In general, highly abundant bacteria in the Mn environment are considered MnOB [29,30,31], but it is crucial to clearly identify MnOB from other functional bacteria. This is because the Mn cycle frequently occurs in conjunction with other cycles of substances present in the Mn environment, resulting in the coexistence of functional bacteria. The results indicated that functional bacteria involved with Mn/Fe oxidation–reduction [32], nitrogen transformation [33], and sulfur oxidation [34] can be discovered in the Mn environment at the same time. As previously observed, these functional bacteria are likely to contribute significantly to the overall bacterial community or have a positive relationship to MnOB [31,34]. Therefore, it is essential to accurately identify the MnOB group.
Table 1. The MnOB-related group detected from various Mn environments.
Table 1. The MnOB-related group detected from various Mn environments.
Environmental ConditionsThe MnOB GroupReferences
RiverBacillus, Pseudomonas, Sphingomonas, Hyphomicrobium, Cyanobacteria[31]
Microbacterium, Agromyces, Mycobacterium, Arthrobacter, Pseudomonas, Burkholderiales[35]
Aurantimonas, Rhodobacter, Bacillus, Shewanella[36]
LakeMetallogenium, Leptothrix, Siderocapsa, Naumaniella, Bacillus, Pseudomonas[37]
Bacillus, Pseudomonas, Afipia[35]
WellAcinetobacter sp., Bacillus, Sphingobacterium sp.[38]
Pseudomonas, Burkholderiales[35]
Mn depositsBacillus, Exiguobacterium, Staphylococcus, Brevibacterium, Alcanivorax sp.[39]
Hyphomicrobium, Leptolyngbya[40]
Sphingomonas, Flavobacterium, Janthinobacterium, Acinetobacter[30]
SeawaterBacillus sp. FF-1[41]
Citreicella manganoxidans sp. nov.[42]
Marinobacter manganoxydans MnI7-9[43]
Activated sludgeBrevibacillus[44]
Serratia marcescens[45]
BiofilterPseudoalteromonas sp.[46]
Flavobacterium, Brevundimonas, Stenotrophomonas[47]
Leptothrix, Pseudomonas, Methylibium[18]
Drinking water systemBacillus[35]
Lysinibacillus, Bacillus, Pseudomonas, Brevundimonas[48]

2.2. MnOB Recognition

The identification of MnOB is intended to confirm their oxidation capacity or taxonomic status. After the Mn2+ is oxidized, the insoluble bioMnOx, which appears as black–brown, is often coated on the surface of MnOB [15,49]. As a result, the most direct method of determining oxidation ability is morphological observation. For example, in a biofilter packed with quartz sand and continually fed with Mn2+, MnOB were successfully enhanced when the sand transformed from white to black–brown [18]. The approach can also be used to identify MnOB in a culture medium. However, when Fe2+ and Mn2+ coexist in water or a culture medium, MnOB must be distinguished from iron oxidizing bacteria, as the biogenic iron oxides may coat the bacterial surface and appear similar to bioMnOx. In such cases, the reductive Leucoberbelin blue (LBB) dye, which interacts with Mn oxides to form a blue solution, can be utilized for further chemical detection.
The most common genotypic approach for determining the taxonomic status of MnOB is 16S rRNA gene sequence technology, which includes DNA extraction, polymerase chain reaction (PCR) amplification, and sequencing. The primers used in the PCR should be specific for a particular MnOB genus. However, it is currently challenging to design enough specific primers to detect the numerous MnOB groups. This explains why universal primers are commonly utilized in most investigations. The universal primers allow researchers to explore the most relevant or dominant MnOB communities in a system [19,34], but this is dependent on the sequences deposited in the database.
Moreover, the process by which MnOB oxidize Mn2+ is unclear. It is well accepted that the oxidation of Mn2+ is an extracellular enzymatic reaction employing various multicopper oxidases (MCOs). Several genes have been identified as being involved in the coding of MCOs for Mn2+ oxidation, including mofA in L. discophora SS-1, genes in the ccm operon of P. putida MnB1, mnxA, B, C, D, E, F, and G in P. putida SG-1; cumA in P. putida GB-1 [5]; moxA in Pedomicrobium sp. ACM 3067; mokA in Lysinibacillus strain MK-1; cotA in Bacillus pumilus WH4 [17]; cueO amplified from Escherichia coli [50]; and copA in B. panacihumi MK-8 [51]. Therefore, primers based on these functional proteins can be used to identify MnOB with the same protein gene, but further classification of MnOB still requires the assistance of 16S rRNA gene sequencing.

3. Engineering Application of MnOB for Mn2+ Removal

3.1. The Bioprocess

Groundwater, which serves as a source of drinking water, usually contains Mn2+ because of the anaerobic environment. The Mn2+ concentration should be controlled below 0.10 mg/L or 0.05 mg/L [52]. As illustrated in Figure 1, MnOB can be used to remove Mn2+ from groundwater. Figure 3 shows the bioprocess, in which a sand biofilter was built as the functional unit, for removing Mn2+ from groundwater. This bioprocess is simpler than conventional drinking water purification processes (coagulation, precipitation, filtration, and disinfection) and no chemical oxidants are required. Li et al. [53] estimated that it might save CNY 50 million in construction costs and CNY 12,000 in daily operation and maintenance costs for a 12 × 104 m3/d water supply plant. Given that switching from a non-biological to a biological process can greatly enhance processing capacity and save operational costs by up to 80% [10], this provides a technical transformation strategy for the groundwater treatment facility.

