AEM Accepts, published online ahead of print on 15 August 2014 Appl. Environ. Microbiol. doi:10.1128/AEM.02289-14 Copyright © 2014, American Society for Microbiology. All Rights Reserved.

1

Enhanced uranium immobilization and reduction by Geobacter sulfurreducens biofilms

2 3

Dena L. Cologgi,a* Allison M. Speers,a Blair A. Bullard,a Shelly D. Kelly,b and Gemma Regueraa#

4 5

Department of Microbiology and Molecular Genetics, Michigan State University, East Lansing,

6

Michigan a; EXAFS Analysis, Bolingbrook, Illinois b

7 8

Running Head: Uranium reduction by Geobacter biofilms

9 10

Keywords: metal reduction, uranium, biofilms, pili, c-cytochromes, bioremediation

11 12

#

Address correspondence to Gemma Reguera, [email protected]

13 14

* Present address: Dena Cologgi, department of Civil and Environmental Engineering,

15

University of Alberta, Edmonton, Alberta, Canada.

16 17

18

ABSTRACT

19

Biofilms formed by dissimilatory metal reducers are of interest to develop permeable biobarriers

20

for the immobilization of soluble contaminants such as uranium. Here we show that biofilms of

21

the model uranium-reducing bacterium Geobacter sulfurreducens immobilized substantially

22

more U(VI) than planktonic cells and for longer periods of time, reductively precipitating it to a

23

mononuclear U(IV) phase involving carbon ligands. The biofilms also tolerated high and

24

otherwise toxic concentrations (up to 5 mM) of uranium, consistent with a respiratory strategy

25

that also protected the cells from uranium toxicity. The enhanced ability of the biofilms to

26

immobilize uranium correlated only partially with the biofilm biomass and thickness and

27

depended greatly on the area of the biofilm exposed to the soluble contaminant. By contrast,

28

uranium reduction depended on the expression of Geobacter conductive pili and, to a lesser

29

extent, to the presence of the c-cytochrome OmcZ in the biofilm matrix. The results support a

30

model in which the electroactive biofilm matrix immobilizes and reduces the uranium in the top

31

stratum. This mechanism prevents the permeation and mineralization of uranium in the cell

32

envelope, thereby preserving essential cellular functions and enhancing the catalytic capacity of

33

Geobacter cells to reduce uranium. Hence, the biofilms provide cells with a physical and

34

chemical protective environment for the sustained immobilization and reduction of uranium that

35

is of interest for the development of improved strategies for the in situ bioremediation of

36

environments impacted by uranium contamination.

37 38

INTRODUCTION

39

In their natural environment microorganisms are most often found as surface-attached

40

communities or biofilms (1, 2). Biofilm cells are encased in a matrix of exopolymeric substances

41

(EPS) such as polysaccharides, nucleic acids, lipids, and proteins, which promote adhesion to

42

surfaces and mechanically stabilize the multilayered communities (3). In addition, the EPS

2

43

matrix functions as a hydrated, catalytic microenvironment for the cells that promotes the

44

adsorption of nutrients and, in some cases, their extracellular processing to facilitate their

45

assimilation (3). Although the biofilm structure is dynamic and can change to minimize mass

46

transfer limitations (4), gradients of nutrients and waste products may form within the biofilms

47

and, as a result, the chemical composition of the matrix, and the physiology of the biofilm cells,

48

is heterogeneous (5). This unique physical and chemical microenvironment also makes the

49

biofilm cell physiology substantially different from that of planktonic cells and leads to some

50

general, biofilm-specific traits such as increased resistance to antimicrobials (6), greater

51

catalytic rates (7), and increased metabolic productivity (8).

52

Multimetal resistance is also a common biofilm trait (9) and, for this reason, biofilms have

53

attracted interest for applications in metal bioremediation (10). The polyionic nature of the

54

biofilm matrix can provide, for example, 20- to 30-fold more charged groups for metal sorption

55

than planktonic cells (11). This limits the diffusion of the metals within the biofilm and their

56

permeation inside the cells, thus reducing the susceptibility of the biofilm cells to metal toxicity

57

compared to their planktonic counterparts (9). The chemical and physiological heterogeneity of

58

the biofilms also establishes pH and redox gradients across the biofilm matrix, which may

59

influence metal speciation (9). The chemistry of some metals is such that their speciation affects

