downloaded from on april 5, 2020 by guest · 141 heated conductivity cell. a koh gradient was...

31
1 1 Running Title: Innate NDFO capacity of nitrate-reducing bacteria 1 2 Fe(II) oxidation is an innate capability of nitrate-reducing bacteria involving 3 abiotic and biotic reactions 4 5 Hans K. Carlson 1 , Iain C. Clark 2 , Steven J. Blazewicz 3 , Anthony T. Iavarone 4 , 6 and John D. Coates 1,# 7 8 Department of Plant and Microbial Biology 1 , Department of Civil and 9 Environmental Engineering 2 , Department of Environmental Science, Policy 10 and Management 3 , QB3/Chemistry Mass Spectrometry Facility 4 , University 11 of California, Berkeley, CA 94720. 12 13 #Corresponding author: John D. Coates. Mailing address: University of 14 California, Berkeley, Department of Plant and Microbial Biology, Berkeley, CA 15 94720. Phone 510-643-8455. Fax: 510-642-4995. E-mail: 16 [email protected]. 17 18 Author contributions: H.K.C., I.C.C., and J.D.C. designed research. H.K.C., 19 A.T.I., I.C.C., and S.B. performed research and analyzed data. H.K.C., I.C.C. 20 and J.D.C. wrote the paper. 21 22 The authors declare no conflict of interest. 23 24 Copyright © 2013, American Society for Microbiology. All Rights Reserved. J. Bacteriol. doi:10.1128/JB.00058-13 JB Accepts, published online ahead of print on 17 May 2013 on May 2, 2020 by guest http://jb.asm.org/ Downloaded from

Upload: others

Post on 01-May-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

1

1

Running Title: Innate NDFO capacity of nitrate-reducing bacteria 1

2

Fe(II) oxidation is an innate capability of nitrate-reducing bacteria involving 3

abiotic and biotic reactions 4

5

Hans K. Carlson1, Iain C. Clark2, Steven J. Blazewicz3, Anthony T. Iavarone4, 6

and John D. Coates1,# 7

8

Department of Plant and Microbial Biology1, Department of Civil and 9

Environmental Engineering2, Department of Environmental Science, Policy 10

and Management3, QB3/Chemistry Mass Spectrometry Facility4, University 11

of California, Berkeley, CA 94720. 12

13

#Corresponding author: John D. Coates. Mailing address: University of 14

California, Berkeley, Department of Plant and Microbial Biology, Berkeley, CA 15

94720. Phone 510-643-8455. Fax: 510-642-4995. E-mail: 16

[email protected]. 17

18

Author contributions: H.K.C., I.C.C., and J.D.C. designed research. H.K.C., 19

A.T.I., I.C.C., and S.B. performed research and analyzed data. H.K.C., I.C.C. 20

and J.D.C. wrote the paper. 21

22

The authors declare no conflict of interest. 23

24

Copyright © 2013, American Society for Microbiology. All Rights Reserved.J. Bacteriol. doi:10.1128/JB.00058-13 JB Accepts, published online ahead of print on 17 May 2013

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 2: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

2

2

Abstract 25

Phylogenetically diverse species of bacteria can catalyze the oxidation of ferrous 26

iron (Fe(II)) coupled to nitrate (NO3-) reduction, often referred to as nitrate-27

dependent iron oxidation (NDFO). Very little is known about the biochemistry of 28

NDFO, and though growth benefits have been observed, mineral encrustations 29

and nitrite accumulation likely limit growth. Acidovorax ebreus, like other species 30

in the Acidovorax genus, is proficient at catalyzing NDFO. Our results suggest 31

that the induction of specific Fe(II) oxidoreductase proteins is not required for 32

NDFO. No upregulated periplasmic or outer membrane redox active proteins, 33

like those involved in Fe(II) oxidation by acidophilic iron oxidizers or anaerobic 34

photoferrotrophs, were observed in proteomics experiments. We demonstrate 35

that while “abiotic” extracellular reactions between Fe(II) and biogenic NO2-/NO 36

can be involved in NDFO, intracellular reactions between Fe(II) and periplasmic 37

components are essential to initiate extensive NDFO. We present evidence that 38

an organic co-substrate inhibits NDFO, likely by keeping periplasmic enzymes in 39

their reduced state or by stimulating metal efflux pumping, or both, and that 40

growth during NDFO relies on the capacity of a nitrate-reducing bacterium to 41

overcome the toxicity of Fe(II) and reactive nitrogen species. Based on our data 42

and evidence in the literature, we postulate that all respiratory nitrate-reducing 43

bacteria are innately capable of catalyzing NDFO. Our findings have implications 44

for a mechanistic understanding of NDFO, the biogeochemical controls on 45

anaerobic Fe(II) oxidation, and the production of NO2-, NO and N2O in the 46

environment. 47

Introduction 48

Although microbial nitrate-dependent Fe(II) oxidation (NDFO) has been 49

known for over two decades and has been demonstrated to play a critical role in 50

both the extant (1) and ancient (2) global iron cycle, almost nothing is known of 51

the underlying biochemical or genetic mechanisms involved (3). NDFO microbes 52

have been demonstrated to oxidize both soluble and insoluble Fe(II) (1, 4-6), and 53

to produce a variety of insoluble mixed valence iron mineral products (1, 5, 7, 8). 54

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 3: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

3

3

Some evidence already exists that the capacity for NDFO is widespread 55

and likely innate to all nitrate-reducing bacteria. Bacterial species that couple the 56

oxidation of Fe(II) to nitrate reduction have been isolated from a wide range of 57

habitats and are phylogenetically diverse indicating their environmental 58

prevalence and importance (5, 9, 10). Recently, species from the genus 59

Acidovorax have dominated enrichment cultures and been isolated as proficient 60

nitrate-dependent Fe(II) oxidizers (11-14), however, even E. coli has been 61

demonstrated to be capable of Fe(II) oxidation coupled to nitrate reduction (15). 62

Although NDFO is based on thermodynamically favorable redox reactions (16, 63

17), only a few studies suggest a growth enhancement from this metabolism (12, 64

18). This suggests that NDFO may primarily be inadvertent or a detoxification 65

strategy. 66

While the toxicity of Fe(II) to biological systems under aerobic conditions is 67

widely appreciated, less is understood about Fe(II) toxicity under anaerobic 68

conditions. Some studies have demonstrated that Fe(II) is toxic at low 69

concentrations to anaerobic bacteria, such as anoxygenic phototrophs (19) or 70

streptococci (20). Several possibilities for anaerobic Fe(II) toxicity include 71

inhibition of the F-ATPase (20), binding to membranes (21), disrupting protein 72

stability or replacing active-site metal cofactors (22), and oxidation and 73

precipitation of insoluble Fe(III) on cellular components to impair nutrient uptake 74

(13). Under NDFO conditions, redox transformations of nitrogen oxides can 75

produce intermediates, such as nitric oxide, which is capable of binding to and 76

reacting with heme cofactors (23) or Fe-S clusters (24) or, in the presence of 77

transition metals, nitrosating protein thiols to inhibit or alter protein activity (25). 78

Here we present the results of proteomic and physiological experiments 79

used to understand the mechanism of NDFO in the model organism Acidovorax 80

ebreus. Like several other species in the Acidovorax genus (12-14), A. ebreus 81

was isolated based on its capacity for NDFO (11). Previous studies have 82

reported a growth advantage from NDFO by Acidovorax species in both batch 83

and continuous flow cultures (12, 14), in addition to the accumulation of nitrite 84

(NO2-) and periplasmic mineral encrustations (26). Based on our findings with A. 85

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 4: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

4

4

ebreus, we postulate that all nitrate respiring organisms are innately capable of 86

catalyzing NDFO, but the survival and growth benefit obtained is dependent on 87

the ability to overcome Fe(II), NO2- and NO toxicity. Our findings have 88

implications for understanding the evolution of Fe(II) oxidation as a microbial 89

metabolism, the impact of Fe(II) on nitrate reducing communities and the 90

influence of Fe(II) on the product distribution of microbial nitrate reduction. 91

