+ All Categories
Home > Documents > The reasonable man attempts to adapt ... - reading.ac.uk

The reasonable man attempts to adapt ... - reading.ac.uk

Date post: 11-Nov-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
305
The reasonable man attempts to adapt himself to suit the world: the unreasonable one persists in trying to adapt the world to suit himself. Therefore all progress depends upon the unreasonable man. George Bernard Shaw
Transcript
Page 1: The reasonable man attempts to adapt ... - reading.ac.uk

The reasonable man attempts to adapt himself to suit the world: the unreasonable one persists in trying to adapt the world to suit himself. Therefore all progress depends upon the unreasonable man.

George Bernard Shaw

Page 2: The reasonable man attempts to adapt ... - reading.ac.uk

UNIVERSITY OF READING

School of Animal and Microbial Sciences

The General Amino Acid Permease of Rhizobium leguminosarum biovar viciae

by

David L. Walshaw

Submitted in partial fulfilment of the requirement for the degree of Docter of Philosophy 1995

Page 3: The reasonable man attempts to adapt ... - reading.ac.uk

I declare that this thesis is my own account of my research and that this work

has not previously been submitted for a degree at any University. However, I

would like to acknowledge the help I received from the undergraduate project

students Adam Wilkinson and Mathias Mondy in restriction mapping the

cosmid pRU3004, under my joint supervision with Dr P.S.Poole.

David Walshaw

Page 4: The reasonable man attempts to adapt ... - reading.ac.uk

TABLE OF CONTENTS

Page 5: The reasonable man attempts to adapt ... - reading.ac.uk

Page

CHAPTER 1 LITERATURE REVIEW 1

1.1 INTRODUCTION 2 1.1.1 Rhizobia 2 1.1.2 Nodule formation and structure 2

1.2 BACTEROID METABOLISM 5 1.2.1 Carbon sources supplied to the bacteroid 5 1.2.2 The TCA cycle in the bacteroid 8 1.2.3 The role of poly-β-hydroxybutyrate biosynthesis 13 1.2.4 The malate-aspartate shuttle 14 1.2.5 Other roles of amino acids in bacteroid metabolism 19

1.3 AMINO ACID TRANSPORT IN BACTERIA 27

1.4 ABC TRANSPORTERS 29 1.4.1 Overall structure of ABC transporters 29 1.4.2 The transmembrane domains 32 1.4.3 The ATP-binding domains 35 1.4.4 Periplasmic binding proteins 40 1.4.5 Mechanism of solute translocation 44 1.4.6 The role of binding protein-dependent transporters 49 1.4.7 Regulation of ABC transporters 49

1.5 AMINO ACID TRANSPORT IN RHIZOBIUM 51

1.6 REGULATION INVOLVING NTRC 53

CHAPTER 2 MATERIALS AND METHODS 58

2.1.1 Bacterial strains 59 2.1.2 Culture conditions 68 2.1.3 DNA and genetic manipulations 68 2.1.4 Mutagenesis 69 2.1.5 Transport Assays 70 2.1.6 Isolation of periplasmic fractions and protein gel

electrophoresis

71 2.1.7 Protein binding assays 71 2.1.8 Enzyme assays 72 2.1.9 Metabolite excretion assays 73 2.1.10 Intracellular concentrations 74

Page 6: The reasonable man attempts to adapt ... - reading.ac.uk

Page

2.1.11 Protein determination 75 2.1.12 Plant Assays 75

CHAPTER 3 THE CLONING AND CHARACTERIZATION OF THE GENERAL AMINO ACID PERMEASE OF RHIZOBIUM LEGUMINOSARUM STRAIN 3841

76

3.1 INTRODUCTION 77

3.2 RESULTS 78 3.2.1 Isolation of cosmid pRU3024 carrying the general

amino acid permease genes of Rhizobium leguminosarum strain 3841

78 3.2.2 Restriction mapping, sub-cloning and mutational

analysis of pRU3024

80 3.2.3 Nucleotide sequence of the 5.4kb MluI-ClaI

fragment of pRU135

85 3.2.4 Coding regions of the nucleotide sequence from

pRU189

97 3.2.5 Other features of the nucleotide sequence from

pRU189

106 3.2.6 Mutation of the general amino acid permease 107 3.2.7 Mapping of promoter sites in the aap operon by

complementation analysis

111 3.2.8 Transcription levels of aap genes 113 3.2.9 Amino acid uptake in strains RU542, RU543, RU634

and RU636

114 3.2.10 Growth of strain RU543 on amino acids as sole

source of carbon and nitrogen

116 3.2.11 Amino acid uptake in strain RU640 118 3.2.12 Expression of the R. leguminosarum general amino

acid permease in E. coli

118 3.2.13 Physical properties of the aapJ gene product 120 3.2.14 Effect of aapJ on amino acid uptake in 3841 122 3.2.15 Effect of aapQMP on amino acid uptake in strains

3841

123 3.2.16 Specificity of the aapJ gene product 124 3.2.17 Substrate-binding activity of AapJ 125

Page 7: The reasonable man attempts to adapt ... - reading.ac.uk

Page

3.2.18 Amino acid exchange 130 3.2.19 Plant properties of strains RU542, RU543, RU634,

and RU636

143 3.2.20 Nucleotide sequence of the 0.8kb BamHI fragment

of pRU3024

144 3.2.21 Amino acid transport in strain RU632 147 3.2.22 Plant properties of RU632 149 3.2.23 Nucleotide sequence adjacent to the transposon in

cosmids pRU3053, pRU3082, pRU3083, pRU3084, pRU3085 and pRU3086

150 3.2.24 Amino acid transport in metC mutants of strain 3841 154

3.3 DISCUSSION 155

CHAPTER 4 NITROGEN REGULATION OF THE GENERAL AMINO ACID PERMEASE OF RHIZOBIUM LEGUMINOSARUM STRAIN 3841

161

4.1 INTRODUCTION 162

4.2 RESULTS 163 4.2.1 Effect of the metC-aapJ intergenic region on growth

of strain 3841

163 4.2.2 Effect of nitrogen supply on the transcription of

aapJQM

164 4.2.3 Amino acid uptake in strain RU929 169 4.2.4 Effect of nitrogen supply on the transcription metC

and cysE

170 4.2.5 Sequence analysis of the metC-aapJ intergenic region 173

4.3 DISCUSSION 175

CHAPTER 5 INTER-REGULATION OF THE TCA CYCLE AND THE GENERAL AMINO ACID PERMEASE OF R. LEGUMINOSARUM STRAIN 3841.

177

5.1 INTRODUCTION 178

5.2 RESULTS 179

Page 8: The reasonable man attempts to adapt ... - reading.ac.uk

Page

5.2.1 Aspartate resistant mutants of R. leguminosarum strain 3841

179

5.2.2 Growth of strains RU116, RU118, RU137 and RU156 on succinate and glucose

183

5.2.3 Amino acid transport in strains RU116 and RU156 184 5.2.4 Transductional analysis of strains RU116, RU118,

RU137 and RU156

186 5.2.5 Nucleotide sequence adjacent to the transposon in

strains RU116, RU137 and RU156

187 5.2.6 Activity of TCA cycle enzymes in strains RU116,

RU118, RU137 and RU156

194 5.2.7 Growth of strains RU116, RU137 and RU156 on

arabinose

195 5.2.8 Complementation of strain RU156 196 5.2.9 Effect of pRU3004 on aspartate transport in strains

RU116, RU137, and RU156

197 5.2.10 Southern blot of pRU3004 against RU116, RU137,

RU156 chromosomal DNA

198 5.2.11 Restriction mapping, sub-cloning and mutation of

pRU3004

199 5.2.12 Genes carried by pRU3004 201 5.2.13 β-galactosidase activities from pRU3004 mutants 210 5.2.14 TCA cycle enzyme activities in sucCDAB mutants of

strain 3841

210 5.2.15 Mapping of promoter sites in pRU3004 211 5.2.16 Amino acid excretion by strains RU116, RU156 and

RU543

215 5.2.17 Intracellular concentrations of α-ketoglutarate and

glutamate in strains RU116 and RU156

220 5.2.18 Transcription of the aap operon in sucDA mutants of

strain 3841

222 5.2.19 Plant properties of strains RU116, RU137 and

RU156

224

5.3 DISCUSSION 226

Page 9: The reasonable man attempts to adapt ... - reading.ac.uk

Page

CHAPTER 6 FINAL DISCUSSION 231

6.1.1 The general amino acid permease of Rhizobium leguminosarum

232

6.1.2 TCA cycle enzymes in Rhizobium leguminosarum 237 6.1.3 Methionine biosynthetic enzymes in Rhizobium

leguminosarum

238 6.1.4 Future work 239

REFERENCES

241

Page 10: The reasonable man attempts to adapt ... - reading.ac.uk

ABSTRACT

The four genes, aapJQMP, encoding the general amino acid permease of R.

leguminosarum strain 3841 have been cloned, sequenced, and shown to be

transcribed as a single operon. Sequence homology data indicate that these

genes encode the components of an ABC transporter. However, the

periplasmic binding protein, AapJ, and the two integral membrane

components, AapQ and AapM, are significantly larger than the equivalent

components of previously described ABC transporters of amino acids. The

strong homology of these proteins to sequence from the Escherichia coli genome

sequencing project suggests that E. coli may possess a previously unreported

general amino acid permease. Transcription of the aap operon has been shown

to be negatively regulated by NtrC in response to nitrogen supply.

Mutation of any one of the aap genes resulted in a reduction in the uptake by

strain 3841 of a range of amino acids, including aliphatic amino acids such as

leucine and alanine, and polar amino acids such as glutamate, aspartate and

histidine. Over expression of the aap operon resulted in a marked increase in

the uptake of all the amino acids tested. The results of experiments to

investigate the effect of mutation and over expression of aap genes on amino

acid exchange by strain 3841, appear to indicate that the general amino acid

permease facilitates both uptake and efflux of amino acids. The involvement of

the general permease in amino acid efflux is also indicated by the reduced

glutamate excretion during growth on glucose/NH4Cl/aspartate, exhibited by

an aapJ mutant of strain 3841.

An attempt to isolate general amino acid permease mutants on the basis of

resistance to a toxic concentration of aspartate, led to the discovery that

mutation of genes encoding the TCA cycle enzymes α-ketoglutarate

dehydrogenase and succinyl-CoA synthetase causes almost total abolition of

uptake by the general amino acid permease of strain 3841. This effect has been

shown not to be due to regulation of aap gene expression at the transcriptional

Page 11: The reasonable man attempts to adapt ... - reading.ac.uk

level. Nor does it appear to be the result of substrate-inhibition of the

transporter, despite the fact that α-ketoglutarate dehydrogenase and succinyl-

CoA synthetase mutants are found to accumulate and excrete glutamate. It is

therefore proposed that the general amino acid permease may be subject to

post-translational modification.

The chromosomal region containing contiguous genes that code for malate

dehydrogenase, succinyl-CoA synthetase and α-ketoglutarate dehydrogenase,

has been cloned and partially sequenced.

Mutants in any one of the components of the general amino acid permease

were found to induce pea root nodules that reduce acetylene as effectively as

those of the wild-type strain. α-Ketoglutarate dehydrogenase and succinyl-

CoA synthetase mutants of strain 3841 formed ineffective nodules.

Page 12: The reasonable man attempts to adapt ... - reading.ac.uk

1

CHAPTER 1 LITERATURE REVIEW

Page 13: The reasonable man attempts to adapt ... - reading.ac.uk

2

1.1 INTRODUCTION

1.1.1 Rhizobia

Bacteria of the genus Rhizobium stimulate leguminous plants to develop root

nodules which the bacteria infect and inhabit. Ultimately, the two organisms

establish metabolic co-operation: The bacteria reduce (fix) molecular nitrogen

to ammonia, which is exported to the plant for assimilation; the plant reduces

carbon dioxide to sugars during photosynthesis and translocates these to the

root where they are used to provide the bacteria with a carbon and energy

source.

This dramatic symbiosis is both ecologically and agronomically significant,

leguminous root nodules being by far the largest single source of organic

nitrogen in the global nitrogen cycle. However, the system has an additional

attraction as an area for study. During a complex series of developmental steps,

the bacteria and the plant each influence in the other such fundamental

activities as cell division, gene expression, metabolic function, and cell

morphogenesis. Analysis of these processes may reveal otherwise elusive

components that are parts of plant and bacterial systems for signal

transduction, gene regulation, cell division, and cell wall formation.

1.1.2 Nodule formation and structure

Individual bacterial species and strains nodulate a particular set of host

plants (Table 1.1), and are characterized as having a host range that is either

broad (nodulating many different plants) or narrow (nodulating one or few

hosts). Nodules formed on different plants by different bacteria nonetheless

display striking developmental similarities (Fig. 1.1). Rhizobia attach to the

roots of their host and cause a characteristic curling of the host's root hairs (Yao

& Vincent, 1969; Dazzo & Gardiol, 1984). As this happens, cells in the root

cortex, under the epidermis, start to divide and form the nodule primordium

(Libbenga & Harkes, 1973; Newcomb, 1981). Bacteria trapped in a curled hair,

Page 14: The reasonable man attempts to adapt ... - reading.ac.uk

3

or between a hair and another cell, proliferate and begin to infect the outer plant

cells; as they do so the plant cells are stimulated to produce infection threads

(Callaham & Torrey, 1981). The infection threads ramify and penetrate

individual primordium cells. Bacteria released from infection threads into the

cytoplasm of target cells are enveloped in plant plasma membrane (Robertson et

al., 1978). The bacteria undergo limited DNA replication and division, then

cease both processes. Finally, the endosymbiotic forms of the bacteria, referred

to as bacteroids, begin to fix nitrogen by the action of the enzyme nitrogenase.

Table 1.1 Some Rhizobium-plant associations Bacterial Species Plant Hosts Rhizobium meliloti alfalfa (Medicago sativa) Rhizobium leguminosarum biovar viciae pea (Pisum sativum), vetch (Vicia sativa) biovar trifolii clovers (Trifolium species) biovar phaseoli Phaseolus bean Bradyrhizobium japonicum soybean (Glycine max) Rhizobium spp. NGR234 broad host range; genera including

Vigna, Macroptilium, Lablab, Glycine

Fig. 1.1 Initial stages in the rhizobia-legume symbiosis (Fisher & Long, 1992). Rhizobia attach to host root hairs (a) and cause root hair deformation (b), branching (c) and curling (d). Concomitant mitosis in the root cortex (e) culminates in the formation of the nodule primordium. Bacteria invade the plant cells through a novel structure termed the infection thread (f), a plant-derived tube which delivers the bacteria into individual cells within the nodule primordium. At this point the bacteria differentiate into bacteroids, which can fix atmospheric nitrogen to ammonia.

Page 15: The reasonable man attempts to adapt ... - reading.ac.uk

4

Nitrogenase is irreversibly inactivated by oxygen. However, the plant

derived "peribacteroid membrane" (PBM) contains leghaemoglobin.

Leghaemoglobin binds oxygen then releases it when the local concentration

drops below a certain level, thus providing a high flux for the bacteroid to use

in respiration, but in an environment with low free oxygen (Appleby, 1984).

The specialization of the PBM may also include specific transport or

permeability functions. Since all metabolite exchange between the host and the

bacteroid has to occur across this membrane, it may play an important role in

the regulation of nitrogen fixation.

Page 16: The reasonable man attempts to adapt ... - reading.ac.uk

5

1.2 BACTEROID METABOLISM

1.2.1 Carbon sources supplied to the bacteroid

The process of nitrogen fixation is energetically very costly with the

nitrogenase reaction utilizing as many as 16 molecules of ATP and reductant

equivalent to 8e- per molecule of N2 reduced.

Bacteroids within plant nodule cells are enclosed within the plant derived

PBM, and are dependent on the plant for provision of carbon/energy

substrates.

Nodule homogenates contain a number of carbohydrates and organic acids

in sufficient quantities to be considered as potential carbon substrates for

bacteroid respiration. To be considered seriously, these candidates must be

taken up and metabolized by isolated bacteroids at rates capable of supporting

nitrogenase, and they must be readily able to cross the PBM.

Although it has been shown that nitrogen fixation is fuelled by recently

synthesized sucrose translocated to the nodule (Reibach & Streeter, 1983;

Gordon et al., 1985; Kouchi & Nakaji, 1985), neither sucrose nor the hexoses

resulting from its hydrolysis are readily accumulated by isolated bacteroids.

This was first established for Rhizobium leguminosarum biovar viciae (Hudman &

Glenn, 1980; Glenn & Dilworth, 1981; de Vries et al., 1982) and was later

reported for other rhizobia (Reibach & Streeter, 1984; Saroso et al., 1984;

Salminen & Streeter, 1987b). In contrast to bacteroids, cultured R.

leguminosarum bv. viciae will take up sugars efficiently, indicating that

fundamental changes in carbon transport and metabolism occur in response to

the special environment in nodules (Hudman & Glenn, 1980; de Vries et al.,

1982; San-Francisco & Jacobson, 1986).

Movement of neutral sugars across the PBM of isolated soybean

symbiosomes is only by slow, passive diffusion with rates being inadequate to

support nitrogenase activity (Day & Udvardi, 1989; Udvardi et al., 1990).

Page 17: The reasonable man attempts to adapt ... - reading.ac.uk

6

The distribution of enzymes of carbohydrate metabolism in bacteroids is also

consistent with a minor role for sugars in bacteroid function. For example,

bacteroids appear to lack invertase, as first demonstrated by Robertson and

Taylor (1973), although they may have low sucrose synthase activity (Salminen

& Streeter, 1987b). Low levels of glycolytic enzymes in bacteroids relative to

host cytoplasm have also been demonstrated by numerous groups (Reibach &

Streeter, 1983; Salminen & Streeter, 1987b; Copeland et al., 1989b). Furthermore,

activity of the Entner-Douderoff pathway, a common mechanism for

conversion of hexose to carboxylic acids in Gram-negative bacteria, is very low

in bacteroids, although it is the main pathway of sugar metabolism in cultured

rhizobia (Salminen & Streeter, 1987b). Saroso et al. (1986) have shown that

snake bean bacteroids of cowpea Rhizobium NGR234 do not induce sugar

catabolic enzymes even though snake bean nodules contain significant

concentrations of these substrates and the enzymes systems are inducible in the

free-living form.

The relative unimportance of sugars in N2 fixation is also indicated by

reports that, with one exception (Duncan, 1981), Rhizobium mutants lacking the

capability to metabolize carbohydrates are still able to form effective symbioses

(Arias et al., 1979; Ronson & Primrose, 1979; Cervenansky & Arias, 1984; Glenn

et al., 1984; Arwas et al., 1985; Dilworth et al., 1986; El-Guezzar et al., 1988;

Lafontaine et al., 1989).

The fact that sugars are found in bacteroids is probably due to high

concentrations effecting passive uptake (Reibach & Streeter, 1984) coupled with

the low oxidation rate of these compounds. However, it should be noted that

glucose has been reported to stimulate nitrogenase activity in French bean

bacteroids (Trinchant et al., 1981) and some glucose uptake by French bean

symbiosomes has also been observed (Herrada et al., 1989). Thus French bean

may be an exception to the above generalization.

Page 18: The reasonable man attempts to adapt ... - reading.ac.uk

7

In contrast to neutral sugars, the importance of one or more of the C4-

dicarboxylate tricarboxylic acid (TCA) cycle intermediates, succinate, fumarate

and L-malate in symbiosis has been demonstrated. Numerous dicarboxylic acid

transport (Dct) mutants have been obtained (Table 1.2).

Most of the work has been done with the fast-growing species, but overall

results are remarkably consistent in showing that bacteroids incapable of

dicarboxylate uptake do not fix nitrogen. Furthermore, in contrast to the

sluggish metabolism of sugars, the metabolism of dicarboxylic acids in the

bacteroid is rapid, as illustrated by the relative rates of conversion of labelled

compounds to CO2 by purified bacteroids incubated under microaerobic

conditions (Salminen & Streeter, 1987a).

Table 1.2 Dicarboxylic acid transport mutants of rhizobia red, reduced uptake of dicarboxylic acids and reduced fixation activity; $, exact genetic lesion unknown. Mutants were generated with transposons apart from: #, spontaneous; *, produced with nitrosoguanidine. Rhizobium

species (biovar) Bacterial

phenotype Phenotype in nodule

Reference

leguminosarum

(viciae) Succinate uptake-# Nod+ Fix- Glenn and Brewin (1981)

Dct- Nod+ Fix- Arwas et al. (1985) Dct- Nod+ Fix- Finan et al. (1983) Dctred Nod+ Fixred Finan et al. (1983) meliloti Dct- Nod+ Fix- Engelke et al. (1987) Dctred Nod+ Fixred Engelke et al. (1987) Dct- Nod+ Fix- Bolton et al. (1986) Dct-* Nod+ Fix- Hornez et al. (1989) Dct- Nod+ Fix- Watson et al. (1988) trifolii Dct-* Nod+ Fix- Ronson et al. (1981) leguminosarum

(phaseoli) Dct- Nod+ Fix- LaFontaine et al. (1989)

japonicum Reduced succinate uptake$

Nod+ Fixred Humbeck and Werner (1989)

Page 19: The reasonable man attempts to adapt ... - reading.ac.uk

8

Since dct mutants nodulate and differentiate into bacteroids, it is clear that

they must be able to obtain sufficient carbon from the host plant to fuel such

processes. Despite attempts to determine the identity of the carbon sources

being used under such circumstances, they remain unknown. Double mutants

of R. leguminosarum lacking the Dct system and the ability to catabolize a range

of other carbon sources including C5-, C6-, and C12-sugars (Arwas et al., 1986)

still nodulate as well as the dct- parent strain.

Thus, while it is apparent that dicarboxylic acids are not the sole source of

reducing equivalents utilized for bacteroid formation, the fact that these TCA

cycle intermediates are required for nitrogen fixation implies an important role

for the TCA cycle in the bacteroid.

The observation that nodules induced by a malic enzyme mutant of R.

meliloti fail to reduce acetylene (Driscoll & Finan, 1993) is consistent with this

suggestion, since malic enzyme, which oxidatively decarboxylates malate to

pyruvate (Fig. 1.2), is required for metabolism of dicarboxylates via the TCA

cycle in Rhizobium (Dilworth et al., 1988). In addition, malic enzyme activity has

been demonstrated in R. leguminosarum (McKay et al., 1988) and B. japonicum

(Copeland et al., 1989a; Kimura & Tajima, 1989) bacteroids.

B. japonicum bacteroids have been found to oxidize 14C-labelled succinate,

pyruvate and acetate in a manner consistent with operation of the TCA cycle

(Stovall & Cole, 1978).

1.2.2 The TCA cycle in the bacteroid

The tricarboxylic acid cycle is the major energy-generating pathway in

aerobic heterotrophs, as well as an important source of intermediates for

cellular biosynthesis. It consists of a series of enzyme-catalyzed reactions which

brings about the total oxidation of acetyl units, derived as acetyl-CoA from

pyruvate and other metabolites (Fig. 1.2).

The demands imposed on cell metabolism vary according to the nutritional

environment. Thus in E. coli, which can grow under aerobic and anaerobic

Page 20: The reasonable man attempts to adapt ... - reading.ac.uk

9

conditions, deriving energy from a variety of respiratory and fermentative

processes, the TCA cycle is an inducible pathway which is fully induced only

under strictly aerobic conditions that make the greatest demands on its dual

anabolic and catabolic functions. Under anaerobic conditions all TCA cycle

enzymes in E. coli are repressed to some extent (Gray et al., 1966; Langley &

Guest, 1978; Buck et al., 1986), however the α-ketoglutarate dehydrogenase

complex is particularly affected (Amarasingham & Davis, 1965; Smith &

Neidhardt, 1983). As a result the TCA cycle becomes a branched non-cyclic

pathway (Miles & Guest, 1987; Guest, 1992; Guest & Russell, 1992), in which

carbon flows at a much reduced rate through two routes, an oxidative route

leading to α-ketoglutarate and a reductive route leading to succinate and

succinyl-CoA. These routes fulfil the biosynthetic functions of the TCA cycle.

Page 21: The reasonable man attempts to adapt ... - reading.ac.uk

10

Fig. 1.2 Some potential carbon metabolic pathways in Rhizobium. TCA cycle: CS, citrate synthase; ACN, aconitase; IDH, isocitrate dehydrogenase; KGDH, α-ketoglutarate dehydrogenase; SCS, succinyl-CoA synthetase; SDH, succinate dehydrogenase; FUM, fumerase; MDH, malate dehydrogenase. PHB biosynthesis: KT, β-ketothiolase; HBDH, acetoacetyl-CoA reductase; PHBP, poly-β-hydroxybutyrate synthase. Glutamate degradation: GOGAT, aspartate aminotransferase; GDH, glutamate dehydrogenase; GPT, glutamate-pyruvate aminotransferase; GDC, glutamate decarboxylase; AGT, γ-aminobutyric-glutamic transaminase; SSDH, succinate semialdehyde dehydrogenase. Other: ME, malic enzyme; PDH, pyruvate dehydrogenase.

Page 22: The reasonable man attempts to adapt ... - reading.ac.uk

11

In the nodule Rhizobium is maintained in an oxygen-limited environment.

Under oxygen-limitation, electron transport in the bacterial cell is saturated

with NAD(P)H, and the NAD(P)H:NAD(P) ratio is increased (Jackson &

Dawes, 1976). α-Ketoglutarate dehydrogenase has been shown to be inhibited

by NADH in the Gram-negative bacteria Acetobacter xylinum (Kornfeld et al.,

1977; DeKok et al., 1980) and Acinetobacter sp. (Weitzman, 1972; Hall &

Weitzman, 1977). Furthermore, when the non-symbiotic N2-fixing bacterium,

Azotobacter beijerinckii is subjected to oxygen limitation, α-ketoglutarate

dehydrogenase activity decreases by approximately 70% (Jackson & Dawes,

1976). Levels of NADH and NAD in anaerobically isolated B. japonicum

bacteroids are compatible with possible inhibition of α-ketoglutarate

dehydrogenase (Tajima & Kouzai, 1989; Salminen & Streeter, 1990), and this

enzyme may be a key point for regulation of TCA cycle function in the

bacteroid. Some potential consequences of the inhibition of α-ketoglutarate

dehydrogenase in the bacteroid are discussed in Section 1.2.5.

Activity of other TCA cycle enzymes such as citrate synthase (Kurz & LaRue,

1977), isocitrate dehydrogenase (Kurz & LaRue, 1977; Karr et al., 1984; Irigoyen

et al., 1990), fumarase (Karr et al., 1984) and malate dehydrogenase (DeVries et

al., 1980; Karr et al., 1984; Waters et al., 1985; Kouchi et al., 1988; Irigoyen et al.,

1990) has been detected in bacteroids of various species of Rhizobium and

Bradyrhizobium. The levels of malate dehydrogenase appear to be particularly

high, with activities in R. leguminosarum strain MNF300 bacteroids being 20-fold

greater than those in free-living cells (McKay et al., 1989).

The genes encoding all of the TCA cycle enzymes of E. coli have been cloned,

sequenced and located in the linkage map (Fig. 1.3). There is a major cluster at

16.3 min encoding citrate synthase, succinate dehydrogenase, the specific

components of the 2-oxoglutarate dehydrogenase complex, and succinyl-CoA

synthetase. There is also a smaller cluster at 2.8 min encoding the specific

components of the pyruvate dehydrogenase complex and lipoamide

Page 23: The reasonable man attempts to adapt ... - reading.ac.uk

12

dehydrogenase, the common component of the pyruvate dehydrogenase and 2-

oxoglutarate dehydrogenase complexes. The significance of this clustering is

unclear. The other genes are scattered about the linkage map.

Fig. 1.3 Linkage map of E. coli showing locations of TCA cycle and related genes (Guest & Russell, 1992). Arrows indicate experimentally identified mRNA transcripts (Miles & Guest, 1987). PDHC, pyruvate dehydrogenase complex; CS, citrate synthase; ACN, aconitase; ICDH, isocitrate dehydrogenase complex; ODHC, 2-oxoglutarate dehydrogenase complex; SCS, succinyl-CoA synthetase; SDH, succinate dehydrogenase; FUM, fumarase; MDH, malate dehydrogenase.

Page 24: The reasonable man attempts to adapt ... - reading.ac.uk

13

In the case of Rhizobium relatively few TCA cycle enzymes have been

characterized. The gene encoding isocitrate dehydrogenase in R. meliloti has

recently been cloned and mutated (McDermott & Kahn, 1992). The isocitrate

dehydrogenase mutants were found to form ineffective nodules. This is

compatible with reports that mutants of R. meliloti lacking α-ketoglutarate

dehydrogenase (Duncan & Fraenkel, 1979) or succinate dehydrogenase (Gardiol

et al., 1987) activity also form ineffective nodules. Interestingly, levels of malate

dehydrogenase and succinyl-CoA synthetase in the α-ketoglutarate

dehydrogenase mutant were found to be approximately 4-fold and 5-fold

higher, respectively than in the wild-type (Duncan & Fraenkel, 1979); while

malate dehydrogenase activity in a succinate dehydrogenase mutant isolated by

Finan et al. (1981) was elevated 5-fold.

Malate dehydrogenase from B. japonicum has been purified from bacteroids

and partially characterized (Waters et al., 1985; Emerich et al., 1988).

1.2.3 The role of poly-β-hydroxybutyrate biosynthesis

Bacteroids in some symbioses accumulate large quantities of poly-β-

hydroxybutyrate (PHB). In B. japonicum bacteroids PHB can account for up to

70% of cell dry mass (Wong & Evans, 1971; Bergersen & Turner, 1990). PHB is

derived from acetyl-CoA and its synthesis requires NAD(P)H (Fig. 1.2). Thus,

assuming the TCA cycle fuels the nitrogenase reaction, it appears that PHB

synthesis and nitrogenase compete for energy and reductants.

In Azotobacter oxygen-limitation brings about accumulation of large

quantities of PHB (Senior et al., 1972). Since very low oxygen tensions result in

an increase in the NAD(P)H:NAD(P) ratio (Jackson & Dawes, 1976), and

NADPH and/or NADH inhibit isocitrate dehydrogenase and citrate synthase

(Senior et al., 1972), it has been proposed that in Azotobacter, the formation of

PHB serves a regulatory role (Senior, 1971; Senior et al., 1972). It is suggested

that under oxygen-limitation, accumulated NAD(P)H is channelled into the

Page 25: The reasonable man attempts to adapt ... - reading.ac.uk

14

synthesis of PHB, thereby reducing the inhibition of citrate synthase and

isocitrate dehydrogenase.

By analogy, it has been proposed that PHB biosynthesis in B. japonicum

bacteroids serves a similar purpose (McDermott et al., 1989). Certainly the

results of measurements of pyridine nucleotide redox state in bacteroids

incubated under conditions mimicking the nodule environment are suggestive.

The NADP pool was found to be even more reduced than the NAD pool

(Tajima et al., 1988; Tajima & Kouzai, 1989). Synthesis of PHB may, in part, be

responsible for the lower NADH/NAD ratio since in B. japonicum acetoacetyl-

CoA reductase, the second enzyme in the biosynthetic pathway from acetyl-

CoA to PHB (Fig. 1.2), has a 20-fold lower Km for NADH than NADPH

(Emerich, 1985).

1.2.4 The malate-aspartate shuttle

Various amino acids have been considered as possible carbon sources for

bacteroids. Kahn et al. (1985) focused on glutamate because it is often needed to

induce nitrogen fixation by ex planta cultures of Rhizobium and because

glutamate catabolism can be accompanied by ammonia excretion (O'Gara &

Shanmugam, 1976; Tubb, 1976). To explain why bacteroids continue to fix

nitrogen whilst showing no other signs of nitrogen limitation, (glutamine

synthetase and glutamate synthase (GOGAT) levels are very low in bacteroids,

and the high-affinity ammonia uptake system that is induced by nitrogen

limitation in free-living R. leguminosarum (O'Hara et al., 1985) in not switched

on), Kahn et al. have proposed the model outlined in Fig. 1.4.

In this scheme glutamate (or another nitrogen-containing compound) is fed

to the bacteroid where it is catabolized to yield ammonia or an amino acid,

energy and carbon. The waste products are returned to the plant with no loss

of fixed nitrogen and the plant uses the nitrogen to regenerate glutamate and

complete the cycle. This cycle benefits only the bacteroid since all of the

nitrogen is being recycled. However, if the plant removes some of the ammonia

Page 26: The reasonable man attempts to adapt ... - reading.ac.uk

15

in order to support its own growth, an "appropriate" bacterial response is to

produce more ammonia in order to replenish the fixed nitrogen that serves as a

carrier of carbon and energy. This model elegantly accounts for the repression

of nitrogen assimilatory enzymes in the bacteroid (nitrogen containing

compounds are available to it) and also explains why the bacteroid exports

fixed nitrogen.

Fig 1.4 Nutrient flow between plant and bacteroid, as proposed by Kahn et al. (1985). PBM, peribacteroid membrane; BM, bacteroid membrane.

Page 27: The reasonable man attempts to adapt ... - reading.ac.uk

16

While glutamate can be catabolized in a variety of ways (Fig. 1.2), Kahn et al.

have suggested the operation of a malate-aspartate shuttle between the plant

cytoplasm and the bacteroid (as an extension of the model in Fig. 1.4). In this

pathway malate and glutamate are imported into the bacteroid in exchange for

α-ketoglutarate and aspartate, respectively. Malate is oxidized within the

bacteroid to give oxaloacetate which is then transaminated by glutamate to give

aspartate. α-Ketoglutarate is transaminated in the plant by aspartate to yield

glutamate. The net result is the transfer of NADH into the bacteroid without

the transfer of carbon (Fig. 1.5).

Fig. 1.5 The malate-aspartate shuttle. PBM, peribacteroid membrane; BM, bacteroid membrane.

Page 28: The reasonable man attempts to adapt ... - reading.ac.uk

17

The malate-aspartate shuttle is used to transfer reducing equivalents into

mitochondria (LaNoue & Tischler, 1974; Meijer & Van Dam, 1974; Indiveri et al.,

1987; Dierks & Krämer, 1988) and therefore argue Kahn et al., should not

require novel behaviour from the plant. The needed transport proteins would

be made by the plant but would be located in the PBM instead of the inner

membrane of the mitochondrion.

There has been much debate over the malate-aspartate shuttle as a possible

alternative or supplementary mechanism for the transfer of reducing

equivalents to bacteroids.

Significant levels of malate dehydrogenase (DeVries et al., 1980; Karr et al.,

1984; Waters et al., 1985; Kouchi et al., 1988; McKay et al., 1989; Irigoyen et al.,

1990; Appels & Haaker, 1991) and aspartate aminotransferase (Ryan et al., 1972;

Ryan & Fottrell, 1974; Reynolds et al., 1981; Kouchi et al., 1988; Appels &

Haaker, 1991) have been found in both cytosol and bacteroids of legume

nodules. Malate dehydrogenase is required for TCA cycle activity and cytosolic

aminotransferase might be required for aspartate and asparagine synthesis, so

the presence of these enzymes does not necessarily indicate the operation of a

malate-aspartate shuttle. However, it is difficult to explain high

aminotransferase activity in bacteroids on the basis of aspartate and asparagine

synthesis.

When nitrogenase activity in cultured Bradyrhizobium japonicum was induced

by lowering O2 concentration, etc., aspartate aminotransferase activity nearly

doubled relative to cells in which nitrogenase was repressed (Werner & Stripf,

1978). Furthermore, a mutant of Rhizobium meliloti lacking aspartate

aminotransferase forms nodules which are Fix- (Rastogi & Watson, 1991).

Appels and Haaker (1991) have recently proposed that the main function of a

nodule stimulated cytoplasmic form of aspartate aminotransferase in Rhizobium

leguminosarum strain PRE is participation in a malate-aspartate shuttle. This

suggestion was based on changes in the concentrations of amino acids and

Page 29: The reasonable man attempts to adapt ... - reading.ac.uk

18

organic acids in the incubation medium of pea bacteroids. However, the

findings of Rosendahl et al. (1992) do not support this hypothesis:

Symbiosomes of R. leguminosarum bv. viciae MNF300 incubated with

[14C]malate did excrete [14C]aspartate, and this excretion was increased 3-fold

when unlabelled glutamate was also added as an "exchange substrate" for

aspartate. However, when symbiosomes were incubated with [14C]glutamate

no radioactive α-ketoglutarate could be detected in the incubation medium,

with or without the addition of unlabelled malate.

This disparity in results is apparently not due to impermeability of the

peribacteroid membrane to α-ketoglutarate, since the bacteroid preparations

used in the earlier study were prepared at a high osmotic value and were

reported to contain a majority of intact symbiosomes. However, it is

conceivable that metabolism of carbon storage materials in the bacteroid

affected the carbon pools recorded by Appels and Haaker.

Streeter and Salminen (1990) concluded that an active malate-aspartate

shuttle is not present in B. japonicum bacteroids. Bacteroids supplied with

[14C]malate released only small amounts of [14C]aspartate, and only when

unlabelled glutamate was also provided. Although α-[14C]ketoglutarate was

released by bacteroids when malate, succinate or α-ketoglutarate was supplied

in addition to [14C]glutamate, the formation of labelled α-ketoglutarate was

attributed to a rapid exchange of label via aspartate aminotransferase in the

periplasm.

While glutamate is readily accumulated by isolated Bradyrhizobium japonicum

bacteroids (Salminen & Streeter, 1987a; Udvardi et al., 1988) and will support

their nitrogenase activity (Bergersen & Turner, 1988; Kouchi & Fukai, 1988;

Appels & Haaker, 1991; Kouchi et al., 1991) the observation that the

peribacteroid membrane in soybean and siratro nodules is virtually

impermeable to glutamate (Udvardi et al., 1988; Udvardi & Day, 1988; Ouyang

& Day, 1992) is not compatible with the operation of a malate-aspartate shuttle.

Page 30: The reasonable man attempts to adapt ... - reading.ac.uk

19

However, comparison of respiratory evolution of 14CO2 from [14C]glutamate by

intact symbiosomes with that from naked bacteroids, indicates that glutamate

uptake at higher concentrations may be significant (Kouchi et al., 1991):

Evolution of 14CO2 from [14C]glutamate by B. japonicum symbiosomes was less

than 15% of that by naked bacteroids at a glutamate concentration of 0.1mM,

but about 45% of that at 2mM. On the other hand, rates of respiratory

utilization of [14C]malate by intact symbiosomes were about 60% of those by

naked bacteroids at both concentrations of 0.1 and 2mM. In soybean nodules

estimates of the in vivo concentration of glutamate in the plant cell cytosol range

from 1 to 10mM (Kouchi & Yoneyama, 1986; Streeter, 1987; Bergersen & Turner,

1988). Therefore, the peribacteroid membrane may not work as a potential

barrier to the supply of glutamate to the bacteroids from plant cytosol within

the range of concentrations expected in vivo in the cytosol in nodules.

Glutamate is neither actively accumulated by, nor does it support

nitrogenase activity of, isolated R. meliloti bacteroids (McRae et al., 1989a;

McRae et al., 1989b), although its respiration following passive diffusion into R.

meliloti bacteroids has been reported (Ta et al., 1988).

1.2.5 Other roles of amino acids in bacteroid metabolism

If the operation of an orthodox malate-aspartate shuttle in many symbioses

now seems unlikely, a variety of evidence suggests that glutamate may still

play a central role in bacteroid metabolism.

B. japonicum bacteroids from soybean nodules contain a large pool of

glutamate (Kouchi & Yoneyama, 1986; Streeter, 1987). This fact is consistent

with several studies which indicate that a significant proportion of

malate/succinate supplied to bacteroids is converted to glutamate: Salminen

and Streeter (1987a) provided anaerobically isolated B. japonicum bacteroids

with 14C labelled succinate or malate under microaerobic conditions. After 1

hour 62% of unrespired succinate (32% of the label taken up) and 75% of

unrespired malate (21% of the label taken up) were recovered as glutamate.

Page 31: The reasonable man attempts to adapt ... - reading.ac.uk

20

Significantly, when bacteroids were supplied with [14C]glutamate 97% of

unrespired substrate (46% of the label taken up) was found as glutamate; i.e.

essentially the only fate of glutamate was conversion to CO2. (This has also

been found with R. leguminosarum symbiosomes (Rosendahl et al., 1992)). In

earlier work with B. japonicum bacteroids under low O2 concentrations (Tajima

et al., 1986) [2,3-14C]succinate was found to be incorporated mainly into TCA

cycle acids, with little or no 14C in amino acids. This discrepancy may be due to

the fact that Tajima et al. incubated bacteroids under an Ar-02 mixture, whereas

the later experiments employed N2-O2, so that the bacteroids were actively

fixing N2 during incubation. Indeed glutamate is the most highly labelled

compound in bacteroids following incubation of intact soybean nodules with 15N2 (Ohyama & Kumazawa, 1980).

Kouchi et al. (1991) have also reported high levels of [14C]glutamate in B.

japonicum bacteroids following 20 minute incubations with either [14C]malate or

[14C]glutamate.

A 3-fold increase in glutamate concentration was observed in anaerobically

isolated R. meliloti bacteroids incubated with succinate or malate for 30 minutes

under 4% oxygen plus either argon or nitrogen (Miller et al., 1991). In this case

the bacteroid pool size of free ammonia was found to be sufficient to account

for the glutamate synthesized under argon.

Experiments with intact nodules, in which [14C]malate was generated by the

action of phosphoenolpyruvate carboxylase on 14CO2, appear to confirm the

results obtained with isolated bacteroids (Salminen & Streeter, 1992): Intact

nodules of soybean inoculated with B. japonicum and pea inoculated with R.

leguminosarum were detached and immediately fed 14CO2 for up to 6 minutes.

Bacteroids rapidly purified from these nodules contained most label in malate.

However, the rate of glutamate labelling was 67% of the rate of malate labelling

in the case of B. japonicum bacteroids, while in R. leguminosarum bacteroids

isolated after 6 minutes incubation, glutamate contained 78% of the amount of

Page 32: The reasonable man attempts to adapt ... - reading.ac.uk

21

label found in malate. Since the radioactivity in glutamate in the cytosolic

fraction of the nodules was found to be low, the labelling of glutamate in the

bacteroids would seem indicative of malate metabolism rather than uptake of

labelled glutamate, especially since the uptake of malate by isolated B.

japonicum bacteroids is 3.4 times faster than that of glutamate (Salminen &

Streeter, 1991).

It has been suggested (McDermott et al., 1989; Salminen & Streeter, 1990) that

glutamate synthesis from malate under the oxygen-limiting conditions of the

nodule is the consequence of diversion of α-ketoglutarate away from the TCA

cycle due to inhibition of α-ketoglutarate dehydrogenase by NADH (Section

1.2.2). The fact that no accumulation of radioactivity in α-ketoglutarate is seen

in bacteroids supplied with [14C]malate (Salminen & Streeter, 1990), probably

reflects the equilibrium constant (1014) of the glutamate dehydrogenase

catalysed reaction for the formation of glutamate, and the high ammonia

concentrations in N2-fixing bacteroids (Klucas, 1974; Streeter, 1989). Indeed

ammonia and α-ketoglutarate are extremely efficient inducers of glutamate

dehydrogenase in B. japonicum (Fottrell & Mooney, 1969).

Glutamate may also be formed from α-ketoglutarate via glutamine

synthetase and glutamate synthase, both of which have been detected in R.

leguminosarum and B. japonicum bacteroids (Brown & Dilworth, 1975). In fact,

glutamate synthase activity in B. japonicum bacteroids increases as the symbiosis

progresses, whereas that of glutamate dehydrogenase declines (Stripf &

Werner, 1978); and glutamate synthase activity doubles when B. japonicum is

shifted from nitrogenase repressed to nitrogenase derepressed conditions

(Werner & Stripf, 1978). Furthermore, a NADP-dependent glutamate synthase

deficient mutant of B. japonicum forms ineffective nodules (O'Gara et al., 1984),

while a B. japonicum mutant lacking glutamine synthetase activity fails to

nodulate soybeans (Carlson et al., 1987).

Page 33: The reasonable man attempts to adapt ... - reading.ac.uk

22

However, glutamate synthase does not appear to be significant in R. meliloti

bacteroids, as mutants lacking this enzyme are Fix+ (Kondorosi et al., 1977;

Osburne & Signer, 1980; Lewis et al., 1990). In addition, none of the three forms

of glutamine synthetase, encoded by the genes glnA, glnII and glnT (Section 1.6),

is essential for symbiotic nitrogen fixation in R. meliloti, since strains carrying a

single mutation in any one of these genes are Fix+ (de Bruijn et al., 1989;

Somerville et al., 1989). Furthermore, assuming that glnT is not expressed in

bacteroids, as appears to be the case for R. leguminosarum (Espin et al., 1990), the

Fix+ phenotype of a glnA glnII double mutant (de Bruijn et al., 1989), suggests

that no glutamine synthetase activity is required for nitrogen fixation by R.

meliloti bacteroids.

A possible fate of glutamate derived from α-ketoglutarate is decarboxylation

to form γ-aminobutyrate (GABA), then deamination and oxidation to succinate

(Fig. 1.2). This "GABA-shunt" would provide a detour around α-ketoglutarate

dehydrogenase, removing an equivalent amount of CO2 and yielding the same

end product, succinate, which can re-enter the TCA cycle. While this pathway

appears to operate in R. meliloti bacteroids (Fitzmaurice & O'Gara, 1991; Miller

et al., 1991), the results of tracer experiments and the detection of only extremely

low levels of glutamate decarboxylase activity, suggest that the GABA-shunt is

essentially absent from B. japonicum bacteroids (Salminen & Streeter, 1990;

Kouchi et al., 1991).

There is evidence to suggest that the amino acids aspartate and alanine may

also play a significant part in bacteroid metabolism.

Like glutamate, aspartate is actively accumulated by B. japonicum bacteroids

(Reibach & Streeter, 1984), and at controlled substrate and oxygen

concentrations, stimulates respiration (Salminen & Streeter, 1987a), and

supports nitrogenase activity (Salminen & Streeter, 1987a; Bergersen & Turner,

1988; Kouchi & Fukai, 1988; McRae et al., 1989b; Appels & Haaker, 1991; Kouchi

et al., 1991). B. japonicum bacteroids contain relatively high concentrations of

Page 34: The reasonable man attempts to adapt ... - reading.ac.uk

23

aspartate (Streeter, 1987) but exogenously supplied aspartate appears to be

catabolized without being incorporated into the endogenous aspartate pool

(Salminen & Streeter, 1987a; Tajima & Kouchi, 1990; Kouchi et al., 1991). On

incubation of anaerobically isolated B. japonicum bacteroids with [14C]aspartate

the largest amount of unrespired label is found as glutamate (Salminen &

Streeter, 1987a; Tajima & Kouchi, 1990; Kouchi et al., 1991), and the failure of

amino-oxyacetate, a specific inhibitor of transamination, to affect the utilization

of aspartate (Tajima & Kouchi, 1990; Kouchi et al., 1991), indicates that aspartate

utilization occurs via the TCA cycle after direct deamination by aspartase. In

contrast, glutamate utilization is strongly inhibited by the presence of amino-

oxyacetate, and it is suggested (Tajima & Kouchi, 1990; Kouchi et al., 1991) that

the major fate of exogenously supplied glutamate in B. japonicum bacteroids is

transamination to form aspartate.

While actively accumulated by a low-affinity transport system, aspartate

does not support nitrogenase activity in R. meliloti bacteroids (Miller et al., 1988;

McRae et al., 1989b). However, the observation that a Tn5 induced mutant of R.

meliloti lacking aspartate aminotransferase does not grow on aspartate as a

carbon source and forms Fix- nodules has prompted Rastogi and Watson (1991)

to suggest that aspartate provided by the plant to the bacteria in the nodule is

essential for an effective symbiosis.

