cloning, expression and functional ch aracterization of...

40
1 Cloning, Expression and Functional Characterization of Secondary Amino 1 Acid Transporters of Lactococcus lactis 2 3 Hein Trip 1 , Niels L. Mulder 2 and Juke S. Lolkema 4 Molecular Microbiology, Groningen Biomolecular Sciences and Biotechnology Institute, 5 University of Groningen, Groningen, The Netherlands. 6 7 * correspondence 8 J. S. Lolkema, Centre for Life Sciences, University of Groningen, Nijenborgh 7, 9747AG 9 Groningen, The Netherlands 10 tel. +31503632155; fax. +31503632154; email. [email protected] 11 12 1 Present address: Department of Earth, Environmental, and Life Sciences, TNO, Lange 13 Kleiweg 137, Rijswijk, P.O Box 45, The Netherlands 14 2 Present address: Department of Pediatrics, Center for Liver, Digestive and Metabolic 15 Diseases, University Medical Center Groningen, University of Groningen, Hanzeplein 1, 16 9713 GZ Groningen, The Netherlands. 17 18 running title: Amino acid transporters of L. lactis 19 20 21 22 23 Copyright © 2012, American Society for Microbiology. All Rights Reserved. J. Bacteriol. doi:10.1128/JB.01948-12 JB Accepts, published online ahead of print on 9 November 2012 on May 3, 2018 by guest http://jb.asm.org/ Downloaded from

Upload: vothuy

Post on 14-Feb-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

1

Cloning, Expression and Functional Characterization of Secondary Amino 1

Acid Transporters of Lactococcus lactis 2

3

Hein Trip1 , Niels L. Mulder2 and Juke S. Lolkema 4

Molecular Microbiology, Groningen Biomolecular Sciences and Biotechnology Institute, 5

University of Groningen, Groningen, The Netherlands. 6

7

* correspondence 8

J. S. Lolkema, Centre for Life Sciences, University of Groningen, Nijenborgh 7, 9747AG 9

Groningen, The Netherlands 10

tel. +31503632155; fax. +31503632154; email. [email protected] 11

12

1Present address: Department of Earth, Environmental, and Life Sciences, TNO, Lange 13

Kleiweg 137, Rijswijk, P.O Box 45, The Netherlands 14

2Present address: Department of Pediatrics, Center for Liver, Digestive and Metabolic 15

Diseases, University Medical Center Groningen, University of Groningen, Hanzeplein 1, 16

9713 GZ Groningen, The Netherlands. 17

18

running title: Amino acid transporters of L. lactis 19

20

21

22

23

Copyright © 2012, American Society for Microbiology. All Rights Reserved.J. Bacteriol. doi:10.1128/JB.01948-12 JB Accepts, published online ahead of print on 9 November 2012

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

2

Abstract 24

25

Fourteen genes encoding putative secondary amino acid transporters were identified in the 26

genomes of Lactococcus lactis subsp. cremoris strains MG1363 and SK11 and subsp. lactis 27

strains IL1403 and KF147, twelve of which were common to all four strains. Amino acid 28

uptake in L. lactis cells overexpressing the genes revealed transporters specific for histidine, 29

lysine, arginine, agmatine, putrescine, aromatic amino acids, acidic amino acids, serine and 30

branched chain amino acids. Substrate specificities were demonstrated by inhibition profiles 31

in the presence of excess of the other amino acids. 32

Four knock-out mutants were constructed, lacking the lysine transporter LysP, the histidine 33

transporter HisP (formerly LysQ), the acidic amino acid transporter AcaP (YlcA) or the 34

aromatic amino acid transporter FywP (YsjA). The LysP, HisP and FywP deletion mutants 35

showed drastically decreased rates of uptake of the corresponding substrates at low 36

concentrations. The same was observed for the AcaP mutant with aspartate, but not with 37

glutamate. In rich M17 medium, deletion of none of the transporters affected growth. In 38

contrast, deletion of the HisP, AcaP and FywP transporters did effect growth in a defined 39

media with free amino acids as the sole amino acid source. HisP was essential at low 40

histidine concentration, AcaP in the absence of glutamine. FywP appeared to play a role in 41

retaining intracellularly synthesized aromatic amino acids when these were not added to the 42

medium. Finally, HisP, AcaP and FywP did not play a role in the excretion of accumulated 43

histidine, glutamate or phenylalanine, respectively, indicating the involvement of other 44

transporters. 45

46

47

48

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

3

Introduction 49

50

The lactic acid bacterium Lactococcus lactis is widely used for the production of cheese and 51

buttermilk and is therefore of great commercial importance. Lactic acid bacteria have adapted 52

to nutrient rich environments and lack various biosynthetic pathways. Most dairy and 53

laboratory strains of L. lactis are auxotrophic for a number of amino acids, for example 54

branched chain amino acids and histidine. It is believed that the capacity to synthesize these 55

amino acids was lost during adaptation to milk (1,2), which provides all essential amino acids 56

in the form of its major protein constituent casein (3). Casein is degraded by an efficient 57

proteolytic system that involves protein hydrolysis into a range of peptides of different 58

lengths, transport of some of these peptides via the oligopeptide uptake system Opp and 59

further degradation to free amino acids by intracellular peptidases (4,5). Nevertheless, dairy 60

strains of L. lactis are still able to grow in media containing free amino acids as the sole 61

source of amino acid (6), implying that transport systems for essential, free amino acids have 62

been conserved. A range of transport systems for essential as well as non-essential amino 63

acids has been described that use the proton motive force (by proton symport), the 64

electrochemical gradients of substrate and counter substrate (exchange) or ATP hydrolysis as 65

driving force (7). The exchange systems are in fact not used to supply amino acids for protein 66

biosynthesis, but are part of simple metabolic pathways that generate energy in the form of 67

ATP or proton motive force (pmf) by deimination or decarboxylation of the precursor amino 68

acid combined with precursor/product exchange across the membrane (8,9,10). 69

Most of the knowledge about amino acid transport in L. lactis and other lactic acid bacteria 70

was obtained from studies performed in the years around 1990. Transport studies using whole 71

cells or membrane vesicles revealed secondary uptake systems specific for branched amino 72

acids and methionine (11,12), alanine and glycine (13), lysine (14), serine and threonine, 73

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

4

histidine, cysteine (15) and tyrosine and phenylalanine (16). ATP hydrolysis driven transport 74

systems were found for glutamate and glutamine (17) and possibly proline (18). Only few 75

amino acid transporters have since been cloned and characterized. From L. lactis, only bcaP 76

and glnPQ were cloned and experimentally shown to encode a branched chain amino acid 77

transporter (19) and an ATP driven glutamate/glutamine transporter (20), respectively. Other 78

genes have been annotated as amino acid transporters, but mainly on basis of their homology 79

with other known transporters or their genetic context. 80

Most described bacterial and fungal amino acid transporters that use proton symport or 81

precursor/product exchange as energy coupling mechanism are members of the Amino acid-82

Polyamine-Organocation (APC) superfamily (TC 2.A.3 in the classification of Saier et al. 83