60

their solubility and, therefore, their potential for spread and their bioavailability. This is

61

particularly important for uranium (U), which often persists in contaminated groundwater and

62

sediments as the soluble uranyl cation (UO22+) containing the oxidized U(VI) species. Its

63

solubility facilitates the spread of the contaminant plume far away from the source (12) and

64

results in volumes of contaminated groundwater and sediments too large to permit standard

65

‘excavation and removal’ and ‘pump and treat’ remediation approaches (13). Hence, a

66

promising bioremediation approach would be to immobilize the U(VI) contaminant and reduce it

67

to the U(IV) species, which is less soluble and can precipitate out of solution under the right

68

redox conditions and pH (12). This prevents the migration of the contaminant, reduces its

3

69

bioavailability, and minimizes the potential for exposure to humans and other living components

70

of the ecosystem. 

71

One way to reduce U(VI) to U(IV) is by the action of some dissimilatory metal-reducing

72

bacteria, which can gain energy for growth by coupling the oxidation of various electron donors

73

to the reduction of the uranyl cation (14-16). This metabolic ability can be stimulated in situ with

74

field-scale additions of electron donors and results in the removal of the uranyl cation from the

75

groundwater and its accumulation in the sediments as sparingly soluble, less mobile U(IV)

76

minerals (17-21). Field-scale additions of electron donors often stimulate the growth and activity

77

of dissimilarity metal-reducing microorganisms within the family Geobacteraceae (17, 19, 20,

78

22). Studies with the model representative Geobacter sulfurreducens indicate that the reduction

79

of U by these organisms is extracellular; this prevents the permeation and reductive

80

precipitation of the radionuclide inside the cell envelope, thereby protecting the cell from its

81

toxicity (23). The extracellular reduction of U by G. sulfurreducens is primarily mediated by hair-

82

like protein filaments or pili and, to a lesser extent, by c-cytochromes of the outer membrane

83

(23). The pili of Geobacter are conductive (24) and their expression levels are linearly

84

proportional to the amount of U reductively precipitated by planktonic cells (23). Pili expression

85

in G. sulfurreducens is also required to form an electroactive biofilm matrix (25, 26), where the

86

cells are encased in a exopolysaccharide matrix containing c-cytochromes (27). This suggests

87

that Geobacter biofilms may be particularly suitable to catalyze the speciation of U while

88

providing a microenvironment that also protects cells from its toxicity.

89

Biofilms are particularly prevalent at the matrix-well screen interface and within rock fractures

90

in subsurface environments contaminated with U, thereby influencing its mobility and speciation

91

(13). However, the contribution of Geobacter biofilms to these processes and their potential use

92

as bioremediation tools has not been fully investigated. This contrasts with the availability of

93

studies about U transformations mediated by Geobacter planktonic cells (14, 15, 23, 28, 29) or

94

by biofilms formed by other metal-reducing bacteria such as Desulfovibrio (30, 31) and

4

95

Shewanella (32, 33). Hence, we investigated U transformations mediated by biofilms of G.

96

sulfurreducens. The results indicate that the biofilms have enhanced capacity to immobilize and

97

reduce U compared to planktonic cells and also tolerate exposure to higher concentrations of

98

the contaminant for prolonged periods of time. Furthermore, microscopic, genetic, and

99

spectroscopic studies of the U-reducing biofilms indicated that the radionuclide was reduced

100

extracellularly in a catalytic process influenced by the biofilm structure and the presence of

101

redox components of the biofilm matrix, most significantly the conductive pili. These findings

102

support the notion that Geobacter biofilms contribute to the immobilization and reduction of U in

103

the subsurface and highlight their potential use as permeable biobarriers for the in situ

104

bioremediation of U.

105 106

MATERIALS AND METHODS

107

Strains and culture conditions. Wild-type (WT) Geobacter sulfurreducens PCA (ATCC

108

51573), a pilin-deficient mutant (24) (herein designated pilA), and its genetically complemented

109

strain pRG5::pilA (24) (herein designated pilA+) were routinely cultured in fresh water (FW)

110

medium (34) modified as previously described (23) and supplemented with 15 mM acetate and

111

40 mM fumarate (FWAF). The medium was dispensed into tubes or serum bottles, sparged with

112

N2:CO2 (80:20), sealed with butyl rubber stoppers (Bellco) and aluminum tear-off seals

113

(Wheaton), and autoclaved 30 minutes. Unless otherwise indicated, incubations were at 30 oC.