Recently isolated isocitrate dehydrogenase mutants of R. meliloti (McDermott

& Kahn, 1992) are found to be strict glutamate auxotrophs. Since these mutants

are Nod+, the host must be directly or indirectly providing enough glutamate

for cell growth and maintenance. However, their Fix- phenotype suggests that

either α-ketoglutarate or glutamate is not provided in large quantities or that

glutamate catabolism is of relatively little importance. McDermott and Kahn

have suggested that glutamate obtained via aspartate aminotransferase may be

significant in R. meliloti (α-ketoglutarate is required for production of glutamate

in this way, and isocitrate dehydrogenase mutants cannot synthesize α-

Page 35: The reasonable man attempts to adapt ... - reading.ac.uk

24

ketoglutarate). This explanation could account for both the Fix- phenotype of R.

meliloti GABA-shunt mutants (Fitzmaurice & O'Gara, 1988) and the Fix+

phenotype of R. meliloti glutamate synthase mutants (Kondorosi et al., 1977;

Osburne & Signer, 1980; Lewis et al., 1990), and is also consistent with aspartate

having an important role in R. meliloti bacteroid metabolism.

Alanine is neither actively accumulated by, nor does it support nitrogenase

activity of, isolated R. meliloti bacteroids (McRae et al., 1989a; McRae et al.,

1989b). However, in anaerobically isolated B. japonicum bacteroids incubated

with [14C]succinate under microaerobic conditions for 1 hour, 11% of the label

taken up was recovered as alanine (Salminen & Streeter, 1987a). Curiously, the

same figure after incubation with [14C]malate was only 3%. In their

experiments involving incubation of pea and soybean nodules with 14CO2, in

which malate was the major form of labelled carbon supplied to the bacteroids

(see above), Salminen and Streeter (1992) found that R. leguminosarum

bacteroids accumulated as much label in alanine as in glutamate (74% of the

amount of label found in malate), while in B. japonicum bacteroids the rates of

labelling of alanine and aspartate were 18 and 24% that of malate, respectively.

Several groups have reported excretion of alanine and/or aspartate by

nitrogen fixing bacteroids. Kretovich et al. (1986) incubated bacteroids of

Rhizobium lupini 359a with succinate, malate or fumarate under nitrogen and 6%

oxygen for up to 30 minutes. Aspartate, and smaller amounts of alanine were

found to be excreted into the incubation medium in each case, with the greatest

excretion (37 nmole of aspartate per gram of bacteroids per minute) being

observed when malate was the substrate. Appels and Haaker (1991) noted the

presence of alanine in the medium after incubation of R. leguminosarum strain

PRE bacteroids with malate plus glutamate or aspartate plus α-ketoglutarate

under nitrogen fixing conditions. They also reported glutamate-pyruvate

aminotransferase activity in the bacteroids and proposed the formation of

alanine through transamination of pyruvate by glutamate. (Pyruvate is formed

Page 36: The reasonable man attempts to adapt ... - reading.ac.uk

25

from malate by the action of malic enzyme, and glutamate and malate would be

formed from aspartate and α-ketoglutarate by a reverse malate-aspartate

shuttle).

Kouchi et al. (1991) found marked accumulation of alanine, but not aspartate,

in the medium when anaerobically isolated bacteroids of B. japonicum A1017

were incubated with malate for 20 minutes under N2 and 1.5% O2. When

bacteroids were incubated with glutamate, aspartate, but not alanine,

accumulated in the medium. However, when bacteroids were incubated with

malate plus glutamate the accumulation in the medium of both aspartate and

alanine, particularly aspartate, increased greatly.

Similar results have been obtained with anaerobically isolated R.

leguminosarum symbiosomes from pea nodules (Rosendahl et al., 1992):

Incubation of symbiosomes with 0.5mM [14C]malate under 4% O2 resulted in

14% of the label recovered after 30 minutes being detected in alanine; 7% inside

the symbiosomes and 7% in the surrounding medium. Aspartate plus

glutamate (indistinguishable by TLC) accounted for 7% of the label: 2%

intracellular, 5% extracellular. When symbiosomes were incubated with 0.5mM

[14C]malate plus 0.5mM unlabelled glutamate under the same conditions the

results were as follows: 9% of the label recovered as alanine inside

symbiosomes; 25% as alanine in the incubation medium; 3% as

aspartate/glutamate inside symbiosomes; 17% as aspartate/glutamate in the

incubation medium.

Rosendahl et al. suggest the following explanation of these results: Labelled

malate is transported across the PBM and bacteroid membrane into the

bacteroids where it is converted to oxaloacetate (malate dehydogenase) and

pyruvate (malic enzyme). Transamination from amino donors such as

glutamate then produces radioactive aspartate and alanine. When exogenous

glutamate is added, the release of alanine and aspartate is accelerated because

glutamate taken into bacteroids is functioning as an additional amino donor in

Page 37: The reasonable man attempts to adapt ... - reading.ac.uk

26

transamination, and/or the rates of amino acid transport across the bacteroid

and/or peribacteroid membrane are increased. The latter suggestion is

prompted by reports of common amino acid transporters capable of exchanging

intra- and extracellular pools of amino acids in R. leguminosarum (Poole et al.,

1985) and cowpea Rhizobium (Glenn et al., 1991).

It has been argued (McDermott et al., 1989) that under the oxygen-limited

conditions of the bacteroid where citrate synthase, isocitrate dehydrogenase

and α-ketoglutarate dehydrogenase may be repressed relative to malate

dehydrogenase (Senior, 1971; Jackson & Dawes, 1976; Karr et al., 1984;

Suganuma & Yamamoto, 1987; McKay et al., 1989), oxaloacetate derived from

actively accumulated malate (or succinate) would build up in the absence of

outlets other than the TCA cycle. Since oxaloacetate competitively inhibits

succinate dehydrogenase, removal or utilization of excess oxaloacetate may be

essential in maintaining the TCA cycle. Excretion of aspartate (derived from

oxaloacetate via aspartate aminotransferase), and/or excretion of alanine

(derived from malate via malic enzyme and glutamate pyruvate transaminase)

by the bacteroid could provide a means of preventing accumulation of

oxaloacetate. If this is the case, amino acid transport across the

bacteroid/peribacteriod membrane, the likely rate determining step in the

oxaloacetate depletion process, could regulate bacteroid metabolism. This idea

is not without precedent since the malate-aspartate shuttle in mitochondria is

regulated by an amino acid exchange transporter (Murphy et al., 1979).

Page 38: The reasonable man attempts to adapt ... - reading.ac.uk

27

1.3 AMINO ACID TRANSPORT IN BACTERIA

The prokaryotic amino acid uptake systems reported to date can be divided

into two categories: (i) periplasmic binding protein-independent, secondary

solute transport systems in which the free energy for accumulation of the amino

acid(s) is supplied by an electrochemical gradient (ii) periplasmic binding

protein-dependent transport systems which utilize ATP hydrolysis to energize

translocation.

Amino acid uptake systems in category (i) are usually symporters, mediating

the coupled movement of two (or more) solutes in the same direction.

Hydropathy analyses and other topographic studies on transport proteins of

this type, point, in most cases, to a secondary structure that includes twelve

hydrophobic domains in α-helical configuration, traversing the membrane in a

zig-zag fashion (Poolman & Konings, 1993). Examples of this type of

transporter are the proton/glutamate symporter (GltP) of Bacillus subtilis

(Tolner et al., 1995), the sodium/proline symporter (PutP) of Salmonella

typhimurium (Miller & Maloy, 1990; Poolman & Konings, 1993), and the

sodium/proton/glutamate symporter (GltT) of Bacillus stearothermophilus and

Bacillus caldotenax (Tolner et al., 1992a). Such transporters are not of direct

relevance to the research reported in this thesis and are further discussed only

briefly in Section 1.4.6. Transporters in category (ii) are described in Section 1.4.

Bacterial amino acid transport systems are usually highly specific for a single

amino acid, or group of related amino acids (Halpern, 1974; Landick et al., 1985;

Antonucci & Oxender, 1986). However, general amino acid permeases, which

transport all amino acids, have been found in the eukaryotic microorganisms

Neurospora and Saccharomyces (Pall, 1969; Rao et al., 1975; DeBusk & DeBusk,

1980; Ogilvie-Villa et al., 1981; Grenson et al., 1970; Grenson & Hon, 1972; Eddy,

1982; Jauniaux & Grenson, 1990).

Transport of one amino acid may be mediated by several different permeases

with overlapping specificities. Thus in E. coli three L-glutamate transport

Page 39: The reasonable man attempts to adapt ... - reading.ac.uk

28

systems have been identified: (i) a binding protein-dependent glutamate-

aspartate system (Schellenberg & Furlong, 1977); (ii) a proton/glutamate-

aspartate system (Wallace et al., 1990; Tolner et al., 1992b); (iii) a

sodium/glutamate system (Deguchi et al., 1989; Deguchi et al., 1990; Kalman et

al., 1991). Some potential reasons for such multiplicity of transport systems are

discussed in Section 1.4.6.

Page 40: The reasonable man attempts to adapt ... - reading.ac.uk

29

1.4 ABC TRANSPORTERS

Prokaryotic periplasmic binding protein-dependent permeases are members

of the ABC superfamily of membrane transporters. Members of this

superfamily have extensive sequence and structural similarity, and the

designation ABC transporters recognizes a highly conserved ATP-binding

cassette, a particularly characteristic feature of family members that is

indicative of the use of ATP hydrolysis in the energization of substrate

translocation. The name traffic ATPases has also been suggested for these

transporters (Mimura et al., 1990).

ABC transporters have been identified in both prokaryotes and eukaryotes.

Among the eukaryotic systems are the medically important cystic fibrosis

transmembrane regulator (CFTR) and multidrug resistance (MDR) protein.

Each ABC transporter is relatively specific for a given substrate. However,

different transporters handle a wide variety of substrates including amino

acids, sugars, inorganic ions, polysaccharides and peptides (Ames, 1986;

Furlong, 1987; Higgins et al., 1990). Some ABC transporters are uptake systems,

while others export substrate from the cell: none has yet been identified that

can pump in both directions (Higgins, 1992).

1.4.1 Overall structure of ABC transporters

ABC transporters require the function of multiple polypeptide/protein

domains, organized in a characteristic fashion (Fig. 1.6). The typical transporter

consists of four membrane-associated domains. Two of these domains are

highly hydrophobic and each usually consists of six membrane-spanning

segments. These domains form the pathway through which substrate crosses

the membrane, and are believed to play a significant part in determining the

substrate specificity of the transporter. The other two domains are peripherally

located at the cytoplasmic face of the membrane, bind ATP and couple ATP

hydrolysis to the transport process.

Page 41: The reasonable man attempts to adapt ... - reading.ac.uk

30

Fig 1.6 Structural organization of a typical ABC transporter (Higgins, 1992). The location of the short sequence motif conserved between the transmembrane domains of many of these transporters is indicated by *.

The individual domains of an ABC transporter are frequently expressed as

separate polypeptides, particularly in prokaryotic species. The oligopeptide

transporter of S. typhimurium (Hiles et al., 1987) is an example of this

arrangement (Fig. 1.7A). However, there are examples in which two or more

domains are fused into larger, multifunctional polypeptides, as in the ribose

transporter (Bell et al., 1986) of E. coli (Fig. 1.7B). Many eukaryotic ABC

transporters, such as the MDR protein and the CFTR gene product, have all four

domains fused into a single protein (Fig. 1.7C).

Page 42: The reasonable man attempts to adapt ... - reading.ac.uk

31

Fig. 1.7 Examples of differing domain organization in ABC transporters (Higgins, 1992).

Page 43: The reasonable man attempts to adapt ... - reading.ac.uk

32

In addition to the four core domains, all bacterial ABC transporters that

mediate solute uptake require a substrate-binding protein located in the

periplasm (Figs. 1.7A, B, D). These periplasmic components are essential for the

function of the transporter with which they are associated, although they are

not integral to the process of transmembrane solute translocation itself. By

contrast, there is no evidence that eukaryotic ABC transporters utilize a soluble

receptor (a periplasmic binding protein equivalent) to transfer substrate to the

membrane-bound components of the transporter.

No ABC transporter has yet been shown to function with fewer than the four

core domains, and in the absence of evidence to the contrary, it is assumed that

the four core domains form the basic unit required to mediate solute

translocation. It is generally assumed that a functional transport complex

consists of one of each of the four core domains rather than a larger oligomeric

assembly. However, at present, this is purely an assumption. A few ABC

transporters appear to lack a full complement of domains. For example, the

operon encoding the histidine transporter of S. typhimurium includes only a

single gene, hisP, encoding an ATP-binding component. However,

coimmunoprecipitation of HisQMP has recently shown that this transport

complex contains HisP in a 2:1 ratio with the other domains (Kerppola et al.,

1991), suggesting that HisP functions as a homodimer (Fig. 1.7D). Similar

results have also been obtained for the maltose system of E. coli (Davidson &

Nikaido, 1991). The core components of minimalist transporters such as the

glutamine uptake system in E. coli, which, in addition to a binding protein gene,

glnH, has only one hydrophobic membrane component gene, glnP, and one

ATP-binding component gene, glnQ, (Nohno et al., 1986), probably function as

homodimers, although this has yet to be tested experimentally.

Page 44: The reasonable man attempts to adapt ... - reading.ac.uk

33

1.4.2 The transmembrane domains

The two transmembrane domains of ABC transporters are highly

hydrophobic. Each is predicted, from its sequence, to consist of multiple α-

helical segments that could span the membrane. The majority of transporters

are predicted to have six membrane-spanning segments per domain (a total of

twelve per transporter), with the N- and C-termini on the cytoplasmic face of

the membrane and three extracellular and two intracellular loops (Fig. 1.6). The

available experimental data are consistent with these predictions. Thus the six

predicted transmembrane segments of each domain of the oligopeptide

permease of S. typhimurium have been identified experimentally, using both

biochemical techniques and β-lactamase fusion analysis of OppB and OppC

(Pearce et al., 1992).

A few ABC transporters apparently do not conform to the two-times-six

transmembrane helix paradigm. For example, TnphoA fusion analysis of MalF

from the maltose transporter of E. coli (Froshauer et al., 1988), suggests a total of

eight membrane-spanning regions for this protein, in agreement with

computer-assisted predictions. Alignment with equivalent components of other

transporters (Overduin et al., 1988) indicates that MalF consists of the six

standard transmembrane segments, but has an N-terminal extension with two

additional transmembrane segments (Fig. 1.8). Furthermore, these two

transmembrane segments can be deleted without loss of MalF function

(Ehrmann et al., 1990). Another potential exception is the histidine transporter

of S. typhimurium, which has two hydrophobic components, HisQ and HisM,

each predicted to have five, rather than six, transmembrane segments (Higgins

et al., 1982b; Ames, 1985). These predictions have recently been confirmed

experimentally through TnphoA fusion analysis of HisQ and HisM, and studies

involving use of proteolytic enzymes and antibodies with oriented membrane

vesicles (Kerppola et al., 1991; Kerppola & Ames, 1992). Alignment with other

transporters (Fig. 1.8) indicates that the usual N-terminal transmembrane

segment of each domain may be absent, which places the N-termini on the

Page 45: The reasonable man attempts to adapt ... - reading.ac.uk

34

exterior of the cell. This orientation may indicate that ten (two-times-five)

transmembrane segments provide the minimal unit required to form the

translocation pathway itself; the additional N-terminal transmembrane

segment(s) of most transporters may simply facilitate correct folding, packing,

and orientation within the membrane.

Fig. 1.8 Alignment of OppC, HisQ and MalF, showing membrane-spanning segments as black boxes (Higgins, 1992). The majority of integral membrane components of ABC transporters are exemplified by OppC, which has six transmembrane segments. HisQ (and HisM) appears to lack transmembrane segment 1, while MalF has two additional N-terminal transmembrane segments. The conserved motif in these proteins (indicated by *) is located between transmembrane segments 4 and 5.

Comparison of the amino acid sequences of the hydrophobic components of

one transporter with those of another reveals little or no similarity. This implies

that the structural constraints required for the function of the hydrophobic

domains can be satisfied by a variety of amino acid combinations. The only

significant sequence conservation between the transmembrane domains of

several different ABC transporters is a short motif identified on many bacterial

transporters (Ames, 1985; Dassa & Hofnung, 1985; Kerppola & Ames, 1992;

Saurin et al., 1994), appropriately positioned on a cytoplasmic loop (Figs. 1.6

and 1.8) to interact with the ATP-binding domains (Pearce et al., 1992). Whether

or not it serves this function has yet to be established.

Page 46: The reasonable man attempts to adapt ... - reading.ac.uk

35

The two integral membrane domains of an individual ABC transporter are

generally more closely related to each other than they are to the equivalent

domains of other transporters in the superfamily. Thus the HisQ and HisM

components of the S. typhimurium histidine transporter are closely related to

each other (Higgins et al., 1982b), as are the OppB and OppC components of the

oligopeptide transporter (Hiles et al., 1987). This similarity implies that the two

domains function symmetrically as a pseudodimer. This would be consistent

with the view that the single hydrophobic domain found in some ABC

transporters, may function as a homodimer.

While, for bacterial uptake systems, the periplasmic binding proteins are

important in determining substrate specificity (Section 1.4.4), it is clear that the

transmembrane components also play a role. Binding protein-independent

mutants still exhibit substrate selectivity (Treptow & Shuman, 1985; Petronilli &

Ames, 1991), and mutations that alter the specificity of a transporter invariably

alter the transmembrane domains. Thus, mutations that alter the selectivity of

the histidine transporter of S. typhimurium from L-histidine to L-histidinol delete

four amino acids from a membrane-spanning segment of the transmembrane

domain HisM (Payne et al., 1985). Similarly, mutations that allow the maltose

transporter of E. coli to transport the analogue p-nitrophenyl-α-maltoside (not

normally a substrate) also alter the transmembrane domains (Reyes et al., 1986).

1.4.3 The ATP-binding domains

The ATP-binding domains of ABC transporters are their most characteristic

feature. Each domain is about 200 amino acids long and the domains from

different transporters share considerable sequence identity. The conserved

sequences include two short motifs, the Walker motifs, associated with many

nucleotide-binding proteins (Walker et al., 1982; Higgins et al., 1985; Higgins et

al., 1986).

The ATP-binding domains are highly hydrophilic, include no potential

membrane-spanning segments, and would not normally be expected to span

Page 47: The reasonable man attempts to adapt ... - reading.ac.uk

36

the membrane. Nevertheless, these components appear to be tightly associated

with the cytoplasmic membrane, although the nature of this association is not

clear. In some cases there is evidence to suggest peripheral attachment to the

cytoplasmic face of the membrane. Thus the MalK component of the maltose

transporter of E. coli is found to be cytoplasmic in the absence of the integral

membrane components (Shuman & Silhavy, 1981), while OppF, the ATP-

binding component of the S. typhimurium oligopeptide transporter, is less

accessible to proteolysis from the exterior of the cell than from the cytoplasm

(Gallagher et al., 1989).

Such evidence contrasts with what has been found for HisP from S.

typhimurium, which behaves neither like a typical integral nor a peripheral

membrane protein when it is in the presence of HisQ and HisM (Kerppola et al.,

1991). Interestingly, HisP prefers to be associated with the membrane even in

the absence of HisQ and HisM, but in such a case it assumes an improper

association, acquiring an unusual level of sensitivity to proteases and

detergents, and in general behaving like a typical peripheral membrane protein.

This observation has led to the suggestion that HisP is deeply embedded in the

membrane (Kerppola et al., 1991; Baichwal et al., 1993). A possible explanation

for the association of HisP with the membrane in the absence of HisQ and HisM

may be found in its predicted tertiary structure, which has been modelled to

include a large loop with some amphipathic characteristics (Mimura et al.,

1991). This loop might allow proper insertion of HisP in the membrane in the

presence, and its improper association in the absence, of the hydrophobic

components. The notion that a segment of the ATP-binding domain protrudes

into or through a pore generated by the transmembrane domains, derives

support from the recent finding that HisP is accessible to proteases and to an

impermeant biotinylating reagent from the exterior surface of the membrane

when the complex has been assembled in the presence of HisQ and HisM, but

not in their absence (Baichwal et al., 1993).

Page 48: The reasonable man attempts to adapt ... - reading.ac.uk

37

While the nature of the interaction between the ATP binding component and

the transmembrane components is uncertain, the interaction must be specific as

the ATP-binding domain from one transporter cannot normally replace that of

another. Indeed, the interaction appears to induce a conformational change in

the ATP-binding domain that alters its biochemical properties (Reyes &

Shuman, 1988; Davidson & Nikaido, 1990). The residues involved in these

interactions are unknown, although gene-fusion studies show that specificity

cannot reside in the N-terminus (Schneider & Walter, 1991).

The two attempts at computer-assisted modelling of ATP-binding

components have generated similar structures by entirely different procedures

(Hyde et al., 1990; Mimura et al., 1991). The core of both models is a nucleotide-

binding pocket which includes five hydrophobic β-sheets and the glycine-rich

Walker motif A, appropriately positioned to interact with ATP and mediate

phosphoryl transfer (Fig. 1.9). Extending from the core nucleotide-binding fold

are loops that have no direct counterpart in adenylate kinase. In one model

(Hyde et al., 1990) these sequences are folded as two separate loops (designated

loops 2 and 3); in the other model (Mimura et al., 1991) they are folded as one

large loop (termed the helical domain). While this difference remains to be

clarified experimentally, it is significant that both models predict that the same

sequences protrude from the nucleotide-binding pocket.

Page 49: The reasonable man attempts to adapt ... - reading.ac.uk

38

Fig. 1.9 Schematic representation of the predicted topology of the ATP-binding component of an ABC transporter (Doige & Ames, 1993). The shaded section corresponds to the Walker motif A. "P-loops" such as this are known to be involved in phosphoryl transfer in many nucleotide-binding proteins. Amino acid residues refer to the HisP sequence. The black box represents a bound mononucleotide.

Page 50: The reasonable man attempts to adapt ... - reading.ac.uk

39

The available data support these model structures. In vitro chemical

modification studies using the photolabelling analogue of ATP, 8-azido ATP are

consistent with the proposed nucleotide-binding fold (Mimura et al., 1990).

Structure function analysis of HisP mutants (Petronilli & Ames, 1991; Shyamala

et al., 1991) showed that the ability of mutant proteins to function in transport

and/or bind ATP was related to the location of the mutation within both the

sequence and the predicted structure. A good correlation was found between

loss of ATP-binding ability (and transport) and location of the mutation within

the predicted ATP-binding pocket, while transport-negative mutants within the

helical domain generally did not affect ATP binding. In addition, a set of

interesting mutations of hisP that enable the mutant HisP proteins to facilitate

transport and ATP hydrolysis even in the total absence of the periplasmic HisJ

protein (Ames & Spudich, 1976; Petronilli & Ames, 1991) also supports the

models because they are all located within the ATP-binding pocket.

Presumably the mutations have changed the binding pocket so that it has

acquired a capacity for unregulated ATP hydrolysis (Petronilli & Ames, 1991;

Shyamala et al., 1991; Speiser & Ames, 1991). Interestingly, although mutations

resulting in unregulated ATP hydrolysis have also been found in the maltose

permease of E. coli (Davidson et al., 1992), these are located in the hydrophobic

domains, indicating that such an effect can arise from a disturbance in the

structure of many regions of the membrane-bound complex.

Each ABC transporter has two ATP-binding domains and both are required

for function. Thus elimination of either one of the two domains of the Opp

oligopeptide transporter of S. typhimurium abolishes function (Hiles et al., 1987).

It is not clear, however, whether the two ATP-binding domains are functionally

equivalent. For transporters such as the histidine permease of S. typhimurium

the two ATP-binding subunits are identical (Kerppola et al., 1991), which

suggests equivalency. The estimated stoichiometry of two ATP molecules

hydrolyzed for each molecule of substrate transported (Mimmack et al., 1989) is

Page 51: The reasonable man attempts to adapt ... - reading.ac.uk

40

also consistent with two equivalent domains, each hydrolyzing one ATP

molecule per transport cycle. However, the observation that genetic analyses of

prokaryotic transport systems in which the ATP-binding components are coded

for by two genes (such as livG/livF in E. coli, braF/braG in Pseudomonas

aeruginosa, and oppD/oppF in both E. coli and S. typhimurium) have usually not

identified the second of the two ATP-binding component genes, has led to the

suggestion (Haney & Oxender, 1992) that most missense mutations and protein

fusions in the second gene are functionally silent. Data for CFTR leads to a

similar conclusion. Many cystic fibrosis mutations fall in the first nucleotide-

binding domain (Cutting et al., 1990; Kerem et al., 1990), yet a mutation in the C-

terminal nucleotide-binding domain that might be expected to disrupt ATP

hydrolysis does not appear to inhibit CFTR function (Anderson et al., 1991a).

In conclusion, while the specific role of each nucleotide-binding domain

remains to be clarified, a working model consistent with the available data

predicts that the ABC proteins have a tightly folded core structure that binds

and hydrolyzes ATP; loops that extend from this core structure interact with the

other components of the transporter and couple, presumably via a

conformational change, the energy of ATP hydroysis to transport.

1.4.4 Periplasmic binding proteins

All known prokaryotic ABC transport systems that facilitate uptake utilize a

periplasmic binding protein. Genetic studies have demonstrated that these

periplasmic proteins are absolutely required for the function of the transport

system with which they are associated. However, they are not integral to the

mechanism of transmembrane translocation itself, since mutants of bacterial

uptake systems can be isolated that function in the absence of the periplasmic

component (Shuman, 1982; Speiser & Ames, 1991).

The periplasmic proteins are relatively easy to purify and many have been

studied in considerable detail. The characterized proteins vary in size from 25

kD for HisJ (Higgins & Ames, 1981) to 59 kD for OppA (Hiles & Higgins, 1986).

Page 52: The reasonable man attempts to adapt ... - reading.ac.uk

41

There is little sequence conservation between binding proteins for different

substrates, the only exceptions being the pairs of periplasmic proteins that

interact with the same core transmembrane complex (e.g. the histidine- and

lysine-arginine-ornithine-binding proteins which deliver substrates to the

HisMQP complex in the membrane of S. typhimurium, and the leucine- and

leucine-isoleucine-valine-binding proteins which deliver substrates to the

LivHMGF complex in the membrane of E. coli). Several binding proteins have

been crystallized and their three dimensional structures determined. These

include the leucine-isoleucine-valine-, leucine-, arabinose-, maltose- and

galactose-binding proteins of E. coli (Adams & Oxender, 1989; Quiocho, 1990;

Vyas et al., 1991). All have a similar structure with two globular domains and a

cleft between that forms the substrate binding site. All substrates appear to be

bound via hydrogen bonds (Quiocho, 1986; Pflugrath & Quiocho, 1988). The

Venus flytrap model in which the proteins undergo a conformational change

upon binding substrate that traps the substrate in the cleft between the two

domains (Mao et al., 1982; Sack et al., 1989), has recently been confirmed for the

lysine-arginine-ornithine-binding protein of S. typhimurium, which has been

purified in both the liganded and unliganded forms (Oh et al., 1993). The

structures determined for the two forms suggest the existence of two states for

the unliganded receptor, open empty and closed empty, that are in dynamic

equilibrium with each other, with the substrate capable of binding to one of the

two globular domains. In the presence of substrate, the closed form is

stabilized by additional interactions between the substrate and the second

globular domain, resulting in a closed liganded form (Oh et al., 1993). In effect,

this model postulates that the act of binding the substrate is responsible for

stabilizing the receptor in its closed conformation.

The periplasmic proteins serve as the initial receptor for transport, delivering

substrate to the membrane-bound components. The in vitro binding

specificities and affinities measured for the purified proteins correspond well

Page 53: The reasonable man attempts to adapt ... - reading.ac.uk

42

with in vivo characteristics of the transport process, which implies that binding

provides the rate-limiting step for transport (Miller et al., 1983). Interaction of

the binding protein-substrate complex with the membrane-associated transport

components has been demonstrated both biochemically and genetically.

Prossnitz et al. (1988) used chemical cross-linking reagents to isolate subunits of

the S. typhimurium histidine transport complex that were cross-linked to each

other in vivo and in vitro. They identified cross-linked products by Western

blot, using antibodies to HisJ and HisQ. Several cross-linked forms were

identified, including one that reacted with antibodies against both HisJ and

HisQ. A previously described mutant of HisJ that is able to bind histidine

normally but does not allow transport (Kustu & Ames, 1974), showed

significantly reduced cross-linking to HisQ relative to the wild-type protein,

supporting the suggestion that the interaction between HisJ and HisQ occurred

during the transport event. That the interaction between binding protein and

transmembrane component is specific was indicated by the finding that MalE

(the binding protein from the maltose transporter of E. coli) could not be cross-

linked to HisQ.

Treptow and Shuman (1988) isolated mutants of malE that restored maltose

uptake in strains of E. coli containing a variety of mutations in malF or malG, the

genes encoding the hydrophobic membrane components of the transporter.

Many of the suppressors were found to be allele specific i.e. a malE mutation

selected in one mutant background was found to be unable to restore transport

in other mutant backgrounds. Mutations mapped widely over the malE gene,

with mutations in the N-terminal end of malE tending to suppress malF alleles

and mutants in the C-terminal end tending to suppress malG alleles. These

results suggest that interaction involves both domains of the periplasmic

binding protein and both hydrophobic transmembrane components.

Interaction of the binding protein with the membrane-associated transport

components does not occur in the absence of bound substrate (Prossnitz et al.,

Page 54: The reasonable man attempts to adapt ... - reading.ac.uk

43

1988; Doige & Ames, 1993), suggesting that the conformational change

undergone by the periplasmic protein upon binding of the substrate enables

interaction with the appropriate complex of membrane proteins.

The precise function of binding proteins in transport is still not fully

understood. Their role has commonly been ascribed to that of increasing the

effective concentration of substrate in the periplasm. However, Hennge and

Boos (1983) have argued that the effect of a high concentration, in the

periplasm, of binding protein with a high affinity for substrate, is to generate a

high concentration of liganded binding protein, not of ligand itself. Estimates

of the concentration of the binding protein in the periplasm vary from 0.5 to

10mM (Silhavy et al., 1975; Hengge & Boos, 1983; Ames, 1986). With

dissociation constants for ligands in the range of 10-5 to 10-7 M, this means that

even when low concentrations of ligand are present in the medium, the

concentration of liganded binding protein in the periplasm can be in the

millimolar range (Hengge & Boos, 1983). The concentration of the free ligand in

the periplasm will be the same as it is outside the periplasm. This by itself

effectively demonstrates that the membrane complex recognizes the liganded

binding protein and not the free ligand, since strains not expressing binding

proteins do not show appreciable transport from the high-affinity system, even

when the ligand is present in millimolar concentrations, whereas wild-type

transport systems show transport affinities in the micromolar range (Hengge &

Boos, 1983; Treptow & Shuman, 1985).

It has been suggested that binding proteins facilitate the movement of

substrate within the periplasm. The periplasm of Gram-negative bacteria is a

gel-like matrix and , at low substrate concentrations, diffusion may be limiting

(Brass et al., 1986). However, the recent identification of binding proteins in

Mycoplasma and other Gram-positive species, which do not have a periplasm

(Dudler et al., 1988; Gilson et al., 1988; Alloing et al., 1990; Perego et al., 1991;

Rudner et al., 1991), argues against this. An adaptation of this hypothesis is that

Page 55: The reasonable man attempts to adapt ... - reading.ac.uk

44

the binding proteins enhance transport by restricting diffusion to two, rather

than three, dimensions. Diffusion of the binding protein-substrate complex in

two dimensions is expected to enhance the efficiency at which substrate is

delivered to the membrane transport complex, compared with three-

dimensional diffusion of the unbound substrate in solution. In Gram-positive

species the binding proteins are anchored to the membrane by a lipid group,

and diffusion is consequently restricted to two dimensions. In Gram-negative

species the dimensions of the periplasm also effectively restrict diffusion of the

binding protein to two dimensions. Indeed, the periplasmic binding protein of

the maltose transporter in E. coli still functions efficiently when anchored to the

membrane via a non-cleavable signal sequence (Fikes & Bassford, 1987).

It is also possible that the periplasmic protein imposes directionality on

transport. No ABC transporter has yet been found that can mediate both

uptake and export of a substrate, yet comparison of the membrane-associated

proteins of uptake and export systems does not identify any feature that allows

the two to be distinguished. Since all known bacterial uptake systems require a

periplasmic component, while no export system does, it is conceivable that the

presence of the binding protein determines the directionality of the transporter.

Consistent with such a suggestion is the fact that interaction of the binding

protein-substrate complex with the membrane-associated domains is required

for ATP hydrolysis (Bishop et al., 1989; Petronilli & Ames, 1991), which implies

an induced conformational change in the membrane-associated components of

the transporter.

1.4.5 Mechanism of solute translocation

The presence of conserved ATP-binding motifs (Walker motifs) in all

characterized ABC proteins indicates a role for ATP in solute translocation by

ABC transporters, and recent studies have provided good evidence that ATP

hydrolysis by the transporters themselves provides the driving force for solute

accumulation by binding protein-dependent systems in Gram-negative bacteria.

Page 56: The reasonable man attempts to adapt ... - reading.ac.uk

45

The OppD and HisP proteins of S. typhimurium, and the MalK protein of E.

coli, have been shown to bind ATP and/or a variety of ATP-affinity analogues

(Hobson et al., 1984; Higgins et al., 1985). The Km of the histidine transporter for

ATP is estimated to be about 200µM, appropriately less than the normal

intracellular ATP concentration (Ames et al., 1989). Furthermore, the ABC

proteins not only bind ATP, but ATP-binding is essential for function. Thus

mutation of the ATP-binding site of several ABC transporters inhibits activity

(Hiles et al., 1987; Joshi et al., 1989; Higgins et al., 1990; Ames & Joshi, 1991;

Shyamala et al., 1991). Additionally, in cell membrane-derived vesicle systems,

histidine and maltose transport by E. coli shows an absolute requirement for

ATP (Dean et al., 1989; Prossnitz et al., 1989).

A requirement for ATP does not necessarily imply that hydrolysis energizes

transport. ATP binding could, potentially, serve a structural or regulatory role.

However, non-hydrolyzable analogues of ATP are unable to support active

transport (Ames et al., 1989), and in vivo studies on several E. coli ABC

transporters demonstrated that ATP hydrolysis is dependent upon, and occurs

concomitantly with, substrate translocation (Mimmack et al., 1989), while no

ATP consumption was observed for non-ABC transporters. Reconstitution of

transporters in vesicles (Ames et al., 1989; Dean et al., 1989; Prossnitz et al., 1989)

and in proteoliposomes (Bishop et al., 1989; Davidson & Nikaido, 1990) also

revealed the complete dependence of transport on ATP hydrolysis, and vice

versa.

There is, therefore, no doubt that ATP hydrolysis is required for transport

and that domains of the transporters themselves can bind ATP. However,

although the simplest interpretation of these facts is that the ABC domains

themselves hydrolyze ATP and directly couple the energy of hydrolysis to the

transport process, this has yet to be formally demonstrated.

Besides ATP, GTP and CTP can also energize histidine transport in vesicles

(Bishop et al., 1989). However, the low cytoplasmic pools of these nucleotides

Page 57: The reasonable man attempts to adapt ... - reading.ac.uk

46

and their poor affinities for the transport proteins (Hobson et al., 1984; Higgins

et al., 1985) make it unlikely that they serve a significant role in energizing

transport in vivo.

The stoichiometry of ATP hydrolysis is not firmly established. In vesicle

systems, stoichiometries of 5 (Bishop et al., 1989) and 1.4 to 17 (Davidson &

Nikaido, 1990) have been reported. These high and variable values may be due

to damage incurred by the complex during reconstitution resulting in the

uncoupling of hydrolysis and transport (Bishop et al., 1989). Estimates using

whole cells suggest a stoichiometry of close to two ATP molecules hydrolysed

per molecule of substrate transported (Mimmack et al., 1989).

While it has been demonstrated that ATP is only hydrolyzed upon

interaction of the periplasmic binding protein with the membrane-bound

complex (Bishop et al., 1989; Petronilli & Ames, 1991), it is not known how the

signal for ATP hydrolysis is transduced. Similarly, the mechanism by which

the energy of ATP hydrolysis is coupled to the transport process is unknown.

There is no evidence that a phosphorylated protein intermediate is involved

(Ames & Nikaido, 1981), and it is generally assumed that ATP binding and

hydrolysis induces a conformational change in the ATP-binding domain, which

is transmitted, via domain-domain interactions, to the transmembrane subunits

that mediate translocation across the membrane. The helical domain (loop 2/3

region) of the ATP-binding components is a good candidate for involvement in

this conformational transduction, particularly since mutations in this loop

region can uncouple ATP hydrolysis from the transport event, without altering

ATP binding or hydrolysis (Petronilli & Ames, 1991).

The nature of the interaction between substrate and transmembrane domains

is also uncertain. At one extreme the transmembrane domains might simply

form a pore in the membrane which the ATP-binding domains serve to open

and close. However, the transmembrane domains do not simply form an open

channel in the absence of the ATP-binding domains (Higgins, 1992).

Page 58: The reasonable man attempts to adapt ... - reading.ac.uk

47

Furthermore, such an arrangement could not account for the directionality, and

ability to concentrate solute against a gradient, of ABC transporters. Classical

kinetic analyses of transporters imply a conformational change exposing

substrate-binding sites to opposing sides of the membrane (Stein, 1990). A

better model may therefore be that substrate interacts with a binding site

located within a pore-like structure generated by the transmembrane domains,

and subsequent ATP hydrolysis reorientates this site to expose it to the opposite

face of the membrane. However, it is equally possible that reorientation of the

binding site is facilitated by substrate-binding and that ATP hydrolysis resets

the orientation in order to impose directionality and the ability to actively

accumulate substrate.

While the great majority of ABC transport systems mediate active transport,

two apparently typical ABC transporters, CFTR and MDR, have recently been

associated with chloride channel activity (Anderson et al., 1991b; Kartner et al.,

1991; Gill et al., 1992; Valverde et al., 1992). This has highlighted the problem of

defining the physical differences between channels and transporters.

Transporters are generally considered to be enzyme-like, interacting

stoichiometrically with their substrate and undergoing defined conformational

changes during each transport cycle. In contrast, channels are viewed as holes

which, when open, allow non-stoichiometric passage of molecules with

appropriate characteristics. This distinction is based on several experimental

observations: (i) Kinetic and biochemical characterization of several

transporters has revealed enzyme-like intermediate states during each transport

cycle, often interpreted as the alternate exposure of a substrate-binding site at

each face of the membrane (Stein, 1990). (ii) Many transporters have been

shown to be stoichiometrically coupled to other events such as ATP hydrolysis

or proton movement. (iii) The turnover number of transporters is restricted by

the enzyme-like conformational changes. In contrast, channels are ultimately

diffusion-limited, and their turnover number can be several orders of

Page 59: The reasonable man attempts to adapt ... - reading.ac.uk

48

magnitude greater than has ever been measured for a transporter. (iv)

Channels simply facilitate equilibration of substrate in response to

concentration and electrochemical gradients. Transporters, in contrast, can

utilize energy to concentrate substrate against a gradient.

In the case of ABC systems the difference between a transporter and a

channel may reside in the organization of the transmembrane domains. At one

extreme , the translocation pathway formed by the transmembrane domains of

ABC transporters and channels may be very similar, with a small structural

change allowing a transporter to function as a channel. At the other extreme,

the transmembrane pathways may be very different. One suggestion (Higgins,

1992) is that transport activity might be associated with the monomeric

membrane-bound complex, while channel function might involve a different

transmembrane translocation pathway generated by the oligomeric association

of complexes (Fig. 1.10).

Fig. 1.10 Schematic representation of transmembrane domains of an ABC transporter viewed from the membrane surface, showing the pathway through the membrane (*) provided by monomeric complexes (A), and a potential alternative pathway (B) provided by oligomerization (Higgins, 1992).

Page 60: The reasonable man attempts to adapt ... - reading.ac.uk

49

1.4.6 The role of binding protein-dependent transporters

In general ABC transporters are low capacity but high affinity (Km 0.01-1µM)

systems and can accumulate substrate against very large concentration

gradients (>10,000-fold). This is in contrast to members of the twelve-

membrane-spanning-helix family of transporters (Section 1.3), which as a

general rule are low affinity but high capacity systems that, for thermodynamic

reasons, cannot accumulate substrate against large concentration gradients

(Hengge & Boos, 1983).

For some substrates, E. coli possesses both an ABC transporter and a twelve-

transmembrane-helix transporter. The reason for this apparent duplication of

effort may be that under certain growth conditions changing energy status

might favour the use of transporters energized by alternate means (twelve-

transmembrane-helix transporters are energized by the electrochemical

gradient or by co- or countertransport). However, it has also been proposed

(Higgins, 1992; Doige & Ames, 1993) that the kinetic properties of twelve-

transmembrane-helix transporters make them suitable for bulk uptake of

carbon and nitrogen sources for growth, while those of ABC transporters

indicate a scavenging role that becomes important in environments where

nutrients are present at extremely low concentrations. In the specific case of

amino acid uptake, it has been suggested that ABC transporters may be

involved in the recapture of biosynthetically produced amino acids that would

otherwise be lost from the cell (Ames, 1972; Antonucci & Oxender, 1986).

1.4.7 Regulation of ABC transporters

As with other bacterial uptake systems, ABC transporters are commonly

regulated at the transcriptional level (Haney & Oxender, 1992), and many are

only expressed in the presence of their specific substrates. However, the

activity of ABC transporters, once expressed, may also be regulated. Uptake of

sugars such as maltose (Nelson & Postma, 1984) is inhibited by direct

Page 61: The reasonable man attempts to adapt ... - reading.ac.uk

50

interaction between a component of glucose metabolism (Enzyme III) and the

ABC transporter, such that glucose is used as the preferred carbon source.

Regulation of transport through alteration of binding protein affinity by

phosphorylation has also been described (Celis, 1984): E. coli was incubated

with 32Pi and an osmotic shock fluid separated by gel chromatography. One

labelled fraction that exhibited binding of arginine was shown to contain

phosphorylated lysine-arginine-ornithine-binding protein. Unphosphorylated

lysine-arginine-ornithine-binding protein was also recovered and was found to

have a 50-fold greater affinity for substrate than the phosphorylated protein

(dissociation constants for arginine of the modified and unmodified proteins

are 5.0µM and 0.1µM, respectively). A mutant has been isolated that does not

express the kinase that phosphorylates the binding protein (Celis, 1990). This

kinase, which has been purified (Celis, 1990), serves to deactivate the binding

protein and limit entry of basic amino acids into the cell.

There is also evidence that the intracellular level of substrate may regulate

uptake. In a study of histidine transport in reconstituted proteoliposomes,

which demonstrates the unidirectional nature of substrate translocation in this

system (Doige & Ames, 1993), it was found that accumulated histidine did not

exit from the proteoliposomes, and that incorporation of ADP, inorganic

phosphate, and histidine inside the proteoliposomes did not result either in the

exit of histidine or the synthesis of ATP. Furthermore, both the internally

accumulated histidine and the hydrolytically produced ADP inhibited

transport, suggesting that the internal pool of histidine (and ADP) regulates

translocation, presumably via a substrate-recognizing site on the cytoplasmic

surface of the membrane-bound complex (Doige & Ames, 1993).

Page 62: The reasonable man attempts to adapt ... - reading.ac.uk

51

1.5 AMINO ACID TRANSPORT IN RHIZOBIUM

Poole et al. (1985) studied the stereospecific and kinetic properties of L-

glutamate transport in R. leguminosarum biovar viciae strain 3841. They found

not only that glutamate uptake is competitively inhibited by a wide range of

other amino acids, but also that intracellular L-leucine can be exchanged out of

strain 3841 by L-glutamate. These results strongly suggest that strain 3841

possesses a general amino acid transport system.

The fact that L-glutamate transport failed to show any substantial deviation

from Michaelis-Menten kinetics suggests that the general permease may be the

only kinetically significant transport system for L-glutamate in strain 3841. The

existence of specific transport systems for glycine and leucine, operating in

addition to the common system in strain 3841, was indicated by the failure of

large excesses of L-glutamate to completely inhibit uptake of these substrates.

Glenn et al. (1991) studied proline transport in cowpea Rhizobium NGR234,

and found that a 5-fold excess of many amino acids (including glutamate)

caused a 70-90% inhibition of L-proline uptake. In addition, [14C]proline taken

up by NGR234 cells over 3 minutes, was found to be exchangeable with

extracellular valine, histidine or isoleucine but not glutamate. This suggests

that proline is carried by a general amino acid transporter. The fact that

intracellular proline is not excreted in the presence of extracellular glutamate

may indicate that glutamate transport is distinct from that of proline, or may

reflect the relative affinities of these substrates for the common carrier.

Glutamate transport in cowpea Rhizobium MNF2030 and R. leguminosarum

biovar trifolii MNF1000 is inhibited 50-100% by the addition of a 10-fold excess

of a variety of other amino acids (Jin et al., 1990). However, kinetic analysis of

glutamate uptake in these strains indicates the presence of two glutamate

transport systems in both cases (a high affinity (low Km) system with a

relatively low capacity, and a low affinity (high Km) system with a greater

capacity).

Page 63: The reasonable man attempts to adapt ... - reading.ac.uk

52

In a recent study (Watson et al., 1993), aspartate transport in R. meliloti was

found to be mediated by at least two systems; the Dct system (apparent Km for

aspartate transport of 10mM) and a second system (apparent Km of 1.5mM)

which was competitively inhibited by glutamate. There was also some kinetic

evidence for the existence of a third, high affinity, system. In dct mutants

growing on aspartate, high initial rates of aspartate uptake preceded the lower

steady-state value. This was taken to be indicative of the presence of an internal

aspartate pool in R. meliloti, and a rapid initial equilibration of aspartate inside

the cell with that outside by an exchange process.

The capacity of the glutamate-aspartate system is sufficient to allow growth

of R. meliloti dct mutants on aspartate as nitrogen source, but not as both carbon

and nitrogen source. Furthermore, this system is not directly induced by

aspartate, suggesting that it may be regulated, at least partly, by nitrogen

requirement. Previously reported experiments with chemostat cultures have

demonstrated that glutamate uptake in R. leguminosarum strain 3841 is

regulated by nitrogen availability (Poole et al., 1987).

Page 64: The reasonable man attempts to adapt ... - reading.ac.uk

53

1.6 REGULATION INVOLVING NTRC

Transcriptional regulation of nitrogen metabolism in enteric bacteria has

been found to involve five genes (Reitzer & Magasanik, 1987; Merrick, 1988):

ntrA (also designated glnF and rpoN), which codes for the alternative sigma

factor, σ54; ntrBC (also designated glnLG), which are located downstream of

glnA (the structural gene for glutamine synthetase) and code for the NtrB (also

called NRII) and NtrC (also called NRI) proteins belonging to the family of

histidine kinase sensors and response regulators (Stock et al., 1989); and finally

glnD and glnB, which code for a uridylyltransferase and PII protein,

respectively. In E. coli, glnB and glnD are expressed independently of the

nitrogen status of the cell and of σ54 (van Heeswijk et al., 1993). The glnA-ntrBC

operon has three promoters; one, ntrBp, located between glnA and ntrB, and

two, glnAp1 and glnAp2, located upstream of glnA. Transcription initiated at

the σ70-dependent promoters glnAp1 and ntrBp serves to maintain a low

intracellular concentration of glutamine synthetase, NtrB and NtrC (Reitzer &

Magasanik, 1986) under conditions of nitrogen-excess. Under nitrogen-

limitation expression of the glnA-ntrBC operon is primarily due to glnAp2.

Initiation of transcription from this promoter requires σ54 and the upstream

binding of the activator protein NtrC.

