46
1 NasT-mediated antitermination plays an essential role in the regulation of the 1 assimilatory nitrate reductase operon in Azotobacter vinelandii 2 Running title: Induction of nasAB in Azotobacter vinelandii 3 4 Baomin Wang 1 # , Leland S. Pierson III 2 , Christopher Rensing 1 , Malkanthi K. 5 Gunatilaka 1 , and Christina Kennedy 1 6 7 8 1 The School of Plant Sciences, University of Arizona, Tucson, Arizona 85721, USA 9 2 Department of Plant Pathology and Microbiology, Texas A&M University, College 10 Station, Texas, USA 11 12 13 14 15 16 17 Current address: Center for Agricultural and Environmental Biotechnology, RTI 18 International, Research Triangle Park, NC 27709 19 # Corresponding author. Current address: 20 Department of Medicine, University of Virginia, Charlottesville, VA 22908. 21 E-mail: [email protected] 22 Copyright © 2012, American Society for Microbiology. All Rights Reserved. Appl. Environ. Microbiol. doi:10.1128/AEM.01720-12 AEM Accepts, published online ahead of print on 6 July 2012 on March 12, 2020 by guest http://aem.asm.org/ Downloaded from

Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

1

NasT-mediated antitermination plays an essential role in the regulation of the 1

assimilatory nitrate reductase operon in Azotobacter vinelandii 2

Running title: Induction of nasAB in Azotobacter vinelandii 3

4

Baomin Wang1 #, Leland S. Pierson III2, Christopher Rensing 1�, Malkanthi K. 5

Gunatilaka1, and Christina Kennedy1 6

7 8

1The School of Plant Sciences, University of Arizona, Tucson, Arizona 85721, USA 9

2Department of Plant Pathology and Microbiology, Texas A&M University, College 10

Station, Texas, USA 11

12

13

14

15

16

17

� Current address: Center for Agricultural and Environmental Biotechnology, RTI 18

International, Research Triangle Park, NC 27709 19

#Corresponding author. Current address: 20

Department of Medicine, University of Virginia, Charlottesville, VA 22908. 21

E-mail: [email protected] 22

Copyright © 2012, American Society for Microbiology. All Rights Reserved.Appl. Environ. Microbiol. doi:10.1128/AEM.01720-12 AEM Accepts, published online ahead of print on 6 July 2012

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 2: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

2

Abstract 23

Azotobacter vinelandii is a well-studied model system for nitrogen fixation in bacteria. 24

Regulation of nitrogen fixation in A. vinelandii is independent of NtrB/NtrC, a conserved 25

nitrogen regulatory system in proteobacteria. Previous work showed that a ntrC mutation 26

in A. vinelandii resulted in a loss of induction of assimilatory nitrate and nitrite reductases 27

encoded by the nasAB operon. In addition to NtrC, several other proteins, including 28

NasT, a protein containing a potential RNA-binding domain ANTAR (AmiR and NasR 29

transcription antitermination regulators), have been implicated in nasAB regulation. In 30

this work, we characterize the sequence upstream of nasA and identify several DNA 31

sequence elements, including two potential NtrC binding sites and a putative intrinsic 32

transcriptional terminator upstream of nasA that are potentially involved in nasAB 33

regulation. Our analyses confirm that the nasAB promoter, PnasA, is under NtrC control. 34

However, unlike NtrC-regulated promoters in enteric bacteria, PnasA shows high activity 35

in the presence of ammonium; in addition, the PnasA activity is altered in the nifA¯ 36

background. We discuss the implication of these results on NtrC-mediated regulation in 37

A. vinelandii. Our study provides direct evidence that induction of nasAB is regulated by 38

NasT-mediated antitermination, which occurs within the leader region of the operon. The 39

results also support the hypothesis that NasT binds the promoter proximal hairpin of 40

nasAB for its regulatory function, which contributes to the understanding of the 41

regulatory mechanism of ANTAR-containing antiterminators. 42

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 3: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

3

Introduction 43

Azotobacter vinelandii is a well-studied gram-negative, aerobic diazotroph. Expression of 44

nitrogen-fixation genes in A. vinelandii is tightly regulated by NifA, a transcriptional 45

activator under the control of NifL, an anti-activator protein that detects cellular nitrogen 46

level and redox state (18, 34). In addition, NifL is a receptor of the nitrogen signal from 47

GlnK, a PII-like protein (27, 40, 41, 54, 63) that is modulated by the global nitrogen 48

sensor protein GlnD (15, 43, 57), a bifunctional uridylyltransferase/UMP-removing 49

(UTase/UR) enzyme (30, 41). 50

51

A. vinelandii encodes NtrB/NtrC (64), a conserved two-component regulatory system in 52

proteobacteria involved in control of nitrogen gene expression. In enteric bacteria, 53

NtrB/NtrC, GlnK, and GlnD constitute a global nitrogen regulatory system (45). As a 54

transcriptional activator, NtrC directly controls the expression of the genes related to 55

nitrogen assimilation and metabolism. The activity of NtrC is regulated by NtrB (32), a 56

histidine kinase under the direct control of the PII protein (49). When the cellular 57

nitrogen status is low, NtrC is phosphorylated by NtrB (29, 31). Phosphorylated NtrC 58

activates transcription of target nitrogen genes. In Klebsiella pneumonia, NtrC tightly 59

regulates the expression of nifLA (20, 44), which in turn regulates nif gene transcription 60

in response to oxygen and the nitrogen signal from GlnK (9, 18, 37). However, in A. 61

vinelandii, nifLA is expressed constitutively (8), independent of NtrB/NtrC (64). 62

63

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 4: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

4

Although ntrC is not involved in nif regulation in A. vinelandii, mutational analyses 64

showed that a functional ntrC is required for synthesis of assimilatory nitrate/nitrite 65

reductases (64), encoded by the nasAB operon (52). In addition to NtrC, nasAB 66

expression requires σ54 (56, 64), an alternative sigma factor required by NtrC-regulated 67

promoters. It was also shown that nasAB expression was repressed in the presence of 68

ammonium and induced by the presence of nitrate or nitrite. These observations led to the 69

hypothesis that NtrC in A. vinelandii regulates nasAB expression analogously to NtrC-70

mediated regulation in enteric bacteria (52). However, this hypothesis disagrees with an 71

earlier observation that the promoter of nifLA from K. pneumonia, which is regulated 72

solely by NtrC, was highly expressed in A. vinelandii in the presence of high 73

concentrations of ammonium (33). 74

75

An earlier study identified the gene products of a second operon, nasST, located 10-kb 76

upstream of nasAB and involved in nasAB induction (24). The results of mutational 77

analysis suggest that NasT plays a positive role, while NasS plays a negative role, in 78

nasAB regulation. The predicted NasT protein contains a putative RNA-binding domain 79

that is conserved in several antiterminators, suggesting that the protein may have 80

antitermination functions (22, 39, 47, 61). 81

82

The results from earlier studies also suggest that nasB has a negative autoregulatory role 83

in operon induction, which is evidenced by the finding that nasAB exerts high expression 84

when nasB is mutated even in the absence of nitrate/nitrite induction (24, 52). Adding to 85

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 5: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

5

the complexity of nasAB regulation is the possibility of competition for molybdenum 86

between nitrogenase and nitrate reductase (NasB), which in turn affects nasAB regulation 87

(25). 88

89

The aim of the present study was to elucidate the molecular mechanism of nasAB 90

induction. We determined the promoter of nasAB and characterized its properties. Our 91

results demonstrate that NasT-mediated antitermination plays the essential role in 92

regulatin of nasAB. 93

94

Materials and Methods 95

96

Strains, plasmids, and growth conditions 97

The strains and plasmids used in this study are listed in Table S1. A. vinelandii UW136 98

and its derivatives were grown at 30 °C in modified Burk's nitrogen-free salts medium 99

(48) supplemented with 1% sucrose (BS). When needed, BS medium was supplemented 100

with the following fixed nitrogen sources: ammonium acetate, 15 mM; urea, 10 mM; 101

