• Aucun résultat trouvé

Long-term continuous production of H2 in a microbial electrolysis cell (MEC) treating saline wastewater

N/A
N/A
Protected

Academic year: 2021

Partager "Long-term continuous production of H2 in a microbial electrolysis cell (MEC) treating saline wastewater"

Copied!
9
0
0

Texte intégral

(1)

O

pen

A

rchive

T

oulouse

A

rchive

O

uverte

(OATAO)

OATAO is an open access repository that collects the work of some Toulouse

researchers and makes it freely available over the web where possible.

This is an author’s version published in:

http://oatao.univ-toulouse.fr/

20586

To cite this version:

Carmona-Martínez, Alessandro Alfredo and Trably, Eric and Milferstedt, Kim and

Lacroix, Rémy and Etcheverry, Luc and Bernet, Nicolas Long-term continuous

production of H2 in a microbial electrolysis cell (MEC) treating saline wastewater.

(2015) Water Research, 81. 149-156. ISSN 0043-1354

Any correspondance concerning this service should be sent to the repository administrator:

tech-oatao@listes-diff.inp-toulouse.fr

(2)

Long-term continuous production of H

2

in a microbial electrolysis cell

(MEC) treating saline wastewater

Alessandro A. Carmona-Martínez

a

, Eric Trably

a

, Kim Milferstedt

a

, R!

emy Lacroix

b

,

Luc Etcheverry

c

, Nicolas Bernet

a,*

aINRA, UR0050, Laboratoire de Biotechnologie de l'Environnement, Avenue des Etangs, Narbonne F-11100, France b6T-MIC Ing!enieries, 4 rue Brindejonc des Moulinais, 31500 Toulouse, France

cLaboratoire de G!enie Chimique, Universit!e de Toulouse, CNRS, 4 all!ee Emile Monso, 31432 Toulouse, France

Keywords: Biohydrogen

Microbial electrolysis cell Saline wastewater

Scanning electron microscopy High-throughput sequencing Geoalkalibacter

a b s t r a c t

A biofilm-based 4 L two chamber microbial electrolysis cell (MEC) was continuously fed with acetate under saline conditions (35 g/L NaCl) for more than 100 days. The MEC produced a biogas highly enriched in H2(!90%). Both current (10.6 ± 0.2 A/m2Anodeor 199.1 ± 4.0 A/m3MEC) and H2production

(201.1 ± 7.5 LH2/m2Cathode$d or 0.9 ± 0.0 m3H2/m3MEC$d) rates were highly significant when considering the saline operating conditions. A microbial analysis revealed an important enrichment in the anodic biofilm with five main bacterial groups: 44% Proteobacteria, 32% Bacteroidetes, 18% Firmicutes and 5% Spirochaetes and 1% Actinobacteria. Of special interest is the emergence within the Proteobacteria phylum of the recently described halophilic anode-respiring bacteria Geoalkalibacter (unk. species), with a relative abundance up to 14%. These results provide for the first time a noteworthy alternative for the treatment of saline effluents and continuous production of H2.

1. Introduction

A microbial electrolysis cell (MEC) is a well established microbial electrochemical technology (MET) used to produce high amounts of H2(Logan et al., 2008). A MEC profits from the activity of

anode-respiring bacteria (ARB) embedded in anodic biofilms to break down the organic matter and harvest electrons that are chemically used at the abiotic chemical cathode to produce H2. Nevertheless,

the use of a biotic cathode is also possible as recently reported by

Batlle-Vilanova et al. (2014)on the comparison of abiotic and biotic cathodes for microbial electrolysis of H2. Usually, most of the MEC

tests are performed using fresh wastewaters (i.e.: non saline) as anodic solutions that allow the successful development of elec-troactive biofilms from domestic wastewater (WW). Consequently, the application of such developed biofilms in MECs is mainly limited to the treatment of domestic WW. A successful treatment of other types of effluents such as saline WW which are significantly produced by the seafood, petroleum and leather industry will require the development of specific biofilms capable to work under

such halophilic conditions. The enrichment of moderate halophilic biofilms composed of ARB could provide an alternative for the treatment of saline WWs which represent 5% of the WW generated worldwide and for which their treatment is usually limited by technical or economic constraints (Lefebvre and Moletta, 2006). However, another issue that emerges from an extensive overview of the literature on MECs (Logan et al., 2008) is the lack of studies working under saline conditions. Moreover, real WW treatment will require continuously operated MECs that will make necessary the development of a well attached-mature biofilm as an essential prerequisite for long-term operation of ARB biofilm-based METs.

Although the scalability of the MEC technology is of the main interest within the MET-related scientific community only a few attempts have been done to bring the MEC technology from the bench-via the pilot-to the full-scale size. In this context it is worth mentioning the works byEscapa et al. (2015), Gil-Carrera et al. (2013), Heidrich et al. (2013, 2014)who have very recently done significant attempts to scale-up the technology by operating 100 L size MECs fed with real domestic WW. Independently of the drawbacks encountered by these two engineering research groups, they have shown for the first time that the MEC technology is finally capable of converting low strength domestic WW to produce

* Corresponding author.

(3)

a gas stream highly enriched in hydrogen (~100%) and most importantly at very similar operational conditions as the ones of conventional WW treatment plants.

The aim of the present work was to operate a stable biofilm-based MEC in continuous mode for more than 100 days under sa-line conditions. A mature and moderate halophilic anodic ARB biofilm was first enriched from saline anoxic sediments under potentiostatic control during three acetate fed-batch cycles and then, operated under continuous mode. The long-term operation corroborated not only that a mature biofilm had well developed on the anode but also it confirmed the overall stability of the system. At the end of the experiment, the anode was removed and 1) the microbial composition by 16S rDNA sequencing and 2) biofilm attachment by scanning electron microscopy were analyzed. To the best of the authors' knowledge, this is the first work reporting such a long-term performance of a continuously operated MEC under saline conditions.