(21)). Here we set out to clone and functionally express the APC family members of L. lactis 84

and to determine their substrate specificity. Four mutants, lacking one of the identified 85

transporter genes, were constructed and used to demonstrate the role of the transporters in the 86

physiology of the cell. 87

88

89

90

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

5

Material and Methods 91

92

Media and growth conditions. 93

L. lactis NZ9000 (22) and JP9000 (23), strains derived from strain MG1363 carrying the 94

nisRK genes in the pepN and pseudo_10 loci, respectively, were used for nisin inducible 95

expression of amino acid transporter genes and as parental strains of transporter deletion 96

mutants. L. lactis cells were grown at 30 °C in M17 media supplemented with 25 mM 97

glucose (hereafter referred to as GM17) and containing 5 μg/ml chloramphenicol when 98

appropriate, or in SA media, a chemically defined medium (6) containing free amino acids 99

unless stated otherwise. E. coli, used as a cloning host for deletion constructs, was grown in 100

LB media containing 150 μg/ml erythromycin. 101

102

Plasmid and strain construction. 103

Genes encoding putative amino acid transporters were amplified from chromosomal DNA of 104

L. lactis MG1363 and IL1403 (in case of aguD) using primers listed in Supplementary Table 105

S1, and cloned behind the nisin inducible promoter PnisA in pNZ8048 (24), using NcoI or PagI 106

restriction sites at the translational start of the genes and a XbaI restriction site approximately 107

200 bp downstream the genes. The resulting plasmids were transformed to L. lactis NZ9000 108

for nisin inducible expression. 109

Markerless deletion of lysP, lysQ, llmg_1452 (ylcA in strain IL1403; see supplementary Table 110

S2) and llmg_2011 (ysjA in IL1403) was performed with a two-step integration and excision 111

system using plasmid pCS1966 (25), which is based on positive selection of an erythromycin 112

resistance marker and negative selection against a 5’-fluoroorotic acid transporter encoded by 113

oroP. For deletion of lysQ, up- and downstream flanking regions of approximately 700 bp, 114

overlapping with the first 50 bp and last 100 bp of the gene, were amplified from 115

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

6

chromosomal DNA using primers lysQ-fl1-fw, lysQ-fl2-rv, lysQ-fl2-fw and lysQ-fl2-rv 116

(Supplementary Table S1) and cloned adjacently into pCS1966 using E. coli DH5α as a 117

cloning host. The resulting plasmid was transformed into L. lactis NZ9000. Erythromycin 118

resistant colonies were tested for integration at the lysQ site using PCR. Correct clones were 119

grown overnight in non-selective medium to allow for recombination and excision of the 120

plasmid backbone including either the wild type or the disrupted lysQ gene, resulting in lysQ 121

deletion or return to the wild type situation, respectively. Cells were spread on SA medium 122

agar plates containing 10 μg/ml 5’-fluoroorotic acid for selection against the oroP gene and 123

colonies were tested by colony PCR for deletion of lysQ. For deletion of lysP, llmg_2011 and 124

llmg_1452, approximately 2.6 kb fragments of the respective genes including approximately 125

700 bp up- and downstream regions, were amplified using primers lysP-/2011-/1452-fl1-fw 126

and lysP-/2011-/1452-fl2-rv (Table S1) and cloned into pCS1966. From the resulting 127

constructs, approximately 1.2 kb internal fragments of the transporter genes were deleted by 128

amplifying the whole constructs except the internal fragments using primers lysP-/2011-129

/1452-fl1-rv and lysP-/2011-/1452-fl2-fw (Table S1). The PCR products were self-ligated 130

after digestion with restriction enzymes, resulting in the deletion constructs, which were 131

transformed into E. coli DH5α for propagation. The plasmids were transformed to L. lactis 132

JP9000, a MG1363 derivative containing the nisRK genes in the pseudo_10 locus (23). The 133

procedure was continued as described for deletion of lysQ. 134

135

Amino acid transport assays. 136

L. lactis strains harbouring expression constructs or the empty vector pNZ8048 were grown 137

in GM17 to mid-exponential phase (OD600 = 0.6) and induced with 5 ng/ml nisin. After 1 138

hour cells were harvested, washed and resuspended to an OD600 of 2 in ice-cold 100 mM 139

potassium phosphate buffer, pH 6.0, containing 0.2 % glucose and kept on ice until use. 140

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

7

Samples of 100 μl of the cell suspension were incubated for 5 min at 30 °C under continuous 141

stirring followed by the addition of 14C-labelled amino acids (Ala, Arg, Asn, Asp, Glu, His, 142

Ile, Leu, Lys, Phe, Pro, Ser, Tyr and Val) or 14C-labelled putrescine (Perkin Elmer) to final 143

concentrations ranging from 1 to 5 μM, depending on the amino acid. In case of putrescine 144

uptake by cells expressing aguD, 50 μM (final concentration) of unlabelled agmatine was 145

added 1 minute after allowing cells to accumulate 14C-putrescine (present at 4.7 μM). Uptake 146

was stopped by addition of 2 ml ice cold 0.1 M LiCl and filtration through a 0.45-μm-pore-147

size nitrocellulose filter (BA85; Schleicher & Schuell GmbH). The filter was washed once 148

with 2 ml 0.1 mM LiCl and submerged in Emulsifier Scintillator Plus scintillation fluid 149

(Packard Bioscience). Radioactivity was measured by scintillation counting in a Tri-Carb 150

2000CA liquid scintillation analyser (Packard Instruments). 151

152

Measurement of amino acid and dipeptide concentrations using RP-HPLC 153

Samples were run on a Shimadzu high-speed HPLC Prominence UFLC and later analysed 154

using the LC Solutions 1.24 SP1 software from Shimadzu (Kyoto, Japan). Samples taken at 155

different time points from cell suspensions in 100 mM potassium phosphate buffer, pH 6, 156

containing 10 mM glucose and histidine-leucine, glycine-glutamate or phenylalanine-valine 157

at 5 mM concentrations were centrifuged and supernatants were subjected to derivatization 158

by DEEMM as described by Pudlik and Lolkema (26). Detection of aminoenone derivatives 159

was performed in an Alltech Platinum EPS C18 column with dimensions of 250 x 4.6 mm, 160

operated at 25 °C through a binary gradient using eluent A (25 mM acetate pH 5.8, 0.02 % 161

sodium azide) and eluent B (a 80:20 mixture of acetonitrile and methanol) as described (39). 162

External standards were prepared by mixing dipeptide and the two corresponding free amino 163

acids at concentrations of 4 mM and 0.2 mM or 1 mM and 50 μM, respectively. 164

165

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

8

Results 166

167

Secondary amino acid transporters of Lactococcus lactis 168

The genomes of L. lactis strains IL1403, MG1363, KF147, and SK11 were screened for 169

homologues of amino acid transporters by BLAST searches using different members of the 170

Amino Acid-Polyamine-Organocation (APC) superfamily (21) as query sequences. Fourteen 171

proteins were identified that showed significant homology (more than 20 % amino acid 172

sequence identity) with known APC family members (Table 1 and Supplementary Table S2). 173

Twelve of the proteins were found in all four strains, whereas YlcA (IL1403 nomenclature) 174

was not present in strain SK11 and YrfD was present only in the subsp. lactis strains IL1403 175

and KF147. Sequence identities between the putative transporter proteins varied from 64 % 176

(LysQ and LysP) down to insignificant values below 20% (e.g. ArcD1 and YlcA; see the 177

phylogenetic tree in Fig. 1). All fourteen protein sequences shared the hydropathy profile 178

typical of the APC family members, indicating the same fold containing 12 transmembrane 179

segments and with the N- and C-termini located in the cytoplasm (27,28). The ArcD2 and 180

ArcD1 proteins have one and two additional transmembrane segments, respectively. The 181

arcD1, arcD2, gadC and yrfD genes are located in clusters encoding a putative arginine 182

deiminase (ADI) pathway (arcD1 and arcD2) (29), a glutamate decarboxylation pathway 183