114

Biofilm cultures. Biofilms were either grown on 6-well, cell-culture-treated polystyrene

115

plates (Corning) or on glass coverslips assembled vertically, four at a time, onto rubber

116

stoppers, as described previously (26). Briefly, the glass coverslips used for the stopper

117

assemblies were first acid-washed overnight in a bath of 50/50 (v/v) HCl/NO3- or 15% (v/v)

118

HCl/H2O and rinsed thoroughly with ddH2O before inserting them into slits cut on inverted

119

stoppers. Each coverslip assembly was placed upside down (coverslips down) in a sterile 50 ml

120

conical tube (Corning) and submerged in FW medium lacking vitamins, minerals, acetate, and

5

121

fumarate before sterilizing it by autoclaving, as previously described (26). The assembly was

122

then transferred to a clean conical tube. The biofilm chambers (either wells of plates or conical

123

tubes with coverslip assemblies) were filled with 6 or 20 ml of FWAF medium, respectively, and

124

inoculated to a final OD600 of 0.04 with an early stationary-phase FWAF culture. The chambers

125

were incubated at 30 °C for 24, 48, or 72 h, as specified, inside a vinyl glove bag (Coy Labs)

126

with a H2:CO2:N2 (7:10:83) atmosphere. At the end of the incubation period, the culture

127

supernatant fluids were decanted from the biofilm vessels and the biofilms were washed once

128

with sterile, anaerobically-prepared wash buffer (29).

129

When indicated, the biofilm biomass in the samples was estimated as total cell protein.

130

To do this, the biofilm biomass was scraped off from the surface with a sterile plastic spatula,

131

harvested by centrifugation (5 min, 12,000 x g), and the biofilm cells were then lysed in 2M

132

NaOH at 100 oC for 1 h. The lysate solution was allowed to cool before neutralizing it with an

133

equal volume of 2M HCl. After another cycle of centrifugation to remove cellular debris, the

134

supernatant, which contained the soluble protein released after lysing the cells, was analyzed

135

for protein content using a Pierce Microplate BCA Protein Assay Kit (reducing reagent

136

compatible, Thermo Scientific) with BSA standards, according to the manufacturer’s

137

specifications. Protein was measured as an OD562 on a Tecan Sunrise Plate Reader (Tecan,

138

Inc.).

139 140

U immobilization assays with resting planktonic cell suspensions and biofilms.

141

The ability of planktonic or biofilm cells to immobilize U was assayed by monitoring the removal

142

of U(VI), provided as uranyl acetate (1 mM, or as indicated), at 30 oC by resting planktonic cells

143

and biofilms, using protocols adapted from those described previously (23, 29). Planktonic cells

144

were harvested from mid-exponential cultures (OD600, 0.3-0.5) grown for 48 h under pili-inducing

145

conditions (25 oC) and suspended in 10 ml of reaction buffer (OD600, 0.1) as described before

146

(23). Resting biofilms were prepared by first growing the surface-attached communities in

6

147

FWAF medium for 24, 48, or 72 h on the coverslip-stopper assemblies described above. The

148

culture broth of the biofilm culture was then decanted gently and the coverslip assembly was

149

rinsed gently with sterile, anaerobically-prepared wash buffer (29). To initiate the assay, the

150

assemblies were submerged upside down in 50-ml conical tubes containing 20 ml of reaction

151

buffer (23, 29) supplemented with 20 mM sodium acetate and 1 mM uranyl acetate (Electron

152

Microscopy Sciences), which had been prepared from stock solutions in 30 mM bicarbonate

153

buffer. Heat-killed planktonic cells (23) and uninoculated biofilm controls were also included to

154

rule out any abiotic removal activity and/or absorption. Resting planktonic cells, resting biofilms,

155

and the controls were incubated for up to 24 h at 30 °C. Supernatant samples (500 l) were

156

periodically removed during incubation, filtered (0.22 µm Millex-GS filter, Millipore), acidified in

157

2% nitric acid (500 l), and stored at -20 °C. All procedures were performed inside an anaerobic

158

glove bag, as described above. The concentration of U(VI) in the acidified samples was

159

measured using a Platform Inductively Coupled Plasma Mass Spectrometer (ICP-MS)

160

(Micromass,Thermo Scientific) or a Kinetic Phosphorescence Analyzer (KPA) (Chemchek) and

161

provided a measure of the total U immobilized in each sample. For comparisons, the amount of

162

U immobilized by the planktonic and biofilm cells was also normalized by the cell biomass,

163

which was calculated as the amount of total cell protein released from the samples after lysis

164

with 2M NaOH (100 oC for 1 h) and neutralization with 2M HCl, as described above.

