Next Article in Journal
Characteristics and Treatment of Wastewater from the Mercaptan Oxidation Process: A Comprehensive Review
Next Article in Special Issue
Progress in Nitrogen Removal in Bioelectrochemical Systems
Previous Article in Journal
Rheological Properties of Aqueous Dispersions of Bacterial Cellulose
Previous Article in Special Issue
Bioelectrochemical Methane Production from Food Waste in Anaerobic Digestion Using a Carbon-Modified Copper Foam Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhanced Anaerobic Digestion by Stimulating DIET Reaction

1
Department of Civil Engineering, Inha University, 100 Inha-ro, Nam-gu, Incheon 22212, Korea
2
Department of Environmental Engineering, Korea Maritime and Ocean University, Busan 49112, Korea
3
Department of Energy Engineering, Gyeongnam National University of Science and Technology, Dongjin-ro 33, Jinju 52725, Korea
*
Author to whom correspondence should be addressed.
Processes 2020, 8(4), 424; https://doi.org/10.3390/pr8040424
Submission received: 11 March 2020 / Revised: 29 March 2020 / Accepted: 30 March 2020 / Published: 3 April 2020
(This article belongs to the Special Issue Development and Applications of Bioelectrochemical Systems)

Abstract

:
Since the observation of direct interspecies electron transfer (DIET) in anaerobic mixed cultures in 2010s, the topic “DIET-stimulation” has been the main route to enhance the performance of anaerobic digestion (AD) under harsh conditions, such as high organic loading rate (OLR) and the toxicants’ presence. In this review article, we tried to answer three main questions: (i) What are the merits and strategies for DIET stimulation? (ii) What are the consequences of stimulation? (iii) What is the mechanism of action behind the impact of this stimulation? Therefore, we introduced DIET history and recent relevant findings with a focus on the theoretical advantages. Then, we reviewed the most recent articles by categorizing how DIET reaction was stimulated by adding conductive material (CM) and/or applying external voltage (EV). The emphasis was made on the enhanced performance (yield and/or production rate), CM type, applied EV, and mechanism of action for each stimulation strategy. In addition, we explained DIET-caused changes in microbial community structure. Finally, future perspectives and practical limitations/chances were explored in detail. We expect this review article will provide a better understanding for DIET pathway in AD and encourage further research development in a right direction.

1. Introduction

Anaerobic digestion (AD) is a traditional biological process utilized for treating various organic wastes including wastewater and complex solid wastes. Via AD, organic matters are converted to biogas, rich in methane (CH4), while nutrients-rich residues suitable for high-quality compost are generated as byproduct [1,2,3]. AD process consists of up to four consecutive steps, i.e., hydrolysis, acidogenesis, acetogenesis, and methanogenesis, in which the rate-limiting step can be different depending on the type of treated feedstock [4]. However, the syntrophic acetogenesis was often considered as a bottleneck that is highly governing the whole rate of AD process [5,6]. Syntrophic acetogens convert C2–C6 organic acids/alcohols to electron donors of low-molecular weight, mainly H2 and acetic acid, which would be consumed by methanogens. This is a well-known indirect interspecies electron transfer (IIET) route, and many attempts were made to enhance this reaction. In particular, the importance of keeping low H2 partial pressure for acquiring stable IIET was addressed in many studies [7,8,9]. However, a new route, transferring electrons directly between acetogens or acidogens to methanogens, called direct interspecies electron transfer (DIET), was recently introduced (Figure 1).
The first observation of DIET was done by Summers et al. [10], between two Geobacter sp. (G. metallireducens and G. sulfurreducens) in metabolizing ethanol. Geobacter is a well-known anaerobic metal oxidizer and can produce electrically conductive pili (e-pili), which can exchange electrons among different cells [11]. Later on, to our knowledge, the first clear evidence of DIET for CH4 production was observed by Rotaru et al. [12]. When G. metallireducens was cocultured with Methanosaeta harundinacea fed with ethanol, it was possible to generate > 1 mole CH4/mole ethanol, which cannot be done theoretically. From one mole of ethanol, G. metallireducens can generate 1 mole of acetate with 2 moles of H2. Methanosaeta harundinacea cannot metabolize H2, and thus, the theoretical maximum CH4 yield is 1 mole CH4/mole ethanol. From this result, the authors indicated DIET between two species, which was further confirmed by metatranscriptomic analysis. Thereafter, not only Geobacter, but the presence of many other electro-active bacteria (EAB) was detected in AD broth, and the research to stimulate DIET reaction is extensively going on [13,14,15,16].
Through the studies of last ten years, it was found that DIET reaction could be stimulated either by the supplementation of a conductive material (CM) and/or applying external voltage (EV) (Figure 2) [17,18,19,20]. Basically, EAB can oxidize organics, and then the released electrons are directly transferred through the extracellular cytochromes or the e-pili of EAB to methanogens, which reduce CO2 into CH4 [12]. Almost all of the DIET-related studies agreed that e-pili are composed of PilA protein [21,22,23,24]. Supplementing CM, specifically magnetite, could interestingly mediate the electron transfer among cells through substituting the pilus-associated cytochrome OmcS that are commonly found in G. sulfurreducens [25]. Li et al. [26] observed an increase in electron transfer rate from 0.0017 ± 0.0003 to 0.0056 ± 0.0015 s−1, as a result of carbon cloth supplementation. Additionally, DIET could be stimulated through the establishment of two electrodes inside AD reactor, under slight EV with/without the presence of CM [3,27,28,29]. Such systems were called as bioelectrochemical systems merged with anaerobic digestion (BES-AD), which were initially applied for hydrogen generation [30,31,32,33]. However, the produced hydrogen was easily further converted to CH4 on the cathode, while the overall CH4 production rate was increased. The interesting thing found here was that the amounts of CH4 generated through the electrodes in BES-AD systems were limited, while the majority of CH4 production was assigned to the bulk of the broth, indicating active DIET reaction happening in the bulk [3,27,28,29].
Recent DIET-related review articles have mainly focused on the evidences of DIET and the addition of CM for stimulating DIET [34,35,36,37]; while the application of EV for stimulating DIET and corresponding microbial community change were not addressed. In this review, we are combing various points regarding DIET, i.e., DIET-caused thermodynamic favorability, recent researches that employed CM and/or EV for DIET stimulation, DIET-driven microbial community change, and future insights for DIET-related hot topics. Further, herein, we are giving a special focus for the contradictory results and debatable points that were found in recent DIET-related research articles. This review is expected to draw a panoramic understanding for various aspects related to DIET and figure out the weakness/challenging points in our knowledge about DIET.

2. DIET-Caused Theoretical Advantages

From the thermodynamic perspective, it was basically expected that DIET can provide more energy benefits for the syntrophic partners than that secured by IIET. This is because in the case of DIET, there is no need for the generation and diffusion of metabolites, where these reactions consume energy [38]. In addition, under such conditions, there is no potential problems from accumulation of H2, and as a result, faster electron transfer is expected [39]. As clearly shown in Table 1, DIET can provide a more thermodynamically favorable route for producing CH4 from simple, toxic, and complex substrates, comparted to IIET. The reason for this difference is that Gibbs free energy of H2 is zero, while H+ has Gibbs free energy of −39.8 kJ/mol at pH 7 [40]. Therefore, DIET is advantageous for gaining more energy, facilitating cell growth and finally enhancing reaction rate. In addition, recent studies could calculate the external electron transfer rates per cell pair (cp) values, through mathematical modeling, which were 44.9 × 103 and 5.24 × 103 e cp−1 s−1 for DIET and H2-based IIET, respectively [12,41,42].
From energetic standpoint, some energy loss, however, was observed in the case of DIET occurrence; this was ascribed to the activation energy needed for electron donating and accepting redox cofactors (such as cytochromes) in order to transfer electrons [41]. It is worthy to highlight that the aforementioned electron transfer rate constants that were evaluated by Storck’s group [41] were based on the respiration of Shewanella sp. on the electrodes. Considering that this Shewanella sp. has a limited electrical current density, and the transfer rate has high impact upon energy loss [44], it can be concluded that the above-stated values might be lower than the real values associated with electron transfer under DIET-stimulated conditions. Future advanced experiments and mathematical models are needed in order to accurately evaluate electron transfer constants for DIET and IIET.
The dominance of DIET over other potential syntrophic metabolism routes might happen as a response for the broth reactions. In other words, DIET can be a more favored route for microorganism in order to encounter environmental conditions [38]. For instance, when overall energy yield from a certain reaction is limited, the consumption of energy for generating metabolite, as electron shuttles, can be unfavored; instead, DIET in such case can be the optimal choice for energy saving [10,38]. For example, DIET could dominate over other IIET routes, when the system was subjected to acidic inhibition [45,46], toxic substrate [29,43], sulfate-rich wastewater [47], or ammonia containing wastewater [48].
So far, investigating the accurate number of electrons released from one organism and received by another one is still difficult, especially in the case of mixed cultures; therefore, the determination of the accurate energy gain for each microorganism is still limited. The determination of in-situ free energy potential, associated with DIET, is required for evaluating the portion of free energy that are available/consumed during DIET.