2. Materials and methods 2.1. General conditions

All chemicals were of analytical or biochemical grade and were purchased from Sigma-Aldrich and Merck. If not stated otherwise, all potentials provided in this manuscript refer to the SCE reference electrode (KCl 3.0 M, þ240 mV vs. SHE, Materials Mates, La Guilleti"ere 38700 Sarcenas, France). All media preparations were adjusted to pH 7.0. Bioelectrochemical experiments were con-ducted under potentiostatic control and incubations were per-formed at 37.0#C.

2.2. Metabolite and biogas analysis

The model substrate was acetate and its concentration was determined by liquid injection into a gas chromatograph (Clarus 580 GC, PerkinElmer). Biogas composition (CH4, CO2, H2 and N2)

was determined using a gas chromatograph (Clarus 580, Perkin Elmer) coupled to thermal catharometer detector, as described elsewhere (Qu!em!eneur et al., 2012). The biogas was periodically analyzed for its H2S content using a second gas chromatograph

(Clarus 480, Perkin Elmer) equipped with an RtUBond-SiOH col-umn and a molecular sieve (RtMolsieve). Operating conditions were as follows: the carrier gas was helium at a pressure of 250.0 kPa under a flow rate of 4.0 mL min$1; temperature of the injector and the detector was fixed at 150.0 #C. Throughout the

manuscript the average gas composition is reported. During Phase II the gas was daily sampled. During Phase III the gas was sampled before increasing/changing the substrate concentration. Averages and standard deviations are calculated taking into account two samples at each point. More details regarding the follow-up of gas production are given inFig. S5 and S6.

2.3. Origin of inoculum

Usually, MEC studies make use of domestic wastewater (i.e., NaCl concentrations % 1.0 g/L) as source of microorganisms. With the objective to operate the MEC under moderate saline conditions (e.g., 35.0 g/L NaCl), saline sediments (10.0% w/v) were used as inoculum (pH of 7.8 ± 0.2, 36.1 ± 3.5 g of volatile solids per gram of sediments and a value of conductivity of 93.6 ± 12.1 mS/cm). Sed-iments were collected in a lagoon receiving wastewaters from a salt factory (Salins de Saint Martin, Gruissan, France: 43#5034.2300, 3#40

57.4700) and stored at room temperature as reported elsewhere

(Pierra et al., 2015).

2.4. Growth medium during fed-batch and continuous operation A specific saline growth medium (35.0 g/L NaCl), namely modified-Starkey medium and previously developed for the se-lection of fermentative bacteria, was used throughout the study to promote the growth of electroactive biofilm on the electrode sur-face (Rafrafi et al., 2013). This medium was used since it has been previously utilized in our research group for successful develop-ment of electroactive biofilms under saline conditions ( Carmona-Martínez et al., 2013).

Modified Starkey medium contained the following (per liter of tap water): 35.0 g NaCl, 0.5 g K2HPO4, 2.0 g NH4Cl, 7.6 g MES buffer,

0.2 g Yeast Extract, 1 mL of oligo-elements solution ( Carmona-Martínez et al., 2013) and different COD concentrations as fed ac-etate (0.64e6.42 g COD/L) of acetic acid as electron donor (see also

Table S2). Although the model substrate used here was acetate, real saline WW is characterized for its high content of organic matter (Lefebvre and Moletta, 2006). The treatment of such WW in a MEC might follow conventional degradation steps: hydrolysis, acido-genesis, acetogenesis and finally, electron harvesting by halophilic ARB. During fed-batch chronoamperometric biofilm growth, the medium was supplemented with 50.0 mM of 2-(N-morpholino) ethanesulfonic acid (MES) to buffer the pH at around 7.0. For continuous MEC operation the medium inlet was adjusted at pH 7.0 with 1 M NaOH. Either during fed-batch or continuous operation, pH of the effluent slightly increased and remained relatively neutral (7.4 ± 0.2).

2.5. Bench scale 4L microbial electrolysis cell (MEC) configuration A scheme of the unstirred up-flow microbial electrolysis cell (MEC) is presented inFig. 1. The MEC body was a polycarbonate cylinder constituted of several parts (from outside to inside): (1) a water jacket (20.0 cm diameter and 30.0 cm height) used to keep a constant temperature of 37.0#C (see cross-section and top view in

Fig. 1); (2) the MEC body (16.0 cm diameter and 30.0 cm height) that contained both anode and cathode, had a total volume of 6.0 L and a final working volume of 4.0 L (i.e., after placing electrodes). As anode material, a flexible 1 cm thick piece of graphite felt (12.0 cm diameter and 20.0 cm height) with a projected surface area of 754.0 cm2 (grade RVG 4000, Mersen, France) was used. The graphite felt was slightly compressed and cylinder-shaped by a welded grid of titanium as electron collector, with two wires piercing the upper part of the MEC body to allow the connection with the potentiostat. An anionic exchange membrane (AEM) was placed between anode and cathode, as shown inFig. S1(FAA-PK, FuMA-Tech GmbH, Germany). The presence of a membrane in a MEC avoids H2recycling at the anode. Furthermore, the use of a

AEM usually leads to a better electrochemical performance in comparison to other ion exchange membranes (Rozendal et al., 2008). As cathode material (188.0 cm2 of catholyte exposed sur-face area), an empty cylindrical stainless steel 254SMO tube with an external diameter of 2.1 cm was located at the center of the MEC (Flowell, France). Through the cathode, a 2.0 mm diameter hole was drilled at the top of the electrode to collect the biogas. The volume of biogas was daily measured by displacement of water in a grad-uated column. Additionally, a SCE reference electrode was used to monitor a constant applied potential of þ200.0 mV (þ444.0 mV vs. SHE) at the anode.

