DataPDF Available
1
Supporting information
Periodic polarization of electroactive biofilms
increases current density and charge carriers
concentration while modifying biofilm structure
Xu Zhanga, Antonin Prévoteaua,*, Ricardo O. Lourob, Catarina M. Paqueteb, Korneel
Rabaeya,*
a. Center for Microbial Ecology and Technology (CMET), Ghent University, Coupure Links
653, 9000, Ghent, Belgium.
b. Universidade Nova de Lisboa, Av da Republica (EAN), 2780-157 Oeiras, Portugal.
Corresponding Authors
*Phone: 32 (0)9 264 59 76 ; e-mail: Korneel.Rabaey@UGent.be.
*Phone: 32 (0)9 264 59 73 ; e-mail: Antonin.Prevoteau@UGent.be.
Pages:25
Figures: 20
Table: 2
2
Table of Contents
Fig. S1. Scheme of the 8-electrode reactor and uncompensated resistance (Ru) recorded
for the 8-electrode (n = 5 measurements per electrode). .................................................... 3
Fig. S2. Quasi-saturation in acetate. Impact of rotating speed and acetate concentration
on the steady state current of a continuously polarized EAB. ............................................ 4
Fig. S3. Limited impact of capacitive current in the total amount of charge recorded per
half-period of polarization (Qcycle). ..................................................................................... 5
Fig. S4. Chronoamperograms recorded during periodic polarization for the triplicates
EAB-10s, after 10 days of growth. ..................................................................................... 6
Table S5. Qcycle and Γ for the EABs under periodic polarization. ...................................... 7
Fig. S6. Growth of mature EABs for the preliminary experiment. .................................... 8
Fig. S7. Electrochemical responses of mature EABs with different tc. .............................. 9
Fig. S8. Evolution of OCPf and estimated fraction of redox cofactors reduced at the
electrode interface (CR/CT) with different tc. .................................................................... 10
Fig. S9. Cyclic voltammograms of mature EABs for preliminary experiment. ............... 11
Fig. S10. Schematized, conjectured concentration profiles for the reduced form of the
redox charge carrier .......................................................................................................... 12
Fig. S11. Representative transient current at the start of discharge for the different
periodic polarizations. ....................................................................................................... 13
Fig. S12. Recording the apparent steady-state catalytic current for EABs grown under
periodic polarization. ........................................................................................................ 14
Fig. S13. Charge surface coverage estimation (Γ, in mole- cm2)..................................... 15
Fig. S14. Representative non-turnover CVs of EABs grown under continuous (control)
and periodic polarization. .................................................................................................. 16
Fig. S15. Proportional relationship between jss and Γ (top); and between jss and C
(bottom). ............................................................................................................................ 17
Fig. S16. Representative Cottrell plots for the discharge of the different EABs. ............. 19
Fig. S17. SDS-PAGE gel (15 % acrylamide) with denatured DNA samples for each EAB.
........................................................................................................................................... 20
Fig. S18. Ratio of heme over total protein content (top) and optical absorption spectra of
heme from the different EABs (bottom). .......................................................................... 21
Fig. S19. Calibration curves of total protein. .................................................................... 22
Table S20. Morphologic characteristics of the EABs. ..................................................... 23
Fig. S21. Microbial community composition of EABs at the end of the 10 days operation.
........................................................................................................................................... 24
Fig. S22. Fraction of Geobacter determined by Illumina sequencing of different EABs. 26
3
Fig. S1. Scheme of the 8-electrode reactor and uncompensated resistance (Ru) recorded for
the 8-electrode (n = 5 measurements per electrode).
Considering the non-central position of the reference electrode in the reactor, the
uncompensated resistance between each working electrode and the reference electrode was
estimated by applying a current interruption method (Bard and Faulkner 2001). Ru was similar
for each electrode, between 13 and 20 Ω. Considering the maximum current generation of
EABs (Imax ~ 2 mA), the maximum difference in ohmic drop (= Imax × Ru) was small (7 Ω × 2
mA = 14 mV) and could not substantially impact the electrochemical measurements since the
potential of polarization (− 0.1 V) is well above the potential at which the anodic plateau is
reached (around − 0.2 V, see Fig. S3). The high reproducibility of electrochemical recordings
between randomly distributed triplicates further confirms that the electrode position did not
impact the measurements.