NaNO2, 5 mM; KNO3 10 mM. E. coli DH5α was grown on Luria-Bertani broth or agar 102

medium at 37 °C. Media were supplemented with antibiotics where appropriate: for A. 103

vinelandii, carbenicillin (20 µg ml-1) and gentamicin (0.05 µg ml-1); for E. coli, 104

carbenicillin (50 µg ml-1) and gentamicin (15 µg ml-1). 105

106

Oligonucleotides 107

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 6: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

6

Oligonucleotides used in this study were purchased from Integrated DNA Technologies, 108

Inc. (Coralville, IA) (Table S2). 109

110

Bioinformatics 111

DNA secondary structure analysis was performed using Mfold (66); the putative σ54 112

binding site and NtrC binding sites were identified using PromScan (http://molbiol-113

tools.ca/promscan/). The sequence of the A. vinelandii genome is available at the website 114

http://www.azotobacter.org/. Protein sequence BLAST against the Conserved Domain 115

Database (CDD) was performed at the NCBI website 116

http://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi.The sequence alignment was 117

created using the EMMA program of the EMBOSS package, and manipulated in 118

GeneDoc (http://www.psc.edu/biomed/genedoc). 119

120

DNA manipulation 121

Plasmid isolations were carried out using the GeneJET Plasmid miniprep kit (Fermentas, 122

Glen Burnie, MD). Restriction enzyme digestions, ligations, cloning, and DNA 123

electrophoresis were performed using standard protocols (55). DNA fragments were 124

purified from agarose gels using the QIAquick Gel Extraction Kit (QIAGEN, Valencia, 125

CA). 126

127

Transformations 128

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 7: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

7

For general cloning, plasmids were transformed into chemically competent E. coli DH5α 129

(55). A. vinelandii transformations were performed on competence medium as described 130

previously (3). 131

132

Cloning the nasAB promoter and surrounding sequence 133

A. vinelandii UW136 genomic DNA was isolated and purified as described previously 134

(53). Purified DNA was digested with XhoI and separated by agarose gel electrophoresis. 135

DNA fragments of approximately 7.8-kb were excised from the gel, purified, and ligated 136

into the XhoI site of pBluescript II KS(+). After transformation into E. coli DH5α, 137

positive clones were identified by colony PCR using primers P35 and P36 (Table S2). 138

The plasmid pWB30 was sequenced using primers P26 and P27 (Table S2) to confirm the 139

cloned sequence. 140

141

For colony PCR in this study, GoTaq Green Master Mix (Promega) was used, the PCR 142

was carried out as follows: 92oC for 2 min, then 30 circles (92oC for 1 min, 55oC for 1 143

min, 68oC for 1 min for extension) for amplification, 1 extension step at 72oC for 5 min. 144

145

Construction of the ntrC deletion mutant 146

A 2.4-kb fragment containing ntrC was amplified from the A. vinelandii chromosome 147

using primer pairs P1 and P2 (Table S2) and cloned into the pGEM-T vector (Promega, 148

Madison, WI). The resulting plasmid was linearized with SacI, blunt-ended using T4 149

DNA polymerase, and self-ligated to yield pWB680. PCR amplification of this plasmid 150

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 8: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

8

was performed using primer pair P3 and P4 (Table S2), the PCR product was digested 151

with SacI and ligated with the SacI-Gm-SacI cassette from pTnMod-OGm, giving rise to 152

pWB685. The orientation of the Gmr cassette with respect to ntrC in pWB685 was 153

confirmed by sequencing. Plasmid pWB685 was transformed into A. vinelandii and a 154

carbenicillin sensitive and gentamicin resistance transformant was selected on BS 155

medium supplied with ammonium acetate (BSN). The ΔntrC::Gm allelic replacement 156

mutation in the transformant was confirmed by colony PCR using the primer pair P1 and 157

P2 (Table S2). 158

159

Construction of lacZ fusion reporterss for A. vinelandii 160

Transcriptional and translational lacZ fusion probes for A. vinelandii were constructed, as 161

described below, such that the two plasmids differed primarily in the multiple cloning 162

sites (MCSs) preceding the lacZ start codon (Fig. 2). 163

164

To construct the backbone of the probes, the β-lactamase (bla) gene and the pMB1 165

replicon were amplified from the plasmid pBluescript II KS(+) using the primer pairs P16 166

and P17 (Table S2). The PCR product was gel purified and self-ligated, giving rise to the 167

plasmid pWhite. A 2.8-kb vnf sequence was amplified from the plasmid pJW1 using 168

primer pairs P33 and P34 (Table S2), digested with SacI and XhoI and cloned into the 169

SacI-SalI sites of pWhite, giving rise to the plasmid pWvnf. 170

171

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 9: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

9

For the translational fusion probe, a fragment containing four tandem copies of the E. coli 172

transcriptional terminator rrnB1 was amplified from pPROBE-NT using primer pairs P8 173

and P9 (Table S2)and cloned into the SphI site of pIC20H using blunt end ligation to 174

create pICT4. The 3-kb BamHI-lacZ-XhoI fragment was released from the plasmid 175

pSUP102::Tn5-B21 using BamHI and XhoI and cloned into the BamHI-XhoI sites of 176

pBluescript II KS(+), yielding the plasmid pBlue-lacZ. The XbaI-lacZ-XhoI fragment, 177

containing the MCS, was removed from pBlue-lacZ and cloned into the XbaI-XhoI sites 178

of pICT4, creating the probe cassette in pIC-lacZ. pIC-lacZ was digested by HindIII and 179

a ~4-kb HindIII-rrnB1-lacZ-HindIII fragment was cloned into the HindIII site of pWvnf, 180

leading to the probe plasmid pVnflacZa. 181

182

Plasmid pKT2lacZ was digested with BamHI and SalI and the BamHI-lacZ-SalI 183

fragment, containing a Shine-Dalgarno (SD) sequence, was cloned into the BamHI-XhoI 184

site of pVnflacZa, leading to the transcriptional fusion probe pVnflacZb, 185

186

Successful insertion of foreign sequences into both probe plasmids was verified by PCR 187

using primers P14 and P25 (Table S2), which amplified the region between the rrnB1 188

terminators and the 70th bp in the lacZ ORF. 189

190

Construction of site-directed and deletion mutations 191

Site specific substitution and deletion mutations were constructed using the overlap 192

extension method based on two rounds of PCR (28). In the first round, two fragments 193

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 10: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

10

were amplified from the wild-type sequence using two pairs of primers: one of each 194

complementary primer pair contained the desired mutation. In the second round, the two 195

products of the first reaction were used as templates. The amplification was performed 196

with two external primers to create a PCR product containing the desired mutation that is 197

the fusion of two fragments from the first round of PCR. PCR reactions were carried out 198

with pfu DNA polymerase (Stratagene) as follows: 95°C for 1 min, then 25 circles (92oC 199

for 1 min, 55oC for 1 min, 72oC for 1min), followed by an extension step at 72 oC for 5 200

min. 201

202

Construction of plasmids for expression analysis in E. coli MC1061 203

nasT was amplified from pMAS20 using primer pairs P37 and P38 (Table S2). The PCR 204

product was digested with EcoRI and BglII and cloned into the EcoRI-BamHI site of the 205

plasmid pDK6 to construct pDK6-T. Using a similar strategy, pDK6-S was assembled by 206

amplifying nasS from pMAS20 using primer pairs P39 and P40 (Table S2) and cloning 207

into pDK6. 208

209

The lacZ reporter plasmid pBTW was engineered from plasmid pBT. The DNA fragment 210

containing chloramphenicol resistance gene, p15A replicon, and lacUV5 promoter was 211

amplified from pBT using primer pairs P41 and P42 (Table S2), and digested with NotI 212

and XhoI. The resulting fragment was linked with the NotI-lacZ-XhoI cassette from the 213

plasmid pblue-lacZ to generate pBTW. The tested DNA sequence can be cloned into the 214

NotI-BamHI region of pBTW for expression analysis. 215

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 11: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