2.6. MEC start-up and operation: phase I, II and III

At all times the anode potential was constantly fixed at þ200.0 mV vs. SCE (þ444.0 mV vs. SHE) instead of adding a specific voltage to the MEC circuit (see pleaseAppendix). During

(4)

Phase I, the MEC was initially operated in fed-batch mode for at least three chronoamperometric cycles to grow an anodic biofilm (Fig. 2). By using a VSP Potentiostat/Galvanostat interfaced to a VMP3B-80 Current Booster unit (BioLogic Science Instruments, France) it was possible to guarantee an homogenous distribution of the potential along the anode electrode. To initiate the biofilm growth, the MEC was first inoculated with 10.0% w/v of saline sediments and then fed with modified Starkey medium supple-mented with a progressive increase of COD concentration as fed acetate: 0.64, 0.96 and 1.28 g/L (Fig. 2). Before every fed-batch cycle, the medium was vigorously flushed with N2gas (purity ! 99.9999,

Linde France S.A.) to provide anaerobic conditions for at least 30.0 min using a commercial air stone (or aquarium bubbler). During these fed-batch cycles, current production was monitored as a direct signal of the development of the electroactive biofilm on the anode. During Phase II and after chronoamperometric proof of the presence of a mature biofilm on the electrode (Fig. 2), the MEC was operated in continuous mode at a flow rate of 7.5 L d$1, cor-responding to an HRT of about 12.0 h. For continuous MEC opera-tion, no special precautions were taken to remove oxygen from the medium and no sediments were added to the medium composi-tion. Finally, during Phase III, an increase in COD concentration as fed acetate from 0.64 to 6.42 g/L was tested to evaluate whether the MEC could improve its overall performance.

2.7. DNA extraction and MiSeq sequencing of anodic biofilm samples

At the end of the long-term continuous operation (after 100.0 days), the MEC was opened and the anodic biofilm was vertically sampled at three different locations (from bottom to top: at 6.0, 12.0 and 18.0 cm) to evaluate the homogeneity of the microbial population regardless of its location on the electrode material (see

Fig. S2). Due to the homogenous distribution of the applied po-tential we obtained an uniform composition along the anode as indicated by the high similarity in CE-SSCP profiles (Fig. S2). The biomass was heat-treated and underwent bead-beating as done in

Rochex et al. (2008)to facilitate DNA extraction and purification with the QIAamp DNA mini kit (QIAGEN, Courtaboeuf, France). DNA amount and purity in extracts was confirmed by spectrophotom-etry (Infinite NanoQuant M200, Tecan, Austria). Extracted DNA was stored at $20#C until further use. The V3-4 region of the 16S rRNA

gene was amplified with the forward primer CTTTCCCTA-CACGACGCTCTTCCGATCTTACGGRAGGCAGCAG and the reverse primer GGAGTTCAGACGTGTGCTCTTCCGATCTTACCAGGGTATCTAA TCCT plus the respective linkers over 30 amplification cycles at an annealing temperature of 65.0#C. In a second PCR reactor of 12

cycles, an index sequence was added using the primers AATGA-TACGGCGACCACCGAGATCTACACTCTTTCCCTACACGAC and CAAG CAGAAGACGGCATACGAGAT-index-GTGACTGGAGTTCAGACGTGT. The resulting PCR products were purified and loaded onto the Illumina MiSeq cartridge according to the manufacturer's in-structions for sequencing of paired 250 bp reads. Sequencing-related work was done at the GeT PlaGe sequencing center of the genotoul life science network in Toulouse, France (get.genotoul.fr). 38100 forward and reverse sequences were retained after assembly and quality checking using an slightly modified version of the Standard Operation Procedure byKozich et al. (2013)in Mothur version 1.33.0. SILVA release 102 as provided bySchloss et al. (2009)

was used for alignment and as taxonomic outline. Using Mothur, representative sequences of deltaproteobacterial OTUs (at the 3.0% level) with a relative abundance of more than two percent were aligned with select deltaproteobacterial sequences (seeTable S1) from the 16S rRNA-based phylogenetic tree harboring all type

Fig. 1. From left to right: diagram of the continuous up-flow two chamber microbial electrolysis cell (not a scale) and schematic representation of hydrogen production in this system. RE and AEM acronyms stand for reference electrode and anionic exchange membrane, respectively. At the cathode, the possible hydrogen producing reaction would be as follows: 2H2O þ 2e$/ 2OH$þH2. Scheme drawn with modifications fromRozendal et al. (2007).

Fig. 2. Phase I: progressive chronoamperometric biofilm development during three fed-batch cycles at increasing acetic acid concentrations and at applied potential of þ200 mV vs. SCE (KCl 3.0 M).

(5)

strains of all species with validly published names up to June 2013 (The All-species Living Tree, release LTPS115 (Yarza et al., 2008)). The software PAUP* (version 4.0b10) was used to infer a phylogeny using the criterion of maximum parsimony (Swofford, 1989). Bootstrap support was calculated using 1500 repetitions. SumTrees (version 3.3.1) of the DendroPy package (version 3.12.0) was used to map bootstrap values to the best phylogeny (Sukumaran and Holder, 2010). Sequences of most abundant OTUs found in the biofilm were deposited in the NCBI Genbank database under the following accession name: “PRJNA273803”.

2.8. Biofilm analysis by scanning electron microscopy (SEM) The anode was removed from the MEC and vertically sampled for SEM after 100 days of operation. Graphite felt electrode samples (0.5 ' 0.5 cm) were carefully fixed in a 2.5% glutaraldehyde solution in a 0.1 M sodium cacodylate buffer (0.1 M, pH 7.4) overnight at 4#C.

After two serial wash steps in sodium cacodylate buffer (0.2 M), samples were dehydrated in a graded ethanol series, desiccated on a Leica EM CPD300 critical point dryer and then coated with Plat-inum on a Leica EM MED020 ion sputter. Finally, electrode samples were observed with a FEI Quanta 250 scanning electron micro-scope, with an accelerating voltage of 5.0 kV.

2.9. Calculations

The performance of the MEC was assessed in terms of geo-metric/volumetric current density and geogeo-metric/volumetric hydrogen production rate as follows. While the working volume of the MEC (ca. 4 L) was considered for both volumetric calculations (in m3H2/m3MEC$d1and A/m3MEC), the projected surface area of the

anode (754.0 cm2) and cathode (188.0 cm2) were used to calculate the geometric current density and geometric H2production rate (in

A/m2Anode and L H2/m2Cathode$d1), respectively. Coulombic effi-ciency was calculated in terms of electrons recovered from the substrate as follows (Logan et al., 2008): CE ¼ (MsI)/(F besq

Dc).