165 166

U valence and speciation by X-ray Absorption Spectroscopy (XAS) analyses. The

167

valence and speciation of the U immobilized by resting biofilms was estimated by XANES (X-ray

168

Absorption Near Edge Spectroscopy) and EXAFS (Extended X-ray Absorption Fine Structure

169

Analysis), respectively. For these analyses, biofilms grown on coverslip-stopper assemblies and

170

exposed to U for 24 h, as described above, were rinsed gently with wash buffer, scraped off the

171

coverslips, and suspended in 2 ml of reaction buffer. The biofilm biomass was then harvested

7

172

by centrifugation (12,000 x g, 10 min), loaded into custom-made plastic holders, and stored at -

173

80 °C (23). All procedures were carried out in an anaerobic chamber and samples were kept

174

frozen during XAS measurements. The XAS measurements were performed with a

175

multielement Ge detector in fluorescence mode using the PNC-CAT beamline 20-BM at the

176

Advanced Photon Source (Argonne National Laboratory) and standard beamline parameters, as

177

described elsewhere (35). XANES measurements were used to calculate the fraction of the

178

immobilized U that had been reduced to U(IV) by linear combination fitting of the spectrum with

179

U(VI) and U(IV) standards. The spectra were energy aligned using a simultaneously measured

180

uranyl nitrate standard. The U LIII-edge EXAFS spectrum was also collected for the biofilm

181

samples and modeled to determine the atomic coordination about U, as previously described

182

(23). The number of axial oxygen (Oax) atoms calculated in the EXAFS spectral analyses was

183

also used to estimate the amount of U(VI) and U(IV) in the samples, as there are two double-

184

bonded Oax atoms for each U(VI) atom in the uranyl cation (Oax=U=Oax) and none for U(IV)

185

(36).

186 187

Vitality fluorescent assays. The respiratory activity of 48-h old biofilms after exposure

188

to various concentrations (1, 2.5, and 5 mM) of uranyl acetate for 24 h was assayed using the

189

fluorescent RedoxSensor vitality kit (Invitrogen) in reference to biofilms not exposed to uranyl

190

acetate (0 mM controls), with protocols adapted from those described previously (23). The

191

biofilms were first grown for 48 h on coverslip assemblies, rinsed gently with reaction buffer, and

192

incubated in reaction buffer in the presence or absence of the uranyl acetate. The experiment

193

also included negative controls (0* mM) in which the 0 mM biofilms were first treated for 5 min

194

with electron transport chain uncouplers (a mix of sodium azide [10 mM] and carbonyl cyanide

195

3-chlorophenylhydrazone or CCCP [10 μM]) provided in RedoxSensor vitality kit (Invitrogen).

196

After 24 h of incubation, the reaction buffer was decanted from the tubes and the assemblies

8

197

were washed once with wash buffer. The biofilm biomass was scraped from the coverslips and

198

suspended in 1 ml of reaction buffer, vortexed briefly, mixed 1:1 with the Redox dye solution,

199

and incubated at room temperature for 10 min before measuring the dye’s fluorescence (490

200

nm excitation, 520 nm emission) using a SpectraMax M5 plate reader (Molecular Devices).

201

Separate aliquots of the biofilm samples were stained with SYTO 9 (Invitrogen) following

202

manufacturer’s recommendations to determine the total biofilm biomass. The respiratory activity

203

of the biofilms was estimated as the vitality index, i.e., the relative fluorescence emission from

204

the RedoxSensor dye normalized by the fluorescence emission from the Syto9 dye for 6

205

replicate samples (3 biological samples and two technical replicates for each). Statistically

206

significant (p < 0.05, one star; p < 0.005, two stars) changes in vitality indexes were determined

207

in pairwise comparisons to the 0 mM control biofilms using the t-test function of the Microsoft

208

Excel software.