11

216

The DNA fragment containing the nasA leader extending into the 162nd nucleotide of 217

nasA was amplified from the plasmid pWB552 using primer pairs P6 and P43 (Table S2), 218

and cloned into the NotI-BamHI region of pBTW to construct pWB555. Using a similar 219

strategy, DNA fragments were amplified from plasmids pWB650, pWB664, and 220

pWB908 and cloned into pBTW, respectively, to create pBW901, pBW902, and 221

pBW910. 222

223

β-galactosidase assay 224

β-galatosidase activity in E. coli was assayed as described by Miller (1972). β-225

galatosidase activity in A. vinelandii was assayed as following: cultures (2 ml) were 226

centrifuged (3,000 x g) and rinsed once with 1 ml Z buffer (60 mM Na2HPO4, 40 mM 227

NaH2PO4, 10 mM KCl, 1 mM MgSO4, 10 mM dithiothreitol), cell pellets were dissolved 228

in 1 ml Z buffer, and 0.2 ml of cell solutions were used for β-galactosidase assays as 229

described previously and reported as Miller Units (65). 230

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 12: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

12

Results 231

Characterization of the sequence upstream of the NasA start codon 232

233

Genome annotation indicated that the nasAB operon consists of four open reading frames 234

(ORFs) (58) instead of two as identified previously (52). For consistency, we refer to 235

these 4 ORFs as the nasAB operon.One newly identified ORF (Avin23360) is located 236

between NasA and NasB, and the other (Avin23340) is located downstream of NasB. 237

These two ORFs are named NasC and NasH, respectively. Amino acid similarity analysis 238

(Blastp) suggested that the predicted proteins NasC and NasH are homologous to 239

subunits of nitrite reductase (47.57 % identity) and Uroporphyrinogen-III 240

methylase (52.77% identity), respectively, that might be involved in synthesis of 241

prosthetic group of nitrite reductase. Compared to NasA, which is 2,448-nt in length, 242

both NasC and NasH are small (324-nt and 735-nt, respectively). 243

244

The sequence around the NasA start codon was determined originally by Dr. Maria 245

Tortolero and verified by the A. vinelandii genome sequence (58). Using PromScan 246

software (http://molbiol-tools.ca/promscan/), we identified a potential σ54 binding 247

sequence (5’-CTGGCACAGCCCCTGCA-3’, conserved nucleotides are underlined) (5) 248

146-nt upstream of the NasA start codon (Fig. 1a). Two regions of dyad symmetry 249

homologous to the consensus NtrC binding sequence 5’-TGCACCNNNTGGTGCA-3’ 250

(21) were identified 80-nt upstream of the σ54 binding site. The arrangement of the 251

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 13: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

13

predicted σ54 and NtrC binding sites closely resembles the organization of NtrC-regulated 252

promoters in enteric bacteria (36). 253

254

Transcription initiation of σ54 dependent promoters usually begins 12 nt downstream of 255

the σ54 binding site followed closely by the translation initiation region (4). However, the 256

sequence between the predicted transcriptional initiation site and the NasA start codon is 257

135 nt, raising the possibility that additional promoters exist upstream of nasA. 258

Bioinformatic analysis of this region did not identify other potential sigma factor binding 259

sites. Another possibility is that the nasAB transcript has a long 5’ leader sequence with 260

potential regulatory function. Post-transcriptional regulation of many bacterial operons 261

occurs within the 5’ end leader transcript regions and is often used to fine-tune gene 262

expression. Analysis in support of this revealed several potential regulatory features 263

upstream of nasA. 264

265

RNA secondary structure prediction by Mfold (66) revealed three potential hairpin 266

structures located between the putative σ54 binding site and the NasA start codon (Fig. 1 267

A and B). Hairpin I has a 10 bp stem and a hexanucleotide AACGUG loop (ΔG=-13.5 268

kcal/mol), and an adjacent hairpin II has a 6 bp stem and a hexanucleotide ACAGAA 269

loop (ΔG=-8.5 kcal/mol). These two hairpins are separated by one T nucleotide. Hairpin 270

III has a high GC content (ΔG=-28 kcal/mol) and is situated immediately upstream of a 271

poly(T) sequence 25-nt upstream of the NasA start codon. The combination of the hairpin 272

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 14: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

14

and poly(T) sequence resembles an intrinsic transcriptional terminator structure. This 273

suggested the possibility that antitermination is involved in nasAB regulation. 274

275

In addition to the three potential hairpin structures, we identified a small putative ORF 276

encoding the octapeptide MDKGVLAG within this leader region. The small ORF begins 277

at the last base of the middle hairpin II and ends within the loop region of the terminator 278

hairpin (Fig. 1A). 279

280

Construction of LacZ probes specific for A. vinelandii 281

282

A. vinelandii lacks β-galactosidase, enabling lacZ to be used as an expression reporter. 283

Previously, lacZ was integrated into the genomic copy of nasA or nasB for expression 284

studies (52). Using the reporter integrated into the genome of bacteria instead of within a 285

plasmid in trans ensures the consistency of the gene copy number. However, integration 286

of the reporter gene into the nasAB structural genes may disrupt normal operon 287

expression. Previous work showed that lacZ integrated into nasB resulted in higher 288

expression of nasB even in the absence of nitrate induction (24, 52). Based on these 289

considerations, we designed lacZ reporters that can be integrated into A. vinelandii 290

genome, but outside of nasAB region, via homologous recombination. For convenience, 291

we selected the vnf locus for reporter plasmid integration. The vnf genes encode an 292

alternative nitrogenase that uses vanadium as the metal cofactor (7). In addition, vnf is 293

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 15: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

15

not essential for bacterial viability and its expression is inhibited in the presence of 294

molybdenum (7), which was included in the medium used in these studies. 295

296

We constructed two reporter plasmids, pVnflacZa and pVnflacZb, for analysis of nasAB 297

expression (Fig. 2). Plasmid pVnflacZa is a translational fusion vector used to determine 298

the effects of mutations within the region identified upstream of NasA on translation of 299

NasAB. This region with Shine-Dalgarno (SD) sequence was cloned into the multiple 300

cloning site (MCS) XbaI-BcuI-BamHI of pVnflacZa to form an in-frame translational 301

fusion with the LacZ ORF. The lacZ sequence begins with ATGGATCC, where 302

GGATCC is a BamHI restriction site. Plasmid pVnflacZb is a transcriptional fusion 303

vector used to measure promoter activity. This plasmid contains the MCS site XbaI-BcuI-304

BamHI-XbaI followed by a SD sequence (5’-AGGAGGT-3’), located 8-nt upstream of 305

the LacZ start codon. For both plasmids, a four tandem terminator rrnB1 sequence T1(4) 306

was inserted upstream of MCSs to block background lacZ expression (46). 307

308

In addition to the ~2.8 kb vnf sequence, each plasmid contains a pMB1 replicon for 309

propagation in E. coli and a bla gene (encoding β-lactamase) for ampicillin or 310

carbenicillin selection in E. coli and A. vinelandii. We found that the integrated plasmids 311

were stably maintained within the A. vinelandii genome for more than 30 generations 312

even in the absence of antibiotic selection (data not shown). 313

314

Time course expression of the translation fusion nasA΄–΄lacZ in A. vinelandii 315

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 16: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

16

316

A 613-nt sequence starting with nucleotide 190 upstream of the potential NtrC binding 317

sites and ending with the 162nd nucleotide of the nasA gene was selected for further 318

analysis. This fragment was cloned into the XbaI-BamHI region of pVnflacZa, resulting 319

in plasmid pWB552 with an in-frame fusion between nasA΄ and ΄lacZ. One transformant 320

with pWB552 integrated in the genome in the vnf locus was selected for lacZ expression 321

analysis. The transformant was grown in medium containing different nitrogen sources. 322

At multiple time points, culture samples were harvested and analyzed for β-galactosidase 323

activity (Fig. 3). 324

325

In the presence of ammonium, the cellular β-galactosidase activity remained at 326

background levels throughout growth. In contrast, in the presence of nitrate or nitrite, β-327

galactosidase activity increased over time and reached maximal levels after 6 hours, in 328

agreement with the expression profile of lacZ integrated into nasA published previously 329

(42, 52, 64). In the absence of fixed nitrogen sources, A. vinelandii can utilize 330

atmospheric N2 for growth. Under these conditions, β-galactosidase activity remained at 331

background levels, indicating nasAB translation is low even in the absence of ammonium. 332

333

Induction of nasAB by nitrate and nitrite occurs via antitermination 334

335

After confirming that the nasAB sequence selected is suitable for studies of nasAB 336

regulation, we tested whether the predicted transcriptional terminator structure upstream 337