3.2. The Aeration Unit

Because iron (Fe) and Mn are frequently found together in groundwater, and because the problems they cause are comparable and can be removed simultaneously by biofilters, they are discussed together. Bio-oxidation of Fe2+ or Mn2+ requires very little dissolved oxygen (DO). According to Equations (3)–(5), 0.14 mg and 0.29 mg DO are required per mg of Fe2+ and Mn2+, respectively. Therefore, there are no strict limits on the procedure as long as the aeration unit provides appropriate DO. In general, a simple aeration technique, such as falling water aeration, can provide the necessary level of DO. It was found that groundwater with high concentrations of Fe2+ ~ 15 mg/L and Mn2+ ~ 2.0 mg/L could be efficiently treated under a low DO of 4–5 mg/L [54].
4Fe2+ + O2 → 4Fe3+ +2O2−
2Mn2+ + O2 → 2Mn4+ +2O2−
[O2] = 0.14[Fe2+] + 0.29[Mn2+]
High aeration is thought to increase pH, hence accelerating the oxidation of Fe2+ or Mn2+ [55]. It was discovered by Hoyland et al. [56], however, that biofilter columns operating at pH 6.3 and 6.7 began to remove Mn2+ earlier than those operating at higher pH. This is to be expected, given that inorganic carbon is a carbon source for MnOB [22]; yet, high aeration removes more inorganic carbon from groundwater. Excessive aeration appears to be unnecessary because the MnOB is active at both weak and neutral pH conditions. This should be taken into account when designing the aeration unit.
In addition to Fe2+ and Mn2+, ammonia may be present in groundwater. The nitrification process by which nitrifiers convert ammonia to nitrate dominates ammonia removal. As the nitrification process takes 4.57 mg of DO per mg of ammonia, the presence of ammonia pressures the aeration unit; however, the actual demand for DO was lower than the theoretical calculation estimate. This demonstrated that autotrophic anammox bacteria, which do not require DO, contribute to the conversion of ammonia to nitrogen gas and nitrate, accounting for 48.5%, 46.6% [57], and 15.92% [28] of the removal, respectively. This provides an optimization design proposal for the aeration unit. Furthermore, an additional aerator should be installed at the bottom of the filter bed if the aeration unit fails to supply enough DO. This constant aeration may weaken the interception ability of the filter bed, resulting in metal oxide residues in the effluent. To ensure the quality of the water, a second filtration step is required after the biofilters.

3.3. The Biofilter

The biofilter is the central component of the bioprocess, and its start-up and performance are of particular interest. The main drawback of biofilters appears to be their lengthy start-up phase, during which Mn2+ is ineffectively removed. This phase might last for weeks or months. According to previous studies, the start-up time could be shortened by using backwashing sludge or mature biofilter-supporting materials as inocula [30]. Moreover, during the maturation stage of biofilters, the usage of special filtering media with Mn2+ adsorption capacity can ensure the quality of the water [58]. The biological removal of Mn2+ has been widely employed all around the world. It shows that the biofilter can still be operated to remove Mn2+, even at low temperatures of 4 °C [4], 3–4 °C [59,60], and 8–14.8 °C [61].
The biofilters are frequently designed to operate at a lower rate of ~2 m/h at start-up period [54]. Actually, they have a very high treatment load following the start-up period. Štembal et al. [62] observed that at an average Mn2+ concentration of 1.06 mg/L, the biofilter’s filtration rate can reach 24 m/h. Cheng et al. [63] discovered that when fed 8 °C groundwater containing total Fe of 5–10 mg/L and ammonia of 0.9–1.3 mg/L, a pilot-scale biofilter working at 6 m/h could tolerate a maximum Mn2+ concentration of about 10 mg/L. Fe2+ and Mn2+ can be removed simultaneously in a one-stage biofilter, where Fe2+ is removed in the top filter layer and Mn2+ is bio-oxidized and removed in the lower layer. However, the effect of Fe2+ on Mn2+ removal should be considered. This is mostly due to the dissolution of bioMnOx into Mn2+ by Fe2+ in the biofilter bed [64], which needs a thicker filter bed or a longer start-up period to remove Mn2+. It was shown that the start-up time of a one-stage biofilter required more than 6 months at Fe2+ concentrations up to 12 mg/L, compared to 1–3 months at Fe2+ values of 7 mg/L. In newly constructed biofilters, mature sand that is coated with bioMnOx or MnOB is usually dispersed as inoculum on the upper filter layer. The redox interaction between high Fe2+ and Mn oxides should be carefully monitored.

4. The Widespread Application of MnOB-Produced BioMnOx

4.1. Characterization of BioMnOx

BioMnOx is widely used because of its excellent physicochemical properties. Powder X-ray diffraction (XRD) patterns are used to investigate the phase properties. This implies that nearly all bioMnOx has low crystallinity and an amorphous structure, as evidenced by its disordered structure [18,50,65,66,67,68]. When compared to standard JCPDS or PDF cards, these fresh bioMnOx compounds contain one or more characteristic peaks of the model compounds, indicating the precursor of Mn ore minerals [18].
The particle size range of bioMnOx is at the nanoscale [49], but it aggregates to a larger micrometer scale as aging time increases. According to Zhou and Fu [69], the surface area of bioMnOx, which ranges from 98 to 224 m2/g, is frequently greater than that of chemically synthesized MnO2. This is inconsistent with our findings that the bioMnOx produced by a biofilter has a surface area of 39.1 m2/g [18], while the as-prepared MnO2 shows similar values of 35.41 and 39.29 m2/g [70]. The larger the particle size, the smaller the specific surface area. The aggregation of bioMnOx with prolonged aging time is mostly responsible for the reduced surface area in biofilters. Furthermore, crystallinity also increases with aging time, resulting in crystal phase succession [71]. As illustrated in Equations (1) and (2), a larger surface area provides more adsorption sites for pollutants, and a lower crystallinity promotes electron transport in redox processes. That is, using freshly generated bioMnOx for pollution control is preferable.
Elemental valence states can be determined via X-ray photoelectron spectroscopy (XPS). The XPS spectra demonstrate that bioMnOx has multiple valences including Mn2+, Mn3+, and Mn4+. The redox reactions are driven by Mn4+ and Mn3+, as described in Equations (1) and (2). Their high content suggests that more organic compounds can be attacked, taking more electrons, but the content is affected by MnOB types or cultivation conditions [72]. In any case, the low crystallinity, relatively high surface area, and reactivity of bioMnOx allow it to be employed as an adsorbent, oxidant, and catalyst. It can be concluded that the primary applications of bioMnOx in water environment remediation focus on metal adsorption, the decolorization of organic dyes, and the degradation of emerging pollutants, as summarized below.