The cellular nitrogen status, as indicated by the intracellular glutamine:α-

ketoglutarate ratio, is sensed by the uridylyltransferase. It responds to

nitrogen-limitation (low glutamine:α-ketoglutarate ratio) or nitrogen-excess

(high glutamine:α-ketoglutarate ratio) by uridylylating or deuridylylating PII,

respectively (Bueno et al., 1985; Holtel & Merrick, 1989). In its unuridylylated

form, PII interacts with NtrB causing it to be a protein phosphatase (Bueno et al.,

1985; Holtel & Merrick, 1989). However, uridylylation of PII prevents this

interaction allowing NtrB to act as a kinase towards NtrC by transferring

phosphate from ATP, and it is the phosphorylated form of NtrC that acts as an

activator at glnAp2 (Ninfa & Magasanik, 1986; Reitzer & Magasanik, 1987;

Page 65: The reasonable man attempts to adapt ... - reading.ac.uk

54

Holtel & Merrick, 1988; Merrick, 1988). Thus nitrogen-limitation leads to

increased synthesis of glutamine synthetase, NtrB and NtrC.

Glutamine synthetase is the key enzyme in the high-affinity, glutamine

synthetase/glutamate 2-oxoglutarate amino transferase, pathway for ammonia

assimilation, and the same cascade that controls the activity of NtrB as a kinase,

also controls post-translational regulation of glutamine synthetase. Glutamine

synthetase can be converted into a less active form by adenylylation. The

adenylyltransferase (product of the glnE gene) which catalyzes both the

adenylylation and deadenylylation of glutamine synthetase is controlled by PII.

Unuridylylated PII (nitrogen-excess conditions) interacts with the

adenylyltransferase enabling it to catalyze the adenylylation of glutamine

synthetase, while uridylylated PII (nitrogen-limited conditions) causes the

adenylyltransferase to catalyze deadenylylation. Thus the post-translational

regulation of glutamine synthetase mirrors its transcriptional regulation.

In addition to acting as a positive regulator, NtrC can also act as a negative

regulator. Indeed, both glnAp1 and ntrBp are negatively controlled by NtrC. At

ntrBp an NtrC binding site overlaps the transcriptional start and the -10 region

of the promoter, so that binding of NtrC impedes RNA polymerase binding and

ntrBC expression (Reitzer & Magasanik, 1983; Dixon, 1984; Ueno-Nishio et al.,

1984; MacFarlane & Merrick, 1985). In Klebsiella pneumoniae, mutations of NtrB

that result in constitutive phosphorylation of NtrC, are found also to cause

strong repression of ntrBp. This repression is independent of nitrogen status,

and is greater than the repression observed in an ntrB deletion strain

(MacFarlane & Merrick, 1987), suggesting that at ntrBp, phosphorylated NtrC

binds more effectively than the unphosphorylated form.

At glnAp1 in E. coli and S. typhimurium two NtrC binding sites with potential

for inhibiting expression from this promoter have been identified (Ames &

Nikaido, 1985; Hirschman et al., 1985; Reitzer & Magasanik, 1985). Site 1

overlaps the -35 region of the promoter, while site 2 overlaps the transcriptional

Page 66: The reasonable man attempts to adapt ... - reading.ac.uk

55

start site. In K. pneumoniae only site 2 is present (Dixon, 1984; Hawkes et al.,

1985).

In addition to controlling the level of glutamine synthetase, the Ntr system

also regulates the expression of other genes related to the nitrogen status of the

cell, including genes involved in amino acid uptake (Merrick, 1988).

In S. typhimurium uptake of histidine, arginine, lysine and ornithine is

mediated by a single membrane associated complex composed of the proteins

HisQ, HisM and HisP in a 1:1:2 ratio (Section 1.4.1). This complex interacts with

two periplasmic binding proteins; one, HisJ is specific for histidine, while the

other, ArgT, binds arginine, lysine and ornithine (Higgins et al., 1982b). The

argT gene lies directly upstream of the hisJQMP operon, and is transcribed,

under the control of the promoter argTr, as a monocistronic unit. Transcription

of hisJQMP is controlled by the promoter dhuA, which is located between argT

and hisJ. The observation that transcription initiated by dhuA is elevated in an

ntrB mutant (Higgins & Ames, 1982; Stern et al., 1984), has led to the erroneous

suggestion that dhuA is negatively regulated by NtrC (Haney & Oxender, 1992).

In fact, studies of lacZ fusions to argTr and dhuA in ntrA- and ntrC- strains have

shown both these promoters to be σ54-dependent, and activated by NtrC under

nitrogen-limitation (Ames & Nikaido, 1985). Mutational analysis has enabled

the site of action of NtrC at argTr to be described in detail (Schmitz et al., 1988).

Uptake of glutamine, glutamate and aspartate in S. typhimurium is also under

nitrogen control, and expression of genes encoding transporters of these amino

acids is likely to be regulated by the Ntr system (Kustu et al., 1979).

Glutamine transport in E. coli is positively regulated by NtrC, with nitrogen-

limitation resulting in increased expression of the glnHPQ operon (Nohno et al.,

1986; Claverie-Martin & Magasanik, 1991).

The pattern of Ntr control in Rhizobiaceae differs from that found in enteric

bacteria, with glnA and ntrBC being located in separate operons. In R.

leguminosarum biovar phaseoli, ntrBC are found in the operon ORF1-ntrBC which

Page 67: The reasonable man attempts to adapt ... - reading.ac.uk

56

is transcribed from a σ70-dependent promoter independently of nitrogen status

(Patriarca et al., 1993; Amar et al., 1994). The function of ORF1 is currently

unknown. The increase in transcription of ntrC observed in an ntrC mutant has

been taken to indicate that this operon is subject to negative autoregulation by

NtrC, and a putative repressor-binding site for NtrC has been identified in the

promoter region for the operon (Patriarca et al., 1993). These findings indicate

that, unlike the situation in enteric bacteria, changes in the level of NtrC

phosphorylation, and not its intracellular concentration, are sufficient to

activate or repress transcription from its target promoter(s) in R. leguminosarum

biovar phaseoli.

At least two glutamine synthetase isozymes, GSI and GSII, encoded by the

glnA and glnII genes, respectively, are present in Rhizobiaceae (Darrow & Knots,

1977; Fuchs & Keister, 1980; Filser et al., 1986; Colonna-Romano et al., 1987;

Patriarca et al., 1992), with a third isozyme GSIII, encoded by glnT, having been

identified in R. leguminosarum (Chiurazzi et al., 1992) and R. meliloti (de Bruijn et

al., 1989). GSI is similar to the glutamine synthetase of enteric bacteria and is

subject to adenylylation. Transcription of glnA, as part of the glnBA operon, can

be initiated by two promoters, one upstream of glnB and one located between

glnB and glnA (Chiurazzi & Iaccarino, 1990). Although in free living cells, glnA

is mostly transcribed from the promoter upstream of glnB, and the activity of

this promoter is positively regulated by NtrC (Chiurazzi & Iaccarino, 1990),

synthesis of GSI is found not to be greatly affected by nitrogen supply (Carlson

et al., 1987; Szeto et al., 1987; Rossi et al., 1989; Chiurazzi & Iaccarino, 1990; Amar

et al., 1994). In contrast, synthesis of the eukaryote-like GSII is fully dependent

on positive control by NtrC in response to nitrogen source availability, with

glnII being transcribed as a monocistronic unit (Carlson et al., 1987; Martin et al.,

1988; de Bruijn et al., 1989; Rossi et al., 1989; Shatters et al., 1989; Patriarca et al.,

1992). The increase in transcription of glnII and glnB observed in a glnB mutant

of R. leguminosarum biovar phaseoli, has been shown not to result from an

Page 68: The reasonable man attempts to adapt ... - reading.ac.uk

57

increase in the level of NtrC, and is consistent with the suggestion that changes

in the degree of phosphorylation of NtrC are solely responsible for the

regulation of NtrC-controlled promoters in this species (Amar et al., 1994).

It has been suggested that the different patterns of regulation observed for

ntrBC, glnBA and glnII may be the result of differences in the affinity of NtrC for

the upstream activator sequences of glnBA and glnII, and the putative

repressor-binding site for ORF1-ntrBC (Patriarca et al., 1993). Alternatively, it is

also possible that additional, as yet unidentified, regulatory factor(s) could play

a role in the transcriptional regulation of some or all of these genes.

Page 69: The reasonable man attempts to adapt ... - reading.ac.uk

58

CHAPTER 2 MATERIALS AND METHODS

Page 70: The reasonable man attempts to adapt ... - reading.ac.uk

59

2.1.1 Bacterial strains

The bacterial strains, plasmids and bacteriophages used in this work are

listed in Table 2.1.

Table 2.1 Bacterial strains, plasmids and bacteriophages used

Strain, phage or plasmid

Description Source or Reference

Bacterium R. leguminosarum 3841 StrR derivative of strain 300 Johnston and

Beringer (1975) RU116 Strain 3841 sucD::Tn5 This work RU117 Aspartate toxic escape, putative

Tn5 mutant of 3841 This work

RU118 Aspartate toxic escape Tn5 mutant of 3841

This work

RU126 Aspartate toxic escape, putative Tn5 mutant of 3841

This work

RU137 Strain 3841 phbC::Tn5 This work RU140 Aspartate toxic escape, putative

Tn5 mutant of 3841 This work

RU151 Aspartate toxic escape, putative Tn5 mutant of 3841

This work

RU154 Aspartate toxic escape, putative Tn5 mutant of 3841

This work

RU156 Strain 3841 sucA::Tn5 This work RU158 Aspartate toxic escape, putative

Tn5 mutant of 3841 This work

RU216 3841 containing genomic Tn5(KmR) insert introduced by transduction from strain RU116

This work

RU237 3841 containing genomic Tn5(KmR) insert introduced by transduction from strain RU137

This work

RU256 3841 containing genomic Tn5(KmR) insert introduced by transduction from strain RU156

This work

RU368 Strain 3841 containing pMP220 This work RU438 Strain 3841 containing pRU3024 This work RU439 Strain 3841 containing pRU135 This work RU441 Strain 3841 containing pRU3026 This work RU442 Strain 3841 containing pRU3027 This work RU443 Strain 3841 containing pRU3028 This work

Page 71: The reasonable man attempts to adapt ... - reading.ac.uk

60

Strain, phage or plasmid

Description Source or Reference

RU444 Strain RU116 containing

pRU3004 This work

RU449 Strain RU137 containing pRU3004

This work

RU453 Strain RU156 containing pRU3004

This work

RU500 Strain 3841 containing pRU3029 This work RU502 Strain 3841 containing pRU3031 This work RU503 Strain 3841 containing pRU3032 This work RU504 Strain 3841 containing pRU3033 This work RU506 Strain 3841 containing pRU3035 This work RU510 Strain 3841 containing pRU3039 This work RU515 Strain 3841 containing pRU3044 This work RU517 Strain 3841 containing pRU3046 This work RU519 Strain 3841 containing pRU3048 This work RU522 Strain 3841 containing pRU3049 This work RU526 Strain 3841 containing pRU3053 This work RU541 pRU3026 homogenote in strain

3841 This work

RU542 pRU3027 homogenote in strain 3841

This work

RU543 pRU3028 homogenote in strain 3841

This work

RU622 Strain 3841 containing pAR36A This work RU631 pRU3029 homogenote in strain

3841 This work

RU632 pRU3031 homogenote in strain 3841

This work

RU633 pRU3032 homogenote in strain 3841

This work

RU634 pRU3035 homogenote in strain 3841

This work

RU635 pRU3039 homogenote in strain 3841

This work

RU636 pRU3046 homogenote in strain 3841

This work

RU637 pRU3048 homogenote in strain 3841

This work

RU638 pRU3049 homogenote in strain 3841

This work

RU639 pRU3053 homogenote in strain 3841

This work

RU640 Strain 3841 containing pRU191 This work

Page 72: The reasonable man attempts to adapt ... - reading.ac.uk

61

RU641 Strain 3841 containing pRU192 This work

Page 73: The reasonable man attempts to adapt ... - reading.ac.uk

62

Strain, phage or plasmid

Description Source or Reference

RU724 pRU3067 homogenote in strain

3841 This work

RU725 pRU3059 homogenote in strain 3841

This work

RU726 pRU3069 homogenote in strain 3841

This work

RU733 pRU3061 homogenote in strain 3841

This work

RU735 Strain RU542 containing pRU191 This work RU736 Strain RU542 containing pRU307 This work RU737 Strain RU542 containing pRU308 This work RU738 Strain RU542 containing pRU309 This work RU739 Strain RU542 containing pRU310 This work RU740 Strain RU542 containing pRU311 This work RU741 Strain RU542 containing pRU312 This work RU742 Strain RU542 containing pRU313 This work RU743 Strain RU634 containing pRU191 This work RU744 Strain RU634 containing pRU307 This work RU745 Strain RU634 containing pRU308 This work RU746 Strain RU634 containing pRU309 This work RU747 Strain RU634 containing pRU310 This work RU748 Strain RU634 containing pRU311 This work RU749 Strain RU634 containing pRU312 This work RU750 Strain RU634 containing pRU313 This work RU751 Strain RU636 containing pRU191 This work RU752 Strain RU636 containing pRU308 This work RU753 Strain RU636 containing pRU309 This work RU754 Strain RU636 containing pRU310 This work RU755 Strain RU636 containing pRU312 This work RU756 Strain RU636 containing pRU313 This work RU757 Strain RU543 containing pRU191 This work RU758 Strain RU543 containing pRU309 This work RU759 Strain RU543 containing pRU310 This work RU760 Strain RU543 containing pRU313 This work RU889 Strain RU116 containing

pRU3028 This work

RU897 Strain RU156 containing pRU3028

This work

RU891 Strain RU116 containing pAR36A

This work

RU899 Strain RU156 containing pAR36A

This work

RU913 Strain 3841 containing pRK415-1 This work

Page 74: The reasonable man attempts to adapt ... - reading.ac.uk

63

RU914 Strain 3841 containing pIJ1891 This work

Page 75: The reasonable man attempts to adapt ... - reading.ac.uk

64

Strain, phage or plasmid

Description Source or Reference

RU915 Strain 3841 containing pRU309 This work RU916 Strain 3841 containing pRU310 This work RU917 Strain 3841 containing pRU313 This work RU918 Strain 3841 containing pRU388 This work RU919 Strain 3841 containing pRU389 This work RU929 Strain 3841 ntrC::Ω Reid (1995) RU972 Strain 3841 containing pRU3080 This work RU974 Strain 3841 containing pRU3082 This work RU975 Strain 3841 containing pRU3083 This work RU976 Strain 3841 containing pRU3084 This work RU977 Strain 3841 containing pRU3085 This work RU978 Strain 3841 containing pRU3086 This work RU980 Strain RU929 containing

pRU3028 This work

RU981 Strain RU929 containing pRU3031

This work

RU982 Strain RU929 containing pRU3033

This work

RU983 Strain RU929 containing pRU3035

This work

RU984 Strain RU929 containing pRU3046

This work

RU986 Strain RU929 containing pRU3082

This work

RU988 Strain RU929 containing pRU3086

This work

RU990 Strain 3841 containing pRU393 This work RU999 pRU3082 homogenote in strain

3841 This work

RU1000 pRU3084 homogenote in strain 3841

This work

RU1001 pRU3086 homogenote in strain 3841

This work

RU1002 Strain RU116 containing pMP220

This work

RU1003 Strain RU156 containing pMP220

This work

RU1013 Strain RU929 containing pRU3024

This work

RU1017 pRU3028 homogenote in strain RU929

This work

RU1018 pRU3035 homogenote in strain RU929

This work

Page 76: The reasonable man attempts to adapt ... - reading.ac.uk

65

Strain, phage or plasmid

Description Source or Reference

RU1019 pRU3046 homogenote in strain

RU929 This work

RU1024 Strain RU116 containing pRU3024

This work

RU1025 Strain RU156 containing pRU3024

This work

RU1027 pRU3033 homogenote in strain RU929

This work

RU1029 pRU3082 homogenote in strain RU929

This work

RU1030 pRU3086 homogenote in strain RU929

This work

RU1069 Strain 3841 containing pRU3089 This work RU1070 Strain 3841 containing pRU3090 This work RU1071 Strain 3841 containing pRU3091 This work E. coli 803 met- gal-. Wood (1966) DH5α supE44 ∆lacU169 (φ80 lacZ∆M15)

hsdR17 recA1 endA1 gyrA96 thi-1 relA1.

Hanahan (1983); Bethesda Research Laboratories (1986)

JC5412 Low glutamate uptake: does not grow on glutamate as sole carbon and nitrogen source

Willetts and Mount (1969)

MC1061 hsdR mcrB araD139 ∆(araABC-leu)7679 ∆lacX74 galU galK rpsL thi

Meissner et al. (1987)

S17-1 pro hsdR recA [RP4-2(Tc::Mu) (Km::Tn7)]; RP4 integrated into its chromosome.

Simon et al. (1983)

RU1050 Strain JC5412 containing pRU310 This work Plasmid pAR36A R. leguminosarum glnII::lacZ

translational fusion in pMP220 Patriarca et al. (1992)

pBC KS+ Phagemid, pUC19 derivative, f1 origin of replication, ColE1 replicon; CmR

Stratagene Research Laboratories.

pBluescript SK+ Phagemid, pUC19 derivative, f1 origin of replication, ColE1 replicon; AmpR

Stratagene Research Laboratories.

pIJ1891 pLAFR3 containing the pUC118 polylinker; TcR

A. Downie

Page 77: The reasonable man attempts to adapt ... - reading.ac.uk

66

Strain, phage or plasmid

Description Source or Reference

pJQ200 pACYC derivative, P15A origin

of replication; GmR Quandt and Hynes (1993)

pLAFR1 Wide host range P-group cloning vector, mobilizable RK2 cosmids; TcR

Friedman et al. (1982)

pMP220 IncP transcriptional fusion vector; TcR

Spaink et al. (1987)

pPHJI1 P-group chaser plasmid; GmR Hirsch and Beringer (1984)

pRK2013 ColE1 replicon with RK2 tra genes; helper plasmid used for mobilizing P- and Q-group plasmids; NmR KmR.

Figurski and Helinski (1979)

pRK415-1 Broad host range P-group cloning vector; TcR

Keen et al. (1988)

pSP72 Promoterless cloning vector; AmpR

Promega

pSUP202-1::Tn5 mob; KmR Simon et al. (1983) pRU32 pBluescript SK+ carrying 9.6kb

Tn5-bearing EcoRI fragment from strain RU116

This work

pRU34 pBluescript SK+ carrying 10.2kb Tn5-bearing EcoRI fragment from strain RU156

This work

pRU36 Sub-clone of pRU32 generated by BamHI digestion followed by ligation

This work

pRU37 Sub-clone of pRU32 generated by HindIII digestion followed by ligation

This work

pRU40 Sub-clone of pRU34 generated by BamHI digestion followed by ligation

This work

pRU41 Sub-clone of pRU34 generated by HindIII digestion followed by ligation

This work

pRU99 pBluescript SK+ carrying 8.9kb Tn5-bearing EcoRI fragment from strain RU137

This work

pRU100 Sub-clone of pRU99 generated by HindIII digestion followed by ligation

This work

Page 78: The reasonable man attempts to adapt ... - reading.ac.uk

67

pRU101 pBC KS+ carrying 3.2kb KmR BamHI fragment from pRU99

This work

Page 79: The reasonable man attempts to adapt ... - reading.ac.uk

68

Strain, phage or plasmid

Description Source or Reference

pRU133 pRK415-1 carrying 3.2kb EcoRI

fragment from pRU3024

pRU135 pRK415-1 carrying 10.2kb HindIII fragment from pRU3024

This work

pRU149 pRK415-1 carrying 6.0kb HindIII-SstI fragment from pRU135

This work

pRU165 Sub-clone of pRU135 generated by XbaI-MluI digestion followed by blunted-end ligation.

This work

pRU180 pRK415-1 carrying 4.6kb SstI fragment from pRU3004

This work

pRU181 pRK415-1 carrying 2.5kb SstI fragment from pRU3004

This work

pRU182 pRK415-1 carrying 6.3kb PstI fragment from pRU3004

This work

pRU185 Sub-clone of pRU135 generated by XbaI-ClaI digestion followed by blunted-end ligation.

This work

pRU186 pRK415-1 carrying 5.2kb NcoI fragment from pRU135 cloned in BamHI site

This work

pRU189 pBluescript SK+ carrying 5.4kb MluI-ClaI fragment from pRU135, cloned in EcoRV site

This work

pRU190 pBluescript SK+ carrying 5.4kb MluI-ClaI fragment from pRU135, cloned in EcoRV site in opposite orientation to pRU189

This work

pRU191 pRK415-1 carrying 5.4kb XbaI-HindIII fragment from pRU189

This work

pRU192 pRK415-1 carrying 5.4kb XbaI-HindIII fragment from pRU190

This work

pRU194 pJQ200 carrying 10.7kb Tn5-lacZ-bearing salI fragment from pRU3059

This work

pRU222-247 Ordered BstUI deletions of pRU189

This work

pRU248-275 Ordered BstUI deletions of pRU190

This work

pRU276 pRK415-1 carrying 4.4kb EcoRI fragment from pRU3004

This work

Page 80: The reasonable man attempts to adapt ... - reading.ac.uk

69

pRU277 pRK415-1 carrying 3.8kb EcoRI fragment from pRU3004

This work

Page 81: The reasonable man attempts to adapt ... - reading.ac.uk

70

Strain, phage or plasmid

Description Source or Reference

pRU278 pRK415-1 carrying 3.8kb EcoRI

fragment from pRU3004 cloned in opposite orientation to pRU278

This work

pRU307 pRK415-1 carrying 2.0kb XbaI-HindIII fragment from pRU234

This work

pRU308 pRK415-1 carrying 3.1kb XbaI-HindIII fragment from pRU239

This work

pRU309 pRK415-1 carrying 4.4kb XbaI-HindIII fragment from pRU246

This work

pRU310 pIJ1891 carrying 5.4kb XbaI-HindIII fragment from pRU189

This work

pRU311 pIJ1891 carrying 2.0kb XbaI-HindIII fragment from pRU234

This work

pRU312 pIJ1891 carrying 3.1kb XbaI-HindIII fragment from pRU239

This work

pRU313 pIJ1891 carrying 4.4kb XbaI-HindIII fragment from pRU246

This work

pRU383 pBluescript SK+ carrying 8.8kb BamHI Tn5-lacZ-bearing fragment from pRU3031

This work

pRU384 pBluescript SK+ carrying 8.4kb BamHI Tn5-lacZ-bearing fragment from pRU3033

This work

pRU386 pIJ1891 carrying 5.0kb lacZ-bearing PstI fragment from pRU3068

This work

pRU387 pIJ1891 carrying 5.0kb lacZ-bearing PstI fragment from pRU3068 cloned in opposite orientation to pRU386

This work

pRU388 pRK415-1 carrying 1.8kb KpnI-BamHI fragment from pRU256

This work

pRU389 pIJ1891 carrying 1.8kb KpnI-BamHI fragment from pRU256

This work

pRU393 pMP220 carrying 1.0kb EcoRI-PstI fragment from pRU189

This work

pRU394 pSP72 carrying 11.8kb Tn5-lacZ-bearing salI fragment from pRU3061

This work

pRU395 pSP72 carrying 8.9kb Tn5-lacZ-bearing salI fragment from pRU3069

This work

Page 82: The reasonable man attempts to adapt ... - reading.ac.uk

71

Strain, phage or plasmid

Description Source or Reference

pRU396 pSP72 carrying 10.2kb Tn5-lacZ-

bearing salI fragment from pRU3070

This work

pRU397 pRK415-1 carrying 6.3kb PstI fragment from pRU3004 cloned in opposite orientation to pRU182

This work

pRU398 pRK415-1 carrying 4.4kb EcoRI fragment from pRU3004 cloned in opposite orientation to pRU276

This work

pRU3004 pLAFR1 cosmid containing mdh-sucCDAB from strain 3841

This work

pRU3024 pLAFR1 cosmid containing aapJQMP from strain 3841

This work

pRU3026 pRU3024 aapQ::Tn5-lacZ This work pRU3027 pRU3024 aapP::Tn5-lacZ This work pRU3028 pRU3024 aapJ::Tn5-lacZ This work pRU3029 pRU3024 aapQ::Tn5-lacZ This work pRU3031 pRU3024 cysE::Tn5-lacZ This work pRU3032 pRU3024 aapM::Tn5-lacZ This work pRU3033 pRU3024 cysE::Tn5-lacZ This work pRU3035 pRU3024 aapM::Tn5-lacZ This work pRU3039 pRU3024 aapQ::Tn5-lacZ This work pRU3044 pRU3024::Tn5-lacZ This work pRU3046 pRU3024 aapQ::Tn5-lacZ This work pRU3048 pRU3024 aapJ::Tn5-lacZ This work pRU3049 pRU3024::Tn5-lacZ This work pRU3053 pRU3024 metC::Tn5-lacZ This work pRU3059 pRU3004 sucC::Tn5-lacZ This work pRU3061 pRU3004 sucA::Tn5-lacZ This work pRU3067 pRU3004 sucA::Tn5-lacZ This work pRU3068 pRU3004 sucA::Tn5-lacZ This work pRU3069 pRU3004 sucB::Tn5-lacZ This work pRU3070 pRU3004 mdh::Tn5-lacZ This work pRU3075 pRU3004 sucA::Tn5-lacZ This work pRU3076 pRU3004 mdh::Tn5-lacZ This work pRU3080 pRU3024::Tn5-lacZ This work pRU3082 pRU3024 metC::Tn5-lacZ This work pRU3083 pRU3024::Tn5-lacZ This work pRU3084 pRU3024 metC::Tn5-lacZ This work pRU3085 pRU3024::Tn5-lacZ This work pRU3086 pRU3024 metC::Tn5-lacZ This work

Page 83: The reasonable man attempts to adapt ... - reading.ac.uk

72

pRU3089 pRU3024::Tn5-lacZ This work

Page 84: The reasonable man attempts to adapt ... - reading.ac.uk

73

Strain, phage or plasmid

Description Source or Reference

pRU3090 pRU3024::Tn5-lacZ This work pRU3091 pRU3024::Tn5-lacZ This work Bacteriophage λ::Tn5-lacZ λ carrying the Tn5-B20

transposon Simon et al. (1989)

RL38 Generalized transducing phage of R. leguminosarum

Buchanan-Wollaston (1979)

2.1.2 Culture conditions

R. leguminosarum strains were grown at 28°C on either TY (Beringer, 1974) or

acid minimal salts (AMS) medium derived from that of Brown and Dilworth

(1975), the changes being; potassium phosphate (0.5mM), MgSO4 (2mM) and

buffering provided by MOPS (20mM) pH 7.0. All carbon and nitrogen sources

were at 10 mM unless otherwise stated. Y medium used in transductions was

prepared as previously described by Sherwood (1970). Antibiotics were used at

the following concentrations in µg ml-1 unless otherwise stated; gentamicin 20,

kanamycin 40, spectinomycin 100 streptomycin 500, and tetracycline 2 (in

AMS), 5 (in TY).

E. coli strains were grown at 37°C on LB, with antibiotic concentrations in µg

ml-1 as follows; ampicillin 50, chloramphenicol 10 gentamicin 5, kanamycin 25,

and tetracycline 10.

2.1.3 DNA and genetic manipulations

All routine DNA analysis was performed essentially according to Sambrook

et al. (1989). Southern transfer of DNA to positively charged nylon membrane

(Boehringer Mannheim) and hybridisation were done using an Amersham ECL

kit according to the manufacturers instructions. Conjugations were performed

using either Escherichia coli strain S17-1 as the donor strain according to Simon et

Page 85: The reasonable man attempts to adapt ... - reading.ac.uk

74

al. (1983), or as triparental matings according to Figurski et al. (1979) with either

E. coli strain 803 or DH5α as the donor, and strain 803 containing pRK2013

providing the transfer functions. Transductions were performed according to

Buchanan-Wollaston (1979) using the phage RL38. Transductants were selected

for on TY agar containing kanamycin (80 µg ml-1). DNA sequencing was

performed by the cycle sequencing method using a Promega fmol kit according

to the manufacturers instructions. Nucleotide sequences of non-routine

sequencing primers used are as follows:

P0: GTTCAGGACGCTACTTG

P15: GGATCCATAATTTTTTCCTCC

P16: AAGATAAGACAACGGAAAAGG

P21: ATGGGTCAGGCGGGTGTTG

P22: GTCGCAAATGTCACTATGG

Computer-assisted sequence analysis was performed using GCG software.

2.1.4 Mutagenesis

Transposon mutagenesis was carried out on R. leguminosarum bv. viciae 3841

with Tn5 using the suicide vector pSUP202-1 as described (Simon et al., 1983).

Mutations in pRU3024 and pRU3004 were produced by first transforming the

cosmid into E. coli strain MC1061. Transformants were mutagenised with Tn5-

lacZ by using λ containing the transposon derivative B20 essentially as

described (Simon et al., 1989). Kanamycin resistant colonies were pooled, and

the cosmids isolated by the alkaline lysis technique. The cosmids were

transformed into E. coli strain DH5α and kanamycin resistant colonies purified.

Cosmid DNA was isolated from each purified strain and the location and

orientation of transposons determined by restriction mapping.

To create chromosomal mutations, mutated cosmids were conjugated into R.

leguminosarum strain 3841. After purification, the incompatible plasmid pPHJI1

was conjugated into each strain and the homogenotes isolated by the technique

of Ruvkun and Ausubel (1981).

Page 86: The reasonable man attempts to adapt ... - reading.ac.uk

75

2.1.5 Transport Assays

For R. leguminosarum strains, cells were prepared and transport assays

performed as previously described (Poole et al., 1985), using in each case, a total

substrate concentration of 25µM. In exchange experiments, cell suspensions

were incubated in 50µM labelled AIB at 28°C for between 2 and 20 min prior to

the addition of unlabelled AIB or glutamate to a concentration of 4mM, CCCP

to a concentration of 50µM, or KCN to a concentration of 10mM. Subsequent

0.1ml samples for Millipore filtration and counting, were taken at intervals of 1,

2, 5 or 10 min for up to 50 min.

For E. coli strains, cells grown on LB to A660 0.5-0.7 were harvested, washed

twice in 50mM potassium phosphate, pH6.9, containing 0.5mM MgSO4 (buffer

A) and resuspended to a final A660 of approximately 10 in the same buffer. This

cell suspension was stored at 37°C with shaking at 250 rpm. Uptake was

assayed at 37°C, after 10-fold dilution of the cells into buffer A containing

10mM glucose, and incubation for 1 min. Uptake was initiated by addition of

labelled substrate to a final concentration of 25µM. Samples were removed at 1

min intervals for 2 min, and cells collected by Millipore filtration under

vacuum. Filters were washed rapidly with 2 x 3ml ice-cold buffer A, and the

radioactivity counted in Beckman "Ready Safe" scintillation fluid.

The specific activities of labelled substrates in the assays were: L-[U-

14C]aspartic acid (354 MBq mmol-1), L-[U-14C]glutamic acid (357 MBq mmol-1),

L-[U-14C]alanine (348 MBq mmol-1), L-[U-14C]histidine (359 MBq mmol-1), L-[U-

14C]leucine (359 MBq mmol-1), L-[35S]methionine (370 MBq mmol-1) and D-[U-

14C]glucose (358 MBq mmol-1).

Incorporation experiments were performed as previously described (Poole et

al., 1985).

2.1.6 Isolation of periplasmic fractions and protein gel electrophoresis

R. leguminosarum strains were grown in AMS containing 50µM potassium

phosphate and periplasmic proteins released by lysosyme/EDTA treatment as

Page 87: The reasonable man attempts to adapt ... - reading.ac.uk

76

described by Glenn and Dilworth (1979). Crude periplasmic fractions were

concentrated on microconcentrators (Microcon-10, Amicon) as necessary.

Samples were subjected to SDS-PAGE as previously described (Laemmli, 1970),

using a 12.5% gel.

Cell fractions for use in assays of marker enzymes were prepared as follows:

After the removal of periplasmic proteins, cells were resuspended in 30mM

TRIS/HCl buffer pH8.0 containing 20%(w/v) sucrose and 1mM DTT, and

disrupted by two passages through a French press at 69000 kPa. Following

centrifugation at 30000 g for 20 min, the supernatant was used for enzyme

assay.

2.1.7 Protein binding assays

R. leguminosarum strains were grown and periplasmic proteins isolated as for

SDS-PAGE. Each crude extract was dialysed overnight in 3 x 3l of 5mM

HEPES, pH7.2 before being concentrated to approximately 1 mg ml-1 on an

Ultrafree-20 (Millipore) filter (10000 MWCO). Substrate-binding by the

concentrated extracts was assayed by three techniques.

Ammonium sulphate precipitation was performed essentially as described

by Richarme and Kepes (1983). Periplasmic extract (100 µg protein) was

incubated in 10 µM labelled substrate at 28°C for 10 min. Proteins were

precipitated by the rapid addition of 10 volumes of ice-cold saturated

ammonium sulphate, and collected by Millipore filtration under vacuum.

Filters were washed with 2 x 3ml saturated ammonium sulphate, and counted

in Beckman "Ready Safe" scintillation fluid.

Detection of binding activity by direct polyacrylamide gel electrophoresis of

the ligand-protein complex in non-denaturing conditions was carried out as

described by Le Rudulier et al. (1991). Periplasmic protein extract (20 µg) was

incubated in 10 µM labelled substrate in a total volume of 25 µl, for 30 min at 28

°C. Laemmli's sample buffer (Laemmli, 1970) without SDS and β-

mercaptoethanol was added, and the samples subjected to PAGE using a 12.5%

Page 88: The reasonable man attempts to adapt ... - reading.ac.uk

77

gel and omitting SDS. The analyses were performed with a constant voltage of

200V for approximately 55 min. The gels were quickly dried on Whatman 3MM

paper and exposed to X-ray film for 10 days.

In the third technique, 420µl of protein extract was incubated in 7µM labelled

amino acid in a total assay volume of 450µl, for 30 min at 28°C. Following

incubation, 50µl of the assay was removed and scintillation counted. The

remaining 400µl was spun through a Microcon-10 (Amicon) microconcentrator

at 5°C and the radioactivity in 50µl of the filtrate counted. The amount of

substrate bound to protein was calculated by comparison of the amount of

radioactivity present in the filtrate to that in the incubation mixture prior to

spinning. Controls containing no protein were performed to determine any

background binding of substrate by the filter in the microconcentrator.

Specific activities of substrates in the assays were: L-[U-14C]aspartic acid (676

MBq mmol-1), L-[U-14C]glutamic acid (688 MBq mmol-1), L-[U-14C]alanine (655

MBq mmol-1), L-[U-14C]histidine (695 MBq mmol-1), L-[U-14C]leucine (695 MBq

mmol-1), and L-[35S]methionine (3.01 GBq mmol-1)).

2.1.8 Enzyme assays

Cultures of R. leguminosarum strains were harvested at a cell density of

approximately 5 x 108 cells ml-1, washed and resuspended in 40 mM HEPES,

pH7.0, containing 1 mM DTT. Cells were disrupted by two passages through a

French press at 69000 kPa. Following centrifugation at 30000 g for 20 min, the

supernatant was used for enzyme assay. Citrate synthase, α-ketoglutarate

dehydrogenase, isocitrate dehydrogenase and succinyl-CoA synthetase were

assayed according to Reeves et al. (1971). Malate dehydrogenase was assayed

by the technique of Saroso et al. (1986).

β-Galactosidase fusions were assayed according to Miller (1972) with the

modifications described by Poole et al. (1994a). Alkaline phosphatase activity

was measured as described by de Maagd and Lugtenberg (1986).

Page 89: The reasonable man attempts to adapt ... - reading.ac.uk

78

2.1.9 Metabolite excretion assays

Cells of R. leguminosarum strains grown to A600 0.5-0.7 on AMS containing the

pertinent carbon and nitrogen sources, were harvested aseptically, washed once

in AMS and resuspended to an A600 of approximately 1 in AMS containing the

carbon and nitrogen sources used for initial growth. Resupended cells were

incubated at 28°C and samples removed at 0, 60, 150 and 270 min. Samples

were centrifuged at 11000 g for 15 min and the supernatants assayed for one or

more metabolites.

Glutamate was determined by a technique derived from Bernt and

Bergmeyer (1974). Each assay contained 75 µmol glycine, 60 µmol hydrazine

monohydrate, 4 µmol NAD+, 3 units of glutamate dehydogenase and 250 µl of

sample, in a total volume of 1.5 ml. Assays were incubated for 90 min at 37°C

prior to reading of the absorbance at 340 nm and determining the glutamate

concentrations from a standard curve prepared between 20 and 100 nmol.

Minus enzyme controls were run as blanks.

Aspartate was measured by a procedure adapted from Bergmeyer et al.

(1974). Each assay contained 150 µmol potassium phosphate (pH 7.2), 0.3 µmol

NADH, 15 µmol 2-oxoglutarate, 2 units of aspartate aminotransferase, 2 units of

malate dehydrogenase and 250 µl of sample, in a total volume of 1.5 ml. Assays

were incubated and read as for the glutamate determination.

Alanine was determined in a final volume of 1.5 ml containing 75 µmol

glycine, 60 µmol hydrazine monohydrate, 4 µmol NAD+, 0.5 unit of alanine

dehydrogenase and 250 µl of sample. Assays were incubated and read as for

the glutamate determination.

α-Ketoglutarate assays contained 150 µmol potassium phosphate (pH 7.2),

0.3 µmol NADH, 75 µmol aspartate, 2 units of aspartate aminotransferase, 2

units of malate dehydrogenase and 250 µl of sample, in a total volume of 1.5 ml.

Assays were incubated and read as for the glutamate determination.

Page 90: The reasonable man attempts to adapt ... - reading.ac.uk

79

2.1.10 Intracellular concentrations

500 ml cultures of R. leguminosarum strains were grown on glucose/NH4Cl,

harvested by centrifugation at a cell density of approximately 5 x 108 cells ml-1,

washed three times in AMS, and resuspended in 15 ml 100 mM HEPES, pH7.2,

all operations being carried out at 5°C. Cells were disrupted by two passages

through a French press at 69000 kPa, centrifuged at 30000 g, 5°C, for 20 min,

and the supernatant assayed for protein content. 2.5 ml of ice-cold 10% TCA

was added to 5 ml of supernatant. After centrifugation at 3500 g, 5°C, for 15

min, the pH of the supernatant was adjusted to 7.0 and the volume made up to

8 ml. Samples of the neutralized supernatant were assayed for glutamate and

α-ketoglutarate as described in Section 2.1.9.

2.1.11 Protein determination

The protein concentration of whole cells was determined by the method of

Lowry et al. (1951), using bovine serum albumin as standard. Protein

concentrations of extracts used in enzyme assays and PAGE were determined

by the method of Bradford (1976).

2.1.12 Plant Assays

Seeds of Pisum sativum c.v. meteor were surface sterilised, germinated,

inoculated with R. leguminosarum strains, and grown as described by Poole et al.

(1994b). Four weeks after inoculation, plants were harvested and acetylene

reduction carried out on whole plants as described by Trinick et al. (1976).

Sample nodules were removed and surface sterilised by immersion in calcium

hypochlorite (0.7%) for 10 min. Nodules were then washed three times in

sterile distilled water, crushed and bacteria streaked on TY agar. Isolated

bacteria were subsequently replica plated and screened for appropriate

antibiotic markers.

Page 91: The reasonable man attempts to adapt ... - reading.ac.uk

80

CHAPTER 3 THE CLONING AND CHARACTERIZATION OF THE GENERAL AMINO ACID PERMEASE OF RHIZOBIUM LEGUMINOSARUM STRAIN 3841

Page 92: The reasonable man attempts to adapt ... - reading.ac.uk

81

3.1 INTRODUCTION

Observations that Rhizobium bacteroids under nitrogen-fixing conditions

excrete alanine and/or aspartate (Kretovich et al., 1986; Appels & Haaker, 1991;

Kouchi et al., 1991; Rosendahl et al., 1992), are compatible with the proposal that

oxidation of the dicarboxylate malate by the bacteroid is coupled to

transamination and excretion of aspartate, forming a malate-aspartate shuttle in

the nodule (Kahn et al., 1985). Alternatively, amino acids may be excreted in

order to alleviate inhibition of the TCA cycle by preventing accumulation of

ketoacids and reducing equivalents in the bacteroid. Whichever pathway

operates, amino acid transport across the bacteroid/peribacteroid membrane is

likely to be crucial to nitrogen fixation, and may regulate bacteroid metabolism

(Murphy et al., 1979).

The unusual properties of the high affinity glutamate uptake system of R.

leguminosarum strain 3841 - a broad specificity for structurally unrelated amino

acid side chains, and an apparent ability to mediate exchange between

intracellular and extracellular amino acid pools - are compatible with a role in

nutrient exchange in the nodule, and make this transporter an interesting

subject for study in its own right.

In this chapter the cloning, sequencing and mutation of the genes encoding

the general amino acid permease of R. leguminosarum strain 3841 are described.

The results of experiments to investigate the involvement of this transporter in

amino acid exchange are discussed with regard to both the structure of the

transporter, and models of nutrient exchange in the nodule.

Page 93: The reasonable man attempts to adapt ... - reading.ac.uk

82

3.2 RESULTS

3.2.1 Isolation of cosmid pRU3024 carrying the general amino acid permease genes of Rhizobium leguminosarum strain 3841

The strategy employed in cloning the general amino acid permease genes of

Rhizobium leguminosarum strain 3841 is based on the potential for a strain

carrying additional copies of these genes to show increased labelling when

grown on [14C]glutamate.

Strain 3841 containing a strain 3841 chromosomal library (as EcoRI fragments

in pLAFR1 (Downie et al., 1983)) was grown on acid minimal salts agar

containing glucose (carbon source), NH4Cl (nitrogen source), [14C]glutamate,

and aspartate. Labelling by growth on [14C]glutamate is dependent on

glutamate incorporation and will only be indicative of glutamate uptake [by the

general amino acid permease] if glutamate uptake is the rate limiting-step in

incorporation. Hence, aspartate, which has been shown to cause a severe

reduction in uptake by the general amino acid permease (Reid, 1995), was

included in the growth media. After 2-3 days growth, colonies were lifted onto

nitro-cellulose filters, which were dried and exposed to X-ray film. On visual

inspection of the autoradiograph of approximately 3000 colonies, one colony

that apparently exhibited increased labelling was discerned with difficulty.

Following its purification from the selection media, some of the uptake

properties of this strain, RU438, were investigated (Table 3.1). It can be seen

that the uptake of all the amino acids tested is elevated in strain RU438 in

comparison to the wild-type. The increase in the transport rate of the non-

metabolizable amino acid α-aminoisobutyrate (AIB) is particularly significant

as it indicates that the increase in amino acid uptake is not the result of

increased metabolic drag. (AIB is not incorporated into TCA-precipitated

material from either strain 3841 or strain RU438 following uptake of AIB (Table

3.2)). The fact that glucose uptake is not increased in RU438 suggests that this

strain is affected in amino acid transport specifically.

Page 94: The reasonable man attempts to adapt ... - reading.ac.uk

83

Table 3.1 Rates of amino acid transport in R. leguminosarum strains grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. ND, not determined.

Substrate Strain 3841 RU438 RU439 RU640 RU641 RU913 L-Glutamate 5.6±0.4 19.3±1.1 38.1±4.0 43.9±4.3 38.1±2.6 4.9±0.2 L-Aspartate 3.9±0.1 13.8±0.1 ND 43.9±3.3 ND 3.0±0.8 L-Alanine 6.4±0.7 18.7±1.2 ND 63.7±2.7 ND 6.6±0.8 AIB 4.3±0.6 21.2±2.7 40.4±6.5 70.3±5.3 ND 5.0±0.3 L-Histidine 5.4±0.5 15.7±2.0 ND 40.0±1.2 ND 5.5±0.7 L-Leucine 5.3±0.2 10.8±1.3 ND 54.5±4.8 ND 5.6±0.2 L-Methionine 9.0±0.1 ND ND 37.7±2.8 ND ND D-Glucose 39.5±3.2 29.1±2.3 38.7±2.2 46.2±6.9 ND 44.5±2.1

Table 3.2 Labelling of TCA precipitated material by [14C]AIB and [14C]Alanine in R. leguminosarum strains 3841 and RU438. Cells grown on glucose/NH4Cl/aspartate were washed and resuspended in minimal salts in the presence of either [14C]AIB or [14C]alanine. After the time intervals shown, two identical samples were taken. One of these was added to ice-cold 10%(w/v) TCA and the amount of label in the precipitate determined. The label in whole cells was obtained from the second sample. Values are the result from a single experiment and are expressed as percentage of added label. Substrate Assay Time (min) and strain 1 5 10 AIB 3841 Whole cells 2.0 7.9 14.2 RU438 Whole cells 7.0 28.1 41.8 3841 TCA precipitate <0.1 <0.1 <0.1 RU438 TCA precipitate <0.1 <0.1 <0.1 L-Alanine 3841 Whole cells 2.0 10.0 16.6 RU438 Whole cells 6.3 21.0 27.4 3841 TCA precipitate 0.3 3.5 10.3 RU438 TCA precipitate 0.4 2.5 6.6

Page 95: The reasonable man attempts to adapt ... - reading.ac.uk

84

In order to confirm that the observed transport phenotype of RU438 was

caused by the cosmid, pRU3024, carried by the strain, rather than a

chromosomal mutation in the host, pRU3024 DNA was isolated from strain

RU438 and used to transform E. coli strain DH5α. Cosmid pRU3024 was then

conjugated from the resulting E. coli strain back into strain 3841. The transport

phenotype of the resulting R. leguminosarum strain with respect to the

previously tested amino acids and glucose was found to be the same as that of

strain RU438 (data not shown).

3.2.2 Restriction mapping, sub-cloning and mutational analysis of pRU3024

Restriction analysis of pRU3024 was carried out using BamHI, EcoRI, and

HindIII. The resulting map of strain 3841 DNA is shown in Fig. 3.1. This map

does not show all of the 17.1kb insert DNA. Two small EcoRI fragments (2.1kb

and 0.7kb) which lie outside the region shown were not mapped.

The 6.6kb and 3.2kb BamHI fragments, the 6.3kb, 4.8kb and 3.2kb EcoRI

fragments, and the 10.2kb and 2.9kb HindIII fragments of pRU3024 were cloned

in the broad host range, medium copy number vector pRK415-1. These sub-

clones were then conjugated into strain 3841 and the rate of glutamate uptake

by the resulting strains investigated. For the purposes of comparison glutamate

uptake is taken to be representative of general amino acid uptake. All strains

showed glutamate transport similar to the wild-type (data not shown) except

strain RU439, which contains the clone of the 10.2kb HindIII fragment

(pRU135). Strain RU439 exhibits similar rates of glutamate, AIB and glucose

uptake to those of strain RU438 (Table 3.1).

Page 96: The reasonable man attempts to adapt ... - reading.ac.uk

85

Fig. 3.1 Restriction map of pRU3024 with sub-clones below. The shaded arrows indicate the direction of transcription initiation from the lac promoter in the vector of the sub-clones. Restriction sites are: B, BamHI; C, ClaI; D, DraI; E, EcoRI; H, HindIII; M, MluI; N, NcoI; Nt, NotI; S, SstI; Sp, SspI.

Page 97: The reasonable man attempts to adapt ... - reading.ac.uk

86

Plasmid pRU135 was subjected to further restriction analysis (Fig. 3.1) and a

variety of sub-clones in pRK415-1 created. The effect of these sub-clones on

glutamate uptake in strain 3841 was investigated. Two sub-clones of pRU135

capable of producing the same effect on amino acid transport in strain 3841 as

pRU135 itself (data not shown) were found to be pRU165, a clone of the 8.7kb

HindIII-MluI fragment, and pRU185, a clone of the 6.9kb HindIII-ClaI fragment

(Fig. 3.1). By contrast, pRU149 and pRU186, clones of the 6.0kb HindIII-SstI and

5.2kb NcoI fragments of pRU135, respectively (Fig. 3.1), have no effect on amino

acid transport. From these results it was concluded that the DNA responsible

for increasing amino acid uptake in strain 3841 is contained within the 5.4kb

region of pRU135 that is bounded by the unique MluI and ClaI sites (Fig. 3.1).