Where Ms(g/mol) is the molecular weight of the substrate, I (A) is

the current, F is the Faraday's constant, bes(4) is the number of

electrons per mol of substrate, q (L/s) is the flow rate and

Dc (g/L) is

the substrate concentration change. The efficiency of the MEC was assessed on the basis of total recovered moles of hydrogen versus the theoretical amount of H2 that should be produced at the

cathode (Logan et al., 2008). Thus, the cathodic hydrogen recovery was calculated as rcat¼

h

H2/hCEfor the MEC operated in continuous

mode during Phase II and III (Batlle-Vilanova et al., 2014). The number of moles of hydrogen was calculated as

h

H2¼(nH2P)/(R T),

where

n

H2(L) is the volume of hydrogen measured, P (atm) is the

atmospheric pressure measured in the laboratory, R ((L$atm)/ (mol$K)) is the ideal gas constant and T (K) is the temperature. The theoretical amount of hydrogen that could be produced from the measured current was calculated as

h

CE¼Cp/(2 F), where Cpis the

total coulombs calculated by integrating the current over time, 2 is the moles of electrons per mole of hydrogen and F is the Faraday's constant.

3. Results and discussion

3.1. Phase I: MEC start-up and development of a moderate halophilic anodic biofilm

In order to obtain a mature anodic biofilm, the up-flow MEC was operated in fed-batch mode for three cycles and the current pro-duction was monitored as sole indicator of the growth of the electroactive biofilm. The moderate halophilic conditions were ensured by feeding the MEC with saline (35.0 g/L NaCl) modified

Starkey medium. The MEC was re-inoculated two more times with 10% of sediments at the beginning of each cycle.Fig. 2shows the chronoamperometric (CA) current production, i.e.: the electro-activity of the biofilm at increasing COD concentration as fed ace-tate ranging from 0.64 to 1.28 g/L. As shown, the biofilm started to grow after approximately 1 day of poising the anode at þ200.0 mV vs. SCE (þ444.0 mV vs. SHE). This was illustrated as an exponential-like current production trend, characteristic of the growth of anode respiring bacteria (ARB) able to produce significant sustained cur-rent densities (j > 1 A/m2) and mature biofilms. Such growth is a major prerequisite for long-term continuous operation of a biofilm-based MET such as an MEC. In the first cycle, a short lag phase of 1 day was observed (Fig. 2). In the 2nd and 3rd fed-batch cycles, the current immediately started as a clear indicator of the presence of an efficient electroactive biofilm on the anode material (see microscopic analysis section).

3.2. Phase II: MEC continuous operation performance

After successful biofilm formation during three batch cycles (Fig. 2), the MEC was switched from batch to continuous feeding mode with an HRT of 12.0 h. The MEC performance of the long-term operation is presented inFig. 3. Since no sediment was added to the medium all along the continuous feeding, the observed current was only resulting from the biofilm activity. Instead of adding a specific voltage to the MEC circuit, a potentiostat was used to constantly poise the anode potential at þ200.0 mV vs. SCE (Nam et al., 2011). Consequently, the observed potential at the cathode slightly oscil-lated around $1363.6 ± 165.5 mV during all continuous operation. Such electrochemical conditions were favorable to H2production at

the cathode. The average gas composition at the cathode was close to 90% H2, i.e. (in %): H287.8 ± 1.5, CO23.7 ± 0.3 and CH42.0 ± 0.1.

The low concentrations of O2(3.2 ± 0.6%) and N2(3.3 ± 0.9%) were

very likely caused by the manual sampling procedure. The minor amounts of detected CH4were probably due to gas diffusion from

the anodic chamber through the membrane. This result clearly points out that the acetate was efficiently converted at the anode to electrons that finally were converted into H2 at the cathode.

Moreover, the amount of biogas being produced at the anode was negligible in comparison to the gas produced at the cathode. Thus, no liquid displacement was observed in the column used to mea-sure biogas. Nonetheless, to conduct a more detailed analysis of the composition of the gas atmosphere at the anodic chamber, gas samples were regularly analyzed. Methane was the main com-pound detected since no methanogenesis inhibitor was added, i.e., (in %): CH468.1 ± 12.7, CO218.1 ± 5.9, H211.7 ± 5.7, O20.3 ± 0.3, N2

1.8 ± 1.4.

Once the continuous flow of medium was enabled (Fig. 3), the overall production gradually increased with a stable performance around days 30e35 with maximal current densities of 2.3 ± 0.0 A/ m2Anodeor 43.5 ± 0.7 A/m3MEC, and hydrogen production rates of

51.4 ± 1.9 LH2/m2Cathode$d or 0.2 ± 0.0 m3H2/m3MEC$d. Interestingly, the initial pH adjustment to neutrality was enough to keep a rela-tively stable neutral environment in the anodic chamber (pH: 7.4 ± 0.2). In MECs where an AEM is used the working principle of the membrane charge transport relies in the transport of (ideally) hydroxyl ions from cathode to anode (Rozendal et al., 2008). Some of the available hydroxyl ions produced due to the reduction of H2O

migrated to the anode chamber where they have slightly increased the anolyte's pH. The stable neutral pH was probably due to the short HRT of 12.0 h that allowed the ARB to convert the substrate into electrons and later into H2without significantly reducing the

overall pH. In contrast, a highly alkaline pH of ca. 11e12 was observed at the cathodic chamber. The pH unbalance observed in our experiments (anode: 7.0e7.5/cathode: 11.0e12.0) was likely

(6)

caused by the incapability of negative charged ions like OH$

to sufficiently cross the membrane and then migrate to the anode chamber. A possible explanation for the loss of membrane perfor-mance might be the formation of biofouling which can block the available pores of the membrane (Erable et al., 2012).

The MEC-based process described here has been used to convert a readily substrate by ARB such acetate into H2under saline

con-ditions. However, to fully demonstrate that such technology could be considered as a novel alternative treatment of saline effluents, our MEC-based process should be tested with real saline waste-water. A work that has shown very promising results and that it is currently under development (Marone et al., 2014).