4.2. Adsorption and Oxidation of Metals

Because of the cation vacancies in the crystal structure, bioMnOx is negatively charged [49]. These negative charges can be compensated for through cation intercalation and sorption, indicating the possibility of heavy metal removal (Figure 4). It is amazing that MnOB can withstand high concentrations of heavy metals. In a culture medium containing 100 mg/L Mn2+, Wan et al. [19] evaluated the heavy metal removal capacity of a MnOB consortium. The removal of Mn2+ was reduced by 0.4%, 87.5%, 7.7%, and 22.4%, respectively, with the addition of Fe3+ of 56 mg/L, Co2+ of 56 mg/L, Ni2+ of 58.7 mg/L, and Zn2+ of 65 mg/L; however, 72.0% of Fe3+, 12.6% of Co2+, 44.1% of Ni2+, and 90.4% of Zn2+ could be removed simultaneously. Meanwhile, Cu2+ did not inhibit Mn2+ removal by MnOB at values ranging from 6.4 mg/L to 96 mg/L until it reached 128 mg/L. Bacterial cell walls, extracellular polymeric sheaths, and bioMnOx can absorb heavy metals, but the capability of the latter is around two orders of magnitude higher than that of the others [73]. To a certain extent, bioMnOx, which has a high capacity for adsorbing heavy metals, can protect MnOB from toxicity. Moreover, it indicates that bioMnOx produced by the Pseudomonas putida strain MnB1 has a seven to eight times higher adsorption capacity for Pb2+, Cd2+, and Zn2+ than abiotic MnOx (birnessite) [74]. In comparison to abiotic MnOx (todorokite), Bacillus sp. WH4-produced bioMnOx has a maximum Cd adsorption capacity that is approximately 2.96 times greater [75]. By modifying the zeolite with bioMnOx, the removal of Pb2+, Cd2+, and Zn2+ may also be improved by 36.4–70.5% [76].
Although arsenic (As) is not a heavy metal, it is frequently examined alongside heavy metals. In aqueous environments, As3+ and As5+ are the two most common forms; however, As3+ exhibits greater metal toxicity. In Figure 4, the oxidation of As3+ to As5+ can reduce the metal toxicity of As. It shows that bioMnOx, whose oxidation rate (k1 = 0.23 min−1) is much higher than that of abiotic MnOx, can achieve quick As3+ oxidation and, subsequently, As5+ adsorption [77,78]. Liu et al. [79] indicated that the tolerant concentrations of bioMnOx for heavy metals As5+ and Mn2+ are 500 mg/L and 120 mg/L, respectively. This explains how biofilters may simultaneously remove Mn and As from groundwater. In a similar way, bioMnOx also transforms the extremely toxic Co2+ [80] and TI+ [81] into less hazardous metal oxides. On the contrary, the oxidation of Cr3+ by bioMnOx, which results in the formation of more mobile Cr6+, increases the toxicity.

4.3. Dye Decolorization

Up to 15% of the dyes used in various industries can be observed in industrial effluents. These dye residues are stable in water and are harmful to living organisms in aqueous environments; thus, they should be removed [82]. Adsorption is believed to be the most tested, environmentally friendly, and effective method [83]. In comparison to chemically produced adsorbents, the synthesis of bioMnOx requires no additional energy input, indicating a low-cost adsorbent. It shows that for each gram of bioMnOx produced from a mixed MnOB consortium, 22 mg of methylene blue and 23.8 mg of crystal violet could be decolorized, respectively [19]. In addition, the bioMnOx produced by Marinobacter sp. MnI7-9 has a surface area of 27.68 m2/g and a maximum adsorption capacity of 115.61 mg/g for indigo carmine [84].
The combination of adsorption and oxidation promotes decolorization, but the pH conditions must be carefully optimized. This is because the redox processes mediated by bioMnOx are highly dependent on solution pH. Equations (1) and (2) demonstrate that the bioMnOx has a higher oxidation capability at lower pHs. The release of Mn2+ from both self-leaching and reduction of the bioMnOx results in secondary contamination and a loss of reactivity under the strongly acidic pH conditions. MnOB can regenerate bioMnOx by the reoxidation of Mn2+ [85], but it is ineffective in strongly acidic environments. Due to the resorption of Mn2+ by bioMnOx, the release of Mn2+ can be decreased or absent under weakly acidic pH circumstances. As the pH of the solution increases, the oxidation capacity decreases significantly, and adsorption could become dominant. However, the catalytic reactions mediated by Mn or oxygen vacancies, as described below, may still be involved in the oxidation of dyes.

4.4. Control of Organic Contaminants

Microorganic contaminants (MOCs) have recently been frequently found in water bodies, indicating possible negative impacts on the ecosystem. These MOCs are stable in water and usually difficult to degrade directly through bioactivity. As a result, physicochemical treatment approaches such as ozonation, advanced oxidation processes (AOPs), and adsorption for degradation or removal have been developed. One of the physicochemical processes is the decomposition of MOCs utilizing bioMnOx, which is accomplished indirectly by MnOB. The use of MnOB appears to be an eco-friendly and cheap alternative. To the best of our knowledge, as illustrated in Figure 5, these MOCs that can be degraded include, but are not limited to, antibiotics [68,72], endocrine disruptors [86,87], and pesticides [88]. It implies that the process mediated by MnOB, which has the advantage of combining adsorption and degradation, has a broad spectrum of performance.
Amorphous bioMnOx is more reactive than commercial MnO2. As an illustration, the data showed that MnO2 barely decomposed carbofuran, whereas MnOB indirectly degraded 90.63% of it under the same conditions [89]. Previous research has shown that the two reactions described by Equations (1) and (2) are commonly acknowledged as MOC degradation mechanisms. In fact, in the liquid phase, the oxygen or Mn vacancies of bioMnOx can catalyze the oxidation of water or oxygen to form reactive oxygen species, functioning similarly to the AOPs (Figure 5). The catalytic capacity increases with worsening crystallinity, which suggests more vacancies. However, the crystallinity of bioMnOx rises as it grows or ages. This could explain why newly formed bioMnOx is more reactive than commercial MnO2. The AOP-like degradation process induced by bioMnOx should be studied further, since it has the potential to contribute significantly to degradation. During the degradation of tetracycline hydrochloride by MnO2 nanomaterials, Pal et al. [90] observed that the efficiency of tetracycline hydrochloride degradation by MnO2 nanoparticles decreased from 66% to 27% and 37% in the presence of the reactive oxygen species scavengers sodium azide and t-butanol, respectively.

5. Further Concerns and Challenges

5.1. Groundwater Purification

The biofiltration process has been developed and engineered for at least 30 years to remove Mn2+ from groundwater. Although Mn2+ is found in groundwater alongside Fe2+, ammonia, or As, simultaneous removal of these contaminants has been successful. Nitrate concentrations can reach 20 mg/L or greater in some regions, and must be decreased to less than 10 mg/L. According to conventional nitrogen removal theory, carbon sources and anoxic conditions are required in a biofilter for nitrate denitrification. The biofilm can reasonably satisfy the anoxic conditions, since the MnOB has fewer DO needs. However, carbon supplies in groundwater are typically limited; adding carbon sources to groundwater complicates the treatment process and, more importantly, the faster growth of the heterotrophic denitrifying bacteria may take up the required space for MnOB. The simultaneous removal of Mn2+ and nitrate, therefore, poses new difficulties for the bioprocess.
Furthermore, the increase of pH of groundwater by over-aeration promotes the chemical oxidation of Mn2+. However, it seems that the excessive aeration is unnecessary because it does not substantially accelerate bio-oxidation. Due to the existence of autotrophic nitrogen removal bacteria, even if ammonia increases the DO consumption, the real demand for DO is lower than that of complete nitrification. As a result, determining how to accurately supply DO to save energy is a challenge.