92

Materials and Methods: 93

Media and culture conditions 94

Anaerobic cultures of Acidovorax ebreus strain TPSY, formerly 95

Diaphorobacter sp. TPSY (11) and Azospira suillum strain PS were grown 96

organotrophically in anoxic bicarbonate buffered basal medium (BBM) (27), pH 97

6.8 at 37 °C with 10 mM sodium nitrate and varying concentrations of sodium 98

acetate and harvested in late log-phase. BBM contained the following 99

components (per Liter): 0.25 g NH4Cl, 0.6 g NaH2PO4, 0.1 g KCl, and 2.52 g 100

NaHCO3 with the addition of vitamins and minerals according to Bruce et 101

al.,1999 (27) and the appropriate concentration of electron donors and acceptors. 102

Addition of millimolar Fe(II) to the growth medium immediately formed a 103

white precipitate leaving 1-2 mM soluble Fe(II). This precipitate was shown to be 104

predominantly vivianite by Kappler et al., 2005 (13). Unless otherwise indicated, 105

our growth and cell suspension experiments utilized this mixed soluble and 106

insoluble (vivianite) Fe(II) form. 107

Fe(II) oxidation cell suspensions. Anaerobic cultures of Acidovorax ebreus 108

were grown organotrophically in bicarbonate-buffered basal media (BBM) with 10 109

mM sodium nitrate and 6.25 mM sodium acetate and harvested in late log-phase. 110

Cells were washed three times in BBM without donors or acceptors under an 111

N2:CO2 (80%:20% v:v) headspace. In cell suspensions, cells were inoculated into 112

fresh media at ~5x108 cells/mL with electron donor and acceptor concentrations 113

as indicated in the text. When indicated, chloramphenicol was added to cell 114

suspensions at a concentration of 100 μg/mL. Chloramphenicol was found to 115

completely inhibit growth of A. ebreus at concentrations above 50 μg/mL. When 116

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 5: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

5

5

indicated, Fe(II) in the form of Fe(II)-NTA, Fe(II)-NTA-agarose or synthetic 117

vivianite was added from anoxic stock solutions. Fe(II)-NTA-agarose was 118

prepared by adding FeCl2 from sterile anoxic stock solutions to NTA-agarose 119

beads (Qiagen). The beads were washed 5x with sterile water under anoxic 120

conditions. Synthetic vivianite was made by chemical precipitation. FeCl2 was 121

added to 10 mM phosphate buffer. The precipitate was washed with 10 mM 122

phosphate buffer and kept as a suspension under anoxic conditions. To assess 123

the capacity of NO and NO2- to oxidize and solubilize Fe(II) from synthetic 124

vivianite, anoxic aqueous stock solutions of sodium nitrite or diethylamine 125

NONOate (Cayman chemical) in 0.1 M NaOH were added to BBM containing 126

synthetic vivianite. NONOates, or diazeniumdiolates, are stable at alkaline pH, 127

but decompose at neutral pH to release nitric oxide (28). DEA-NONOate decays 128

at pH 7.4 and 25 °C with a 16-minute half-life to release 2 moles of NO per mole 129

of parent compound. Appropriate controls for the effect of diethylamine (the 130

product of NONOate decay) and sodium hydroxide were used to show that 131

neither of these compounds, in the concentrations added to generate NO, altered 132

the oxidation state or solubility of Fe(II) in synthetic vivianite. 133

Analytical techniques 134

Nitrate, nitrite, and acetate were measured with an ion chromatograph 135

(ICS-2100, Dionex, Sunnyvale, California) using a method developed to quantify 136

all three analytes simultaneously. Samples for ion chromatography were diluted 137

in 10 mM NaOH, filtered and stored at -20 °C until analysis. The guard and 138

analytical columns were IonPac AG16 (4 x 50mm) and IonPac AS16 (4 x 139

250mm), respectively, with an ASRS-300 4-mm suppressor system and a DS6 140

heated conductivity cell. A KOH gradient was generated using the EGC III KOH 141

generator at an isocratic flow rate of 1.5 ml/min. The KOH concentration was 1.5 142

mM from 0-7 min, ramped to 10 mM from 7-13 min, was held at 10 mM from 13-143

16 min, ramped to 35 mM from 16-17 min, was held at 35 mM from 17-27 min 144

and ramped back down to 1.5 mM from 27-30 min. 145

Soluble and insoluble Fe(II) were measured separately to characterize the 146

oxidation of each Fe(II) form in cultures and to avoid analytical artifacts due to 147

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 6: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

6

6

rapid reactions which occur between NO2- and Fe(II) in acidic solution and the 148

regeneration of NO2- through reactions between NO and O2 under oxic 149

conditions. Klueglein and Kappler (29) provide an excellent discussion of this 150

problem and other alternatives to avoid it. At each time point, 0.5 mL of culture 151

media was removed in the anaerobic chamber (Coy) and centrifuged to separate 152

insoluble Fe(II) from soluble Fe(II) and NO2- in solution. Soluble Fe(II) was 153

directly measured with the ferrozine assay and insoluble Fe(II) was resuspended 154

in 0.5 M HCl for 24 hrs and Fe(II) was quantified using the ferrozine assay (30). 155

Control experiments in which the insoluble Fe(II) was washed with anoxic buffer 156

to remove residual nitrite prior to resuspension in 0.5 M HCl gave identical 157

ferrozine results. 158

Two techniques were used for measurement of nitric oxide (NO) gas. For 159

headspace NO analysis, culture headspace was anaerobically sampled and 160

injected into the purge cell of a Nitric Oxide Analyzer 280i (Sievers). For solution 161

NO, samples of culture solution were sampled and injected into a purge solution 162

in anoxic vials (31). The headspace from these vials was injected into the NOA. 163

The purge solution releases NO from iron complexes in solution and contains 5.5 164

mL glacial acetic acid, 1.5 mL 0.8 M potassium ferricyanide and 1 mL 0.1 M 165

sulfanilamide in 2 M glacial acetic acid. The ferricyanide oxidizes iron and 166

releases NO and sulfanilamide reacts rapidly with NO2- to prevent further NO 167

release in the acid solution. NO gas standards were used to generate a standard 168

curve for quantification. 169

For measurement of N2O, culture headspace (2 mL) was collected with a 170

gas tight syringe and injected directly into a Hewlett Packard HP6890 (Agilent 171

Technologies Inc., Santa Clara, CA) gas chromatograph fitted with a Hayesep 172

DB 100/120 (1/16” x 1.5 m) column which fed into a Hayesep DB 120/140 (1/16” 173

x 2.0 m) column leading to a pulse discharge detector (PDD). The PDD was 174

calibrated for N2O using a 5 point-standard curve, and a single standard was 175

analyzed hourly thereafter to correct for instrumental drift. 176

Growth was monitored in cultures by direct counts of acridine orange-177

stained cells. At timepoints, 500 μL of cultures was fixed in 3.7% formaldehyde. 178

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 7: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

7

7

Cells were then diluted in water or 100 mM oxalate buffer. 1 mL of diluted cells 179

were stained with acridine orange (final concentration 1 μg/mL), and applied to a 180

polycarbonate black filter in a glass vacuum-filter apparatus (Fisher). The filter 181

was washed with 5 mL of oxalate buffer and cells counted by microscope 182

inspection with a 100x magnification oil-immersion lens using an Axioimager M1 183

microscope (Zeiss). 184

Trypsin shaving proteomics sample preparation 185

For proteomics, cells from NDFO and organoheterotrophic cultures were 186

harvested anaerobically at timepoints between 8-48 hours. At each timepoint, 50 187

mL of culture was centrifuged anaerobically, treated with oxalate buffer to 188

remove Fe(II) from cells and washed 2x with 100 mM ammonium bicarbonate 189

buffer. Cells were resuspended in two aliquots of 500 μL. One sample was 190

lysed (for a whole cell lysate proteome) and the other was kept intact (for a 191

trypsin-shaved proteome) under anaerobic conditions. 200 nanograms of 192

Trypsin Gold porcine protease (Pierce) was added to lysed and intact cells and 193

samples were incubated for 45 minutes at 37 °C . After 45 minutes, both the 194

lysed and intact cells were centrifuged at 10,000 rcf and the supernatant was 195

treated in two sequential 30-minute steps with 1 mM dithiothreitol and 1 mM 196

iodoacetamide then further digested overnight with 200 nanograms of trypsin at 197

37 °C. Peptides from overnight digests were concentrated with 100 μL C18 OMIX 198

tips (Agilent) and eluted with 85% acetonitrile:0.1% trifluoroacetic acid in water. 199