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 17: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

17

of nasA is involved in operon regulation. We constructed a hairpin III deletion mutant, 338

which has the entire hairpin III deleted (Fig. 1B and 4). The hairpin III deletion sequence 339

was then cloned into the XbaI-BamHI region of pVnflacZa to construct a translation 340

fusion. The new construct differed from pWB552 only by the hairpin III region deleted. 341

Analysis of β-galactosidase activity demonstrated that deletion of hairpin III resulted in 342

increased lacZ expression under all conditions tested, indicating that the hairpin III has a 343

negative role in nasAB regulation. This result is consistent with the hypothesis that 344

hairpin III and the poly(T) sequence immediately downstream of it constitute an intrinsic 345

transcription terminator. However, the observation of high β-galactosidase activity in the 346

presence of ammonium contrasts with the assumption that the promoter of nasAB is 347

subjected to ammonium repression, which is mediated by NtrC. When introduced into a 348

ntrC mull mutant, the translational fusion showed trace amount of expression under all 349

nitrogen conditions tested, suggesting that nasAB promoter is under NtrC regulation 350

(Table 1). 351

352

The use of antitermination to regulate assimilatory nitrate reductase operon induction was 353

identified previously in K. oxytoca (pneumonia) M5al (38). The leader region of the 354

assimilatory nitrate reductase operon nasFEDCBA (referred to as nasF) in K. oxytoca 355

(pneumonia) M5al features a factor-independent terminator and one promoter-proximal 356

hairpin required for antitermination (38). The promoter-proximal hairpin is the potential 357

binding site of antiterminator protein NasR and plays an important role in nitrate/nitrite-358

induced antitermination (13, 14). Although no sequence similarity between nasF leader 359

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 18: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

18

sequence and the sequence upstream of nasA was identified, the independence of 360

potential hairpin I and II from hairpin III (Fig. 1B) resembles that of the promoter-361

proximal hairpin and terminator hairpin within the leader of nasF. To determine whether 362

disruption of predicted hairpins I and II structures altered nasAB expression, we 363

constructed two hairpin deletion mutants. These two constructs contain partial deletions 364

within hairpins I and II, respectively (Fig. 1B and 4). Secondary structure analyses 365

indicated that the partial deletion in each hairpin did not significantly alter the potential 366

free energy and formation of the adjacent hairpins. Each of the individual sequences with 367

deletions was cloned into the XbaI-BamHI region of pVnflacZa to construct translation 368

fusions, analogous to the construction of hairpin III deletion mutant. β-galactosidase 369

activity analysis showed that disruption of hairpin I abolished lacZ expression under all 370

tested nitrogen conditions, while disruption of hairpin II reduced lacZ expression when 371

induced by nitrite, but did not reduce expression to background levels (Fig. 4). These 372

results suggest that both hairpins I and II play a positive role in nasAB induction. 373

374

The loss of lacZ expression in the hairpin I deletion in the presence of nitrate or nitrite 375

could be due to disruption of nitrate/nitrite-induced antitermination, or due to disruption 376

of the integrity of the nasAB promoter, which prevents transcription initiation. A separate 377

mutant with all three hairpins deleted was constructed (pWB643 in Fig. 5). The deletion 378

of all three hairpins resulted in high β-galactosidase activity detected under all tested 379

nitrogen conditions, indicating that the integrity of nasAB promoter was retained when all 380

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 19: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

19

three hairpins were deleted. Thus the lack of lacZ expression of the hairpin I deletion is 381

most likely due to disruption of nitrate/nitrite-induced antitermination. 382

383

For several bacterial operons involved in amino acid biosynthesis, translation of small 384

ORFs located within the 5’ leader mRNA region are involved in ribosome-mediated 385

antitermination (23, 26). To test whether the putative ORF located within the nasAB 386

leader region behaved in an analogous manner, the start codon (ATG) of the eight amino 387

acid peptide ORF was converted into a TAG stop codon, eliminating potential translation 388

of this small ORF. The construct with the mutated leader region was cloned into 389

pVnflacZa and introduced into A. vinelandii. β-galactosidase assays showed that this 390

mutation had similar lacZ expression as the wild-type control (Fig. 4), indicating that this 391

putative small ORF had no role in nasAB regulation. 392

393

nasAB has a NtrC-dependent promoter 394

395

To further define the position of the nasAB promoter(s), we constructed a series of 5’ end 396

deletions upstream of the potential σ54 binding region in pWB643 and cloned them into 397

the MCS of pVnflacZa (Fig. 1A and 5). β-galactosidase assays showed that deletions 398

ending at nucleotide -168 and -131, both of which maintained the putative NtrC binding 399

sites, exhibited slightly lower but similar levels of β-galactosidase activity as the control 400

pWB643. A deletion that removed one putative NtrC binding site upstream of base -108 401

resulted in a 90% reduction in β-galactosidase activity, suggesting this sequence is 402

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 20: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

20

directly involved in promoter function. Deletion to nucleotide -43, which removes the 403

other putative NtrC binding site plus some downstream sequences, caused a 95% 404

reduction in β-galactosidase activity as compared to the full-length control. This result is 405

similar to the deletion analysis of the nifLA promoter in K. pneumonia, where deletions 406

reaching nucleotide -28 led to a 93% reduction of promoter activity (19). In addition, 407

substitution of the conserved dinucleotide pairs GG/GC with AA/AT in the potential σ54 408

binding site reduced β-galactosidase activity to background levels under all tested 409

nitrogen conditions (Table 1), suggesting that the nasAB promoter requires σ54 for 410

transcription. Taken together, we conclude that expression of nasAB is driven by an 411

NtrC-regulated promoter, PnasA. 412

413

NasT in A. vinelandii is homologous to RNA-binding antiterminators AmiR and EutV 414

415

NasT (GenBank: CAA58582) consists of two domains (24): an N-terminal domain 416

(amino acids 5–117) that is homologous to the REC (receiver)domain of the response 417

regulator of two-component regulatory systems (47), and a C-terminal domain (amino 418

acids 141–184) that is homologous to an RNA-binding domain, ANTAR (39, 61). The 419

overall structure of NasT is homologous to AmiR (51) and EutV (16) (Fig. 6), two RNA-420

binding antiterminators identified in Pseudomonas aeruginosa and Enterococcus 421

faecalis, respectively. Protein sequence homology analysis using the Water program of 422

EMBOSS (European Molecular Biology Open Software Suite) indicated that NasT and 423

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 21: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

21

AmiR share 47.4% similarity and 24.7% identity, while NasT and EutV share 58.3% 424

similarity and 28.9% identity. 425

426

NasT is required for nasAB antitermination 427

428

We constructed a bacterial heterologous expression system to test whether NasT 429

functions within the leader region of nasAB. The system is comprised of a protein-430

expressing plasmid pDK6, a newly constructed lacZ reporter plasmid, and the E. coli 431

strain MC1061, in which the lac operon has been deleted (12). pDK6 features a tac 432

promoter upstream of a multiple cloning site (MCS) and a lacIQ gene, which prevents tac 433

from being activated in the absence of the inducer isopropylthio-β-D-galactoside (IPTG) 434

(1). For our purposes, we cloned nasT into the EcoRI-BamHI region of the pDK6 MCS. 435

The lacZ reporter plasmid was derived from pBT, a component of the BacterioMatch II 436

Two-Hybrid System (Stratagene). pBT is a low copy number plasmid in E. coli and 437

features a lacUV5 promoter upstream of the λ cI gene. We replaced the cI gene in pBT 438

with the lacZ reporter cassette, which contains an XbaI-BamHI cloning site preceding 439

lacZ, resulting in the reporter plasmid pBTW. lacZ in the reporter cassette starts with 440

ATGGATCC, where GGATCC is a BamHI restriction site. We cloned the nasACBH 441

leader sequence extending into the 162nd nucleotide of nasA into the XbaI-BamHI site of 442

pBTW, resulting in the plasmid pWB555, which contains the translational fusion 443

lacUV5- nasA΄-΄lacZ. In addition, we constructed three mutants with deletions at 444

different hairpins of the leader region for expression comparison (Fig. 7). 445

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 22: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