The coulombic efficiency (CE) of the system was evaluated as well as the cathodic hydrogen recovery (rcat), a well-accepted

parameter to quantify to what extent the measured H2was

actu-ally recovered from the observed current.

In Phase II, the maximum CE and rcatwere found at day 14 of the

continuous operation with a value close to 80.0% and 40.0% (day 14), respectively. This observation indicates that at day 14 i) at least 80.0% of the substrate was biologically recovered as electrons and ii) that 40.0% of those electrons were further converted into H2 at the cathode. Although the COD removal stabilized (from day 30) at 75% both rcatand CE rapidly decreased. Thus, both rcatand CE are not

representative of the continuous operation during Phase II in which the MEC was operated under a single substrate concentration of 1.28 g/L COD in fed acetate. Such rapid decrease and stabilization of CE (~10%) and rcat (close to zero) suggests that i) parallel metabolic reactions occurred as a possible electron sink (Batlle-Vilanova et al., 2014) and ii) there was an important loss of electrons probably due to detected but unquantifiable CH4production (~70%) in the anode

chamber.

3.3. Phase III: increasing acetate concentration test under continuous feeding

After day 54 (seeFig. 3), an increase in substrate concentration was tested to evaluate whether the MEC could improve its overall performance.Fig. 4shows the current density and H2production

rates after 50 days of continuous operation of the MEC under increasing concentrations of acetate (see alsoFig. S3). As expected, the performance of the MEC showed a linear increasing trend depending on the amount of substrate fed. The maxima average values of both current densities (10.6 ± 0.2 A/m2Anode or

199.01 ± 4.0 A/m3MEC) and H2rates (201.1 ± 7.5 L/m2Cathode$d or 0.9 ± 0.0 m3H2/m3MEC$d) were obtained when the MEC was continuously operated with a COD concentration as fed acetate of 6.42 g/L. Usually the current density increases with the addition of higher substrate concentrations until saturation kinetics is observed. When considering that no saturation is reached, it might suggest that the MEC was capable to convert even higher concen-trations of substrate.

During Phase III we observed a similar trend as the one observed in Phase II regarding COD removal, CE and rcat. Although an

important fraction of the COD in fed acetate was removed (50e70%) the values obtained for CE (~20%) and rcat(~20%) indicated again an

undetected gas leakage in the system.

The relatively stable CE (Fig. 4), with the exception of the value at 0.64 g/L COD due to the difference in current density production (seeFig. S3), indicated that the MEC was able to convert the sub-strate into electrons with a stable but low efficiency of 20%. When considering the conversion of substrate into H2 (rcat), it can be

noticed that although the rcatincreased during phase III, the highest

value of rcatwas 24%.

Although rcat possibly increased due to the lack of substrate

limitation, rcatis still considered lower than previously reported

values in the literature. Possible reasons for this relatively low value of rcat(%25%) during this test and after considering that the MEC

was fed with synthetic medium, could be the followings: first, the architectural design of the MEC could have had an impact on the overall substrate conversion. The MEC was fed from the bottom of the system (Fig. 1) and the graphite felt anode was placed in a parallel orientation with respect to the direction of the flow. Thus, it is likely that at higher concentrations of substrate there was not proper mass transfer through the material. Second, it is probable that the graphite felt anode material suffer of biofouling due to its open pore structure that provided not only a habitat for the growth of ARB but also a trap for cellular material.

The findings presented here are in good agreement with pre-vious and current literature in terms of current density (j), H2

production rate, Coulombic efficiency (CE) and cathodic H2

recov-ery (rcat) (Kundu et al., 2013). However, the scarcity of works

reporting the operation of bench scale liter MECs for the production of H2 under continuous flow conditions prohibits an objective

comparison of results (with the few exceptions of pilot scale MECs, e.g.: (Cusick et al., 2011)). Nonetheless, the MEC performance is highly important, especially when considering the long-term

(7)

operation (>100 d) of the MEC operated under saline conditions. These results are even more remarkable when compared with optimized lab scale MECs that are specifically designed to, e.g.: i) substantially increase mass transfer, ii) reduce electrode spacing, iii) enhance the H2 evolution reaction with highly catalytically

active cathode materials or iv) improve the electrodes arrangement (Cheng and Logan, 2011; Jeremiasse et al., 2011; Liang et al., 2011). 3.4. Comparison of the obtained performance with highly scalable MECs

Only a few significant attempts have been done to scale-up the technology by operating 100 L size MECs fed with real domestic WW (Escapa et al., 2015; Gil-Carrera et al., 2013; Heidrich et al., 2013, 2014). When compared our results to the recently reported performance by scalable-MECs one can clearly notice that our MEC has outperformed those reported inTable S3. The main difference among these works and ours is the high conductivity used here (9 S/m) due to the addition of 35.0 g/L of NaCl in comparison to domestic WW that typically has a conductivity of only 1 S/m.

A possible explanation why our MEC outperformed those re-ported inTable S3might be the constant experimental conditions used here, e.g.: (i) potentiostatically fixed anodic applied potential (þ200 mV vs. SCE), (ii) constant temperature (37#C), (iii) neutral

pH conditions (7.0) and (iv) use of a synthetic medium containing a non fermentable substrate such as acetate.

Although current densities and H2production rates reported by

Escapa et al. (2015), Gil-Carrera et al. (2013), Heidrich et al. (2013, 2014)are still considered low but promising (Table S3), these two engineering research groups have shown for the first time that the MEC technology is finally capable of converting real WW into a gas stream highly enriched in H2(~100%) and most importantly at very

similar operational conditions as the ones of conventional WW treatment plants.

3.5. Structure and composition of the anodic bacterial community with an emphasis on the deltaproteobacteria

Using Illumina MiSeq technology, 38100 partial 16S ribosomal DNA gene sequences were obtained out of which 44% were

assigned to the phylum Proteobacteria, 32% to Bacteroidetes, 18% to Firmicutes, 5% to Spirochaetes and 1% to Actinobacteria. An anno-tated abundance table is given in theAppendix as Table S1. Several microorganisms within the deltaproteobacteria class have been proved to possess electroactivity. These microorganisms repeatedly appear in anodic biofilm communities and in well performing mi-crobial electrochemical technologies (METs) that produce current densities ! 1 A/m2(Yates et al., 2012). We selected one represen-tative sequence from each of deltaproteobacterial operational taxonomic units (OTUs) with an abundance of more or equal to 2.0% for the inference of a phylogenetic tree (Fig. 5). The four OTUs in

Fig. 5make up more than 34% of all sequences.