Acetonitrite was removed by vacuum centrifugation (speed-vac) and peptides 200

analyzed by LC-MS/MS (see supplemental methods). 201

Proteomics Data Analysis 202

Normalized peptide counts were used as semi-quantitative measure of 203

relative protein abundance. Validation of this approach has been published (32, 204

33). Peptide counts were normalized by dividing the observed number of 205

peptides for a given protein by the total number of peptides observed in a 206

sample. For statistical analysis, normalized peptide counts in intact and lysed 207

samples were pooled for a given growth condition by combining all timepoints 208

from across the growth curve. Student’s t-test was used to compare the 209

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 8: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

8

8

normalized peptide counts from NDFO and organotrophic growth conditions. P-210

values≤0.05 were considered to be significant. 211

212

Results and Discussion 213

Evidence against an inducible Fe(II) oxidoreductase in A. ebreus 214

Previous studies on anaerobic Fe(II) oxidation by phototrophic organisms 215

demonstrated induction of putative Fe(II) oxidoreductase proteins and enhanced 216

Fe(II) oxidation when cells were pre-cultured in the presence of Fe(II)-NTA (34). 217

In contrast, pre-culturing of A. ebreus on Fe(II)-NTA was not required for NDFO 218

of mixed insoluble (vivianite/siderite) and soluble Fe(II) in chloramphenicol-219

treated washed cell suspensions without acetate, indicating a basal capacity for 220

NDFO (Fig. 1a). Similar results were obtained in the absence of chloramphenicol 221

(Supp. Fig. S1). Furthermore, pre-culturing in the presence of Fe(II)-NTA did not 222

enhance the rate or extent of Fe(II) oxidation of mixed soluble and insoluble 223

Fe(II) (Fig. 1a). 224

In the model acidophilic Fe(II) oxidizer Acidothiobacillus ferroxidans, outer 225

membrane multiheme c-type cytochromes and rusticyanin, a periplasmic copper 226

protein, are proposed to be involved in Fe(II) oxidation (35). In phototrophic iron 227

oxidizers Rhodobacter sp. SW2 and Rhodopseudomonas palustris TIE-1, outer 228

membrane porins (PioB), periplasmic multiheme c-type cytochromes 229

(PioA/FoxE), and periplasmic high-potential iron proteins (PioC) or quinoproteins 230

(FoxY) are involved (35). Similar c-type cytochrome-based mechanisms are 231

proposed for microaerophilic Fe(II) oxidation (36). While a number of c-type 232

cytochrome genes are present in the A. ebreus genome, there are no multiheme 233

c-type cytochromes or close homologs of the pio/fox genes (35). The genes 234

Dtpsy_1207, Dtpsy_1208, and Dtpsy_2198 have homology with multicopper 235

oxidase (MCO) family proteins, some of which are known to be involved in Mn(II) 236

oxidation, but these proteins typically utilize oxygen as a co-substrate to oxidize 237

metals (37), and were never observed in proteomics experiments; as such, they 238

are unlikely to be involved in NDFO by A. ebreus. 239

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 9: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

9

9

To search for an inducible Fe(II) oxidoreductase, we compared the whole 240

cell lysate and trypsin-shaved proteomes of organotrophic (5 mM acetate/10 mM 241

nitrate) and NDFO (10 mM Fe(II)/5 mM acetate/10 mM nitrate) growth cultures of 242

A. ebreus at 8, 18, 24, 32 and 48 hrs. Representative growth curves and data on 243

the transformation of Fe(II), acetate, and nitrogen oxides are presented in 244

Supplementary Figure S2. Trypsin shaving has previously been shown to be 245

useful for identifying surface exposed proteins, including cytochromes, in other 246

bacteria (38, 39), and we used it to increase our chance of identifying outer 247

membrane proteins which may be contacting Fe(II) and catalyzing cell surface 248

Fe(II) oxidation. When the proteomes of cells harvested at timepoints between 8 249

and 48 hours from NDFO and organotrophic A. ebreus growth cultures are 250

compared, significantly increased abundance (p≤0.05) is observed for heavy 251

metal RND (resistance-nodulation-division) family efflux pumps, Dtpsy_1460 252

(p=0.018), Dtpsy_1461 (p=0.0030) and Dtpsy_1462 (p= 0.00070), (Fig. 2, 253

Dataset S1). In the set of extracytoplasmic proteins that were observed 254

exclusively in NDFO cultures (Dataset S1E), only a few peptides were observed, 255

but efflux pumps are highly represented. The expression of heavy metal efflux 256

pumps in response to anaerobic Fe(II) stress is a novel finding, and suggests 257

that these proteins may be involved in exporting metals other than those 258

commonly considered to be substrates (40, 41). Furthermore, because the RND 259

family efflux pumps identified in A. ebreus are proton/solute antiporters, the 260

energetic cost is putatively less than an ATP-consuming transporter, and RND 261

family efflux pumps are able to load substrate from either the periplasm or 262

cytoplasm (42). 263

The proteomics data also suggest a cytoplasmic response to 264

redox/nitrosative stress. Although the p-values are higher (Dataset S1), 265

dihydrolipoamide dehydrogenase subunits of the pyruvate dehydrogenase 266

complex, Dtpsy_1140 (p=0.074) and Dtpsy_1658 (p=0.070), which are known to 267

be sensitive to nitric oxide (NO) (43), were more abundant. Also more abundant 268

were proteins involved in fatty acid synthesis, enoyl-CoA hydratase, Dtpsy_2933 269

(p=0.11), and a member of the 2-nitropropane dioxygenase family, Dtpsy_2296 270

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 10: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

10

10

(p=0.055) which is closely related to a Neisseria gonorrhoeae homolog 271

(NGO1024) putatively involved in anaerobic, NO-dependent fatty acid synthesis 272

(44). Aconitase, Dtpsy_1066 (p= 0.25), which is sensitive to nitrosative stress 273

(45), and the large subunit of NO reductase Dtpsy_0109 (p=0.13) (46), were also 274

observed more often in Fe(II) cultures, but the p-values are higher (Dataset S1). 275

Two proteins, Dtpsy_0452 (p=0.030) and Dtpsy_1098 (p=0.045), were 276

significantly less abundant in NDFO cultures of A. ebreus (Dataset S1). Given 277

that peptides from core periplasmic and inner-membrane respiratory proteins 278

were observed in the proteomics experiments, it is unlikely that a respiratory 279

ferroxidase was missed. Several periplasmic c-type cytochromes were observed 280

(Dataset S1), including Dtpsy_0754, a component of the cytochrome bc1 281

complex, but there was no significant difference in their abundance between 282

organotrophic and NDFO proteomes (Dataset S1). Overall, the proteomic 283

response to Fe(II) is consistent with a cell envelope and cytoplasmic stress 284

response to Fe(II) and nitrogen oxides rather than induction of a respiratory 285

metabolism. 286

287

Fe(II) is toxic to nitrate reducing cultures of A. ebreus 288

In support of Fe(II) toxicity to nitrate-reducing cells, when 0.1 mM Fe(II) 289

was added to organotrophic cultures of A. ebreus, a growth lag of between 12-18 290

hours was observed (Fig. 1b). Similar concentrations of Fe(II) are inhibitory to 291

the growth of Rhodobacter capsulatus grown under humic acid oxidizing 292

phototrophic conditions (19). A lag was also observed in NDFO cultures of A. 293

ebreus containing higher Fe(II) concentrations (8 to 10 mM) (Supp. Fig. S2b, c), 294

and NO2- and NO both accumulated to at least ten-fold higher levels in the NDFO 295

cultures relative to organotrophic (no Fe) cultures (Supp. Fig. S2d, e). The 296

growth lag in NDFO cultures was partially relieved by addition of NTA with an 297

accompanying increase in the rate of Fe(II) oxidation (Supp. Fig. S3a, b). We 298

also observed less NO accumulation in the headspace of NDFO cultures grown 299

with Fe(II)-NTA compared with unchelated Fe(II) (Supp. Fig. S3c). NO toxicity is 300

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 11: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