5.2. Mn2+-Supplied Biofilter Application

A biofilter supplied with Mn2+ can be developed for further engineering applications of MnOB. On the one hand, the quality of water bodies, including surface water and groundwater, is complicated by exogenous emerging contaminants. During the biofiltration process, these pollutants, in concentrations ranging from ng/L to μg/L, may be adsorbed or degraded by bioMnOx. Importantly, because the amount of bioMnOx determines the efficiency, the fed Mn2+ concentrations can be adjusted to ensure it. Regrettably, this has rarely been reported in previous studies, and it should be studied more.
Moreover, bioMnOx is also employed to control high levels of contaminants present in wastewater, such as heavy metals and organic compounds, as illustrated in Figure 4 and Figure 5, but the majority of the relevant study was conducted in flasks. Mn2+ is commonly released during the decomposition of large quantities of organic compounds, resulting in the loss of bioMnOx and reactivity. In addition, during the adsorption procedure, the bioMnOx may become saturated. As mentioned, biofiltration is a feasible solution to these issues because it not only re-oxidizes the released Mn2+ but also maintains the reactivity of bioMnOx by oxidizing the fed Mn2+, allowing for continuous adsorption and degradation in situ, whereas Mn2+ must be completely oxidized to avoid secondary pollution.

6. Conclusions

Mn2+ is commonly found in groundwater at concentrations above the required level for drinking. Hence, it is necessary to remove Mn2+ from groundwater. The MnOB-based bioprocess consisting of aeration, biofiltration, and disinfection has been discussed, and the literature survey indicates that Mn2+ can be effectively removed without any chemical oxidants. The aeration strength and filter bed should be optimally designed, in accordance with the concentrations of ammonia and Fe, for energy saving and start-up purposes. Further research should focus on the nitrate removal from groundwater by this bioprocess. In addition, a Mn2+-supplied biofilter capable of producing bioMnOx can be further developed for water remediation applications such as metal adsorption, dye decolorization, and organic substance degradation.