OTU 1 (with an abundance of 14.1%) was closely related to the well-known anode-respiring bacterium (ARB) Geoalkalibacter sub-terraneus. G. subterraneus is capable of attaching to electrode ma-terials under similar experimental conditions as the ones used here, i.e. saline medium, 37.0 #C, pH 7.0, constant applied potential,

Fig. 4. Phase III: continuous MEC subjected to a concentration variation test between days 50 and 100 of operation. Data are based on experiments of at least two chro-noamperometric repetitions with the same electroactive biofilm (seeFig. S3).

Fig. 5. Phylogeny of highly abundant OTUs from this study (relative abundance in parenthesis) and select typestrains of electroactive proteobacteria with an emphasis on deltaproteobacteria. The Firmicute Thermincola was used as outgroup. Only bootstrap support of more than 75% is mapped onto the appropriate nodes. Accession numbers for type strains are listed inTable S1 in the Appendix.

(8)

among others (Carmona-Martínez et al., 2013). Based on its posi-tion within the phylogeny, OUT 1 very likely possesses electroactive properties.

The close relatedness between OTU 4 (8.2%) and Desulfuromonas acetoxidans suggests that OTU 4, as D. acetoxidans, may be an acetate-oxidizing ARB. D. acetoxidans has been previously detected in electroactive biofilms enriched from sediments and preliminary characterized as a pure culture (Bond et al., 2002). Similarly, it was inferred that OUT 10 possesses an electroactive function within the anodic biofilm community as an additional ARB enriched in this study due to the relatedness of OTU 10 (2.1%) to Desulfuromusa ferrireducens and Geopsychrobacter electrodiphilus, two psychro-philic dissimilatory iron and anode respiring bacteria (Holmes et al., 2004; Vandieken et al., 2006). The fact that several ARB belonging to the Deltaproteobacteria class coexisted within the biofilm could have been facilitated by the three dimensional geometry of the graphite felt anode. Its open pore structure may be a trap for cellular material and thus provide spatially separated habitats supporting growth of several ARB.

The inference of electroactivity to OTU 3 (10.0%) was not straightforward from its position in the phylogeny. Instead, or-ganisms grouped into OTU 3 might be sulfate-reducing bacteria as the sequences are phylogenetically closely related to Desulfovibrio halophilus (Caumette et al., 1991) and Desulfovibrio marinus (Ben Dhia Thabet et al., 2007). OTU 3 may be able to oxidize H2and/or

acetate and use sulfate as terminal electron acceptor under mod-erate halophilic conditions. However, this reaction has unlikely happened as only traces of H2S (i.e.: H2S % 0.07 ± 0.01%) were

detected in the anodic chamber. Microbial reduction of the only sulfate source introduced to the reactor by traces in the oligo-elements solution could therefore not explain the quantity of OTU 3 sequences. Nonetheless, the initial 23 g of MES buffer added during Phase I of biofilm growth could have provided an additional source of sulfate for the sulfate-reducing bacteria to persist. 3.6. Qualitative imaging of biofilm deposition on the anode

In order to evaluate the biofilm shape and deposition on the anodic surface, scanning electron microscopy (SEM) was used to conduct a qualitative inspection as described in the experimental section. SEM was chosen over confocal laser scanning microscopy (CLSM) due to the three dimensional arrangement of the electrode material which would prohibit a reliable quantitative analysis.

Fig. 6shows representative SEM pictures of multiple samples of the anodic biofilm developed after more than 100 days of continuous operation (see alsoFig. S4). As expected, most of the carbon felt fibers were covered by amorphous layers. Although these layers did not resemble to those observed in pure cultures or highly enriched

anodic biofilms in which at higher magnifications a characteristic bacterial shape can be observed, the layers do indicate the presence of a microbial biofilm attached to the electrode material.

In the case of pure culture or highly enriched biofilms derived from domestic wastewaters, the bacterial cells usually accumulate as densely packed amorphous layers that completely cover the electrode material. In the present work, saline sediments were used as inoculum source and the medium contained an important amount of NaCl (35.0 g/L). Therefore, the amorphous aggregates observed inFig. 6andFig. S4could be caused not only by a mi-crobial biofilm but also by the organic matter contained in the sediments and/or salt precipitates possibly formed due to the long-term operation. Such deposit could have decreased the total available surface area of the electrode. Nonetheless, when closely analyzing the pictures with the 10

mm scale bar in

Figs. 6andS4, several rod shaped structures can be seen. Such characteristic shape structures resemble those of the main bacteria detected in the microbial analysis of the anodic biofilms. At the 10

mm scale bar

pictures, some kind of web-like extracellular material can be seen and can likely be attributed due to sample preparation as suggested by others (Badalamenti et al., 2013). To affirm whether this extra-cellular material could be classified as a microbial nanowire, further measurements are required to proof that these putative wires actually perform the so called “long-range electron transport” in a catalytically active microbial biofilm (Malvankar and Lovley, 2012). 4. Conclusions

The continuous operation of a MEC for more than 100 days and under saline conditions showed a significant H2production rate. A

biofilm was first enriched in batch mode for three chro-noamperometric cycles. The biogas stream was highly enriched in H2(>90%). When considering that the system was operated under

saline conditions, both current and H2production rates were highly

remarkable and promising for further application. Although the COD removal was important in the system, the electrons from COD were not found either as CE or rcat. Such lost of electrons evidenced

parallel metabolic reactions and an important leak in the system. The microbiological analysis of the biofilm showed a significant selection with mainly Deltaproteobacteria colonizing the anode. This MEC-based process is a promising alternative to other tech-nologies to treat saline wastewaters.