11

11

likely less of a problem in the presence of Fe(II)-NTA which forms a stable NO 301

complex with a sub-micromolar Kd (47). 302

303

Possible mechanisms for NDFO by A. ebreus and other nitrate-reducing 304

bacteria 305

In the absence of an inducible Fe(II) oxidoreductase, one or a combination 306

of several mechanisms for NDFO are possible, including (i) direct electron 307

donation from Fe(II) to respiratory complexes if the appropriate electron acceptor 308

is available (i.e. NO3- for Nar, NO2

- for Nir, etc), (ii) extracellular or intracellular 309

mineral phase Fe(II) oxidation coupled to the reduction of NO3- or biogenic NO2

- 310

and NO (48) (green rusts have been shown to form during NDFO by Acidovorax 311

sp. BoFeN1 (49), (iii) proton dependent abiotic Fe(II) reduction of NO2- at lower 312

pH conditions that exist in the bacterial periplasm, and (iv) Fe(II) oxidation 313

coupled to the reduction of NO2- and NO catalyzed by periplasmic components 314

such as protein thiols (3, 41, 50) or membrane bound iron (51, 52). 315

316

Fe(II) must enter the periplasm for extensive oxidation by A. ebreus 317

NDFO catalysis in the periplasm requires that Fe(II) cross the outer 318

membrane, which, though porous, is negatively charged and will bind Fe(II) (52). 319

A long lag (~6hrs) precedes Fe(II) oxidation (Fig. 1a) and nitrate reduction (Supp. 320

Fig. S5a) in cell suspensions with mixed soluble and insoluble Fe(II), consistent 321

with the outer membrane acting as a barrier to Fe(II) entry into the periplasm 322

(Fig. 1a, Supp. Fig. S1, Supp. Fig. S5a). When A. ebreus cells are lysed, 323

releasing periplasmic components, NDFO commences earlier (Fig. 3a). Fe(II)-324

NTA, which will not adsorb to cell surfaces because the Fe(II) is bound by a 325

chelator, is oxidized rapidly in the first 6 hours of the incubation (Fig. 3b). A more 326

extensive treatment of the effect of metal ligands on accelerating the rate of 327

microbial Fe(II) oxidation was recently published (53), and the authors 328

demonstrate that NO2- will react more rapidly with Fe(II)-NTA than un-chelated 329

Fe(II), suggesting another possible explanation for the rapid rate of Fe(II)-NTA 330

oxidation by A. ebreus (Fig. 3b) (53). However, Fe(II)-NTA-agarose beads, 331

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 12: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

12

12

which cannot penetrate the outer-membrane, showed no oxidation by A. ebreus 332

cell suspensions (Fig. 3b). This result suggests that Fe(II)-NTA must enter the 333

periplasm to be oxidized, and argues against the involvement of inducible outer-334

membrane redox active proteins or a secreted oxidant in NDFO of Fe(II)-NTA by 335

A. ebreus. Previous observations from a scanning transmission x-ray 336

microscopy (STXM) study found that the rate of Fe(II) oxidation by Acidovorax 337

BoFeN1 is faster at the cells, and specifically in the periplasm, compared to at 338

the extracellular Fe(II) (26). This is consistent with periplasmic access of Fe(II) 339

being required for robust NDFO by members of the Acidovorax genus. 340

We sought to understand how NDFO occurs in the periplasm. 341

Cytochrome complexes of intact cells are reduced by Fe(II)-NTA (Supp. Fig. S4), 342

as has previously been observed with other organisms (5, 18) and acetate is not 343

required for NDFO in cell suspensions (Fig 1a). Thus, we hypothesized that 344

respiratory complexes could directly mediate NDFO. To test this, we deprived 345

cells of molybdenum, the metal cofactor of the nitrate reductase, Nar. A. ebreus 346

cells grown anaerobically in the absence of molybdenum under nitrite-reducing 347

conditions, but with NO3- available to induce Nar expression, were unable to 348

either grow by organotrophic NO3- reduction (data not shown) or oxidize Fe(II) 349

coupled to NO3- reduction in cell suspensions (Fig. 3c). These results suggest 350

that either Nar directly catalyzes NO3- reduction coupled to Fe(II) oxidation (54) 351

and/or NO2- is produced to further react with Fe(II) through the abiotic and biotic 352

mechanisms previously discussed. 353

354

Both nitrite and nitric oxide can oxidize extracellular solid phase Fe(II) 355

Previously, it was observed that synthetic solid-phase iron phosphate 356

(vivianite) in the absence of soluble Fe(II) is not oxidized by growth cultures of 357

Acidovorax species (13, 26). In contrast, we found that small amounts of 358

synthetic vivianite (~0.5 mM Fe(II) over the course of 24 hours) can be oxidized 359

by cell suspensions of A. ebreus, but only in the presence of an organic co-360

substrate (Fig. 4a). As the cell suspensions were treated with chloramphenicol, it 361

is unlikely that the cells produced an organic chelator as a result of acetate 362

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 13: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

13

13

amendment. An alternative option is that intermediates of nitrate respiration 363

escape the periplasm and solubilize and/or oxidize solid-phase iron. Candidates 364

for a secreted oxidant include NO2- or, when soluble Fe(II) is also present in the 365

system, Fe(III) (13, 26). Some NDFO organisms, such as Pseudogulbenkiania 366

ferroxidans (11) or Pseudogulbenkiania sp. MAI-1 (53), release higher 367

concentrations of extracellular NO2- during organotrophic nitrate reduction, which 368

can react with extracellular Fe(II) especially in the presence of catalytic mineral 369

phases, such as green rusts (21, 48, 49), or chelating ligands (53). However, NO 370

should also be considered. NO is produced during organotrophic nitrate 371

reduction by A. ebreus (Supp. Fig. S2e), and other nitrate reducers (55, 56). NO 372

oxidizes Fe(II) and is a good Fe(II) ligand that will displace a number of anionic 373

ligands (57). NO oxidizes synthetic vivianite Fe(II) faster than NO2- (Fig. 4c). 374

After 15 minutes, ~0.5 mM synthetic vivianite Fe(II) is oxidized by 0.5 mM DEA 375

NONOate, but no noticeable synthetic vivianite Fe(II) oxidation occurred in the 376

presence of 1 mM NO2- (Fig. 4c). Even low steady-state concentrations of NO 377

may be enough to account for the small amount of synthetic vivianite oxidized 378

when acetate is added to dense cell suspensions (Fig. 4a). NO can also 379

solubilize Fe(II) from synthetic vivianite (Fig. 4b) and increase the rate of Fe(II) 380

oxidation via the periplasmic reactions previously discussed. We see 381

accumulation of solution NO (Supp. Fig. S6a) in later phases of cultures along 382

with a rebound in the concentration of soluble Fe(II) (Supp. Fig. S6b). The 383

observation that NO, and to a lesser extent NO2- can oxidize solid phase 384

extracellular Fe(II) (Fig. 4c) is ostensibly at odds with the observation that Fe(II)-385

NTA agarose was not oxidized by A. ebreus (Fig. 3b). However, NO binds to 386

Fe(II)-NTA with nanomolar affinity to form stable Fe(II)-NTA-NO complexes (58), 387

but does not oxidize the chelated Fe(II) (47). Although abiotic oxidation of Fe(II)-388

NTA by NO2- is faster than the un-chelated form at neutral pH (53), negligible 389

amounts of NO2- are produced by A. ebreus during organotrophic denitrification 390

(Supp. Fig. S2d). 391

It has been proposed that a local low pH around iron oxidizing cells could 392

keep Fe(III) soluble and allow it to diffuse out of cells to react with extracellular 393

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 14: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

14

14

vivianite (26). However, we have direct evidence for NDFO-associated efflux 394

pumps from proteomics (Fig. 2, Dataset S1), suggesting a possible active 395

mechanism for secreting Fe(III) from cells to react with extracellular Fe(II) 396

minerals through an electron exchange mechanism. RND family efflux pumps 397

have relaxed specificity and can transport a wide range of metal ions (40). This 398

electron exchange mechanism may involve reactions between extracellular 399

colloidal iron and extracellular minerals (13, 26). Previous reports that found 400

NDFO of solid phase Fe(II) in batch cultures could be explained by a combination 401

of the solubilization of Fe(II) by NO, oxidation by NO, NO2- and Fe(III), and lysis 402

and cryptic growth in long incubations (4, 5, 59). 403

404

Acetate is utilized before Fe(II) as an electron donor for nitrate-reduction. 405