22

446

Beta-galactosidase expression assays showed that lacUV5- nasA΄-΄lacZ in E. coli has a 447

basal level of expression, and deletion of the transcriptional terminator hairpin increased 448

the level of β-galactosidase activity 15-fold, confirming the negative role of the 449

transcriptional terminator in E. coli. In the presence of co-expressed NasT, the level of β-450

galactosidase activity increased by more than 30-fold, suggesting that NasT is 451

synthesized as a positive regulator. The positive role of NasT is compromised when the 452

integrity of hairpin I or II was disrupted: partial deletion of hairpin I reduced β-453

galactosidase activity to the basal level, while partial deletion of hairpin II reduced β-454

galactosidase activity by 87%. The results are consistent with in vivo analyses (Fig. 4). 455

We conclude that NasT acts within the leader region of nasAB, and its function requires 456

hairpins I and II. 457

458

The activity of the nasAB promoter is altered in the nifA¯ background 459

460

Expression analysis of the translational nasA΄-΄lacZ fusion showed that nasA΄-΄lacZ 461

integrated into the genomes of the nif ¯ strain UW1 (nifA–) and the wild-type strain 462

UW136 had a similar expression pattern under the conditions tested (Table 1). However, 463

β-galactosidase activity levels induced by nitrate or nitrite in UW1 were much higher 464

than in UW136. This expression difference led to the hypothesis that PnasA has lower 465

activity in a functional nif system. 466

467

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 23: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

23

The DNA sequence -316 to +16 that contained PnasA and upstream sequences but lacking 468

the three hairpin sequences was inserted into the MCS upstream of lacZ in the 469

transcriptional reporter pVnflacZb, giving rise to the transcriptional PnasA–lacZ fusion. In 470

UW136, PnasA–lacZ showed similar expression under all tested conditions, and was not 471

expressed in a ntrC mutant background (Table 2), similar to the expression profiles of the 472

hairpin III deletion or three-hairpin-deletion derivatives (Fig. 4 and 5). However, in 473

UW1, PnasA–lacZ showed a very different expression profile. Although the transcriptional 474

fusion in UW1 showed similar expression levels as in UW136 when supplied with 475

ammonium, replacement of ammonium by other nitrogen sources led to 4-fold or higher 476

levels of lacZ expression, suggesting the activity of PnasA is under partial repression in the 477

presence of ammonium. When atmospheric N2 was used as the sole nitrogen source, 478

UW1 did not grow as evidenced by the lack of an increase in cell number, but PnasA–lacZ 479

maintained strong expression. Urea is a nitrogen source that can be used by A. vinelandii 480

without repressing nitrogen fixation (65). The results from this experiment lead to the 481

conclusion that the activity of PnasA is altered in the nifA¯ background. 482

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 24: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

24

Discussion 483

484

Our study showed that the reported lack of expression of nasAB, the assimilatory nitrate 485

reductase operon in A. vinelandii in the presence of ammonium was not due to the 486

repression of the promoter PnasA, but due to transcriptional termination within the nasAB 487

leader region. The arrangement of NtrC and σ54 binding sites of PnasA is similar to that of 488

NtrC-regulated promoters in enteric bacteria (36). However, the high activity of PnasA in 489

the presence of ammonium distinguishes it from NtrC-regulated promoters in enteric 490

bacteria (45). We attribute this unusual activity of PnasA to NtrC-mediated regulation in A. 491

vinelandii. Although we cannot exclude the possibility that other regulatory factors might 492

also be involved in nasAB activation, the abolishment of nasAB promoter activation in an 493

ntrC mutant suggests the role of other potential regulatory factors is secondary to NtrC. 494

Relaxed NtrC regulation in the presence of ammonium also explains the previous 495

observation that the promoter of nifL from K. pneumonia, which is regulated solely by 496

NtrC, also was highly expressed in A. vinelandii in the presence of high concentrations of 497

ammonium (33). 498

499

Although PnasA showed similar activity in the presence of ammonium, nitrogen, urea, 500

nitrite or nitrate in the wild-type strain UW136, the promoter showed much higher 501

activity (4-17-fold) in UW1 (nifA−) when ammonium is removed or replaced with other 502

nitrogen sources (Table 2). Since UW1 and UW136 only differ in nifA function, it is 503

possible that NifA directly regulates PnasA. However, no potential NifA binding sequence 504

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 25: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

25

TGT-N10-ACA (10) was identified within PnasA or its upstream region, suggesting that the 505

regulatory role of nifA on the regulation of PnasA might be indirect. Activity analysis of 506

both nitrogenase and PnasA in future studies may bring new insight into nitrogen 507

regulation in A. vinelandii. 508

509

The increase of PnasA activity in UW1 in the absence of ammonium resembles the relief of 510

ammonium repression on NtrC-regulated promoters in enteric bacteria. We speculate that 511

the ammonium from the medium or nitrogen fixation increases cellular nitrogen status, 512

resulting in the reduced activity of NtrC. It was demonstrated previously that GlnK in A. 513

vinelandii transduces nitrogen signal from GlnD to NifL (40, 41, 54, 63). It is possible 514

that GlnK also transduces a nitrogen signal to NtrBC, a cascade similar to the general 515

nitrogen regulatory system (GlnD→GlnK→NtrBC) in enteric bacteria (45). 516

517

Antitermination mediated by NasT plays an essential role in nasAB induction. The 518

functional and structural features of NasT suggest that it belongs to the same family of 519

regulators as AmiR and EutV. The direct ANTAR-RNA interaction has been previously 520

confirmed in AmiR (50), EutV (22), and NasR (13) by means of electrophoretic mobility 521

shift assays. Unlike the REC-ANTAR structure, the N-terminus of NasR in K. oxytoca 522

(pneumonia) M5al is a nitrate- and nitrite-sensing (NIT) domain , and the binding of 523

ligand at NIT triggers the antitermination function of NasR (13, 60). 524

525

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 26: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

26

The mechanism of ANTAR-mediated antitermination is still largely unknown. Alignment 526

of multiple RNAs targeted by EutV show a 13 nt shared conserved sequence 527

(AGCAANGRRGCUY) (22); this sequence overlaps with the left stem of the terminator 528

hairpin and the right stem of a putative low-stability antitermination hairpin (2). This 529

conserved feature led to the hypothesis that binding of EutV to the conserved sequence 530

increases the stability of the antiterminator structure, preventing formation of a terminator 531

hairpin (2). 532

533

The putative EutV binding sequence is not identified in the leader region of nasACBH. 534

The results of our deletion mutational analyses indicate that hairpins I and II, which are 535

separated from the terminator hairpin, are potential binding sites of NasT. Similarly, the 536

potential binding site of NasR is a promoter-proximal hairpin separated from the 537

terminator (14). Thus, the model of ANTAR-related antitermination reconciling all these 538

findings remains to be defined in future studies. 539

540

In the absence of nitrate and nitrite induction, the translational nasA΄-΄lacZ fusion showed 541

low levels of expression (Fig. 4). Although low, these expression levels are still higher 542

compared to the levels produced by the hairpin I deletion mutant, suggesting that read-543

through of the terminator may occur at a low level. We speculate that low levels of 544

nasAB expression in the absence of nitrate and/or nitrite might be related to the regulatory 545

role of NasB. Previous analysis suggested that NasB has a negative role on nasAB 546

expression, and deletion of nasB resulted in high expression of nasA even without 547

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 27: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

27

nitrate/nitrite induction (24, 52). How NasB is involved in nasAB operon regulation 548

remains to be determined. 549

550

In summary, identification of NasT-mediated antitermination in nasAB regulation 551

expands our understanding of nitrate/nitrite reductases regulation in bacteria and 552

ANTAR-containing antiterminators. The study also sheds additional insights into 553

nitrogen regulation in A. vinelandii, and indicates that it is much more complicated than 554

previously thought. nasAB represents the first NtrC-regulated operon in A. vinelandii that 555

has been studied in detail, and it will be of interest to learn how other NtrC-regulated 556

operons in this bacterium regulate their expression in a substrate specific manner. 557

558

Acknowledgements 559

We thank Bentley A. Fane and Yeou-Cherng Bor for discussions and critical review of 560

the manuscript. 561

Dedication (in memoriam): This manuscript is dedicated to the memory of Christina 562