209 210

Microscopy. An Scanning Electron Microscope (SEM) was used to image 48-h old WT

211

biofilms, which were grown on coverslip assemblies as described above except that round (12-

212

mm diameter), rather than square, glass coverslips were used. The biofilms were incubated at

213

30 oC in reaction buffer in the presence or absence of 1 mM uranyl acetate for 24 h and washed

214

once in wash buffer before fixing the cells at 4°C for 1-2 h in 4% glutaraldehyde. After this, the

215

biofilms were rinsed briefly in 0.1 M sodium phosphate buffer and then dehydrated in a series of

216

ethanol washes (25%, 50%, 75%, 95%, 10 minutes each) followed by three 10-min washes in

217

100% ethanol. The samples were critical point dried using a Blazers 010 critical point dryer

218

(Blazers Union Ltd.) with liquid CO2 as the transitional fluid. Once dry, the coverslips containing

219

the biofilm samples were mounted on aluminum stubs using epoxy glue and coated with ~10

220

nm of osmium using a NEOC-AT osmium coater (Meiwafosis Co., Ltd.). Samples were imaged

9

221

with a JEOL JSM-7500F SEM equipped with an Energy Dispersive Spectroscopy (EDS) 30mm2

222

detector crystal, which was used for elemental analyses.

223

Confocal Laser Scanning Microscopy (CLSM) was also used to examine the biofilms

224

and generate 3D-images for structural analyses. The WT biofilms were grown for 24, 48, and 72

225

h on 6-well plates and the pilA and pilA+ biofilms were grown for 48 h. At the end of the

226

incubation period, the culture supernatant fluid from each well was carefully decanted, and the

227

biofilms were stained for 15 min with the LIVE/DEAD® BacLight™ Bacterial Viability Kit

228

(Invitrogen) dye solution, following the manufacturer’s recommendations. After staining, the

229

biofilms were washed once in phosphate buffer saline (PBS) and imaged on a Zeiss Pascal

230

LSM microscope (Carl Zeiss Microscopy, LLC) equipped with an Achroplan 40x/0.80W

231

submersible objective. Images were collected every 1.14 µm, and side and top projections of

232

the biofilms were created using the Zeiss LSM Image Browser software (Carl Zeiss Microscopy,

233

LLC). Approximately 6-10 distinct fields of view (1,024 x 1,024 pixels, 0.22 µm/pixel) were

234

imaged for each of three independent biofilm replicates and the micrographs were analyzed with

235

the COMSTAT image analysis software using connected volume filtration to remove noise in the

236

data, as described previously (4). The biofilm structural parameters quantified with the

237

COMSTAT analyses are listed in Table S2.

238 239

SDS-PAGE and heme-stain of proteins of the biofilm EPS matrix. The

240

exopolysaccharide matrix (EPS) of 48-h biofilms grown on 6-well plates was extracted using a

241

modification of previously described protocols (27, 37). Briefly, biofilms were scraped off the

242

wells, collected in reaction buffer, and the solution was centrifuged for 10 min at 13,000 x g.

243

After decanting the supernatant fluid, the biofilm pellet was suspended in 1/5 volume of TNE (10

244

mM Tris-HCl [pH 7.5], 100 mM NaCl, 5 mM EDTA) and vortexed for 1 min. SDS was then

245

added to a final concentration of 0.1% (w/v), and the solution was mixed at room temperature

246

for 5 min. The samples were then passed 10 times through an 18G needle, and centrifuged at

10

247

15,500 x g for 20 min to collect the insoluble sheared biological fraction. The resulting pellet,

248

which predominantly contained the biofilm matrix and associated proteins, was washed 5 times

249

to remove any SDS before suspending it in 10 mM Tris-HCl buffer (pH 7.5). 

250

Heme-containing proteins in the EPS biofilm matrix were also analyzed, as previously

251

described (27). The protein content in the matrix sample was determined with the bicinchoninic

252

acid (BCA) protein assay, as described above, and the equivalent of 20 µg of protein of EPS

253

sample was boiled in SDS sample buffer for 10 min before electrophoretic analysis in a 12%

254

Mini-Protean TGX gel (BioRad) at 250 V for 30 min. The Novex Sharp markers (Invitrogen)

255

were used as molecular weight standards. Heme-containing proteins were visualized on the gel

256

after N,N,N’,N’-tetramethylbenzidine staining, as described previously (23, 38). A duplicate gel

257

was run in parallel and stained for total protein using Coomassie Brilliant Blue G-250 (BioRad)

258

according to the manufacturer’s recommendations.