Excess acetate inhibits NDFO and NO2-, NO and N2O accumulation 406

Previously, it was noted that acetate is consumed before Fe(II) in NDFO 407

growth cultures (5, 60) and this is also true for A. ebreus (Supp. Fig. S2b). For 408

other Acidovorax species in growth culture under donor-limited conditions, 409

increasing acetate concentrations increased the extent of iron oxidation (12), 410

likely due, at least in part, to increasing biomass. However, we observed that 411

under acceptor-limited conditions in growth cultures, excess acetate decreased 412

the rate and extent of Fe(II) oxidation and NO2-, NO and N2O accumulation by A. 413

ebreus (Supp. Fig. S2, S7a). To understand the effect of acetate on iron 414

oxidation without growth or cell number as confounding variables, we performed 415

cell suspensions with chloramphenicol. Remarkably, the presence of acetate 416

also controls Fe(II) oxidation and NO2- and NO accumulation in cell suspensions 417

(Fig. 5, Supp. Fig. S5). In donor/acceptor balanced conditions (10 mM Fe(II), 5 418

mM acetate, 10 mM nitrate), the most rapid phase of Fe(II) oxidation occurred 419

once acetate was depleted to below 0.5 mM (between 24-36 hrs) (Fig. 5a). A 420

similar rapid phase of Fe(II) oxidation at acetate concentrations below 0.5 mM is 421

observed in NDFO growth cultures (Supp. Fig. S2b). However, in acceptor-422

limited conditions (10 mM Fe(II), 10 mM acetate, 10 mM nitrate), this effect was 423

greatly diminished because acetate was consumed prior to over Fe(II) (Fig. 5b, 424

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 15: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

15

15

Supp. Fig. S2c, Supp. Fig. S5b). Consistent with the observations made with A. 425

ebreus, acetate is consumed before Fe(II) in NDFO growth cultures of Azospira 426

suillum strain PS (5) (Supp. Fig. S7b) and excess acetate inhibits Fe(II) oxidation 427

under acceptor-limited conditions (Supp. Fig. S7c). 428

A recent paper from Klueglein and Kappler (29) makes the claim that 429

enzymatic NDFO is unlikely to occur in Acidovorax sp. BoFeN1 in part based on 430

the fact that the organism does not oxidize an initial concentration of 7 mM Fe(II) 431

in a growth culture with 5 mM acetate (40 mM electron equivalents), 100 μM 432

NO3- (0.5 mM electron equivalents) and ~20mM N2O (40 mM electron 433

equivalents) (5% added to headspace). If acetate is preferentially consumed by 434

Acidovorax sp. BoFeN1, very few electron equivalents of N2O will remain for 435

Fe(II) oxidation to proceed coupled to N2O reduction. The authors mention, but 436

do not show, the data from a similar growth experiment with 0.5 mM acetate 437

instead of 5 mM acetate. Of course, in this instance, very little growth is 438

expected and the concentration of cells available to catalyze NDFO would be 439

very low. While our results suggest that NDFO, particularly of chelated Fe(II) 440

(Figure 2b), can be catalyzed in the periplasm of nitrate-reducing bacterial cells, 441

more work is needed to define precisely the catalytic sites. 442

443

A model for NDFO as an innate capability of nitrate-reducing bacteria 444

In NDFO growth cultures, when acetate is present, periplasmic respiratory 445

complexes remain reduced and NO/NO2- concentrations remain low. Also, metal 446

efflux pumps may function to keep Fe(II) out of the periplasm (Fig. 2) (61). When 447

acetate is depleted and respiratory complexes are oxidized, Fe(II) efflux is 448

slowed and periplasmic NDFO reactions can proceed. Fe(II) oxidation coupled 449

to reduction of intermediates of nitrate reduction may be directly catalyzed by 450

Nar, Nir and Nor (54), or other periplasmic components or through abiotic 451

reactions that can be accelerated by ligands (53). If the rate of Fe(II) oxidation 452

and Fe(III) mineral precipitation is faster than the rate of efflux, precipitation of 453

Fe(II) minerals on periplasmic proteins will lead to their inhibition and NO2-, NO, 454

and N2O accumulation (Supp. Fig. S2d-f) (13, 14, 62). In our data, NO2-, NO and 455

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 16: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

16

16

N2O did not accumulate to their maximum steady-state level and acetate was not 456

completely consumed in NDFO cultures or cell suspensions until after the more 457

rapid phase of Fe(II) oxidation (Supp. Fig. S2b, S2d-f, Fig. 5, Supp. Fig. S5). 458

This result suggests that a catalyst of Fe(II) oxidation is depleted alongside a 459

catalyst of NO2-, NO and N2O reduction. As no inducible Fe(II) oxidoreductase 460

appears to be present in A. ebreus (Dataset S1), the simplest explanation of our 461

results is that an important site of Fe(II) oxidation is the same as the site of NO2-, 462

NO and N2O reduction (i.e. the redox-active proteins Nar, Nir, Nor and Nos and 463

other redox-active components of the periplasm). A model for NDFO consistent 464

with our results and results in the literature is presented in Figure 6. 465

We postulate that the periplasmic components of all nitrate-reducing 466

bacteria are innately capable of catalyzing a baseline level of NDFO. The rate 467

and extent of NDFO will depend on differences in the respiratory chain 468

components, the Fe(II) form, the accumulation of respiratory intermediates (NO2-, 469

NO), the concentration of the organic co-substrate and the capacity of the 470

bacteria to overcome Fe(II) and nitrogen oxide toxicity through efflux pumps and 471

detoxification mechanisms. While an inducible Fe(II) oxidoreductase may exist, 472

our results suggest that it is not necessary for a nitrate-reducing organism to be 473

proficient at catalyzing NDFO. 474

Some previous work suggests a growth benefit from NDFO. In batch 475

cultures the effect is subtle (12, 14), but continuous flow may enhance the growth 476

benefit (14). We did not observe a significant growth benefit due to NDFO in our 477

batch cultures of A. ebreus (Supp. Fig. S8), but it may be that any energetic gain 478

from Fe(II) oxidation is largely masked by the energetic cost of coping with Fe(II) 479

toxicity. We have recently proposed several possible mechanisms whereby an 480

organism could gain energetic benefit from NDFO (3). Future work should focus 481

on demonstrating which, if any, of these dominates in different bacteria. 482

Preliminary proteomics data suggest that efflux pumps are induced during 483

NDFO in both A. suillum and P. ferroxidans, but genetic knockouts and other 484

experiments are needed to precisely define their role. Also, more studies are 485

needed on the reactivity of NO and NO2- with various geochemically important 486

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 17: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

17

17

Fe(II) forms. An awareness of the accumulation of NO is important to a thorough 487

understanding of the influence of Fe(II) on biological systems. For example, 488

recent work has emphasized the importance of NO in regulating bacterial motility 489

and biofilm formation by environmentally important species (25, 63), including the 490

environmentally important Fe(III) reducer, Shewanella oneidensis (64), but 491

almost nothing is known about the biogeochemical controls on NO production in 492

the environment. Finally, both NO and N2O are well-known atmospheric 493

pollutants, and the emissions of NO and N2O from diverse environments have 494

been compared (65, 66). However, to our knowledge, no study has assessed 495

the importance of organic electron donors in controlling NO and N2O emissions in 496

Fe(II) rich systems. Our results suggest that NDFO may be a major contributor 497

to the flux of nitrogen oxide intermediates in Fe(II)-rich systems limited for 498

organic electron donor. 499

Acknowledgements 500

The authors would like to thank members of the Coates lab for stimulating 501

discussions and critical comments on the experiments and manuscript, Mary 502

Firestone for advice on experiments, Michael Marletta, Lars Plate and other 503

members of the Marletta lab for assistance with measuring nitric oxide and 504

stimulating discussions, the Energy Biosciences Institute for funding, and the 505

National Institutes of Health for LC-MS instrumentation (grant 1S10RR022393-506

01). We would also like to thank anonymous reviewers for helpful insights and 507

comments. 508

509

510

511

Figure Legends: 512 Fig. 1. NDFO of mixed soluble and insoluble Fe(II) is not inducible, but 513

Fe(II) is toxic to nitrate-reducing A. ebreus. a. Concentration of Fe(II) 514

(soluble and vivianite) in chloramphenicol treated washed cell suspensions of 515

Acidovorax ebreus. Cells were pre-grown organotrophically (10 mM acetate, 10 516

mM nitrate) in the absence of Fe(II) or under NDFO conditions in the presence of 517

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 18: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