Kennedy (1945-2009). 563

564

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 28: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

28

References 565

566 1. Amann, E., J. Brosius, and M. Ptashne. 1983. Vectors bearing a hybrid trp-lac 567

promoter useful for regulated expression of cloned genes in Escherichia coli. 568 Gene 25:167-78. 569

570 2. Baker, K. A., and M. Perego. 2011. Transcription anti-termination by a 571

phosphorylated response regulator and cobalamin-dependent termination at a B12 572 riboswitch contribute to ethanolamine utilization in Enterococcus faecalis. J. 573 Bacteriol.:JB.00217-11. 574

575 3. Bali, A., G. Blanco, S. Hill, and C. Kennedy. 1992. Excretion of ammonium by 576

a nifL mutant of Azotobacter vinelandii fixing nitrogen. Appl. Environ. 577 Microbiol. 58:1711-8. 578

579 4. Barrios, H., B. Valderrama, and E. Morett. 1999. Compilation and analysis of 580

sigma(54)-dependent promoter sequences. Nucleic Acids Res 27:4305-13. 581 582 5. Beynon, J., M. Cannon, V. Buchanan-Wollaston, and F. Cannon. 1983. The 583

nif promoters of Klebsiella pneumonia have a characteristic primary structure. 584 Cell 34:665-671. 585

586 6. Bishop, P. E., D. M. Jarlenski, and D. R. Hetherington. 1980. Evidence for an 587

alternative nitrogen fixation system in Azotobacter vinelandii. Proc. Natl. Acad. 588 Sci. USA 77:7342-6. 589

590 7. Blanco, G., M. Drummond, P. Woodley, and C. Kennedy. 1993. Sequence and 591

molecular analysis of the nifL gene of Azotobacter vinelandii. Mol. Microbiol. 592 9:869-79. 593

594 8. Buchanan-Wollaston, V., M. C. Cannon, J. L. Beynon, and F. C. Cannon. 595

1981. Role of the nifA gene product in the regulation of nif expression in 596 Klebsiella pneumonia. Nature 294:776-8. 597

598 9. Buck, M., S. Miller, M. Drummond, and R. Dixon. 1986. Upstream activator 599

sequences are present in the promoters of nitrogen fixation genes. Nature 600 320:374-378. 601

602 10. Casadaban, M. J., and S. N. Cohen. 1980. Analysis of gene control signals by 603

DNA fusion and cloning in Escherichia coli. J. Mol. Biol. 138:179-207. 604 605

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 29: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

29

11. Chai, W., and V. Stewart. 1998. NasR, a novel RNA-binding protein, mediates 606 nitrate-responsive transcription antitermination of the Klebsiella oxytoca M5al 607 nasF operon leader in vitro. J. Mol. Biol. 283:339-51. 608

609 12. Chai, W., and V. Stewart. 1999. RNA sequence requirements for NasR-610

mediated, nitrate-responsive transcription antitermination of the Klebsiella 611 oxytoca M5al nasF operon leader. J. Mol. Biol. 292:203-216. 612

613 13. Contreras, A., M. Drummond, A. Bali, G. Blanco, E. Garcia, G. Bush, C. 614

Kennedy, and M. Merrick. 1991. The product of the nitrogen fixation regulatory 615 gene nfrX of Azotobacter vinelandii is functionally and structurally homologous 616 to the uridylyltransferase encoded by glnD in enteric bacteria. J. Bacteriol. 617 173:7741-9. 618

619 14. Del Papa, M. F., and M. Perego. 2008. Ethanolamine activates a sensor histidine 620

kinase regulating its utilization in Enterococcus faecalis. J. Bacteriol. 190:7147-621 56. 622

623 15. Dixon, R., and D. Kahn. 2004. Genetic regulation of biological nitrogen fixation. 624

Nat. Rev. Micro. 2:621-631. 625 626 16. Drummond, M., J. Clements, M. Merrick, and R. Dixon. 1983. Positive 627

control and autogenous regulation of the nifLA promoter in Klebsiella pneumonia. 628 Nature 301:302-307. 629

630 17. Espin, G., A. Alvarez-Morales, F. Cannon, R. Dixon, and M. Merrick. 1982. 631

Cloning of the glnA, ntrB and ntrC genes of Klebsiella pneumonia and studies of 632 their role in regulation of the nitrogen fixation (nif) gene cluster. Mol. Gen. Genet. 633 186:518-24. 634

635 18. Ferro-Luzzi Ames, G., and K. Nikaido. 1985. Nitrogen regulation in Salmonella 636

typhimurium. Identification of an ntrC protein-binding site and definition of a 637 consensus binding sequence. EMBO J 4:539-47. 638

639 19. Fox, K. A., A. Ramesh, J. E. Stearns, A. Bourgogne, A. Reyes-Jara, W. C. 640

Winkler, and D. A. Garsin. 2009. Multiple posttranscriptional regulatory 641 mechanisms partner to control ethanolamine utilization in Enterococcus faecalis. 642 Proc. Natl. Acad. Sci. USA 106:4435-40. 643

644 20. Gollnick, P., and P. Babitzke. 2002. Transcription attenuation. Biochimica et 645

Biophysica Acta (BBA) - Gene Structure and Expression 1577:240-250. 646 647

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 30: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

30

21. Gutierrez, J. C., F. Ramos, L. Ortner, and M. Tortolero. 1995. nasST, two 648 genes involved in the induction of the assimilatory nitrite-nitrate reductase operon 649 (nasAB) of Azotobacter vinelandii. Mol. Microbiol. 18:579-91. 650

651 22. Gutierrez, J. C., E. Santero, and M. Tortolero. 1997. Ammonium repression of 652

the nitrite-nitrate (nasAB) assimilatory operon of Azotobacter vinelandii is 653 enhanced in mutants expressing the nifO gene at high levels. Mol. Gen. Genet. 654 255:172-179. 655

656 23. Henkin, T. M., and C. Yanofsky. 2002. Regulation by transcription attenuation 657

in bacteria: how RNA provides instructions for transcription 658 termination/antitermination decisions. BioEssays 24:700-707. 659

660 24. Hill, S., S. Austin, T. Eydmann, T. Jones, and R. Dixon. 1996. Azotobacter 661

vinelandii NIFL is a flavoprotein that modulates transcriptional activation of 662 nitrogen-fixation genes via a redox-sensitive switch. Proc. Natl. Acad. Sci. USA 663 93:2143-8. 664

665 25. Ho, S. N., H. D. Hunt, R. M. Horton, J. K. Pullen, and L. R. Pease. 1989. Site-666

directed mutagenesis by overlap extension using the polymerase chain reaction. 667 Gene 77:51-59. 668

669 26. Jiang, P., and A. J. Ninfa. 1999. Regulation of autophosphorylation of 670

Escherichia coli nitrogen regulator II by the PII signal transduction protein. J. 671 Bacteriol. 181:1906-1911. 672

673 27. Jiang, P., J. A. Peliska, and A. J. Ninfa. 1998. Enzymological characterization 674

of the signal-transducing Uridylyltransferase/Uridylyl-Removing Enzyme (EC 675 2.7.7.59) of Escherichia coli and its interaction with the PII Protein. Biochemistry 676 37:12782-12794. 677

678 28. Kamberov, E. S., M. R. Atkinson, and A. J. Ninfa. 1995. The Escherichia coli 679

PII Signal Transduction Protein Is Activated upon Binding 2-Ketoglutarate and 680 ATP. J. Bio. Chem. 270:17797-17807. 681

682 29. Keener, J., and S. Kustu. 1988. Protein kinase and phosphoprotein phosphatase 683

activities of nitrogen regulatory proteins NTRB and NTRC of enteric bacteria: 684 roles of the conserved amino-terminal domain of NTRC. Proc. Natl. Acad. Sci. 685 USA 85:4976-80. 686

687 30. Kennedy, C., and M. H. Drummond. 1985. The use of cloned nif regulatory 688

elements from Klebsiella pneumonia to examine nif regulation in Azotobacter 689 vinelandii. J. Gen. Microbiol. 131:1787-1795. 690

691

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 31: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