(gadC) (30) and an agmatine deiminase (AgDI) pathway (yrfD) (9), respectively, and, 184

although not demonstrated experimentally, their functions were annotated as 185

arginine/ornithine, glutamate/γ-aminobutyric acid and agmatine/putrescine exchangers, 186

respectively. In contrast, 8 of the other putative amino acid transporter genes appear to be 187

monocistronic, with neighbouring genes of unknown function or functions that could not be 188

directly related to amino acid metabolism. The exception is formed by the adjacent ydgB and 189

ydgC genes that are separated by a 71 bp non-coding fragment which does not contain a 190

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

9

potential terminator or clear promoter sequence. These genes may be transcribed as one 191

mRNA. 192

The brnQ gene was identified in strains MG1363, SK11 and KF147 but missing in IL1403. 193

BrnQ has been proposed to be a branched chain amino acid transporter (19,31) but is not 194

related in amino acid sequence to the APC family transporters. Instead, BrnQ belongs to the 195

LIVCS family of branched chain amino acid/cation symporters (TC 2.A.26) (21). 196

Nevertheless, hydropathy profile alignment suggested that the two families are distantly 197

related and belong to the same structural class (MemGen class ST[2]) (27,28). No other 198

homologues of the LIVCS family were present in any of the four L. lactis strains. Members 199

of the DAACS family (TC 2.A.23), another family that contains many amino acid 200

transporters, were completely absent from the L. lactis genomes. 201

202

Cloning and characterization of the amino acid transporters 203

The putative amino acid transporter genes except for ctrA (aka bcaP), which was well 204

characterized as a branched chain amino acid transporter before (19), were expressed in the 205

MG1363 derived strain NZ9000 using the NICE expression system (24). Cells expressing the 206

genes were screened for enhanced uptake by comparing the uptake of fourteen amino acids 207

and putrescine (only in case of yrfD) to the uptake by control cells harbouring the empty 208

vector pNZ8048. Uptake was measured after 10 sec and with the amino acids present at 209

concentrations ranging from 0.9 μM to 5 μM. Expression of ten transporter genes resulted in 210

increased uptake of one or more amino acids (Table 1), while the remaining four genes 211

(yagE, yibG, yshA and gadC) showed no increased activity with any of the amino acids (data 212

not shown). The positive clones were studied in more detail by following uptake of the 213

identified substrate(s) with time (Fig. 2, left panels; Fig. 3) and by measuring the inhibition of 214

uptake by excess unlabelled amino acids (Fig. 2, right panels). 215

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

10

Overexpression of lysP resulted in a strongly increased initial rate of uptake of lysine (Fig. 216

2A). The inhibition profile showed that LysP has narrow substrate specificity; none of the 217

other amino acids inhibited the uptake of lysine significantly. A 10-fold increase in the initial 218

rate of uptake of histidine was observed upon overexpression of lysQ (Fig. 2B). The 219

inhibition profile indicated that LysQ is mainly a histidine transporter but has low affinity for 220

the other basic amino acids arginine and lysine. The transporter was renamed HisP. The 221

adjacent genes ydgC and ydgB resulted in 20-fold and 5-fold increased initial rates of uptake 222

of serine, respectively (Fig. 2C, D). Uptake by cells expressing ydgC was followed by a rapid 223

efflux of label from the cells that was not observed in the case of ydgB or in the control cells. 224

The nature of this efflux is unclear. The inhibition studies revealed a different substrate 225

specificity of the two transporters. YdgC showed in addition to serine, high affinity for 226

threonine and cysteine, while serine uptake by YdgB was most strongly inhibited by alanine 227

and glycine, whereas threonine and cysteine had much less effect. Transporters YdgC and 228

YdgB were renamed SerP1 and SerP2, respectively. Initial rates of uptake of aspartate and 229

glutamate increased 30-fold and 7-fold, respectively, by cells expressing ylcA (Fig. 2E). The 230

inhibition profile of glutamate uptake revealed that the transporter is specific for the two 231

acidic amino acids. The transporter was renamed AcaP. Strongly increased rates of uptake of 232

phenylalanine and tyrosine were observed upon expressing ysjA. The inhibition profile of 233

tyrosine uptake showed that YsjA also has affinity for tryptophan and that the transporter is 234

highly specific for aromatic amino acids. YsjA was renamed FywP. The ArcD1 and ArcD2 235

genes found in the ADI cluster resulted in 3- to 4-fold increased initial rates of arginine 236

uptake (Fig. 2G, H). Also the uptake of ornithine, the end product of the ADI pathway and 237

counter substrate in arginine/ornithine exchange, was increased more than 10-fold (Table 1). 238

The inhibition profiles of both transporters indicated a main substrate specificity for arginine 239

with low affinity for lysine and histidine. The branched chain amino acids isoleucine, leucine 240

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

11

and valine were taken up in cells expressing brnQ at a two- to four-fold higher initial rate 241

than in control cells (Fig. 3A). Although affinity for other amino acids was not determined, 242

BrnQ appears to be a branched chain amino acid transporter as suggested before (19) and 243

similar to the characterized homologue in Lactobacillus delbrueckii (31). Recently, the yrfD 244

gene present in strain IL1403 but not in MG1363, was proposed to encode the 245

precursor/product exchanger AguD that would mediate the uptake of extracellular agmatine 246

in exchange with intracellular putrescine, the product of the AgDI pathway (9). While control 247

cells did not show significant putrescine uptake activity, as observed previously (32), 248

expression of aguD enabled the cells to take up putrescine with an initial rate of 249

approximately 1.4 nmol/min/mg protein (Fig. 3B). Addition of a 10-fold excess of unlabelled 250

agmatine to cells that had accumulated putrescine led to a rapid efflux of the latter 251

demonstrating the exchange mode of transport. 252

253

Deletion of lysP kills high affinity lysine uptake without effecting growth on free amino acids 254

The lysP gene was deleted from L. lactis JP9000, a MG1363 derivative, using a two-step 255

integration and excision system (47). Lysine uptake activity in resting cells of strain JP9000 256

and JP9000∆lysP grown to mid-exponential phase in GM17 was measured at a lysine 257

concentration of 1.6 μM (Fig. 4A). The initial rate of uptake in the mutant had decreased 258

drastically by approximately 20-fold to 0.5 nmol/min/mg protein suggesting that LysP is the 259

major lysine transporter in L. lactis at low lysine concentrations, but that at least one other, 260

possibly low affinity, uptake system for lysine is present. 261

GM17 is a rich medium containing approximately 2.5 mM lysine and most other free amino 262

acids in the mM range as well as peptides derived from casein by partial hydrolysis (33). 263

Growth characteristics of the wild type and mutant strains in GM17 medium were the same 264

(not shown). Growth of the wild type strain JP9000 in SA medium, a chemically defined 265

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

12

medium containing free amino acids (6) was slightly affected when the standard 1.4 mM of 266

lysine was omitted from the medium. The cells grew a little slower and to a 10% lower 267

optical density indicating that the lysine biosynthesis capacity is not a bottleneck in growth 268

(Fig. 4B). In line with these results, growth of the JP9000∆lysP strain was similar to the wild 269

type strain under the two conditions. A concentration as low as 56 μM lysine in the medium 270

was sufficient to overcome the small growth defect in the wild type strain, while the mutant 271

did not respond to this concentration (Fig. 4B). Apparently, uptake of lysine by LysP was 272

responsible for the growth enhancement at low lysine concentrations observed for the parent 273

strain. 274

275

HisP is essential for growth on free amino acids at low histidine concentrations 276