Concomitant to the above restriction analysis and sub-cloning, pRU3024 was

subjected to saturation Tn5-lacZ mutagenesis. Mutated cosmids were screened

for their effect on amino acid transport in strain 3841 (Table 3.3). Nine cosmids

(pRU3026, pRU3027, pRU3028, pRU3029, pRU3032, pRU3035, pRU3039,

pRU3046 and pRU3048) were isolated that produce no effect on amino acid

uptake. Restriction analysis of these cosmids revealed that in all cases the

transposon had inserted in the 5.4kb MluI-ClaI region found in pRU135 (Fig.

3.2). It was therefore decided to sequence this 5.4kb of DNA.

Page 98: The reasonable man attempts to adapt ... - reading.ac.uk

87

Table 3.3 Rates of glutamate uptake by R. leguminosarum strains grown on glucose/NH4Cl/aspartate. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the result from a single experiment except those for strains 3841 and RU438 which are the mean±SEM of determinations from three or more independent cultures. Strain Uptake 3841 3.2±0.1 RU438 (3841/pRU3024) 19.9±1.2 RU441 (3841/pRU3026) 3.6 RU442 (3841/pRU3027) 5.7 RU443 (3841/pRU3028) 3.1 RU500 (3841/pRU3029) 4.5 RU502 (3841/pRU3031) 8.0 RU503 (3841/pRU3032) 5.3 RU504 (3841/pRU3033) 8.4 RU506 (3841/pRU3035) 4.9 RU510 (3841/pRU3039) 3.7 RU515 (3841/pRU3044) 20.8 RU517 (3841/pRU3046) 3.0 RU519 (3841/pRU3048) 3.0 RU522 (3841/pRU3049) 9.4 RU526 (3841/pRU3053) 15.7 RU972 (3841/pRU3080) 18.8 RU974 (3841/pRU3082) 21.7 RU975 (3841/pRU3083) 19.7 RU976 (3841/pRU3084) 20.5 RU977 (3841/pRU3085) 19.0 RU978 (3841/pRU3086) 19.7 RU1069 (3841/pRU3089) 19.8 RU1070 (3841/pRU3090) 22.7 RU1071 (3841/pRU3091) 21.1

Page 99: The reasonable man attempts to adapt ... - reading.ac.uk

88

Fig. 3.2 Restriction map showing location of transposons in mutants of cosmid pRU3024. The locations of the Tn5-lacZ insertions are flagged with the number of the cosmid in which they occur. Each flag points in the direction of transcription of the lacZ gene in the transposon. Filled flags represent active fusions. Restriction sites are: B, BamHI; C, ClaI; E, EcoRI; H, HindIII; M, MluI.

Page 100: The reasonable man attempts to adapt ... - reading.ac.uk

89

Three mutated pRU3024 cosmids were isolated that do not cause the increase

in glutamate uptake in strain 3841 that is associated with pRU3024 itself, but do

increase uptake to a lesser degree (Table 3.3). In one of these cosmids,

pRU3049, the insert lies close to one end of the previously described 5.4kb MluI-

ClaI region (Fig. 3.2), and the phenotype produced by this cosmid is likely to be

due to the proximity of the transposon to the promoter region for the genes

lying between these restriction sites (Section 3.2.6). Intriguingly however, the

transposons in the other two cosmids in this class, pRU3031 and pRU3033, are

both located in a 0.8kb BamHI fragment approximately 3.5kb away from the

5.4kb MluI-ClaI region (Fig. 3.2). Since cosmids containing mutations in the

intervening 3.5kb region, for example pRU3083, pRU3085, and pRU3086 (Fig.

3.2), produce the same increase in glutamate uptake in 3841 as pRU3024 (Table

3.3), it appears that pRU3024 contains two distinct regions of DNA that affect

amino acid uptake in strain 3841.

3.2.3 Nucleotide sequence of the 5.4kb MluI-ClaI fragment of pRU135

A variation of the ordered deletion strategy described by Robson et al. (1986)

was employed in determining the nucleotide sequences of both strands of the

5.4kb MluI-ClaI fragment from pRU135:

The 5.4kb fragment was cloned, in both orientations, into the EcoRV site of

pBluescript SK+, creating plasmids pRU189 and pRU190. Each plasmid was cut

at its unique SmaI site within the polylinker, partially digested with BstUI and

ligated (Fig. 3.3). E. coli strain DH5α was transformed with the resulting DNA

and AmpR transformants were screened for a range of plasmids between 3.0kb

and 8.4kb that retained the XbaI site of the pBluescript SK+ polylinker.

Page 101: The reasonable man attempts to adapt ... - reading.ac.uk

90

Fig. 3.3 Schematic representation of the strategy employed to obtain ordered deletions of pRU189 and pRU190 for sequencing. The unfilled arrow indicates the direction of sequencing from primers Reverse and SK. Restriction sites are: X, XbaI; S, SmaI.

Page 102: The reasonable man attempts to adapt ... - reading.ac.uk

91

Sequencing of the insert DNA in these plasmids (which are illustrated in Fig.

3.4) using Reverse and SK primers to the pBluescript SK+ polylinker, provided

the nucleotide sequence shown in Fig. 3.5. (In the case of pRU247 an additional

primer (P16), to the end of the sequence provided by the SK primer, was

needed to obtain sequence overlapping that from the next deletion, pRU246).

In the course of this sequencing it was found that sequence obtained from

pRU190 lacked the G at position 438 of the sequence in Fig. 3.5. That this was

due to the deletion of a base during the construction of pRU190 rather than the

presence of an erroneous additional base in pRU189, was demonstrated by

sequencing cosmid pRU3024 itself in the vicinity of the ambiguity, using two

primers (P21 and P22), one to each strand approximately 100bp upstream of the

problematic base.

In order to be certain that the sequenced DNA was capable of enhancing

amino acid uptake in strain 3841, the insert DNA from both pRU189 and

pRU190 was excised as a 5.4kb HindIII-XbaI fragment and in each case cloned

into pRK415-1. The two resulting plasmids, pRU191 and pRU192, respectively,

were then conjugated into strain 3841 and glutamate uptake in the resulting

strains (RU640 and RU641, respectively) measured (Table 3.1). Both strains

show a greater than six-fold increase in glutamate transport in comparison to

the wild-type.

Page 103: The reasonable man attempts to adapt ... - reading.ac.uk

92

Fig. 3.4 Schematic representation of ordered deletions of pRU189 and pRU190. Figures in parentheses indicate the position of the relevant BstUI site for each deletion with reference to the sequence in Fig. 3.5. Arrows indicate the direction of sequencing from primers Reverse and SK.

Page 104: The reasonable man attempts to adapt ... - reading.ac.uk

93

Page 105: The reasonable man attempts to adapt ... - reading.ac.uk

94

Page 106: The reasonable man attempts to adapt ... - reading.ac.uk

95

Page 107: The reasonable man attempts to adapt ... - reading.ac.uk

96

Page 108: The reasonable man attempts to adapt ... - reading.ac.uk

97

Page 109: The reasonable man attempts to adapt ... - reading.ac.uk

98

Page 110: The reasonable man attempts to adapt ... - reading.ac.uk

99

Page 111: The reasonable man attempts to adapt ... - reading.ac.uk

100

Page 112: The reasonable man attempts to adapt ... - reading.ac.uk

101

3.2.4 Coding regions of the nucleotide sequence from pRU189

The deduced amino acid sequences from four open reading frames (ORFs)

found in the nucleotide sequence from pRU189 are shown in Fig. 3.5. All of

these ORFs are transcribed in the same direction. Proposed protein starts are

based on the presence of methionine codons which are preceded by sequence

resembling a ribosome binding site, and on comparisons of the proteins to

similar proteins in other bacteria.

Sequence comparisons suggest that the polypeptides encoded by the four

ORFs constitute the components of an ABC transporter, with particular

homology being found to carriers of amino acids and opines. Consequently, we

propose the gene designation aap (for amino acid permease). The letter

assigned to an individual gene (Fig. 3.5) is based on that of the corresponding

gene in the histidine transport operon of Salmonella typhimurium.

The deduced amino acid sequence of the aapP gene product (molecular

weight 29kD) exhibits extensive homology to the ATP-binding proteins of other

bacterial amino acid transporters. The strongest identities are to GlnQ from

Bacillus stearothermophilus (61.4%), GlnQ, HisP and ArtP from E. coli (52.1%,

50.2% and 43.9%, respectively), GluA from Cornyebacterium glutamicum (61.8%),

HisP from S. typhimurium (50.0%), and OccP and NocP from Agrobacterium

tumefaciens (46.4% and 49.0%, respectively). AapP also contains the ATP/GTP-

binding site motif A and the ABC transporters family signature (Fig. 3.5) that

are characteristic of the ATP-binding proteins of ABC transporters (Higgins et

al., 1986; Higgins, 1992).

The C-terminal ends of AapQ and AapM show significant homology to

integral membrane proteins from other ABC transporters. In particular, both

proteins contain the recently determined consensus sequence (Saurin et al.,

1994) characteristic of this type of protein (Fig. 3.5). AapQ exhibits 30.6%

identity over 121 amino acids to OccM from A. tumefaciens and 33.0% identity

over 115 amino acids to HisM from E. coli, while AapM is 28.4% identical over

221 residues to HisM from S. typhimurium and 27.7% identical over 223 amino

Page 113: The reasonable man attempts to adapt ... - reading.ac.uk

102

acids to GlnP from E. coli. However, both AapM and AapQ are larger by 110 to

180 amino acids than other previously described amino acid transport proteins

of this type, with their predicted molecular weights being 42kD and 43kD,

respectively.

AapJ exhibits similarities to binding proteins. The homology to the

corresponding component of the glutamine transport system of B.

stearothermophilus (29.5% identity over 245 residues) is greatest (although AapJ

possesses an additional 96 amino acids at the C-terminal end). At the N-

terminal end of the protein a putative signal sequence (Fig. 3.5) is identified by

the GCG program SIGCLEAVE, indicating translocation of AapJ through the

cytoplasmic membrane.

Interestingly, the most significant homology to each of the Aap proteins is

provided by the deduced polypeptide from one of four adjacent unidentified

ORFs in the 67.4 to 76.0 minute region of the E. coli K12 chromosome (EMBL

accession number U18997). Each aap gene product shows identity to a different

E. coli ORF as follows: AapJ, 58.4%; AapQ, 48.9%; AapM, 50.7%; AapP, 71.0%.

Indeed, the genes occur in the same order in E. coli as they do in R.

leguminosarum. The homology of the E. coli ORFs to AapJ, AapQ and AapM is

particularly striking since these proteins are significantly bigger, and only

approximately 30% identical to the corresponding components of previously

reported ABC transporters of amino acids.

AapJ is also 57.6% identical to the translation of an incomplete unidentified

ORF from Pseudomomas fluorescens (EMBL accession number D00852; (Hong et

al., 1991) ). An alignment of AapJ and the AapJ-like proteins from E. coli and P.

fluorescens is shown in Fig. 3.6.

Page 114: The reasonable man attempts to adapt ... - reading.ac.uk

103

Rl MKNKLLSAAI--GAAVLAVGASAASATTLSDVKAKGFVQCGVNTGLTGFAAPDASGNWAGFD

Ec EKDDDSHTGC--RQRAACRCKSGAAGATLDAVQKKGFVQCGISDGLPGFSYADADGKFSGID

Pf MKVLKSTLAIFTAAAVLGVSGFAQAGATLDAVQKKGFVQCGVSDGLPGFSVPDASGKILGID

Pf2 MKVLKSTLAIVCAAAVLGVSGFAQAGATLDAVQKKGFVQCGVSDGLPGFSV-----------

# ~ ~ ~~~## #~ #######~~ ## ##~ ~## #~ #~#

Rl VDFCKAVASAVFGDPTKVKYTPTNAKERFTALQSGEIDVLSRNTTWTINRDTALGFNFRPVT

Ec VDICRGVAAAVFGDDTKVKYTPLTAKERFTALQSGEVDLLSRNTTWTSSRDAGMGMAFTGVT

Pf ADYCRAVAAAVFGDATKVKFSQLNAKERFTALQSGEVDILSRNTTMTSSRDAGMGLKF----

Pf2 --------------------------------------------------------------

# #~~##~##### ####~~ ############~#~###### # ~##~~~# # ##

Rl YYDGQGFMVRKGLNVKSALELSGAAICVQSGTTTELNLADYFKTNNLQYNPVVFENLPEVNA

Ec YYDGIGFLTHDKAGLKSAKELDGATVCIQAGTDTELNVADYFKANNMKYTPVTFDRSDESAK

Pf --------------------------------------------------------------

Pf2 --------------------------------------------------------------

#### ##~ ~ ~### ## ##~~#~#~## ####~#####~##~~# ## #~ #

Rl AYDAGRCDVYTTDQSGLYSLRLTLKNPDEHIILPEIISKEPLGPAVRQGDDQWFDIVSWTAY

Ec ALESGRCDTLASDQSQLYALRIKLSNPAEWIVLPEVISKEPLGPVVRRGDDEWFSIVRWTLF

Pf --------------------------------------------------------------

Pf2 --------------------------------------------------------------

# ~~#### ~~### ##~##~ # ## # #~###~######## ##~###~## ## ## ~

Rl ALINAEEFGITQANVDEMKNSPN-PDIKRFLGSETDTKIGTDLGLTNDWAANVIKGVGNYGE

Ec AMLNAEEMGINSQNVDEKAANPATPDMAHLLGKEGDY--GKDLKLDNKWAYNIIKQVGNYSE

Pf --------------------------------------------------------------

Pf2 --------------------------------------------------------------

#~~#### ## #### ~# ##~ ~~## # # # ## # # ## #~## ####~#

Rl IFERNIGQGSPLKIARGLNALWNKGGIQYAPPVR

Ec IFERNVGSESPLKIKRGQNNLWNNGGIQYAPPVR

Pf ----------------------------------

Pf2 ----------------------------------

#####~# ##### ## # ###~##########

Fig. 3.6 Alignment of AapJ from R. leguminosarum (Rl) with similar proteins from E. coli (Ec, EMBL accession number U18997) and P. fluorescens (Pf, EMBL accession number D00852; Pf2, (Hong et al., 1991)). Alignment was performed using the program ALIEN. #, identical amino acids; ~, conservative substitutions

Page 115: The reasonable man attempts to adapt ... - reading.ac.uk

104

The EMBL entry U18997 indicates a possible frame shift in the region of the

sequence corresponding to aapJ, and an arbitrary deletion of the two nucleotides

at positions 199844-5 in this sequence yields an E. coli protein which now also

exhibits significant homology to the other two AapJ proteins over the first

twenty residues, including a methionine start at the same position as that

suggested for the R. leguminosarum protein.

Hydropathy analysis of the deduced amino acid sequences of AapQ and

AapM by the method of Engelman et al. (1986), incorporating the "positive-

inside rule" of von Heijne (1986; 1992), predicts eight membrane-spanning

regions for each protein (Figs. 3.7 and 3.8). Kyte and Doolittle (1982) plots

predict nine membrane-spanning regions for each protein, with the additional

transmembrane segment occurring between the sixth and seventh membrane-

spanning regions of the Engelman plot, in each case. This is a greater number

of transmembrane segments than has been reported for the corresponding

components of other ABC transporters of amino acids. Indeed, six membrane-

spanning regions has been suggested as the paradigm for each integral

membrane component of an ABC transporter (Higgins, 1992).

Page 116: The reasonable man attempts to adapt ... - reading.ac.uk

105

Fig. 3.7 Topology prediction for AapM generated by the program TOP-PRED (Claros & von Heijne, 1994). The hydrophobicity profile (top) calculated by the method of Engelman et al. (1986) with a full window of 21 amino acids and a core window of 11 amino acids, indicates 6 certain and 3 putative membrane-spanning regions. The most likely topology (bottom), selected from the various possible topologies that include all the certain transmembrane segments, but either include or exclude each of the putative ones, is that which has the greatest degree of bias in the distribution of positively charged residues. LL, loop length; KR, number of lysine and arginine residues; KR Diff, positive charge difference.

Page 117: The reasonable man attempts to adapt ... - reading.ac.uk

106

Fig. 3.8 Topology prediction for AapQ generated by the program TOP-PRED (Claros & von Heijne, 1994). The hydrophobicity profile (top) calculated by the method of Engelman et al. (1986) with a full window of 21 amino acids and a core window of 11 amino acids, indicates 8 certain membrane-spanning regions. The most likely topology, on the basis of the distribution of positively charged residues, is illustrated (bottom). LL, loop length; KR, number of lysine and arginine residues; KR Diff, positive charge difference.

Page 118: The reasonable man attempts to adapt ... - reading.ac.uk

107

Sequence alignment of AapM and AapQ with each other, and the integral

membrane components of ABC transporters that mediate uptake of polar amino

acids or glutamine, reveals a region of 63 amino acids that contains 18 well

conserved residues (Fig. 3.9). On the basis of the topologies illustrated in Figs.

3.7 and 3.8, the conserved residues in both AapM and AapQ lie within two

membrane-spanning regions and the cytoplasmic loop that connects them. In

the case of AapM it is transmembrane segments 4 and 5 (Fig. 3.7) that are

involved, while for AapQ segments 2 and 3 (Fig. 3.8) contain the conserved

amino acids. The reason for this difference is not clear, however in the case of

AapQ, the location of the conserved residues may indicate a significant

difference to MalF, the integral membrane component of the ABC transporter of

maltose in E. coli. In MalF the presence of eight transmembrane segments has

been demonstrated experimentally (Froshauer et al., 1988). However, the two

N-terminal transmembrane segments can be deleted without loss of protein

function (Ehrmann et al., 1990). This has given credence to the suggestion that

MalF consists of the six standard membrane-spanning regions, with an

extension of two transmembrane segments at the N-terminus (Overduin et al.,

1988; Higgins, 1992). The presence of conserved residues in transmembrane

segment 2 (Fig. 3.8) of AapQ, suggests that deletion of the two N-terminal

membrane-spanning regions of this protein is likely to impair function.

In all the proteins in Fig. 3.9 except those from general permeases, the

conserved region is located very close to the N-terminus. Since in AapM and

AapQ from R. leguminosarum, these conserved residues apparently constitute

two transmembrane segments and a cytoplasmic loop, it might be predicted

that the first two membrane-spanning regions of the non-general permeases in

Fig. 3.9 are joined by a cytoplasmic loop, and that therefore the N-terminus of

these proteins is periplasmic. Hydropathy analysis of each of these proteins by

both the method of Engelman et al. (1986), and that of Kyte and Doolittle (1982),

taking account of the "positive-inside rule" of von Heijne (1986; 1992), predicts

Page 119: The reasonable man attempts to adapt ... - reading.ac.uk

108

five membrane-spanning regions, with a periplasmic N-terminus and a

cytoplasmic C-terminus in the majority of cases. Most exceptions arise from use

of the Engelman scale, which leads to the prediction of four transmembrane

segments for OccQ, YckJ, ArtQ, GltJ, HisM, and GlnP from E. coli and R.

prowazekii, while a Kyte and Doolittle plot results in a prediction of six

transmembrane segments for GltK, and GlnP from R. prowazekii. However, a

five-transmembrane-segment topology has been demonstrated experimentally

for HisM and HisQ (Kerppola et al., 1991; Kerppola & Ames, 1992), and in

addition to producing a consistent location for the N-terminal conserved region,

the five-transmembrane-segment topology places the "integral membrane

component signature" sequence (Saurin et al., 1994) in the cytoplasm in each of

the proteins analyzed. This is consistent with the known location of this

conserved sequence in other ABC transporters (Higgins, 1992). It may therefore

be the case that five, rather than six, membrane-spanning regions is the norm

for the integral membrane components of ABC transporters of this group of

amino acids.

Some of the residues in the region illustrated in Fig. 3.9 may be involved in

substrate specificity, since spontaneous mutants of S. typhimurium in which the

specificity of the histidine transporter is altered from L-histidine to L-histidinol,

are found to contain a deletion from HisM of four amino acids, corresponding

to positions 25-28 of this region (Payne et al., 1985).

In addition, there is a correlation between the nature of the amino acid that

replaces the leucine at position 53 of the conserved region (Fig. 3.9), and the

nature of the substrate transported by the proteins in which this leucine is not

conserved. In NocQ, OccQ, ArtQ and HisQ, each of which constitutes one half

of the integral membrane complex of a transporter of (substituted) basic amino

acid(s), the leucine at position 53 is replaced by an acidic amino acid, either

glutamate or aspartate (Fig. 3.9). Similarly, in one of the two integral membrane

components of the general amino acid permease of both E. coli (see below) and

Page 120: The reasonable man attempts to adapt ... - reading.ac.uk

109

R. leguminosarum, this leucine is replaced by proline (Fig. 3.9). In both the

integral membrane components (or the sole integral membrane component) of

the remaining transporters in Fig. 3.9, all of which transport either glutamate or

glutamine, the leucine at position 53 is either conserved or replaced by a similar

neutral amino acid.

For each of the integral membrane proteins in Fig. 3.9, position 53 of the

conserved region is predicted to lie within a transmembrane segment, as are

residues 25-28 of the conserved sequence in HisM. Thus this data appears to be

compatible with substrate specificity being determined within the pore formed

by the integral membrane components of the transporter.

Page 121: The reasonable man attempts to adapt ... - reading.ac.uk

110

Page 122: The reasonable man attempts to adapt ... - reading.ac.uk

111

As well as the four complete ORFs in the sequence in Fig. 3.5, another

incomplete ORF which starts at base 351 and is transcribed in the opposite

direction to the aap operon was investigated. The deduced amino acid sequence

from this ORF (Fig. 3.5) shows significant homology to the N-terminus of MetC

from Bordetella avium and E. coli (39.6% and 33.6% identity, respectively).

Furthermore, translations of sequence obtained using a primer to the 5' end of

the Tn5-lacZ in mutated pRU3024 cosmids in which the insertion lies

downstream of the start of this ORF , also show homology to MetC (Section

3.2.23). It therefore seems likely that this ORF encodes beta-cystathionase in R.

leguminosarum.

3.2.5 Other features of the nucleotide sequence from pRU189

Downstream of the stop codon at the end of aapP there is an inverted repeat,

centred at positions 5134-5, with potential for forming a G/C-rich stem-loop

structure (∆G = -156.5 kJ mol-1, calculated according to Tinoco et al. (1973))

followed by at least four T residues. These are the characteristics of a rho-

independent terminator (Rosenberg & Court, 1979; Platt & Bear, 1983).

The intergenic region between aapJ and aapQ also contains an inverted repeat

(centred at a position 1773) with potential for forming a stem-loop structure (∆G

= -128.9 kJ mol-1, calculated according to Tinoco et al. (1973)). Such stem-loop

structures are probably involved in differential gene expression in polycistronic

operons (Higgins et al., 1988), and this role has been suggested for extragenic

palindromic sequences following the coding region for the periplasmic protein

HisJ in the histidine transport operon of S. typhimurium (Higgins et al., 1982a;

Higgins et al., 1982b; Stern et al., 1988).

Page 123: The reasonable man attempts to adapt ... - reading.ac.uk

112

3.2.6 Mutation of the general amino acid permease

Tn5-lacZ mutants of strain 3841 were generated by homogenotization

(Ruvkun & Ausubel, 1981) of mutated pRU3024 cosmids that showed impaired

ability to increase amino acid uptake. The location of the mutation in each of

the resulting strains, which was confirmed by Southern blotting (Fig. 3.10), is

illustrated in Fig. 3.11. In the case of mutations lying within the sequenced

5.4kb MluI-ClaI region, the position of the transposon was mapped precisely by

sequencing the corresponding cosmid DNA adjacent to the insert, using a

primer (P15) to the 5' end of Tn5-B20 (Fig. 3.5).

Page 124: The reasonable man attempts to adapt ... - reading.ac.uk

113

Fig. 3.10 Southern blots of Tn5-B20 mutants of strain 3841. A: HindIII digested chromosomal DNA probed with the 10.2kb insert from pRU135. The signal above the 10.2kb band in the 3841 lane is due to incomplete digestion. In each of the mutants the 10.2kb band is replaced by two bands (or a presumed doublet in the case of strain RU636), of the expected size. (Tn5-B20 is ~8.3kb and contains an internal HindIII fragment of ~2.8kb). B: BamHI digested chromosomal DNA probed with the 3.2kb insert from pRU133 (Fig. 3.1). Of the two bands expected to replace the ~0.8kb band of the wild-type in the strain RU632 lane, one is calculated to be only ~0.2kb, with approximately half this DNA being derived from the transposon. This fragment presumably contains insufficient DNA homologous to the probe to be detected.

Page 125: The reasonable man attempts to adapt ... - reading.ac.uk

114

Fig. 3.11 Restriction map of pRU3024 and mutants derived from it. Boxes represent coding regions, the unfilled arrows indicating direction of transcription. The locations of Tn5-lacZ insertions are flagged with the number of the mutant strain (and cosmid) in which they occur. Each flag points in the direction of transcription of the lacZ gene in the transposon. Filled flags represent active fusions. Restriction sites are: B, BamHI; C, ClaI; E, EcoRI; H, HindIII; M, MluI.

Page 126: The reasonable man attempts to adapt ... - reading.ac.uk

115

The uptake rates of glutamate and glucose for the mutants were determined

(Table 3.4).

Table 3.4 Rates of glutamate and glucose uptake by R. leguminosarum strains grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. ND, not determined. Strain Relevant Substrate genotype L-Glutamate D-Glucose 3841 Wild-type 5.6±0.4 39.5±3.2 RU541 aapQ::Tn5-lacZ 1.7±0.1 40.0±1.6 RU542 aapP::Tn5-lacZ 1.6±0.3 36.9±1.9 RU543 aapJ::Tn5-lacZ 1.2±0.6 33.3±1.8 RU631 aapQ::Tn5-lacZ 2.0±0.1 ND RU632 cysE::Tn5-lacZ 4.6±0.3 39.5±3.2 RU633 aapM::Tn5-lacZ 1.8±0.5 ND RU634 aapM::Tn5-lacZ 1.8±0.8 44.7±3.2 RU635 aapQ::Tn5-lacZ 2.0±0.2 ND RU636 aapQ::Tn5-lacZ 1.9±0.6 37.4±0.7 RU637 aapJ::Tn5-lacZ 1.5±0.1 ND RU638 See text 2.8±0.5 ND

With the exception of RU632 and RU638, all the strains show an

approximately 3-fold reduction in glutamate uptake in comparison to the wild-

type, while glucose uptake, in the strains for which it was measured, is

unimpaired. This phenotype is consistent with mutation of the general amino

acid permease, and four of these mutants RU542, RU543, RU634 and RU636,

having mutations in aapP, aapJ, aapM and aapQ, respectively, where chosen for

further study.

Glutamate uptake in strain RU638 is impaired in comparison to the wild-type

strain, but is not as severely affected as that in strain RU543. This is consistent

with the location of the start of aapJ being as proposed in Fig. 3.5, with the

Page 127: The reasonable man attempts to adapt ... - reading.ac.uk

116

mutation in RU638 lying upstream of aapJ and potentially affecting the

promoter region. (Alternative ATG starts for aapJ result in the transposon in

RU638 lying within aapJ. If this were the case, glutamate uptake in strains

RU543 and RU638 would be expected to be identical).

The mutation in strain RU632 and its effects are discussed elsewhere

(Sections 3.2.20 and 3.2.21).

3.2.7 Mapping of promoter sites in the aap operon by complementation analysis

In order to determine the location of promoters for the aap genes, the ability

of plasmid-borne copies of aapJQMP, aapQMP, aapMP and aapP to complement

chromosomal mutations in aapJQMP was investigated.

While strain 3841 will grow on minimal salts agar containing glutamate as

the sole source of carbon and nitrogen, strains RU542, RU543, RU634 and

RU636 were found to be unable to grow on this medium. It was therefore

possible to use growth on glutamate as a test for complementation of mutations

in the aap operon.

Clones in both pRK415-1 and pIJ1891 of aapJQMP, aapQMP, aapMP and aapP

were each constructed from an appropriate deletion of pRU189 previously

generated for sequencing, the insert 3841 DNA being excised from pBluescript

SK+ as a HindIII-XbaI fragment (Fig. 3.12). In plasmid pRK415-1 the cloned

genes are in the opposite orientation to that of the lac promoter in the vector,

whereas in pIJ1891 transcription of the genes carried by the inserts can be

initiated by this promoter. (The lac promoter is constitutive in R. leguminosarum

(Labes et al., 1990)). The cosmid vector pIJ1891 was employed in cloning aap

genes downstream of the lac promoter, since its low copy number minimises the

possibility of over expression of these genes from the foreign promoter.

Each of the clones was conjugated into RU542, RU543, RU634 and RU636 and

the resulting strains tested for growth on minimal salts agar containing

glutamate as the sole carbon and nitrogen source (Fig. 3.12).

Page 128: The reasonable man attempts to adapt ... - reading.ac.uk

117

Fig. 3.12 Complementation of aap mutants for growth on 10mM glutamate as sole source of carbon and nitrogen. Growth of strains (right) containing plasmids (left) is scored: +, good growth; +/-, poor growth; -, no growth. * On incubation for a further 4-5 days these strains show good (+) growth. The shaded arrows indicate the direction of transcription initiation from the lac promoter in the vector of the plasmids. Restriction sites are: Bs, BstUI; C, ClaI; M, MluI.

Page 129: The reasonable man attempts to adapt ... - reading.ac.uk

118

The complete aap operon carried by either of the vectors complements a

chromosomal mutation in any of the aap genes, whilst plasmids lacking a

particular gene fail to complement a mutation in that gene. Mutations in aapQ

and aapM are not complemented by any of the clones in pRK415-1 other than

that carrying the complete aap operon, suggesting that the first three genes in

the operon are under the control of a single promoter at the start of the operon.

This is confirmed by the ability of aapQMP cloned in pIJ1891, under the control

of the lac promoter, to complement mutations in both aapQ and aapM.

Likewise, aapMP complements a mutation in aapM. Since a mutation in aapP is

eventually complemented by aapQMP, aapMP or aapP cloned in pRK415-1, it

would appear that there may be some weak promoter activity directly

upstream of aapP. Enhanced transcription of aapP relative to that of aapQM is

consistent with the expected association of two copies of AapP with one each of

AapQ and AapM within the transporter (Kerppola et al., 1991; Davidson &

Nikaido, 1991).

3.2.8 Transcription levels of aap genes

β-Galactosidase assays were carried out on the reverse mutants previously

described in Section 3.2.6, with cells being grown under both nitrogen-excess

(glucose/NH4Cl) and nitrogen-limited (glucose/glutamate) conditions.

The results (Table 3.5) for the aap mutants are consistent with the direction of

transcription for the aap operon proposed in Section 3.2.4. Only those aap

mutants in which the lacZ gene in the transposon is in the same orientation as

that proposed for the mutated aap gene show appreciable β-galactosidase

activity. Active fusions of aap genes show the expected repression of

transcription under nitrogen-excess conditions (Poole et al., 1985; Poole et al.,

1987), though whether strains mutated in their major glutamate uptake system

are likely to be simply nitrogen-limited when grown on glutamate as the

nitrogen source, may need to be taken into account when considering absolute

values (Section 4.2.2).

Page 130: The reasonable man attempts to adapt ... - reading.ac.uk

119

Table 3.5 β-galactosidase activities in aap mutants of R. leguminosarum strain 3841. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant Growth Conditions genotype Glucose/NH4Cl Glucose/Glutamate 3841 Wild-type 15±2 22±3 RU541 aapQ::Tn5-lacZ 23±5 30±2 RU542 aapP::Tn5-lacZ 15±1 27±1 RU543 aapJ::Tn5-lacZ 155±23 690±39 RU631 aapQ::Tn5-lacZ 14±1 21±1 RU632 cysE::Tn5-lacZ 139±10 269±23 RU633 aapM::Tn5-lacZ 16±5 14±1 RU634 aapM::Tn5-lacZ 50±5 309±46 RU635 aapQ::Tn5-lacZ 38±3 123±6 RU636 aapQ::Tn5-lacZ 75±9 325±50 RU637 aapJ::Tn5-lacZ 164±41 568±37 RU638 See text 30±2 35±1

There is a considerable difference in activity between the fusions in strains

RU543/RU637 (aapJ) and those in strains RU635/RU636 (aapQ) and RU634

(aapM), both under nitrogen-excess and nitrogen-limited conditions. This result

is in accordance with the expectation that the periplasmic binding protein will

be expressed at a greater level than the membrane components of the

transporter (Higgins et al., 1982b). Since aapJ, aapQ and aapM are under the

control of a single promoter (Section 3.2.7), the difference in the expression of

these genes is presumably due to transcriptional attenuation by the putative

stem-loop located between aapJ and aapQ (Section 3.2.5).

3.2.9 Amino acid uptake in strains RU542, RU543, RU634 and RU636

The rates of uptake of a range of amino acids by strains RU542, RU543,

RU634 and RU636 grown on both glucose/NH4Cl and glucose/glutamate were

measured (Tables 3.6 and 3.7).

Page 131: The reasonable man attempts to adapt ... - reading.ac.uk

120

Table 3.6 Rates of amino acid transport in R. leguminosarum strains grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures.

Substrate Strain 3841 RU542 RU543 RU634 RU636 L-Glutamate 5.6±0.4 1.6±0.3 1.2±0.6 1.8±0.8 1.9±0.6 L-Aspartate 3.9±0.1 0.7±0.2 0.8±0.2 0.9±0.3 1.0±0.3 L-Alanine 6.4±0.7 2.6±0.4 2.8±0.5 3.8±0.5 3.5±0.8 AIB 4.3±0.6 2.0±0.6 2.0±0.4 2.7±0.6 3.1±1.0 L-Histidine 5.4±0.5 1.9±0.4 2.0±0.3 2.5±0.3 2.0±0.3 L-Leucine 5.3±0.2 3.5±0.4 3.5±0.2 3.7±0.7 3.6±0.2 L-Methionine 9.0±0.1 5.3±0.9 4.6±0.2 4.2±0.1 5.0±1.0 D-Glucose 39.5±3.2 36.9±1.9 33.3±1.8 44.7±3.2 37.4±0.7

Table 3.7 Rates of amino acid transport in R. leguminosarum strains grown on glucose/glutamate. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures.

Substrate Strain 3841 RU542 RU543 RU634 RU636 L-Glutamate 26.9±2.9 5.1±1.0 5.4±1.1 3.8±0.4 3.9±2.0 L-Aspartate 18.9±0.8 2.1±0.4 2.0±0.3 1.7±0.2 2.1±0.1 L-Alanine 28.8±1.7 6.0±1.2 4.9±1.2 5.3±1.3 6.3±1.1 AIB 37.8±4.3 12.0±2.3 9.3±0.7 10.0±2.4 11.5±0.7 L-Histidine 23.4±2.0 6.3±0.2 6.0±0.8 5.1±0.3 5.5±0.3 L-Leucine 27.2±6.5 10.1±1.1 8.6±0.8 8.3±0.5 9.8±0.4 L-Methionine 15.9±0.6 6.7±0.1 6.1±0.1 6.2±0.2 6.8±0.3 D-Glucose 42.4±0.6 22.8±2.1 17.8±0.8 16.3±1.6 18.3±0.7

Page 132: The reasonable man attempts to adapt ... - reading.ac.uk

121

Uptake of glutamate, aspartate and histidine by the mutants is severely

affected. This suggests that the general permease is the main high affinity

transport system for these amino acids in strain 3841. Alanine, AIB, leucine and

methionine uptake is much less significantly affected and the existence of

additional, specific systems for transporting these substrates seems likely from

previous kinetic studies (Poole et al., 1985).

Initially, the uptake results from cells grown with glutamate as the nitrogen

source (i.e. nitrogen-limited) seem to suggest a greater contribution by the

general permease to the transport of alanine, AIB, leucine and methionine.

However, glucose uptake in the mutants is also substantially reduced under

these conditions, presumably because cells lacking the major glutamate uptake

system struggle to grow on glutamate. It is therefore uncertain whether the low

amino acid uptake rates in mutants grown on glutamate are due directly to the

loss of the general amino acid uptake system, or whether a significant

proportion of the reduction in uptake is attributable to a global decrease in

transport.

3.2.10 Growth of strain RU543 on amino acids as sole source of carbon and nitrogen

Strain 3841 will grow on minimal salts agar containing alanine, histidine or

proline as the sole source of carbon and nitrogen. Growth of strain RU543 on

minimal salts agar containing each of these amino acids as sole carbon and

nitrogen source was therefore investigated, in an attempt to assess the

contribution of the general amino acid permease to the uptake of these

substrates in strain 3841. The results are shown in Table 3.8. Growth of strain

RU543 on the carbon/nitrogen sources glucose/NH4Cl and glutamate was

tested as a positive and a negative control, respectively.

Page 133: The reasonable man attempts to adapt ... - reading.ac.uk

122

Table 3.8 Growth of R. leguminosarum strain RU543 on individual amino acids as sole source of carbon and nitrogen. Growth medium was AMS agar with C/N sources added at 20mM except for glucose and NH4Cl which were at 10mM. +++, good growth; ++, moderate growth; +, poor growth; -, no growth. C/N Source Strain 3841 RU543 Glucose/NH4Cl +++ +++ Glutamate +++ - Alanine ++ ++ Histidine + + Proline +++ +

The ability to grow on proline is significantly reduced in strain RU543

compared to that in strain 3841. This is clearly not due to the mutation in strain

RU543 causing a general impairment of growth, since growth on

glucose/NH4Cl is unaffected. It therefore seems likely that the retardation in

growth of strain RU543 is due to reduced proline uptake, suggesting that the

general amino acid permease is responsible for a significant proportion of

proline uptake in strain 3841 under these conditions.

Growth of strain RU543 on both alanine and histidine is unimpaired. In the

case of alanine this is consistent with the earlier evidence (Section 3.2.9; (Poole et

al., 1985)) for the existence of at least one alanine carrier in strain 3841 besides

the general amino acid permease. However, for histidine the result is

somewhat surprising considering the extent of the reduction in transport of this

substrate in strain RU543 (Section 3.2.9). It is noticeable that the wild-type

strain itself grows only poorly on histidine (giving growth comparable to that of

strain RU543 on proline), so one explanation could be that uptake of nutrient is

not the rate determining step in growth on histidine, and the reduced histidine

uptake in the mutant is still sufficient to support growth at the wild-type rate.

Another possibility is that in addition to the general amino acid permease strain

Page 134: The reasonable man attempts to adapt ... - reading.ac.uk

123

3841 possesses another histidine carrier that is either of low affinity (and hence

insignificant under transport assay conditions where the substrate is only

25µM), or only induced under certain metabolic conditions, such as those

arising from growth on histidine as sole carbon and nitrogen source.

3.2.11 Amino acid uptake in strain RU640

In order to further investigate the broad specificity of the cloned amino acid

permease, the amino acid transport properties of strain RU640 (which contains

additional copies of the aap operon carried by pRU191) grown on

glucose/NH4Cl were studied (Table 3.1).

For the amino acids tested, strain RU640 shows a 4- to 16-fold increase in

uptake compared to the wild-type. This includes aliphatic amino acids such as

leucine and alanine, in addition to polar amino acids such as glutamate,

aspartate and histidine. That this increase is not caused by the presence of the

vector pRK415-1 is demonstrated by the uptake values for strain RU913 (strain

3841/pRK415-1) which are very similar to those of strain 3841. In addition,

glucose transport in strain RU640 shows no significant increase, indicating that

enhancement of uptake is restricted to amino acids.

Strain RU640 grows extremely poorly on glucose/glutamate and uptake

rates for both glutamate and glucose are found to be lower than wild-type

under these conditions (data not shown). This apparent unhealthiness of strain

RU640 is thought to be due to the intolerably high number of copies of the

amino acid transporter that are likely to be present in cells carrying additional

copies of the aap operon under non-repressing conditions.

3.2.12 Expression of the R. leguminosarum general amino acid permease in E. coli

Following the discovery that the four proteins that comprise the general

amino acid permease in R. leguminosarum show high homology to four

Page 135: The reasonable man attempts to adapt ... - reading.ac.uk

124

unidentified proteins from E. coli, it was decided to investigate the effect of the

aap operon, expressed via the lac promoter, on amino acid transport in E. coli.

E. coli strain JC5412, which lacks appreciable glutamate uptake, was

transformed with pRU310 DNA and a TetR transformant purified, to generate

strain RU1050. Glutamate, Histidine and AIB uptake by strain RU1050 grown

in both LB, and LB containing IPTG, was measured (Table 3.9).

Table 3.9 Rates of amino acid transport in E. coli strains JC5412 and RU1050 grown on LB. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Substrate Strain JC5412 RU1050 RU1050+IPTG L-Glutamate 0.74±0.02 1.06±0.21 1.22±0.10 L-Histidine 0.64±0.24 0.91±0.02 1.16±0.13 AIB 0.02±0.01 0.80±0.41 1.61±0.81

Uptake of all three amino acids is elevated in strain RU1050 in comparison to

strain JC5412, with further elevation of uptake apparent in strain RU1050 grown

in the presence of IPTG. Transport of AIB is particularly informative since in

the absence of the general amino acid permease uptake of this substrate is

barely detectable, while in its presence AIB uptake is similar to that of both

glutamate and histidine. This is consistent with functional expression of the

general amino acid permease from R. leguminosarum strain 3841 in E. coli strain

JC5412.

Page 136: The reasonable man attempts to adapt ... - reading.ac.uk

125

3.2.13 Physical properties of the aapJ gene product

Periplasmic proteins from strains 3841, RU542, RU543, RU640, RU913 and

RU918 grown phosphate-limited on glucose/NH4Cl were isolated by

lysosyme/EDTA treatment and subjected to SDS-PAGE (Fig. 3.13).

Fig. 3.13 SDS-PAGE gel of periplasmic proteins from strains 3841, RU542, RU543, RU640, RU913 and RU918.

The gel shows a band at approximately 33kD in the 3841, RU542 and RU913

lanes, that is greatly enhanced in the RU640 and RU918 lanes and is apparently

missing from the RU543 lane. Since strains RU640 and RU918 (Section 3.2.14)

carry additional copies of aapJ, while strains 3841, RU542 and RU913 carry only

the chromosomal copy of aapJ, and strain RU543 is an aapJ mutant, this is

compatible with the band at ~33kD being AapJ. The apparent size of this

Page 137: The reasonable man attempts to adapt ... - reading.ac.uk

126

presumed aapJ gene product is slightly smaller than the 34.5kD that is expected

from the sequence, however, a discrepancy between the calculated molecular

weight and the molecular weight determined on an SDS gel is not unusual

(Noel et al., 1979; Higgins et al., 1982b).

In order to confirm that the protein extracts run on the gel in Fig. 3.13

contained only periplasmic proteins, the activity of the cytoplasmic marker

enzyme malate dehydrogenase was measured in these extracts and compared

to the activity found in the cell fractions remaining after removal of the

periplasmic extracts. Activities of the periplasmic marker enzyme alkaline

phosphatase were also measured (Table 3.10).

Table 3.10 Activities of cytoplasmic and periplasmic marker enzymes in periplasmic protein fractions of R. leguminosarum strains prepared for SDS-PAGE. Activities are expressed as nmol min-1 for the total sample of cells, equivalent to 20ml of culture A600 1.0. Values are the result from a single experiment. UD, undetectable. Strain and Malate dehydrogenase Alkaline phosphatase fraction assayed Activity % of total Activity % of total 3841 Periplasm UD UD 1141 53.7 French-pressed cells 862 100.0 984 46.3 Total 862 100.0 2125 100.0 RU542 Periplasm UD UD 353 18.3 French-pressed cells 823 100.0 1579 81.7 Total 823 100.0 1932 100.0 RU543 Periplasm UD UD 302 17.5 French-pressed cells 753 100.0 1421 82.5 Total 753 100.0 1723 100.0 RU640 Periplasm UD UD 1071 51.3 French-pressed cells 997 100.0 1018 48.7

Page 138: The reasonable man attempts to adapt ... - reading.ac.uk

127

Total 997 100.0 2089 100.0

Page 139: The reasonable man attempts to adapt ... - reading.ac.uk

128

Since no malate dehydrogenase activity was observed in the periplasmic

fractions, whereas this activity was high in the French pressed samples, it can

be concluded that the samples run on the gel in Fig. 3.13 do not contain

intracellular proteins. The alkaline phosphatase data indicate that periplasmic

proteins were released from 17-54% of cells in the preparation of these samples.

The RU640 lane of the gel in Fig. 3.13 was scanned using a LKB Ultroscan XL

densitometer. The optical density of the aapJ band relative to the total optical

density of the lane, indicated that AapJ constitutes ~15% of the total periplasmic

protein in strain RU640.

3.2.14 Effect of aapJ on amino acid uptake in 3841

Since the increase in amino acid uptake by 3841 strains carrying additional

copies of the aap operon might be due solely to the binding of increased

quantities of substrate by the additional periplasmic binding protein without

attendant transport, the effect of aapJ on amino acid uptake was studied.

The 1.8kb insert in pRU256, which contains aapJ and its upstream promoter

region, was excised as a BamHI-KpnI fragment and cloned in both pRK415-1

and pIJ1891, creating pRU388 and pRU389, respectively. (Plasmid pRU256 is a

deletion of pRU190 generated for sequencing). In pRU389, but not pRU388,

transcription of aapJ can be initiated by the lac promoter in the vector. Both

pRU388 and pRU389 were conjugated into strain 3841, creating strains RU918

and RU919, respectively, and glutamate uptake by these strains was measured

(Table 3.11). (That expression of aapJ carried by pRU388 is not affected by the

single base pair deletion in this plasmid that will have been inherited from

pRU190 (Section 3.2.3), is demonstrated by the enhanced AapJ band in the

RU918 lane of the SDS-PAGE gel in Fig. 3.13, and by the ability of pRU192 to

enhance amino acid uptake in strain 3841 (Section 3.2.3)).

Page 140: The reasonable man attempts to adapt ... - reading.ac.uk

129

Table 3.11 Rates of L-glutamate transport in R. leguminosarum strains grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Uptake Host Plasmid 3841 Wild-type None 5.6±0.4 RU913 Wild-type Vector only 4.9±0.2 RU640 Wild-type aapJQMP 43.9±4.3 RU918 Wild-type aapJ 4.4±0.4 RU915 Wild-type aapQMP 4.5±0.1 RU914 Wild-type Vector only 5.3±0.7 RU916 Wild-type aapJQMP 23.0±1.2 RU919 Wild-type aapJ 2.6±0.3 RU917 Wild-type aapQMP 5.1±0.5

In contrast to strains RU640 and RU916 (strain 3841/pRU310), neither RU918

or RU919 shows an increase in glutamate uptake in comparison to strains

RU913 and RU914 (strain 3841/pIJ1891), respectively. This indicates that in the

wild-type the rate of substrate transport by the membrane components of the

carrier is not limited by the amount of periplasmic binding protein available for

ferrying substrate. Crucially, the additional substrate binding capacity

provided by increased amounts of binding protein does not produce an

apparent increase in uptake. The latter conclusion is confirmed by the

observation that apparent glutamate uptake is not increased in either of the

strains RU543 or RU542 carrying either pRU388 or pRU389 (data not shown).

3.2.15 Effect of aapQMP on amino acid uptake in strains 3841

In order to determine whether additional copies of the membrane

components of the general amino acid permease alone are capable of increasing

amino acid uptake in strain 3841, pRU309 and pRU313 were conjugated into

3841, creating strains RU915 and RU917, respectively.