Acknowledgments

This research was financially supported by the French National Research Agency (Project: ANR-09-BioE-10 D!efiH12). A.A.C.M. gratefully acknowledges G. G!evaudan for the processing of biofilm

Fig. 6. Exemplary scanning electron microscopy pictures of the anodic electroactive biofilm observed after more than 100 days of MEC continuous operation. Scale bars are indicated in the bottom right for all pictures.

(9)

samples used for plotting CE-SSCP profiles/pyrosequencing and C. Pouzet from the “Plateforme d'imagerie FR-AIB, TRI-Genotoul” for technical assistance with SEM.

References

Badalamenti, J.P., Krajmalnik-Brown, R., Torres, C.I., 2013. Generation of high current densities by pure cultures of anode-respiring Geoalkalibacter spp. under alka-line and saalka-line conditions in microbial electrochemical cells. mBio 4 (3), e00144e00113.

Batlle-Vilanova, P., Puig, S., Gonzalez-Olmos, R., Vilajeliu-Pons, A., Ba~neras, L., Balaguer, M.D., Colprim, J., 2014. Assessment of biotic and abiotic graphite cathodes for hydrogen production in microbial electrolysis cells. Int. J. Hydrog. Energy 39 (3), 1297e1305.

Ben Dhia Thabet, O., Fardeau, M.L., Suarez-Nu~nez, C., Hamdi, M., Thomas, P., Ollivier, B., Alazard, D., 2007. Desulfovibrio marinus sp. nov., a moderately halophilic sulfate-reducing bacterium isolated from marine sediments in Tunisia. Int. J. Syst. Evol. Microbiol. 57 (9), 2167e2170.

Bond, D.R., Holmes, D.E., Tender, L.M., Lovley, D.R., 2002. Electrode-reducing mi-croorganisms that harvest energy from marine sediments. Science 295 (5554), 483e485.

Carmona-Martínez, A.A., Pierra, M., Trably, E., Bernet, N., 2013. High current density via direct electron transfer by the halophilic anode respiring bacterium Geo-alkalibacter subterraneus. Phys. Chem. Chem. Phys. 15 (45), 19699e19707.

Caumette, P., Cohen, Y., Matheron, R., 1991. Isolation and characterization of Desulfovibrio halophilus sp. nov., a halophilic sulfate-reducing bacterium iso-lated from Solar Lake (Sinai). Syst. Appl. Microbiol. 14 (1), 33e38.

Cheng, S., Logan, B.E., 2011. High hydrogen production rate of microbial electrolysis cell (MEC) with reduced electrode spacing. Bioresour. Technol. 102 (3), 3571e3574.

Cusick, R., Bryan, B., Parker, D., Merrill, M., Mehanna, M., Kiely, P., Liu, G., Logan, B., 2011. Performance of a pilot-scale continuous flow microbial electrolysis cell fed winery wastewater. Appl. Microbiol. Biotechnol. 89 (6), 2053e2063.

Erable, B., F!eron, D., Bergel, A., 2012. Microbial catalysis of the oxygen reduction reaction for microbial fuel cells: a review. ChemSusChem 5 (6), 975e987.

Escapa, A., San-Martín, M.I., Mateos, R., Mor!an, A., 2015. Scaling-up of membrane-less microbial electrolysis cells (MECs) for domestic wastewater treatment: bottlenecks and limitations. Bioresour. Technol. 180 (0), 72e78.

Gil-Carrera, L., Escapa, A., Carracedo, B., Mor!an, A., G!omez, X., 2013. Performance of a semi-pilot tubular microbial electrolysis cell (MEC) under several hydraulic retention times and applied voltages. Bioresour. Technol. 146 (0), 63e69.

Heidrich, E.S., Dolfing, J., Scott, K., Edwards, S.R., Jones, C., Curtis, T.P., 2013. Pro-duction of hydrogen from domestic wastewater in a pilot-scale microbial electrolysis cell. Appl. Microbiol. Biotechnol. 97 (15), 6979e6989.

Heidrich, E.S., Edwards, S.R., Dolfing, J., Cotterill, S.E., Curtis, T.P., 2014. Performance of a pilot scale microbial electrolysis cell fed on domestic wastewater at ambient temperatures for a 12 month period. Bioresour. Technol. 173 (0), 87e95.

Holmes, D.E., Nicoll, J.S., Bond, D.R., Lovley, D.R., 2004. Potential role of a novel psychrotolerant member of the family Geobacteraceae, geopsychrobacter electrodiphilus gen. Nov., sp. nov., in electricity production by a marine sedi-ment fuel cell. Appl. Environ. Microbiol. 70 (10), 6023e6030.

Jeremiasse, A.W., Bergsma, J., Kleijn, J.M., Saakes, M., Buisman, C.J.N., Cohen Stuart, M., Hamelers, H.V.M., 2011. Performance of metal alloys as hydrogen evolution reaction catalysts in a microbial electrolysis cell. Int. J. Hydrog. Energy 36 (17), 10482e10489.

Kozich, J.J., Westcott, S.L., Baxter, N.T., Highlander, S.K., Schloss, P.D., 2013. Devel-opment of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq illumina sequencing platform. Appl. Environ. Microbiol. 79 (17), 5112e5120.

Kundu, A., Sahu, J.N., Redzwan, G., Hashim, M.A., 2013. An overview of cathode material and catalysts suitable for generating hydrogen in microbial electrolysis cell. Int. J. Hydrog. Energy 38 (4), 1745e1757.

Lefebvre, O., Moletta, R., 2006. Treatment of organic pollution in industrial saline wastewater: a literature review. Water Res. 40 (20), 3671e3682.

Liang, D.-W., Peng, S.-K., Lu, S.-F., Liu, Y.-Y., Lan, F., Xiang, Y., 2011. Enhancement of hydrogen production in a single chamber microbial electrolysis cell through anode arrangement optimization. Bioresour. Technol. 102 (23), 10881e10885.

Logan, B.E., Call, D., Cheng, S., Hamelers, H.V.M., Sleutels, T.H.J.A., Jeremiasse, A.W., Rozendal, R.A., 2008. Microbial electrolysis cells for high yield hydrogen gas production from organic matter. Environ. Sci. Technol. 42 (23), 8630e8640.