3. CM as a Tool for DIET Pathway Stimulation

Based on their capability for DIET stimulation, CM are widely utilized for enhancing the efficiency and stability of CH4 production [42,49]. Utilized CM can be classified into carbon-based conductive materials (CBCM) and iron-based conductive materials (IBCM). [49]. CM could compensate the lack of vital cell components, needed for completion of DIET, such as pili and c-type cytochrome [50,51,52]. In ethanol-fed methanogenic co-culture that consisted of Methanosarcina barkeri and G. sulfurreducens, which had deficiency of pili and pili-associated cytochrome OmcS, no CH4 could be produced unless CM, i.e., biochar, was supplemented [52]. Chen’s group inferred that the presence of biochar enabled 86% of the electrons produced during ethanol oxidation to be utilized for CH4 generation; while in the absence of biochar, neither ethanol oxidation nor CH4 generation was observed [52]. In addition, when the pilin-associated c-type cytochrome (OmcS), required for DIET, was not available in Geobacter species, magnetite could compensate such lack, and reduce the need for cell to produce OmcS [25]. Similarly, when PilA gene was deleted or yielded very limited pili, Geobacter strains grew and made co-cultures in the presence of granular activated carbon (GAC); this observation would never happen without GAC role [53,54].
Table 2 shows the positive/negative impact of CM upon CH4 productivity for various wastes, treated in batch and continuous treatment systems, specifically up-flow anaerobic sludge blanket reactor (UASB). Generally, when CMs are added to batch tests, lag phase reduction and CH4 production rate increase could always be observed. However, the increase in CH4 production yield is not necessarily achieved [48,55,56,57,58,59,60,61]. Even, slight negative impact (−4%, compared to control) upon cumulative CH4 production yield was observed by magnetite addition [62]. But, some reports showed that an increase in CH4 yield can be also observed after CM supplementation [43,63]. DIET-stimulated digesters were also able to sustain against the harsh operational conditions, e.g., accumulation of sulfate [47] or ammonia [48] or phenol [29]. Consequently, previous studies highlighted that complex substrates, e.g., dog food, sugar industrial wastewater, and leachate of municipal solid waste, can be effectively treated under DIET-stimulated conditions [49,64,65,66,67]. For justifying these results, authors referred that CM’s presence might secure higher stability for the reactors and resulted in preventing the inhibition-causing problems. Further reason was the enrichment of either DIET-capable or incapable methanogens. In addition, some of those studies referred to the negativity of reaction Gibbs free energy value as a potential reason for enhanced performance. Finally, all these studies highlighted that DIET pathway saves energy, which is reflected on either overcoming the harsh conditions and/or generating higher volumes of CH4.
The different enhancement performances, implemented by CBCM and IBCM can be explained based on the difference in size and structure of such CM [68]. In specific, nano-sized IBCM are probably attached to the bacterial cells, since CM have smaller size than bacterial cells [25,68]. On the other hand, microbes can be efficiently attached to the surface of CBCM, providing electron conduits for many microorganisms [49,51,52]. This strategy is more preferred for microbes, from metabolic standpoint, since there is no energy consumption for conductive pili synthesis [69]. This approach can be utilized for providing the efficiency of using the DIET-based syntrophy in anaerobic digestors with various configurations [70].
For interpreting the impact of CM upon methanogenesis, two important factors need to be considered, i.e., electrical conductivity and oxidation redox potential (ORP). Electrical conductivity of the utilized microbial consortia is supposed to be higher in CM-amended reactors, compared to control [16,53,55,67]. Such increases can be related to the improved e-pili secretion, which is implemented by EAB that are enriched in the presence of CM [67,71,72]. Furthermore, ORP could be lowered down, to more negative values, as a result of CM addition [73]; this is beneficial for CH4 generation process, which need ORP of (−200 to −400 mv) in order to be successfully implemented [74]. From thermodynamic standpoint, the supplementation of CM, which are featured of their negative redox potential, into AD broth can lead to directing reactions to DIET route [75]. In addition, it was found that the types and activities of e-pili and cytochromes, expressed from Geobacter and Pelobacter species for electron transfer, are different based on the surface potential, which is linked to the supplemented CM [76].
Notwithstanding, the enhancing impact, implemented by CM, might be also assigned to the enhanced biomass colonization, since CM usually have large surface that is capable for cell adhesion, especially in the case of nano-sized materials [77,78,79]. This can justify the enhanced CH4 productivity, as CBCM are supplemented (Table 2). In case of IBCM, the electrical conductivity might have more remarkable impact than surface area [43,80]. Further study concluded that magnetite could play double role, i.e., (i) stimulating the up-regulated secretion of key enzymes that are involved in the substrate degradation and (ii) inducing the expression of DIET-related proteins [60]. Other studies confirmed that IBCM, specifically, magnetite and ferrihydrite, can act as oxidizing agents and enhance substrate degradation [20,81]. The latter role is not indeed related to DIET stimulation. Instead, it might be achieved at the expense of DIET stimulation, since IBCM are going to be consumed in such reactions instead of stimulating DIET. More research is needed in the future in order to highlight the relation between the enhancement, done by IBCM, and the characteristics of this CM, such as morphology and crystallinity.

4. Electrical Energy Input as Tool for DIET Pathway Stimulation

In BES-AD, slight EV is applied in order to induce the vital reactions [82], and as a consequence, DIET reaction, and the growth of DIET-capable species are stimulated, directly transferring electrons from EAB species to anode surface [83]. Table 3 shows the impact of various EV upon CH4 from various wastes. Clearly, the slight EV application (−0.3–−1.8 V vs. standard hydrogen electrode (SHE)) led to significant enhancements in CH4 production (20–1360%). Previous studies showed that optimal range can be (−0.3–−1.4 V vs. SHE) [84] or (−0.2–−0.8 V vs. SHE) [85]. Variation in the optimal EV was found in the literature. For instance, EV of −0.3 V vs. SHE was found to be effective for the treatment of phenol [29] and waste activated sludge (WAS) [86]; however, when EV was changed to -0.6 V vs. SHE, no change in the enhanced performance was observed in the case of phenol, while negative impact was denoted in the case of WAS. Other studies showed different results while optimizing EV. In specific, EV ranges of (−0.5–−1.5) and (−0.6–−1.2 V vs. SHE) were tested for treating glucose and WAS, where optimal EV was found to be −1.0 and −0.8 V vs. SHE, respectively [87,88].
It was reported that EV application could reduce the time for stabilizing reaction. However, this is not necessarily leading to higher CH4 yield. In this regard, Park group [93] treated highly concentrated food waste in AD reactor, where EV of −0.3 V vs. SHE was applied. Expectedly, both stabilization time and CH4 production rate were accelerated by 4.0 and 1.7 times faster than control, respectively. However, the final CH4 production was the same for EV-supplemented reactor and control. Similarly, our research group treated phenol and could not observe significant difference in CH4 productivity between EV-applied UASB and control UASB, at low OLR (<5.3 kg chemical oxygen demand (COD)/m3/d). However, the significant impact was observed at higher OLRs [29]. Another research group highlighted that the electrode polarization could boost volumetric CH4 production and CH4 content by 53% and 20.6%, respectively [19]. This implies that EV application can be the key for various merits in AD. The proposed mechanisms for the achieved enhancements under DIET-stimulated conditions were linked mainly to the increased abundance of EAB, along with methanogens, and the advanced electron transfer rate. The effective contribution to DIET enhancement seems to be more related to broth consortia than electrode biofilm, since the amount of CH4 released from electrodes was very limited, compared to that from the broth [3,27,28,29]. Specifically, CH4 generation from the electrodes was found to be lower than 5% of the total generated CH4 [3,29].
Generally, the required EV was decided by considering both electrode potential for H2 evolution, which differs based on the electrode material, and electrode potential at which the reduced products could be harvested [95,96]. For instance, when CH4 production was implemented using a pure culture of Methanobacterium-like archaeon, attached to graphite cathode, the selected EV was −0.4 V vs. SHE [97]. Such EV was more positive than the potential for H2 generation using graphite as electrode material [98,99].
The effectiveness of BES-AD was found to be cardinally linked to the utilized electrode material. Generally, the utilized electrodes are required to have high particular characteristics, such as large surface area, electrochemical active surface, and acceptable electrical conductivity [100,101]. Further, biocompatibility and low cost are additional recommended characteristics [102]. Carbon materials seem to be the most utilized electrode material (Table 2). This might be because of carbon characteristics, e.g., low cost, high conductivity, good biocompatibility, and high capability for biofilm formation [103,104]. For enhancing the aforementioned features, carbon-based electrodes were modified by nano-sized particles of transition elements as (Co and P) [91], Pt [92], Fe [86], and Ni [93] (Table 2).