Reactor 1
Ru / Ω
Electrode
Ru / Ω
E1
13.1 ± 0.9
E5
20.0 ± 0.4
E2
15.8 ± 0.8
E6
17.4 ± 0.8
E3
19.3 ± 0.8
E7
17.3 ± 1.1
E4
18.8 ± 0.9
E8
13.8 ± 0.7
Reactor 2
Ru / Ω
Electrode
Ru / Ω
E1
13.9 ± 0.9
E5
20.2 ± 1.3
E2
17.0 ± 1.0
E6
19.8 ± 1.1
E3
18.5 ± 1.1
E7
17.5 ± 0.6
E4
19.3 ± 0.9
E8
14.2 ± 1.3
4
1
2
Fig. S2. Quasi-saturation in acetate. Impact of rotating speed and acetate concentration on the
3
steady state current of a continuously polarized EAB.
4
The steady state current increased about 2.6 % when the rotation speed of the magnetic stirrer
5
was increased from 100 to 500 rpm (top). The current declined by 3.2 % after adding 7 mM
6
acetate, from 23 to 30 mM (bottom). These small relative impacts of convection and acetate
7
addition (even negative in the latter case) illustrate the quasi-saturation of acetate for the
8
reaction rate.
9
5
10
11
12
Fig. S3. Limited impact of capacitive current in the total amount of charge recorded per half-
13
period of polarization (Qcycle).
14
The applied polarization potential of an EAB-10s was decreased from + 0.3 V to 0.1 V with
15
a negative increment of 0.1 V every 3 periods (panel (a), the red lines show the polarization
16
half-periods). The polarization potential did not impact the value of the OCPf which stays
17
constant at 0.50 V (dotted black line), suggesting that the applied potential did not impact
18
significantly the charging of the EAB during the OCP half-period. This was expected since all
19
the applied potentials are located on the plateau current corresponding to the electron mass-
20
transfer limitation (see panel (c)).
21
The corresponding amounts of charges Qcycle recorded for 10 s of polarization at the different
22
potentials are plotted on panel (b). Qcycle increased quasi-linearly with the polarization
23
potential because of the increasing capacitive current, itself dependent of the amplitude of the
24
potential step between OCPf (− 0.5 V) and the applied potential. However, the variation of
25
Qcycle with the applied potential is particularly small, increasing only by 0.46 %, when the
26
applied potential increased from − 0.1 V to + 0.3 V. This proves that the extreme majority of
27
the charges collected by the electrode is faradaic and originates from the metabolism of the
28
EAB.
29
6
30
Fig. S4. Chronoamperograms recorded during periodic polarization for the triplicates EAB-
31
10s, after 10 days of growth.
32
The CAs shows 3 successive periods. The CAs for the 3 EABs are superimposed, showing a
33
high level of reproducibility within triplicates. This reproducibility is representative of the
34
data collected for the different groups of EABs.
35
0
1
2
3
020 40 60
j/ mA cm−2
t/ s
E4
E5
E6
7
EAB-1s
EAB-10s
EAB-60s
EAB-300s
Qcycle [mC.cm2]
0.90 ± 0.03
12.33 ± 0.12
4.70 ± 0.25
22.02 ± 0.64
Γ [108 mole-.cm2]
3.08 ± 0.11
3.73 ± 0.17
0.26 ± 0.02
0.19 ± 0.01
Qcycle
F×Γ×T1/2 [s1]
0.30 ± 0.00
0.34 ± 0.02
0.31 ± 0.01
0.39 ± 0.01
36
Table S5. Qcycle and Γ for the EABs under periodic polarization.
37
Qcycle is the amount of charge collected per surface of electrode during one respective half-
38
period of polarization (turnover condition). Γ is the surface coverage in electrically connected
39
electron carriers of the EABs (obtained under non-turnover conditions). All values are
40
averaged for the four last days of recording, when a quasi-steady state was reached. F is
41
Faraday constant. The ratio [Qcycle/(F. Γ.T1/2)] therefore corresponds to the amount of the EAB
42
charge capacity (in coulomb storable per EAB) discharged by the EAB into the anode during
43
1 s of average discharge. Interestingly, this amount was always comprised between 0.3 and
44
0.39 for these EABs under periodic polarization.
45
46
8
47
Fig. S6. Growth of mature EABs for the preliminary experiment.
48
Chronoamperograms of the 8 electrodes grown at 0.1 V vs. Ag/AgCl with initially 24 mM
49
acetate as substrate. All 8 electrodes showed good reproducibility of current generation. The
50
catholyte was replaced daily to avoid an excessive pH increase which might degrade the ion
51
exchange membrane. Anolyte was refreshed once after current decayed below ~20% of
52
maximum. Different charging time measurements were performed at the current maximum in
53
the second batch phase (~ 250 h).
54
55
56
57
58
59
60
9
61
Fig. S7. Electrochemical responses of mature EABs with different tc.