Page 141: The reasonable man attempts to adapt ... - reading.ac.uk

130

Neither strain shows an increase in glutamate uptake in comparison to the

strain 3841 containing the corresponding vector alone (Table 3.11). That

pRU309 produces no effect is unsurprising since in this plasmid aapQ and aapM

are promoterless (Fig. 3.12). However, in pRU313 aapQMP are expressed via

the lac promoter in the vector (as demonstrated in the complementation

experiments described in Section 3.2.7), so a wild-type value for glutamate

uptake by strain RU917 indicates that the presence of additional copies of the

membrane components of the carrier, in the absence of increased amounts of

binding protein, does not lead to an increase in amino acid uptake.

Taken in conjunction with the results from Section 3.2.14, this result

demonstrates that the amino acid uptake rates observed in strain RU640 are due

to the intracellular accumulation of substrate rather than merely the binding of

substrate to individual components of the transporter. It also suggests that in

the wild-type cell the relative quantities of the periplasmic binding protein and

the membrane components of the amino acid permease are optimal.

3.2.16 Specificity of the aapJ gene product

The high amino acid uptake rates for strain RU640, and the wild-type level of

glutamate transport observed for strain RU917, are consistent with the idea that

the functioning of the general amino acid permease is dependent on the ability

of a single periplasmic binding protein (the aapJ gene product) to bind the

complete range of amino acids, rather than there being a number of binding

proteins interacting with the membrane component of the transporter, each of

which is specific to one, or a small number of related substrates. The latter

possibility is well documented for the histidine/LAO system from Salmonella

typhimurium and the LS/LIV-I system from E. coli (Higgins et al., 1982b; Adams

et al., 1990).

To investigate this issue further the uptake of a range of amino acids by

strain RU760 (Fig. 3.12) was measured (Table 3.12).

Page 142: The reasonable man attempts to adapt ... - reading.ac.uk

131

Table 3.12 Rates of amino acid transport in R. leguminosarum strains grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 (mg protein)-1. Values are the mean±SEM of determinations from three or more independent cultures. Substrate Strain 3841 RU543 RU760 L-Glutamate 5.6±0.4 1.2±0.6 1.2±0.6 L-Aspartate 3.9±0.1 0.8±0.2 0.4±0.2 L-Alanine 6.4±0.7 2.8±0.5 3.3±0.2 AIB 4.3±0.6 2.0±0.4 1.6±0.7 L-Histidine 5.4±0.5 2.0±0.3 2.4±0.1 L-Leucine 5.3±0.2 3.5±0.2 2.1±0.5

In strain RU760 the chromosomal copy of the aap operon is not functional

because of the transposon in aapJ, but functional copies of aapQMP are provided

by pRU313. If binding proteins other than the aapJ gene product are supporting

amino acid transport by the aapQMP gene products, strain RU760 would be

expected to exhibit increased uptake of a sub-set of amino acids in comparison

to strain RU543. This assumes that the genes encoding other binding proteins

do not lie immediately downstream of the aap operon where they could

potentially be affected by the transposon in aapJ. However, no significant

difference is observed between strains RU760 and RU543 in the uptake of any of

the amino acids tested.

3.2.17 Substrate-binding activity of AapJ

In an endeavour to demonstrate directly the broad specificity of the

periplasmic binding protein component of the general amino acid permease,

amino acid-binding activity in the periplasmic protein fraction from strains

3841, RU640 and RU543 was studied.

As an initial experiment an attempt was made to assay the substrate-binding

affinity for glutamate using ammonium sulphate precipitation (Richarme &

Page 143: The reasonable man attempts to adapt ... - reading.ac.uk

132

Kepes, 1983; Dahl & Manson, 1985; Higgins et al., 1987). Periplasmic proteins

from each of the strains 3841, RU640 and RU543 were isolated as described in

Section 2.1.6. The crude extracts were dialysed overnight in 5mM HEPES,

pH7.2 and concentrated, before being assayed as described in Section 2.1.7.

However, no glutamate-binding activity was detected by this method in any of

the extracts. This was presumed to be due to the glutamate-binding protein

complex failing to survive the precipitation step.

The second approach to the problem employed non-denaturing

polyacrylamide gel electrophoresis, a technique which has been used

successfully in the demonstration of substrate binding by other periplasmic

proteins (Nobile & Deshusses, 1988; Le Rudulier et al., 1991; Talibart et al., 1994).

Dialysed and concentrated periplasmic protein extracts from strains 3841,

RU640 and RU543 were prepared as above. Each of these extracts was then

incubated with each of the substrates [14C]glutamate, [14C]aspartate,

[14C]alanine, [14C]AIB, [14C]histidine and [14C]leucine, before being subjected to

non-denaturing PAGE and autoradiography as described in Section 2.1.7.

In a trial non-denaturing PAGE gel of crude periplasmic protein extracts, the

presumed Aap binding protein band is clearly visible in strain 3841, absent

from strain RU543, and greatly enhanced in strain RU640 (Fig. 3.14).

Page 144: The reasonable man attempts to adapt ... - reading.ac.uk

133

Fig. 3.14 Non-denaturing PAGE gel of periplasmic proteins from strains 3841, RU543 and RU640.

However, the autoradiographs of the gels run following incubation of

protein extracts with radioactive amino acids indicate no labelling of this band

by any of the substrates tested. Nor was labelling of any other periplasmic

protein apparent except in the case of leucine. In the autoradiograph of the gel

loaded with proteins incubated with leucine, each of the three lanes

corresponding to the different strains exhibits an identical single band,

produced by a protein that runs further on the gel than the Aap binding protein

(Fig. 3.15).

Page 145: The reasonable man attempts to adapt ... - reading.ac.uk

134

Fig. 3.15 Autoradiograph of non-denaturing PAGE gel of periplasmic proteins from strains 3841, RU543 and RU640 pre-incubated with [14C]leucine. The control contained no protein.

It is probable that this band corresponds to the binding protein from a

leucine specific transporter and that the highly specific interaction between this

protein and its substrate is strong enough to enable the protein-substrate

complex to survive the electrophoresis conditions. The same electrophoresis

conditions are apparently too harsh for the Aap binding protein-substrate

complex, which is perhaps not surprising given that a binding protein with a

very broad specificity may have a relatively "loose" binding site.

It seemed possible that the use of less stringent assay conditions than had

been previously employed might allow substrate-binding by the Aap binding

protein to be observed. Unfortunately the equilibrium dialysis method (Argast

& Boos, 1979; May et al., 1986) was not financially feasible if a number of

potential substrates were to be screened. The following approach was therefore

adopted: Periplasmic protein extracts from strains 3841, RU640 and RU543

were each dialysed, concentrated, and incubated with substrates exactly as

described for the non-denaturing gel experiment. The radioactivity in each

Page 146: The reasonable man attempts to adapt ... - reading.ac.uk

135

incubation mixture was determined by the scintillation counting of a small

sample. The incubation mixtures were then each spun through a

microconcentrator with a molecular weight cut-off below the size of AapJ, and

the radioactivity of the filtrates determined. Comparison of the radioactivity

present in the filtrate to that in the incubation mixture prior to spinning,

enabled the amount of substrate bound to protein to be calculated (Table 3.13).

Table 3.13 Binding of amino acids by periplasmic fractions from cells of R. leguminosarum strains grown on glucose/NH4Cl. Binding is expressed as pmol [mg protein]-1. Values are the mean±SEM of three or more replicate assays. UD, undetectable. Substrate Strain 3841 RU543 RU640 L-Glutamate 265±6 UD 1824±100 L-Aspartate 208±93 UD 1315±37 L-Alanine 534±37 535±67 594±35 AIB UD UD UD L-Histidine 357±86 UD 1470±242 L-Leucine 74±58 36±4 173±103 L-Methionine 62±54 22±13 122±28

The results for glutamate, aspartate and histidine demonstrate binding of

each of these substrates by the Aap binding protein. In each case there was

binding of the substrate by the periplasmic protein extract from the wild-type

strain 3841, substantially increased binding in strain RU640 (which carries

additional copies of aapJ), but no binding in strain RU543 (an aapJ mutant).

Leucine- and methionine-binding appear to follow a similar pattern, but the

level of binding is significantly lower. The binding of leucine by the extract

from strain RU543 is consistent with the existence of a leucine-specific binding

protein in strain 3841, as indicated by the non-denaturing PAGE gel of

Page 147: The reasonable man attempts to adapt ... - reading.ac.uk

136

periplasmic proteins incubated with this amino acid. Binding of alanine is

evident for each of the strains 3841, RU640 and RU543. The fact that the level of

alanine binding is the same in each of the three strains indicates that rather than

being due to the Aap binding protein, this binding is presumably caused by one

or more alanine-specific proteins. There was no detectable binding of AIB by

the periplasmic protein extracts from any of the three strains.

These results are in line with the amino acid uptake data for the aap mutants

RU542, RU543, RU634 and RU636 (Section 3.2.9). The level of binding by AapJ

of ligands such as leucine, alanine and AIB, for which the general amino acid

permease is apparently not the only significant carrier, is found to be lower than

that for substrates such as glutamate, aspartate and histidine for which this is

the predominant uptake system. That no binding at all could be detected for

alanine and AIB is perhaps not surprising as kinetic evidence suggests that the

general permease has a lower affinity for these substrates (Poole et al., 1985),

and assay conditions such as pH and ionic strength may not have been optimal

for the binding of these amino acids.

3.2.18 Amino acid exchange

Cells of R. leguminosarum strain 3841 have previously been found to exhibit

high rates of exchange of pre-loaded intracellular amino acids with amino acids

in the surrounding medium (Poole et al., 1985). In an attempt to determine the

role of the general amino acid permease in this process, it was decided to

investigate the effect of both mutation and over expression of the aap operon on

amino acid exchange by strain 3841.

Having been allowed to accumulate [14C]AIB, cells of each of the strains 3841,

RU543 and RU640 were incubated in an excess of AIB or glutamate, or had

carbonyl cyanide m-chlorophenylhydrazone (CCCP) added, and the

intracellular concentrations of [14C]AIB monitored over time. The duration of

pre-incubation with [14C]AIB was varied between strains so that the cells of

each strain contained approximately equal concentrations of the radioactive

Page 148: The reasonable man attempts to adapt ... - reading.ac.uk

137

substrate immediately prior to exchange, thus allowing unambiguous

comparison of efflux rates. AIB was chosen as the exchange substrate in these

experiments because it is not metabolized by R. leguminosarum strain 3841

(Section 3.2.1). Consequently the results are not complicated by incorporation

of radioactivity into cell proteins or loss of label as 14CO2.

In the wild-type strain 3841, the intracellular concentration of [14C]AIB

rapidly decreases on addition of either extracellular AIB or extracellular

glutamate (Fig. 3.16). This decrease is presumed to be the result of exchange

between the intracellular and extracellular amino acid pools (Poole et al., 1985).

In the case of exchange with glutamate, some recovery in the internal

[14C]AIB concentration is observed at longer incubation times. This may be due

to uptake of [14C]AIB by a carrier with affinity for AIB but not glutamate,

probably an alanine transporter, after glutamate has saturated the exchange

system on both sides of the membrane. This effect has been seen before in the

exchange of intracellular leucine with extracellular glutamate (Poole et al., 1985).

In strain RU543 the rate of exchange of internal [14C]AIB with external AIB

(homologous exchange) is lower than that in strain 3841, while loss of internal

[14C]AIB due to the presence of external glutamate (heterologous exchange) is

very slight in comparison to that in the wild-type (Fig. 3.17). By contrast, in

strain RU640 the rates of both homologous and heterologous exchange are

increased relative to those in strain 3841 (Fig. 3.18). The fact that in strain

RU543 homologous exchange is not as significantly affected as heterologous

exchange, points to the existence of at least one additional exchange system

specific to AIB. This is likely to be an alanine transporter.

Page 149: The reasonable man attempts to adapt ... - reading.ac.uk

138

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.16 Efflux of AIB from cells of R. leguminosarum strain 3841. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition; , CCCP addition. The time of addition is indicated by the arrow.

Page 150: The reasonable man attempts to adapt ... - reading.ac.uk

139

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.17 Efflux of AIB from cells of R. leguminosarum strain RU543. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition; , CCCP addition. The time of addition is indicated by the arrow.

Page 151: The reasonable man attempts to adapt ... - reading.ac.uk

140

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.18 Efflux of AIB from cells of R. leguminosarum strain RU640. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition; , CCCP addition. The time of addition is indicated by the arrow.

Page 152: The reasonable man attempts to adapt ... - reading.ac.uk

141

The proton ionophore CCCP has a profound effect on the level of

intracellular [14C]AIB in all three strains (Figs. 3.16, 3.17 and 3.18). The rapid

efflux of AIB from CCCP-poisoned cells suggests that maintaining the internal

concentration of AIB against a concentration gradient is energy dependent.

However, nothing can be deduced about the energy requirements of AIB efflux.

Curtailing energy dependent processes may result in active transport systems

becoming simply "open pores" in the cytoplasmic membrane, through which

substrates are free to diffuse. This idea could explain the observation that

CCCP treatment results in similar very rapid rates of AIB efflux from all three

strains, if it is assumed that the differences in the number of copies of the amino

acid transporter between strains, are insignificant in comparison to the total

number of transporters in each strain. That the cytoplasmic membrane has not

simply fallen apart upon treatment with CCCP is shown by the fact that the

level of intracellular [14C]AIB is not reduced to zero. The residual intracellular

[14C]AIB is lost when cells are treated with CCCP together with an excess of

unlabelled AIB (Fig. 3.19).

That the effect of CCCP is due to the loss of energy dependent processes

rather than some specific property of CCCP itself, is demonstrated by the effect

of the cytochrome oxidase inhibitor CN- on cells loaded with [14C]AIB (Fig.

3.20). It can be seen that although CN- poisoning takes longer to affect cells, it

also results in the efflux of AIB.

Page 153: The reasonable man attempts to adapt ... - reading.ac.uk

142

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.19 Effect of excess extracellular AIB on CCCP-induced efflux of AIB from cells of R. leguminosarum strain RU640. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition; , CCCP + AIB addition. The time of addition is indicated by the arrow.

Page 154: The reasonable man attempts to adapt ... - reading.ac.uk

143

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

20

40

60

80

Fig. 3.20 Efflux of AIB from cells of R. leguminosarum strain RU640. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition; , KCN addition. The time of addition is indicated by the arrow.

Page 155: The reasonable man attempts to adapt ... - reading.ac.uk

144

The observation that amino acid exchange is reduced in strain RU543 and

increased in strain RU640 relative to the wild-type, suggests that the general

amino acid permease is involved in the exchange process. However, it is not

clear whether the efflux, as well as the uptake of amino acids, is facilitated by

the general amino acid permease itself.

Two possible models of the exchange process are illustrated in Fig 3.21.

Fig. 3.21 Schematic representation of two possible mechanisms for exchange of amino acids by R. leguminosarum: A, uptake via the general amino acid permease (AAP) and efflux via a second carrier; B, uptake and efflux via the general amino acid permease.

Page 156: The reasonable man attempts to adapt ... - reading.ac.uk

145

In A (Fig. 3.21), amino acid uptake is catalyzed by the general amino acid

permease, with amino acid efflux being facilitated by an entirely separate

carrier. In B, the general amino acid permease is responsible for both the

uptake and the efflux of amino acids. In both models it is suggested that uptake

and efflux are continuous competing processes, with a net build up of substrate

in the cytoplasm resulting from a greater rate of uptake than efflux. Exchange

is explained by assuming that the added extracellular excess of unlabelled

substrate effectively monopolizes uptake, while labelled substrate continues to

leave the cytoplasm via the efflux system. In the case of heterologous exchange,

the final, steady state intracellular concentrations will be a function of the

relative affinities of the two substrates for the uptake and efflux systems.

It is hard to explain the observed effects on amino acid exchange of varying

the copy number of the general amino acid permease, on the basis of model A

in Fig. 3.21. For example, if this model were correct, then the additional copies

of the general amino acid permease carried by strain RU640, would result in

increased uptake capacity without a concomitant rise in the potential for efflux.

It would therefore be expected that in strain RU640 the intracellular

concentration of the chaser substrate would rise more rapidly than in the wild-

type strain, resulting, in the case of homologous exchange at least, in a faster

rise in competition for the efflux system from the chaser, and hence a lower rate

of exchange. In fact, precisely the opposite is observed in practice (compare

Figs. 3.16 and 3.18), which suggests that the general amino acid permease is

intimately involved in the efflux process.

The mechanism by which exchange might occur in the case where the

general amino acid permease provides both uptake and efflux facilities (B, Fig.

3.21) is not known. It seems clear that the general amino acid permease actively

imports amino acids. However, since previously studied ABC transporters are

reported to be uniporters, catalysing transport of substrates in one direction

only (Higgins, 1992), it seems unlikely that this transporter also energizes amino

Page 157: The reasonable man attempts to adapt ... - reading.ac.uk

146

acid efflux. One alternative possibility is that amino acids simply leak out of

the cytoplasm through the general amino acid permease. This may be

particularly feasible for a relatively small amino acid such as AIB, since the

broad specificity of the amino acid permease presumably necessitates it

providing a relatively "wide" pore in the cytoplasmic membrane.

If efflux of intracellular AIB is the result of diffusion through the general

permease, it might be anticipated that the periplasmic binding protein of the

transporter would be unnecessary for such efflux. In order to investigate this

possibility, exchange of AIB in strain RU760 was examined (Fig. 3.22) and

compared to that in strain RU543.

Strain RU760 carries functional copies of aapQMP but not aapJ (Section 3.2.7),

whereas in strain RU543 aapJQM are not transcribed. Therefore, if the three

proteins AapQ, AapM and AapP alone permit efflux of AIB, the rates of

exchange in strain RU760 are expected to be greater than those in strain RU543.

This does not appear to be the case (compare Figs. 3.17 and 3.22), suggesting

that AapJ is essential for efflux via the general amino acid permease.

AapP also appears to be essential to exchange, as loss of pre-loaded [14C]AIB

from strain RU542 on addition of external glutamate or AIB (Fig. 3.23), occurs at

a similar rate to that observed for strain RU543 (Fig. 3.17).

Page 158: The reasonable man attempts to adapt ... - reading.ac.uk

147

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.22 Efflux of AIB from cells of R. leguminosarum strain RU760. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition. The time of addition is indicated by the arrow.

Page 159: The reasonable man attempts to adapt ... - reading.ac.uk

148

Time (min)

0 10 20 30 40 50 60 70

Intra

cellu

lar l

abel

led

AIB

(nm

ol [m

g pr

otei

n]-1

)

0

10

20

30

40

50

60

70

Fig. 3.23 Efflux of AIB from cells of R. leguminosarum strain RU542. The data shown is the result from a single representative experiment. , No addition; , AIB addition; , L-glutamate addition. The time of addition is indicated by the arrow.

Page 160: The reasonable man attempts to adapt ... - reading.ac.uk

149

3.2.19 Plant properties of strains RU542, RU543, RU634, and RU636

Strains RU542, RU543, RU634, and RU636 all nodulated peas, with 25 nodule

isolates of each strain all retaining kanamycin resistance. Acetylene reduction

rates (Table 3.14) show that each of these strains reduce acetylene, and hence

presumably fix nitrogen, at least as well as the wild-type strain 3841.

Table 3.14 Rates of acetylene reduction by pea nodules harbouring aap and cysE mutants of R. leguminosarum strain 3841 Reduction rates are expressed as µmol h-1 per plant. Values are the mean±SEM of three or more independent determinations. Strain Relevant Acetylene reduction genotype 3841 Wild-type 0.45±0.02 RU542 aapP::Tn5-lacZ 0.63±0.03 RU543 aapJ::Tn5-lacZ 0.56±0.05 RU632 cysE::Tn5-lacZ 0.83±0.01 RU634 aapM::Tn5-lacZ 0.56±0.05 RU636 aapQ::Tn5-lacZ 0.66±0.01

Page 161: The reasonable man attempts to adapt ... - reading.ac.uk

150

3.2.20 Nucleotide sequence of the 0.8kb BamHI fragment of pRU3024

In an attempt to determine the nature of the gene (or genes) mutated in

pRU3031 (and hence strain RU632) and pRU3033 it was decided to sequence as

much of the DNA adjacent to the transposons in these cosmids as was

conveniently possible. The aim being to compare the sequence obtained with

sequence databases.

To this end, BamHI restriction fragments from each of pRU3031 and

pRU3033 were randomly cloned into the BamHI site of pBluescript SK+. In each

case the resulting DNA was used to transform E. coli strain DH5α, and KmR,

AmpR transformants were selected. In this way, one clone containing the

majority of the transposon plus the cosmid DNA from the end of the

transposon to the next BamHI site, was isolated from each of pRU3031 and

pRU3033. These clones are pRU383 and pRU384, respectively (Fig. 3.24).

The insert DNA in each of pRU383 and pRU384 was sequenced using the KS

primer to the pBluescript SK+ polylinker and a primer (P0) to the insertion

sequence of the transposon (Fig. 3.24). Additional sequence was obtained by

sequencing pRU3033 DNA directly using a primer (P15) to the 5' end of Tn5-

B20. The total sequence obtained, much of which is from one strand only, is

shown in Fig. 3.25.

Page 162: The reasonable man attempts to adapt ... - reading.ac.uk

151

Fig. 3.24 Restriction map of the cysE region of pRU3024 and mutants derived from it. Tn5-B20 insertions are labelled with the designation of the mutated cosmid in which they occur. Sub-clones of the two mutated cosmids in pBluescript SK+ are shown below. The unfilled arrow indicates the direction of transcription of cysE. The labelled arrows indicate the primers used to obtain the sequence in Fig. 3.25. Restriction site is: B, BamHI.

Page 163: The reasonable man attempts to adapt ... - reading.ac.uk

152

Page 164: The reasonable man attempts to adapt ... - reading.ac.uk

153

The direction of transcription of the DNA in Fig. 3.25 was determined from

the observation that the fusions in both pRU3031 and pRU3033 are active (Fig.

3.2). Of the three possible reading frames for the sequence, that leading to the

deduced amino acid sequence given in Fig. 3.25 appeared the most likely on the

basis of the occurrence of stop codons. Screening of the GenEmbl sequence

database with this amino acid sequence reveals 60.9%, 60.3% and 40.1% identity

to CysE from S. typhimurium, E. coli and Bacillus subtilis, and 39.7% identity to

NifP from Azotobacter chroococcum. (Screening of the database with the deduced

amino acid sequences from the two other possible reading frames reveals no

significant homologies). Both cysE and nifP encode serine acetyltransferase.

Thus the sequence homology data suggest that the mutated gene in pRU3031

and pRU3033 encodes serine acetyltransferase in R. leguminosarum, particularly

since the gene directly upstream of, and transcribed divergently from, the aap

operon apparently also encodes an enzyme on the methionine biosynthetic

pathway (Sections 3.2.4 and 3.2.23).

3.2.21 Amino acid transport in strain RU632

Both pRU3031 and pRU3033 are substantially impaired in their ability to

increase amino acid uptake in strain 3841 as compared to the parent cosmid

pRU3024 (Table 3.3). However, strain RU632, the chromosomal mutant of

strain 3841 derived from pRU3031 (Section 3.2.6), grown on glucose/NH4Cl,

shows no reduction in the uptake of any of the amino acids tested in

comparison to strain 3841 (Table 3.15).

When grown on glucose/glutamate however, strain RU632 shows

considerably reduced rates of uptake for all the amino acids tested except

methionine, for which uptake is only slightly below the wild-type level (Table

3.15). The relatively high rates of methionine transport in strain RU632 under

these conditions may be due to increased expression of methionine-specific

uptake systems initiated by the low intracellular methionine levels that

Page 165: The reasonable man attempts to adapt ... - reading.ac.uk

154

presumably arise in a methionine biosynthetic mutant under nitrogen-

limitation.

Table 3.15 Rates of amino acid transport in R. leguminosarum strain RU632 grown on glucose/NH4Cl and glucose/glutamate. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Substrate Glucose/NH4Cl Glucose/glutamate 3841 RU632 3841 RU632 L-Glutamate 5.6±0.4 4.6±0.3 26.9±2.9 10.8±2.7 L-Aspartate 3.9±0.1 4.3±1.5 18.9±0.8 5.6±0.9 L-Alanine 6.4±0.7 6.7±0.6 28.8±1.7 12.2±1.1 AIB 4.3±0.6 6.6±0.4 37.8±4.3 16.5±2.1 L-Histidine 5.4±0.5 6.6±2.1 23.4±2.0 10.4±1.7 L-Leucine 5.3±0.2 5.1±1.1 27.2±6.5 8.6±2.1 L-Methionine 9.0±0.1 9.0±0.5 15.9±0.6 11.7±0.5 D-Glucose 39.5±3.2 44.1±1.8 42.4±0.6 25.2±2.6

Transport of glucose is also impaired in strain RU632 grown on

glucose/glutamate, but not to the same degree as that of amino acids. It

therefore appears that although some of the reduction in amino acid transport

in strain RU632 under nitrogen-limited conditions can be attributed to a global

effect of the mutation, there is also a specific effect on the general amino acid

permease. This is consistent with the observation that pRU3031 and pRU3033

do not increase amino acid transport in strain 3841 to the same degree as

pRU3024 (Table 3.3). However, it is not clear why a chromosomal mutation of

cysE has no effect on amino acid uptake under nitrogen-excess conditions when,

under the same conditions, mutation of this gene in pRU3024 reduces the ability

of this cosmid to increase amino acid uptake.

The role of serine acetyltransferase in amino acid uptake in strain 3841 is

obscure. It seems unlikely that a reduction in methionine biosynthesis per se is

Page 166: The reasonable man attempts to adapt ... - reading.ac.uk

155

responsible for the effect on the general amino acid permease, since loss of beta-

cystathionase is expected to have at least as much of an effect on methionine

production as inactivation of serine acetyltransferase (Fig. 3.26), and metC

mutants of strain 3841 exhibit wild-type amino acid transport (Section 3.2.24).

Unlike a metC mutation, a cysE mutation is expected to be detrimental to

cysteine biosynthesis (Fig. 3.26), and it is plausible that a change in the

intracellular cysteine concentration has an effect on the nitrogen regulation of

amino acid transport. However, neither the cysE or metC mutants are

methionine or cysteine auxotrophs. Indeed, it is interesting that CysE shows

homology to NifP, and it is possible that R. leguminosarum, like Azotobacter

(Evans et al., 1991), possesses two genes encoding serine acetyltransferase. Thus

it may be the case in strain 3841, that one gene encoding serine acetyltransferase

enables sufficient cysteine biosynthesis for normal amino acid transport under

nitrogen-excess, but that expression of two serine acetyltransferase genes is

required to allow a wild-type response of amino acid transport to nitrogen-

limitation.

3.2.22 Plant properties of RU632

Strain RU632 nodulated peas, and all of 25 nodule isolates retained

kanamycin resistance. The acetylene reduction rate for this strain is shown in

Table 3.14, and from comparison of this to the wild-type rate, it is apparent that

the mutation in RU632 does not cause a decrease in acetylene reduction, and

hence presumably nitrogen fixation.

Page 167: The reasonable man attempts to adapt ... - reading.ac.uk

156

Fig. 3.26 Methionine biosynthesis in microorganisms (Bender, 1978).

3.2.23 Nucleotide sequence adjacent to the transposon in cosmids pRU3053, pRU3082, pRU3083, pRU3084, pRU3085 and pRU3086

In the hope of identifying R. leguminosarum strain 3841 genes lying between

aapJ and cysE, nucleotide sequence adjacent to the transposon in pRU3053,

pRU3082, pRU3083, pRU3084, pRU3085 and pRU3086 was obtained by

sequencing cosmid DNA directly using a primer (P15) to the 5' end of Tn5-B20.

Translations of these sequences were screened for homology to the deduced

proteins from the GenBank and EMBL sequence databases.

Sequences obtained from pRU3053, pRU3082, pRU3084 and pRU3086 were

found to overlap, and were combined to produce one continuous length of

sequence. In addition, bases 1-351 of the sequence from pRU189 (Fig. 3.5), a

translation of which shows homology to MetC (Section 3.2.4), overlap the

sequence from pRU3084. Combination of the cosmid sequence with the 351 bp

Page 168: The reasonable man attempts to adapt ... - reading.ac.uk

157

from pRU189 yields the nucleotide sequence in Fig. 3.27. The deduced

polypeptide from this sequence (Fig. 3.27) shows 41.5% and 37.7% identity to

beta-cystathionase from E. coli and Bordetella avium, respectively. It therefore

seems likely that the gene directly upstream of aapJ in strain 3841 is the R.

leguminosarum equivalent of metC. This gene is transcribed divergently from

aapJ.

The sequences from pRU3083 and pRU3085 did not show significant

homology to any of the sequences in the databases searched.

Page 169: The reasonable man attempts to adapt ... - reading.ac.uk

158

Page 170: The reasonable man attempts to adapt ... - reading.ac.uk

159

Page 171: The reasonable man attempts to adapt ... - reading.ac.uk

160

3.2.24 Amino acid transport in metC mutants of strain 3841

In contrast to the cysE mutant RU632, the metC mutants RU639, RU999,

RU1000 and RU1001 generated by homogenotization of cosmids pRU3053,

pRU3082, pRU3084 and pRU3086, respectively, in strain 3841, exhibit no

significant reduction in glutamate transport under either nitrogen-excess or

nitrogen-limited conditions (Table 3.16).

Table 3.16 Rates of glutamate uptake by metC mutants of R. leguminosarum strain 3841 grown on glucose/NH4Cl and glucose/glutamate. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the result from a single experiment except those for strains 3841 and RU639 which are the mean±SEM of determinations from three or more independent cultures. Strain Growth conditions Glucose/NH4Cl Glucose/glutamate 3841 5.6±0.4 26.9±2.9 RU639 4.5±0.1 21.0±0.6 RU999 4.7 30.5 RU1000 4.8 29.5 RU1001 4.4 27.8

It is interesting that attempts to make mutants of strain 3841 by

homogenotization of pRU3083 and pRU3085 were unsuccessful, suggesting that

mutation of the gene(s) carrying the transposon in these cosmids may be lethal.

Page 172: The reasonable man attempts to adapt ... - reading.ac.uk

161

3.3 DISCUSSION

The sequence homology data strongly suggests that the products of the four

complete R. leguminosarum strain 3841 genes carried by pRU189 comprise the

components of an ABC transporter of amino acids. A mutation in any one of

these four genes leads to a decrease in the transport of all of the amino acids

tested, and the ability of this transporter to carry a broad range of amino acids is

further demonstrated by the uptake rates for strain RU640. It is therefore

reasonable to conclude that the genes carried by pRU189 encode a general

amino acid permease, and the gene designation aap is suggested.

The general amino acid permease appears to be the main high-affinity uptake

system for glutamate, aspartate, histidine and proline in free-living cells of R.

leguminosarum strain 3841. This transporter also accounts for a significant

percentage of alanine, leucine and methionine uptake, although there are

clearly other specific uptake systems for these substrates, as indicated by earlier

kinetic studies (Poole et al., 1985).

The uptake of amino acids by strain RU760 and the results of substrate

binding assays on periplasmic fractions from cells of strains 3841, RU543 and

RU640, suggest that the general amino acid permease utilizes a single

periplasmic binding protein of broad specificity (the aapJ gene product). Such

broad specificity is unusual, and the protein, of apparent molecular weight

~33kD, shows little overall homology to any previously reported periplasmic

binding protein. In particular, AapJ shows no homology to OppA, the

periplasmic binding protein from the oligopeptide transporter of S. typhimurium

(Hiles and Higgins, 1986; Hiles et al., 1987), which also shows broad substrate

specificity, but is considerably larger (52kD) than AapJ.

The rate of exchange by strain 3841 of intracellular AIB with extracellular

AIB or glutamate appears to be directly related to the number of copies of the

aap operon present, and it is hard to envisage a mechanism for exchange in

which the general amino acid permease provides only uptake functions, that

Page 173: The reasonable man attempts to adapt ... - reading.ac.uk

162

can account for this observation. Since ABC transporters are reported to be

uniporters (Higgins, 1992), the most likely possibility seems to be that the

general amino acid permease actively facilitates amino acid uptake but also

allows passive efflux. If this is the case, then the pore formed by the transporter

must be accessible to intracellular substrate. Whether access to the pore can be

gained from the cytoplasm at all times, or is dependent upon the binding

and/or uptake of extracellular substrate, is a critical question for the

functioning of this and possibly other ABC transporters.

The requirement of AapJ for efflux suggests that binding of external

substrate is necessary for this process. An alternative possibility is that in the

wild-type unliganded binding protein interacts with the membrane complex,

and the absence of AapJ results in a change in the conformation of the

membrane components that prevents efflux. However, this seem unlikely since

the binding protein of the histidine uptake system of S. typhimurium appears

only to interact with the membrane components of the transporter in its

liganded form (Prossnitz et al., 1988; Prossnitz et al., 1989).

While the reduction in amino acid exchange in an aapP mutant of strain 3841

may indicate that ATP hydrolysis is required to fuel efflux, it seems more likely

that the energy is needed to open a membrane pore upon binding of a liganded

binding protein complex. It is also possible that the loss of AapP alters the

conformation of the remaining components of the transporter so that efflux can

no longer occur, particularly since it is generally assumed that AapP induces a

conformational change in the transmembrane subunits upon ATP hydrolysis

(Higgins, 1992; Doige & Ames, 1993).

The physiological significance of the exchange capability of the general

amino acid permease is unknown. Clearly the conditions employed in the

exchange experiments described here are unlikely to be encountered in nature,

however, relatively high intracellular concentrations of substrates may

ordinarily be present in cells as a result of active accumulation or endogenous

Page 174: The reasonable man attempts to adapt ... - reading.ac.uk

163

synthesis. Consequently, intracellular substrate will flow out of the cell down a

concentration gradient if suitable pores are present in the cytoplasmic

membrane. If such pores are provided by an uptake system upon

binding/uptake of an external substrate, then there is potential for regulated

exchange of substrates, in which the passage of substrates across the membrane

is determined by the differential affinities of the substrates for a common

carrier.

Certainly the exchange capacity of the general amino acid permease is not

required for nitrogen fixation by strain 3841, as aap mutants induce pea nodules

that reduce acetylene as effectively as those induced by the wild-type. Since the

general amino acid permease appears to be the major high affinity glutamate

transporter in strain 3841, this observation suggests that nitrogen fixation in

Rhizobium leguminosarum is not fuelled via a malate-aspartate shuttle in the

nodule as proposed by Kahn et al. (1985). However, the possibility that

alternative glutamate/aspartate transporters are induced under symbiotic

conditions means that the operation of this shuttle can not be completely

discounted.

In the case of Rhizobium leguminosarum it is alanine rather than aspartate that

is found to be excreted from bacteroids and symbiosomes under nitrogen-fixing

conditions (Appels & Haaker, 1991; Rosendahl et al., 1992). If this excretion of

alanine is the primary mechanism for maintaining sufficient flux through the

TCA cycle to fuel nitrogen fixation in the bacteroid, then the fact that the

general amino acid permease is not required for an effective symbiosis is

unsurprising: Alanine uptake by aap mutants indicates the existence of one or

more alternative alanine carriers, and the results of exchange experiments on

strain RU543, using the alanine analogue AIB, suggest that such a carrier (or

carriers) allows efflux of substrate.

The striking homology between the deduced amino acid sequences of the aap

genes and those of four unidentified ORFs from the 67.4 to 74.0 minute region

Page 175: The reasonable man attempts to adapt ... - reading.ac.uk

164

of the E. coli K12 chromosome, suggests that E. coli possesses a general amino

acid permease. Similarly, the fact that the aapJ gene product shows extremely

high homology to the polypeptide deduced from what is available of the

sequence of an unidentified ORF in P. fluorescens, but shows little homology to

other known binding proteins, makes it likely that a homologue of this

transporter is also found in Pseudomonas.

E. coli strain D2W is reported to have five separate systems for dicarboxylic

amino acid transport (Schellenberg & Furlong, 1977). These are: (i) a binding-

protein-independent, Na+-dependent, glutamate specific system; (ii) a binding-

protein-dependent, Na+-independent system for transport of glutamate and

aspartate; (iii) a binding-protein-independent, Na+-independent glutamate-

aspartate system; (iv) a binding-protein-independent, aspartate-specific system;

and (v) a dicarboxylic acid transport system that carries aspartate in addition to

malate, fumarate and succinate. Genes corresponding to systems (i) and (iii)

from E. coli K12 and/or E. coli B have been cloned and sequenced (Deguchi et

al., 1989; Deguchi et al., 1990; Kalman et al., 1991; Wallace et al., 1990; Tolner et

al., 1992b), and the nucleotide sequence of three genes encoding the membrane

components of a binding-protein-dependent glutamate-aspartate transporter

(corresponding to (ii) above) from E. coli K12 has been entered in the EMBL

database (accession number U10981, B. Wallace (1994)). However, the latter

sequence is not the same as that found in the 67.4 to 74.0 minute region of the

K12 chromosome, and although the deduced ATP-binding protein from this

sequence shows 60.2% identity to AapP, the other two membrane components

show less than 30% identity to either AapQ or AapM. It therefore seems

possible that the general amino acid permease-like transporter is a previously

uncharacterized system for glutamate and aspartate uptake in E. coli.

Since Aap-like proteins are so highly conserved between Escherichia coli,

Pseudomonas fluorescens and Rhizobium leguminosarum it seems reasonable to

conclude that this transporter has an important physiological role, and may be

Page 176: The reasonable man attempts to adapt ... - reading.ac.uk

165

common to a wide range of Gram-negative bacteria. However, the fact that

mutation of aap genes is not lethal to R. leguminosarum, and that the Aap-like

system in E. coli has apparently gone undetected, suggests that this system is

not prominent under laboratory culture conditions.

TFASTA and BLAST searches indicate homology of AapQ and AapM to the

integral membrane components of ABC transporters of polar amino acids or

glutamine. In addition to the "integral membrane component signature"

(Saurin et al., 1994), a sequence alignment of AapQ and AapM with these

proteins, reveals a region of 63 amino acids containing 18 highly conserved

residues, that is located, in the majority of cases, at the N-terminal end of the

protein. This suggests that amino acid transporters of this type may constitute a

sub-family of the ABC superfamily.

The presence of two conserved regions in proteins belonging to this sub-

family, provides a constraint for topological models, since such regions are

expected to occur in similar locations (i.e. periplasmic, transmembrane, or

cytoplasmic) in each protein. Predicted topologies containing eight

transmembrane segments in the case of the general permeases, and five

membrane-spanning regions in the case of the other known members of the

sub-family, are consistent with this constraint. Indeed, a five-transmembrane-

segment topology for HisM and HisQ from S. typhimurium has been confirmed

experimentally (Kerppola et al., 1991; Kerppola & Ames, 1992). In these

topologies the N-terminal conserved region spans two transmembrane

segments and a connecting cytoplasmic loop, while the integral membrane

component signature is located in a cytoplasmic loop.

The N-terminal conserved region may be involved in substrate specificity,

since for one position in this sequence there appears to be a correlation between

the nature of the substrate translocated, and the nature of the amino acid found

at that position; while deletion of four other amino acids from this N-terminal

Page 177: The reasonable man attempts to adapt ... - reading.ac.uk

166

region in HisM of S. typhimurium, results in a change of substrate specificity

from L-histidine to L-histidinol (Payne et al., 1985).

On the basis of sequence homology, the gene lying immediately upstream of

the aap operon in R. leguminosarum strain 3841, is likely to be metC, encoding

beta-cystathionase. This gene is transcribed divergently from the aap operon,

and assuming it is of a similar size to its E. coli counterpart, its 3' end lies

approximately 2.4kb upstream of the probable start of a gene, designated cysE,

which yields a deduced polypeptide with significant homology to serine

acetyltransferase. Both these enzymes are involved in methionine biosynthesis.

However, the β-galactosidase activities produced by Tn5-lacZ mutants of

pRU3024 in which the transposon has inserted between the 3' end of metC and

the start of cysE (pRU3083 and pRU3085), are consistent with transcription of a

gene(s) in the opposite orientation to metC and cysE (Fig. 3.2), suggesting that

these genes are not contained in one operon.

Mutation of cysE, but not metC, results in a reduction in uptake by the

general amino acid permease under nitrogen-limitation. Such a reduction in

transport might occur if the aap operon is regulated in response to a metabolite

level(s) that is altered as a consequence of the mutation of cysE, but not metC.

The fact that a mutation in cysE only affects amino acid uptake under nitrogen-

limitation, is compatible with the possible existence of two genes encoding

serine acetyl transferase in R. leguminosarum, as suggested by the homology of

CysE to NifP from Azotobacter: The expression of one gene encoding serine

acetyltransferase may be sufficient to allow wild-type amino acid transport

under nitrogen excess, however the level of serine acetyltransferase resulting

from expression of two genes, may be required to enable a normal response of

amino acid transport to nitrogen-limitation.

Page 178: The reasonable man attempts to adapt ... - reading.ac.uk

167

CHAPTER 4 NITROGEN REGULATION OF THE GENERAL AMINO ACID PERMEASE OF RHIZOBIUM LEGUMINOSARUM STRAIN 3841

Page 179: The reasonable man attempts to adapt ... - reading.ac.uk

168

4.1 INTRODUCTION

It has previously been reported that glutamate uptake by R. leguminosarum

strain 3841 is significantly lower under conditions of nitrogen-excess than it is

under nitrogen-limitation (Poole et al., 1985; Poole et al., 1987). The cloning and

mutation of the genes encoding the general amino acid permease of strain 3841,

described in Chapter 3, has shown that this transporter is the major high-

affinity uptake system for glutamate in this strain.

It was therefore decided to investigate the regulation of the aap operon in

response to nitrogen supply. In particular, the role of the nitrogen response

regulator NtrC was examined. The results are discussed in this chapter.

Page 180: The reasonable man attempts to adapt ... - reading.ac.uk

169

4.2 RESULTS

4.2.1 Effect of the metC-aapJ intergenic region on growth of strain 3841

For the purpose of monitoring transcription of the aap operon in R.

leguminosarum strain 3841, a lacZ-fusion to the N-terminus of aapJ and its

upstream promoter region was created by cloning a 1.0kb EcoRI-PstI fragment

from pRU189 into the corresponding sites in the polylinker of the expression

vector pMP220.

The resulting plasmid, pRU393 (Fig. 4.1), was conjugated into strain 3841,

creating strain RU990. The intention was to measure β-galactosidase activity in

strain RU990 grown on both glucose/NH4Cl and glucose/glutamate. However,

the presence of pRU393 appears to be extremely detrimental to strain 3841, and

growth of strain RU990 on minimal media was deemed too poor to yield any

meaningful results.

Fig. 4.1 Map of pRU383 showing the location and orientation of the promoterless lacZ gene in the vector.

Since strain RU368 (3841/pMP220) exhibits no difference in growth to that of

the wild-type, it can be concluded that it is the additional copies of the

intergenic region between metC and aapJ, carried by pRU393, that are

deleterious to the growth of strain RU990. A possible reason for this effect is

that one or more transcriptional regulators that normally bind to the native

Page 181: The reasonable man attempts to adapt ... - reading.ac.uk

170

metC-aapJ intergenic region, and perhaps other promoter regions in the

chromosome, are bound to plasmid-borne intergenic regions in RU990, leading

to potentially harmful changes in the regulation of certain genes in this strain.

Indeed, it has previously been observed that over expression of the aap operon

can cause poor growth of strain 3841 (Section 3.2.11).

4.2.2 Effect of nitrogen supply on the transcription of aapJQM

Since the aapJ::lacZ reporter fusion, pRU393, could not be used, it was

decided to use the aapJ::lacZ, aapQ::lacZ and aapM::lacZ fusions in cosmids

pRU3028, pRU3046 and pRU3035, respectively, to study transcription of the aap

operon. The disadvantage to the use of these fusions is the potentially

consequential introduction into the host strain of additional copies of the other

genes carried by the cosmid. Indeed, all these cosmids necessarily contain the

metC-aapJ intergenic region. However, unlike pRU393, none of these cosmids

appears to affect the growth of strain 3841, suggesting that the smaller numbers

of the metC-aapJ intergenic region introduced into the host by the cosmids, in

comparison to pRU393 (due to the lower copy number of the cosmids), do not

have such a significant impact on the cell.

Strains RU443, RU506 and RU517, generated by conjugating cosmids

pRU3028, pRU3035 and pRU3046, respectively, into strain 3841, were grown

under both nitrogen-excess (glucose/NH4Cl) conditions and nitrogen-limited

(glucose/glutamate) conditions, and β-galactosidase activities measured in each

case (Table 4.1).

The transcription-attenuating effect of the putative stem-loop located

between aapJ and aapQ is evident in the relative level of β-galactosidase activity

produced by pRU3028 (aapJ::lacZ) compared to that from pRU3035 (aapM::lacZ)

and pRU3046 (aapQ::lacZ) under the same growth conditions. However, all

three fusions exhibit a similar degree of repression under nitrogen-excess

conditions.

Page 182: The reasonable man attempts to adapt ... - reading.ac.uk

171

Table 4.1 Effect of ntrC on plasmid-borne aap::lacZ fusion activity in R. leguminosarum strain 3841. Strains were grown on AMS containing either glucose/NH4Cl or glucose/glutamate at 10mM. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Growth conditions Host Cosmid Glucose/NH4Cl Glucose/glutamate RU438 (3841/pRU3024) Wild-type aapJQMP 40±5 43±1 RU443 (3841/pRU3028) Wild-type aapJ::Tn5-lacZ 3325±279 7923±666 RU506 (3841/pRU3035) Wild-type aapM::Tn5-lacZ 752±208 2103±184 RU517 (3841/pRU3046) Wild-type aapQ::Tn5-lacZ 1221±265 2853±331 RU1013 (RU929/pRU3024) ntrC::Ω aapJQMP 34±1 34±2 RU980 (RU929/pRU3028) ntrC::Ω aapJ::Tn5-lacZ 7278±546 8256±1738 RU983 (RU929/pRU3035) ntrC::Ω aapM::Tn5-lacZ 2113±527 1484±309 RU984 (RU929/pRU3046) ntrC::Ω aapQ::Tn5-lacZ 2096±517 2121±350

Page 183: The reasonable man attempts to adapt ... - reading.ac.uk

172

Lower rates of uptake by the general amino acid permease of strain 3841

grown on glucose/NH4Cl compared to those of the same strain grown on

glucose/glutamate, have been previously reported for both batch cultures

(Poole et al., 1985) and chemostat cultures (Poole et al., 1987). However, the

reduction in transcription (~3-fold) between glutamate-grown and NH4Cl-

grown cells indicated by the results in Table 4.1, is significantly less than the

reduction in glutamate transport observed between corresponding batch

cultures (~5-fold) and chemostat cultures (~17-fold). This discrepancy is

discussed later in this Section.

In order to investigate whether the repression of the aap operon under

nitrogen-excess conditions is mediated by ntrC, cosmids pRU3028, pRU3035

and pRU3046 were conjugated into strain RU929, an ntrC interposon mutant of

strain 3841. The resulting strains, RU980, RU983 and RU984, respectively, were

grown on both glucose/NH4Cl and glucose/glutamate, and β-galactosidase

activities measured in each case (Table 4.1).

After nitrogen-limited growth, the activities of all three fusions is similar in

the strain RU929 background to that in the strain 3841 background. However,

under nitrogen-excess conditions, transcription of aapJQM in strain RU929

occurs at a similar level to that obtaining under nitrogen-limitation. This

suggests that NtrC may negatively regulate the aap operon in R. leguminosarum

strain 3841. Certainly the data in Table 4.1 are indicative of the involvement in

some way of NtrC in the regulation of the aap operon, since both nitrogen-

limitation, which results in the phosphorylation of NtrC, and mutation of NtrC,

result in a similar increase in the transcription of aapJQM from that found in a

nitrogen-excess wild-type background.

In view of the effect of pRU393 on the growth of strain 3841, it was

considered that the presence of additional copies of the metC-aapJ intergenic

region in the strains in Table 4.1 may have affected the β-galactosidase activities

Page 184: The reasonable man attempts to adapt ... - reading.ac.uk

173

observed in these strains. It was therefore decided to study the activities of a

lacZ fusion to the chromosomal copy of each of aapJQM.