31

31. Kennedy, C., and R. L. Robson. 1983. Activation of nif gene expression in 692 Azotobacter by the nifA gene product of Klebsiella pneumonia. Nature 301:626-8. 693

694 32. Kustu, S., A. K. North, and D. S. Weiss. 1991. Prokaryotic transcriptional 695

enhancers and enhancer-binding proteins. Trends Biochem. Sci. 16:397-402. 696 697 33. Leonardo, J. M., and R. B. Goldberg. 1980. Regulation of nitrogen metabolism 698

in glutamine auxotrophs of Klebsiella pneumonia. J Bacteriol 142:99-110. 699 700 34. Lin, J. T., and V. Stewart. 1996. Nitrate and Nitrite-mediated Transcription 701

Antitermination Control of nasF(Nitrate Assimilation) Operon Expression in 702 Klebsiella pneumonia M5al. J. of Mol. Biol. 256:423-435. 703

704 35. Lin, J. T., and V. Stewart. 1998. Nitrate assimilation by bacteria. Adv. Microb. 705

Physiol. 39:1-30, 379. 706 707 36. Little, R., V. Colombo, A. Leech, and R. Dixon. 2002. Direct interaction of the 708

NifL regulatory protein with the GlnK signal transducer enables the Azotobacter 709 vinelandii NifL-NifA regulatory system to respond to conditions replete for 710 nitrogen. J. Biol. Chem. 277:15472-81. 711

712 37. Little, R., F. Reyes-Ramirez, Y. Zhang, W. C. van Heeswijk, and R. Dixon. 713

2000. Signal transduction to the Azotobacter vinelandii NIFL-NIFA regulatory 714 system is influenced directly by interaction with 2-oxoglutarate and the PII 715 regulatory protein. EMBO J. 19:6041-50. 716

717 38. Luque, F., E. Santero, J. R. Medina, and M. Tortolero. 1987. An effective 718

mutagenic method in Azotobacter vinelandii. Microbiologia 3:45-9. 719 720 39. Meletzus, D., P. Rudnick, N. Doetsch, A. Green, and C. Kennedy. 1998. 721

Characterization of the glnK-amtB operon of Azotobacter vinelandii. J. Bacteriol. 722 180:3260-4. 723

724 40. Merrick, M. J. 1983. Nitrogen control of the nif regulon in Klebsiella 725

pneumonia: involvement of the ntrA gene and analogies between ntrC and nifA. 726 EMBO J 2:39-44. 727

728 41. Merrick, M. J., and R. A. Edwards. 1995. Nitrogen control in bacteria. 729

Microbiol. Rev 59:604-22. 730 731 42. Miller, W. G., J. H. J. Leveau, and S. E. Lindow. 2000. Improved gfp and inaZ 732

Broad-Host-Range Promoter-Probe Vectors. Molecular Plant-Microbe 733 Interactions 13:1243-1250. 734

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 32: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

32

43. Morth, J. P., V. Feng, L. J. Perry, D. I. Svergun, and P. A. Tucker. 2004. The 735 crystal and solution structure of a putative transcriptional antiterminator from 736 mycobacterium tuberculosis. Structure 12:1595-1605. 737

738 44. Newton, J. W., P. W. Wilson, and R. H. Burris. 1953. Direct demonstration of 739

ammonia as an intermediate in nitrogen fixation by Azotobacter. J. Biol. Chem. 740 204:445-51. 741

742 45. Ninfa, A. J., and B. Magasanik. 1986. Covalent modification of the glnG 743

product, NRI, by the glnL product, NRII, regulates the transcription of the 744 glnALG operon in Escherichia coli. Proc. Natl. Acad. Sci. USA 83:5909-13. 745

746 46. Norman, R. A., C. L. Poh, L. H. Pearl, B. P. O'Hara, and R. E. Drew. 2000. 747

Steric hindrance regulation of the Pseudomonas aeruginosa amidase operon. J. 748 Biol. Chem. 275:30660-7. 749

750 47. O'Hara, B. P., R. A. Norman, P. T. Wan, S. M. Roe, T. E. Barrett, R. E. 751

Drew, and L. H. Pearl. 1999. Crystal structure and induction mechanism of 752 AmiC-AmiR: a ligand-regulated transcription antitermination complex. EMBO J. 753 18:5175-86. 754

755 48. Ramos, F., G. Blanco, J. C. Gutiérrez, F. Luque, and M. Tortolero. 1993. 756

Identification of an operon involved in the assimilatory nitrate-reducing system of 757 Azotobacter vinelandii. Mol. Microbiol. 8:1145-1153. 758

759 49. Robson, R. L., J. A. Chesshyre, C. Wheeler, R. Jones, P. R. Woodley, and J. 760

R. Postgate. 1984. Genome size and complexity in Azotobacter chroococcum. J. 761 Gen. Microbiol 130:1603-12. 762

763 50. Rudnick, P., C. Kunz, M. K. Gunatilaka, E. R. Hines, and C. Kennedy. 2002. 764

Role of GlnK in NifL-mediated regulation of NifA activity in Azotobacter 765 vinelandii. J. Bacteriol. 184:812-20. 766

767 51. Sambrook, J., and D. W. Russell. 2001. Molecular Cloning: A Laboratory 768

Manual Third ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. 769 770 52. Santero, E., F. Luque, J. R. Medina, and M. Tortolero. 1986. Isolation of ntrA-771

like mutants of Azotobacter vinelandii. J. Bacteriol. 166:541-4. 772 773 53. Santero, E., A. Toukdarian, R. Humphrey, and C. Kennedy. 1988. 774

Identification and characterization of two nitrogen fixation regulatory regions, 775 nifA and nfrX, in Azotobacter vinelandii and Azotobacter chroococcum. Mol. 776 Microbiol. 2:303-314. 777

778

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 33: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

33

54. Setubal, J. C., P. dos Santos, B. S. Goldman, H. Ertesvag, G. Espin, L. M. 779 Rubio, S. Valla, N. F. Almeida, D. Balasubramanian, L. Cromes, L. Curatti, 780 Z. Du, E. Godsy, B. Goodner, K. Hellner-Burris, J. A. Hernandez, K. 781 Houmiel, J. Imperial, C. Kennedy, T. J. Larson, P. Latreille, L. S. Ligon, J. 782 Lu, M. Maerk, N. M. Miller, S. Norton, I. P. O'Carroll, I. Paulsen, E. C. 783 Raulfs, R. Roemer, J. Rosser, D. Segura, S. Slater, S. L. Stricklin, D. J. 784 Studholme, J. Sun, C. J. Viana, E. Wallin, B. Wang, C. Wheeler, H. Zhu, D. 785 R. Dean, R. Dixon, and D. Wood. 2009. Genome sequence of Azotobacter 786 vinelandii, an obligate aerobe specialized to support diverse anaerobic metabolic 787 processes. J. Bacteriol. 191:4534-45. 788

789 55. Shu, C. J., L. E. Ulrich, and I. B. Zhulin. 2003. The NIT domain: a predicted 790

nitrate-responsive module in bacterial sensory receptors. Trends Biochem. Sci. 791 28:121-4. 792

793 56. Shu, C. J., and I. B. Zhulin. 2002. ANTAR: an RNA-binding domain in 794

transcription antitermination regulatory proteins. Trends Biochem. Sci. 27:3-5. 795 796 57. Soderback, E., F. Reyes-Ramirez, T. Eydmann, S. Austin, S. Hill, and R. 797

Dixon. 1998. The redox- and fixed nitrogen-responsive regulatory protein NIFL 798 from Azotobacter vinelandii comprises discrete flavin and nucleotide-binding 799 domains. Mol. Microbiol. 28:179-92. 800

801 58. Toukdarian, A., and C. Kennedy. 1986. Regulation of nitrogen metabolism in 802

Azotobacter vinelandii: isolation of ntr and glnA genes and construction of ntr 803 mutants. EMBO J. 5:399-407. 804

805 59. Walmsley, J., and C. Kennedy. 1991. Temperature-dependent regulation by 806

molybdenum and vanadium of expression of the structural genes encoding three 807 nitrogenases in Azotobacter vinelandii. Appl. Environ. Microbiol. 57:622-4. 808