18

18

Fe(II)-NTA (10 mM Fe(II)-NTA, 5 mM acetate, 10 mM nitrate) b. Growth 518

expressed as OD 600 in organotrophic cultures (5 mM acetate, 10 mM nitrate) of 519

Acidovorax ebreus containing 0 mM, 0.01 mM or 0.1 mM Fe(II). Points represent 520

averages and error bars represent the standard deviations of triplicates. 521

Fig. 2. Heavy metal efflux pumps are more abundant in NDFO cultures. 522

Normalized peptide counts for Dtpsy_1460, Dtpsy_1461 and Dtpsy_1462 in 523

NDFO and organotrophic cultures at 8, 18, 24, 36, and 48 hours. Top row: Box 524

plots showing the lower extreme, lower quartile, median, upper quartile and 525

upper extreme of normalized peptide counts. Data from the time courses were 526

pooled from trypsin-shaving proteomics experiments. Data is overlayed on the 527

box plots. Bottom row: Normalized peptide count data at 8, 18, 24, 36, and 48 528

hours in NDFO and organotrophic cultures. 529

Fig. 3. Periplasmic reactions are essential to initiate extensive NDFO. a. 530

Concentration of Fe(II) in NDFO cell suspensions of lysed or intact Acidovorax 531

ebreus in the presence of chloramphenicol in media containing 10 mM Fe(II), 10 532

mM nitrate. b. Concentration of Fe(II) as Fe(II)-NTA or Fe(II)-NTA agarose 533

beads in NDFO cell suspensions (10 mM Fe(II), 5 mM acetate, 10 mM nitrate) in 534

the absence of chloramphenicol. c. Concentration of Fe(II) in NDFO cell 535

suspensions of molybenum-depleted Acidovorax ebreus in the presence of 536

chloramphenicol in either molybdenum-free or molybdenum-replete media (10 537

mM Fe(II), 10 mM nitrate). Cells of A. ebreus were pre-grown in molybdenum-538

free media (5 mM acetate, 10 mM nitrite, 10 mM nitrate) and were incapable of 539

growth by organotrophic nitrate reduction (not shown). Points represent 540

averages and error bars represent the standard deviations of triplicates. 541

Fig. 4. Extracellular reactions between NO2- and NO and mineral-phase 542

Fe(II) can contribute to NDFO. a. Concentration of Fe(II) with synthetic 543

vivianite as the sole Fe(II) source in NDFO cell suspensions of Acidovorax 544

ebreus (4 mM Fe(II), 10 mM nitrate and 0 mM acetate or 5 mM acetate) in the 545

presence of chloramphenicol. b. Concentration of soluble Fe(II) after 15 minutes 546

in cell-free basal media containing 4 mM Fe(II) in the form of synthetic vivianite 547

and either 1 mM sodium nitrite (NO2-) or 0.5 mM diethylamine NONOate (NO) or 548

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 19: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

19

19

both compounds. Controls contained diethylamine and sodium hydroxide and 549

showed no Fe(II) oxidation over the time of the experiment. c. Concentration of 550

total Fe(II) after 15 minutes and 24 hours in cell-free basal media containing 4 551

mM Fe(II) in the form of synthetic vivianite and either 1 mM sodium nitrite or 0.5 552

mM diethylamine NONOate (NO releasing compound) or both compounds. 553

Controls contained diethylamine and sodium hydroxide and showed no Fe(II) 554

oxidation over the time of the experiment. Points represent averages and error 555

bars represent the standard deviations of triplicates. 556

Fig. 5. Acetate is consumed prior to Fe(II) as an electron donor in nitrate-557

reducing cultures. a. Concentrations of Fe(II), acetate, nitrate, and nitrite in 558

NDFO cell suspensions of Acidovorax ebreus (10 mM Fe(II), 5 mM acetate, 10 559

mM nitrate) in the presence of chloramphenicol. Points represent averages and 560

error bars represent the standard deviations of triplicates. b. Fe(II) 561

concentrations in NDFO cell suspensions of Acidovorax ebreus with 10 mM 562

Fe(II), 10 mM nitrate and 0 mM acetate, 5 mM acetate or 10 mM acetate. Points 563

represent averages and error bars represent the standard deviations of 564

triplicates. 565

Fig. 6. A mechanistic model for nitrate-dependent Fe(II) oxidation by A. 566

ebreus. Across a representative NDFO growth curve dominant electron flow 567

pathways are proposed for before (a), during (b) and after (c) the most rapid 568

phase of Fe(II) oxidation. Dominant reactions in each panel are indicated by bold 569

arrows, minor/inhibited reactions are indicated by dashed arrows. a. In the 570

presence of acetate, respiratory complexes remain reduced by electrons from the 571

quinone pool and enzymatic NO3-, NO2

-, NO and N2O reduction proceeds. 572

Periplasmic Fe(II) concentrations are low due to metal efflux pumping , and few 573

reactions between Fe(II) and reduced respiratory complexes occur. Some 574

extracellular reactions between Fe(II) and NO2-/NO may proceed depending on 575

how much NO2- and NO accumulate outside of the cell. b. Once acetate has 576

dropped to a low level, Fe(II) will accumulate in the periplasm and be oxidized 577

coupled to nitrogen oxide reduction at respiratory complexes and other 578

periplasmic components in the most rapid phase of Fe(II) oxidation. c. As Fe(II) 579

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 20: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

20

20

is oxidized at respiratory complexes, mineral precipitates will begin to form in 580

association with these proteins, leading to their inhibition and the accumulation of 581

NO2-, NO and N2O. Extracellular NO2

- and NO can oxidize solid phase Fe(II), but 582

NO can also solubilize Fe(II) which could diffuse back into cells to be oxidized. 583

Ultimately, periplasmic Fe(III) precipitation damages respiratory complexes and 584

other periplasmic sites, leading to a slower rate of NDFO, primarily driven by 585

abiotic reactions. 586

587 588 589 590

References Cited 591 592

1. Weber, K. A., L. A. Achenbach, and J. D. Coates. 2006. 593 Microorganisms pumping iron: anaerobic microbial iron oxidation and 594 reduction. Nat. Rev. Microbiol. 4:752-64. 595

2. Canfield, D. E., M. T. Rosing, and C. Bjerrum. 2006. Early anaerobic 596 metabolisms. Philos. Trans. R. Soc. Lond. B Biol. Sci. 361:1819-34; 597 discussion 1835-6. 598

3. Carlson, H. K., I. C. Clark, R. A. Melnyk, and J. D. Coates. 2012. 599 Toward a mechanistic understanding of anaerobic nitrate-dependent iron 600 oxidation: balancing electron uptake and detoxification. Front. Microbiol. 601 3:57. 602

4. Weber, K. A., F. W. Picardal, and E. E. Roden. 2001. Microbially 603 catalyzed nitrate-dependent oxidation of biogenic solid-phase Fe(II) 604 compounds. Environ. Sci. Technol. 35:1644-1650. 605

5. Chaudhuri, S. K., J. G. Lack, and J. D. Coates. 2001. Biogenic 606 magnetite formation through anaerobic biooxidation of Fe(II). Appl. 607 Environ. Microbiol. 67:2844-2848. 608

6. Widdel, F., S. Schnell, S. Heising, A. Ehrenreich, B. Assmus, and B. 609 Schink. 1993. Ferrous iron oxidation by anoxygenic phototrophic bacteria. 610 Nature 362:834-836. 611

7. Machulla, G., J. Thieme, and J. Niemeyer. 1998. Interaction of 612 microorganisms with soil colloids observed by X-ray microscopy, p. II-21 – 613 II-28. In J. Thieme, G. Schmahl, D. Rudolph, and E. Umbach (ed.), X-Ray 614 Microscopy and Spectroscopy. Springer-Verlag, Heidelberg. 615

8. Straub, K. L., and B. E. E. Buchholz-Cleven. 1998. Enumeration and 616 detection of anaerobic ferrous iron-oxidizing, nitrate-reducing bacteria 617 from diverse European sediments. Appl. Environ. Microbiol. 64:4846-618 4856. 619

9. Straub, K. L., M. Benz, B. Schink, and F. Widdel. 1996. Anaerobic, 620 nitrate-dependent microbial oxidation of ferrous iron. Appl. Environ. 621 Microbiol. 62:1458-1460. 622

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 21: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