To this end, strains RU1017, RU1018 and RU1019, containing mutations in

both ntrC and aapJ, aapM or aapQ, respectively, were generated by

homogenotization of pRU3028, pRU3035 and pRU3046 in strain RU929. The β-

galactosidase activity resulting from the Tn5-lacZ mutations in aapJQM in these

strains, and in strains RU543, RU634 and RU636 (which contain only the

corresponding single mutations of aapJ, aapM and aapQ, respectively), was

investigated after growth on both glucose/NH4Cl and glucose/glutamate

(Table 4.2).

Table 4.2 Effect of ntrC on chromosomal aap::lacZ fusion activity in R. leguminosarum strain 3841. Strains were grown on AMS containing either glucose/NH4Cl or glucose/glutamate at 10mM. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Growth conditions Glucose/NH4Cl Glucose/glutamate 3841 Wild-type 15±2 22±3 RU543 aapJ::Tn5-lacZ 155±23 690±39 RU634 aapM::Tn5-lacZ 50±5 309±46 RU636 aapQ::Tn5-lacZ 75±9 325±50 RU929 ntrC::Ω 27±1 23±1 RU1017 ntrC::Ω aapJ::Tn5-lacZ 603±106 678±49 RU1018 ntrC::Ω aapM::Tn5-lacZ 250±38 278±40 RU1019 ntrC::Ω aapQ::Tn5-lacZ 352±49 344±52

β-galactosidase activities in strains RU543, RU634 and RU636 exhibit the

expected repression of the aap operon under nitrogen-excess. However, in

contrast to the results obtained from cosmid-borne fusions, the approximately

5-fold difference in activity between glucose/NH4Cl and glucose/glutamate

Page 185: The reasonable man attempts to adapt ... - reading.ac.uk

174

cultures of these strains, is consistent with the scale of repression demonstrated

by amino acid uptake in batch cultures.

One potential problem in the interpretation of the data in Table 4.2 is the

possibility that strains mutated in their major high-affinity glutamate uptake

system may not be simply nitrogen-limited when grown on glutamate as the

sole nitrogen source. However, the observation that glucose/glutamate

cultures of the mutant strains in Table 4.2 show no significant difference in

growth to those of their parental strains, suggests that mutation of aap genes has

little effect on the utilization of glutamate as the sole nitrogen source in strain

3841. Furthermore, addition of glutamate to chemostat cultures of strain 3841

growing nitrogen-limited on NH4Cl has been previously shown to result in a

slight increase in glutamate uptake (Poole et al., 1987). It therefore seems

unlikely that the loss of glutamate uptake in aap mutants leads to an increase in

the transcription of the aap operon.

The ntrC-aapJQM double mutants exhibit the complete loss of repression of

the aap operon under nitrogen-excess previously observed for cosmid-borne

fusions in a strain RU929 background. The similarity of the β-galactosidase

activities in these mutants grown on glucose/NH4Cl to the corresponding

activities following glucose/glutamate growth, would be hard to explain if the

activity observed in glucose/glutamate cultures was affected by the reduced

glutamate uptake in these strains.

The ratio between the nitrogen-limited and nitrogen-excess levels of

transcription of the aap operon as indicated by the activity of chromosomal

fusions (Table 4.2), is more compatible with the observed rates of glutamate

transport under the two conditions (Poole et al., 1985), than that obtained from

cosmid fusions (Table 4.1). In the case of the cosmid fusions, the ratio of

nitrogen-limited to nitrogen-excess values is too low. This leads to the

conclusion that either the β-galactosidase activities in strains RU443, RU506 and

RU517 grown on glucose/NH4Cl are artificially high or the activities in

Page 186: The reasonable man attempts to adapt ... - reading.ac.uk

175

glucose/glutamate grown RU443, RU506, RU517, RU980, RU983 and RU984,

and glucose/NH4Cl grown RU980, RU983 and RU984 are artificially low. (β-

galactosidase activities from the cosmid fusions are expected to be greater than

those from the chromosomal fusions due to their greater copy number). If the

aap operon is negatively regulated by the binding of NtrC to the upstream

intergenic region, then the binding of the finite pool of NtrC in the cell to an

increased number of metC-aapJ intergenic regions carried by the cosmids in

RU443, RU506 and RU517, might be expected to result in the partial

derepression of the amino acid permease in these strains under nitrogen-excess.

This would account for relatively high β-galactosidase activities in strains

RU443, RU506 and RU517 grown on glucose/NH4Cl.

4.2.3 Amino acid uptake in strain RU929

As an alternative measure of the effect of the mutation of ntrC on the

transcription of the aap operon in R. leguminosarum strain 3841, glutamate

uptake in strain RU929 was determined under both nitrogen-excess and

nitrogen-limited conditions (Table 4.3).

Table 4.3 Rates of glutamate uptake R. leguminosarum strain RU929 grown on glucose/NH4Cl and glucose/glutamate. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant Growth conditions genotype Glucose/NH4Cl Glucose/glutamate 3841 Wild-type 5.6±0.4 26.9±2.9 RU929 ntrC::Ω 21.4±1.3 22.4±3.7

The elevation of glutamate transport in strain RU929 grown on

glucose/NH4Cl to a similar level to that found in strain 3841 grown on

Page 187: The reasonable man attempts to adapt ... - reading.ac.uk

176

glucose/glutamate, is consistent with the aapJQM::lacZ fusion data discussed in

Section 4.2.2.

4.2.4 Effect of nitrogen supply on the transcription metC and cysE

Since mutation of cysE has an effect on amino acid uptake in strain 3841

(Section 3.2.21), it was decided to investigate the nitrogen regulation of this

gene. The nitrogen regulation of metC, which is transcribed divergently to the

aap operon, and like cysE, encodes an enzyme involved in methionine

biosynthesis, was also investigated.

Cosmids pRU3031 and pRU3033, carrying Tn5-lacZ mutated copies of cysE,

and cosmids pRU3082 and pRU3086, carrying Tn5-lacZ mutated copies of metC,

were conjugated into both strain 3841 and strain RU929. β-galactosidase

activities were measured in the resulting strains grown on both glucose/NH4Cl

and glucose/glutamate (Table 4.4).

β-galactosidase activities from corresponding chromosomal fusions were

also studied. Strains RU632, RU999 and RU1001 containing Tn5-lacZ mutations

in cysE, metC and metC, respectively, were generated by homogenotization of

pRU3031, pRU3082 and pRU3086 in strain 3841. (The cysE mutant derived

from pRU3033 could not be isolated). The ntrC-metC double mutants RU1029

and RU1030 were generated by the homogenotization in strain RU929 of

pRU3082 and pRU3086, respectively; while homogenotization of pRU3033 in

strain RU929 generated the ntrC-cysE double mutant RU1027 (homogenotes of

pRU3031 in RU929 could not be isolated). β-galactosidase activities in these

strains were measured under both nitrogen-excess and nitrogen-limited

conditions (Table 4.5).

Page 188: The reasonable man attempts to adapt ... - reading.ac.uk

177

Table 4.4 β-galactosidase activities in R. leguminosarum strains 3841 and RU929 containing pRU3031, pRU3033, pRU3080 and pRU3086. Strains were grown on AMS containing either glucose/NH4Cl or glucose/glutamate at 10mM. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Growth conditions Host Cosmid Glucose/NH4Cl Glucose/glutamate RU438 (3841/pRU3024) Wild-type metC cysE 40±5 43±1 RU502 (3841/pRU3031) Wild-type cysE::Tn5-lacZ 878±189 1313±156 RU504 (3841/pRU3033) Wild-type cysE::Tn5-lacZ 1324±103 2096±103 RU974 (3841/pRU3082) Wild-type metC::Tn5-lacZ 967±183 1188±64 RU978 (3841/pRU3086) Wild-type metC::Tn5-lacZ 1010±140 1133±107 RU1013 (RU929/pRU3024) ntrC::Ω metC cysE 34±1 34±2 RU981 (RU929/pRU3031) ntrC::Ω cysE::Tn5-lacZ 1213±150 1285±134 RU982 (RU929/pRU3033) ntrC::Ω cysE::Tn5-lacZ 2118±168 2162±267 RU986 (RU929/pRU3082) ntrC::Ω metC::Tn5-lacZ 1197±129 1256±409 RU988 (RU929/pRU3086) ntrC::Ω metC::Tn5-lacZ 1055±47 1378±34

Page 189: The reasonable man attempts to adapt ... - reading.ac.uk

178

Table 4.5 β-galactosidase activities in R. leguminosarum strains RU632, RU999, RU1001, RU1026, RU1027, RU1029 and RU1030. Strains were grown on AMS containing either glucose/NH4Cl or glucose/glutamate at 10mM. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Growth conditions Glucose/NH4Cl Glucose/glutamate 3841 Wild-type 15±2 22±3 RU632 cysE::Tn5-lacZ 139±10 269±23 RU999 metC::Tn5-lacZ 91±4 193±10 RU1001 metC::Tn5-lacZ 78±9 172±31 RU929 ntrC::Ω 27±1 23±1 RU1027 ntrC::Ω cysE::Tn5-lacZ 289±28 309±44 RU1029 ntrC::Ω metC::Tn5-lacZ 190±32 152±9 RU1030 ntrC::Ω metC::Tn5-lacZ 181±19 201±51

The regulation of cysE and metC in response to nitrogen supply results in a

similar pattern for the transcription of these genes to that observed for the aap

operon (Section 4.2.2). The data from the chromosomal fusions indicate

repression of both cysE and metC under nitrogen-excess, though the magnitude

of this effect (an approximately 2-fold reduction in transcription) is less than

that observed for the aap operon. However, as was the case for the aap operon,

mutation of ntrC results in the loss of this repression, suggesting that NtrC

negatively regulates transcription of cysE and metC.

The difference between the nitrogen-excess and nitrogen-limited levels of

transcription of cysE and metC as indicated by the β-galactosidase activities

produced by the cosmid-borne fusions to these genes in the strain 3841

background (Table 4.4), is not as great as that indicated by the chromosomal

fusions. This may be explained by the presence of additional copies of the

metC-aapJ intergenic region in strains containing the cosmids, leading to

derepression of genes negatively regulated by NtrC, as was suggested to

Page 190: The reasonable man attempts to adapt ... - reading.ac.uk

179

account for the discrepancy between the chromosomal and cosmid fusion data

for transcription of the aap operon (Section 4.2.2).

The difference in β-galactosidase activities produced by pRU3031 and

pRU3033 under the same conditions (Table 4.4) appears to reflect the relative

distances of the fusions in these cosmids from the N-terminus of cysE (Fig. 3.24),

and may indicate occasional incomplete transcription of this gene.

4.2.5 Sequence analysis of the metC-aapJ intergenic region

The nucleotide sequence between the methionine start of metC and that of

aapJ, contains a putative NtrC binding site (Fig 4.2). This binding site could

potentially regulate both promoters.

Page 191: The reasonable man attempts to adapt ... - reading.ac.uk

180

Page 192: The reasonable man attempts to adapt ... - reading.ac.uk

181

4.3 DISCUSSION

The data presented in this chapter suggest that the aap operon of R.

leguminosarum strain 3841 is negatively regulated by NtrC. The results from

both glutamate uptake assays and aapJQM::lacZ transcriptional fusion studies

show that the aap operon is fully derepressed in an ntrC mutant of strain 3841

under both nitrogen-excess and nitrogen-limited conditions. This is in contrast

to the wild-type strain in which nitrogen-excess leads to an approximately 5-

fold repression of the aap operon relative to the nitrogen-limited state. In

addition, a putative NtrC-binding site can be identified in the metC-aapJ

intergenic region.

While transcriptional activation by NtrC, under nitrogen-limitation, of genes

encoding amino acid transporters in enteric bacteria has been observed (Ames

& Nikaido, 1985; Nohno et al., 1986; Schmitz et al., 1988; Claverie-Martin &

Magasanik, 1991), negative regulation of an amino acid uptake system by NtrC

has not been reported previously. However, negative regulation by NtrC in R.

leguminosarum is not unprecedented, as NtrC has been found to repress

transcription of the operon containing ntrBC in this bacterium (Patriarca et al.,

1993).

Details of the mechanism by which NtrC regulates aap gene expression in

response to nitrogen supply have not been investigated. Any model has to

account for the fact that both loss of NtrC (through mutation), and

phosphorylation of NtrC (the presumed result of nitrogen-limitation), lead to

the derepression of the aap operon.

One potential explanation is that phosphorylated NtrC (NtrC-P) has a

significantly lower affinity for the binding sites that impair transcription of

aapJQMP than the unphosporylated form. Thus the phosphorylation of NtrC

under nitrogen-limiting conditions would lead to increased expression of these

genes. However, this seems unlikely, particularly since in K. pneumoniae, NtrC-

P is reported to bind more effectively as a negative regulator of ntrBC

Page 193: The reasonable man attempts to adapt ... - reading.ac.uk

182

expression, than NtrC (MacFarlane & Merrick, 1987). An alternative possibility

is that NtrC-P, but not NtrC, has a lower affinity for the binding sites upstream

of aapJQMP, than it does for those involved in controlling the expression of

other genes, such as glnII, which has been shown to be positively regulated by

NtrC (Carlson et al., 1987; Martin et al., 1988; de Bruijn et al., 1989; Rossi et al.,

1989; Shatters et al., 1989; Patriarca et al., 1992). In this case, if the total

concentration of NtrC (phosphorylated and unphosphorylated) in the cell is

largely unaffected by the nitrogen supply, then under nitrogen-limitation, the

amount of NtrC (in the form of NtrC-P) available to bind to the promoter region

of aapJQMP will be reduced, and transcription of these genes consequently

increased. The observation that, in contrast to enteric bacteria where NtrC

activates its own transcription under nitrogen-limited conditions (Reitzer &

Magasanik, 1987; Merrick, 1988), transcription of ntrC in R. leguminosarum is

essentially independent of nitrogen status (Patriarca et al., 1993; Amar et al.,

1994), is therefore consistent with this model.

The presence of additional, plasmid-borne, copies of the metC-aapJ intergenic

region in strain 3841, results in severe retardation of growth. This effect might

be caused by the binding of one or more regulators to the extra binding sites

provided by the additional intergenic regions. However, the binding of NtrC to

such additional binding sites is unlikely to be solely responsible for the effect,

since the introduction into strain 3841, of pAR36A, which carries an NtrC-

binding site upstream of the glnII promoter of R. leguminosarum in the same

plasmid as carried the metC-aapJ intergenic region, is not detrimental to growth.

This suggests that at least one other regulator may bind to the metC-aapJ

intergenic region.

The cysE::lacZ and metC::lacZ fusion data suggest that both cysE and metC are

negatively regulated by NtrC in response to nitrogen supply, presumably in a

similar fashion to the aap operon.

Page 194: The reasonable man attempts to adapt ... - reading.ac.uk

183

CHAPTER 5 INTER-REGULATION OF THE TCA CYCLE AND THE GENERAL AMINO ACID PERMEASE OF R. LEGUMINOSARUM STRAIN 3841.

Page 195: The reasonable man attempts to adapt ... - reading.ac.uk

184

5.1 INTRODUCTION

The excretion of the amino acids alanine and aspartate by bacteroids under

nitrogen-fixing conditions has been widely reported (Kretovich et al., 1986;

Appels & Haaker, 1991; Kouchi et al., 1991; Rosendahl et al., 1992), and a

possible role for this excretion in the regulation of the TCA cycle in the

bacteroid has been discussed (Section 1.2.5).

The results from amino acid exchange experiments presented in chapter 3,

demonstrated the potential for the involvement of the general amino acid

permease from R. leguminosarum strain 3841 in the excretion of amino acids.

However, the Fix+ phenotype of aap mutants suggests either that the general

amino acid permease is not significantly involved in excretion by the bacteroid

or that such excretion is not necessary for bacteroid function.

In this chapter the isolation of α-ketoglutarate dehydrogenase and succinyl-

CoA synthetase mutants of R. leguminosarum strain 3841, which exhibit

dramatically reduced rates of uptake by the general amino acid permease, is

described. The results of experiments performed to investigate this apparent

inter-regulation of the TCA cycle and the general amino acid permease are

discussed.

Page 196: The reasonable man attempts to adapt ... - reading.ac.uk

185

5.2 RESULTS

5.2.1 Aspartate resistant mutants of R. leguminosarum strain 3841

Prior to the isolation of cosmid pRU3024, one strategy employed to obtain

amino acid transport mutants of R. leguminosarum was the screening of

transposon mutants of strain 3841 for resistance to a lethal concentration of a

toxic amino acid analogue. The logic being that at least some of the mutants

able to survive such toxic conditions might derive their immunity from an

inability to take up the poison. This approach has been used successfully to

obtain amino acid transport mutants in other gram-negative bacteria (Ames et

al., 1977; Kay, 1971; Weiner & Heppel, 1971; Oxender, 1972; Halpern, 1974;

Schellenberg & Furlong, 1977; Masters & Hong, 1981; Payne et al., 1985; Dila &

Maloy, 1986; Yamato et al., 1990).

In the case of R. leguminosarum however, common toxic amino acid

analogues were ineffective: The aspartate analogue β-hydroxyaspartate which

is toxic to E. coli and has been used to obtain mutants deficient in aspartate

transport in this organism (Kay, 1971; Schellenberg & Furlong, 1977) was found

not to be toxic to strain 3841 at financially viable concentrations. γ-

Glutamylhydrazide, an analogue of glutamine which strongly inhibits 5-amino

imidazole ribonucleotide synthetase (Schroeder et al., 1969), an enzyme of the

purine biosynthetic pathway, has been used at concentrations of 190-500µM to

select for glutamine transport mutants in E. coli (Weiner & Heppel, 1971;

Masters & Hong, 1981). Unfortunately, this analogue appears to be unstable in

acid minimal salts and is unable to survive in this media long enough to do

more than retard the growth of R. leguminosarum over the first 1-2 days, even

when added at concentrations in excess of 10mM. Some of the larger colonies

that grew on plates containing 10mM γ-glutamylhydrazide, which had been

inoculated with Tn5 mutants of strain 3841, were picked and the strains

purified. However, none of these strains exhibited a phenotype suggestive of

mutation of the general amino acid permease, and were not studied further.

Page 197: The reasonable man attempts to adapt ... - reading.ac.uk

186

Following the lack of success with toxic analogues, the possibility of using an

amino acid itself as the toxic agent was investigated. This type of approach has

been employed successfully in the isolation of dicarboxylate transport (dct)

mutants of R. leguminosarum by selecting for growth on 100mM succinate

(Glenn & Brewin, 1981).

The toxicity of the amino acids L-alanine, L-aspartate, L-glutamate and L-

serine towards R. leguminosarum strain 3841 was tested (Table 5.1).

Table 5.1 Growth of R. leguminosarum strain 3841 in the presence of increasing concentrations of four different amino acids. Growth medium was AMS agar with glucose/NH4Cl added at 10mM as the C/N source. +++, good growth; ++, moderate growth; +, poor growth; -, no growth.

Concentration Amino acid L-Alanine L-Aspartate L-Glutamate L-Serine

10mM +++ +++ +++ +++ 25mM + ++ +++ +++ 50mM - ++ +++ ++ 100mM - - ++ -

Although alanine shows the greatest toxicity, kinetic evidence suggests that

the general amino acid permease is not the only significant high affinity

transporter for this substrate in strain 3841 (Poole et al., 1985). Mutation of the

general amino acid permease is therefore less likely to allow escape from

alanine toxicity than it is to allow escape from a toxic level of a substrate for

which the general amino acid permease is the major transporter. Consequently,

it was decided to attempt to isolate amino acid permease mutants by selecting

for growth on 100mM aspartate.

Subsequent work in this laboratory (P.S.Poole, unpublished data) has shown

that at concentrations of 10mM or above, aspartate is transported at appreciable

Page 198: The reasonable man attempts to adapt ... - reading.ac.uk

187

rates by the dct system of R. leguminosarum. This suggests that mutation of an

aspartate-carrying amino acid transporter is unlikely to allow escape from

aspartate toxicity, and in fact strain RU543 exhibits the same susceptibility to

aspartate as the wild-type strain (Table 5.2).

Table 5.2 Growth of R. leguminosarum strains RU543 and 3841 in the presence of increasing concentrations of L-aspartate. Growth medium was AMS agar with glucose/NH4Cl added at 10mM as the C/N source. +++, good growth; ++, moderate growth; +, poor growth; -, no growth. Aspartate Strain concentration 3841 RU543 0mM +++ +++ 50mM ++ ++ 100mM - -

However, at the time, the use of aspartate as a toxic agent for selecting amino

acid permease mutants seemed reasonable. Tn5 mutants of strain 3841 were

grown on acid minimal salts agar containing glucose/NH4Cl as the

carbon/nitrogen source, 100mM aspartate, and kanamycin at 80 µg ml-1. The

increased concentration of kanamycin in this medium was found to be

necessary as 100mM aspartate appears to have an inhibitory effect on the action

of this antibiotic. Fifty colonies able to grow on this medium were purified and

the uptake of aspartate and glucose by the resulting strains was investigated.

The results for a representative sample are shown in Table 5.3.

The selection criteria used to isolate these mutants is not expected to yield

only mutants deficient in amino acid uptake. There will presumably be other

ways in which the cell can overcome the toxic effect of aspartate, such as having

reduced aspartate catabolism (if it is a product of this catabolism that causes

death) or increased metabolism (to use up excess aspartate and/or products of

Page 199: The reasonable man attempts to adapt ... - reading.ac.uk

188

its catabolism). Indeed, a proportion of the mutants in Table 5.3 show no

difference in aspartate uptake to the wild-type. Nevertheless, in many of the

mutants in Table 5.3 aspartate uptake is significantly impaired. However, in

every case where there is a reduction in aspartate transport there is a

concomitant, though not so severe, reduction in glucose transport. This is not

the phenotype expected to result from a mutation of the general amino acid

permease and suggests that these mutants are altered in a gene (or genes) with

a more global impact.

Table 5.3 Rates of aspartate and glucose transport in glucose/NH4Cl grown aspartate toxic escape Tn5 mutants of R. leguminosarum strain 3841. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the result from a single experiment except those for strains 3841, RU116 and RU156 which are the mean±SEM of determinations from three or more independent cultures. Strain Substrate L-Aspartate D-Glucose 3841 3.9±0.1 39.5±3.2 RU116 0.6±0.1 19.5±0.6 RU117 4.3 40.3 RU118 4.3 39.6 RU126 2.3 26.0 RU137 1.5 26.2 RU140 3.5 37.5 RU151 2.6 31.7 RU154 0.1 14.1 RU156 0.2±0.1 12.4±1.1 RU158 0.1 11.1

However, since the effect of the mutation on amino acid uptake in some of

these strains is so severe (in strain RU156 for example, aspartate transport is less

than 5% of that in the wild-type strain), and the effect appears to be particularly

marked for amino acid transport (glucose transport in strain RU156 is more

Page 200: The reasonable man attempts to adapt ... - reading.ac.uk

189

than 30% of that in strain 3841), it was decided that some of these mutants

warranted further investigation. As a representative sample of those mutants in

which amino acid uptake is impaired, strains RU116, RU137 and RU156 were

chosen for further study. Some properties of strain RU118, which overcomes

aspartate toxicity without a reduction in aspartate transport were also

investigated.

5.2.2 Growth of strains RU116, RU118, RU137 and RU156 on succinate and glucose

Since strains RU116 and RU156 were found to grow slowly on plates relative

to the wild-type, strains RU116, RU118, RU137 and RU156 were further

characterized by measuring their growth rates on glucose/NH4Cl and

succinate/NH4Cl (Table 5.4).

Table 5.4 Growth rates of aspartate toxic escape Tn5 mutants of R. leguminosarum strain 3841 on glucose/NH4Cl and succinate/NH4Cl. Growth rates are expressed as mean generation times in min. Values are the result from a single experiment. Carbon source Strain 3841 RU116 RU118 RU137 RU156 Glucose 270 405 230 255 990 Succinate 220 315 210 205 305

Growth of strains RU116 and RU156 on glucose is significantly slower than

that of the wild-type strain 3841. However, both mutants are partially rescued

by growth on succinate This suggests that mutations in strains RU116 and

RU156 are more deleterious to glucose catabolism than they are to that of

succinate.

Page 201: The reasonable man attempts to adapt ... - reading.ac.uk

190

The increased growth rate of strain RU118 on glucose relative to that of strain

3841 suggests that the mutation in this strain may result in an enhanced

metabolic rate under these conditions.

5.2.3 Amino acid transport in strains RU116 and RU156

In order to further characterize the effect of the mutation in each of RU116

and RU156 on amino acid transport, uptake of AIB, alanine and glutamate by

these strains grown on glucose/NH4Cl was investigated. Since a significant

difference is apparent in the growth of these strains on glucose compared to

that on succinate (Section 5.2.2), uptake of glutamate and aspartate was also

measured in cells grown on succinate/NH4Cl (Table 5.5).

Page 202: The reasonable man attempts to adapt ... - reading.ac.uk

191

Table 5.5 Rates of amino acid transport in R. leguminosarum strains RU116 and RU156 grown on glucose/NH4Cl and succinate/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. ND, not determined. Growth conditions Substrate and strain L-Aspartate L-Glutamate L-Alanine AIB D-Glucose Succinate Glucose/NH4Cl 3841 3.9±0.1 5.6±0.4 6.4±0.7 4.3±0.6 39.5±3.2 ND RU116 0.6±0.1 0.8±0.3 2.0±0.2 1.2±0.1 19.5±0.6 ND RU156 0.2±0.1 0.3±0.1 1.2±0.1 0.1±0.0 12.4±1.1 ND Succinate/NH4Cl 3841 3.1±0.1 3.6±0.4 ND ND ND 46.7±0.2 RU116 1.3±0.1 1.4±0.4 ND ND ND 31.2±0.6 RU156 0.6±0.1 0.5±0.1 ND ND ND 22.7±3.9

Page 203: The reasonable man attempts to adapt ... - reading.ac.uk

192

The uptake of all the amino acids tested is significantly reduced in strains

RU116 and RU156. Since AIB transport is affected, it appears that this is not a

metabolic effect. Although the reductions in glutamate and aspartate transport

in RU116 and RU156 cells grown on succinate are less than those found in

glucose grown cultures of these strains, the impairment of amino acid uptake is

significantly greater than that of succinate in each case. Furthermore, the effect

on succinate uptake in the succinate grown mutants is less than that on glucose

uptake in the same strains grown on glucose. This suggests that the partial

rescuing of amino acid uptake observed in strains RU116 and RU156 grown on

succinate is the result of a global effect within the cell rather than a specific

effect on amino acid uptake.

5.2.4 Transductional analysis of strains RU116, RU118, RU137 and RU156

Mutations caused by the insertion of Tn5 are stable, and strongly polar (Berg,

1977). If a single copy of Tn5 is present in the genome of a mutant strain, and if

the insertion of this element is solely responsible for the mutant phenotype,

then the mutant phenotype is expected to be 100% cotransducible with the Tn5

kanamycin resistance marker (Beringer et al., 1978). This expectation applies

not only in cases where Tn5 inactivates a single structural gene, but also in

those cases where a single insertion of Tn5 in a polycistronic operon exerts a

polar effect on a larger number of genes.

Therefore, in order to confirm that the insertion of Tn5 was responsible for

the phenotype of the chosen aspartate toxic escape mutants (rather than a

concurrent spontaneous mutation), the kanamycin resistance marker of each of

strains RU116, RU118, RU137 and RU156 was transduced to strain 3841. The

purified transductants were tested for growth on 100mM aspartate and for the

presence of the high-level streptomycin resistance allele of strain 3841. This can

be distinguished from the low-level streptomycin resistance which is encoded

by Tn5 (Putnoky et al., 1983) by plating on medium containing streptomycin at

500 µg ml-1. Five transductants from each mutant strain were each found to be

Page 204: The reasonable man attempts to adapt ... - reading.ac.uk

193

resistant to 100mM aspartate. In addition, aspartate and glucose uptake in the

transductants RU216, RU237 and RU256 from strains RU116, RU137 and

RU156, respectively, were measured. All were found to have transport rates for

these substrates similar to those of their respective transductional donor strains

(Table 5.6).

Table 5.6 Rates of aspartate and glucose transport in transductants from R. leguminosarum strains RU116, RU137 and RU156, grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the result from a single experiment except those for strains 3841, RU116 and RU156 which are the mean±SEM of determinations from three or more independent cultures. ND, not determined. Strain Substrate L-Aspartate D-Glucose 3841 3.9±0.1 39.5±3.2 RU116 0.6±0.1 19.5±0.6 RU216 0.9 17.0 RU137 1.5 26.2 RU237 1.5 ND RU156 0.2±0.1 12.4±1.1 RU256 0.2 11.8

From these results it was concluded that the phenotypes of strains RU116,

RU137 and RU156 are each tightly linked to a single Tn5 insertion.

5.2.5 Nucleotide sequence adjacent to the transposon in strains RU116, RU137 and RU156

In an attempt to determine the identity of the gene(s) mutated in strains

RU116, RU137 and RU156 it was decided to sequence the chromosomal DNA

either side of the Tn5 insert in each case, and screen the resulting sequences

against sequence databases.

Chromosomal DNA from each strain was isolated, restricted with EcoRI and

the resulting restriction fragments randomly cloned into the EcoRI site of

Page 205: The reasonable man attempts to adapt ... - reading.ac.uk

194

pBluescript SK+. In each case the resulting DNA was used to transform E. coli

strain MC1061, and KmR, AmpR transformants were selected. In this way

clones pRU32, pRU99 and pRU34, containing the approximately 10kb, 9kb and

10kb Tn5-bearing EcoRI fragment from the chromosome of strains RU116,

RU137 and RU156, respectively, were isolated.

Since a primer to the insertion sequence of Tn5 was to be used for sequencing

the DNA adjacent to the transposon, it was necessary to create sub-clones of

pRU32, pRU99 and pRU34 containing only one end of Tn5 (and hence only one

copy of the insertion sequence). Tn5 possesses internal BamHI and HindIII sites,

while in the pBluescript polylinker the BamHI and HindIII sites lie either side of

the EcoRI site. Since the DNA inserts in pRU32 and pRU34 contain either no

BamHI or HindIII sites other than those located within Tn5 (pRU34) or only one

additional BamHI site located between the transposon and the BamHI site in the

polylinker (pRU32), BamHI and HindIII digestion of these plasmids followed by

ligation yielded in each case, two sub-clones, each containing a different end of

Tn5 together with the adjacent chromosomal DNA (Fig. 5.1). The existence of

BamHI sites in the insert DNA between the transposon and the HindIII site in

the polylinker in pRU99 means that BamHI digestion and ligation will not yield

a sub-clone containing a Tn5 insertion sequence. In this case the required sub-

clone, pRU101, was obtained by randomly cloning BamHI restriction fragments

from pRU99 into pBC SK+, transforming E. coli strain DH5α with the resulting

DNA and selecting a KmR, CmR transformant. (This strategy makes use of the

fact that the required BamHI fragment carries the complete kanamycin

resistance gene of Tn5). Plasmid pRU100, the other sub-clone of pRU99 used

for sequencing, was provided by HindIII digestion and ligation.

Page 206: The reasonable man attempts to adapt ... - reading.ac.uk

195

Fig. 5.1 Creation of sub-clones of transposon clones from strains RU116 and RU156, for use in sequencing. Restriction sites are: B, BamHI; E, EcoRI; H, HindIII. *, Site present only in pRU32 and pRU37.

Page 207: The reasonable man attempts to adapt ... - reading.ac.uk

196

Sequencing pRU36/pRU37, pRU101/pRU100 and pRU40/pRU41 using a

primer (P0) to the insertion sequence of Tn5 provided the nucleotide sequence

either side of the transposon in strains RU116, RU137 and RU156, respectively

(Figs. 5.2-5.4).

The GenBank and EMBL databases were searched for sequences showing

homology to each of the six possible translations of each sequence. The

translation shown in Fig. 5.2 of the nucleotide sequence from strain RU116

shows 53.7%, 51.6% and 45.5% identity to SucD, the alpha subunit of succinyl-

CoA synthetase, from E. coli, Coxiella burnetti and Thermus aquaticus,

respectively. In addition, the translation shown in Fig. 5.4 of the nucleotide

sequence from strain RU156 shows 76.5%, 71.4%, 69.4%, 69.4% and 58.1%

identity to α-ketoglutarate dehydrogenase (encoded by sucA or an equivalent

gene) from E. coli, Bacillus subtilis, C. burnetti, Azotobacter vinelandii and Homo

sapiens, respectively. These data suggest that both RU116 and RU156 are TCA

cycle mutants.

The translation shown in Fig. 5.3 of the nucleotide sequence from strain

RU137 shows 60.9% identity to poly-beta-hydroxybutyrate synthase (encoded

by phbC) from R. meliloti. This suggests that biosynthesis of poly-beta-

hydroxybutyrate (PHB) is blocked in strain RU137.

Page 208: The reasonable man attempts to adapt ... - reading.ac.uk

197

Page 209: The reasonable man attempts to adapt ... - reading.ac.uk

198

Page 210: The reasonable man attempts to adapt ... - reading.ac.uk

199

Page 211: The reasonable man attempts to adapt ... - reading.ac.uk

200

5.2.6 Activity of TCA cycle enzymes in strains RU116, RU118, RU137 and RU156

In order to obtain experimental evidence for the nature of the mutation in

strains RU116 and RU156, the activities of a selection of TCA cycle enzymes in

these strains were assayed (Table 5.7). The activity of these enzymes in strains

RU118 and RU137 was also measured (Table 5.7).

Table 5.7 TCA cycle enzyme activities in glucose/NH4Cl grown aspartate toxic escape Tn5 mutants of R. leguminosarum strain 3841. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of three or more independent determinations, except the succinyl-CoA synthetase values for strains RU116 and RU156 which are the average of two independent determinations. αKDH, α-ketoglutarate dehydrogenase; SCS, succinyl-CoA synthetase; MDH, malate dehydrogenase; ICDH, isocitrate dehydrogenase; CS, citrate synthase; UD, undetectable; ND, not determined.

Enzyme Strain 3841 RU116 RU118 RU137 RU156 αKDH 109±10 6±2 86±3 42±12 UD SCS 55±21 20 ND ND 252 MDH 4978±833 18830±2524 3838±505 3893±1835 17435±1622 ICDH 864±164 904±94 886±147 890±61 957±66 CS 166±26 227±31 155±24 100±23 123±37

No α-ketoglutarate dehydrogenase activity can be detected in strain RU156,

while strain RU116 lacks succinyl-CoA synthetase activity. This is consistent

with mutation of sucA and sucD, respectively in these strains. The lack of α-

ketoglutarate dehydrogenase activity in strain RU116 is discussed elsewhere

(Section 5.2.14). Interestingly, malate dehydrogenase levels in strains RU116

and RU156 are greatly elevated in comparison to strain 3841. This effect is not

observed for citrate synthase or isocitrate dehydrogenase. One potential

explanation, which is discussed in Section 5.2.15, is that the increase in malate

dehydrogenase activity is a consequence of mdh lying upstream in the same

Page 212: The reasonable man attempts to adapt ... - reading.ac.uk

201

operon as sucA and sucD. The increased activity of succinyl-CoA synthetase in

strain RU156 can be explained similarly (Section 5.2.15).

The activities of α-ketoglutarate dehydrogenase, and to a lesser degree,

citrate synthase are reduced in strain RU137 in comparison to strain 3841, while

malate dehydrogenase and isocitrate dehydrogenase are unaffected. This

suggests that PHB biosynthesis may be involved in the regulation of certain

TCA cycle enzymes. This possibility is discussed elsewhere (Section 5.3).

The activities of all the enzymes assayed in strain RU118 are comparable to

those in strain 3841, suggesting that the mutation in RU118 is unrelated to those

in strains RU116, RU137 and RU156.

5.2.7 Growth of strains RU116, RU137 and RU156 on arabinose

If strains RU116 and RU156 are mutated in succinyl-CoA synthetase and α-

ketoglutarate dehydrogenase, respectively, it is expected that they will be

unable to grow on arabinose as the carbon source, since in R. leguminosarum the

catabolic pathway for this substrate (Fig. 5.5) enters the TCA cycle at α-

ketoglutarate (Dilworth et al., 1986). Consequently, these strains were tested for

growth on acid minimal salts containing arabinose and NH4Cl. Growth of

strain RU137 on this medium was also investigated. Neither strain RU116 nor

strain RU156 was able to grow, whereas strains 3841 and RU137 grew normally.

Page 213: The reasonable man attempts to adapt ... - reading.ac.uk

202

Fig. 5.5 Pathway of catabolism of L-arabinose in R. leguminosarum (Dilworth et al., 1986).

5.2.8 Complementation of strain RU156

Concomitant to the isolation and sequencing of the transposon clones of

strains RU116, RU137 and RU156, the mutation in strain RU156 was

complemented from a strain 3841 chromosomal library.

Page 214: The reasonable man attempts to adapt ... - reading.ac.uk

203

Strain RU156 was found to be unable to grow on glutamate as sole source of

carbon and nitrogen, presumably because glutamate uptake is insufficient

and/or α-ketoglutarate dehydrogenase is one of the principle catabolic

enzymes for glutamate. It was therefore possible to look for complementation

of the mutation in strain RU156 by testing for growth on glutamate. (Growth

on arabinose can be used to test for complementation of the mutation in RU156

(Section 5.2.15), however the inability of this strain to grow on arabinose as the

carbon source was not known when these experiments were performed).

A strain 3841 chromosomal library (as EcoRI fragments in pLAFR1) was

conjugated from E. coli strain 803 into strain RU156. Five transconjugants able

to grow on AMS agar containing glutamate as sole source of carbon and

nitrogen were isolated. Purified cosmids from these transconjugants were used

to transform E. coli strain S17-1 and thence reconjugated into strain RU156. This

was done in order to distinguish between Glu+ revertants and true

transconjugants. On reconjugation, four of the five cosmids were found to

complement strain RU156 for growth on glutamate. Since all four of these

cosmids show identical restriction patterns, only one, pRU3004, was employed

in further experiments.

5.2.9 Effect of pRU3004 on aspartate transport in strains RU116, RU137, and RU156

Cosmid pRU3004 was conjugated from E. coli strain S17-1 into RU116, RU137

and RU156, creating strains RU444, RU449 and RU453, respectively. Rates of

aspartate uptake in these strains indicate that pRU3004 complements strains

RU116 and RU156, but not strain RU137, for aspartate transport (Table 5.8).

Page 215: The reasonable man attempts to adapt ... - reading.ac.uk

204

Table 5.8 Rates of aspartate transport in R. leguminosarum strains RU444, RU449 and RU453 grown on glucose/NH4Cl. Rates of uptake are expressed as nmol min-1 [mg protein]-1. Values are the result from a single experiment except those for strains 3841, RU116 and RU156 which are the mean±SEM of determinations from three or more independent cultures. Strain Uptake 3841 3.9±0.1 RU116 0.6±0.1 RU444 2.7 RU137 1.5 RU449 1.6 RU156 0.2±0.1 RU453 3.5

5.2.10 Southern blot of pRU3004 against RU116, RU137, RU156 chromosomal DNA

Southern blotting of EcoRI digested chromosomal DNA from each of strains

3841, RU116, RU137 and RU156 with pRU3004, demonstrated that pRU3004

contains DNA homologous to the Tn5-bearing EcoRI fragment in both RU116

and RU156, but not to that in RU137.

Although the inserts in pRU32 and pRU34 both appear to be approximately

10kb in size, the fact that pRU32 contains a BamHI site which is absent from

pRU34 indicates that the transposon in strains RU116 and RU156 has inserted

into a different approximately 4kb EcoRI fragment in each case. Both these

fragments must be present in pRU3004 (because pRU3004 carries an EcoRI

fragment of the strain 3841 chromosome).

In other bacteria, sucA and sucD are clustered with genes encoding other

TCA cycle enzymes (Miles & Guest, 1987; Nicholls et al., 1990; Nishiyama et al.,

1991; Guest, 1992; Guest & Russell, 1992). It was therefore anticipated that

several R. leguminosarum TCA cycle genes might be carried by pRU3004, in

addition to sucA and sucD. Consequently, pRU3004 was subjected to further

Page 216: The reasonable man attempts to adapt ... - reading.ac.uk

205

study, in an attempt to establish whether the effect on amino acid transport in

strain 3841 was specific to mutation of α-ketoglutarate dehydrogenase and

succinyl-CoA synthetase, or whether the TCA cycle in general was implicated.

5.2.11 Restriction mapping, sub-cloning and mutation of pRU3004

A combination of sub-cloning and Southern blotting was used to produce the

restriction map of the 13.4kb of insert DNA from pRU3004 shown in Fig. 5.6.

Cosmid pRU3004 was subjected to saturation Tn5-lacZ mutagenesis. The

position of the transposon in mutated cosmids was determined by restriction

analysis. Cosmids in which the mutation is located within the mapped 13.4kb

region of pRU3004 are illustrated in Fig. 5.6.

Southern blotting of pRU3004 with pRU32 and pRU34 demonstrated that the

transposon in each of strains RU116 and RU156 lies within the region of the

strain 3841 chromosome corresponding to the mapped 13.4kb of pRU3004

DNA. The position of these mutations (determined by restriction analysis of

pRU32 and pRU34) is illustrated in Fig. 5.6.

Page 217: The reasonable man attempts to adapt ... - reading.ac.uk

206

Fig. 5.6 Map of pRU3004 and mutants derived from it, with sub-clones of mutants below. The locations of Tn5 and Tn5-lacZ insertions are flagged with the number of the cosmid and/or mutant strain in which they occur. Flags representing Tn5-lacZ insertions point in the direction of transcription of the lacZ gene in the transposon. Filled flags represent active fusions. Restriction sites are: B, BamHI; E, EcoRI; P, PstI; S, SalI; Ss, SstI.

Page 218: The reasonable man attempts to adapt ... - reading.ac.uk

207

5.2.12 Genes carried by pRU3004

The Tn5-B20-bearing salI fragments from pRU3059, pRU3061, pRU3069 and

pRU3070 were each cloned into the salI site of either pJQ200 or pSP72, creating

plasmids pRU194, pRU394, pRU395 and pRU396, respectively (Fig. 5.6).

Sequencing of the insert DNA in these plasmids using a primer (P15) to the 5'

end of Tn5-B20, and either Reverse, SK and KS primers to the pJQ200

polylinker, or T7 primer to the pSP72 polylinker, provided the nucleotide

sequences shown in Figs. 5.7-5.12. Sequencing of pRU41 using Reverse and SK

primers, provided the sequence shown in Fig. 5.13.

One potential translation of each of these sequences (Figs. 5.7-5.13) exhibits

significant homology to one of the TCA cycle enzymes α-ketoglutarate

dehydrogenase, succinyl-CoA synthetase or malate dehydrogenase from other

organisms as indicated in Table 5.9.

The locations of these homologies within the various genes (Table 5.9) are

consistent with the arrangement of Rhizobium genes proposed in Fig. 5.14.

Page 219: The reasonable man attempts to adapt ... - reading.ac.uk

208

1 ACGTAACAAGATCGCACTCATTGGTTCTGGCATGATTGGTGGCACGCTGGCGCACCTCGC 60

R N K I A L I G S G M I G G T L A H L A

61 CGGCCTGAAGGAACTGGGCGACATCGTTCTCTTCGACATCGCGGACGGCATTCCCCAGGG 120

G L K E L G D I V L F D I A D G I P Q G

121 CAAGGGTCTCGATATTTCCCAGTCGTCGCCGGTCGAAGGCTTCGACGTCAATCTGACGGG 180

K G L D I S Q S S P V E G F D V N L T G

181 CGCCAGCGACTATTCCGCGATCGAAGGCGCTGACGTCTGCATCGTCACGCGCGTCGCCCG 240

A S D Y S A I E G A D V C I V T R V A R

241 CAAGCCCGGCATGAGCCGCGATGACCTTCTCGGC 274

K P G M S R D D L L G

Fig. 5.7 Nucleotide sequence from pRU396 obtained using primer P15, and putative translation.

Page 220: The reasonable man attempts to adapt ... - reading.ac.uk

209

1 CCAAGTCGATCGACGAAGTCGTCGCCCATGCCAAGGAAATGCTGGGCAACACGCTGGTGA 60

K S I D E V V A H A K E M L G N T L V T

61 CGGNGCAGACCGGCGAAGCCGGCAAGCAGGTCAACCGCCTGTACATCGAAGACGGCGCCA 120

X Q T G E A G K Q V N R L Y I E D G A N

121 ACATCGCTCGCGAGCTCTATTGCTCGCTGCTGGTCGACCGTTCGGTCGGTCGCGTGGCTT 180

I A R E L Y C S L L V D R S V G R V A X

181 TNGTGGTCTCCACCGAAGGCGGCATGGACATCGAAGCTGTCGCCCACGACACGCCTGAGA 240

V V S T E G G M D I E A V A H D T P E K

241 AGATCCAGACGATCGCCATCGATCCGGAAGCCGGCGTGACGGCTGCCGACGTTGCTGCGA 300

I Q T I A I D P E A G V T A A D V A A I

301 TCTCCAAGGCTTTTGAGCTCGAGGGTGCTGCCGCCGAAGACGCCAAGACGCTTTTT 356

S K A F E L E G A A A E D A K T L F

Fig. 5.8 Combined nucleotide sequence from pRU396 and pRU194 obtained using primers T7 and Universal respectively, and putative translation.

Page 221: The reasonable man attempts to adapt ... - reading.ac.uk

210

1 ATCAAGCTCTACGGCAAGGAGCCGGCTAACTTCTGCGACGTCGGCGGTGGCNCCGGCAAG 60

I K L Y G K E P A N F C D V G G G X G K

61 GAGAAGGTTGCTGCGGCTTTCAAGATCATCACCGCTGTCCCCANGGNCGATGGCATTCTC 120

E K V A A A F K I I T A V P X X D G I L

121 GTCATCATCTTCGGCGGCATCATGAAGTGTCTGTGCAGGGCCTGTGGGCGCNTTGCTGCG 180

V I I F G G I M K C L C R A C G R X A A

181 GTCAAGGAAGTCGGTCTCAAGGTTCCGCTCGTCGTGCGCCTTGAAGGCACCAATGTCGAG 240

V K E V G L K V P L V V R L E G T N V E

241 CTCGGCAAGAAGATCCTGAACGAGTCGGGTNTGGCGATCACGGCGGCTGACGACTTGGAC 300

L G K K I L N E S G X A I T A A D D L D

301 GATGCGGCCAAGAAGATCGTCGCGGCGATCAACGGCTGAGAATGATCATGG 351

D A A K K I V A A I N G * E * S W

Fig. 5.9 Nucleotide sequence from pRU194 obtained using primer P15, and putative translation.

Page 222: The reasonable man attempts to adapt ... - reading.ac.uk

211

1 ACGGCATCTGCGCCTTCCGCCCGGCCTCGACAACACGAGACCAGNAGATGCGCNGCGGAC 60

G I C A F R P A S T T R D Q X M R X G

61 AGGCGGCAATCGGAATGCGGCAGCCGGCCATCAAGCCGGTAAATTCATCAAAGTCAGGAG 120

Q A A I G M R Q P A I K P V N S S K S G

121 GCGGACGGAAGCGTCCGCATAACACCATGGCACGGCAAGAAGCCAACGAGCAGTTTCAGA 180

G G R K R P H N T M A R Q E A N E Q F Q

181 TCACCTCGTTTCTGGATGGCGCCAACGCTGCCTATATCGAGCAGCTCTACGCGC 234

I T S F L D G A N A A Y I E Q L Y A

Fig. 5.10 Nucleotide sequence from pRU194 obtained using primers SK and Reverse, and putative translation.