A hisP deletion mutant was constructed as described above for lysP. Histidine uptake by 277

JP9000∆hisP grown in GM17 medium was strongly reduced to a low, but still significant 278

level when measured at a concentration of 1.5 μM (Fig. 5A). Like observed for LysP above, 279

it appears that HisP is the main high affinity histidine transporter, but that a second transport 280

system, which is expressed at a low level and/or has low affinity for histidine, seems to be 281

present. 282

In rich GM17 medium, which contains approximately 86 μM histidine (33) as well as 283

peptides derived from casein, no significant difference in growth between the wild type and 284

mutant strain was observed (not shown). In the chemically defined SA medium that contains 285

250 μM histidine, JP9000∆hisP grew at a rate of approximately 50% of that of the wild type 286

while the biomass yield was approximately 70% of that of the wild type (Fig. 5B, C). 287

Increasing the concentration of histidine in the medium to 1 mM almost completely restored 288

the growth defect of the mutant which is in line with the presence of a second system able to 289

transport histidine, but with much lower affinity. With no histidine in the medium, both wild 290

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

13

type and ∆hisP mutant cells were unable to grow, demonstrating the inability of L. lactis to 291

synthesize histidine. A histidine concentration as low as 10 μM allowed the wild type strain 292

to grow, but the biomass yield was limited to 40%. A concentration of 50 μM resulted in the 293

same growth as in standard SA medium (Fig. 5B). The mutant strain hardly showed any 294

growth at these two concentrations demonstrating that uptake of histidine by HisP is essential 295

for growth at low concentrations (Fig. 5C). 296

In the absence of free histidine, growth in SA medium of both strains could be fully restored 297

by the addition of 250 μM of the dipeptide histidine-leucine (not shown), demonstrating 298

separate systems for uptake of free histidine and the His-Leu dipeptide, which is probably 299

transported by the di- and tripeptide transporter DtpT (34). 300

301

302

AcaP is essential for growth on free amino acids in the absence of glutamine 303

Resting cells of the ∆acaP mutant JP9000∆acaP grown in SA medium, showed a strongly 304

reduced initial rate of uptake of aspartate at a concentration of 2.2 μM, while the initial rate of 305

uptake of glutamate at 1.9 μM was only marginally affected (Fig. 6A). Since AcaP was 306

shown to transport both aspartate and glutamate (Fig. 2E), the result indicates that a second 307

transporter is present that transports glutamate but not aspartate, and that this transporter is 308

responsible for the uptake of glutamate under the conditions of the experiment. Most likely, 309

the transporter is GlnPQ, an ABC type glutamate/glutamine transporter (20). 310

In GM17 media, rich in all amino acids, no difference in growth between the wild type and 311

acaP deletion mutant was observed. Similarly, in standard SA medium growth of the wild 312

type and mutant strain was not significantly different (compare Fig. 6B and C). SA medium 313

contains 2.1 mM glutamate but no aspartate. Clearly, AcaP plays no role in the uptake of 314

aspartate for biosynthetic purposes during growth. Omitting glutamate from the medium did 315

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

14

not affect growth of wild type. Omitting glutamine, which is present at a concentration of 0.7 316

mM, resulted in only a marginally lower biomass yield. When both glutamate and glutamine 317

were omitted from the medium, no growth was observed (Fig. 6B). Apparently, L. lactis has 318

the capacity to efficiently take up and interconvert glutamate and glutamine. In line, growth 319

of the ΔacaP mutant was not affected when glutamate was omitted from the medium (not 320

shown). The cells are likely to take up glutamine from the medium by an appropriate 321

transporter and internally convert it to glutamate. In contrast, in medium without glutamine, 322

initially, the cells were not able to take up glutamate at a rate high enough to support cell 323

growth demonstrating the role of AcaP as glutamate uptake system during the growth of the 324

wild type cells under the same conditions (Fig. 6C). Apparently, the uptake of glutamate by 325

the mutant cells observed at μM concentrations (Fig. 6A) is not high enough to support the 326

need for glutamate and glutamine in biosynthesis under growth conditions. Remarkably, after 327

a long lag phase of approximately 5 hour the ΔacaP mutant started growing at approximately 328

half of the wild type rate (Fig. 6C). The lag phase indicates an adaptation process, possibly 329

involving the up regulation of a glutamate uptake system other than AcaP (glnPQ). The role 330

of the AcaP transporter during growth on free amino acids in the absence of glutamine is 331

further emphasized by the complete lack of growth of the mutant when, in addition, the 332

concentration of glutamate was reduced to 0.4 mM (Fig. 6C). Growth of the wild type under 333

these conditions was only marginally affected (not shown). 334

335

Deletion of fywP reduces the efficiency of growth at low concentrations of the aromatic 336

amino acids 337

Deletion of the fywP gene from strain JP9000 resulted in a 5-fold lower initial rate of uptake 338

of phenylalanine in cells grown in GM17 medium and when measured at 1 μM concentration 339

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

15

(Fig. 7A). It follows that FywP is the main transporter for phenylalanine uptake under these 340

conditions. 341

No difference in growth was observed between wild type and mutant in rich GM17 medium 342

or chemically defined SA medium which contains 1.2 mM phenylalanine, 0.3 mM tyrosine 343

and 0.5 mM tryptophan. Although these amino acids are known to be not essential for L. 344

lactis MG1363, their presence usually stimulates growth (6,16) suggesting that the rate of 345

cellular biosynthesis is limited. Nevertheless growth of both strains was not affected when the 346

concentration of all three amino acids in SA medium was decreased 5-fold and even when the 347

concentrations were decreased 25-fold into the μM range, the wild type strain grew at the 348

same rate, but to a slightly lower biomass yield (Fig. 7B). Then, the ∆fywP mutant showed an 349

approximately 30% lower growth rate and 20% lower biomass yield, indicating that FywP is 350

important for optimal growth efficiency in an environment with low concentrations of 351

aromatic amino acids. Interestingly, in SA medium with none of the aromatic amino acids 352

present, also a lower growth rate for the fywP mutant was observed which may indicate that 353

FywP plays a role in the recovery of aromatic amino acids that have leaked out of the cells by 354

some other means. 355

356

HisP, AcaP and FywP play no significant role in excretion of accumulated histidine, 357

glutamate and phenylalanine 358

The ∆hisP, ∆acaP and ∆fywP mutants and the wild type were grown to mid-exponential 359

growth phase in SA medium and incubated in a 100 mM potassium phosphate pH 6.0 buffer 360

containing 25 mM glucose and 5 mM of the dipeptides His-Leu (strain NZ9000∆hisP and 361

NZ9000), Gly-Glu (JP9000∆acaP and JP9000) or Phe-Val (JP9000∆fywP and JP9000). The 362

consumption of the dipeptide and excretion of the corresponding free amino acids were 363

analysed by HPLC (Fig. 8). In all cases the dipeptide was taken up efficiently during the first 364

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

16

5 min at rates of 0.17, 0.3 and 0.4 μmol/min/mg protein for His-Leu, Phe-Val and Gly-Glu, 365

respectively. Thereafter, the rates dropped considerably, especially for the Gly-Glu and Phe-366

Val peptides. The free amino acids produced by intracellular peptidase activity appeared 367

readily in the medium at a high rate during the first 5 min after which the rate slowed down in 368

line with the disappearance of the peptide from the medium. The His-Leu and Phe-Val 369

peptides were stoichiometrically converted to the free amino acids by the cells. With the 370