5. Microbial Community Change after DIET Stimulation

Generally, the most commonly studied DIET-implicated genera are Geobacter and Shewanella [10,68,105]. In their pioneering work, Summers et al. [10] observed DIET occurrence in a defined co-culture that consisted of G. metallireducens, as an exoelectrogenic bacteria, and G. sulfurreducens, a fumarate reducer. Later on, metatranscriptomic analysis was employed for confirming DIET occurrence between G. sulfurreducens and G. metallireducens, [106] or G. sulfurreducens and Methanothrix sp. [107]. These studies considered the limited abundance for hydrogenase and/or formate dehydrogenase transcript as strong indicator for DIET occurrence among the syntrophic partners. Further, DIET-capable syntrophic co-culture, which consisted of Methanosaeta sp. and Geobacter sp., was detected in BES-AD system [108].
In addition, the potential DIET occurrence was previously claimed between Geobacter sp. and acetoclastic methanogen (Methanosaeta sp.) [12], hydrogentrophic methanogen (Methanobacterium sp.) [7], or methanogen that can work as both acetoclastic and hydrogentrophic (Methanosarcina sp.) [50]. Apparently, various types of methanogens exhibited DIET capability. However, so far, there is no conclusion regarding which methanogen must be enriched as a result of DIET-stimulating condition, either by CM supplementation and/or EV application. For instance, when the low strength wastewater, which contained glucose as a main carbon source, was treated through BES system, equipped with graphene/polypyrrole electrode, hydrogentrophic methanogens’ dominance was recorded [109]. However, the usage of nano-graphene for the treatment of the same wastewater could induce the presence of Methanosaeta sp. (acetoclastic methanogen) [110] (Table 4). Another informative example can be obtained through exploring the impact of ethanol, utilized as DIET stimulator, upon the syntrophic metabolism of propionate and butyrate [72]. Results exhibited that ethanol’s presence enriched the abundance of Methanosarcina sp. and Methanosaeta sp., irrespective of whether the substrate was propionate or butyrate. However, in case of propionate, both of Methanosarcina sp. and Methanosaeta sp. could be enriched by almost the same ratio. On the other hand, when the substrate was propionate, ethanol’s presence could mainly augment the abundance of Methanosarcina sp. [72]. Like the relation between the supplemented CM and enriched methanogens, the relation between the value of the applied EV and the type of methanogens that can be enriched is still unclear. In addition, the change in the composition of microbial community cannot directly refer to the active type of interspecies electron transfer [80]. Another unclear point is related to the electron receptors of enriched methanogens. Basically, there is an agreement that Geobacter sp. is exporting electrons through its conductive pili [50,111,112]; however, the mechanism by which electrons are received by methanogen is still vaguely understood [12].
In BES-AD studies, various species could be enriched, such as Methanococcus maripaludis [119] and Methanobacterium congolense [120]. The type of enriched consortia might be related to the utilized electrode and the utilized catalyst [121]. For instance, mixed culture of anaerobic consortia could have better respiration condition when stainless-steel was used as cathode, than that while using pure conventional carbon-based materials [122]. Further, microbial growth/colonization on the surface of the electrode is governed by the electrode surface properties since they affect microbes–electrode interaction, which depends upon the formed bond, e.g., Van der Waals force, hydrogen bonding, or electrostatic interaction [101]. For instance, a faster growth for biofilm could be achieved, when the utilized electrode material has a positive charge. This is because of microorganisms’ negative charge [123]. Further, extra-porous and nano-sized materials are perfect scaffold for improved microbial attachment/growth, enhancing biofilm formation [124]. A point of debate is the impact of DIET stimulation upon microbial diversity; some researches confirmed diversity lowering, as a result of DIET stimulation [29,125]; however, other researches referred to increased microbial diversity under DIET-stimulated condition [86,126].
Among the 11 studies shown in Table 2, only 4 research articles detected Geobacter sp. [47,48,57,59]. Furthermore, those four studies revealed very slight abundance for Geobacter sp. (<5% of the total bacterial community detected). Previous review article also highlighted that most of the studies that claimed DIET occurrence could not detect Geobacter sp. [34]. This dilemma might have two possible explanations. Firstly, if the claimed DIET did not really happen, then the enhancements of CH4 productivity, in terms of yield and/or rate of production, might be assigned to other roles played by the utilized DIET-stimulator, such as enhancing biomass colonization [77,78,79] or stimulating organics’ degrading microbes for enzymes’ secretion and substrate utilization [29,60]. Secondly, if the claimed DIET really happened, then the applied DIET stimulation strategy could induce DIET-capable syntrophic bacteria, yet to be confirmed as EAB [34,127]. The latter explanation might agree with recent findings that confirmed DIET can be a common electron transfer option for not only Geobacter and Shewanella but also for various syntrophs such as Pelobacter sp. and Syntrophorhabdus sp., which are potential EAB [3,32]. Future studies need to confirm DIET capability/incapability of the detected species in DIET-stimulated reactors, in order to assure DIET occurrence; in addition, the enrichments in the abundance of other species, different from EAB, must be involved in explaining CH4 productivity and substrate utilization efficiency profiles.

6. Conclusions and Future Perspectives

This review tried to give comprehensive explanation for the recent advanced knowledge regarding DIET, as an alternative metabolic approach for enhancing CH4 productivity from various wastes. DIET reaction could provide theoretically advantageous route due to its more negative value of ΔG0, and there is no need for the generation and diffusion of metabolites. It could be stimulated through the supplementation of CM and/or the application of EV. CBCM could compensate the lack of Geobacter conductive pili, and IBCM could reduce the bacterial need for OmcS production. In applying EV, carbon was the commonly utilized electrode material, because of its high conductivity and low cost. The increased amount of CH4 production from the electrodes was negligible, indicating active DIET reaction occurring in the broth. DIET establishment leads to enriching acetoclastic and/or hydrogentrophic methanogens. Geobacter sp. was absent or slightly present in most of DIET studies that utilized mixed culture as inocula. Away from EAB, other bacterial species that are enriched after DIET stimulation might be implicated in the detected CH4 productivity improvements.
(a)
Engineering perspectives
From an engineering perspective, the economic feasibility of DIET stimulation in AD reactors is still questionable. It can be roughly considered that the cost of CM supplementation, as DIET stimulation approach, might be lower than applying EV, which requires the installation of two electrodes that are recommended to be fabricated using CM. However, further studies need to be implemented in order to compare the economic feasibility of the two DIET stimulation approaches, considering the maintenance cost and CM loss during EV application and CM supplementation, respectively. For reducing the overall cost needed for DIET stimulation, two approaches are suggested. First approach is the usage of relatively cheap CM, such as red mud, or to recover that CM from other wastes, e.g., biochar recovery from digestate by pyrolysis.
Second approach depends on DIET establishing through two steps: (i) ethanol-type fermentation of easily biodegradable substrate, such as polysaccharide and (ii) methanogenesis, where, the ethanol-rich effluent, obtained from the first step was utilized for providing better conditions for DIET stimulation in the second step. Such approach was based on the special favorability of DIET-implicated microorganisms to ethanol, as a substrate [128]. Zhao group denoted that sludge fed with ethanol had higher abundance for Methanothrix sp. and gene for pilA, compared to that when propionate and butyrate were used as feeds. This strategy caused increase in CH4 production from WAS by 36% [129] and 30% [130]. On the other hand, fabrication of electrodes that have low cost can support EV-based DIET performing. These perspectives can help in the commercialization of DIET-based reactors; however, more successful tests for bench-scale reactors are needed. In this regard, Yee et al. [131] provided a detailed practical protocol for the setting of the electrochemical reactors. Future tests have to consider calculating benefit/cost ratio for DIET-stimulated digestor and comparing it to control digestors. Thereafter, large scale application will just be an issue of time.
(b)
Scientific perspectives
Undoubtedly, there is a literature chaos, specifically around CM-based DIET stimulation topic, since DIET occurrence has been claimed in many researches, without revealing vital experimental evidences. Therefore, recent review articles indicated that more experimental observations are required before claiming DIET stimulation [36,80]. Future researches should direct more effort for verifying DIET activation and/or other potential changes in anaerobic growth, which can be implicated in the observed enhancement in CH4 production rate and/or yield. Recently, Van Steendam et al. [132] evaluated some advanced DIET identification methods, including fluorescence in situ hybridization, cyclic voltammetry, scanning, transmission electron microscope, and various (meta-)omic approaches; the usage of combined methods for confirming DIET occurrence was finally recommended. Additionally, Neu et al. [133] presented temperature-resolved THz spectroscopy that can easily measure the conductivity of wild-type pili. Further, Yee et al. [131] explained some electrochemical techniques that can be utilized for determining the electroactivity and the mode of electron transfer during the reactions. The aforementioned studies are referring to the right track of research that is worthy to be followed. More studies are needed to describe strict and simple verification strategy for validating DIET occurrence/absence.
Since they have noticed the electric conductivity of syntrophic aggregates that contained Geobacter sp., Lovley’s team explained the observed long-distance electron transfer based on e-pili, composed of PilA protein, which had metallic-like conductivity and could act as conduit [10,21,22,134,135]. However, a recent work by Wang et al. [136] gave proofs to indicate that the conduit for such long-distance electron exchange process is not e-pili, rather, it can be cytochrome OmcS, in its polymerized chain form. Wang’s group concluded that the role played by PilA is stimulating the excretion of OmcS. This view around the mechanism of electron transfer needs further research to be established. From the way how the e-pili were formed, it seems that we need to understand more about the characteristics of those e-pili. Very recently, an astonishing research work proved that the protein nanowires that were harvested from G. sulfurreducens can form a thin-film device, which is able to generate continuous electric power (0.5 V across a 7 μm thick film) in the ambient environment [137].
The rate of electron transfer seems to be one of the hot points that needs further experimental research and predictive models, since available information in literature about the efficiency and energetic losses, during DIET, are limited. Moreover, more models are needed to witness the response of microbial consortia to harsh environmental changes, and the mechanism of microbial cooperation for metabolism under such conditions. In other words, the question that needs to be answered is to what extent DIET can be the way for the best diet?