Page 223: The reasonable man attempts to adapt ... - reading.ac.uk

212

Page 224: The reasonable man attempts to adapt ... - reading.ac.uk

213

1 GCCTNGNGNGAAGAGNGNGTGAAGATGANNTGCCTTGCNAGACGATNGCNAAGNGNCCTC 60

A X X E E X V K M X C L A R R X X X X L

61 AAGGATGTGCAGAACACNGCCGCCATGCTGACCACCTACAATGAGGTGGACATGAAGGCG 120

K D V Q N T A A M L T T Y N E V D M K A

Fig. 5.12 Nucleotide sequence from pRU395 obtained using primer P15, and putative translation.

1 AATTCCGCATGAAGTTCCACAAGCCTGTTGTGCTCGACCTGTTCTGCTACCGTCGCTACG 60

F R M K F H K P V V L D L F C Y R R Y

61 GCCACAATGAAGGCGACGAACCGTCCTTCACGCAGCCGAAGATGTACAAGGTGATCCGCG 120

G H N E G D E P S F T Q P K M Y K V I R

121 CCCACAAGACCGT 133

A H K T

Fig. 5.13 Nucleotide sequence from pRU41 obtained using primers SK and Reverse, and putative translation.

Page 225: The reasonable man attempts to adapt ... - reading.ac.uk

214

Table 5.9 Homology of deduced polypeptides from pRU3004 to known sequences of TCA cycle enzymes. The number in the left hand column refers to the corresponding shaded area in Fig. 5.14. mdh encodes malate dehydrogenase; sucC and scsB encode the beta subunit of succinyl-CoA synthetase; sucD encodes the alpha subunit of succinyl-CoA synthetase; sucA encodes α-ketoglutarate dehydrogenase; sucB encodes dihydrolipoamide succinyltransferase. Number Sequence Identities 1 Fig. 5.7 40.4% over amino acids 1 to 92 of Mdh in Photobacterium

36.2% over amino acids 1 to 92 of Mdh in E.coli 2 Fig. 5.8 38.1% over amino acids 63 to 180 of SucC in E.coli

44.6% over amino acids 64 to 151 of SucC in C. burnetti 36.3% over amino acids 81 to 171 of ScsB in T. aquaticus

3 Fig. 5.9 54.3% over amino acids 275 to 389 of SucC in E. coli

55.5% over amino acids 273 to 382 of SucC in C. burnetti 38.9% over amino acids 270 to 379 of ScsB in T. aquaticus

4 Fig. 5.2 53.7% over amino acids 223 to 289 of SucD in E. coli

51.6% over amino acids 223 to 286 of SucD in C. burnetti 45.5% over amino acids 222 to 287 of SucD in T. aquaticus

5 Fig. 5.10 57.1% over amino acids 7 to 28 of SucA in C. burnetti 6 Fig. 5.11 62.8% over amino acids 438 to 480 of SucA in C. burnetii

60.5% over amino acids 445 to 487 of SucA in A. vinelandii 58.1% over amino acids 439 to 481 of SucA in E. coli 55.8% over amino acids 443 to 485 of SucA in B. subtilis

7 Fig. 5.12 68.1% over amino acids 679 to 797 of SucA in E. coli

63.7% over amino acids 688 to 810 of SucA in A. vinelandii 63.5% over amino acids 681 to 795 of SucA in C. burnetti 60.3% over amino acids 681 to 803 of SucA in B. subtilis

8 Fig. 5.13 50.0% over amino acids 187 to 222 of SucB in B. subtilis

45.0% over amino acids 172 to 211 of SucB in E. coli 43.2% over amino acids 168 to 204 of SucB in A. vinelandii

Page 226: The reasonable man attempts to adapt ... - reading.ac.uk

215

Fig. 5.14 Putative arrangement of genes in mapped region of pRU3004. Gene locations and orientations are based on homologies of sequenced regions (shaded) to genes in other bacteria as detailed in Table 5.9. Numbering of shaded areas refers to Table 5.9.

Page 227: The reasonable man attempts to adapt ... - reading.ac.uk

216

5.2.13 β-galactosidase activities from pRU3004 mutants

As experimental confirmation of the proposed directions of transcription for

mdh and sucCDAB, β-galactosidase activities produced in a strain 3841

background by corresponding pRU3004 mutants were investigated. Those

cosmids in which lacZ in the transposon is aligned with the proposed direction

of transcription for the corresponding gene show significantly higher activity

than those in which lacZ is in the opposite orientation (Fig. 5.6).

5.2.14 TCA cycle enzyme activities in sucCDAB mutants of strain 3841

In order to investigate directly the nature of the partially sequenced genes

carried by pRU3004, activities of the corresponding TCA cycle enzymes were

assayed in chromosomal mutants of these genes.

The Tn5-lacZ mutants of strain 3841, RU725, RU733, RU724 and RU726 were

generated by homogenotization (Ruvkun & Ausubel, 1981) of cosmids

pRU3059, pRU3061, pRU3067 and pRU3069, respectively. No homogenotes of

cosmids pRU3070 or pRU3076 could be isolated, presumably because mutation

of mdh is lethal.

α-Ketoglutarate dehydrogenase and succinyl-CoA synthetase levels in the

mutant strains were measured (Table 5.10).

Table 5.10 TCA cycle enzyme activities in R. leguminosarum strains RU724, RU725, RU726 and RU733 grown on glucose/NH4Cl. Activities are expressed as nmol min-1 (mg protein)-1. Values are the mean of two independent determinations, except the 3841 values which are the mean±SEM of four independent determinations. αKDH, α-ketoglutarate dehydrogenase; SCS, succinyl-CoA synthetase; UD, undetectable; ND, not determined.

Enzyme Strain 3841 RU724 RU725 RU726 RU733 αKDH 109±10 UD UD UD UD SCS 55±21 ND 22 98 ND

Page 228: The reasonable man attempts to adapt ... - reading.ac.uk

217

The loss of α-ketoglutarate dehydrogenase activity in the sucA (RU724,

RU733) and sucB (RU726) mutants is expected, as is the reduction in succinyl-

CoA synthetase activity in the sucC mutant (RU725). The observation that

strain RU725 exhibits negligible α-ketoglutarate dehydrogenase activity, as

does strain RU116 (Section 5.2.6), can be understood if sucAB are under the

control of the same promoter as sucCD, or if high levels of succinyl-CoA,

accumulated as a result of the loss of succinyl-CoA synthetase, cause feed-back

inhibition of α-ketoglutarate dehydrogenase (Williamson & Cooper, 1980). In

order to distinguish between these two possibilities, the location of promoters

in the mapped 13.4kb region of pRU3004 was investigated (Section 5.2.15).

The elevated succinyl-CoA synthetase level in strain RU726 is discussed

elsewhere (Section 5.2.15).

5.2.15 Mapping of promoter sites in pRU3004

Promoter activity in the region of the mdh and sucCDAB genes of R.

leguminosarum strain 3841 was investigated by testing the ability of various sub-

clones of pRU3004 to complement chromosomal mutations in these genes.

None of the strains RU725, RU733, RU724 and RU726 is able to grow on

arabinose as the carbon source, a phenotype consistent with mutation of α-

ketoglutarate dehydrogenase or succinyl-CoA synthetase (Section 5.2.7). It was

therefore possible to use growth on arabinose as a test for complementation of

the mutation in these strains.

A selection of EcoRI, PstI and SstI fragments of pRU3004 were cloned in the

corresponding site in pRK415-1, creating plasmids pRU180, pRU181, pRU182,

pRU276, pRU277, pRU278, pRU397 and pRU398 (Fig. 5.15). Each of these

plasmids was conjugated into RU116, RU156, RU724, RU725, RU276 and RU733

and the resulting strains tested for growth on arabinose (Table 5.11).

Page 229: The reasonable man attempts to adapt ... - reading.ac.uk

218

Fig 5.15 Sub-clones of pRU3004 used in complementation analysis. Two sub-clones of pRU3068 are also shown. The shaded arrows indicate the direction of transcription initiation from the lac promoter in the vector of each sub-clone. Restriction sites are: E, EcoRI; P, PstI; Ss, SstI.

Page 230: The reasonable man attempts to adapt ... - reading.ac.uk

219

Table 5.11 Growth of sucCDAB mutants of R. leguminosarum strain 3841, containing sub-clones of pRU3004, on arabinose/NH4Cl. Growth medium was AMS agar with the C/N source added at 10mM. +, good growth; -, no growth; ND, not determined. Plasmid Strain RU116 RU156 RU724 RU725 RU726 RU733 None - - - - - - pRU277 - ND ND - ND ND pRU278 - ND ND - ND ND pRU181 - ND ND - ND ND pRU180 ND - - ND - - pRU276 ND - - ND + - pRU398 ND - - ND - - pRU397 ND - - ND + - pRU182 ND ND ND ND - ND

Transcription of the strain 3841 genes carried by pRU181, pRU276, pRU277

and pRU397 can be initiated by the lac promoter in the vector. Therefore the

observation that mutations in strains RU725 and RU116 are not complemented

by pRU277, which carries sucCD, indicates that sucCDA are under the control of

a single promoter upstream of sucC (because a chromosomal insertion in either

sucC or sucD appears to prevent transcription of at least one gene downstream

of sucD). This result is consistent with the observation that strains RU116 and

RU725 exhibit negligible α-ketoglutarate dehydrogenase activity (Sections 5.2.6

and 5.2.14).

The absence of a promoter between the transposon in pRU3068 and the PstI

site upstream of it was confirmed by cloning, in both orientations in pIJ1891, the

lacZ-bearing PstI fragment from pRU3068, and measuring the β-galactosidase

activity produced by the resulting clones in a strain 3841 background. In

pRU386 (Fig. 5.15) transcription of the lacZ fusion can be initiated by the lac

promoter in the vector, and β-galactosidase activity in strain 3841 containing

this plasmid is high (1336±211 nmol min-1 [mg protein]-1). In pRU387 (Fig. 5.15)

Page 231: The reasonable man attempts to adapt ... - reading.ac.uk

220

the direction of transcription of the insert DNA is opposite to the lac promoter

in the vector, so the low level of β-galactosidase activity (95±4 nmol min-1 [mg

protein]-1) observed in strain 3841 containing this plasmid is indicative of a lack

of native promoter activity in the insert. (Cosmid pRU3068, which contains the

same replicon as pRU386 and pRU387, produces β-galactosidase activity of 3119

±195 nmol min-1 [mg protein]-1 in strain 3841).

Since the mutation in strain RU726 is not complemented by pRU398, either

there is no promoter in the intergenic region between sucA and sucB, or a sucB

mutation is polar on a downstream gene required for growth on arabinose.

However, since the mutation in strain RU726 is complemented by pRU276, it

can be concluded that pRU398 does not complement strain RU726 because sucB

lacks a promoter of its own. The fact that pRU397 complements the mutation in

RU726 demonstrates that the gene mutated in this strain lies within the 2.4kb

EcoRI-PstI fragment containing the transposon.

It is not possible to deduce from these complementation data whether the

promoter controlling sucCDAB lies in the intergenic region between mdh and

sucC, or upstream of mdh. However, the latter arrangement might provide an

explanation for the increased malate dehydrogenase activity observed in sucDA

mutants (RU116 and RU156; Table 5.7), and the elevated levels of succinyl-CoA

synthetase in sucAB mutants (RU156 and RU726; Table 5.7 and 5.10). A

mutation in sucCDAB may lead to an increase in the expression of the mutated

sucCDAB in response to a change in the intracellular concentration of one or

more metabolites, such as α-ketoglutarate, in the mutant. If this is the case, then

the location of mdh upstream of sucCDAB, and under the control of the same

promoter, would result in the increased expression of mdh in a sucCDAB

mutant. Similarly, mutation of sucAB would lead to increased sucCD

expression.

Page 232: The reasonable man attempts to adapt ... - reading.ac.uk

221

5.2.16 Amino acid excretion by strains RU116, RU156 and RU543

If, as has been suggested (Section 1.2.5), it is inhibition and/or repression of

α-ketoglutarate dehydrogenase that results in the excretion of amino acids by

bacteroids, it might be expected that mutants of R. leguminosarum lacking α-

ketoglutarate dehydrogenase activity will also excrete amino acids.

Consequently, strains RU116 and RU156 were tested for amino acid excretion.

Supernatants from cultures growing on L-malate/NH4Cl were assayed over

time for alanine, aspartate, glutamate and α-ketoglutarate (Table 5.12). L-

malate was provided as the carbon source because this is the most abundant C4-

dicarboxylate in the nodule (Streeter, 1987), and consequently the most likely

carbon source for the bacteroid (Section 1.2.1).

Although no excretion of alanine or aspartate is apparent, significant

concentrations of glutamate and α-ketoglutarate were detected in the

supernatants from cultures of both strain RU116 and strain RU156. This is in

contrast to the wild-type, for which no excretion of any of the four metabolites

assayed was detected.

The excretion of certain metabolites, including α-ketoglutarate, by bacteria

under particular metabolic conditions is well documented (Tempest & Neijssel,

1992; Kramer, 1994), and the excretion of α-ketoglutarate by R. leguminosarum

strains containing mutations that block the major catabolic pathway for this

substrate is not surprising. However, the excretion of glutamate by strains

RU116 and RU156 suggests that a proportion of the excess α-ketoglutarate is

converted to glutamate.

Page 233: The reasonable man attempts to adapt ... - reading.ac.uk

222

Table 5.12 Excretion of metabolites by R. leguminosarum strains RU116 and RU156 grown on L-malate/NH4Cl. Cells grown on L-malate/NH4Cl were washed and resuspended in AMS containing 10mM L-malate/NH4Cl and incubated at 28°C. After the time intervals shown, the concentrations of each of the four metabolites in samples of supernatants was determined. At the 0 min and 150 min time points, samples of culture were taken and assayed for protein content. This enabled an approximate calculation of the rate of excretion, in nmol min-1 [mg cell dry weight]-1, assuming 0.2 [mg protein] ml-1 ≡ 0.4 mg dry weight of cells (P.S.Poole, unpublished data). Values are the result from a single representative experiment, with the concentration at different times given in µM. Metabolite Time (min) Excretion and strain 0 60 150 270 rate L-Alanine 3841 <10 <10 <10 <10 <1 RU116 <10 <10 <10 <10 <1 RU156 <10 <10 <10 <10 <1 L-Aspartate 3841 <10 <10 <10 <10 <1 RU116 <10 <10 <10 <10 <1 RU156 <10 <10 <10 <10 <1 L-Glutamate 3841 <10 <10 <10 <10 <1 RU116 <10 81 158 477 15 RU156 <10 52 98 138 11 α-Ketoglutarate 3841 <10 <10 <10 <10 <1 RU116 <10 218 805 >1000 79 RU156 <10 404 >1000 >1000 >105

Since aspartate is a potential amino-donor in the production of glutamate

from α-ketoglutarate, it was thought possible that the escape of strains RU116

and RU156 from aspartate toxicity might be due to transamination of aspartate

to glutamate, and subsequent excretion of the glutamate. To investigate this

possibility, excretion of glutamate by strains 3841, RU116 and RU156 grown on

both glucose/NH4Cl and glucose/NH4Cl/aspartate was measured. For the

purpose of comparison to the results from cultures containing malate as the

Page 234: The reasonable man attempts to adapt ... - reading.ac.uk

223

carbon source, supernatants from the glucose/NH4Cl cultures were also

assayed for α-ketoglutarate (Table 5.13).

Table 5.13 Excretion of glutamate and α-ketoglutarate by R. leguminosarum strains 3841, RU116, RU156 and RU543 grown on glucose/NH4Cl and glucose/NH4Cl/aspartate. Cells grown on glucose/NH4Cl or glucose/NH4Cl/(20mM)aspartate were washed and resuspended in AMS containing 10mM glucose/NH4Cl or glucose/NH4Cl/(20mM)aspartate respectively, and incubated at 28°C. After the time intervals shown, the concentrations of the α-ketoglutarate and/or glutamate in samples of supernatants was determined. At the 0 min and 150 min time points, samples of culture were taken and assayed for protein content. This enabled an approximate calculation of the rate of excretion, in nmol min-1 [mg cell dry weight]-

1, assuming 0.2 [mg protein] ml-1 ≡ 0.4 mg dry weight of cells (P.S.Poole, unpublished data). Values are the result from a single representative experiment, with the concentration at different times given in µM. Glc, glucose; NH4, NH4Cl; Asp, aspartate. Metabolite Time (min) Excretion and culture 0 60 150 270 rate L-Glutamate 3841/Glc/NH4 <10 <10 <10 <10 <1 RU116/Glc/NH4 <10 46 179 333 18 RU156/Glc/NH4 <10 <10 65 120 8 RU543/Glc/NH4 <10 <10 <10 <10 <1 3841/Glc/NH4/Asp <10 34 106 225 10 RU116/Glc/NH4/Asp <10 195 602 >1000 59 RU156/Glc/NH4/Asp <10 170 443 863 49 RU543/Glc/NH4/Asp <10 <10 18 55 2 α-Ketoglutarate 3841/Glc/NH4 <10 <10 <10 <10 <1 RU116/Glc/NH4 <10 38 49 55 5 RU156/Glc/NH4 <10 95 195 311 24

Excretion of glutamate by strains RU116 and RU156 is substantially increased

by the presence of aspartate in the medium, consistent with increased

transamination of α-ketoglutarate. This result supports the notion that these

strains are able to survive high levels of aspartate because the increased

availability of intracellular α-ketoglutarate enables the toxic aspartate to be

Page 235: The reasonable man attempts to adapt ... - reading.ac.uk

224

removed by transamination. The increased rate of glutamate excretion by

strains RU116 and RU156 grown on glucose/NH4Cl/aspartate appears to be

greater than can be accounted for by the rate of α-ketoglutarate excretion by

these strains grown on glucose/NH4Cl, given the small apparent size of the

intracellular pool of α-ketoglutarate (Section 5.2.17). However, metabolism, via

the TCA cycle, of oxaloacetate generated from the transamination of α-

ketoglutarate by aspartate, will presumably lead to the generation of additional

α-ketoglutarate in cells grown on glucose/NH4Cl/aspartate. The rate of

excretion of α-ketoglutarate by strains RU116 and RU156 grown on malate is

significantly greater than that found when glucose is the carbon source. This is

consistent with greater carbon flux through the TCA cycle during growth on the

TCA cycle intermediate, malate.

Interestingly, the presence of aspartate in the medium results in excretion of

glutamate, but not α-ketoglutarate, by strain 3841. It seems likely that the

excreted glutamate is derived from α-ketoglutarate, and its synthesis is

presumably the result of an increased supply of amino-donor (aspartate) for the

transamination of α-ketoglutarate, and/or increased production (via the TCA

cycle) of α-ketoglutarate from oxaloacetate generated from aspartate by

transamination. It is also possible that the presence of aspartate in the medium

allows intracellular glutamate to be exchanged out of the cell (Section 3.2.18).

The observation that α-ketoglutarate is not excreted by 3841 grown on

glucose/NH4Cl/aspartate suggests that, in this case, the rate of glutamate

synthesis equals the rate of α-ketoglutarate production. In strains RU116 and

RU156 grown on glucose/NH4Cl and malate/NH4Cl, the rate of α-

ketoglutarate generation presumably exceeds the maximum rate of

transamination to glutamate, and consequently α-ketoglutarate is excreted. The

fact that, while the rates of α-ketoglutarate excretion by these strains differ

substantially between growth on malate and growth on glucose, the rates of

Page 236: The reasonable man attempts to adapt ... - reading.ac.uk

225

glutamate excretion are very similar for both carbon sources, is consistent with

this suggestion.

Since strain 3841 excretes glutamate when grown on

glucose/NH4Cl/aspartate, it was possible to investigate the role of the general

amino acid permease in the excretion of glutamate, by assaying glutamate

excretion by strain RU543 (an aapJ mutant) grown on glucose/NH4Cl/aspartate

(Table 5.13).

Glutamate excretion by strain RU543 is significantly reduced in comparison

to strain 3841. It is unlikely that this result is a consequence of an effect on

aspartate uptake in strain RU543, since aspartate uptake in R. leguminosarum is

mediated by the dct system at the concentration of aspartate employed in this

experiment (Section 5.2.1). Strain RU543 grows more slowly on

glucose/NH4Cl/aspartate than strain 3841 (mean generation times are 1095 min

and 660 min, respectively), and a proportion of the reduction in glutamate

excretion observed in RU543 may be attributable to the reduced metabolic rate

in this strain. However, a potential explanation of the poor growth of RU543 on

glucose/NH4Cl/aspartate, is that this strain can not easily excrete the glutamate

it synthesizes under these conditions. These data therefore appear to indicate

the involvement of the general amino acid permease in the excretion of

glutamate from strain 3841 grown on glucose/NH4Cl/aspartate. In view of the

results of amino acid exchange experiments on strain 3841 (Section 3.2.18), and

the presence of aspartate in the medium, glutamate may be exported via an

exchange mechanism.

The protocol used in this research to prepare cells for transport assays

involves the final resuspension of cells in minimal salts medium lacking carbon

and nitrogen sources, and incubation for between one and six hours prior to the

assay(s). It therefore seemed possible (though unlikely, given the lack of carbon

and nitrogen sources) that the low amino acid transport values obtained for

strains RU116 and RU156 might be the result of competitive inhibition of the

Page 237: The reasonable man attempts to adapt ... - reading.ac.uk

226

general permease by glutamate excreted from these strains into the medium

during the incubation period before the assays. However, no glutamate

excretion could be detected from cells of strains RU116 and RU156 resuspended

and incubated for five hours in minimal salts containing no carbon or nitrogen

source (data not shown). In addition, glutamate transport values for cells of

either strain RU116 or strain RU156, spun down and resuspended in fresh salts

immediately prior to assaying, were found to be identical to those of cells

treated in the usual way before being assayed (data not shown). It is evident

therefore, that the impaired amino acid transport in strains RU116 and RU156 is

not due to the excretion of glutamate by these strains.

5.2.17 Intracellular concentrations of α-ketoglutarate and glutamate in strains RU116 and RU156

The fact that α-ketoglutarate and glutamate are excreted by strains RU116

and RU156 suggests that the intracellular concentrations of these substrates

may be elevated in these mutants. It seems improbable that an elevated

intracellular concentration of any metabolite will be maintained during an

incubation without carbon or nitrogen of up to six hours, so it is unlikely that

direct inhibition of the general amino acid permease by, for example,

intracellular glutamate, is the reason for the reduced amino acid uptake in

sucDA mutants of strain 3841. However, an elevated level of intracellular

glutamate or α-ketoglutarate may be the signal that initiates down-regulation of

the amino acid permease.

It was therefore decided to ascertain whether intracellular glutamate or α-

ketoglutarate concentrations are in fact elevated in strains RU116 and RU156.

The assays were carried out on cells grown on glucose/NH4Cl (Table 5.14).

Page 238: The reasonable man attempts to adapt ... - reading.ac.uk

227

Table 5.14 Intracellular concentrations of glutamate and α-ketoglutarate in R. leguminosarum strains 3841, RU116 and RU156 grown on glucose/NH4Cl. Concentrations are calculated assuming: 1g dry weight of cells ≡ 1.45ml (Dilworth & Glenn, 1982); 0.2 [mg protein] ml-1 ≡ 0.4 mg dry weight of cells (P.S.Poole, unpublished data); the periplasmic space constitutes 20% total cell volume (Dilworth & Glenn, 1982). Values are the result from a single experiment and are given in mM. Metabolite Strain 3841 RU116 RU156 L-Glutamate 4 12 52 α-Ketoglutarate <1 <1 <1

Since glutamine is likely to have been hydrolysed to glutamate during the

preparation of the samples in which intracellular substrate concentrations were

assayed (Section 2.1.10), the figures given in Table 5.14 strictly represent the

combined intracellular concentrations of glutamate and glutamine. However,

since glutamate is known to be excreted by strains RU116 and RU156 under the

growth conditions employed in this experiment, it is probable that a significant

proportion of the elevated figure for intracellular glutamate in each of these

strains (Table 5.14) is attributable to glutamate rather than glutamine. If this is

the case, the involvement of the intracellular glutamate level in the regulation of

the general amino acid permease seems likely.

Very little intracellular α-ketoglutarate was detected for all three strains

RU116, RU156 and 3841. This seems unlikely to be an experimental artefact

since α-ketoglutarate was shown to be stable under the conditions used to

prepare samples (Section 2.1.10), and the results for intracellular glutamate are

as anticipated. It therefore appears that the intracellular pool of α-ketoglutarate

in the wild-type is low under these conditions, and this is reflected in the ready

excretion of this metabolite by the mutants.

Page 239: The reasonable man attempts to adapt ... - reading.ac.uk

228

5.2.18 Transcription of the aap operon in sucDA mutants of strain 3841

In order to determine whether transcriptional regulation mediates the

reduction in amino acid transport observed in succinyl-CoA synthetase and α-

ketoglutarate dehydrogenase mutants of strain 3841, the activity of an aapJ::lacZ

transcriptional fusion in strain RU116 and RU156 backgrounds was

investigated.

Cosmid pRU3028, a Tn5-lacZ mutant of pRU3024 in which the transposon

has inserted in aapJ such that lacZ is in the same orientation as the mutated gene

(Fig. 3.2), was conjugated into strains RU116 and RU156. β-Galactosidase

activity in the resulting strains, RU889 and RU897, respectively, was measured

under both nitrogen-excess and nitrogen-limited conditions (Table 5.15). The β-

galactosidase activity produced by pRU3024 in strain RU116 and RU156

backgrounds (strains RU1024 and RU1025, respectively) was also measured as a

control.

β-Galactosidase activities in strains RU889 and RU897 are generally very

similar to those in strain RU443 (3841/pRU3028), although activity in strain

RU897 grown on glucose/glutamate is somewhat reduced, perhaps because

this strain grows poorly under these conditions. Certainly, the activity of the

fusion in strains RU889 and RU897 does not indicate a decrease in the

transcription of the aap operon sufficient to account for the severe reduction in

amino acid transport observed in strains RU116 and RU156 grown on

glucose/NH4Cl (Section 5.2.3).

Page 240: The reasonable man attempts to adapt ... - reading.ac.uk

229

Table 5.15 β-galactosidase activities in R. leguminosarum strains RU116 and RU156 containing pRU3028 and pAR36A. Strains were grown on AMS containing either glucose/NH4Cl or glucose/glutamate at 10mM. Activities are expressed as nmol min-1 [mg protein]-1. Values are the mean±SEM of determinations from three or more independent cultures. Strain Relevant genotype Growth conditions Host Cosmid/plasmid Glucose/NH4Cl Glucose/glutamate RU438 (3841/pRU3024) Wild-type aapJQMP 40±5 43±1 RU1024 (RU116/pRU3024) sucD::Tn5 aapJQMP 32±2 40±3 RU1025 (RU156/pRU3024) sucA::Tn5 aapJQMP 41±4 34±4 RU443 (3841/pRU3028) Wild-type aapJ::Tn5-lacZ 3325±279 7923±666 RU889 (RU116/pRU3028) sucD::Tn5 aapJ::Tn5-lacZ 3225±209 6484±118 RU897 (RU156/pRU3028) sucA::Tn5 aapJ::Tn5-lacZ 3853±234 4726±259 RU368 (3841/pMP220) Wild-type Promoterless lacZ 62±3 105±1 RU1002 (RU116/pMP220) sucD::Tn5 Promoterless lacZ 111±21 50±7 RU1003 (RU156/pMP220) sucA::Tn5 Promoterless lacZ 37±11 40±1 RU622 (3841/pAR36A) Wild-type glnII:lacZ 101±14 531±79 RU891 (RU116/pAR36A) sucD::Tn5 glnII:lacZ 270±6 879±66 RU899 (RU156/pAR36A) sucA::Tn5 glnII:lacZ 53±4 513±42

Page 241: The reasonable man attempts to adapt ... - reading.ac.uk

230

It is interesting that, like the wild-type strain, both mutants exhibit repression

of the aap operon when grown on NH4Cl as the nitrogen source. This result is

consistent with the finding that in strains RU116 and RU156, as in strain 3841,

the intracellular α-ketoglutarate concentration is low (Section 5.2.17). Strains

lacking α-ketoglutarate dehydrogenase activity might have been expected to

have an elevated intracellular concentration of α-ketoglutarate and hence an

elevated α-ketoglutarate to glutamine ratio, resulting in a phenotype

compatible with apparent nitrogen-limitation (Section 1.6), even under

conditions of nitrogen-excess. This is apparently not the case.

As confirmation of this observation β-galactosidase activities in strains

RU116 and RU156 carrying the R. leguminosarum glnII::lacZ reporter fusion

pAR36A (Patriarca et al., 1992) were measured (Table 5.15). pAR36A activity in

the RU116 and RU156 backgrounds (strains RU891 and RU899, respectively)

responds to a change from nitrogen-limited to nitrogen-excess conditions as it

does in the 3841 background (strain RU622).

While it is apparent that the aap operon is not affected at the transcriptional

level by a mutation in sucDA, inhibition of the transporter by intracellular

glutamate (or indeed any substrate) also appears to be unlikely (Section 5.2.16).

It seems reasonable to propose that mutation of α-ketoglutarate dehydrogenase

or succinyl-CoA synthetase may lead to post-translational modification of the

amino acid permease, and it is plausible that this modification is initiated by an

increase in intracellular glutamate. Post-translational regulation, in the form of

phosphorylation of the binding protein, has previously been reported for the

lysine-arginine-ornithine uptake system of E. coli (Celis, 1984; Celis, 1990).

5.2.19 Plant properties of strains RU116, RU137 and RU156

Although both strain RU116 and strain RU156 nodulate peas, the nodules are

extremely small and green. No kanamycin resistant bacteria could be recovered

from nodules produced by these strains, and acetylene reduction was not

tested. These results are consistent with an earlier report that an α-

Page 242: The reasonable man attempts to adapt ... - reading.ac.uk

231

ketoglutarate dehydrogenase mutant of R. meliloti formed ineffective nodules

(Duncan & Fraenkel, 1979).

The pea nodules produced by strain RU137 appeared on average to be larger

than those produced by strain 3841, however acetylene reduction by RU137

nodules was not investigated. All 25 nodule isolates of this strain were found to

have retained kanamycin resistance. Bean nodules induced by a PHB synthase

mutant of Rhizobium etli have been found to exhibit increased nitrogenase

activity (M.A.Cevallos, personal communication). In contrast, PHB

biosynthesis mutants of R. meliloti have been reported to have similar symbiotic

traits to their parental strain (Povolo et al., 1994).

Page 243: The reasonable man attempts to adapt ... - reading.ac.uk

232

5.3 DISCUSSION

Mutation of the TCA cycle enzymes α-ketoglutarate dehydrogenase and

succinyl-CoA synthetase in R. leguminosarum strain 3841 leads to a severe

reduction in amino acid uptake by the mutants. Uptake of other substrates

such as glucose or succinate is also impaired in such mutants, probably as a

result of general dehabilitation produced by the disruption of central

metabolism, however the particularly marked reduction in amino acid transport

suggests that there is a specific effect on this system. This is in accordance with

the suggestion that it is the co-ordinated operation of the TCA cycle and amino

acid transport that regulates the flow of carbon and energy through the

bacteroid (Section 3.1).

The lacZ fusion data indicate that there is no change in the transcription

levels of the aap operon in sucDA mutants, and the conditions encountered by

cells prior to transport assays effectively eliminate the possibility of intracellular

substrate inhibition of the general amino acid permease. This suggests that the

general amino acid permease may be subject to post-translational modification.

The intracellular glutamate concentration was found to be elevated in sucDA

mutants grown on glucose/NH4Cl, and it is possible that this elevation is the

signal for the down-regulation of the general amino acid permease. In addition,

glutamate is excreted by sucDA mutants grown on both glucose/NH4Cl and

malate/NH4Cl. These results are consistent with several studies which indicate

that a significant proportion of malate supplied to bacteroids under nitrogen-

fixing conditions, is converted to glutamate (Salminen & Streeter, 1987a; Kouchi

et al., 1991; Miller et al., 1991; Salminen & Streeter, 1992). Such studies have led

to the suggestion that glutamate synthesis in the bacteroid is a consequence of

inhibition of α-ketoglutarate dehydrogenase by NADH (McDermott et al., 1989;

Salminen & Streeter, 1990). Although complete loss of α-ketoglutarate

dehydrogenase activity is artificial, the accumulation of glutamate by free-living

sucDA mutants appears to support this hypothesis.

Page 244: The reasonable man attempts to adapt ... - reading.ac.uk

233

Although no excretion of glutamate by bacteroids has been reported, such

excretion does add credence to the suggestion that inhibition and/or repression

of central metabolic enzymes in the bacteroid may lead to excretion of amino

acids (McDermott et al., 1989; Rosendahl et al., 1992). Indeed, inhibition of

pyruvate dehydrogenase, an enzyme with strong structural and functional

similarities to α-ketoglutarate dehydrogenase (Miles & Guest, 1987; Guest &

Russell, 1992), would lead to the accumulation of pyruvate. This might explain

the observed excretion of alanine by bacteroids: Transamination of pyruvate

followed by excretion of the resulting alanine, could relieve pyruvate

accumulation. This hypothesis is supported by the observation that alanine

dehydrogenase and alanine aminotransferase activity has been detected in

bacteroids but not free-living cells of R. leguminosarum strain 3841 (P.S.Poole,

unpublished data).

A further indication of the potential importance of transamination in

determining the fate of metabolites and hence regulating growth of R.

leguminosarum, is provided by the finding that addition of aspartate to the

growth medium leads to an increase in glutamate excretion by sucDA mutants.

This is consistent with an increase in transamination of α-ketoglutarate due to

the presence of additional amino-donor, although increased synthesis of α-

ketoglutarate from oxaloacetate generated by transamination of aspartate, is

also likely to be a contributory factor.

Although α-ketoglutarate is excreted by sucDA mutants, there is apparently

no significant increase in the intracellular concentration of this metabolite in

these mutants. This is consistent with a report that radioactivity failed to

accumulate in α-ketoglutarate in bacteroids supplied with labelled malate, an

effect that was ascribed to rapid conversion of α-ketoglutarate to glutamate

(Salminen & Streeter, 1990).

The significantly reduced rate of glutamate excretion by the aapJ mutant

RU543 grown on glucose/NH4Cl/aspartate, in comparison to that of strain

Page 245: The reasonable man attempts to adapt ... - reading.ac.uk

234

3841, suggests that the general amino acid permease is involved in glutamate

excretion. This is consistent with results from amino acid exchange

experiments which indicate that efflux of intracellular AIB in the presence of an

excess of extracellular amino acid is dependent on the general permease

(Section 3.2.18). It is possible that glutamate excretion occurs via exchange with

aspartate.

In view of the apparent involvement of the general amino acid permease in

the excretion of glutamate, it is interesting that strains RU116 and RU156 are

able to excrete glutamate despite exhibiting severely reduced glutamate uptake

(Table 5.5). The low rate of glutamate uptake in these strains does not appear to

be due to repression of the aap operon, or substrate inhibition, and one

possibility is that the amino acid permease is covalently modified so as to

reduce uptake but maintain efflux; although this would argue against an

exchange mechanism. Modification of the periplasmic binding protein could

potentially alter uptake without affecting efflux, and the regulation of amino

acid uptake via post-translational modification of a binding protein has been

reported previously (Celis, 1984; Celis, 1990). Furthermore, it has been

suggested that a function of high affinity ABC transporters of amino acids may

be to recapture biosynthetically produced amino acids as they escape from the

cell (Ames, 1972; Antonucci & Oxender, 1986). If the bacteroid is required to

excrete amino acids, this may provide a rationale for reducing uptake, but

maintaining efflux, via the general amino acid permease of R. leguminosarum, in

circumstances which result in the accumulation of intracellular amino acid.

In strain 3841, the gene encoding the TCA cycle enzyme, malate

dehydrogenase, is found to be located immediately upstream of the genes

encoding succinyl-CoA synthetase, which in turn lie immediately upstream of

the α-ketoglutarate dehydrogenase genes. This arrangement of genes is

different from that found in E. coli, where mdh is located in an entirely different

part of the chromosome to sucABCD (Guest, 1992; Guest & Russell, 1992), but

Page 246: The reasonable man attempts to adapt ... - reading.ac.uk

235

appears to be similar to that in T. aquaticus where sucCD are clustered with mdh

(Nicholls et al., 1990; Nishiyama et al., 1991), although the locations of sucAB are

not known in this case. The occurrence of the genes for the three enzymes in a

single operon in R. leguminosarum could account for the elevated malate

dehydrogenase levels observed in succinyl-CoA synthetase and α-ketoglutarate

dehydrogenase mutants, and the increased succinyl-CoA synthetase activity

found in α-ketoglutarate dehydrogenase mutants, if a mutation in sucCDAB

leads to an increase in the expression of the mdh-sucCDAB operon. One

potential reason for such an increase in expression, is that the mdh-sucCDAB

operon may be regulated in response to metabolite level(s) that are altered as a

result of the mutation.

A similar pattern of enzyme activities has been observed previously (but not

explained) for an α-ketoglutarate dehydrogenase mutant of R. meliloti (Duncan

& Fraenkel, 1979). Interestingly, a succinate dehydrogenase mutant of R.

leguminosarum is reported to exhibit a 5-fold increase in malate dehydrogenase

activity (Finan et al., 1981). This suggests that a mutation in sdhABCD has a

similar effect on the regulation of mdh expression to a sucCDAB mutation.

Attempts to isolate malate dehydrogenase mutants of strain 3841 were

unsuccessful, probably because loss of this enzyme from central metabolism is

lethal. However, if, as suspected, the promoter that controls sucCDAB

expression lies upstream of mdh, then an mdh mutant would not provide insight

into the question of whether it is the overall disruption of the TCA cycle, or the

specific loss of α-ketoglutarate dehydrogenase, that leads to severely reduced

amino acid uptake.

It seems likely that it is the lowered α-ketoglutarate dehydrogenase activity

in strain RU137, leading to some accumulation of glutamate, that is the cause of

the impaired amino acid transport in this strain. Indeed, across the strains

RU116, RU137 and RU156, there is a good correlation between the degree of

reduction in amino acid transport and that in α-ketoglutarate dehydrogenase

Page 247: The reasonable man attempts to adapt ... - reading.ac.uk

236

activity. However, it is not clear why α-ketoglutarate dehydrogenase activity is

reduced in strain RU137. If mdh and sucAB are indeed in the same operon, then

any regulation of α-ketoglutarate dehydrogenase must be post-transcriptional,

since malate dehydrogenase activity in strain RU137 is unaffected (Table 5.7).

It has been proposed that biosynthesis of PHB in the oxygen-limited

environment of Bradyrhizobium japonicum bacteroids serves to lower

NAD(P)H/NAD(P) ratios, thereby reducing the inhibition of certain TCA cycle

enzymes including α-ketoglutarate dehydrogenase (McDermott et al., 1989).

Under such circumstances a mutation that prevents PHB biosynthesis may lead

to increased inhibition of α-ketoglutarate dehydrogenase as a result of

increased NAD(P)H levels. If PHB production occurs in free-living cells of R.

leguminosarum under aerobic conditions, then this may provide an explanation

of the phenotype of strain RU137 with regard to α-ketoglutarate activity.

Another possibility that also implies a role for PHB biosynthesis in free-living

cells, is that mutation of phbC affects acetyl-CoA concentrations in the cell.

Acetyl-CoA plays a significant role in determining the level of TCA cycle

intermediates due to its allosteric activation of the enzyme pyruvate

carboxylase, which catalyzes the conversion of pyruvate to oxaloacetate. It is

therefore conceivable that alteration of the acetyl-CoA level may lead to

reduced α-ketoglutarate dehydrogenase activity.

Reduction in α-ketoglutarate dehydrogenase activity is the probable reason

for the escape from aspartate toxicity of strains RU116, RU137 and RU156. A

plausible escape mechanism is the removal of toxic aspartate through the

transamination of α-ketoglutarate to glutamate by aspartate aminotransferase.

Glutamate is less toxic to strain 3841 than aspartate (Section 5.2.1), and is also

excreted.

Page 248: The reasonable man attempts to adapt ... - reading.ac.uk

237

CHAPTER 6 FINAL DISCUSSION

Page 249: The reasonable man attempts to adapt ... - reading.ac.uk

238

6.1.1 The general amino acid permease of Rhizobium leguminosarum

The general amino acid permease of Rhizobium leguminosarum is encoded by

four genes, aapJQMP. Mutation of any one of these genes results in a decrease

in the uptake of amino acids as diverse as leucine, aspartate and histidine, while

over expression of all four genes together leads to a substantial increase in

transport of all the amino acids tested.

The results from amino acid uptake assays on mutants indicate that the

general amino acid permease is the main high-affinity uptake system for

glutamate, aspartate and histidine in free-living cells of R. leguminosarum strain

3841. While this transporter also accounts for a significant percentage of

alanine, leucine and methionine uptake, the residual transport rates in aap

mutants indicate that there are other specific uptake systems for these

substrates. The failure of an aapJQM mutant to grow on proline, suggests that

the general permease is the major uptake system for this amino acid.

Complementation studies indicate that the genes aapJQMP are transcribed

from a single promoter upstream of aapJ, although there may be some

additional weak promoter activity upstream of aapP. The nucleotide sequence

contains a region of dyad symmetry downstream of aapP, that has the

characteristics of a rho-independent transcriptional terminator. It is therefore

likely that these four genes comprise an operon.

The sequence homology data, and in particular the presence of known

signature sequences, suggest that aapJQMP encode the components of an ABC

transporter. The periplasmic binding protein, AapJ, and the two integral

membrane components, AapQ and AapM, are significantly larger than the

equivalent components of previously described ABC transporters of amino

acids. This increased size may be a function of the broad specificity of this

transporter. The deduced polypeptides from the aapJQMP genes exhibit strong

homology to sequence from the Escherichia coli genome sequencing project,

Page 250: The reasonable man attempts to adapt ... - reading.ac.uk

239

suggesting that E. coli may possess a previously unreported general amino acid

permease.

TFASTA and BLAST searches indicate homology of AapQ and AapM to the

integral membrane components of ABC transporters of polar amino acids or

glutamine. In addition to the "integral membrane component signature"

(Saurin et al., 1994), a sequence alignment of AapQ and AapM with these

proteins, reveals a region of 63 amino acids containing 18 highly conserved

residues, that is located, in the majority of cases, at the N-terminal end of the

protein. This suggests that amino acid transporters of this type may constitute a

sub-family of the ABC superfamily.

Predicted topologies containing eight transmembrane segments in the case of

the general permeases, and five membrane-spanning regions in the case of the

other known members of this sub-family, suggest that the N-terminal

conserved region spans two transmembrane segments and a connecting

cytoplasmic loop, while the integral membrane component signature is located

in a cytoplasmic loop. This is consistent with the experimentally determined

topologies of HisM and HisQ of the histidine transporter from S. typhimurium

(Kerppola et al., 1991; Kerppola & Ames, 1992).

For one of the conserved leucines in the N-terminal region, there is a

correlation between the nature of the substrate translocated by proteins in

which it is substituted, and the nature of the amino acid that substitutes. It is

therefore possible that the residue at this position, which is predicted to lie in a

transmembrane segment, is involved in determining substrate specificity.

Furthermore, the deletion from HisM of four amino acids lying in the other

predicted transmembrane segment in this N-terminal region, has been found to

alter the specificity of the histidine transporter in S. typhimurium, from L-

histidine to L-histidinol (Payne et al., 1985).

The rate of exchange of intracellular and extracellular amino acids by strain

3841 is dependent upon the number of copies of the aap operon present,

Page 251: The reasonable man attempts to adapt ... - reading.ac.uk

240

suggesting that the general permease facilitates efflux of amino acids, in

addition to uptake. Translocation of substrate in both directions has not been

previously reported for an ABC transporter, and the most likely explanation is

that the general amino acid permease actively imports amino acids but also

allows outward diffusion. Thus, while the reduction in amino acid exchange in

an aapP mutant of strain 3841 may indicate that ATP hydrolysis is required to

fuel efflux, it seems more likely that the energy is needed to open a membrane

pore upon binding of a liganded binding protein complex. Whether

intracellular substrate has access to such a pore at all times, or is dependent

upon the binding and/or uptake of extracellular substrate for efflux, is an

important mechanistic question, of potential relevance to other ABC

transporters. Since amino acid exchange is reduced in a strain lacking only

AapJ, and assuming that, as in the histidine transporter of S. typhimurium

(Prossnitz et al., 1988; Prossnitz et al., 1989), unliganded binding protein does

not interact with the membrane-bound complex, it is reasonable to conclude

that binding of extracellular substrate, at least, is required for efflux of

intracellular substrate.

The physiological significance of the exchange capability of the general

amino acid permease is unknown. Certainly this capability is not required for

nitrogen fixation by strain 3841, as aap mutants induce pea nodules that reduce

acetylene as effectively as those induced by the wild type. Since the general

amino acid permease appears to be the major high affinity glutamate

transporter in strain 3841, this observation suggests that nitrogen fixation in

Rhizobium leguminosarum is not fuelled via a malate-aspartate shuttle in the

nodule as proposed by Kahn et al. (1985). However, the possibility that

alternative glutamate/aspartate transporters are induced under symbiotic

conditions means that the operation of this shuttle can not be completely

discounted.

Page 252: The reasonable man attempts to adapt ... - reading.ac.uk

241

The results from both aapJQM::lacZ transcriptional fusion studies and

glutamate uptake assays on a ntrC mutant, suggest that the aap operon of R.

leguminosarum strain 3841 is negatively regulated by NtrC in response to

nitrogen supply. In batch culture, nitrogen-excess leads to an approximately 5-

fold repression of the aap operon relative to the nitrogen-limitation.

One potential explanation of this repression that fits the available data, is

that, in contrast to NtrC, phosphorylated NtrC has a lower affinity for the

binding sites upstream of aapJQMP, than it does for those involved in

controlling the expression of other genes, such as glnII, which has been shown

to be positively regulated by NtrC (Carlson et al., 1987; Martin et al., 1988; de

Bruijn et al., 1989; Rossi et al., 1989; Shatters et al., 1989; Patriarca et al., 1992). If

this is the case, then because in R. leguminosarum transcription of ntrC is

essentially independent of the nitrogen status of the cell (Patriarca et al., 1993;

Amar et al., 1994), under nitrogen-limitation, the amount of NtrC (in the form of

NtrC-phosphate) available to bind to the promoter region of the aap operon will

be reduced, and transcription of aapJQMP consequently increased.

Uptake by the general amino acid permease of R. leguminosarum strain 3841

may also be regulated in response to activity of the TCA cycle enzyme α-

ketoglutarate dehydrogenase. Abolition of α-ketoglutarate dehydrogenase

activity through mutation of sucA, results in the almost total loss of uptake by

the general amino acid permease. Mutation of sucD, a gene located upstream of

sucA in the same operon and encoding the alpha subunit of succinyl-CoA

synthetase, reduces α-ketoglutarate dehydrogenase activity by 95% and amino

acid uptake via the general permease by 86%. Mutation of phbC, the gene

encoding poly-beta-hydroxybutyrate synthase, reduces α-ketoglutarate

dehydrogenase activity by 61%, possibly as a result of altering intracellular

NAD(P)H or acetyl-CoA levels, and also causes a 61% reduction in aspartate

uptake.

Page 253: The reasonable man attempts to adapt ... - reading.ac.uk

242

Data from an aapJ::lacZ fusion indicate that there is no change in the

transcription levels of the aap operon in sucDA mutants, and the conditions

encountered by cells prior to transport assays effectively eliminate the

possibility of intracellular substrate inhibition of the general amino acid

permease. This suggests that the general amino acid permease may be subject

to post-translational modification in these mutants.

The intracellular glutamate concentration is elevated in sucDA mutants

grown on glucose/NH4Cl, and it is possible that this elevation is the signal for

the down-regulation of the general amino acid permease. In addition,

glutamate is excreted by sucDA mutants grown on both glucose/NH4Cl and

malate/NH4Cl. These results are consistent with suggestions that the observed

synthesis of glutamate in bacteroids is a consequence of inhibition of α-

ketoglutarate dehydrogenase by NADH (McDermott et al., 1989; Salminen &

Streeter, 1990).