Gly-Glu peptide, the rate of glutamate excretion was 2 times lower than that of glycine 371

excretion during the first 5 min suggesting that part of the free intracellular glutamate was 372

further converted. No products other than glycine, glutamate and glycine-glutamate were 373

detected in the medium. More importantly, for none of the conversions, a difference was 374

observed between the wild type strains and the transporter mutants suggesting that HisP, 375

AcaP nor FywP played a role in the excretion of accumulated amino acids. 376

377

378

379

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

17

Discussion 380

381

Since the emergence of genome sequencing and with the improvement of recombinant DNA 382

techniques, surprisingly little work has been carried out on the identification of amino acid 383

transport systems in lactic acid bacteria, more specifically in the model LAB L. lactis. Most 384

of the knowledge still originates from extensive transport studies in whole cells and 385

membrane vesicles derived from wild type L. lactis strains, performed in the late eighties and 386

early nineties of the previous century (8,14,16,17,35). Only two amino acid transport systems 387

have since been cloned for expression and functional characterization: the ABC-type 388

glutamate/glutamine transporter GlnPQ, encoded by glnP and glnQ (20) and BcaP, a 389

secondary, branched chain amino acid transporter, encoded by bcaP (19). Here, 13 genes of 390

L. lactis MG1363 and one of L. lactis IL1403 (aguD), all but one (brnQ) encoding members 391

of the APC family (21) and annotated either as unknown or as amino acid transporters based 392

on homology, were cloned and overexpressed in L. lactis NZ9000. Substrates were identified 393

(Fig. 2 and 3), and the specificity towards all proteinogenic amino acids was determined (Fig. 394

2, right panels). Four transporters, SerP1, SerP2, AcaP and FywP, for which no function was 395

assigned based on homology, are entirely newly identified amino acid transporters. SerP1 and 396

SerP2 transport serine, AcaP and FywP acidic amino acids and aromatic amino acids, 397

respectively. FywP has the same substrate specificity but, in sequence, is not closely related 398

to the aromatic amino acid transporter AroP. Hence, the different names. LysP and HisP 399

(formerly LysQ) were identified as lysine and histidine transporters, respectively, which was 400

predicted previously based on regulatory sequences in their promoter regions (36,37). The 401

predicted functions of AguD as an agmatine/putrescine exchanger (9) and ArcD1 and ArcD2 402

as arginine/ornithine exchangers (29) were also confirmed. BrnQ, which is not homologous 403

to the other APC family transporters, was shown to be a branched chain amino acid 404

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

18

transporter, in agreement with transport studies with a brnQ knock-out strain described 405

previously (19). No activity could be determined for four genes as overexpression did not 406

result in increased uptake of any of the amino acids. The specificity of GadC that was 407

predicted to be a glutamate/γ-butyric acid exchanger (30) could not be confirmed in this 408

study. 409

Much of the data presented here is in agreement with earlier reports on transport of amino 410

acids studied in wild type cells or membrane vesicles of L. lactis. The presence of separate 411

secondary transport systems for lysine (14), histidine (15), serine and threonine (15), and 412

phenylalanine and tyrosine (16) is confirmed in here by the identification and characterization 413

of LysP, HisP, SerP1 and SerP2, and FywP. In contrast, a secondary transporter for aspartate 414

and glutamate uptake, i.e. AcaP, was not reported before. Instead, the ABC transporter 415

GlnPQ with high affinity for glutamate and glutamine and low affinity for aspartate has been 416

described (20,38). The physiological relevance for having both an ATP-driven glutamate 417

transporter together with a secondary, probably proton symport driven glutamate transporter 418

is unclear, but may be related to the importance of glutamate and glutamine in nitrogen 419

metabolism in the cell. 420

The physiological role of the LysP, HisP, AcaP and FywP transporters was demonstrated 421

using knock-out mutants. Uptake in resting cells demonstrated that the lysP, hisP, acaP and 422

fywP genes encode the major transporters for lysine, histidine, aspartate and tyrosine, 423

respectively, when the substrates were present at μM concentrations. For all these substrates, 424

unknown, low affinity transport systems appeared to be present as well. Glutamate uptake in 425

the acaP mutant, although being a substrate of AcaP, was similar to that in wild type, 426

indicating that a second transport system, most likely GlnPQ, is the major glutamate transport 427

system under these conditions. In chemically defined media containing free amino acids in a 428

range of 0.3 to 3.4 mM but no protein and peptides as additional sources of amino acids (SA 429

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

19

medium (6)), growth of the mutants was similar to that of the wild type strain, which is in line 430

with the presence of additional low affinity uptake systems. The role of the transporters in 431

growth became apparent at low concentration but the effect was strongly dependent on the 432

capacity of L. lactis to synthesize the particular amino acid. While HisP was essential for 433

growth at low concentrations of histidine (Fig. 5B), LysP only restored a small growth defect 434

observed in the wild type when no lysine was present (Fig. 4B). The phenotype of the fywP 435

mutant was in between; a significantly lower growth rate was observed at low concentrations 436

of the aromatic amino acids and even without them (Fig. 7B). FywP transports aromatic 437

amino acids that can be synthesized by L. lactis itself. The effect at low concentrations is in 438

agreement with earlier observations that addition of phenylalanine, and in general, addition of 439

some non-essential free amino acids stimulates growth of L. lactis MG1363 in defined media 440

(6,16). Jensen et al. (39) demonstrated that in the intracellular amino acid pool of L. lactis 441

MG1363 growing in SA medium (containing free amino acids), only aspartate was derived to 442

a substantial proportion (38%) from de novo biosynthesis and that all other amino acids were 443

largely taken up from the medium. This probably reflects regulation to avoid the higher 444

energy costs of biosynthesis compared to the uptake of external free amino acids. The 445

observed difference between the wild type and fywP mutant in SA medium without aromatic 446

amino acids suggests that FywP is involved in retaining these amino acids in the cytoplasm. 447

Then, aromatic amino acids that would leak out of the cells by passive diffusion, would be 448

taken up again by FywP. Excretion or leakage of phenylalanine and tyrosine was indeed 449

observed by Whipp et al. (40) in an E. coli mutant strain that lacks aromatic amino acid 450

transporters. The situation in milk, with casein as the main amino acid source, is more 451

complicated. While casein is preferred over free amino acids as source of amino acids using 452

its efficient proteolytic system (3), it has been demonstrated at the same time that some 453

strains of L. lactis do not reach their maximal growth rate in milk and that this can be 454

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

20

overcome by addition of certain amino acids (41). The implication is that in complex media 455

containing casein as well as free amino acids (for example released by other LAB in mixed 456

fermentations) amino acid transporters still play a role. In this light it is not surprising that 457

dairy strains of L. lactis that have lost their biosynthetic pathways for histidine and branched 458

chain amino acids during their adaptation to milk (1,2), have retained the transporters that can 459

import these as free amino acids, to not become fully dependent on casein as amino acid 460

source. 461

During fermentation in milk, some amino acids are excreted by L. lactis (42,43,44) indicating 462

that the amino acid composition of the peptides taken up after hydrolysis of casein by PrtP is 463

not matching the requirements for biosynthesis. Excretion of the excess amino acids has been 464

proposed to be coupled to the excretion of a proton (proton symport) which would contribute 465

to the generation of metabolic energy in the form of a proton motive force (pmf; energy 466

recycling model). A potential role for HisP, AcaP and FywP as the secondary transporters 467

involved in this process was investigated by feeding the cells dipeptides and measuring the 468

release of the free amino acids in wild type and mutants. The dipeptides histidine-leucine, 469

glycine-glutamate and phenylalanine-valine were efficiently taken up and the free amino 470

acids produced intracellularly by hydrolysis were readily excreted. No difference between the 471

knock-out mutants and wild type was observed suggesting that either a different transport 472

system or passive diffusion was responsible for the release from the cells. Passive diffusion 473

has previously been shown to be a mechanism of excretion of accumulated neutral amino 474

acids (12,45), however, it is unlikely that glutamate diffuses to a significant extent across the 475

membrane. Instead it was demonstrated that in Corynebacterium glutamicum glutamate 476

excretion is mediated by NCgl1221, a mechanosensitive channel which can also mediate 477

betaine transport (46,47,48). A deletion mutant accumulated up to 10-fold higher intracellular 478

glutamate concentrations. In the same organism, other transport systems were found 479