62
(a) Evolution of OCP during a 60 s charge under turnover condition; (b) Discharge
63
chronoamperograms for tc = 5 s (orange circle) and tc = 300 s (red circle).
64
During the charging time interval, the OCP of the electrode quickly decreases because of the
65
charge accumulation at the vicinity of the electrode due to reduction of the charge carriers (a).
66
Discharge currents after 5s or 300 s of charging decayed sharply, which reflected a quick
67
release of stored electrons. The minimum current of transient curves after 300 s charge is
68
smaller than final current after 60 s discharge, which suggests a longer recovery time of
69
metabolic activity after longer charging time.
70
71
10
72
Fig. S8. Evolution of OCPf and estimated fraction of redox cofactors reduced at the electrode
73
interface (CR/CT) with different tc.
74
Performed on mature EABs previously grown under continuous polarization. The proportion
75
of redox cofactors reduced at the electrode interface was estimated via the Nernst equation
76
assuming a single redox cofactor of E1/2 = − 0.35 V vs. Ag/AgCl (Fig. S2).
77
11
78
Fig. S9. Cyclic voltammograms of mature EABs for preliminary experiment.
79
The turnover (a) and nonturnover (b) CVs were recorded at 50 mV s1. The turnover CV
80
presented a typical sigmoidal shape. The apparent midpoint potential (E1/2) of the mature EAB
81
was determined by the first derivative and half-wave potential of the turnover CV at − 0.35 V
82
vs. Ag/AgCl, which is consistent with previous studies (Virdis et al. 2016; Zhang et al. 2017).
83
84
12
85
Fig. S10. Schematized, conjectured concentration profiles for the reduced form of the redox
86
charge carrier (Cred) during a transition from a polarization at − 0.1 V vs Ag/AgCl (purple) to
87
a charge under OCP (under turnover conditions). The scheme assumes a constant
88
concentration along the EAB = [Cred] + [Cox], where Cox represents the oxidized form of the
89
charge carrier (Schrott et al.).
90
The 𝑥 axis shows the normal distance from the electrode surface, within the EAB. At t 0
91
(polarization), a steady-state concentration gradient is expected (purple line), with all the
92
redox cofactors oxidized at the electrode surface x = 0 (plateau current of the polarization
93
curve i.e. electron transport limitation). Once the polarization is switched to OCP, the electron
94
sink (anode) is turned off. With increasing time of charge, the concentration profiles should
95
flatten (by electron diffusion, initially toward the electrode interface) and increase (by
96
microbial metabolic activity along the EAB, providing new electrons to the redox matrix). It
97
can be reasonably conjectured that the fraction of reduced cofactor at x = 0 (estimated by the
98
Nernst equation) gives an estimation of a minimum state of charge of the EAB.
99
100
13
101
Fig. S11. Representative transient current at the start of discharge for the different periodic
102
polarizations.
103
The corresponding half-periods of charge/discharge are stated on their respective panel. The
104
CAs were recorded after 220 h of growth under intermittent operation.
105
Note that the local minimum in the transient current for the EAB with T1/2 of 60 s occurs
106
before the one for the EAB with T1/2 of 300 s (1.2 s vs. 1.9 s, respectively). This suggests a
107
longer recovery of the microbial (electro)activity when the microbial metabolism was
108
interrupted for a longer time (once the EAB is fully charged).
109
110
14
111
Fig. S12. Recording the apparent steady-state catalytic current for EABs grown under
112
periodic polarization.
113
The steady-state catalytic current of each electrode was recorded daily, after 300 s of
114
chronoamperometry at 0.1 V vs. Ag/AgCl. The measurement allows an objective
115
comparison of the electroactivity of the different EABs because the initial current of discharge
116
is transient. The respective periodic polarizations are restarted at the end of the recording. The
117
figure shows one representative chronoamperogram for each kind of EAB recorded after 220
118
h of growth.
119
15
120
Fig. S13. Charge surface coverage estimation (Γ, in mole- cm2).
121
The amount of charges storable in the EABs was determined under nonturnover conditions by
122
background subtracted chronocoulometry monitoring a full EAB discharge at − 0.1 V (after
123
having fully charged the EAB at − 0.6 V vs. Ag/AgCl for 300 s) (Zhang et al. 2017).
124
16
125
Fig. S14. Representative non-turnover CVs of EABs grown under continuous (control) and
126
periodic polarization.
127
Recorded at 5 mV s1 at the end of the experiment (10 days). The substantial increase of the
128
redox features for EAB-1s and EAB-10s already suggests a higher concentration of charge
129
carriers and/or a faster electron transport within these EAB when compared to the control.
130
The opposite is true for EAB-60s and EAB-300s.