21

21

10. Hafenbradl, D., M. Keller, R. Dirmeier, R. Rachel, P. Robnagel, S. 623 Burggraf, H. Huber, and K. O. Stetter. 1996. Ferroglobus placidus gen. 624 nov., sp. nov. a novel hyperthermophilic archaeum that oxidizes Fe2+ at 625 neutral pH under anoxic conditions. Arch. Microbiol. 166:308-314. 626

11. Byrne-Bailey, K. G., K. A. Weber, A. H. Chair, S. Bose, T. Knox, T. L. 627 Spanbauer, O. Chertkov, and J. D. Coates. 2010. Completed genome 628 sequence of the anaerobic iron-oxidizing bacterium Acidovorax ebreus 629 strain TPSY. J. Bacteriol. 192:1475-6. 630

12. Muehe, E. M., S. Gerhardt, B. Schink, and A. Kappler. 2009. 631 Ecophysiology and the energetic benefit of mixotrophic Fe(II) oxidation by 632 various strains of nitrate-reducing bacteria. FEMS Microbiol. Ecol. 70:335-633 343. 634

13. Kappler, A., B. Schink, and D. K. Newman. 2005. Fe(III)-mineral 635 formation and cell encrustation by the nitrate-dependent Fe(II)-oxidizer 636 strain BoFeN1. Geobiology 3:235-245. 637

14. Chakraborty, A., E. E. Roden, J. Schieber, and F. Picardal. 2011. 638 Enhanced growth of Acidovorax sp. strain 2AN during nitrate-dependent 639 Fe(II) oxidation in batch and continuous-flow systems. Appl. Environ. 640 Microbiol. 77:8548-8556. 641

15. Brons, H. J., W. R. Hagen, and A. J. Zehnder. 1991. Ferrous iron 642 dependent nitric oxide production in nitrate reducing cultures of 643 Escherichia coli. Arch. Microbiol. 155:341-7. 644

16. Koppenol, W. H. 1996. Thermodynamics of reactions involving nitrogen-645 oxygen compounds. Methods Enzymol. 268:7-12. 646

17. Dutton, A. S., J. M. Fukuto, and K. N. Houk. 2005. Theoretical reduction 647 potentials for nitrogen oxides from CBS-QB3 energetics and (C)PCM 648 solvation calculations. Inorg. Chem. 44:4024-8. 649

18. Weber, K. A., D. B. Hedrick, A. D. Peacock, J. C. Thrash, D. C. White, 650 L. A. Achenbach, and J. D. Coates. 2009. Physiological and taxonomic 651 description of the novel autotrophic, metal oxidizing bacterium, 652 Pseudogulbenkiania sp. strain 2002. Appl. Microbiol. Biotechnol. 83:555-653 65. 654

19. Poulain, A. J., and D. K. Newman. 2009. Rhodobacter capsulatus 655 Catalyzes Light-Dependent Fe(II) Oxidation under Anaerobic Conditions 656 as a Potential Detoxification Mechanism. Appl. Environ. Microbiol, 657 75:6639-6646. 658

20. Dunning, J. C., Y. Ma, and R. E. Marquis. 1998. Anaerobic killing of oral 659 streptococci by reduced, transition metal cations. Appl. Environ. Microbiol. 660 64:27-33. 661

21. Picardal, F. 2012. Abiotic and Microbial Interactions during Anaerobic 662 Transformations of Fe(II) and NOx. Front. Microbiol. 3:112. 663

22. Gurd, F. R., and P. E. Wilcox. 1956. Complex formation between metallic 664 cations and proteins, peptides and amino acids. Adv. Protein Chem. 665 11:311-427. 666

23. McCleverty, J. A. 2004. Chemistry of nitric oxide relevant to biology. 667 Chem. Rev. 104:403-18. 668

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 22: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

22

22

24. Harrop, T. C., Z. J. Tonzetich, E. Reisner, and S. J. Lippard. 2008. 669 Reactions of synthetic [2Fe-2S] and [4Fe-4S] clusters with nitric oxide and 670 nitrosothiols. J. Am. Chem. Soc. 130:15602-10. 671

25. Spiro, S. 2007. Regulators of bacterial responses to nitric oxide. FEMS 672 Microbiol. Rev. 31:193-211. 673

26. Miot, J., K. Benzerara, G. Morin, A. Kappler, S. Bernard, M. Obst, C. 674 Ferard, F. Skouri-Panet, J. M. Guigner, N. Posth, M. Galvez, G. E. 675 Brown, and F. Guyot. 2009. Iron biomineralization by anaerobic 676 neutrophilic iron-oxidizing bacteria. Geochim. Cosmochim. Acta 73:696-677 711. 678

27. Bruce, R. A., L. A. Achenbach, and J. D. Coates. 1999. Reduction of 679 (per)chlorate by a novel organism isolated from paper mill waste. Environ. 680 Microbiol. 1:319-29. 681

28. Keefer, L. K., R. W. Nims, K. M. Davies, and D. A. Wink. 1996. 682 "NONOates" (1-substituted diazen-1-ium-1,2-diolates) as nitric oxide 683 donors: convenient nitric oxide dosage forms. Methods Enzymol. 268:281-684 93. 685

29. Klueglein, N., and A. Kappler. 2012. Abiotic oxidation of Fe(II) by 686 reactive nitrogen species in cultures of the nitrate-reducing Fe(II) oxidizer 687 Acidovorax sp. BoFeN1 - questioning the existence of enzymatic Fe(II) 688 oxidation. Geobiology 11(2):180-190. 689

30. Stookey, L. L. 1970. Ferrozine - a New Spectrophotometric Reagent for 690 Iron. Anal. Chem. 42:779-&. 691

31. Nagababu, E., S. Ramasamy, D. R. Abernethy, and J. M. Rifkind. 692 2003. Active nitric oxide produced in the red cell under hypoxic conditions 693 by deoxyhemoglobin-mediated nitrite reduction. J. Biol. Chem. 278:46349-694 56. 695

32. Zhu, W., J. W. Smith, and C. M. Huang. 2010. Mass spectrometry-based 696 label-free quantitative proteomics. J. Biomed. Biotechnol. 2010:840518. 697

33. Zhang, B., N. C. VerBerkmoes, M. A. Langston, E. Uberbacher, R. L. 698 Hettich, and N. F. Samatova. 2006. Detecting differential and correlated 699 protein expression in label-free shotgun proteomics. J. Proteome Res. 700 5:2909-18. 701

34. Jiao, Y., and D. K. Newman. 2007. The pio operon is essential for 702 phototrophic Fe(II) oxidation in Rhodopseudomonas palustris TIE-1. J. 703 Bacteriol. 189:1765-1773. 704

35. Bird, L. J., V. Bonnefoy, and D. K. Newman. 2011. Bioenergetic 705 challenges of microbial iron metabolisms. Trends Microbiol. 19:330-40. 706

36. Singer, E., D. Emerson, E. A. Webb, R. A. Barco, J. G. Kuenen, W. C. 707 Nelson, C. S. Chan, L. R. Comolli, S. Ferriera, J. Johnson, J. F. 708 Heidelberg, and K. J. Edwards. 2011. Mariprofundus ferrooxydans PV-1 709 the first genome of a marine Fe(II) oxidizing Zetaproteobacterium. PLoS 710 One 6:e25386. 711

37. Tebo, B. M., H. A. Johnson, J. K. McCarthy, and A. S. Templeton. 712 2005. Geomicrobiology of manganese(II) oxidation. Trends Microbiol. 713 13:421-428. 714

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 23: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

23

23

38. Carlson, H. K., A. T. Iavarone, A. Gorur, B. S. Yeo, R. Tran, R. A. 715 Melnyk, R. A. Mathies, M. Auer, and J. D. Coates. 2012. Surface 716 multiheme c-type cytochromes from Thermincola potens and implications 717 for respiratory metal reduction by Gram-positive bacteria. Proc. Natl. 718 Acad. Sci. U S A 109:1702-7. 719

39. Tjalsma, H., L. Lambooy, P. W. Hermans, and D. W. Swinkels. 2008. 720 Shedding & shaving: disclosure of proteomic expressions on a bacterial 721 face. Proteomics 8:1415-28. 722