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

21

mediating excretion of lysine and arginine (LysE) (49), threonine and serine (ThrE) (49) and 480

branched chain amino acids and methionine (BrnFE) (50,51). In L. lactis MG1363, only 481

distant homologues of NCgl1221 (llmg_0881 gene) and BrnFE (azlC-llmg_0882 gene pair) 482

can be found. Interestingly, llmg_0881 is located downstream of the gadCB operon (but 483

transcribed in the opposite direction), putatively encoding the glutamate/γ-aminobutyrate 484

pathway proteins (30). Apart from homologues of these systems, other types of carriers 485

mediating excretion of amino acids may be present. 486

487

488

489

Acknowledgements 490

491

We thank Agata Pudlik for assistance with the HPLC analysis of amino acids and dipeptides. 492

This work was supported by the European Community’s Seventh Framework Programme 493

grant agreement No. 211441-BIAMFOOD and by the Netherlands Organization for Scientific 494

Research (NWO-ALW). 495

496

497

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

22

Table 1. Amino acid transporters of L. lactis ssp. lactis and cremoris. 498

Proteina Transporter Initial rate of uptakeb

(nmol/min/mg)

Substrate(s)c

substrate concentration

(μM)

wild type

cells

overexpressing

cells

LysQ HisP histidine 1.5 0.7 4.6 His, Arg, Lys

LysP LysP lysine 1.6 11.5 48 Lys

YdgC SerP1 serine 4.2 2.2 40 Ser, Thr, Cys

YdgB SerP2 serine 4.2 2.2 9.6 Ser, Ala, Gly

YibG - -d - - - -

YsjA FywP phenylalanine

tyrosine

1.0

1.0

0.2

0.2

7.5

9.6

Phe, Tyr, Trp

CtrA BcaP - - - - Leu, Ile, Vale

YagE - -d - - - -

ArcD1 ArcD1 arginine

ornithine

1.5

17.5

0.7

7.5

2.8

108

Arg, Orn, His

ArcD2 ArcD2 arginine

ornithine

1.5

17.5

0.2

7.5

0.9

72

Arg, Orn

GadC GadC -d - - - -

YshA - -d - - - -

YlcA AcaP aspartate

glutamate

2.2

1.9

0.2

0.4

6.6

3

Asp, Glu

YrfD AguD putrescine 4.7 0.2 1.4 Agm, Put

BrnQ BrnQ leucine

isoleucine

valine

1.4

1.5

1.8

0.5

0.7

0.2

1.0

1.2

1.3

Leu, Ile, Val

499 a Annotation of L. lactis ssp. lactis IL1403 genome except for BrnQ. 500 b Uptake measured during the first 10 sec. 501 c Data from uptake, inhibition and exchange studies (see also Fig. 2 and 3) 502

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

23

d No significant difference between control cells and overexpressing cells were found for any 503

of the amino acids 504 e See reference (19) 505

506

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

24

References 507

508

1. Delorme C, Godon JJ, Ehrlich SD, Renault P. 1993. Gene inactivation in Lactococcus 509

lactis: histidine biosynthesis. J. Bacteriol. 175:4391-4399. 510

2. Godon JJ, Delorme C, Bardowski J, Chopin MC, Ehrlich SD, Renault P. 1993. Gene 511

inactivation in Lactococcus lactis: branched-chain amino acid biosynthesis. J. Bacteriol. 512

175:4383-4390. 513

3. Kunji ER, Hagting A, De Vries CJ, Juillard V, Haandrikman AJ, Poolman B, 514

Konings WN. 1995. Transport of beta-casein-derived peptides by the oligopeptide transport 515

system is a crucial step in the proteolytic pathway of Lactococcus lactis. J. Biol. Chem. 516

270:1569-1574. 517

4. Poolman B, Kunji ER, Hagting A, Juillard V, Konings WN. 1995. The proteolytic 518

pathway of Lactococcus lactis. Soc. Appl. Bacteriol. Symp. Ser. 24:65S-75S. 519

5. Savijoki K, Ingmer H, Varmanen P. 2006. Proteolytic systems of lactic acid bacteria. 520

Appl. Microbiol. Biotechnol. 71:394-406. 521

6. Jensen PR, Hammer K. 1993. Minimal Requirements for Exponential Growth of 522

Lactococcus lactis. Appl. Environ. Microbiol. 59:4363-4366. 523

7. Konings WN. 2002. The cell membrane and the struggle for life of lactic acid bacteria. 524

Antonie Van Leeuwenhoek 82:3-27. 525

8. Driessen AJ, Molenaar D, Konings WN. 1989. Kinetic mechanism and specificity of the 526

arginine-ornithine antiporter of Lactococcus lactis. J. Biol. Chem. 264:10361-10370. 527

9. Ladero V, Rattray FP, Mayo B, Martin MC, Fernandez M, Alvarez MA. 2011. 528

Sequencing and transcriptional analysis of the biosynthesis gene cluster of putrescine-529

producing Lactococcus lactis. Appl. Environ. Microbiol. 77:6409-6418. 530

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

25

10. Molenaar D, Bosscher JS, ten Brink B, Driessen AJ, Konings WN. 1993. Generation 531

of a proton motive force by histidine decarboxylation and electrogenic histidine/histamine 532

antiport in Lactobacillus buchneri. J. Bacteriol. 175:2864-2870. 533

11. Driessen AJ, de Jong S, Konings WN. 1987. Transport of branched-chain amino acids 534

in membrane vesicles of Streptococcus cremoris. J. Bacteriol. 169:5193-5200. 535

12. Driessen AJ, Hellingwerf KJ, Konings WN. 1987. Mechanism of energy coupling to 536

entry and exit of neutral and branched chain amino acids in membrane vesicles of 537

Streptococcus cremoris. J. Biol. Chem. 262:12438-12443. 538

13. Driessen AJ, Kodde J, de Jong S, Konings WN. 1987. Neutral amino acid transport by 539

membrane vesicles of Streptococcus cremoris is subject to regulation by internal pH. J. 540