Author Contributions

Conceptualization, Y.C.; investigation, K.Y.; writing—original draft preparation, Y.B.; writing—review and editing, Y.C. and B.T.; visualization, C.Q.; project administration, X.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Shandong Province, grant number ZR2021ME059.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gustin, M.S.; Bank, M.S.; Bishop, K.; Bowman, K.; Branfireun, B.; Chételat, J.; Eckley, C.S.; Hammerschmidt, C.R.; Lamborg, C.; Lyman, S.; et al. Mercury biogeochemical cycling: A synthesis of recent scientific advances. Sci. Total Environ. 2020, 737, 139619. [Google Scholar] [CrossRef] [PubMed]
  2. Yu, C.; Xie, S.; Song, Z.; Xia, S.; Åström, M.E. Biogeochemical cycling of iron (hydr-)oxides and its impact on organic carbon turnover in coastal wetlands: A global synthesis and perspective. Earth-Sci. Rev. 2021, 218, 103658. [Google Scholar] [CrossRef]
  3. Wang, G.; Hambly, A.C.; Wang, G.; Tang, K.; Andersen, H.R. Engineered manganese redox cycling in anaerobic–aerobic MBBRs for utilisation of biogenic manganese oxides to efficiently remove micropollutants. Chem. Eng. J. 2022, 446, 136998. [Google Scholar] [CrossRef]
  4. Husnain, S.M.; Asim, U.; Yaqub, A.; Shahzad, F.; Abbas, N. Recent trends of MnO2-derived adsorbents for water treatment: A review. New J. Chem. 2020, 44, 612–696. [Google Scholar] [CrossRef]
  5. Brouwers, G.J.; Vijgenboom, E.; Corstjens, P.L.A.M.; De Vrind, J.P.M.; De Jong, E.W. Bacterial Mn2+ Oxidizing Systems and Multicopper Oxidases: An Overview of Mechanisms and Functions. Geomicrobiol. J. 2000, 17, 1–24. [Google Scholar] [CrossRef]
  6. Cheng, Y.; Zhang, S.S.; Huang, T.L.; Cheng, L.J.; Yao, X. Effects of coagulants on the catalytic properties of iron-manganese co-oxide filter films for ammonium and manganese removal from surface water. J. Clean. Prod. 2020, 242, 118494. [Google Scholar] [CrossRef]
  7. Zhang, R.F.; Huang, T.L.; Wen, G.; Tian, X.; Tang, Z.C. Removal of ammonium and manganese from surface water using a MeOx filter system as a pretreatment process. Environ. Technol. 2021. [Google Scholar] [CrossRef] [PubMed]
  8. Xing, X.; Huang, T.; Cheng, Y.; Hu, R.; Wen, G.; Li, K. The simultaneous removal of ammonium and manganese from surface water in South China by manganese co-oxide film. Toxics 2023, 11, 22. [Google Scholar] [CrossRef]
  9. Hou, Q.X.; Zhang, Q.; Huang, G.X.; Liu, C.Y.; Zhang, Y. Elevated manganese concentrations in shallow groundwater of various aquifers in a rapidly urbanized delta, south China. Sci. Total Environ. 2020, 701, 134777. [Google Scholar] [CrossRef]
  10. Tekerlekopoulou, A.G.; Pavlou, S.; Vayenas, D.V. Removal of ammonium, iron and manganese from potable water in biofiltration units: A review. J. Chem. Technol. Biotechnol. 2013, 88, 751–773. [Google Scholar] [CrossRef]
  11. Jones, M.R.; Luther, G.W.; Tebo, B.M. Distribution and concentration of soluble manganese(II), soluble reactive Mn(III)-L, and particulate MnO2 in the Northwest Atlantic Ocean. Mar. Chem. 2020, 226, 103858. [Google Scholar] [CrossRef]
  12. Li, Q.; Xie, L.; Jiang, Y.; Fortner, J.D.; Yu, K.; Liao, P.; Liu, C. Formation and stability of NOM-Mn(III) colloids in aquatic environments. Water Res. 2019, 149, 190–201. [Google Scholar] [CrossRef]
  13. Li, L.; Wei, D.; Wei, G.; Du, Y. Product identification and the mechanisms involved in the transformation of cefazolin by birnessite (δ-MnO2). Chem. Eng. J. 2017, 320, 116–123. [Google Scholar] [CrossRef]
  14. Wang, A.; Wang, H.; Deng, H.; Wang, S.; Shi, W.; Yi, Z.; Qiu, R.; Yan, K. Controllable synthesis of mesoporous manganese oxide microsphere efficient for photo-Fenton-like removal of fluoroquinolone antibiotics. Appl. Catal. B Environ. 2019, 248, 298–308. [Google Scholar] [CrossRef]
  15. Tebo, B.M.; Bargar, J.R.; Clement, B.G.; Dick, G.J.; Murray, K.J.; Parker, D.; Verity, R.; Webb, S.M. Biogenic manganese oxides: Properties and mechanisms of formation. Annu. Rev. Earth Planet. Sci. 2004, 32, 287–328. [Google Scholar] [CrossRef] [Green Version]
  16. Katsoyiannis, I.A.; Zouboulis, A.I. Biological treatment of Mn(II) and Fe(II) containing groundwater: Kinetic considerations and product characterization. Water Res. 2004, 38, 1922–1932. [Google Scholar] [CrossRef] [PubMed]
  17. Tang, W.; Liu, Y.; Gong, J.; Chen, S.; Zeng, X. Analysis of manganese oxidase and its encoding gene in Lysinibacillus strain MK-1. Process Saf. Environ. Prot. 2019, 127, 299–305. [Google Scholar] [CrossRef]
  18. Cai, Y.; He, J.; Zhang, J.; Li, J. Antibiotic contamination control mediated by manganese oxidizing bacteria in a lab-scale biofilter. J. Environ. Sci. 2020, 98, 47–54. [Google Scholar] [CrossRef]
  19. Wan, W.; Xing, Y.; Qin, X.; Li, X.; Liu, S.; Luo, X.; Huang, Q.; Chen, W. A manganese-oxidizing bacterial consortium and its biogenic Mn oxides for dye decolorization and heavy metal adsorption. Chemosphere 2020, 253, 126627. [Google Scholar] [CrossRef] [PubMed]
  20. Nealson, K.H. The Manganese-Oxidizing Bacteria. In The Prokaryotes: Volume 5: Proteobacteria: Alpha and Beta Subclasses; Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K., Stackebrandt, E., Eds.; Springer: New York, NY, USA, 2006; pp. 222–231. ISBN 978-0-387-30745-9. [Google Scholar]
  21. Gregory, E.; Staley, J.T. Widespread distribution of ability to oxidize manganese among freshwater bacteria. Appl. Environ. Microbiol. 1982, 44, 509–511. [Google Scholar] [CrossRef]
  22. Yu, H.; Leadbetter, J.R. Bacterial chemolithoautotrophy via manganese oxidation. Nature 2020, 583, 453–458. [Google Scholar] [CrossRef] [PubMed]
  23. Mouchet, P. From conventional to biological removal of iron and manganese in France. J. Am. Water Works Ass. 1992, 84, 158–167. [Google Scholar] [CrossRef]
  24. Therdkiattikul, N.; Ratpukdi, T.; Kidkhunthod, P.; Chanlek, N.; Siripattanakul-Ratpukdi, S. Manganese-contaminated groundwater treatment by novel bacterial isolates: Kinetic study and mechanism analysis using synchrotron-based techniques. Sci. Rep. 2020, 10, 13391. [Google Scholar] [CrossRef] [PubMed]