40. Nies, D. H. 2003. Efflux-mediated heavy metal resistance in prokaryotes. 723 FEMS Microbiol. Rev. 27:313-39. 724

41. Flores-Santana, W., D. J. Salmon, S. Donzelli, C. H. Switzer, D. 725 Basudhar, L. Ridnour, R. Cheng, S. A. Glynn, N. Paolocci, J. M. 726 Fukuto, K. M. Miranda, and D. A. Wink. 2011. The specificity of nitroxyl 727 chemistry is unique among nitrogen oxides in biological systems. Antioxid. 728 Redox. Signal. 14:1659-74. 729

42. Nikaido, H., and Y. Takatsuka. 2009. Mechanisms of RND multidrug 730 efflux pumps. Biochimi. Biophys. Acta-Proteins and Proteomics 1794:769-731 781. 732

43. Richardson, A. R., E. C. Payne, N. Younger, J. E. Karlinsey, V. C. 733 Thomas, L. A. Becker, W. W. Navarre, M. E. Castor, S. J. Libby, and F. 734 C. Fang. 2012. Multiple targets of nitric oxide in the tricarboxylic acid cycle 735 of Salmonella enterica serovar typhimurium. Cell Host Microbe 10:33-43. 736

44. Isabella, V. M., and V. L. Clark. 2011. Identification of a conserved 737 protein involved in anaerobic unsaturated fatty acid synthesis in Neiserria 738 gonorrhoeae: implications for facultative and obligate anaerobes that lack 739 FabA. Mol. Microbiol. 82:489-501. 740

45. Jones-Carson, J., J. R. Laughlin, A. L. Stewart, M. I. Voskuil, and A. 741 Vazquez-Torres. 2012. Nitric oxide-dependent killing of aerobic, 742 anaerobic and persistent Burkholderia pseudomallei. Nitric Oxide 27:25-743 31. 744

46. Rock, J. D., M. J. Thomson, R. C. Read, and J. W. B. Moir. 2007. 745 Regulation of denitrification genes in Neisseria meningitidis by nitric oxide 746 and the repressor NsrR. J. Bacteriol. 189:1138-1144. 747

47. Schneppensieper, T., A. Wanat, G. Stochel, and R. van Eldik. 2002. 748 Mechanistic information on the reversible binding of NO to selected iron(II) 749 chelates from activation parameters. Inorg. Chem. 41:2565-73. 750

48. Hansen, H. C. B., S. Guldberg, M. Erbs, and C. B. Koch. 2001. Kinetics 751 of nitrate reduction by green rusts - effects of interlayer anion and Fe(II): 752 Fe(III) ratio. Appl. Clay Sci. 18:81-91. 753

49. Pantke, C., M. Obst, K. Benzerara, G. Morin, G. Ona-Nguema, U. 754 Dippon, and A. Kappler. 2012. Green rust formation during Fe(II) 755 oxidation by the nitrate-reducing Acidovorax sp. strain BoFeN1. Environ. 756 Sci. Technol. 46:1439-46. 757

50. Wong, P. S., J. Hyun, J. M. Fukuto, F. N. Shirota, E. G. DeMaster, D. 758 W. Shoeman, and H. T. Nagasawa. 1998. Reaction between S-759

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 24: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

24

24

nitrosothiols and thiols: generation of nitroxyl (HNO) and subsequent 760 chemistry. Biochemistry 37:5362-71. 761

51. Coby, A. J., and F. W. Picardal. 2005. Inhibition of NO3- and NO2- 762 reduction by microbial Fe(III) reduction: Evidence of a reaction between 763 NO2- and cell surface-bound Fe2+. Appl. Environ. Microbiol. 71:5267-764 5274. 765

52. French, S., D. Puddephatt, M. Habash, and S. Glasauer. 2012. The 766 dynamic nature of bacterial surfaces: Implications for metal-membrane 767 interaction. Crit. Rev. Microbiol 39(2):196-217. 768

53. Kopf, S. H., C. Henny, and D. K. Newman. 2013. Ligand-Enhanced 769 Abiotic Iron Oxidation and the Effects of Chemical versus Biological Iron 770 Cycling in Anoxic Environments. Environ. Sci. Technol. 47:2602-11. 771

54. Lack, J. G., S. K. Chaudhuri, R. Chakraborty, L. A. Achenbach, and J. 772 D. Coates. 2002. Anaerobic biooxidation of Fe(II) by Dechlorosoma 773 suillum. Microbiol. Ecol. 43:424-31. 774

55. Ye, R. W., B. A. Averill, and J. M. Tiedje. 1994. Denitrification: 775 production and consumption of nitric oxide. Appl. Environ. Microbiol. 776 60:1053-8. 777

56. Price, M. S., L. Y. Chao, and M. A. Marletta. 2007. Shewanella 778 oneidensis MR-1 H-NOX regulation of a histidine kinase by nitric oxide. 779 Biochemistry 46:13677-83. 780

57. Mcdonald, C. C., W. D. Phillips, and H. F. Mower. 1965. An Electron 781 Spin Resonance Study of Some Complexes of Iron Nitric Oxide and 782 Anionic Ligands. J. Am. Chem. Soc. 87:3319-&. 783

58. Kumaraswamy, R., K. Sjollema, G. Kuenen, M. van Loosdrecht, and 784 G. Muyzer. 2006. Nitrate-dependent [Fe(II)EDTA](2-) oxidation by 785 Paracoccus ferrooxidans sp nov., isolated from a denitrifying bioreactor. 786 Systematic and Applied Microbiology 29:276-286. 787

59. Shelobolina, E., H. Xu, H. Konishi, R. Kukkadapu, T. Wu, M. Blothe, 788 and E. Roden. 2012. Microbial Lithotrophic Oxidation of Structural Fe(II) 789 in Biotite. Appl. Environ. Microbiol. 78:5746-52. 790

60. Benz, M., A. Brune, and B. Schink. 1998. Anaerobic and aerobic 791 oxidation of ferrous iron at neutral pH by chemoheterotrophic nitrate-792 reducing bacteria. Arch. Microbiol. 169:159-65. 793

61. Long, F., C. C. Su, M. T. Zimmermann, S. E. Boyken, K. R. 794 Rajashankar, R. L. Jernigan, and E. W. Yu. 2010. Crystal structures of 795 the CusA efflux pump suggest methionine-mediated metal transport. 796 Nature 467:484-U140. 797

62. Miot, J., K. Maclellan, K. Benzerara, and N. Boisset. 2011. Preservation 798 of protein globules and peptidoglycan in the mineralized cell wall of nitrate-799 reducing, iron(II)-oxidizing bacteria: a cryo-electron microscopy study. 800 Geobiology 9:459-70. 801

63. Wang, Y., and E. G. Ruby. 2011. The roles of NO in microbial symbioses. 802 Cell Microbiol 13:518-26. 803

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 25: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

25

25

64. Plate, L., and M. A. Marletta. 2012. Nitric Oxide Modulates Bacterial 804 Biofilm Formation through a Multicomponent Cyclic-di-GMP Signaling 805 Network. Mol. Cell 46:449-60. 806

65. Venterea, R. T., and D. E. Rolston. 2000. Nitric and nitrous oxide 807 emissions following fertilizer application to agricultural soil: Biotic and 808 abiotic mechanisms and kinetics. J. Geophys. Res.-Atmospheres 809 105:15117-15129. 810

66. Beaulieu, J. J., J. L. Tank, S. K. Hamilton, W. M. Wollheim, R. O. Hall, 811 Jr., P. J. Mulholland, B. J. Peterson, L. R. Ashkenas, L. W. Cooper, C. 812 N. Dahm, W. K. Dodds, N. B. Grimm, S. L. Johnson, W. H. McDowell, 813 G. C. Poole, H. M. Valett, C. P. Arango, M. J. Bernot, A. J. Burgin, C. 814 L. Crenshaw, A. M. Helton, L. T. Johnson, J. M. O'Brien, J. D. Potter, 815 R. W. Sheibley, D. J. Sobota, and S. M. Thomas. 2011. Nitrous oxide 816 emission from denitrification in stream and river networks. Proc. Natl. 817 Acad. Sci. U S A 108:214-9. 818

819 820

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 26: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 27: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 28: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 29: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 30: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from

Page 31: Downloaded from on April 5, 2020 by guest · 141 heated conductivity cell. A KOH gradient was generat ed using the EGC III KOH 142 generator at an isocratic flow rate of 1. 5 ml/min

on May 2, 2020 by guest

http://jb.asm.org/

Dow

nloaded from