Bacteriol. 169:2748-2754. 541

14. Driessen AJ, van Leeuwen C, Konings WN. 1989. Transport of basic amino acids by 542

membrane vesicles of Lactococcus lactis. J. Bacteriol. 171:1453-1458. 543

15. Driessen AJ. 1987. Amino acid transport in lactic streptococci. PhD thesis. University of 544

Groningen, Groningen, The Netherlands. 545

16. Poolman B. Konings WN. 1988. Relation of growth of Streptococcus lactis and 546

Streptococcus cremoris to amino acid transport. J. Bacteriol. 170:700-707. 547

17. Poolman B, Smid EJ, Konings WN. 1987. Kinetic properties of a phosphate-bond-548

driven glutamate-glutamine transport system in Streptococcus lactis and Streptococcus 549

cremoris. J. Bacteriol. 169:2755-2761. 550

18. Molenaar D, Hagting A, Alkema H, Driessen AJ, Konings WN. 1993. Characteristics 551

and osmoregulatory roles of uptake systems for proline and glycine betaine in Lactococcus 552

lactis. J. Bacteriol. 175:5438-5444. 553

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

26

19. Den Hengst CD, Groeneveld M, Kuipers OP, Kok J. 2006. Identification and 554

functional characterization of the Lactococcus lactis CodY-regulated branched-chain amino 555

acid permease BcaP (CtrA). J. Bacteriol. 188:3280-3289. 556

20. Schuurman-Wolters GK, Poolman B. 2005. Substrate specificity and ionic regulation 557

of GlnPQ from Lactococcus lactis. An ATP-binding cassette transporter with four 558

extracytoplasmic substrate-binding domains. J. Biol. Chem. 280:23785-23790. 559

21. Saier MH Jr. 2000. Families of transmembrane transporters selective for amino acids 560

and their derivatives. Microbiology 146:1775-1795. 561

22. Kuipers OP, de Ruyter PG, Kleerebezem M, de Vos WM. 1998. Quorum sensing-562

controlled gene expression in lactic acid bacteria. J. Biotechnol. 64:15–21. 563

23. Pinto JP, Zeyniyev A, Karsens H, Trip H, Lolkema JS, Kuipers OP, Kok J. 2011. 564

pSEUDO, a genetic integration standard for Lactococcus lactis. Appl. Environ. Microbiol. 565

77:6687-6690. 566

24. De Ruyter PG, Kuipers OP, de Vos WM. 1996. Controlled gene expression systems for 567

Lactococcus lactis with the food-grade inducer nisin. Appl. Environ. Microbiol. 62:3662-568

3667. 569

25. Solem C, Defoor E, Jensen PR, Martinussen J. 2008. Plasmid pCS1966, a new 570

selection/counterselection tool for lactic acid bacterium strain construction based on the oroP 571

gene, encoding an orotate transporter from Lactococcus lactis. Appl. Environ. Microbiol. 572

74:4772-4775. 573

26. Pudlik AM. Lolkema JS. 2012. Rerouting Citrate Metabolism in Lactococcus lactis: 574

Citrate Driven Transamination. Appl. Environ. Microbiol. 78, 6665-6673 575

27. Lolkema JS, Slotboom DJ. 1998. Estimation of structural similarity of membrane 576

proteins by hydropathy profile alignment. Mol. Membr. Biol. 15:33-42. 577

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

27

28. Lolkema JS, Slotboom DJ. 2008. The major amino acid transporter superfamily has a 578

similar core structure as Na+-galactose and Na+-Leucine transporters. Mol. Membr. Biol. 579

25:567-570. 580

29. Budin-Verneuil A, Maguin E, AuffrayY, Ehrlich DS, Pichereau V. 2006. Genetic 581

structure and transcriptional analysis of the arginine deiminase (ADI) cluster in Lactococcus 582

lactis MG1363. Can. J. Microbiol. 52:617-622. 583

30. Sanders JW, Leenhouts K, Burghoorn J, Brands JR, Venema G, Kok J. 1998. A 584

chloride-inducible acid resistance mechanism in Lactococcus lactis and its regulation. Mol. 585

Microbiol. 27:299-310. 586

31. Stucky K, Hagting A, Klein JR, Matern H, Henrich B, Konings WN, Plapp R. 1995. 587

Cloning and characterization of brnQ, a gene encoding a low-affinity, branched-chain amino 588

acid carrier in Lactobacillus delbruckii subsp. lactis DSM7290. Mol. Gen. Genet. 249:682-589

690. 590

32. Romano A, Trip H, Lonvaud-Funel A, Lolkema JS, Lucas PM. 2012. Evidence of 591

two functionally distinct ornithine decarboxylation systems in lactic acid bacteria. Appl. 592

Environ. Microbiol. 78:1953-1961. 593

33. Guimont C. 2002. Change of free amino acids in M17 medium after growth of 594

Streptococcus thermophilus and identification of a glutamine transport ATP-binding protein. 595

Int. Dairy J. 12:729-736. 596

34. Fang G, Konings WN, Poolman B. 2000. Kinetics and substrate specificity of 597

membrane-reconstituted peptide transporter DtpT of Lactococcus lactis. J. Bacteriol. 598

182:2530-2535. 599

35. Driessen AJ. 1989. Secondary transport of amino acids by membrane vesicles derived 600

from lactic acid bacteria. Antonie Van Leeuwenhoek 56:139-160. 601

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

28

36. Rodionov DA, Vitreschak AG, Mironov AA, Gelfand MS. 2003. Regulation of lysine 602

biosynthesis and transport genes in bacteria: yet another RNA riboswitch? Nucleic Acids 603

Res. 31:6748-6757. 604

37. Vitreschak AG, Lyubetskaya EV, Shirshin MA, Gelfand MS, Lyubetsky VA. 2004. 605

Attenuation regulation of amino acid biosynthetic operons in proteobacteria: comparative 606

genomics analysis. FEMS Microbiol. Lett. 234:357-370. 607

38. Poolman B, Driessen AJ, Konings WN. 1987. Regulation of solute transport in 608

streptococci by external and internal pH values. Microbiol. Rev. 51:498-508. 609

39. Jensen NB, Christensen B, Nielsen J, Villadsen J. 2002. The simultaneous biosynthesis 610

and uptake of amino acids by Lactococcus lactis studied by (13)C-labeling experiments. 611

Biotechnol. Bioeng. 78:11-16. 612

40. Whipp MJ, Halsall DM, Pittard AJ. 1980. Isolation and characterization of an 613

Escherichia coli K-12 mutant defective in tyrosine- and phenylalanine-specific transport 614

systems. J. Bacteriol. 143:1-7. 615

41. Hugenholtz J, Dijkstra M, Veldkamp H. 1987. Amino acid limited growth of starter 616

cultures in milk. FEMS Microbiol. Ecol. 45:191-198. 617

42. Cretenet M, Laroute V, Ulve V, Jeanson S, Nouaille S, Even S, Piot M, Girbal L, Le 618

LY, Loubiere P, Lortal S, Cocaign-Bousquet M. 2011. Dynamic analysis of the 619

Lactococcus lactis transcriptome in cheeses made from milk concentrated by ultrafiltration 620

reveals multiple strategies of adaptation to stresses. Appl. Environ. Microbiol. 77:247-257. 621

43. Juillard V, Le BD, Kunji ER, Konings WN, Gripon JC, Richard J. 1995. 622

Oligopeptides are the main source of nitrogen for Lactococcus lactis during growth in milk. 623

Appl. Environ. Microbiol. 61:3024-3030. 624

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

29

44. Niven GW, Knight DJ, Mulholland F. 1998. Changes in the concentrations of free 625

amino acids in milk during growth of Lactococcus lactis indicate biphasic nitrogen 626

metabolism. J. Dairy Res. 65:101-107. 627

45. Anderson JJ, Oxender DL. 1978. Genetic separation of high- and low-affinity transport 628

systems for branched-chain amino acids in Escherichia coli K-12. J. Bacteriol. 136:168-174. 629