  25. Hope, C.K.; Bott, T.R. Laboratory modelling of manganese biofiltration using biofilms of Leptothrix discophora. Water Res. 2004, 38, 1853–1861. [Google Scholar] [CrossRef] [PubMed]
  26. Burger, M.S.; Krentz, C.A.; Mercer, S.S.; Gagnon, G.A. Manganese removal and occurrence of manganese oxidizing bacteria in full-scale biofilters. J. Water Supply Res. Technol.-AQUA 2008, 57, 351–359. [Google Scholar] [CrossRef]
  27. Cheng, Q.F.; Huang, Y.; Nengzi, L.C.; Zhang, J. Performance and microbial community profiles in pilot-scale biofilter for the simultaneous removal of ammonia, iron and manganese at different manganese concentrations. Bioproc. Biosyst. Eng. 2019, 42, 741–752. [Google Scholar] [CrossRef]
  28. Yang, H.; Li, D.; Zeng, H.; Zhang, J. Long-term operation and autotrophic nitrogen conversion process analysis in a biofilter that simultaneously removes Fe, Mn and ammonia from low-temperature groundwater. Chemosphere 2019, 222, 407–414. [Google Scholar] [CrossRef]
  29. Hasan, H.A.; Abdullah, S.R.S.; Kofli, N.T.; Kamarudin, S.K. Effective microbes for simultaneous bio-oxidation of ammonia and manganese in biological aerated filter system. Bioresour. Technol. 2012, 124, 355–363. [Google Scholar] [CrossRef]
  30. Cai, Y.A.; Li, D.; Liang, Y.; Luo, Y.; Zeng, H.; Zhang, J. Effective start-up biofiltration method for Fe, Mn, and ammonia removal and bacterial community analysis. Bioresour. Technol. 2015, 176, 149–155. [Google Scholar] [CrossRef]
  31. Hirano, S.; Ito, Y.; Tanaka, S.; Nagaoka, T.; Oyama, T. Microbial community composition in iron deposits and manganese crusts formed in riverine environments around the Aso area in Japan. Res. Microbiol. 2020, 171, 271–280. [Google Scholar] [CrossRef]
  32. Johnson, D.B.; Pakostova, E. Dissolution of Manganese (IV) Oxide Mediated by Acidophilic Bacteria, and Demonstration That Manganese (IV) Can Act as Both a Direct and Indirect Electron Acceptor for Iron-Reducing Acidithiobacillus spp. Geomicrobiol. J. 2021, 38, 570–576. [Google Scholar] [CrossRef]
  33. Xie, H.; Yang, Y.; Liu, J.; Kang, Y.; Zhang, J.; Hu, Z.; Liang, S. Enhanced triclosan and nutrient removal performance in vertical up-flow constructed wetlands with manganese oxides. Water Res. 2018, 143, 457–466. [Google Scholar] [CrossRef]
  34. Zhao, X.; Liu, B.; Wang, X.; Chen, C.; Ren, N.; Xing, D. Single molecule sequencing reveals response of manganese-oxidizing microbiome to different biofilter media in drinking water systems. Water Res. 2020, 171, 115424. [Google Scholar] [CrossRef]
  35. Marcus, D.N.; Pinto, A.; Anantharaman, K.; Ruberg, S.A.; Kramer, E.L.; Raskin, L.; Dick, G.J. Diverse manganese(II)-oxidizing bacteria are prevalent in drinking water systems. Environ. Microbiol. Rep. 2017, 9, 120–128. [Google Scholar] [CrossRef]
  36. Anderson, C.R.; Davis, R.E.; Bandolin, N.S.; Baptista, A.M.; Tebo, B.M. Analysis of in situ manganese(II) oxidation in the Columbia River and offshore plume: Linking Aurantimonas and the associated microbial community to an active biogeochemical cycle. Environ. Microbiol. 2011, 13, 1561–1576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zakharova, Y.R.; Parfenova, V.V.; Granina, L.Z.; Kravchenko, O.S.; Zemskaya, T.I. Distribution of iron- and manganese-oxidizing bacteria in the bottom sediments of Lake Baikal. Inland Water Biol. 2010, 3, 313–321. [Google Scholar] [CrossRef]
  38. Li, C.; Wang, S.; Du, X.; Cheng, X.; Fu, M.; Hou, N.; Li, D. Immobilization of iron- and manganese-oxidizing bacteria with a biofilm-forming bacterium for the effective removal of iron and manganese from groundwater. Bioresour. Technol. 2016, 220, 76–84. [Google Scholar] [CrossRef] [PubMed]
  39. Sujith, P.P.; Mourya, B.S.; Krishnamurthi, S.; Meena, R.M.; Bharathi, P.A.L. Mobilization of manganese by basalt associated Mn(II)-oxidizing bacteria from the Indian Ridge System. Chemosphere 2014, 95, 486–495. [Google Scholar] [CrossRef]
  40. Park, J.H.; Kim, B.; Chon, C. Characterization of iron and manganese minerals and their associated microbiota in different mine sites to reveal the potential interactions of microbiota with mineral formation. Chemosphere 2018, 191, 245–252. [Google Scholar] [CrossRef] [PubMed]
  41. Wu, J.H.; Kang, F.; Wang, Z.K.; Song, L.; Guan, X.Y.; Zhou, H. Manganese removal and product characteristics of a marine manganese-oxidizing bacterium Bacillus sp. FF-1. Int. Microbiol. 2022, 25, 701–708. [Google Scholar] [CrossRef]
  42. Rajasabapathy, R.; Mohandass, C.; Dastager, S.G.; Liu, Q.; Li, W.; Colaço, A. Citreicella manganoxidans sp. nov., a novel manganese oxidizing bacterium isolated from a shallow water hydrothermal vent in Espalamaca (Azores). Antonie Leeuwenhoek 2015, 108, 1433–1439. [Google Scholar] [CrossRef]
  43. Wang, H.; Li, H.; Shao, Z.; Liao, S.; Johnstone, L.; Rensing, C.; Wang, G. Genome Sequence of Deep-Sea Manganese-Oxidizing Bacterium Marinobacter manganoxydans MnI7-9. J. Bacteriol. 2012, 194, 899–900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhao, X.; Wang, X.; Liu, B.; Xie, G.; Xing, D. Characterization of manganese oxidation by Brevibacillus at different ecological conditions. Chemosphere 2018, 205, 553–558. [Google Scholar] [CrossRef] [PubMed]
  45. Queiroz, P.S.; Barboza, N.R.; Cordeiro, M.M.; Leão, V.A.; Guerra-Sá, R. Rich growth medium promotes an increased on Mn(II) removal and manganese oxide production by Serratia marcescens strains isolates from wastewater. Biochem. Eng. J. 2018, 140, 148–156. [Google Scholar] [CrossRef]
  46. Subari, F.; Kamaruzzaman, M.A.; Abdullah, S.R.S.; Hasan, H.A.; Othman, A.R. Simultaneous removal of ammonium and manganese in slow sand biofilter (SSB) by naturally grown bacteria from lake water and its diverse microbial community. J. Environ. Chem. Eng. 2018, 6, 6351–6358. [Google Scholar] [CrossRef]
  47. Hou, D.; Zhang, P.; Wei, D.; Zhang, J.; Yan, B.; Cao, L.; Zhou, Y.; Luo, L. Simultaneous removal of iron and manganese from acid mine drainage by acclimated bacteria. J. Hazard. Mater. 2020, 396, 122631. [Google Scholar] [CrossRef]