Author Contributions

Conceptualization: D.-H.K.; Writing—Original Draft Preparation: A.M. and S.I., Writing—Review & Editing: D.-H.K., Y.-C.S. and Y.A.; Supervision: D.-H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by INHA UNIVERSITY Research Grant (INHA-61668-01).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Im, S.; Petersen, S.; Lee, D.; Kim, D. Effects of storage temperature on CH4 emissions from cattle manure and subsequent biogas production potential. Waste Manag. 2020, 101, 35–43. [Google Scholar] [CrossRef]
  2. Shin, S.; Im, S.; Mostafa, A.; Lee, M.; Yun, Y.; Oh, S.; Kim, D. Effects of pig slurry acidification on methane emissions during storage and subsequent biogas production. Water Res. 2019, 152, 234–240. [Google Scholar] [CrossRef]
  3. Im, S.; Yun, Y.; Song, Y.; Kim, D. Enhanced anaerobic digestion of glycerol by promoting DIET reaction. Biochem. Eng. J. 2019, 142, 18–26. [Google Scholar] [CrossRef]
  4. Feng, Q.; Song, Y.; Kim, D.; Kim, M.; Kim, D. Influence of the temperature and hydraulic retention time in bioelectrochemical anaerobic digestion of sewage sludge. Int. J. Hydrog. Energy 2018, 44, 2170–2179. [Google Scholar] [CrossRef]
  5. Shen, L.; Zhao, Q.; Wu, X.; Li, X.; Li, Q.; Wang, Y. Interspecies electron transfer in syntrophic methanogenic consortia: From cultures to bioreactors. Renew. Sustain. Energy Rev. 2016, 54, 1358–1367. [Google Scholar] [CrossRef]
  6. Meslé, M.; Dromart, G.; Oger, P. Microbial methanogenesis in subsurface oil and coal. Res. Microbiol. 2013, 164, 959–972. [Google Scholar] [CrossRef]
  7. Zhao, Z.; Li, Y.; Quan, X.; Zhang, Y. Towards engineering application: Potential mechanism for enhancing anaerobic digestion of complex organic waste with different types of conductive materials. Water Res. 2017, 115, 266–277. [Google Scholar] [CrossRef]
  8. Cazier, E.; Trably, E.; Steyer, J.; Escudie, R. Reversibility of hydrolysis inhibition at high hydrogen partial pressure in dry anaerobic digestion processes fed with wheat straw and inoculated with anaerobic granular sludge. Waste Manag. 2019, 85, 498–505. [Google Scholar] [CrossRef]
  9. Cazier, E.; Trably, E.; Steyer, J.; Escudie, R. Biomass hydrolysis inhibition at high hydrogen partial pressure in solid-state anaerobic digestion. Bioresour. Technol. 2015, 190, 106–113. [Google Scholar] [CrossRef] [Green Version]
  10. Summers, Z.; Fogarty, H.; Leang, C.; Franks, A.; Malvankar, N.; Lovley, D. Direct exchange of electrons within aggregates of an evolved syntrophic coculture of anaerobic bacteria. Science 2010, 80, 819–824. [Google Scholar] [CrossRef] [Green Version]
  11. Liu, X.; Shi, L.; Gu, J. Microbial electrocatalysis: Redox mediators responsible for extracellular electron transfer. Biotechnol. Adv. 2018, 36, 1815–1827. [Google Scholar] [CrossRef]
  12. Rotaru, A.; Shrestha, P.; Liu, F.; Shrestha, M.; Shrestha, D.; Embree, M.; Zengler, K.; Wardman, C.; Nevin, K.; Lovley, D. A new model for electron flow during anaerobic digestion: Direct interspecies electron transfer to Methanosaeta for the reduction of carbon dioxide to methane. Energy Environ. Sci. 2014, 7, 408–415. [Google Scholar] [CrossRef]
  13. Xiao, L.; Liu, F.; Liu, J.; Li, J.; Zhang, Y.; Yu, J.; Wang, O. Nano-Fe3O4 particles accelerating electromethanogenesis on an hour-long timescale in wetland soil. Environ. Sci. Nano 2018, 5, 436–445. [Google Scholar] [CrossRef]
  14. Zhang, J.; Lu, Y. Conductive Fe3O4 nanoparticles accelerate syntrophic methane production from butyrate oxidation in two different lake sediments. Front. Microbiol. 2016, 7, 1–9. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, W.; Zhang, J.; Lu, Y. Stimulation of carbon nanomaterials on syntrophic oxidation of butyrate in sediment enrichments and a defined coculture. Sci. Rep. 2018, 8, 1–13. [Google Scholar] [CrossRef]
  16. Li, H.; Chang, J.; Liu, P.; Fu, L.; Ding, D.; Lu, Y. Direct interspecies electron transfer accelerates syntrophic oxidation of butyrate in paddy soil enrichments. Environ. Microbiol. 2015, 17, 1533–1547. [Google Scholar] [CrossRef]
  17. Ye, J.; Hu, A.; Ren, G.; Chen, M.; Tang, J.; Zhang, P.; Zhou, S.; He, Z. Enhancing sludge methanogenesis with improved redox activity of extracellular polymeric substances by hematite in red mud. Water Res. 2018, 134, 54–62. [Google Scholar] [CrossRef]
  18. Piao, D.; Song, Y.; Kim, D. Bioelectrochemical enhancement of biogenic methane conversion of coal. Energies 2018, 11, 2577. [Google Scholar] [CrossRef] [Green Version]
  19. Feng, Q.; Song, Y.; Yoo, K.; Kuppanan, N.; Subudhi, S.; Lal, B. Polarized electrode enhances biological direct interspecies electron transfer for methane production in upflow anaerobic bioelectrochemical reactor. Chemosphere 2018, 204, 186–192. [Google Scholar] [CrossRef]
  20. Kato, S.; Hashimoto, K.; Watanabe, K. Methanogenesis facilitated by electric syntrophy via (semi)conductive iron-oxide minerals. Environ. Microbiol. 2012, 14, 1646–1654. [Google Scholar] [CrossRef]
  21. Lovley, D.; Walker, D. Geobacter protein nanowires. Front. Microbiol. 2019, 10, 2078. [Google Scholar] [CrossRef] [Green Version]
  22. Ueki, T.; Walker, D.; Tremblay, P.; Nevin, K.; Ward, J.; Woodard, T.; Nonnenmann, S.; Lovley, D. Decorating the outer surface of microbially produced protein nanowires with peptides. ACS Synth. Biol. 2019, 8, 1809–1817. [Google Scholar] [CrossRef]
  23. Malvankar, N.; Lovley, D. Microbial nanowires for bioenergy applications. Curr. Opin. Biotechnol. 2014, 27, 88–95. [Google Scholar] [CrossRef]
  24. Tan, Y.; Adhikari, R.; Malvankar, N.; Ward, J.; Nevin, K.; Woodard, T.; Smith, J.; Snoeyenbos-West, O.; Franks, A.; Tuominen, M.; et al. The low conductivity of Geobacter uraniireducens pili suggests a diversity of extracellular electron transfer mechanisms in the genus Geobacter. Front. Microbiol. 2016, 7, 1–10. [Google Scholar] [CrossRef]
  25. Liu, F.; Rotaru, A.; Shrestha, P.; Malvankar, N.; Nevin, K.; Lovley, D. Magnetite compensates for the lack of a pilin-associated c-type cytochrome in extracellular electron exchange. Environ. Microbiol. 2015, 17, 648–655. [Google Scholar] [CrossRef]
  26. Li, J.; Xiao, L.; Zheng, S.; Zhang, Y.; Luo, M.; Tong, C.; Xu, H.; Tan, Y.; Liu, J.; Wang, O.; et al. A new insight into the strategy for methane production affected by conductive carbon cloth in wetland soil: Beneficial to acetoclastic methanogenesis instead of CO2 reduction. Sci. Total Environ. 2018, 643, 1024–1030. [Google Scholar] [CrossRef]
  27. Song, Y.; Joicy, A.; Jang, S. Direct interspecies electron transfer in bulk solution significantly contributes to bioelectrochemical nitrogen removal. Int. J. Hydrog. Energy 2018, 44, 1–11. [Google Scholar] [CrossRef]
  28. Feng, Q.; Song, Y.; Yoo, K.; Kuppanan, N.; Subudhi, S.; Lal, B. Bioelectrochemical enhancement of direct interspecies electron transfer in upflow anaerobic reactor with effluent recirculation for acidic distillery wastewater. Bioresour. Technol. 2017, 241, 171–180. [Google Scholar] [CrossRef]
  29. Mostafa, A.; Im, S.; Lee, M.; Song, Y.; Kim, D. Enhanced anaerobic digestion of phenol via electrical energy input. Chem. Eng. J. 2020, 389, 124501. [Google Scholar] [CrossRef]
  30. Ahn, Y.; Im, S.; Chung, J. Optimizing the operating temperature for microbial electrolysis cell treating sewage sludge. Int. J. Hydrog. Energy 2017, 42, 27784–27791. [Google Scholar] [CrossRef]
  31. Cheng, S.; Hamelers, H. Critical review microbial electrolysis cells for high yield hydrogen gas production from organic matter. Environ. Sci. Technol. 2008, 42, 8630–8640. [Google Scholar]
  32. Guo, X.; Liu, J.; Xiao, B. Bioelectrochemical enhancement of hydrogen and methane production from the anaerobic digestion of sewage sludge in single-chamber membrane-free microbial electrolysis cells. Int. J. Hydrog. Energy 2013, 38, 1342–1347. [Google Scholar] [CrossRef]