46. Hashimoto K, Nakamura K, Kuroda T, Yabe I, Nakamatsu T, Kawasaki H. 2010. 630

The protein encoded by NCgl1221 in Corynebacterium glutamicum functions as a 631

mechanosensitive channel. Biosci. Biotechnol. Biochem. 74:2546-2549. 632

47. Nakamura J, Hirano S, Ito H, Wachi M. 2007. Mutations of the Corynebacterium 633

glutamicum NCgl1221 gene, encoding a mechanosensitive channel homolog, induce L-634

glutamic acid production. Appl. Environ. Microbiol. 73:4491-4498. 635

48. Nottebrock D, Meyer U, Krämer R, Morbach S. 2003. Molecular and biochemical 636

characterization of mechanosensitive channels in Corynebacterium glutamicum. FEMS 637

Microbiol. Lett. 218:305-309. 638

49. Simic M, Sahm H, Eggeling L. 1996. A new type of transporter with a new type of 639

cellular function: L-lysine export from Corynebacterium glutamicum. Mol. Microbiol. 640

22:815-826. 641

50. Kennerknecht N, Sahm H, Yen MR, Patek M, Saier MH Jr, Eggeling L. 2002. Export 642

of L-isoleucine from Corynebacterium glutamicum: a two-gene-encoded member of a new 643

translocator family. J. Bacteriol. 184:3947-3956. 644

51. Trotschel C, Deutenberg D, Bathe B, Burkovski A, Krämer R. 2005. Characterization 645

of methionine export in Corynebacterium glutamicum. J. Bacteriol. 187:3786-3794. 646

647

648

649

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

30

Figure legends 650

651

Figure 1. Phylogenetic tree of all APC family members and BrnQ (LIVCS family) of L. 652

lactis ssp. lactis (IL1403 and KF147) and ssp. cremoris (MG1363 and SK11). Nomenclature 653

of strain IL1403 was used, except for BrnQ. Newly assigned names are shown between 654

parantheses. Protein alignment and tree building was carried out using ClustalX. Amino acid 655

sequence identities between proteins were indicated. 656

657

Figure 2. Substrate specificity of amino acid transporters. Cells of L. lactis NZ9000 658

containing the empty vector pNZ8048 or pNZ8048 with lysP (A), lysQ (B), ydgC (C), ydgB 659

(D), ylcA (E), ysjA (F), arcD1 (G) or arcD2 (H) cloned behind the nisin inducible promoter, 660

were induced for 1 hour with 5 ng/ml nisin and tested for uptake of 14C-labeled amino acids 661

at a concentration of 1 – 5 μM (see Table 1). Left panels: time dependent uptake in cells 662

harbouring the empty vector (closed symbols) or the cells expressing a transporter (open 663

symbols). The substrates are indicated at the y-axis. In E, glutamate uptake is shown in 664

circles and aspartate uptake in triangles. In F, phenylalanine uptake is shown in circles and 665

tyrosine uptake in triangles. Right panels: substrate specificity as determined by inhibition of 666

substrate (indicated at the y-axis) by 1 mM of unlabelled amino acids, indicated at the x-axis 667

with one letter codes. Uptake was determined after 10 seconds of incubation with labelled 668

substrate. The control, i.e. without addition of unlabelled amino acid, was set to 100 % (last 669

bar, Co). Error bars represent standard deviations of the means of two experiments. 670

671

Figure 3. Amino acid uptake via BrnQ (A) and AguD (B). (A) Cells expressing brnQ (open 672

symbols) or harboring the empty vector pNZ8048 (closed symbols) were incubated with 14C-673

leucine (squares), 14C-isoleucine (triangles) or 14C-valine (circles) at concentrations shown in 674

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

31

Table 1. (B) Cells expressing aguD or harbouring pNZ8048 were incubated with 4.7 μM 14C-675

putrescine. To demonstrate agmatine/putrescine exchange, 50 μM of unlabelled agmatine was 676

added to cells that were allowed to accumulate 14C-putrescine for 1 minute (triangles). 677

678

Figure 4. Phenotype of lysP deletion mutant. (A) Uptake of 14C-lysine (at a concentration of 679

1.6 μM) in wild type cells (strain JP9000, open circles) or lysP mutant cells (closed circles). 680

Cells were pregrown to mid-exponential phase in GM17 medium. (B) Growth of JP9000 681

(triangles) and JP9000∆lysP (circles) in SA medium containing 56 μM lysine (closed 682

symbols) or no lysine (open symbols). Shown is a representative experiment of at least two 683

independent experiments. 684

685

Figure 5. Phenotype of hisP deletion mutant. (A) Uptake of 14C-histidine (at a concentration 686

of 1.5 μM) in wild type cells (open circles) or hisP mutant cells (closed circles). Cells were 687

pregrown to mid-exponential phase in GM17 medium. (B and C) Growth of NZ9000 (B) and 688

NZ9000∆lysQ (C) in SA medium containing 1 mM (●), 250 μM (○), 50 μM (▼), 10 μM (∆) 689

or no histidine (■). Shown is a representative experiment of at least two independent 690

experiments. 691

692

Figure 6. Phenotype of acaP deletion mutant. (A) Uptake of 14C-glutamate (circles) or 693

aspartate (triangles) in wild type cells (open symbols) or acaP mutant cells (closed symbols). 694

Glutamate and aspartate were present at 1.9 and 2.2 μM, respectively. Cells were grown to 695

mid-exponential phase in SA medium. (B) Growth of JP9000 in standard SA medium (closed 696

circles) and SA medium from which either glutamate (closed triangles), glutamine (open 697

circles) or both (open triangles) were omitted. (C) Growth JP9000∆acaP in standard SA 698

medium (closed triangles), SA medium without glutamine(open circles) and SA medium 699

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

32

without glutamine and containing 0.4 mM glutamate rather than the standard 2.1 mM (open 700

triangles). Shown is a representative experiment of at least two independent experiments. 701

702

Figure 7. Phenotype of fywP deletion mutant. (A) Uptake of 14C-phenylalanine (at a 703

concentration of 1 μM) in wild type cells (open circles) or ∆fywP mutant cells (closed 704

circles). Cells were grown to mid-exponential phase in GM17 medium. (B) Growth of 705

JP9000 (closed symbols) and JP9000∆fywP (open symbols) in SA medium containing 0.24 706

mM, 0.06 mM and 0.1 mM (circles) or 48 μM, 12 μM and 20 μM (triangles) of 707

phenylalanine, tyrosine and tryptophan, respectively or without any of these amino acids 708

(squares). Shown is a representative experiment of at least two independent experiments. 709

710

Figure 8. Dipeptide uptake and amino acid excretion by wild type and transporter mutant 711

strains. Cells grown to mid-exponential phase in SA medium were harvested and resuspended 712

to an OD600 of 2 in 100 mM KPi pH 6.0 buffer, containing 25 mM glucose and 5 mM 713

histidine-leucine (A), glycine-glutamate (B) or phenylalanine-leucine (C) and incubated at 714

30°C. Concentrations of free amino acids and dipeptide in the buffer were determined from 715

samples taken at indicated time points. Open symbols, wild type strain NZ9000 (A) or 716

JP9000 (B and C); closed symbols, NZ9000∆lysQ (A), JP9000∆ylcA (B) or JP9000∆ysjA (C). 717

Squares: dipeptide concentrations; circles: histidine (A), glycine (B) or phenylalanine 718

concentrations (C); triangles: leucine (A), glutamate (B) or valine concentrations (C). Amino 719

acid concentrations are indicated on the left y-axes; dipeptide concentrations on the right y-720

axes. 721

722

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on May 3, 2018 by guest

http://jb.asm.org/

Dow

nloaded from