  48. Cerrato, J.M.; Falkinham, J.O.; Dietrich, A.M.; Knocke, W.R.; Mckinney, C.W.; Pruden, A. Manganese-oxidizing and -reducing microorganisms isolated from biofilms in chlorinated drinking water systems. Water Res. 2010, 44, 3935–3945. [Google Scholar] [CrossRef]
  49. Li, H.; Wu, Y.; Tang, Y.; Fang, B.; Luo, P.; Yang, L.; Jiang, Q. A manganese-oxidizing bacterium-Enterobacter hormaechei strain DS02Eh01: Capabilities of Mn(II) immobilization, plant growth promotion and biofilm formation. Environ. Pollut. 2022, 309, 119775. [Google Scholar] [CrossRef]
  50. Su, J.; Deng, L.; Huang, L.; Guo, S.; Liu, F.; He, J. Catalytic oxidation of manganese(II) by multicopper oxidase CueO and characterization of the biogenic Mn oxide. Water Res. 2014, 56, 304–313. [Google Scholar] [CrossRef]
  51. Zeng, X.; Zhang, M.; Liu, Y.; Tang, W. Manganese(II) oxidation by the multi-copper oxidase CopA from Brevibacillus panacihumi MK-8. Enzyme Microb. Technol. 2018, 117, 79–83. [Google Scholar] [CrossRef]
  52. Patil, D.S.; Chavan, S.M.; Oubagaranadin, J.U.K. A review of technologies for manganese removal from wastewaters. J. Environ. Chem. Eng. 2016, 4, 468–487. [Google Scholar] [CrossRef]
  53. Li, D.; Zhang, L.; Wang, H.; Yang, H.; Wang, B. Operational performance of biological treatment plant for iron and manganese removal. J. Water Supply Res. Technol. 2005, 54, 15–24. [Google Scholar] [CrossRef]
  54. Zeng, H. Biological Purification of Iron, Manganese and Ammonia with High Concentration in Groundwater and Engineering Application. Ph.D. Thesis, Harbin Institute of Technology, Harbin, China, 2010. Available online: https://kns.cnki.net/KCMS/detail/detail.aspx?dbname=CDFD0911&filename=2011013022.nh (accessed on 16 March 2011).
  55. Pacini, V.A.; Ingallinella, A.M.; Sanguinetti, G. Removal of iron and manganese using biological roughing up flow filtration technology. Water Res. 2005, 39, 4463–4475. [Google Scholar] [CrossRef] [PubMed]
  56. Hoyland, V.W.; Knocke, W.R.; Falkinham, J.O.; Pruden, A.; Singh, G. Effect of drinking water treatment process parameters on biological removal of manganese from surface water. Water Res. 2014, 66, 31–39. [Google Scholar] [CrossRef] [PubMed]
  57. Cai, Y.A.; Li, D.; Liang, Y.; Zeng, H.; Zhang, J. Autotrophic nitrogen removal process in a potable water treatment biofilter that simultaneously removes Mn and NH4+-N. Bioresour. Technol. 2014, 172, 226–231. [Google Scholar] [CrossRef]
  58. Piispanen, J.K.; Sallanko, J.T. Mn(II) removal from groundwater with manganese oxide-coated filter media. J. Environ. Sci. Health Part A 2010, 45, 1732–1740. [Google Scholar] [CrossRef]
  59. Cai, Y.A.; Li, D.; Liang, Y.; Zeng, H.; Zhang, J. Operational parameters required for the start-up process of a biofilter to remove Fe, Mn, and NH3-N from low-temperature groundwater. Desalin. Water Treat. 2016, 57, 3588–3596. [Google Scholar] [CrossRef]
  60. Wand, L.; Li, D.; Zeng, H. Simultaneous purification of high-iron, manganese and ammonia nitrogen from low temperature groundwater in biological filter. China Environ. Sci. 2019, 39, 3300–3307. [Google Scholar]
  61. Dangeti, S.; Roshani, B.; Rindall, B.; Mcbeth, J.M.; Chang, W. Biofiltration field study for cold Fe(II)- and Mn(II)-rich groundwater: Accelerated Mn(II) removal kinetics and cold-adapted Mn(II)-oxidizing microbial populations. Water Qual. Res. J. 2017, 52, 229–242. [Google Scholar] [CrossRef]
  62. Štembal, T.; Markić, M.; Ribičić, N.; Briški, F.; Sipos, L. Removal of ammonia, iron and manganese from groundwaters of northern Croatia—Pilot plant studies. Process Biochem. 2005, 40, 327–335. [Google Scholar] [CrossRef]
  63. Cheng, Q.; Li, D.; Li, X.; Ren, Y.; Zhang, J. The maximum manganese concentration of groundwater containing high concentration of iron, manganese and nitrogen. J. Harbin Inst. Technol. 2014, 46, 20–24. [Google Scholar]
  64. Zeng, H.; Li, D.; Gao, Y.; Zhao, Y.; Li, C.; Zhang, J. Redox relationship between Fe and Mn in biological filter layer for iron and manganese removal. China Water Wastewater 2010, 26, 86–88. [Google Scholar]
  65. Du, Z.; Li, K.; Zhou, S.; Liu, X.; Yu, Y.; Zhang, Y.; He, Y.; Zhang, Y. Degradation of ofloxacin with heterogeneous photo-Fenton catalyzed by biogenic Fe-Mn oxides. Chem. Eng. J. 2020, 380, 122427. [Google Scholar] [CrossRef]
  66. Tian, N.; Tian, X.; Nie, Y.; Yang, C.; Zhou, Z.; Li, Y. Biogenic manganese oxide: An efficient peroxymonosulfate activation catalyst for tetracycline and phenol degradation in water. Chem. Eng. J. 2018, 352, 469–476. [Google Scholar] [CrossRef]
  67. Comert, S.; Tepe, O. Production and characterization of biogenic manganese oxides by manganese-adapted Pseudomonas putida NRRL B-14878. Geomicrobiol. J. 2020, 37, 753–763. [Google Scholar] [CrossRef]
  68. Zhou, N.; Liu, D.; Min, D.; Cheng, L.; Huang, X.; Tian, L.; Li, D.; Yu, H. Continuous degradation of ciprofloxacin in a manganese redox cycling system driven by Pseudomonas putida MnB-1. Chemosphere 2018, 211, 345–351. [Google Scholar] [CrossRef] [PubMed]
  69. Zhou, H.; Fu, C. Manganese-oxidizing microbes and biogenic manganese oxides: Characterization, Mn(II) oxidation mechanism and environmental relevance. Rev. Environ. Sci. Biotechnol. 2020, 19, 489–507. [Google Scholar] [CrossRef]
  70. Cai, Y.; He, J. Degradation of ciprofloxacin by the Mn cycle system (MnCS): Construction, characterization and bacterial analysis. Environ. Res. 2021, 195, 110860. [Google Scholar] [CrossRef]
  71. Tang, Y.; Webb, S.M.; Estes, E.R.; Hansel, C.M. Chromium(iii) oxidation by biogenic manganese oxides with varying structural ripening. Environ. Sci. Process. Impacts 2014, 16, 2127–2136. [Google Scholar] [CrossRef]
  72. Tu, J.; Yang, Z.; Hu, C.; Qu, J. Characterization and reactivity of biogenic manganese oxides for ciprofloxacin oxidation. J. Environ. Sci. 2014, 26, 1154–1161. [Google Scholar] [CrossRef]
  73. Boonfueng, T.; Axe, L.; Yee, N.; Hahn, D.; Ndiba, P.K. Zn sorption mechanisms onto sheathed Leptothrix discophora and the impact of the nanoparticulate biogenic Mn oxide coating. J. Colloid Interf. Sci. 2009, 333, 439–447. [Google Scholar] [CrossRef] [PubMed]