259 260

RESULTS AND DISCUSSION

261

Enhanced U immobilization and tolerance by biofilms. The kinetics of U

262

immobilization were investigated by measuring the amount of U, provided as uranyl acetate,

263

removed from solution by resting 48-h old biofilms and in reference to pili-expressing planktonic

264

cells (Fig. 1A). Heat-killed planktonic cell controls and uninoculated biofilm vessel controls were

265

also included to rule out any abiotic precipitation of U either due to components of the medium

266

(such as phosphate) or adsorption to the biofilm coverslip assemblies and vessels used for the

267

assays (Fig. 1A). The kinetics of U removal by the biofilms was linear throughout the 24-h long

268

assay (~10 µM/h; R² = 0.967). By contrast, the planktonic cells stopped removing U from

269

solution after 12 h and all the U immobilized by the planktonic cell biomass was solubilized

270

again after 24 h of incubation. The enhanced capacity of the biofilms to immobilize U compared

271

to planktonic cells (Fig. 1A) cannot be explained by differences in cell biomass because the

272

rates of U immobilization per h, once normalized by the total cell protein, were 6-times higher in

11

273

the biofilms (~2.8 µmol U immobilized per mg protein per h) than in planktonic cells (~0.45

274

µmol/mg/h). Furthermore, the biofilms sustained the rates of U immobilization 2 times longer (24

275

h) than their planktonic counterparts. As a result, the yields of U immobilization by the biofilms at

276

the end of the 24-h assay were 12-times greater (~66 µmol of U immobilized per mg of biofilm

277

biomass protein) than the maximum yields measured in the planktonic cells after 12 h (~5.4

278

µmol of U per mg of protein).

279

The sustained immobilization of U by 48-h old biofilms for 24 h suggests that the biofilm

280

cells remained viable and metabolically active throughout the assay. To investigate this, we

281

used the fluorogenic RedoxSensorTM dye to measure the respiratory activities of the biofilms

282

after exposure to increasing concentrations of U (1, 2.5, and 5 mM) for 24 h and in reference to

283

biofilms incubated similarly but without uranium (0 mM) (Fig 1B). Interestingly, the vitality index

284

of the biofilms exposed to 1 mM U was greater than the unexposed 0 mM controls.

285

Furthermore, increasing the concentrations of U (2.5 and 5 mM) did not decrease the

286

respiratory activities of the biofilms to or below unexposed biofilm levels. The measured vitality

287

correlated well with the respiratory activity of the biofilms because pretreating the biofilms with

288

sodium azide and CCCP (0* mM controls), which are chemicals that uncouple the electron

289

transport chain, reduced the vitality by 80%. The dye used in these experiments yields green

290

fluorescence when modified by the bacterial reductases, which are mostly located in the

291

electron transport system of the cell envelope (39, 40). It is unlikely that the increases in the

292

biofilm’s respiratory rate with U were due to a cellular response of the cells to U toxicity,

293

because the vitality index also serves as a proxy of viability and overall cell health in G.

294

sulfurreducens cells after U exposure (23). For example, exposure of pili-expressing planktonic

295

cells to 1 mM U for only 6 h reduces the respiratory activity of the cells by 70%, because of the

296

toxic effects of the U that permeates and mineralizes in the cell envelope (23). Yet when the

297

cells are encased in the protective biofilm microenvironment they tolerated high, otherwise toxic

298

levels of U, and their metabolic activities were stimulated as well.

12

299

U reduction by the biofilms. As the reaction of interest during U bioremediation is the

300

reduction of the soluble U(VI) species to the less mobile U(IV), we used XANES to investigate if

301

any of the U immobilized by the biofilms had been reduced to U(IV). For these experiments, the

302

48-h old biofilms studied in Fig. 1A were collected at the end of the 24-h U challenge and the

303

valence of the biofilm-associated U was analyzed by XANES in reference to U(VI) and U(IV)

304

standards. Approximately 65% (± 5%) of the U immobilized by the biofilms was in the reduced

305

state, U(IV) (Table S1). Consistent with the reductive precipitation of U by the biofilms, SEM-

306

EDS analyses revealed needle-like, extracellular precipitates coating the microcolonies (Fig. 2A)

307

that contained U (Fig. S1). The preferential localization of U minerals on the top regions of the

308

biofilm is consistent with the metabolic stratification of cells within electroactive biofilms of G.