131
17
132
133
134
135
136
Fig. S15. Proportional relationship between jss and Γ (top); and between jss and C (bottom).
137
Values of jss are the one averaged for the four last days of operation, when performances were
138
stabilized. Values of Γ and C were obtained at the end of the experiment (10 days).
139
This first order kinetics with respect to C suggests that the concentration of charge carriers
140
was a limiting factor for the production of the biocatalytic current. Assuming an analogy with
141
a model of redox conduction for oxidoreductase entrapped in redox polymers (Bartlett and
142
Pratt 1995), and in our particular case of substrate saturation, this would mean that the main
143
rate determining step for the value of jss was either:
144
(i) the electron transfer from the biocatalyst to the extracellular redox matrix. In this case
145
most redox cofactors along the EAB are almost exclusively oxidized since the
146
18
kinetics of electron transport along the EAB would be much faster in comparison (flat
147
steady state concentration profile where [Cred] ~ 0 and [Cox] ~ C). This case
148
corresponds to the kinetics regime I described by Bartlett and Pratt in their model
149
(Bartlett and Pratt 1995). An experimental observation of the case (i) could be the thin,
150
still growing EAB observed by Liu and Bond where most c-type cytochrome were
151
recorded as oxidized along the EAB (by spectroelectrochemistry) (Liu and Bond
152
2012).
153
(ii) A second case is when the current is simultaneously limited by the kinetics of electron
154
transfer from the biocatalyst to the redox matrix and by the kinetics of electron
155
transport along the redox matrix, both mechanisms being interrelated. In this case the
156
concentration profile of the reduced cofactor substantially increases with the normal
157
distance from the electrode (see above a representative schematized example: purple
158
curve in Figure S10, or a real measurement across a thick, mature EAB (Liu and
159
Bond 2012) performed by Liu and Bond). This case corresponds to the kinetics
160
regime II described by Bartlett and Pratt in their model (Bartlett and Pratt 1995).
161
Under substrate saturation and in a well-buffered medium, the case (i) would imply a
162
continuous growth of the EAB thickness (until it would finally reach the case (ii)!). Since
163
EABs typically reach a maximal thickness, and since a substantial redox gradient has been
164
commonly observed across mature EABs, we believe that the case (ii) appears more
165
reasonable for mature EABs having reached their maximum current. That said, this very short
166
discussion was based on an analogy with a “simple” model of redox conduction for hydrogels
167
where biocatalysts and redox cofactors are homogeneously distributed. Anodic EABs
168
certainly show another level of complexity, including the very possible implication of pili in
169
the electron transport processes.
170
19
171
Fig. S16. Representative Cottrell plots for the discharge of the different EABs.
172
The charge transport parameters (C × Dapp1/2) were extracted from Cottrell slopes, as
173
previously described (Zhang et al. 2017).
174
20
175
Fig. S17. SDS-PAGE gel (15 % acrylamide) with denatured DNA samples for each EAB.
176
The column on the right contains the protein markers (NZYColour Protein Marker II from
177
NZYTECH). The samples are control EABs from reactor 1 (A) and reactor 2 (B); EABs-1s
178
(C), EABs-10s (D), EABs-60s (E) and EABs-300s (F). Gel was revealed with heme staining
179
to show c-type cytochromes.
180
21
181
182
Fig. S18. Ratio of heme over total protein content (top) and optical absorption spectra of
183
heme from the different EABs (bottom).
184
The molar absorption coefficient was ε410nm = 125,000 M1 cm1 in 100 mM sodium
185
phosphate buffer, pH = 7.6.
186
22
187
Fig. S19. Calibration curves of total protein.
188
By using Pierce BCA Protein Assay Kit, the working range is from 5 to 250 µg mL1 (SD
189
stands for standard deviation with n = 3).
190
The working reagent was prepared by mixing part A (50 parts) with part B (1 part) from BCA
191
reagent kit. Then 80 µL of sample or bovine serum albumin (BSA) as standard was added into
192
1.6 mL working reagent and mixed well. A water bath was used to heat samples at 60 °C for
193
30 min (enhanced procedure of working range 5 ~ 250 µg mL1). After cooling down to room
194
temperature, the absorbance of all samples and standards were measured at 562 nm.
195
196
197
198
199
200
201
202
23
Table S20. Morphologic characteristics of the EABs.
203
‘Mushroom’ outer diameter means the maximal length of mushroom-like structure, while
204
inner diameter represents the maximum length of the inside hole.