  33. Kyazze, G.; Popov, A.; Dinsdale, R.; Esteves, S.; Hawkes, F.; Premier, G.; Guwy, A. Influence of catholyte pH and temperature on hydrogen production from acetate using a two chamber concentric tubular microbial electrolysis cell. Int. J. Hydrog. Energy 2010, 35, 7716–7722. [Google Scholar] [CrossRef]
  34. Yin, Q.; Wu, G. Advances in direct interspecies electron transfer and conductive materials: Electron flux, organic degradation and microbial interaction. Biotechnol. Adv. 2019, 37, 107443. [Google Scholar] [CrossRef]
  35. Baek, G.; Kim, J.; Kim, J.; Lee, C. Role and potential of direct interspecies electron transfer in anaerobic digestion. Energies 2018, 11, 107. [Google Scholar] [CrossRef] [Green Version]
  36. Baniamerian, H.; Isfahani, P.G.; Tsapekos, P.; Alvarado-Morales, M.; Shahrokhi, M.; Vossoughi, M.; Angelidaki, I. Application of nano-structured materials in anaerobic digestion: Current status and perspectives. Chemosphere 2019, 229, 188–199. [Google Scholar] [CrossRef]
  37. Baek, G.; Kim, J.; Lee, C. A review of the effects of iron compounds on methanogenesis in anaerobic environments. Renew. Sustain. Energy Rev. 2019, 113, 109282. [Google Scholar] [CrossRef]
  38. Lovley, D.R. Reach out and touch someone: Potential impact of DIET (direct interspecies energy transfer) on anaerobic biogeochemistry, bioremediation, and bioenergy. Rev. Environ. Sci. Biotechnol. 2011, 10, 101–105. [Google Scholar] [CrossRef] [Green Version]
  39. Viggi, C.; Rossetti, S.; Fazi, S.; Paiano, P.; Majone, M.; Aulenta, F. Magnetite particles triggering a faster and more robust syntrophic pathway of methanogenic propionate degradation. Environ. Sci. Technol. 2014, 48, 7536–7543. [Google Scholar] [CrossRef]
  40. Cheng, Q.; Call, D. Hardwiring microbes: Via direct interspecies electron transfer: Mechanisms and applications. Environ. Sci. Process. Impacts 2016, 18, 968–980. [Google Scholar] [CrossRef]
  41. Storck, T.; Virdis, B.; Batstone, D. Modelling extracellular limitations for mediated versus direct interspecies electron transfer. ISME J. 2016, 10, 621–631. [Google Scholar] [CrossRef] [Green Version]
  42. Lovley, D. Happy together: Microbial communities that hook up to swap electrons. ISME J. 2017, 11, 327–336. [Google Scholar] [CrossRef] [Green Version]
  43. Mostafa, A.; Im, S.; Song, Y.; Kang, S.; Kim, D. Enhanced anaerobic digestion of long chain fatty acid by adding magnetite and carbon nanotubes. Microorganisms 2020, 8, 333. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, Y.; Li, J.; Liu, F.; Yan, H.; Li, J.; Zhang, X. Reduction of Gibbs free energy and enhancement of Methanosaeta by bicarbonate to promote anaerobic syntrophic butyrate oxidation. Bioresour. Technol. 2018, 267, 209–217. [Google Scholar] [CrossRef]
  45. Zhao, Z.; Zhang, Y.; Li, Y.; Dang, Y.; Zhu, T.; Quan, X. Potentially shifting from interspecies hydrogen transfer to direct interspecies electron transfer for syntrophic metabolism to resist acidic impact with conductive carbon cloth. Chem. Eng. J. 2017, 313, 10–18. [Google Scholar] [CrossRef]
  46. Feng, Q.; Song, Y.; Yoo, K.; Kuppanan, N.; Subudhi, S.; Lal, B. Influence of neutralization in acidic distillery wastewater on direct interspecies electron transfer for methane production in an upflow anaerobic bioelectrochemical reactor. Int. J. Hydrog. Energy 2017, 42, 27774–27783. [Google Scholar] [CrossRef]
  47. Li, Y.; Zhang, Y.; Yang, Y.; Quan, X.; Zhao, Z. Potentially direct interspecies electron transfer of methanogenesis for syntrophic metabolism under sulfate reducing conditions with stainless steel. Bioresour. Technol. 2017, 234, 303–309. [Google Scholar] [CrossRef]
  48. Zhuang, L.; Ma, J.; Yu, Z.; Wang, Y.; Tang, J. Magnetite accelerates syntrophic acetate oxidation in methanogenic systems with high ammonia concentrations. Microb. Biotechnol. 2018, 11, 710–720. [Google Scholar] [CrossRef]
  49. Liu, F.; Rotaru, A.; Shrestha, P.; Malvankar, N.; Nevin, K.; Lovley, D. Promoting direct interspecies electron transfer with activated carbon. Energy Environ. Sci. 2012, 5, 8982–8989. [Google Scholar] [CrossRef] [Green Version]
  50. Rotaru, A.; Shrestha, P.; Liu, F.; Markovaite, B.; Chen, S.; Nevin, K.; Lovley, D. Direct interspecies electron transfer between Geobacter metallireducens and Methanosarcina barkeri. Appl. Environ. Microbiol. 2014, 80, 4599–4605. [Google Scholar] [CrossRef] [Green Version]
  51. Chen, S.; Rotaru, A.; Liu, F.; Philips, J.; Woodard, T.; Nevin, K.; Lovley, D. Carbon cloth stimulates direct interspecies electron transfer in syntrophic co-cultures. Bioresour. Technol. 2014, 173, 82–86. [Google Scholar] [CrossRef] [Green Version]
  52. Chen, S.; Rotaru, A.; Shrestha, P.; Malvankar, N.; Liu, F.; Fan, W.; Nevin, K.; Lovley, D. Promoting interspecies electron transfer with biochar. Sci. Rep. 2014, 4, 5019. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, S.; Chang, J.; Lin, C.; Pan, Y.; Cui, K.; Zhang, X.; Liang, P.; Huang, X. Enhancement of methanogenesis via direct interspecies electron transfer between Geobacteraceae and Methanosaetaceae conducted by granular activated carbon. Bioresour. Technol. 2017, 245, 132–137. [Google Scholar] [CrossRef] [PubMed]
  54. Ueki, T.; Nevin, K.; Rotaru, A.; Wang, L.; Ward, J.; Woodard, T.; Lovley, D. Geobacter strains expressing poorly conductive pili reveal constraints on direct interspecies electron transfer mechanisms. mBio 2018, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  55. Yan, W.; Shen, N.; Xiao, Y.; Chen, Y.; Sun, F.; Kumar Tyagi, V.; Zhou, Y. The role of conductive materials in the start-up period of thermophilic anaerobic system. Bioresour. Technol. 2017, 239, 336–344. [Google Scholar] [CrossRef]
  56. Yan, W.; Lu, D.; Liu, J.; Zhou, Y. The interactive effects of ammonia and carbon nanotube on anaerobic digestion. Chem. Eng. J. 2019, 372, 332–340. [Google Scholar] [CrossRef]
  57. Ye, J.; Hu, A.; Ren, G.; Zhou, T.; Zhang, G.; Zhou, S. Red mud enhances methanogenesis with the simultaneous improvement of hydrolysis-acidification and electrical conductivity. Bioresour. Technol. 2018, 247, 131–137. [Google Scholar] [CrossRef]
  58. Baek, G.; Kim, J.; Cho, K.; Bae, H.; Lee, C. The biostimulation of anaerobic digestion with (semi) conductive ferric oxides: Their potential for enhanced biomethanation. Environ. Biotechnol. 2015, 99, 10355–10366. [Google Scholar] [CrossRef]
  59. Pérez, C.; DeGrandpre, M.; Lagos, N.; Saldías, G.; Cascales, E.; Vargas, C. Effect of ferrihydrite biomineralization on methanogenesis in an anaerobic incubation from paddy soil. J. Geophys. Res. Biogeosci. 2015, 120, 673–692. [Google Scholar] [CrossRef]
  60. Jing, Y.; Wan, J.; Angelidaki, I.; Zhang, S.; Luo, G. iTRAQ quantitative proteomic analysis reveals the pathways for methanation of propionate facilitated by magnetite. Water Res. 2017, 108, 212–221. [Google Scholar] [CrossRef]
  61. Yamada, C.; Kato, S.; Ueno, Y.; Ishii, M.; Igarashi, Y. Conductive iron oxides accelerate thermophilic methanogenesis from acetate and propionate. J. Biosci. Bioeng. 2015, 119, 678–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Baek, G.; Kim, J.; Lee, C. A long-term study on the effect of magnetite supplementation in continuous anaerobic digestion of dairy effluent—Enhancement in process performance and stability. Bioresour. Technol. 2016, 222, 344–354. [Google Scholar] [CrossRef]
  63. Li, L.; Tong, Z.; Fang, C.; Chu, J.; Yu, H. Response of anaerobic granular sludge to single-wall carbon nanotube exposure. Water Res. 2015, 70, 1–8. [Google Scholar] [CrossRef]