Malvankar, N.S., Lovley, D.R., 2012. Microbial nanowires: a new paradigm for bio-logical electron transfer and bioelectronics. ChemSusChem 5 (6), 1039e1046.

Marone, A., Carmona-Martínez, A.A., Sire, Y., Trably, E., Bernet, N., Steyer, J.-P., 2014. Treatment of olive brine wastewater by bioelectrochemical systems, Fundaci!on General de la Universidad de Alcal!a. University of Alcal!a.

Nam, J.-Y., Tokash, J.C., Logan, B.E., 2011. Comparison of microbial electrolysis cells operated with added voltage or by setting the anode potential. Int. J. Hydrog. Energy 36 (17), 10550e10556.

Pierra, M., Carmona-Martínez, A.A., Trably, E., Godon, J.-J., Bernet, N., 2015. Specific and efficient electrochemical selection of Geoalkalibacter subterraneus and Desulfuromonas acetoxidans in high current-producing biofilms. Bio-electrochemistry.http://dx.doi.org/10.1016/j.bioelechem.2015.02.003.

Qu!em!eneur, M., Hamelin, J., Barakat, A., Steyer, J.-P., Carr"ere, H., Trably, E., 2012. Inhibition of fermentative hydrogen production by lignocellulose-derived compounds in mixed cultures. Int. J. Hydrog. Energy 37 (4), 3150e3159.

Rafrafi, Y., Trably, E., Hamelin, J., Latrille, E., Meynial-Salles, I., Benomar, S., Giudici-Orticoni, M.-T., Steyer, J.-P., 2013. Sub-dominant bacteria as keystone species in microbial communities producing bio-hydrogen. Int. J. Hydrog. Energy 38 (12), 4975e4985.

Rochex, A., Godon, J.-J., Bernet, N., Escudi!e, R., 2008. Role of shear stress on composition, diversity and dynamics of biofilm bacterial communities. Water Res. 42 (20), 4915e4922.

Rozendal, R.A., Hamelers, H.V.M., Molenkamp, R.J., Buisman, C.J.N., 2007. Perfor-mance of single chamber biocatalyzed electrolysis with different types of ion exchange membranes. Water Res. 41 (9), 1984e1994.

Rozendal, R.A., Sleutels, T.H.J.A., Hamelers, H.V.M., Buisman, C.J.N., 2008. Effect of the Type of Ion Exchange Membrane on Performance, Ion Transport, and pH in Biocatalyzed Electrolysis of Wastewater, pp. 1757e1762.

Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann, M., Hollister, E.B., Lesniewski, R.A., Oakley, B.B., Parks, D.H., Robinson, C.J., Sahl, J.W., Stres, B., Thallinger, G.G., Van Horn, D.J., Weber, C.F., 2009. Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl. Environ. Microbiol. 75 (23), 7537e7541.

Sukumaran, J., Holder, M.T., 2010. DendroPy: a Python library for phylogenetic computing. Bioinformatics 26 (12), 1569e1571.

Swofford, D.L., 1989. Phylogenetic Analysis Using Parsimony. Illinois Natural History Survey, Champaign.

Vandieken, V., Mußmann, M., Niemann, H., Jørgensen, B.B., 2006. Desulfuromonas svalbardensis sp. nov. and Desulfuromusa ferrireducens sp. nov., psychrophilic, Fe(III)-reducing bacteria isolated from Arctic sediments, Svalbard. Int. J. Syst. Evol. Microbiol. 56 (5), 1133e1139.

Yarza, P., Richter, M., Peplies, J., Euzeby, J., Amann, R., Schleifer, K.-H., Ludwig, W., Gl€ockner, F.O., Rossell!o-M!ora, R., 2008. The all-species living tree project: a 16S rRNA-based phylogenetic tree of all sequenced type strains. Syst. Appl. Micro-biol. 31 (4), 241e250.

Yates, M.D., Kiely, P.D., Call, D.F., Rismani-Yazdi, H., Bibby, K., Peccia, J., Regan, J.M., Logan, B.E., 2012. Convergent development of anodic bacterial communities in microbial fuel cells. ISME J. 6 (11), 2002e2013.

Figure

Fig. 2. Phase I: progressive chronoamperometric biofilm development during three fed-batch cycles at increasing acetic acid concentrations and at applied potential of þ200 mV vs
Fig. 5. Phylogeny of highly abundant OTUs from this study (relative abundance in parenthesis) and select typestrains of electroactive proteobacteria with an emphasis on deltaproteobacteria

Références

Documents relatifs

Le but des recherches était, dans un premier temps, la caractérisation des DIES, en référence à des cadres épistémologiques (Comment on peut construire des concepts en classe ?

ب ( ﺎﯾدﯾﻣوﻛﻟا : موﻬﻔﻣ ﻰﻟإ جوﻟوﻟا لﺑﻗ ﻰﻟإ رﯾﺷﻧ نأ بﺟﯾ ﺎﯾدﯾﻣوﻛﻟا لﺻأ ﻩذﻫ ﺔﻣﻠﻛﻟا » ﺔﯾﻧﺎﻧوﯾ Comédie ، سوزوﯾﻧﯾد ﻪﻟﻹا بﻛاوﻣ

The model can predict the dynamics of the shape and structure (fine root density, coarse root topology and biomass) of the root system either independently of soil conditions

It maps surfaces defined in the Cartesian rectangular domain, such as a tensor-product surface, to polar coordinate parametric domain without loss of continuity.. On the other hand,

Our study has three aims: (1) To document the distribution of exotic ant species in New Caledonia in relation to different habitat types and key habitat variables; (2) To assess

(3)- Dr Ghaouti benmellha ,Eléments du Droit Algerien de la Famille, Tome premier, Le mariage et sa dissolution , Office de Publications universitaires, publi sud, Pari, 1985,

17 The focus on this set of 198 clones confirmed as active hits with simple aliphatic substrates seemed plausible for the fol- lowing reasons: firstly because of the

Sur le site d’étude, ce changement de paradigme est bien visible à partir de 1997, date de création de la Réserve Naturelle Nationale (RNN) de l’île du Rohrschollen, gérée