205
206
Control 1
Control 2
EABs-1s
EABs-10s
EABs-60s
EABs-300s
Average
thickness / µm
36.7 ± 0.6
39.1 ± 4.8
37.6 ± 0.8
35.1 ± 0.0
24.9 ± 1.2
11.2 ± 0.3
‘Mushroom’
outer diameter
size / µm
nd
nd
32 ± 4
25 ± 2
69 ± 9
37 ± 3
Mushroom
inner diameter
(hole) size / µm
nd
nd
nd
nd
17 ± 5
11 ± 1
24
207
Fig. S21. Microbial community composition of EABs at the end of the 10 days operation.
208
209
DNA extraction was performed by using the FastPrep method as previously described
210
(Vaidya et al. 1998). Biofilms were scratched from the electrode surface and stored in a
211
Lysing Matrix E tube (Qbiogene, Alexis Biochemicals, Carlsbad, CA), cell pellets were re-
212
suspended in lysis buffer containing 100 mM Tris/HCl (at pH 8.0), 100 mM EDTA, 100 mM
213
NaCl, 1% (w/v) polyvinylpyrrolidone and 2% (w/v) sodium dodecyl sulfate. Cells were lysed
214
using 0.2 mL beads of 0.1 mm size in a Fast Prep-96 homogenizer (40 s at 1600 rpm, twice).
215
After 1min centrifugation at 18,000 × g, the supernatant was washed using phenol/chloroform
216
(1:1) and chloroform. Next, DNA was precipitated with 1:10 volume sodium acetate (3M) and
217
1 volume isopropyl alcohol at −20 °C more than 1 h. Samples were centrifuged and washed
218
with 80% ethanol, then resuspended in 50 μL of Milli-Q water. The quality and quantity of
219
DNA samples were assessed by polymerase chain reaction (PCR) using the primers described
220
below. PCR products were then separated by electrophoresis on 1.2% agarose gels.
221
The V3V4 regions of the bacterial 16S rRNA genes was sequenced using the Illumina
222
platform (LGC Genomics GmbH, Berlin, Germany) by using 2 × 300 bp paired-end reads and
223
the primers 341F (5’-NNNNNNNNTCCTACGGGNGGCWGCAG) and 785R (5’-
224
NNNNNNNNTGACTACHVGGGTATCTAAKCC) as previously described (Stewardson et
225
al. 2015). Each PCR included DNA extract (~ 5 ng), forward and reverse primer (~15 pmol
226
25
for each) and MyTaq buffer (20 μL containing 1.5 units MyTaq DNA polymerase (Bioline)
227
and 2 μL of BioStabII PCR Enhancer) (Prokopenko et al. 2013). 8-nt barcode sequence was
228
performed for both forward and reverse primers of each sample.
229
230
PCR procedures included: 1) 2 min pre-denaturation at 96 °C; 2) 30 cycles × (15 s at 96 °C,
231
30 s at 50 °C and 60 s at 72 °C). DNA concentration of the amplicons of interest were
232
estimated by gel electrophoresis. Amplicon DNA (~20 ng) of each sample were pooled for an
233
amount of 48 samples each carrying different barcodes. Another 5 cycles were amplified for
234
the low yield PCRs. One volume AMPure XP beads (Agencourt) was used to clean primer
235
dimer and remove other small mispriming products. An additional purification on MinElute
236
columns (Qiagen) were used to purify the amplicon pools. Each purified amplicon pool DNA
237
(~ 100 ng) was taken for Illumina libraries construction by using the Ovation Rapid DR
238
Multiplex System 1-96 (NuGEN). Illumina libraries were pooled and size selected with
239
preparative gel electrophoresis. Sequencing was done on an Illumina MiSeq using v3
240
Chemistry (Illumina). The Mothur community pipeline (Schloss et al. 2009) was used to
241
analyze and cluster 16S rRNA gene sequences into operational taxonomic units (OTUs) as
242
previously reported (Andersen et al. 2015). The analysis procedure included the following
243
processes: clipping 16S rRNA sequences from primers; combining and turning into forward
244
and reverse primer orientation sequences of the fragments (after removing the primer
245
sequences); removing the wrong size sequences and making them unique; aligning the
246
sequences with a V3V4 customized SILVA database which were not match or overhung;
247
removing chimera with UCHIME software (Edgar et al. 2011). Then sequences of
248
taxonomical classification and non-bacterial removal were executed with Silva database v.123.
249
OTUs were picked up by clustering (97%) identity level using cluster split method.
250
26
251
Fig. S22. Fraction of Geobacter determined by Illumina sequencing of different EABs.
252
The microbial communities were characterized for only one EAB per period, except for
253
controls (CT), were n = 2. No trend was observed with respect to the half-period of
254
polarization.