309

sulfurreducens (41). As the biofilms grow, their thickness increases and a gradient of acetate is

310

formed across the biofilm. This concentrates the metabolically active cells in the top regions of

311

the biofilm, where acetate is more available. As a result of this metabolic stratification, U is

312

preferentially reduced in the metabolically-active, top biofilm stratum (41).

313

SEM micrographs also revealed the presence of extracellular filaments interspersed with

314

the U precipitates (Fig. 2B). The dense filament network was more clearly visible in biofilm

315

controls not exposed to U (Fig. 2C). Some filaments had diameters (< 4 nm) that closely

316

matched the diameters reported for the conductive pili of G. sulfurreducens (24), whereas

317

others had larger diameters (ca. 15-20 nm) within the ranges reported for dehydrated EPS

318

fibers (27). The conductive pili of G. sulfurreducens are the primary sites for extracellular

319

immobilization and reduction of U by pili-expressing planktonic cells (23) and are also required

320

for the formation of electroactive biofilms (25, 26). In addition to containing the conductive pili,

321

the EPS matrix of G. sulfurreducens also anchors several c-cytochromes involved in metal

322

reduction (27). Thus, the Geobacter matrix provides an extended matrix for U immobilization but

323

also contains redox-active components which could mediate its reduction.

13

324

U immobilization and reduction during biofilm development. As the biofilms

325

develop, their thickness increases and more biofilm matrix and matrix-associated redox

326

components are available to immobilize and reduce U, respectively. Hence, we hypothesized

327

that older and thicker biofilms would immobilize and reduce more U than younger and thinner

328

biofilms. For this reason, we studied U immobilization and reduction as a function of biofilm age

329

(24-h, 48-h, and 72-h old biofilms) (Table S1). The biofilms were also imaged by confocal

330

microscopy (Fig. 4) and the confocal micrographs were analyzed with the COMSTAT software

331

(4) to estimate the thickness and other structural parameters of the biofilms (Table S2).

332

We first studied U immobilization during biofilm development. Although the capacity of

333

the biofilms to immobilize U increased steadily as they aged (Fig. 3A) and their thickness

334

increased (Fig. 4), the amount of U immobilized per biomass unit did not change significantly

335

during biofilm development (Fig. S3A). For example, young (24-h old) biofilms had less biofilm

336

biomass, thickness, and surface area (cells per biofilm layer) than older (48-h and 72-h old)

337

biofilms (Fig. S4, panels A-C). Yet despite being less dense, they immobilized more U per

338

biofilm biomass unit than older biofilms (Fig. S3A). Interestingly, surface coverage by all of the

339

biofilms was similar and neared confluence (81-92%) (Fig. S4D). The surface to volume ratio

340

also remained relatively constant during biofilm development (Fig. S4E), indicating that the area

341

of the biofilm exposed to the liquid millieu relative to the biofilm volume is the same. This also

342

maintained the biofilm topography constant, as indicated by the similar roughness coefficients

343

measured in all the biofilms (Fig. S4F). Taken together, the results support a model in which

344

achieving biofilm confluence early on in biofilm development is critical to promote U

345

immobilization. As the biofilms age, they adapt their structure to maintain a topography that

346

maintains sufficient areas of the biofilms exposed to the liquid milleu and maintained similar

347

levels of U immobilization per biomass unit. This model also implies that U is preferentially

348

immobilized in the biofilm regions exposed to the liquid milley, which is further supported by the

14

349

preferential localization of U on the outer layers of the biofilms revealed in SEM micrographs

350

(Fig. 2).

351

The fraction of the immobilized U that was reduced to U(IV) was also estimated by

352

XANES (Table S1). Interestingly, the yields of U reduction increased linearly with biofilm age

353

(Fig. 3A), even when normalized by the biofilm biomass (Fig. S3A). None of the biofilm

354

structural parameters determined in the COMSTAT analyses fit the linear trend of U reduction

355

(Fig. S4). Parameters important for U immobilization (surface coverage, surface to volume

356

ration, and roughness coefficient) were, for example, unaffected by biofilm age. The biofilm

357

parameters that changed were those related to biofilm biomass (biofilm biomass, thickness, and

358

surface area) but they all stabilized in the 48-h and 72-h old biofilms rather that increasing

359

linearly. By contrast, the total protein content of the biofilms, which accounts for protein