  64. Ambuchi, J.; Zhang, Z.; Shan, L.; Liang, D.; Zhang, P.; Feng, Y. Response of anaerobic granular sludge to iron oxide nanoparticles and multi-wall carbon nanotubes during beet sugar industrial wastewater treatment. Water Res. 2017, 117, 87–94. [Google Scholar] [CrossRef] [Green Version]
  65. Dang, Y.; Holmes, D.; Zhao, Z.; Woodard, T.; Zhang, Y.; Sun, D.; Wang, L.; Nevin, K.; Lovley, D. Enhancing anaerobic digestion of complex organic waste with carbon-based conductive materials. Bioresour. Technol. 2016, 220, 516–522. [Google Scholar] [CrossRef]
  66. Dang, Y.; Sun, D.; Woodard, T.; Wang, L.; Nevin, K.; Holmes, D. Stimulation of the anaerobic digestion of the dry organic fraction of municipal solid waste (OFMSW) with carbon-based conductive materials. Bioresour. Technol. 2017, 238, 30–38. [Google Scholar] [CrossRef] [PubMed]
  67. Lei, Y.; Sun, D.; Dang, Y.; Chen, H.; Zhao, Z.; Zhang, Y.; Holmes, D. Stimulation of methanogenesis in anaerobic digesters treating leachate from a municipal solid waste incineration plant with carbon cloth. Bioresour. Technol. 2016, 222, 270–276. [Google Scholar] [CrossRef] [PubMed]
  68. Lovley, D. Syntrophy goes electric: Direct interspecies electron transfer. Annu. Rev. Microbiol. 2017, 71, 643–664. [Google Scholar] [CrossRef]
  69. Zhao, Z.; Zhang, Y.; Woodard, T.; Nevin, K.; Lovley, D. Enhancing syntrophic metabolism in up-flow anaerobic sludge blanket reactors with conductive carbon materials. Bioresour. Technol. 2015, 191, 140–145. [Google Scholar] [CrossRef]
  70. Barua, S.; Dhar, B. Advances towards understanding and engineering direct interspecies electron transfer in anaerobic digestion. Bioresour. Technol. 2017, 244, 698–707. [Google Scholar] [CrossRef]
  71. Shen, Y.; Linville, J.; Urgun-Demirtas, M.; Schoene, R.; Snyder, S. Producing pipeline-quality biomethane via anaerobic digestion of sludge amended with corn stover biochar with in-situ CO2 removal. Appl. Energy 2015, 158, 300–309. [Google Scholar] [CrossRef] [Green Version]
  72. Zhao, Z.; Zhang, Y.; Yu, Q.; Dang, Y.; Li, Y.; Quan, X. Communities stimulated with ethanol to perform direct interspecies electron transfer for syntrophic metabolism of propionate and butyrate. Water Res. 2016, 102, 475–484. [Google Scholar] [CrossRef]
  73. Salvador, A.; Martins, G.; Melle-franco, M.; Serpa, R.; Stams, A.; Cavaleiro, A.; Pereira, M.; Alves, M. Carbon nanotubes accelerate methane production in pure cultures of methanogens and in a syntrophic coculture. Environ. Microbiol. 2017, 19, 2727–2739. [Google Scholar] [CrossRef] [Green Version]
  74. Hirano, S.; Matsumoto, N.; Morita, M.; Sasaki, K.; Ohmura, N. Electrochemical control of redox potential affects methanogenesis of the hydrogenotrophic methanogen Methanothermobacter thermautotrophicus. Lett. Appl. Microbiol. 2013, 56, 315–321. [Google Scholar] [CrossRef]
  75. Gu, M.; Yin, Q.; Liu, Y.; Du, J.; Wu, G. New insights into the effect of direct interspecies electron transfer on syntrophic methanogenesis through thermodynamic analysis. Bioresour. Technol. Rep. 2019, 7, 100225. [Google Scholar] [CrossRef]
  76. Ishii, S.; Suzuki, S.; Tenney, A.; Nealson, K.; Bretschger, O. Comparative metatranscriptomics reveals extracellular electron transfer pathways conferring microbial adaptivity to surface redox potential changes. ISME J. 2018, 12, 2844–2863. [Google Scholar] [CrossRef]
  77. Ammar, Y.; Swailes, D.; Bridgens, B.; Chen, J. Influence of surface roughness on the initial formation of biofilm. Surf. Coat. Technol. 2015, 284, 410–416. [Google Scholar] [CrossRef] [Green Version]
  78. Medilanski, E.; Kaufmann, K.; Wick, L.Y.; Wanner, O.; Harms, H. Influence of the surface topography of stainless steel on bacterial adhesion. Biofouling 2002, 18, 193–203. [Google Scholar] [CrossRef]
  79. Habouzit, F.; Gévaudan, G.; Hamelin, J.; Steyer, J.; Bernet, N. Influence of support material properties on the potential selection of Archaea during initial adhesion of a methanogenic consortium. Bioresour. Technol. 2011, 102, 4054–4060. [Google Scholar] [CrossRef]
  80. Martins, G.; Salvador, A.; Pereira, L.; Alves, M. Methane production and conductive materials: A critical review. Environ. Sci. Technol. 2018, 52, 10241–10253. [Google Scholar] [CrossRef] [Green Version]
  81. Yang, Z.; Guo, R.; Shi, X.; Wang, C.; Wang, L.; Dai, M. Magnetite nanoparticles enable a rapid conversion of volatile fatty acids to methane. RSC Adv. 2016, 6, 25662–25668. [Google Scholar] [CrossRef]
  82. Kondaveeti, S.; Min, B. Bioelectrochemical reduction of volatile fatty acids in anaerobic digestion effluent for the production of biofuels. Water Res. 2015, 87, 137–144. [Google Scholar] [CrossRef] [PubMed]
  83. Choi, O.; Sang, B.I. Extracellular electron transfer from cathode to microbes: Application for biofuel production. Biotechnol. Biofuels 2016, 9, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lee, M.; Reddy, C.; Min, B. In situ integration of microbial electrochemical systems into anaerobic digestion to improve methane fermentation at different substrate concentrations. Int. J. Hydrog. Energy 2018, 44, 2–11. [Google Scholar] [CrossRef]
  85. Lee, B.; Park, J.G.; Shin, W.B.; Tian, D.J.; Jun, H.B. Microbial communities change in an anaerobic digestion after application of microbial electrolysis cells. Bioresour. Technol. 2017, 234, 273–280. [Google Scholar] [CrossRef]
  86. Feng, Y.; Zhang, Y.; Chen, S.; Quan, X. Enhanced production of methane from waste activated sludge by the combination of high-solid anaerobic digestion and microbial electrolysis cell with iron-graphite electrode. Chem. Eng. J. 2015, 259, 787–794. [Google Scholar] [CrossRef]
  87. Linji, X.; Wenzong, L.; Yining, W.; Aijie, W.; Shuai, L.; Wei, J. Optimizing external voltage for enhanced energy recovery from sludge fermentation liquid in microbial electrolysis cell. Int. J. Hydrog. Energy 2013, 38, 15801–15806. [Google Scholar] [CrossRef]
  88. Choi, K.; Kondaveeti, S.; Min, B. Bioelectrochemical methane (CH4) production in anaerobic digestion at different supplemental voltages. Bioresour. Technol. 2017, 245, 826–832. [Google Scholar] [CrossRef]
  89. De Vrieze, J.; Gildemyn, S.; Arends, J.; Vanwonterghem, I.; Verbeken, K.; Boon, N.; Verstraete, W.; Tyson, G.; Hennebel, T.; Rabaey, K. Biomass retention on electrodes rather than electrical current enhances stability in anaerobic digestion. Water Res. 2014, 54, 211–221. [Google Scholar] [CrossRef] [Green Version]
  90. Choi, J.; Lee, C. Bioelectrochemical enhancement of methane production in anaerobic digestion of food waste. Int. J. Hydrog. Energy 2018, 44, 1–10. [Google Scholar] [CrossRef]
  91. Hagos, K.; Liu, C.; Lu, X. Effect of endogenous hydrogen utilization on improved methane production in an integrated microbial electrolysis cell and anaerobic digestion: Employing catalyzed stainless steel mesh cathode. Chin. J. Chem. Eng. 2018, 26, 574–582. [Google Scholar] [CrossRef]
  92. Liu, W.; Cai, W.; Guo, Z.; Wang, L.; Yang, C.; Varrone, C.; Wang, A. Microbial electrolysis contribution to anaerobic digestion of waste activated sludge, leading to accelerated methane production. Renew. Energy 2016, 91, 334–339. [Google Scholar] [CrossRef] [Green Version]
  93. Park, J.; Lee, B.; Tian, D.; Jun, H. Bioelectrochemical enhancement of methane production from highly concentrated food waste in a combined anaerobic digester and microbial electrolysis cell. Bioresour. Technol. 2018, 247, 226–233. [Google Scholar] [CrossRef] [PubMed]
  94. Cai, W.; Han, T.; Guo, Z.; Varrone, C.; Wang, A.; Liu, W. Methane production enhancement by an independent cathode in integrated anaerobic reactor with microbial electrolysis. Bioresour. Technol. 2016, 208, 13–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Noori, M.; Vu, M.; Ali, R.; Min, B. Recent advances in cathode materials and configurations for upgrading methane in bioelectrochemical systems integrated with anaerobic digestion. Chem. Eng. J. 2019, 123689. [Google Scholar] [CrossRef]