255
256
Supplementary references:
257
Andersen, S.J., Candry, P., Basadre, T., Khor, W.C., Roume, H., Hernandez-Sanabria, E., Coma, M.,
258
Rabaey, K., 2015. Electrolytic extraction drives volatile fatty acid chain elongation through lactic acid
259
and replaces chemical pH control in thin stillage fermentation. Biotechnology for biofuels 8(1), 1.
260
Bard, A.J., Faulkner, L.R., 2001. Electrochemical Methods: Fundamentals and Applications.
261
Bartlett, P.N., Pratt, K.F.E., 1995. Theoretical treatment of diffusion and kinetics in amperometric
262
immobilized enzyme electrodes Part I: Redox mediator entrapped within the film. Journal of
263
Electroanalytical Chemistry 397(1), 61-78.
264
Edgar, R.C., Haas, B.J., Clemente, J.C., Quince, C., Knight, R., 2011. UCHIME improves sensitivity
265
and speed of chimera detection. Bioinformatics 27(16), 2194-2200.
266
Liu, Y., Bond, D.R., 2012. Long-distance electron transfer by G. sulfurreducens biofilms results in
267
accumulation of reduced c-type cytochromes. ChemSusChem 5(6), 1047-1053.
268
Prokopenko, M.G., Hirst, M.B., De Brabandere, L., Lawrence, D.J.P., Berelson, W.M., Granger, J.,
269
Chang, B.X., Dawson, S., Crane Iii, E.J., Chong, L., Thamdrup, B., Townsend-Small, A., Sigman,
270
D.M., 2013. Nitrogen losses in anoxic marine sediments driven by Thioploca-anammox bacterial
271
consortia. Nature 500(7461), 194-198.
272
Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann, M., Hollister, E.B., Lesniewski, R.A.,
273
Oakley, B.B., Parks, D.H., Robinson, C.J., 2009. Introducing mothur: open-source, platform-
274
independent, community-supported software for describing and comparing microbial communities.
275
Applied and environmental microbiology 75(23), 7537-7541.
276
Schrott, G.D., Ordoñez, M.V., Robuschi, L., Busalmen, J.P., 2014. Physiological Stratification in
277
Electricity-Producing Biofilms of Geobacter sulfurreducens. ChemSusChem 7(2), 598-603.
278
Stewardson, A.J., Gaïa, N., François, P., Malhotra-Kumar, S., Delémont, C., Martinez de Tejada, B.,
279
Schrenzel, J., Harbarth, S., Lazarevic, V., 2015. Collateral damage from oral ciprofloxacin versus
280
nitrofurantoin in outpatients with urinary tract infections: a culture-free analysis of gut microbiota.
281
Clinical Microbiology and Infection 21(4), 344.e341-344.e311.
282
27
Vaidya, R., Tender, L.M., Bradley, G., O'Brien, M.J., 2nd, Cone, M., Lopez, G.P., 1998. Computer-
283
controlled laser ablation: a convenient and versatile tool for micropatterning biofunctional synthetic
284
surfaces for applications in biosensing and tissue engineering. Biotechnology progress 14(3), 371-377.
285
Virdis, B., Millo, D., Donose, B.C., Lu, Y., Batstone, D.J., Kromer, J.O., 2016. Analysis of electron
286
transfer dynamics in mixed community electroactive microbial biofilms. RSC Advances 6(5), 3650-
287
3660.
288
Zhang, X., Philips, J., Roume, H., Guo, K., Rabaey, K., Prévoteau, A., 2017. Rapid and Quantitative
289
Assessment of Redox Conduction Across Electroactive Biofilms by using Double Potential Step
290
Chronoamperometry. ChemElectroChem 4(5), 1026-1036.
291
292
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The mechanism of electron transport across electroactive biofilms (EABs) is of high interest and still a matter of debate. Quantitative assessments of their redox conduction take considerable time and require non-turnover conditions (absence of substrate), which can be detrimental to EABs. Here we measured the charge transport parameters of Geobacter spp. dominated EABs with double potential step chronoamperometry (DPSC) via Cottrell analysis. The DPSC measurement is simpler and much faster than usual techniques and allows the determination of the charge transport parameters even under turnover conditions. The electrochemical responses were well-described by a model of redox conduction only driven by electron diffusion within the EAB. The apparent diffusion coefficient for the electron (Dapp) was measured at ~ 3.2 × 10−7 cm2 s−1, a value similar to those recorded for pure Geobacter sulfurreducens EABs, or for some redox polymers with comparable concentrations of redox centers. This method will be valuable for assessing the impact of EAB characteristics and environmental factors on charge transport ability of the biofilm, and for determining the rate-limiting step(s) for current production.