360

contributed by cells and by proteinaceous components of the biofilm matrix, increased linearly

361

as the biofilms aged (Fig. S5). Biofilm formation is a developmental process, consisting of

362

specific stages such as attachment, microcolony formation, and biofilm maturation (42). As they

363

transition from one stage of biofilm development to the next, the biofilms increase their

364

thickness and biomass but also undergo extensive gene reprogramming so that specific biofilm

365

components and activities are expressed (9, 43). Pili expression in G. sulfurreducens is, for

366

example, required to transition from young, monolayered biofilms to thicker biofilms composed

367

of several layers of cells (26) and this, in turn, is required to maintain the electroactivity of the

368

biofilms (25). Hence, as the Geobacter biofilms age, they express more redox-active

369

components in the biofilm matrix and this, in turn, may increase their capacity to reduce the

370

immobilized U to U(IV).

371 372

Role of the redox components of the biofilm matrix in U immobilization and

373

reduction. As the conducive pili of G. sulfurreducens are the primary site for U immobilization

374

and reduction in planktonic cells (23) and also are required for biofilm formation (25, 26), we

15

375

compared the capacity of 48-h biofilms of the WT, a pilin-deficient mutant (pilA), and its

376

genetically complemented strain pilA+ to immobilize and reduce U (Fig. 3B). The pilA biofilms

377

had a diminished capacity to immobilize and reduce U (p = 0.07) compared to WT biofilms of

378

the same age. The defect was due to the pilus deficiency because complementing the mutation

379

in trans (pilA+) restored the phenotypes and promoted U immobilization and reduction to levels

380

comparable to the WT biofilms.

381

The defect in pili production in the pilA biofilms also prevented the biofilms from growing

382

in thickness (Fig. 4) and lead to ~ 50% decreases in other biomass-related structural

383

parameters such as biomass volume and surface area (cells per biofilm layer) compared the

384

WT biofilms (Table S2 and Fig. S4A-C). These defects are consistent with the reported role of

385

the pili at promoting cell-cell aggregation during biofilm formation and providing the structural

386

support required to build multilayered biofilms (26). However, despite the biomass and pili

387

defects, the pilA biofilms immobilized more U per biomass unit than the WT biofilms.

388

Furthermore, genetic complementation of the pilA mutation in the pilA+ strain resulted in denser

389

biofilms (Fig. 4), consistent with the increased ability of the pilA+ hyperpiliated strains to

390

aggregate (26). Yet despite the increased biofilm biomass, the pilA+ biofilms immobilized less U

391

per biofilm biomass unit than any other biofilm (Fig. S3B). The lack of a clear correlation

392

between biofilm biomass and U immobilization in the pilA, WT, and pilA+ biofilms may reflect

393

differences in diffusion rates of electron donors and acceptors, as more diffusional constraints

394

are expected in thicker (pilA+>WT) than in thinner (pilA) biofilms. Consistent with this, the pilA+

395

biofilms had lower surface to volume ratios than any other biofilms (Fig. S4E and Table S2),

396

indicating that less surface area of the biofilm volume is exposed to the liquid medium and

397

available for electron donors and acceptors diffusion. Furthermore, the roughness coefficient,

398

which provides a measure of the spatial heterogeneity of the biofilms as variations in biofilm

399

thickness, was also lowest in the pilA+ biofilms (Fig, S4F). This indicates that the very dense

400

pilA+ biofilms have a relative uniform topography (very low roughness coefficient, ~0.06),

16

401

whereas the thin pilA biofilms formed the most heterogenous biofilms of all (highest roughness

402

coefficient, ~0.32, almost twice the WT coefficients) (Table S2). Hence, the more heterogenous

403

structure of the pilA and WT biofilms also maximizes the biofilm surface area available for

404

diffusion of electron donors and acceptors.

405

Although the pili-deficient pilA biofilms immobilized more U per biomass unit than any

406

other strain, they reduced a lower fraction (~38%) of the immobilized U to U(IV) than the WT

407

(~63%) and pilA+ (~76%) biofilms per biomas unit (Fig. S3). The capacity of the biofilms to

408

reduce U (pilA

Enhanced uranium immobilization and reduction by Geobacter sulfurreducens biofilms.

Biofilms formed by dissimilatory metal reducers are of interest to develop permeable biobarriers for the immobilization of soluble contaminants such a...
581KB Sizes 1 Downloads 6 Views