  96. Deutzmann, J.; Sahin, M.; Spormann, A. Extracellular enzymes facilitate electron uptake in biocorrosion and bioelectrosynthesis. mBio 2015, 6, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Beese-Vasbender, P.; Grote, J.; Garrelfs, J.; Stratmann, M.; Mayrhofer, K. Selective microbial electrosynthesis of methane by a pure culture of a marine lithoautotrophic archaeon. Bioelectrochemistry 2015, 102, 50–55. [Google Scholar] [CrossRef]
  98. Gregory, K.; Bond, D.; Lovley, D. Graphite electrodes as electron donors for anaerobic respiration. Environ. Microbiol. 2004, 6, 596–604. [Google Scholar] [CrossRef]
  99. Thrash, J.; Coates, J. Review: Direct and indirect electrical stimulation of microbial metabolism. Environ. Sci. Technol. 2008, 42, 3921–3931. [Google Scholar] [CrossRef]
  100. Wei, J.; Liang, P.; Huang, X. Recent progress in electrodes for microbial fuel cells. Bioresour. Technol. 2011, 102, 9335–9344. [Google Scholar] [CrossRef]
  101. Guo, K.; Prévoteau, A.; Patil, S.; Rabaey, K. Engineering electrodes for microbial electrocatalysis. Curr. Opin. Biotechnol. 2015, 33, 149–156. [Google Scholar] [CrossRef] [PubMed]
  102. Nelabhotla, A.; Dinamarca, C. Bioelectrochemical CO2 reduction to methane: MES integration in biogas production processes. Appl. Sci. 2019, 9, 1056. [Google Scholar] [CrossRef] [Green Version]
  103. Li, S.; Cheng, C.; Thomas, A. Carbon-based microbial-fuel-cell electrodes: From conductive supports to active catalysts. Adv. Mater. 2016, 29. [Google Scholar] [CrossRef] [PubMed]
  104. Yee, M.; Snoeyenbos-west, O.; Thamdrup, B.; Ottosen, L.; Rotaru, A. Extracellular electron uptake by two methanosarcina species. Front. Energy Res. 2019, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  105. Kouzuma, A.; Kato, S.; Watanabe, K. Microbial interspecies interactions: Recent findings in syntrophic consortia. Front. Microbiol. 2015, 6, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Shrestha, P.; Rotaru, A.; Summers, Z.; Shrestha, M.; Liu, F.; Lovley, D. Transcriptomic and genetic analysis of direct interspecies electron transfer. Appl. Environ. Microbiol. 2013, 79, 2397–2404. [Google Scholar] [CrossRef] [Green Version]
  107. Rice, M.; Soils, P.; Holmes, D.; Shrestha, P.; Walker, D.; Dang, Y.; Nevin, K.; Woodard, T.; Lovley, D. Metatranscriptomic evidence for direct interspecies electron transfer between Geobacter and Methanothrix species in methanogenic rice paddy soils. Appl. Environ. Microbiol. 2017, 83, 1–11. [Google Scholar]
  108. Zhao, Z.; Zhang, Y.; Wang, L.; Quan, X. Potential for direct interspecies electron transfer in an electric- anaerobic system to increase methane production from sludge digestion. Sci. Rep. 2015, 5, 1–12. [Google Scholar] [CrossRef]
  109. Tian, T.; Qiao, S.; Yu, C.; Yang, Y.; Zhou, J. Low-temperature anaerobic digestion enhanced by bioelectrochemical systems equipped with graphene/PPy- and MnO2 nanoparticles/PPy-modified electrodes. Chemosphere 2019, 218, 119–127. [Google Scholar] [CrossRef]
  110. Tian, T.; Qiao, S.; Li, X.; Zhang, M.; Zhou, J. Nano-graphene induced positive effects on methanogenesis in anaerobic digestion. Bioresour. Technol. 2017, 224, 41–47. [Google Scholar] [CrossRef]
  111. Ziels, R.; Nobu, M.; Sousa, D. Elucidating syntrophic butyrate-degrading populations in anaerobic digesters using stable isotope-informed genome-resolved metagenomics. Msystems 2019, 4, 563387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Mei, R.; Nobu, M.; Narihiro, T.; Yu, J.; Sathyagal, A.; Willman, E.; Liu, W. Novel Geobacter species and diverse methanogens contribute to enhanced methane production in media-added methanogenic reactors. Water Res. 2018, 147, 403–412. [Google Scholar] [CrossRef]
  113. Cruz Viggi, C.; Casale, S.; Chouchane, H.; Askri, R.; Fazi, S.; Cherif, A.; Zeppilli, M.; Aulenta, F. Magnetite nanoparticles enhance the bioelectrochemical treatment of municipal sewage by facilitating the syntrophic oxidation of volatile fatty acids. J. Chem. Technol. Biotechnol. 2019, 94, 3134–3146. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, G.; Li, Q.; Gao, X.; Wang, X. Sawdust-derived biochar much mitigates VFAs accumulation and improves microbial activities to enhance methane production in thermophilic anaerobic digestion. ACS Sustain. Chem. Eng. 2018, 7, 2141–2150. [Google Scholar] [CrossRef]
  115. Lee, J.; Lee, S.; Park, H. Enrichment of specific electro-active microorganisms and enhancement of methane production by adding granular activated carbon in anaerobic reactors. Bioresour. Technol. 2016, 205, 205–212. [Google Scholar] [CrossRef] [PubMed]
  116. Xiao, Y.; Yang, H.; Yang, H.; Wang, H.; Zheng, D.; Liu, Y.; Pu, X.; Deng, L. Improved biogas production of dry anaerobic digestion of swine manure. Bioresour. Technol. 2019, 294, 122188. [Google Scholar] [CrossRef] [PubMed]
  117. Xu, S.; Han, R.; Zhang, Y.; He, C.; Liu, H. Differentiated stimulating effects of activated carbon on methanogenic degradation of acetate, propionate and butyrate. Waste Manag. 2018, 76, 394–403. [Google Scholar] [CrossRef]
  118. Ma, W.; Li, H.; Zhang, W.; Shen, C.; Wang, L.; Li, Y.; Li, Q.; Wang, Y. TiO2 nanoparticles accelerate methanogenesis in mangrove wetlands sediment. Sci. Total Environ. 2020, 713, 136602. [Google Scholar] [CrossRef]
  119. Lohner, S.; Deutzmann, J.; Logan, B.; Leigh, J.; Spormann, A. Hydrogenase-independent uptake and metabolism of electrons by the archaeon Methanococcus maripaludis. ISME J. 2014, 8, 1673–1681. [Google Scholar] [CrossRef] [Green Version]
  120. Mayer, F.; Enzmann, F.; Lopez, A.; Holtmann, D. Performance of different methanogenic species for the microbial electrosynthesis of methane from carbon dioxide. Bioresour. Technol. 2019, 289, 121706. [Google Scholar] [CrossRef]
  121. Kracke, F.; Wong, A.B.; Maegaard, K.; Deutzmann, J.; Hubert, M.; Hahn, C.; Jaramillo, T.; Spormann, A. Robust and biocompatible catalysts for efficient hydrogen-driven microbial electrosynthesis. Commun. Chem. 2019, 2, 1–9. [Google Scholar] [CrossRef] [Green Version]
  122. Ketep, S.; Bergel, A.; Calmet, A.; Erable, B. Stainless steel foam increases the current produced by microbial bioanodes in bioelectrochemical systems. Energy Environ. Sci. 2014, 7, 1633–1637. [Google Scholar] [CrossRef] [Green Version]
  123. Zhang, T.; Nie, H.; Bain, T.; Lu, H.; Cui, M.; Snoeyenbos-West, O.; Franks, A.; Nevin, K.; Russell, T.; Lovley, D. Improved cathode materials for microbial electrosynthesis. Energy Environ. Sci. 2013, 6, 217–224. [Google Scholar] [CrossRef] [Green Version]
  124. Du, J.; Catania, C.; Bazan, G.; Du, J.; Catania, C.; Bazan, G. Modification of abiotic-biotic interfaces with small molecules and nanomaterials for improved bioelectronics modification of abiotic-biotic interfaces with small molecules and nanomaterials for improved bioelectronics. Chem. Mater. 2013, 26, 686–697. [Google Scholar] [CrossRef]
  125. Guo, Z.; Gao, L.; Wang, L.; Liu, W.; Wang, A. Enhanced methane recovery and exoelectrogen-methanogen evolution from low-strength wastewater in an up-flow bio film reactor with conductive granular graphite fillers. Front. Environ. Sci. Eng. 2018, 12, 13. [Google Scholar] [CrossRef]
  126. Cerrillo, M.; Viñas, M.; Bonmatí, A. Unravelling the active microbial community in a thermophilic anaerobic digester-microbial electrolysis cell coupled system under different conditions. Water Res. 2017, 110, 192–201. [Google Scholar] [CrossRef] [Green Version]
  127. Xu, S.; He, C.; Luo, L.; Lü, F.; He, P.; Cui, L. Comparing activated carbon of different particle sizes on enhancing methane generation in upflow anaerobic digester. Bioresour. Technol. 2015, 196, 606–612. [Google Scholar] [CrossRef]