Article
Full-text available
Volatile fatty acids (VFA) are building blocks for the chemical industry. Sustainable, biological production is constrained by production and recovery costs, including the need for intensive pH correction. Membrane electrolysis has been developed as an in situ extraction technology tailored to the direct recovery of VFA from fermentation while stabilizing acidogenesis without caustic addition. A current applied across an anion exchange membrane reduces the fermentation broth (catholyte, water reduction: H 2 O + e − → ½ H 2 + OH − ) and drives carboxylate ions into a clean, concentrated VFA stream (anolyte, water oxidation: H 2 O → 2e − + 2 H + + O 2 ). In this study, we fermented thin stillage to generate a mixed VFA extract without chemical pH control. Membrane electrolysis (0.1 A, 3.22 ± 0.60 V) extracted 28 ± 6 % of carboxylates generated per day (on a carbon basis) and completely replaced caustic control of pH, with no impact on the total carboxylate production amount or rate. Hydrogen generated from the applied current shifted the fermentation outcome from predominantly C2 and C3 VFA (64 ± 3 % of the total VFA present in the control) to majority of C4 to C6 (70 ± 12 % in the experiment), with identical proportions in the VFA acid extract. A strain related to Megasphaera elsdenii (maximum abundance of 57 %), a bacteria capable of producing mid-chain VFA at a high rate, was enriched by the applied current, alongside a stable community of Lactobacillus spp. (10 %), enabling chain elongation of VFA through lactic acid. A conversion of 30 ± 5 % VFA produced per sCOD fed (60 ± 10 % of the reactive fraction) was achieved, with a 50 ± 6 % reduction in suspended solids likely by electro-coagulation. VFA can be extracted directly from a fermentation broth by membrane electrolysis. The electrolytic water reduction products are utilized in the fermentation: OH − is used for pH control without added chemicals, and H 2 is metabolized by species such as Megasphaera elsdenii to produce greater value, more reduced VFA. Electro-fermentation displays promise for generating added value chemical co-products from biorefinery sidestreams and wastes.
Article
Full-text available
So close, but yet so far: G. sulfurreducens c-type cytochromes become reduced as biofilms grow on electrodes beyond a few cell thicknesses, even if the electrode is poised well above the potential required to oxidize all cytochromes. Cytochrome redox state also lags behind rapid potential changes during voltammetry, but only when the films are multiple cell layers thick, as would be expected if diffusional or exchange-based kinetics controls electron transfer between cytochromes.
Article
Full-text available
Chimeric DNA sequences often form during polymerase chain reaction amplification, especially when sequencing single regions (e.g. 16S rRNA or fungal Internal Transcribed Spacer) to assess diversity or compare populations. Undetected chimeras may be misinterpreted as novel species, causing inflated estimates of diversity and spurious inferences of differences between populations. Detection and removal of chimeras is therefore of critical importance in such experiments. We describe UCHIME, a new program that detects chimeric sequences with two or more segments. UCHIME either uses a database of chimera-free sequences or detects chimeras de novo by exploiting abundance data. UCHIME has better sensitivity than ChimeraSlayer (previously the most sensitive database method), especially with short, noisy sequences. In testing on artificial bacterial communities with known composition, UCHIME de novo sensitivity is shown to be comparable to Perseus. UCHIME is >100× faster than Perseus and >1000× faster than ChimeraSlayer. robert@drive5.com Source, binaries and data: http://drive5.com/uchime. Supplementary data are available at Bioinformatics online.
Article
Full-text available
mothur aims to be a comprehensive software package that allows users to use a single piece of software to analyze community sequence data. It builds upon previous tools to provide a flexible and powerful software package for analyzing sequencing data. As a case study, we used mothur to trim, screen, and align sequences; calculate distances; assign sequences to operational taxonomic units; and describe the α and β diversity of eight marine samples previously characterized by pyrosequencing of 16S rRNA gene fragments. This analysis of more than 222,000 sequences was completed in less than 2 h with a laptop computer.
Article
Mixed community electroactive organisms form multi-layered biofilms that are able to produce current densities comparable to those of pure Geobacter sulfurreducens, an extensively studied metal-reducing organism. The long-range electron transfer (ET) inside the biofilms and at the biofilm/electrode interface was proven to be promoted by a network of outer membrane cytochromes (OMCs). In the present work, we investigate the electron transfer process in mixed community biofilms grown on Indium Tin Oxide (ITO) electrodes by combining electrochemical measurements with Confocal Resonance Raman Microscopy (CRRM) under potentiostatic control and during chronoamperometry (CA). This approach allowed direct comparison of the heterogeneous redox process at the biofilm/electrode interface with the long-range OMCs-mediated ET inside the bulk biofilm. Our work shows that: (i) during substrate oxidation, all OMCs are in the reduced state at any distance from the electrode, and no concentration gradient of oxidized OMCs is observed; (ii) the rate constant for the long-range, homogeneous ET (k0hom) is 0.028 s−1, which is considerably lower than that predicted by others under the hypothesis that homogeneous ET is promoted by OMCs alone, and may thus indicate the contribution of alternative fast electron transfer processes; (iii) the metabolic respiration rate is much faster compared to both homogeneous and heterogeneous ET, which have similar rate constants. All in all, our results suggest that differences exist in electron transfer mechanisms between mixed community and G. sulfurreducens electroactive biofilms.