809 60. Zuker, M. 2003. Mfold web server for nucleic acid folding and hybridization 810

prediction. Nucleic Acids Research 31:3406-3415. 811 812

813

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 34: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

37

Figures Legends 814

815

Figure 1 (A) Nucleotide sequence from 477 bases upstream of nasA to the beginning of 816

the NasA ORF. The predicted transcription start site is marked with +1; the putative σ54-817

binding sequence is shown in a light grey box with the conserved GG/GC bases in bold; 818

the putative NtrC protein binding sites are underlined with the consensus NtrC binding 819

sequence shown underneath for comparison; inverted repeats are above the specific 820

sequences with facing arrows; the start and stop codons of the putative octopeptide ORF 821

in the leader region are bold-faced. The predicted poly(T) sequence adjacent to hairpin III 822

is italicized. The 8 codons of the putative octopeptide sequence are indicated and the 823

putative NasA Shine-Dalgarno (SD) sequence has a dashed underline. (B) Predicted 824

secondary palindromic structures within the nasAB leader sequence. RNA sequence 825

numbers are relative to the predicted transcription start site; the free energies (kcal/mol) 826

of the predicted hairpins are indicated at the top of hairpins; the letters with bold faces 827

mark the sequences deleted in the mutant constructs. 828

829

Figure 2. Map of the A. vinelandii lacZ reporter vectors. Each vector can be integrated 830

into the vnf region of A. vinelandii genome to provide stable single copy expression data. 831

Plasmid pVnflacZa is a translational fusion probe containing the XbaI-BcuI-BamHI 832

cloning site, and pVnflacZb is a transcriptional fusion probe containing the XbaI-BcuI-833

BamHI-XbaI site followed by a SD sequence eight nucleotides upstream of the lacZ. 834

MCS, multiple cloning sites; pMB1, replicon region; T1(4), four tandem copies of the T1 835

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 35: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

38

terminator from the E. coli rrnB1 operon (46); vnf, vanadium(V)-containing nitrogenase 836

gene (7). 837

838

Figure 3. Time course of the expression of the translational nasA΄–΄lacZ fusion integrated 839

into the vnf region of the A. vinelandii genome. nasA΄ in the fusion starts with the 840

nucleotide 190 upstream of the potential NtrC binding site and ends with the 162nd 841

nucleotide of the NasA ORF. Supplied nitrogen sources were (*), nitrite;( •), nitrate; (▪), 842

dinitrogen; (◦), ammonium. Bacteria were grown on BS medium supplied with the 843

indicated nitrogen sources. At the times indicated, samples were taken and assayed for β-844

galactosidase expression. 845

846

Figure 4. Effect of deletions and modifications within the leader sequence on nasA΄–847

΄lacZ expression (not drawn to scale). Top line represents the DNA sequence from the 848

nucleotide -316 to +297 (the 162nd nucleotide of NasA); bent arrow and +1 represent the 849

initiation site and direction of transcription; inverted arrows represent the palindromic 850

sequences comprising hairpins I, II, and III, respectively; empty rectangles represent 851

deleted sequences; hollow diamond represents the ATG → TAG nonsense mutation. 852

Cultures were grown in BS medium supplemented with the indicated nitrogen sources, 853

and β-galactosidase activity was measured 7 hrs after inoculation. All data represent 854

mean of three replicates ± standard deviations from a representative experiment. 855

Figure 5. Effect of deletions upstream of the putative σ54 binding site on nasA΄–΄lacZ 856

expression (not drawn to scale). pWB643 has all three hairpins (nt +14 to + 100) within 857

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 36: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

39

the leader region deleted. All other deletions were constructed based on the sequence 858

around the nasA promoter region in the plasmid pWB643. Top line represents the DNA 859

sequence from the nucleotide -316 to +297 (the 162nd nucleotide of NasA); bent arrow 860

and +1 represent the initiation site and direction of transcription; empty rectangles 861

represent three hairpin deletion within the leader region upstream of nasA; vertical arrows 862

indicate the 5’ ends (-316, -168, -131, -108, and -43) of tested sequences mutants. β-863

galactosidase activity assay was performed as described in the legend of Figure 4. 864

865

Figure 6. Sequence alignment of NasT, EutV, AmiR, and REC domain. REC domain 866

sequence (CDD:cd00156) is from the Conserved Domain Database (CDD) at NCBI (the 867

National Center for Biotechnology Information). Protein sequences were obtained from 868

NCBI protein databases: NasT (GenBank: CAA58582), EutV (GenBank: ZP_03948869) 869

from Enterococcus faecalis TX0104, and AmiR (GenBank: CAA32023) from 870

Pseudomonas aeruginosa. (-) and (~) in the figure indicate the gap of alignment; black 871

background indicates conservation of nucleotides in all compared sequences, and gray 872

background indicates conservation in some compared sequences. 873

874

Figure 7. Analysis of NasT functions within the leader region of nasACBH. Top line 875

represents the DNA sequence from the nucleotide +1 to +297 (the 162nd nucleotide of 876

nasA); bent arrow and +1 represent the initiation site and direction of transcription; 877

inverted arrows represent the palindromic sequences comprising hairpins I, II, and III, 878

respectively. Bottom left: nasA΄ fragments in the lacUV5-nasA΄-΄lacZ cassettes. The 879

hollow arrows represent the lacUV5 promoter; the empty rectangles represent the 880

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 37: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

40

deletion within the leader regions. Bottom right: results of expression analyses of nasA΄-881

΄lacZ with/without co-expressed NasS or NasT. β-galactosidase activity was measured 882

after three hours growth in LB medium. All data are presented as the mean values of 883

triplicate samples ± standard deviations from a representative experiment. 884

885

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 38: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

41

TABLE 1. Expression of the translational nasA΄–΄lacZ fusion in A. vinelandiia. 886

887

nasA΄ A. vinelandii

genotype

β-galactosidase activity (Miller Units)b

Nitrogen source

NH4+ N2 NO2

- NO3-

Wild typec Wild type 69 ± 1 96 ± 4 682 ± 9 478 ± 6

Wild type nif - 27 ± 1 72 ± 2 980 ± 12 1262 ± 18

nasA΄ with

GG/GC→AA/AT

substitution in the

σ54 binding site

Wild type <5 5 ± 2 7 ± 1 5 ± 1

Wild type ∆ntrC::Gmr <5 <5 23 ± 1 <5

nasA΄ with hairpin

III deletion ∆ntrC::Gmr <5 <5 <5 <5

888 a nasA΄ refers to the sequence -316 to +297, which includes the nas AB promoter PnasA 889

and the beginning of the NasA ORF. 890

891 b Cultures were grown in BS medium supplemented with the indicated nitrogen sources. 892

β-galactosidase activity was measured 7 hrs after incubation and reported as Miller Units. 893

All data are mean values of triplicate samples ± standard deviations from a representative 894

experiment. 895

896 cRefers to the sequence of the nasA region used in the fusions. 897 898

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 39: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

42

TABLE 2. Expression of the transcriptional PnasA -lacZ fusion in A. vinelandiia. 899

900

Strain genotype

β-galactosidase activity (Miller Units)b

Nitrogen source

NH4+ N2 Urea NO2

- NO3-

Wild type 594 ± 15 307 ± 14 324 ± 10 439 ± 10 394 ± 25

∆ntrC::Gmr <5 8 ± 1 <5 <5 <5

nif - 559 ± 17 5,328 ± 30 3,470 ± 34 2,087

±142 2,677 ± 32

901 a PnasA refers to the sequence -316 to +12 around the nasAB promoter. 902

903

b Cultures were grown in BS medium supplemented with the indicated nitrogen sources. 904

β-galactosidase activity was measured 7 hrs after incubation and reported as Miller 905

Units. All data are mean values of triplicate samples ± standard deviations from a 906

representative experiment. 907

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 40: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 41: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 42: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 43: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 44: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 45: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 46: Downloaded from //aem.asm.org/content/aem/early/2012/07/01/AEM...6 Gunatilaka 1, and Christina Kennedy 1 7 8 9 1The School of Plant Sciences, University of Arizona, Tucson, Arizona

on March 12, 2020 by guest

http://aem.asm

.org/D

ownloaded from