  128. Zhao, Z.; Wang, J.; Li, Y.; Zhu, T.; Yu, Q.; Wang, T.; Liang, S.; Zhang, Y. Why do DIETers like drinking: Metagenomic analysis for methane and energy metabolism during anaerobic digestion with ethanol. Water Res. 2019, 171, 115425. [Google Scholar] [CrossRef]
  129. Zhu, Y.; Zhao, Z.; Zhang, Y. Using straw as a bio-ethanol source to promote anaerobic digestion of waste activated sludge. Bioresour. Technol. 2019, 286, 121388. [Google Scholar] [CrossRef]
  130. Zhao, Z.; Li, Y.; He, J.; Zhang, Y. Establishing direct interspecies electron transfer during laboratory-scale anaerobic digestion of waste activated sludge via biological ethanol-type fermentation pretreatment. ACS Sustain. Chem. Eng. 2018, 6, 13066–13077. [Google Scholar] [CrossRef]
  131. Yee, M.; Deutzmann, J.; Spormann, A.; Rotaru, A. Cultivating electroactive microbes—From field to bench. Nanotechnology 2020, 31, 1–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Van Steendam, C.; Smets, I.; Skerlos, S.; Raskin, L. Improving anaerobic digestion via direct interspecies electron transfer requires development of suitable characterization methods. Curr. Opin. Biotechnol. 2019, 57, 183–190. [Google Scholar] [CrossRef] [PubMed]
  133. Neu, J.; Yi, S.; Gu, Y.; O’Brien, J.; Srikanth, V.; Vu, D.; Schmuttenmaer, C.; Malvankar, N. Terahertz-conductivity in biological nanowire-networks. In Proceedings of the 44th International Conference on Infrared, Millimeter, and Terahertz Waves (IRMMW-THz), Paris, France, 1–6 September 2019; pp. 1–2. [Google Scholar]
  134. Malvankar, N.; Vargas, M.; Nevin, K.; Franks, A.; Leang, C.; Kim, B.; Inoue, K.; Mester, T.; Covalla, S.; Johnson, J.; et al. Tunable metallic-like conductivity in microbial nanowire networks. Nat. Nanotechnol. 2011, 6, 573–579. [Google Scholar] [CrossRef]
  135. Shrestha, P.; Rotaru, A.; Comolli, L.; Berkeley, L. Plugging in or going wireless: Strategies for interspecies electron transfer. Front. Microbiol. 2014, 5, 1–8. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, F.; Gu, Y.; O’Brien, J.; Yi, S.; Yalcin, S.; Srikanth, V.; Shen, C.; Vu, D.; Ing, N.; Hochbaum, A.; et al. Structure of microbial nanowires reveals stacked hemes that transport electrons over micrometers. Cell 2019, 177, 361–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Liu, X.; Gao, H.; Ward, J.; Xiaorong, L.; Yin, B.; Fu, T.; Chen, J.; Lovley, D.; Yao, J. Power generation from ambient humidity using protein nanowires. Nature 2020, 577, 1–5. [Google Scholar] [CrossRef]
Figure 1. Indirect interspecies electron transfer (IIET) using acetate (HAc) and/or hydrogen (H2) vs. direct interspecies electron transfer (DIET) in anaerobic digestion.
Figure 1. Indirect interspecies electron transfer (IIET) using acetate (HAc) and/or hydrogen (H2) vs. direct interspecies electron transfer (DIET) in anaerobic digestion.
Processes 08 00424 g001
Figure 2. DIET stimulation mechanisms through (a) supplementation of conductive material (CM) and (b) applying external voltage (EV).
Figure 2. DIET stimulation mechanisms through (a) supplementation of conductive material (CM) and (b) applying external voltage (EV).
Processes 08 00424 g002
Table 1. Gibbs free energies for various reactions, under indirect interspecies electron transfer (IIET) and direct interspecies electron transfer (DIET) conditions.
Table 1. Gibbs free energies for various reactions, under indirect interspecies electron transfer (IIET) and direct interspecies electron transfer (DIET) conditions.
ReactionType of ReactionEquation∆G0 (kJ/mol)Reference
Ethanol to CH4IIETCH3CH2OH + H2O→ CH3COO + H+ + 2H2+9.7[40]
DIETCH3CH2OH + H2O→ CH3COO + 5H+ + 4e−149.6[40]
Propionate to CH4IIETCH3CH2COO + 3H2O→ CH3COO + HCO3 + H+ + 3H2+76.5[40]
DIETCH3CH2COO + 3H2O→ CH3COO + HCO3 + 7H+ + 6e−162.5[40]
CH4 oxidationIIETCH4 + 3H2O→ HCO3 + H+ + 4H2+135.6[40]
DIETCH4 +3H2O→ HCO3 + 9H+ + 8e−183.1[40]
Phenol to CH4IIETC6H5OH + 5H2O→ 3C2H4O2 + 2H2+150.5[29]
DIETC6H5OH + 5H2O→ 3C2H4O2 + 4H+ + 4e−9.0[29]
Oleate to CH4IIETC18H33O2 + 16H2O→ 9C2H3O2 + 15H2 + 8H++340.9[43]
DIETC18H33O2 + 16H2O→ 9C2H3O2 + 38H+ + 30e−641.1[43]
Table 2. Impact of various conductive material (CM) supplementation upon CH4 productivity.
Table 2. Impact of various conductive material (CM) supplementation upon CH4 productivity.
Conductive MaterialParticle Size (μm)ConcentrationSubstrateOperationImprovement/Deterioration (%) aReference
Granular activated carbon1220–143010 g/LGlucoseBatch+168 (CH4 production rate)[55]
Single wall carbon nanotubes0.001–0.0021 g/LGlucoseBatch+92 (CH4 production rate)[55]
SucroseContinuous– −2 (CH4 production)[63]
Carbon nanotubes0.001–0.0025 g/LGlucoseBatch+2 (CH4 yield)[56]
Stainless steel500–200026 g/LSodium lactateContinuous+8–+25 (CH4 production)[47]
Red mud-20 g/LWaste activated sludgeBatch+36 (CH4 production)[57]
Ferric oxyhydroxide-20 mM FeWheyBatch+173 (CH4 production Rate)[58]
Ferrihydrite-25 mMAcetateBatch+15 (CH4 production)[59]
Magnetite0.05–0.705–25 mMWhey, propionate, acetate, butyrateBatch+36–+203 (CH4 production rate)
−4–+44 (CH4 production)
[48,58,60,61,62]
a Improvement/deterioration (%), compared to control.
Table 3. Impact of various voltages’ application upon CH4 productivity.
Table 3. Impact of various voltages’ application upon CH4 productivity.
Electrode TypeApplied Voltage (V vs. SHE)Electrode Surface Area (cm2)SubstrateCH4 Production Improvement (%)Reference
Carbon felt−0.560Waste activated sludge + Molasses50[89]
Carbon fiber brush−1.0ND aGlucose36[84]
−1.1ND aGlucose30[88]
Carbon nanotubes−1.1ND aFood waste20[90]
Carbon cloth + Cobalt phosphorous catalyst−0.812Mixture of glucose, starch granule, beef extract, xylose, and cellulose48[91]
Ti/Ru alloy mesh plates−1.820Sewage sludge1360[32]
Carbon cloth + Pt−0.8162Waste activated sludge200[92]
Fe tube + Graphite pillar−0.345Waste activated sludge22[86]
Graphite carbon + Ni−0.3ND aFood waste70[93]
Titanium mesh−0.340Glycerol60[3]
Stainless steel−0.8ND aGlycose52[94]
a Not determined.
Table 4. The enrichment of methanogenic consortia after direct interspecies electron transfer (DIET) stimulation.
Table 4. The enrichment of methanogenic consortia after direct interspecies electron transfer (DIET) stimulation.
Main Carbon SourceDIET StimulatorEnriched MethanogensMetabolicReference
Butyrate, propionate, and acetateMagnetite nanoparticlesMethanobacterium sp.Hydrogentrophic[113]
GlucoseBioelectrochemical system with graphene/polypyrroleMethanoregula sp.Hydrogentrophic[109]
Food waste and Sewage sludgeSawdust-derived biocharMethanothermobacter sp.Hydrogentrophic[114]
GlucoseMagnetiteMethanobacterium sp.Hydrogentrophic[7]
Acetic acidGranular active carbonMethanospirillaceae sp.Hydrogentrophic[115]
GlycerolMagnetiteMethanomassiliicoccus sp.Hydrogentrophic[3]
Swine manureGranular active carbonMethanosaeta sp.Acetoclastic[116]
GlucoseNano-grapheneMethanosaeta sp.Acetoclastic[110]
Ethanol and glucosePowered activated carbonMethanosarcina sp.Acetoclastic[117]
Sodium acetate, sodium propionate, and sodium butyrateTiO2 nanoparticlesMethanobacterium sp. and Methanosarcina sp.Hydrogentrophic and acetoclastic[118]

Share and Cite

MDPI and ACS Style

Mostafa, A.; Im, S.; Song, Y.-C.; Ahn, Y.; Kim, D.-H. Enhanced Anaerobic Digestion by Stimulating DIET Reaction. Processes 2020, 8, 424. https://doi.org/10.3390/pr8040424

AMA Style

Mostafa A, Im S, Song Y-C, Ahn Y, Kim D-H. Enhanced Anaerobic Digestion by Stimulating DIET Reaction. Processes. 2020; 8(4):424. https://doi.org/10.3390/pr8040424

Chicago/Turabian Style

Mostafa, Alsayed, Seongwon Im, Young-Chae Song, Yongtae Ahn, and Dong-Hoon Kim. 2020. "Enhanced Anaerobic Digestion by Stimulating DIET Reaction" Processes 8, no. 4: 424. https://doi.org/10.3390/pr8040424

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