Article
Recent treatment guidelines for uncomplicated urinary tract infections (UTIs) discourage fluoroquinolone prescription because of collateral damage to commensal microbiota, but the ecologic impact of alternative agents has not been evaluated by culture-free techniques. We prospectively collected faecal samples at three time points from ambulatory patients with UTIs treated with ciprofloxacin or nitrofurantoin, patients not requiring antibiotics and household contacts of ciprofloxacin-treated patients. We described changes in gut microbiota using a culture-independent approach based on pyrosequencing of the V3-V4 region of the bacterial 16S rRNA gene. All groups were similar at baseline. Ciprofloxacin had a significant global impact on the gut microbiota whereas nitrofurantoin did not. The end of ciprofloxacin treatment correlated with a reduced proportion of Bifidobacterium (Actinobacteria), Alistipes (Bacteroidetes) and four genera from the phylum Firmicutes (Faecalibacterium, Oscillospira, Ruminococcus and Dialister) and an increased relative abundance of Bacteroides (Bacteroidetes) and the Firmicutes genera Blautia, Eubacterium and Roseburia. Substantial recovery had occurred 4 weeks later. Nitrofurantoin treatment correlated with a reduced relative proportion of the genus Clostridium and an increased proportion of the genus Faecalibacterium. This study supports use of nitrofurantoin over fluoroquinolones for treatment of uncomplicated UTIs to minimize perturbation of intestinal microbiota. Copyright © 2014 European Society of Clinical Microbiology and Infectious Diseases. Published by Elsevier Ltd. All rights reserved.
Article
The elucidation of mechanisms and limitations in electrode respiration by electroactive biofilms is significant for the development of rapidly emerging clean energy production and wastewater treatment technologies. In Geobacter sulfurreducens biofilms, the controlling steps in current production are thought to be the metabolic activity of cells, but still remain to be determined. By quantifying the DNA, RNA, and protein content during the long-term growth of biofilms on polarized graphite electrodes, we show in this work that current production becomes independent of DNA accumulation immediately after a maximal current is achieved. Indeed, the mean respiratory rate of biofilms rapidly decreases after this point, which indicates the progressive accumulation of cells that do not contribute to current production or contribute to a negligible extent. These results support the occurrence of physiological stratification within biofilms as a consequence of respiratory limitations imposed by limited biofilm conductivity.
Article
A comprehensive treatment of the diffusion and reaction within a uniform film containing immobilized enzyme and mediator at an electrode surface is presented. Approximate analytical solutions and the results of numerical simulations using the relaxation method with automatic grid point allocation are compared and found to be in excellent agreement. Case diagrams are presented showing the interrelation between the different limiting cases. Analytical expressions for the current in each limiting case are given, the behaviour of the system is discussed, and the rate-limiting processes are identified. The potential dependence of the current is shown to be a useful diagnostic together with variation in film thickness, enzyme loading, mediator loading and substrate concentration.The model is appropriate to the study of systems such as glucose oxidase entrapped within a redox hydrogel film or immobilized within a polyelectrolyte membrane containing an electrostatically entrapped mediator.
Article
This paper describes laser-based methods for preparing micropatterns of bioactive molecular species in self-assembled monolayers (SAMs) and micropatterns of proteins and other biological molecules immobilized on solid substrates. Applications of these micropatterned surfaces in multianalyte biosensing and tissue engineering are emphasized. The focus of the paper is on the use of a computer-controlled laser ablation system comprising a research-grade inverted optical microscope, a pulsed nitrogen-pumped dye laser emitting at 390 nm, a programmable sample stage, and the computerized control system. The laser system can be implemented in a typical biosensor or tissue culture laboratory to enable the facile and reproducible fabrication of micropatterned surfaces by several methods. Various methods for patterning are discussed with examples given and emphasis placed on (1) laser ablation in the fabrication of photolithography masks, (2) electrochemical patterning of SAMs, and (3) laser desorption of SAMs. The relative merits of each technique are discussed with respect to application in fabrication of active surfaces for biosensing and tissue culture applications.