  74. Zhou, D.; Kim, D.; Ko, S. Heavy metal adsorption with biogenic manganese oxides generated by Pseudomonas putida strain MnB1. J. Ind. Eng. Chem. 2015, 24, 132–139. [Google Scholar] [CrossRef]
  75. Meng, Y.; Zheng, Y.; Zhang, L.; He, J. Biogenic Mn oxides for effective adsorption of Cd from aquatic environment. Environ. Pollut. 2009, 157, 2577–2583. [Google Scholar] [CrossRef] [PubMed]
  76. Kim, D.; Nhung, T.T.; Ko, S. Enhanced adsorption of heavy metals with biogenic manganese oxide immobilized on zeolite. KSCE J. Civ. Eng. 2016, 20, 2189–2196. [Google Scholar] [CrossRef]
  77. Wang, Y.; Tsang, Y.F.; Wang, H.; Sun, Y.; Song, Y.; Pan, X.; Luo, S. Effective stabilization of arsenic in contaminated soils with biogenic manganese oxide (BMO) materials. Environ. Pollut. 2020, 258, 113481. [Google Scholar] [CrossRef]
  78. Liang, G.; Yang, Y.; Wu, S.; Jiang, Y.; Xu, Y. The generation of biogenic manganese oxides and its application in the removal of As(III) in groundwater. Environ. Sci. Pollut. Res. 2017, 24, 17935–17944. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, M.B.; Wang, S.L.; Yang, M.; Ning, X.; Nan, Z.R. Experimental study on treatment of heavy metal-contaminated soil by manganese-oxidizing bacteria. Environ. Sci. Pollut. Res. 2022, 29, 5526–5540. [Google Scholar] [CrossRef]
  80. Aoshima, M.; Tani, Y.; Fujita, R.; Tanaka, K.; Miyata, N.; Umezawa, K. Simultaneous sequestration of Co2+ and Mn2+ by fungal manganese oxide through asbolane formation. Minerals 2022, 12, 358. [Google Scholar] [CrossRef]
  81. Zhang, L.; Yang, Y.; Wu, S.; Xia, F.; Han, X.; Xu, X.; Deng, S.; Jiang, Y. Insights into the synergistic removal mechanisms of thallium(I) by biogenic manganese oxides in a wide pH range. Sci. Total Environ. 2022, 831, 154865. [Google Scholar] [CrossRef]
  82. Moon, S.A.; Salunke, B.K.; Saha, P.; Deshmukh, A.R.; Kim, B.S. Comparison of dye degradation potential of biosynthesized copper oxide, manganese dioxide, and silver nanoparticles using Kalopanax pictus plant extract. Korean J. Chem. Eng. 2018, 35, 702–708. [Google Scholar] [CrossRef]
  83. Su, P.; Wan, Q.; Yang, Y.; Shu, J.; Zhao, H.; Meng, W.; Li, B.; Chen, M.; Liu, Z.; Liu, R. Hydroxylation of electrolytic manganese anode slime with EDTA-2Na and its adsorption of methylene blue. Sep. Purif. Technol. 2022, 278, 119526. [Google Scholar] [CrossRef]
  84. Chen, X.; Pei, Y.J.; Wang, H.; Wang, G.J.; Liao, S.J. Removal of indigo carmine by bacterial biogenic Mn oxides. Adv. Mater. Res. 2013, 864–867, 1779–1783. [Google Scholar] [CrossRef]
  85. Sun, Y.; Zhang, Y.; Li, W.; Zhang, W.; Xu, Z.; Dai, M.; Zhao, G. Combination of the endophytic manganese-oxidizing bacterium Pantoea eucrina SS01 and biogenic Mn oxides: An efficient and sustainable complex in degradation and detoxification of malachite green. Chemosphere 2021, 280, 130785. [Google Scholar] [CrossRef] [PubMed]
  86. Tran, T.N.; Kim, D.G.; Ko, S.O. Efficient removal of 17 alpha-ethinylestradiol from secondary wastewater treatment effluent by a biofilm process incorporating biogenic manganese oxide and Pseudomonas putida strain MnB1. J. Hazard. Mater. 2020, 398, 122810. [Google Scholar] [CrossRef] [PubMed]
  87. Lin, Y.; Liu, H.C.; Wang, X.J. Removal effects and potential mechanisms of bisphenol A and 17 alpha-ethynylestradiol by Biogenic Mn oxides generated by Bacillus sp. WH4. Environ. Sci. Pollut. Res. 2022, 29, 57261–57276. [Google Scholar] [CrossRef] [PubMed]
  88. Luo, J.; Ruan, X.; Chen, W.; Chen, S.; Ding, Z.; Chen, A.; Li, D. Abiotic transformation of atrazine in aqueous phase by biogenic bixbyite-type Mn2O3 produced by a soil-derived Mn(II)-oxidizing bacterium of Providencia sp. J. Hazard. Mater. 2022, 436, 129243. [Google Scholar] [CrossRef]
  89. Liu, Z.; Wang, J.; Qian, S.; Wang, G.; Wang, J.; Liao, S. Carbofuran degradation by biogenic manganese oxides. Bull. Environ. Contam. Toxicol. 2017, 98, 420–425. [Google Scholar] [CrossRef] [PubMed]
  90. Pal, A.; Mahamallik, P.; Saha, S.; Majumdar, A. Degradation of tetracycline antibiotics by advanced oxidation processes: Application of MnO2 nanomaterials. Nat. Resour. Eng. 2017, 2, 32–42. [Google Scholar] [CrossRef]
Figure 1. The diagram of the Mn cycle mediated by MnOB and reducing matters.
Figure 1. The diagram of the Mn cycle mediated by MnOB and reducing matters.
Ijerph 20 01272 g001
Figure 2. Network view map of 859 publications in the literature, generated by using VOSviewer.
Figure 2. Network view map of 859 publications in the literature, generated by using VOSviewer.
Ijerph 20 01272 g002
Figure 3. Processes designed for removal of Mn2+ from groundwater.
Figure 3. Processes designed for removal of Mn2+ from groundwater.
Ijerph 20 01272 g003
Figure 4. Removal of metals by using bioMnOx.
Figure 4. Removal of metals by using bioMnOx.
Ijerph 20 01272 g004
Figure 5. The schematic diagram for the removal of MOCs by bioMnOx.
Figure 5. The schematic diagram for the removal of MOCs by bioMnOx.
Ijerph 20 01272 g005
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

Cai, Y.; Yang, K.; Qiu, C.; Bi, Y.; Tian, B.; Bi, X. A Review of Manganese-Oxidizing Bacteria (MnOB): Applications, Future Concerns, and Challenges. Int. J. Environ. Res. Public Health 2023, 20, 1272. https://doi.org/10.3390/ijerph20021272

AMA Style

Cai Y, Yang K, Qiu C, Bi Y, Tian B, Bi X. A Review of Manganese-Oxidizing Bacteria (MnOB): Applications, Future Concerns, and Challenges. International Journal of Environmental Research and Public Health. 2023; 20(2):1272. https://doi.org/10.3390/ijerph20021272

Chicago/Turabian Style

Cai, Yanan, Kun Yang, Chaochao Qiu, Yunze Bi, Bowen Tian, and Xuejun Bi. 2023. "A Review of Manganese-Oxidizing Bacteria (MnOB): Applications, Future Concerns, and Challenges" International Journal of Environmental Research and Public Health 20, no. 2: 1272. https://doi.org/10.3390/ijerph20021272

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