Indeed the inhibition and/or repression of central metabolic enzymes under

the oxygen-limited conditions of the nodule, may account for the observed

excretion of the amino acids alanine and aspartate by bacteroids under nitrogen

fixing conditions (Kretovich et al., 1986; Appels & Haaker, 1991; Kouchi et al.,

1991; Rosendahl et al., 1992). Such excretion may alleviate inhibition of the TCA

cycle by ketoacids and/or reducing equivalents, which would otherwise

accumulate in the non-growing bacteroid cell.

The potential for the involvement of the general amino acid permease in

amino acid excretion by the bacteroid is indicated by the fact that while strain

3841 is found to excrete glutamate when grown on glucose/NH4Cl/aspartate,

an aapJQM mutant exhibits only a relatively low rate of glutamate excretion

under these conditions. This is compatible with the results from amino acid

exchange experiments which indicate that efflux of intracellular amino acid can

be mediated by the general permease.

Page 254: The reasonable man attempts to adapt ... - reading.ac.uk

243

In addition, sucDA mutants of strain 3841 are able to excrete glutamate

despite the fact that glutamate uptake by these mutants is extremely low. Since

glutamate transport in these mutants is apparently not repressed, or subject to

substrate inhibition, this suggests that the general amino acid permease may be

covalently modified so as to reduce uptake but maintain efflux. Modification of

the periplasmic binding protein could potentially alter uptake without affecting

efflux, and the regulation of amino acid uptake via post-translational

modification of a binding protein has been reported previously (Celis, 1984;

Celis, 1990). If the bacteroid is required to excrete amino acids, such

modification may be required to prevent recapture of excreted amino acids by

this high affinity system, while maintaining its ability to mediate efflux.

However, since aap mutants form effective nodules, it is clear that the amino

acid excretion capacity of the general permease is not required for nitrogen

fixation.

In the case of Rhizobium leguminosarum alanine is the predominant amino

acid found to be excreted from bacteroids and symbiosomes under nitrogen-

fixing conditions (Appels & Haaker, 1991; Rosendahl et al., 1992). If this

excretion of alanine is the primary mechanism for maintaining sufficient flux

through the TCA cycle to fuel nitrogen fixation in the bacteroid, then the fact

that the general amino acid permease is not required for an effective symbiosis

is unsurprising: Alanine uptake by aap mutants indicates the existence of one

or more alternative alanine carriers, and the results of exchange experiments on

an aapJQM mutant, using the alanine analogue AIB, suggest that such a carrier

(or carriers) allows efflux of substrate.

6.1.2 TCA cycle enzymes in Rhizobium leguminosarum

Partial sequencing of a cosmid that complements a mutation in sucA,

suggests that in R. leguminosarum strain 3841 the genes encoding the TCA cycle

enzymes malate dehydrogenase, succinyl-CoA synthetase and α-ketoglutarate

dehydrogenase are clustered, and occur in the order mdh-sucCD-sucAB. The

Page 255: The reasonable man attempts to adapt ... - reading.ac.uk

244

results of complementation studies suggest that sucCDAB are under the control

of a single promoter, and it is possible that the same promoter also initiates

transcription of mdh, although this was not verified.

The presence of mdh-sucCDAB in a single operon could account for the

elevated malate dehydrogenase levels observed in succinyl-CoA synthetase and

α-ketoglutarate dehydrogenase mutants, and the increased succinyl-CoA

synthetase activity found in α-ketoglutarate dehydrogenase mutants, if a

mutation in sucCDAB leads to an increase in the expression of the mdh-sucCDAB

operon. One potential reason for such an increase in expression, is that the mdh-

sucCDAB operon may be regulated in response to metabolite level(s) that are

altered as a result of the mutation.

Attempts to isolate malate dehydrogenase mutants of strain 3841 by

homogenotization of cosmids mutated in mdh, were unsuccessful, suggesting

that loss of this enzyme may be lethal to R. leguminosarum.

6.1.3 Methionine biosynthetic enzymes in Rhizobium leguminosarum

On the basis of sequence homology, the gene lying immediately upstream of

the aap operon in R. leguminosarum strain 3841, is likely to be metC, encoding

beta-cystathionase. Similarly, the gene lying ~3.9kb upstream of the aap operon

is likely to be cysE, encoding serine acetyltransferase. Both these enzymes are

involved in methionine biosynthesis and the genes are transcribed in the

opposite direction to the aap operon. However, metC and cysE are apparently

not contained in the same operon, since activities of lacZ fusions to the metC-

cysE intergenic region of ~2.4kb, are indicative of transcription in the opposite

direction to metC and cysE. Data from cysE::lacZ and metC::lacZ fusions suggest

that both cysE and metC are negatively regulated by NtrC.

Mutation of cysE causes a reduction of amino acid transport in strain 3841

grown on glutamate as the nitrogen source. This indicates that the ability of

amino acid transport in R. leguminosarum to respond normally to nitrogen-

limitation is dependent upon the presence of serine acetyltransferase.

Page 256: The reasonable man attempts to adapt ... - reading.ac.uk

245

However, the homology of CysE to NifP from Azotobacter, and the observation

that a cysE mutant is not a cysteine auxotroph, suggest that R. leguminosarum

may possess two genes encoding serine acetyltransferase. This may account for

the wild-type amino acid transport rates exhibited by a cysE mutant of strain

3841 under nitrogen-excess.

6.1.4 Future work

Because of the unusual characteristics of the R. leguminosarum general amino

acid permease, the general amino acid permeases of other bacteria, particularly

E. coli, probably warrant further investigation.

Since the nucleotide sequence of the DNA flanking the aapJQMP operon of E.

coli is known, the DNA containing these genes could be amplified from

chromosomal DNA by PCR and hence cloned. This would allow investigation

of the effects of mutation and over expression of these genes on amino acid

transport in E. coli. The nucleotide sequence in the two areas of the E. coli

sequence where frame shifts are apparent from the R. leguminosarum sequence,

could also be checked.

It may be possible to investigate the extent of distribution of this transporter

among the bacterial species through a combination of PCR, using degenerate

primers designed from the known aapJ sequences of R. leguminosarum, E. coli

and P. fluorescens, to amplify a band of known size from chromosomal DNA

containing an aapJ gene; and Southern blot, using R. leguminosarum aapJ DNA as

the probe, to confirm the nature of the amplified band.

Further study of the R. leguminosarum general amino acid permease could

yield structural and mechanistic details of general applicability to ABC

transporters. In order to investigate whether the general permease actively

exports amino acids, or simply enables diffusion, the transporter could be

reconstituted in E. coli inside-out membrane vesicles, and amino acid uptake by

these vesicles measured in both the presence and absence of ATP. The topology

of the integral membrane components AapM and AapQ could be investigated

Page 257: The reasonable man attempts to adapt ... - reading.ac.uk

246

by TnphoA fusion analysis, while the possible involvement of particular

residues in these proteins in substrate specificity could be tested through site-

directed mutagenesis.

Since alanine excretion may be important to bacteroid function, and this

excretion may be mediated by an alanine-specific transporter(s) in R.

leguminosarum, the cloning of the gene(s) encoding alanine-specific uptake

system(s) could be attempted. Cloning of this gene(s) would enable the

symbiotic effectiveness of strains mutated in both the general amino acid

permease and alanine-specific transport to be investigated. Strategies for

cloning alanine-specific transporters include looking for increased labelling by

[14C]alanine in an aap mutant carrying an aap mutant chromosomal library; or

obtaining alanine transport mutants by screening for loss of growth on alanine

in transposon mutants of an aap-deleted strain, and then complementing the

mutation from an aap mutant chromosomal library.

The question of whether mdh lies in the same operon as sucCDAB in R.

leguminosarum could be resolved by attempting to complement a sucCDAB

mutation with a clone of sucCDAB that includes the complete mdh-sucC

intergenic region, or by cloning the mdh-sucC intergenic region into a promoter

probe vector. Both these options would require further restriction mapping of

the mdh-sucCDAB region, in order to reveal suitable fragments for cloning.

Investigation of the activity of lacZ fusions to mdh and sucCD, in sucAB mutant

backgrounds, would reveal whether elevated malate dehydrogenase and

succinyl-CoA synthetase activities in α-ketoglutarate dehydrogenase mutants

are the result of increased transcription.

Page 258: The reasonable man attempts to adapt ... - reading.ac.uk

247

REFERENCES

Page 259: The reasonable man attempts to adapt ... - reading.ac.uk

248

Adams, M.D., Wagner, L.M., Graddis, T.J., Landick, R. and Antonucci, T.K.

(1990) Nucleotide sequence and genetic characterization reveal six essential

genes for the LIV-I and LS transport systems of Escherichia coli. Journal of

Biological Chemistry 265: 11436-11443.

Adams, M.D. and Oxender, D.L. (1989) Bacterial periplasmic binding protein

tertiary structures. Journal of Biological Chemistry 264: 15739-15742.

Alloing, G., Trombe, M.-C. and Claverys, J.-P. (1990) The ami locus of the

Gram-positive bacterium Streptococcus pneumoniae is similar to binding protein-

dependent transport operons of Gram-negative bacteria. Molecular Microbiology

4: 633-644.

Amar, M., Patriarca, E.J., Manco, G., Bernard, P., Riccio, A., Lamberti, A., Defez,

R. and Iaccarino, M. (1994) Regulation of nitrogen metabolism is altered in a

glnB mutant strain of Rhizobium leguminosarum. Molecular Microbiology 11: 685-

693.

Amarasingham, C.R. and Davis, B.D. (1965) Regulation of α-ketoglutarate

dehydrogenase formation in Escherichia coli. Journal of Biological Chemistry 240:

3664-3668.

Ames, G.F.-L. (1972) Components of histidine transport. In Biological

Membranes. Proceedings of the 1972 ICN-UCLA Symposium in Molecular Biology.

Fox, C.F. (ed). New York: Academic, pp. 409-426.

Ames, G.F.-L. (1985) The histidine transport system of Salmonella typhimurium.

In Current Topics in Membranes and Transport. New York: Academic, pp. 103-

119.

Ames, G.F.-L. (1986) Bacterial periplasmic transport systems: structure,

mechanism, and evolution. Annual Review of Biochemistry 55: 397-425.

Page 260: The reasonable man attempts to adapt ... - reading.ac.uk

249

Ames, G.F.-L., Nikaido, K., Groarke, J. and Petithory, J. (1989) Reconstitution

of periplasmic transport in inside-out membrane vesicles: energization by ATP.

Journal of Biological Chemistry 264: 3998-4002.

Ames, G.F.-L., Noel, K.D., Taber, H., Spudich, E.N., Nikaido, K., Afong, J. and

Ardeshir, F. (1977) Fine-structure map of the histidine transport genes in

Salmonella typhimurium. Journal of Bacteriology 129: 1289-1297.

Ames, G.F.-L. and Joshi, A. (1991) Energy coupling in bacterial periplasmic

permeases. Journal of Bacteriology 172: 4122-4137.

Ames, G.F.-L. and Nikaido, K. (1981) Phosphate-containing proteins of

Salmonella typhimurium and Escherichia coli. European Journal of Biochemistry 115:

525-531.

Ames, G.F.-L. and Nikaido, K. (1985) Nitrogen regulation in Salmonella

typhimurium. Identification of an ntrC protein-binding site and definition of a

consensus binding sequence. EMBO Journal 4: 539-547.

Ames, G.F.-L. and Spudich, E.N. (1976) Protein-protein interaction in

transport: periplasmic histidine-binding protein J interacts with P protein.

Proceedings of the National Academy of Sciences of the United States of America 73:

1877-1881.

Anderson, M.P., Berger, H.A., Rich, D.P., Gregory, R.J., Smith, A.E. and Welsh,

M.J. (1991a) Nucleotide triphosphates are required to open the CFTR chloride

channel. Cell 67: 775-784.

Anderson, M.P., Rich, D.P., Gregory, R.J., Smith, A.E. and Welsh, M.J. (1991b)

Generation of cAMP activated chloride currents by expression of CFTR. Science

251: 679-682.

Page 261: The reasonable man attempts to adapt ... - reading.ac.uk

250

Antonucci, T.K. and Oxender, D.L. (1986) The molecular biology of amino acid

transport in bacteria. Advances in Microbial Physiology 28: 145-180.

Appels, M.A. and Haaker, H. (1991) Glutamate oxaloacetate transaminase in

pea root nodules -participation in a malate/aspartate shuttle between plant and

bacteroid. Plant Physiology 95: 740-747.

Appleby, C.A. (1984) Leghemoglobin and Rhizobium respiration. Annual

Review of Plant Physiology 35: 443-478.

Argast, M. and Boos, W. (1979) Purification and properties of the sn-glycerol-3-

phosphate-binding protein of Escherichia coli. Journal of Biological Chemistry 254:

10931-10935.

Arias, A., Cervenansky, C., Gardiol, A. and Martinez-drets, G. (1979) Phospho-

glucose isomerase mutant of Rhizobium meliloti. Journal of Bacteriology 137: 409-

414.

Arwas, R., McKay, I.A., Rowney, F.R.P., Dilworth, M.J. and Glenn, A.R. (1985)

Properties of organic acid utilization mutants of Rhizobium leguminosarum strain

300. Journal of General Microbiology 131: 2059-2066.

Arwas, R., Glenn, A.R., McKay, I.A. and Dilworth, M.J. (1986) Properties of

double mutants of Rhizobium leguminosarum which are defective in the

utilization of dicarboxylic acids and sugars. Journal of General Microbiology 132:

2743-2747.

Baichwal, V., Liu, D.X. and Ames, G.F.L. (1993) The ATP-binding component

of a prokaryotic traffic ATPase is exposed to the periplasmic (external) surface.

Proceedings of the National Academy of Sciences of the United States of America 90:

620-624.

Page 262: The reasonable man attempts to adapt ... - reading.ac.uk

251

Bell, A.W., Buckel, S.D., Groarke, J.M., Hope, J.N., Kingsley, D.H. and

Hermodson, M.A. (1986) The nucleotide sequence of the rbsD, rbsA and rbsC

genes of Escherichia coli. Journal of Biological Chemistry 261: 7652-7658.

Bender, D.A. (1978) Amino Acid Metabolism. Chichester: John Wiley & Sons.

Berg, D.E. (1977) Insertion and excision of the transposable kanamycin

resistant determinant. In DNA Insertion Elements, Plasmids and Episomes.

Bukhari, A.I., Shapiro, J.A. and Adhya, S.L. (eds). New York: Cold Spring

Harbor, pp. 205-212.

Bergersen, F.J. and Turner, G.L. (1988) Glutamate as a carbon source for N2-

fixing bacteroids prepared from soybean root nodules. Journal of General

Microbiology 134: 2441-2448.

Bergersen, F.J. and Turner, G.L. (1990) Bacteroids from soybean root nodules:

accumulation of poly-β-hydroxybutyrate during supply of malate and succinate

in relation to N2 fixation in flow-chamber reactions. Proceedings Of The Royal

Society Of London Series B-Biological Sciences 240: 39-59.

Bergmeyer, H.U., Bernt, E., Mollering, H. and Pfleiderer, G. (1974) L-Aspartate

and L-asparagine. In Methods of Enzymatic Analysis. Bergmeyer, H.U. (ed).

New York: Academic Press, pp. 1696-1700.

Beringer, J.E. (1974) R factor transfer in Rhizobium leguminosarum. Journal of

General Microbiology 84: 188-198.

Beringer, J.E., Beynon, J.L., Buchanan-Wollaston, A.V. and Johnston, A.W.B.

(1978) Transfer of the drug-resistance transposon Tn5 to Rhizobium. Nature 276:

633-634.

Page 263: The reasonable man attempts to adapt ... - reading.ac.uk

252

Bernt, E. and Bergmeyer, H.U. (1974) L-Glutamate UV-assay with glutamate

dehydrogenase and NAD. In Methods of Enzymatic Analysis. Bergmeyer, H.U.

(ed). New York: Academic Press, pp. 1704-1708.

Bishop, L., Agbayani, R., Ambudkar, S.V., Maloney, P.C. and Ames, G.F.-L.

(1989) Reconstitution of a bacterial periplasmic permease in proteoliposomes

and demonstration of ATP hydrolysis concomitant with transport. Proceedings

of the National Academy of Sciences of the United States of America 86: 6953-6957.

Bolton, E., Higgisson, B., Harrington, A. and O'Gara, F. (1986) Dicarboxylic

acid transport in Rhizobium meliloti: isolation of mutants and cloning of

dicarboxylic acid transport genes. Archives of Microbiology 144: 142-146.

Bradford, M.M. (1976) A rapid and sensitive method for the quantitation of

microgram quantities of protein utilizing the principle of protein-dye binding.

Analytical Biochemistry 72: 248-254.

Brass, J.M., Higgins, C.F., Foley, M., Rugman, P.A., Birmingham, J. and

Garland, P.B. (1986) Lateral diffusion of proteins in the periplasm of Escherichia

coli. Journal of Bacteriology 165: 787-794.

Brown, C.M. and Dilworth, M.J. (1975) Ammonia assimilation by Rhizobium

cultures and bacteroids. Journal of General Microbiology 86: 39-48.

Buchanan-Wollaston, V. (1979) Generalized transduction in Rhizobium

leguminosarum. Journal of General Microbiology 112: 135-142.

Buck, D., Spencer, M.E. and Guest, J.R. (1986) Cloning and expression of the

succinyl-CoA synthetase genes of Escherichia coli K12. Journal of General

Microbiology 132: 1753-1762.

Page 264: The reasonable man attempts to adapt ... - reading.ac.uk

253

Bueno, R., Pahel, G. and Magasanik, B. (1985) Role of glnB and glnD gene

products in regulation of the glnALG operon of Escherichia coli. Journal of

Bacteriology 164: 816-822.

Callaham, D.A. and Torrey, J.G. (1981) The structural basis for infection of root

hairs of Trifolium repens by Rhizobium. Canadian Journal of Botany 59: 1647-1664.

Carlson, T.A., Martin, G.B. and Chelm, B.K. (1987) Differential transcription of

the two glutamine synthetase genes of Bradyrhizobium japonicum. Journal of

Bacteriology 169: 5861-5866.

Celis, R.T.F. (1984) Phosphorylation in vivo and in vitro of the arginine-

ornithine periplasmic transport protein of Escherichia coli. European Journal of

Biochemistry 145: 403-411.

Celis, R.T.F. (1990) Mutant of Escherichia coli K-12 with defective

phosphorylation of two periplasmic transport proteins. Journal of Biological

Chemistry 265: 1787-1793.

Cervenansky, C. and Arias, A. (1984) Glucose 6-phosphate dehydrogenase

deficiency in pleiotropic carbohydrate-negative mutant strains of Rhizobium

meliloti. Journal of Bacteriology 160: 1027-1030.

Chiurazzi, M., Meza, R., Lara, M., Lahm, A., Defez, R., Iaccarino, M. and Espin,

G. (1992) The Rhizobium leguminosarum biovar phaseoli glnT gene, encoding

glutamine synthetase III. Gene 119: 1-8.

Chiurazzi, M. and Iaccarino, M. (1990) Transcriptional analysis of the glnB-

glnA region of Rhizobium leguminosarum biovar viciae. Molecular Microbiology 4:

1727-1735.

Page 265: The reasonable man attempts to adapt ... - reading.ac.uk

254

Claros, M.G. and von Heijne, G. (1994) TOP-PRED II: An improved software

for membrane protein structure predictions. Computer Applications in the

Biosciences 10: 685-686.

Claverie-Martin, F. and Magasanik, B. (1991) Role of integration host factor in

the regulation of the glnHp2 promoter of Escherichia coli. Proceedings of the

National Academy of Sciences of the United States of America 88: 1631-1635.

Colonna-Romano, S., Riccio, A., Guida, M., Defez, R., Lamberti, A., Iaccarino,

M., Arnold, W., Priefer, U. and Puhler, A. (1987) Tight linkage of glnA and a

putative regulatory gene in Rhizobium leguminosarum. Nucleic Acids Research 15:

1951-1964.

Copeland, L., Quinnell, R.G. and Day, D.A. (1989a) Malic enzyme activity in

bacteroids from soybean nodules. Journal of General Microbiology 135: 2005-

2011.

Copeland, L., Vella, J. and Hong, Z.Q. (1989b) Enzymes of carbohydrate

metabolism in soybean nodules. Phytochemistry 28: 57-61.

Cutting, G.R., Kasch, L.M., Rosenstein, B.J., Zeilinski, J., Tsui, L.-C.,

Antonarakis, S.E. and Kazazian, H.H. (1990) A cluster of cystic fibrosis

mutations in the first nucleotide-binding fold of the cystic fibrosis conductance

regulator protein. Nature 346: 366-369.

Dahl, M.K. and Manson, M.D. (1985) Interspecific reconstitution of maltose

transport and chemotaxis in Escherichia coli with maltose binding protein from

various enteric bacteria. Journal of Bacteriology 164: 1051-1063.

Darrow, R.A.a. and Knots, R.R. (1977) Two forms of glutamine synthesase in

free living and root nodule bacteria. Biochemical and Biophysical Research

Communications 554-559.

Page 266: The reasonable man attempts to adapt ... - reading.ac.uk

255

Dassa, E. and Hofnung, M. (1985) Sequence of gene malG in E. coli K12:

homologies between integral membrane components from binding protein-

dependent transport systems. EMBO Journal 4: 2287-2293.

Davidson, A.L., Shuman, H.A. and Nikaido, H. (1992) Mechanism of maltose

transport in Escherichia-coli-transmembrane signalling by periplasmic binding

proteins. Proceedings of the National Academy of Sciences of the United States of

America 89: 2360-2364.

Davidson, A.L. and Nikaido, H. (1990) Overproduction, solubilization and

reconstitution of the maltose transport system from Escherichia coli. Journal of

Biological Chemistry 265: 4254-4260.

Davidson, A.L. and Nikaido, H. (1991) Purification and characterization of the

membrane-associated components of the maltose transport system from

Escherichia coli. Journal of Biological Chemistry 266: 8946-8951.

Day, D.A. and Udvardi, M.K. (1989) Membrane interface of the Bradyrhizobium

japonicum-Glycine max symbiosis: peribacteroid units from soybean nodules.

Australian Journal of Plant Physiology 16: 69-84.

Dazzo, F.B. and Gardiol, A. (1984) Host specificity in Rhizobium-legume

interactions. In Genes Involved in Microbe-Plant Interactions. New York:

Springer Publishers, pp. 3-31.

de Bruijn, F.J., Rossbach, S., Schneider, M., Ratet, P., Messmer, S., Szeto, W.W.,

Ausubel, F.M. and Schell, J. (1989) Rhizobium meliloti 1021 has three

differentially regulated loci involved in glutamine biosynthesis, none of which

is essential for symbiotic nitrogen fixation. Journal of Bacteriology 171: 1673-

1682.

Page 267: The reasonable man attempts to adapt ... - reading.ac.uk

256

de Maagd, R.A. and Lugtenberg, B. (1986) Fractionation of Rhizobium

leguminosarum cells into outer membrane, cytoplasmic membrane, periplasmic,

and cytoplasmic components. Journal of Bacteriology 167: 1083-1085.

de Vries, G.E., van Brussel, A.A.N. and Quispel, A. (1982) Mechanism and

regulation of glucose transport in Rhizobium leguminosarum. Journal of

Bacteriology 149: 872-879.

Dean, D.A., Fikes, J.D., Gehring, K., Bassford, P.J. and Nikaido, H. (1989)

Active transport of maltose in membrane vesicles obtained from Escherichia coli

cells producing tethered maltose-binding protein. Journal of Bacteriology 171:

503-510.

DeBusk, R.M. and DeBusk, D.E. (1980) Physiological and regulatory properties

of the general amino acid transport system of Neurospora crassa. Journal of

Bacteriology 143: 188-197.

Deguchi, Y., Yamato, I. and Anraku, Y. (1989) Molecular cloning of gltS and

gltP, which encode glutamate carriers of Escherichia coli B. Journal of Bacteriology

171: 1314-1319.

Deguchi, Y., Yamato, I. and Anraku, Y. (1990) Nucleotide sequence of gltS, the

Na+/glutamate symport carrier gene of Escherichia coli B. Journal of Biological

Chemistry 265: 21704-21708.

DeKok, A., Kornfeld, S., Benziman, M. and Milner, Y. (1980) Subunit

composition and partial reactions of the 2-oxoglutarate dehydrogenase complex

of Acetobacter xylinum. European Journal of Biochemistry 106: 49-58.

DeVries, G.E., in T'Veld, P. and Kijne, J.W. (1980) Production of organic acids

in Pisum sativum root nodules as a result of oxygen stress. Plant Science Letters

20: 115-123.

Page 268: The reasonable man attempts to adapt ... - reading.ac.uk

257

Dierks, T. and Krämer, R. (1988) Asymmetric orientation of the reconstituted

aspartate/glutamate carrier from mitochondria. Biochimica et Biophysica Acta

937: 112-126.

Dila, D.K. and Maloy, S.R. (1986) Proline transport in Salmonella typhimurium:

putP permease mutants with altered substrate specificity. Journal of Bacteriology

168: 590-594.

Dilworth, M.J., Arwas, R., McKay, I.A., Saroso, S. and Glenn, A.R. (1986)

Pentose metabolism in Rhizobium leguminosarum MNF300 and in cowpea

Rhizobium NGR234. Journal of General Microbiology 132: 2733-2742.

Dilworth, M.J., McKay, I.A. and Glenn, A.R. (1988) Dicarboxylate metabolism

in rhizobia needs malic enzyme. In Nitrogen Fixation: Hundred Years After.

Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds). New York: Gustav Fischer,

pp. 555

Dilworth, M.J. and Glenn, A.R. (1982) Movements of ammonia in Rhizobium

leguminosarum. Journal of General Microbiology 128: 29-37.

Dixon, R. (1984) Tandem promoters determine regulation of the Klebsiella

pneumoniae glutamine synthetase (glnA) gene. Nucleic Acids Research 12: 7811-

7830.

Doige, C.A. and Ames, G.F.L. (1993) ATP-dependent transport systems in

bacteria and humans: relevance to cystic fibrosis and multidrug resistance.

Annual Review of Microbiology 47: 291-319.

Downie, J.A., Ma, Q.S., Knight, C.D., Hombrecher, G. and Johnston, A.W.B.

(1983) Cloning of the symbiotic region of Rhizobium leguminosarum - the

nodulation genes are between the nitrogenase genes and a nifA-like gene.

EMBO Journal 2: 947-952.

Page 269: The reasonable man attempts to adapt ... - reading.ac.uk

258

Driscoll, B.T. and Finan, T.M. (1993) NAD+-dependent malic enzyme of

Rhizobium meliloti is required for symbiotic nitrogen fixation. Molecular

Microbiology 7: 865-873.

Dudler, R., Schmidhauser, C., Parish, R.W., Wettenhall, R.E.H. and Schmidt, T.

(1988) A Mycoplasma high-affinity transport system and the in vitro

invasivenass of mouse sarcoma cells. EMBO Journal 7: 3963-3970.

Duncan, M.J. (1981) Properties of Tn5-induced carbohydrate mutants in

Rhizobium meliloti. Journal of General Microbiology 122: 61-67.

Duncan, M.J. and Fraenkel, D.G. (1979) α-Ketoglutarate dehydogenase mutant

of Rhizobium meliloti. Journal of Bacteriology 37: 415-419.

Eddy, A.A. (1982) Mechanisms of solute transport. Advances in Microbial

Physiology 23: 1-78.

Ehrmann, M., Boyd, D. and Beckwith, J. (1990) Genetic analysis of membrane

protein topology by a sandwich gene fusion approach. Proceedings of the

National Academy of Sciences of the United States of America 87: 7574-7578.

El-Guezzar, M., Hornez, J.P., Courtois, B. and Derieux, J.C. (1988) Study of a

fructose-negative mutant of Rhizobium meliloti. FEMS Microbiology Letters 49:

429-434.

Emerich, D.W. (1985) Characterization of carbon metabolism in Rhizobium

japonicum bacteroids. In Nitrogen Fixation and CO2 Metabolism. Ludden, P.W.

and Burris, J.E. (eds). New York: Elsevier Science Publishing Co. pp. 21-30.

Page 270: The reasonable man attempts to adapt ... - reading.ac.uk

259

Emerich, D.W., Anthon, G.E., Hayes, R.R., Karr, D.B., Liang, R., Preston, G.G.,

Smith, M.T. and Waters, J.K. (1988) Metabolism of Rhizobium-leguminous plant

nodules with an emphasis on bacteroid carbon metabolism. In Nitrogen

Fixation: Hundred Years After. Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds).

New York: Gustav Fischer, pp. 539-546.

Engelke, T., Jagadish, M.N. and Pühler, A. (1987) Biochemical and genetical

analysis of Rhizobium meliloti mutants defective in C4-dicarboxylate transport.

Journal of General Microbiology 133: 3019-3029.

Engelman, D.M., Steitz, T.A. and Goldman, A. (1986) Identifying nonpolar

transbilayer helices in amino acid sequences of membrane proteins. Annual

Review of Biophysics and Biophysical Chemistry 15: 321-353.

Espin, G., Moreno, S., Wild, M., Meza, R., Iaccarino, M. (1990) A previously

unrecognized glutamine synthetase expressed in Klebsiella pneumoniae from the

glnT locus of Rhizobium leguminosarum. Molecular and General Genetics 223: 513-

516.

Evans, D.J., Jones, R., Woodley, P.R., Wilborn, J.R. and Robson, R.L. (1991)

Nucleotide-sequence and genetic-analysis of the Azotobacter chroococcum

nifUSVWZM gene-cluster, including a new gene (nifP) which encodes a serine

acetyltransferase. Journal of Bacteriology 173: 5457-5469.

Figurski, D.H. and Helinski, D.R. (1979) Replication of an origin-containing

derivative of plasmid RK2 dependent on a plasmid function provided in trans.

Proceedings of the National Academy of Sciences of the United States of America 76:

1648-1652.

Fikes, J.D. and Bassford, P.J. (1987) Export of unprocessed precursor maltose-

binding protein to the periplasm of Escherichia coli cells. Journal of Bacteriology

169: 2352-2359.

Page 271: The reasonable man attempts to adapt ... - reading.ac.uk

260

Filser, M.M.K., Moscatelli, C., Lamberti, A., Vincze, E., Guida, M., Salzano, G.

and Iaccarino, M. (1986) Characterization and cloning of two Rhizobium

leguminosarum genes coding for glutamine synthetase activities. Journal of

General Microbiology 132: 2561-2569.

Finan, T.M., Wood, J.M. and Jordan, C. (1981) Succinate transport in Rhizobium

leguminosarum. Journal of Bacteriology 148: 193-202.

Finan, T.M., Wood, J.M. and Jordan, D.C. (1983) Symbiotic properties of C4-

dicarboxylic acid transport mutants of Rhizobium leguminosarum. Journal of

Bacteriology 154: 1403-1413.

Fisher, R.F. and Long, S.R. (1992) Rhizobium - plant signal exchange. Nature

357: 655-660.

Fitzmaurice, A.M. and O'Gara, F. (1988) Involvement of glutamate as a carbon

source in supporting nitrogen fixation activity in R. meliloti. In Nitrogen

Fixation: Hundred Years After. Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds).

New York: Gustav Fischer, pp. 558

Fitzmaurice, A.M. and O'Gara, F. (1991) Glutamate catabolism in Rhizobium

meliloti. Archives of Microbiology 155: 422-427.

Fottrell, P.F. and Mooney, P. (1969) The regulation of some enzymes involved

in ammonia assimilation by Rhizobium japonicum. Journal of General Microbiology

59: 211-214.

Freidman, A.M., Long, S.R., Brown, S.E., Buikema, W.J. and Ausubel, F.M.

(1982) Construction of the broad host range cosmid cloning vector and its use

in the genetic analysis of Rhizobium mutants. Gene 18: 289-296.

Page 272: The reasonable man attempts to adapt ... - reading.ac.uk

261

Froshauer, S., Green, G.N., Boyd, D., McGovern, K. and Beckwith, J. (1988)

Genetic analysis of the membrane insertion and topology of MalF, a

cytoplasmic membrane protein of Escherichia coli. Journal of Molecular Biology

200: 501-511.

Fuchs, R.L. and Keister, D.L. (1980) Comparative properties of glutamine

synthetases I and II in Rhizobium and Agrobacterium spp. Journal of Bacteriology

144: 641-648.

Furlong, C.E. (1987) Osmotic-shock-sensitive transport systems. In Escherichia

coli and Salmonella typhimurium: Cellular and molecular biology. Neidhardt, F.C.,

Ingraham, J.L., Low, K.B., Magasanik, B., Schaechter, M. and Umbarger, H.E.

(eds). Washington: American Society for Microbiology, pp. 768-796.

Gallagher, M.P., Pearce, S.R. and Higgins, C.F. (1989) Identification and

localization of the membrane-associated, ATP-binding subunit of the

oligopeptide permease of Salmonella typhimurium. European Journal of

Biochemistry 180: 133-141.

Gardiol, A.E., Truchet, G.L. and Dazzo, F.B. (1987) Requirement of succinate

dehydrogenase activity for symbiotic bacteroid differentiation of Rhizobium

meliloti in alfalfa nodules. Applied and Environmental Microbiology 53: 1947-1950.

Gill, D.R., Hyde, S.C., Higgins, C.F., Valverde, M.A., Mintenig, G.M. and

Sepulveda, F.V. (1992) Separation of drug transport and chloride channel

functions of the human multidrug resistance p-glycoprotein. Cell 71: 23-32.

Gilson, E., Alloing, G., Schmidt, T., Claverys, J.-P., Dudler, R. and Hofnung, M.

(1988) Evidence for high affinity binding-protein-dependent transport systems

in Gram-positive bacteria and Mycoplasma. EMBO Journal 7: 3971-3974.

Page 273: The reasonable man attempts to adapt ... - reading.ac.uk

262

Glenn, A.R., McKay, I.A., Arwas, R. and Dilworth, M.J. (1984) Sugar

metabolism and the symbiotic properties of carbohydrate mutants of Rhizobium

leguminosarum. Journal of General Microbiology 130: 239-245.

Glenn, A.R., Holliday, S. and Dilworth, M.J. (1991) The transport and

catabolism of L-proline by cowpea Rhizobium NGR 234. FEMS Microbiology

Letters 82: 307-312.

Glenn, A.R. and Brewin, N.J. (1981) Succinate-resistant mutants of Rhizobium

leguminosarum. Journal of General Microbiology 126: 237-241.

Glenn, A.R. and Dilworth, M.J. (1979) An examination of Rhizobium

leguminosarum for the production of extracellular and periplasmic proteins.

Journal of General Microbiology 112: 405-409.

Glenn, A.R. and Dilworth, M.J. (1981) The uptake and hydrolysis of

disaccharides by fast- and slow-growing species of Rhizobium. Archives of

Microbiology 129: 233-239.

Gordon, A.J., Ryle, G.J.A., Mitchell, D.F. and Powell, C.E. (1985) The flux of 14C-labeled photosynthate through soyabean root nodules during N2 fixation.

Journal of Experimental Botany 36: 756-769.

Gray, C.T., Wimpenny, J.W.T. and Mossman, M.R. (1966) Regulation of

metabolism in facultative bacteria. II. Effects of aerobiosis, anaerobiosis and

nutrition on the formation of Krebs cycle enzymes in Escherichia coli. Biochimica

et Biophysica Acta 117: 33-41.

Grenson, M., Hou, C. and Crabbel, M. (1970) Multiplicity of the amino acid

permeases in Saccharomyces cerevisiae. Journal of Bacteriology 103: 770-777.

Page 274: The reasonable man attempts to adapt ... - reading.ac.uk

263

Grenson, M. and Hon, C. (1972) Ammonia inhibition of the general amino acid

permease and its suppression in NADPH-specific glutamate dehydrogenaseless

mutants of Saccharomyces cerevisiae. Biochemical and Biophysical Research

Communications 48: 749-756.

Guest, J.R. (1992) Oxygen-regulated gene expression in Escherichia coli. Journal

of General Microbiology 138: 2253-2263.

Guest, J.R. and Russell, G.C. (1992) Complexes and complexities of the citric

acid cycle in Escherichia coli. Current Topics in Cellular Regulation 33: 231-247.

Hall, E.R. and Weitzman, P.D.J. (1977) Evidence for allosteric NADH

regulation of cinetobacter α-oxoglutarate dehydrogenase from multiple-

inhibition studies. Biochemical and Biophysical Research Communications 74: 1613-

1617.

Halpern, Y.S. (1974) Genetics of amino acid transport in bacteria. Annual

Review of Genetics 8: 103-133.

Hanahan, D. (1983) Studies on transformation of Escherichia coli with plasmids.

Journal of Molecular Biology 166: 557-580.

Haney, S.A. and Oxender, D.L. (1992) Amino acid transport in bacteria.

International Review Of Cytology 137A: 37-95.

Hawkes, T., Merrick, M. and Dixon, R. (1985) Interaction of purified NtrC

protein with nitrogen regulated promoter from Klebsiella pneumoniae. Molecular

and General Genetics 201: 492-498.

Hengge, R. and Boos, W. (1983) Maltose and lactose transport in Escherichia

coli. Examples of two different types of concentrative transport systems.

Biochimica et Biophysica Acta 737: 443-478.

Page 275: The reasonable man attempts to adapt ... - reading.ac.uk

264

Herrada, G., Puppo, A. and Rigaud, J. (1989) Uptake of metabolites by

bacteriod-containing vesicles and by free bacteroids from french bean nodules.

Journal of General Microbiology 135: 3165-3177.

Higgins, C.F., Ames, G.F.-L., Barnes, W.M., Clement, J.M. and Hofnung, M.

(1982a) A novel intercistronic regulatory element of prokaryotic operons.

Nature 298: 760-762.

Higgins, C.F., Haag, P.D., Nikaido, K., Ardeshir, F., Garcia, G. and Ames, G.F.-

L. (1982b) Complete nucleotide sequence and identification of membrane

components of the histidine transport operon of S. typhimurium. Nature 298:

723-727.

Higgins, C.F., Hiles, I.D., Whalley, K. and Jamieson, D.J. (1985) Nucleotide

binding by membrane components of bacterial periplasmic binding protein-

dependent transport systems. EMBO Journal 4: 1033-1040.

Higgins, C.F., Hiles, I.D., Salmond, G.P.C., Gill, D.R., Downie, J.A., Evans, I.J.,

Holland, I.B., Gray, L., Buckel, S.D., Bell, A.W. and Hermodson, M.A. (1986) A

family of related ATP-binding subunits coupled to many distinct biological

processes in bacteria. Nature 323: 448-450.

Higgins, C.F., Sutherland, L., Cairney, J. and Booth, I.R. (1987) The osmotically

regulated proU locus of Salmonella typhimurium encodes a periplasmic betaine-

binding protein. Journal of General Microbiology 133: 305-310.

Higgins, C.F., Mclaren, R.S. and Newbury, S.F. (1988) Repetitive extragenic

palindromic sequences, mRNA stability and gene expression: evolution by gene

conversion? - A review. Gene 72: 3-14.

Page 276: The reasonable man attempts to adapt ... - reading.ac.uk

265

Higgins, C.F., Hyde, S.C., Mimmack, M.M., Gileadi, U., Gill, D.R. and

Gallagher, M.P. (1990) Binding protein-dependent transport systems. Journal

of Bioenergetics and Biomembranes 22: 571-591.

Higgins, C.F. (1992) ABC transporters: from microorganisms to man. Annual

Review of Cell Biology 8: 67-113.

Higgins, C.F. and Ames, G.F.-L. (1981) Two periplasmic proteins which

interact with a common membrane receptor show extensive homology:

complete nucleotide sequences. Proceedings of the National Academy of Sciences of

the United States of America 78: 6038-6042.

Higgins, C.F. and Ames, G.F.-L. (1982) Regulatory regions of two transport

operons under nitrogen control: Nucleotide sequences. Proceedings of the

National Academy of Sciences of the United States of America 79: 1083-1087.

Hiles, I.D., Gallagher, M.P., Jamieson, D.J. and Higgins, C.F. (1987) Molecular

characterization of the oligonucleotide permease of Salmonella typhimurium.

Journal of Molecular Biology 195: 125-142.

Hiles, I.D. and Higgins, C.F. (1986) Peptide uptake by Salmonella typhimurium.

The periplasmic oligopeptide-binding protein. European Journal of Biochemistry

158: 561-567.

Hirsch, P.R. and Beringer, J.E. (1984) A physical map of pPH1JI and pJB4JI.

Plasmid 12: 139-141.

Hirschman, J., Wong, P.-K., Sei, K., Keener, J. and Kustu, S. (1985) Products of

nitrogen regulatory genes ntrA and ntrC of enteric bacteria activate glnA

transcription in vitro: evidence that the ntrA product is a σ factor. Proceedings of

the National Academy of Sciences of the United States of America 82: 7525-7529.

Page 277: The reasonable man attempts to adapt ... - reading.ac.uk

266

Hobson, A.C., Weatherwax, R. and Ames, G.F.-L. (1984) ATP-binding sites in

the membrane components of histidine permease, a periplasmic transport

system. Proceedings of the National Academy of Sciences of the United States of

America 81: 7333-7337.

Holtel, A. and Merrick, M.J. (1988) Identification of the Klebsiella pneumoniae

glnB gene: nucleotide sequence of wild-type and mutant alleles. Molecular and

General Genetics 215: 134-138.

Holtel, A. and Merrick, M.J. (1989) The Klebsiella pneumoniae PII protein (glnB

gene product) is not absolutely required for nitrogen regulation and is not

involved in NifL-mediated nif gene regulation. Molecular and General Genetics

217: 474-480.

Hong, K.H., Jang, W.H., Choi, K.D. and Yoo, O.J. (1991) Characterization of

Psuedomonas fluorescens carboxylesterase: cloning and expression of the esterase

gene in Escherichia coli. Agricultural and Biological Chemistry 55: 2839-2845.

Hornez, J.P., El-Guezzar, M. and Derieux, J.C. (1989) Succinate transport in

Rhizobium meliloti: characteristics and impact on symbiosis. Current Microbiology

19: 207-212.

Hudman, J.F. and Glenn, A.R. (1980) Glucose uptake by free living and

bacteroid forms of Rhizobium leguminosarum. Archives of Microbiology 128: 72-77.

Humbeck, C. and Werner, D. (1989) Delayed nodule development in a

succinate transport mutant of Bradyrhizobium japonicum. Journal of Plant

Physiology 134: 276-283.

Page 278: The reasonable man attempts to adapt ... - reading.ac.uk

267

Hyde, S.C., Emsley, P., Hartshorn, M., Mimmack, M.M., Gileadi, U., Pearce,

S.R., Gallagher, M.P., Gill, D.R., Hubbard, R.E. and Higgins, C.F. (1990)

Structural model of ATP-binding proteins associated with cystic fibrosis,

multidrug resistance and bacterial transport. Nature 346: 362-365.

Indiveri, C., Krämer, R. and Palmieri, F. (1987) Reconstitution of the

malate/aspartate shuttle from mitochondria. Journal of Biological Chemistry 262:

15979-15983.

Irigoyen, J.J., Sanchez-Diaz, M., Emerich, D.W (1990) Carbon metabolism

enzymes of Rhizobium meliloti cultures and bacteroids and their distribution

within alfalfa nodules. Applied and Environmental Microbiology 56: 2587-2589.

Jackson, F.A. and Dawes, E.A. (1976) Regulation of the tricarboxylic acid cycle

and poly-β-hydroxybutyrate metabolism in Azotobacter beijerinckii grown under

nitrogen or oxygen limitation. Journal of General Microbiology 97: 303-312.

Jauniaux, J. and Grenson, M. (1990) Gap1, the general amino acid permease

gene of Saccharomyces cerevisiae. European Journal of Biochemistry 190: 39-44.

Jin, H.N., Glenn, A.R. and Dilworth, M.J. (1990) How does L-glutamate

transport relate to selection of mixed nitrogen sources in Rhizobium

leguminosarum biovar trifolii MNF1000 and cowpea Rhizobium MNF2030.

Archives of Microbiology 153: 448-454.

Johnston, A.W.B. and Beringer, J.E. (1975) Identification of the Rhizobium

strains in pea root nodules using genetic markers. Journal of General

Microbiology 87: 343-350.

Joshi, A.K., Ahmed, S. and Ames, G.F.-L. (1989) Energy coupling to bacterial

periplasmic transport system: studies in intact Escherichia coli cells. Journal of

Biological Chemistry 264: 2126-2133.

Page 279: The reasonable man attempts to adapt ... - reading.ac.uk

268

Kahn, M.L., Kraus, J. and Sommerville, J.E. (1985) A model of nutrient

exchange in the Rhizobium-legume symbiosis. In Nitrogen Fixation Research

Progress. Evans, H.J., Bottomley, P.J. and Newton, W.E. (eds). Dordrecht:

Martinus Nijhoff, pp. 193-199.

Kalman, M., Gentry, D.R. and Cashel, M. (1991) Characterization of the

Escherichia coli K12 gltS glutamate permease gene. Molecular & General Genetics

225: 379-386.

Karr, D.B., Waters, J.K., Suzuki, F. and Emerich, D.W. (1984) Enzymes of the

poly-β-hydroxybutyrate and citric acid cycles of Rhizobium japonicum bacteroids.

Plant Physiology 75: 1158-1162.

Kartner, N., Hanrahan, J.W., Jensen, T.J., Naismith, A.L., Sun, S., Ackerley, C.A.,

Reyes, E.F., Tsui, L.-C., Rommens, J.M., Bear, C.E. and Riordan, J.R. (1991)

Expression of the cystic fibrosis gene in non-epithelial invertebrate cells

produces a regulated anion conductance. Cell 64: 681-692.

Kay, W.W. (1971) Two aspartate transport systems in Escherichia coli. Journal of

Biological Chemistry 246: 7373-7382.

Keen, N.T., Tamaki, S., Kobayashi, D. and Trollinger, D. (1988) Improved

broad-host-range plasmids for DNA cloning in Gram-negative bacteria. Gene

70: 191-197.

Kerem, B.-S., Zielenski, J., Markiewicz, D., Bozon, D., Gazit, E., Yahav, J.,

Kennedy, D., Riordan, J.R., Collins, F.S., Rommens, J.M. and Tsui, L.-C. (1990)

Identification of mutations in regions corresponding to the two putative

nucleotide (ATP)-binding folds of the cystic fibrosis gene. Proceedings of the

National Academy of Sciences of the United States of America 87: 8447-8451.

Page 280: The reasonable man attempts to adapt ... - reading.ac.uk

269

Kerppola, R.E., Shyamala, U.K., Klebba, P. and Ames, G.F.-L. (1991) The

membrane-bound proteins of periplasmic permeases form a complex:

identification of the histidine permease HisQMP complex. Journal of Biological

Chemistry 266: 9857-9865.

Kerppola, R.E. and Ames, G.F.-L. (1992) Topology of the hydrophobic

membrane-bound components of the histidine periplasmic permease.

Comparison with other members of the family. Journal of Biological Chemistry

267: 2329-2336.

Kimura, I. and Tajima, S. (1989) Presence and characteristics of NADP-malic

enzyme in soybean nodule bacteroids. Soil Science and Plant Nutrition 35: 271-

279.

Klucas, R.V. (1974) Studies on soybean nodule senescence. Plant Physiology 54:

612-616.

Kondorosi, A., Svab, Z., Kiss, G.B. and Dixon, R.A. (1977) Ammonia

assimilation and nitrogen fixation in Rhizobium meliloti. Molecular and General

Genetics 151: 221-226.

Kornfeld, S., Benziman, M. and Milner, Y. (1977) α-Ketoglutarate

dehydrogenase complex of Acetobacter xylinum. Journal of Biological Chemistry

252: 2940-2947.

Kouchi, H., Fukai, K., Katagiri, H., Minamisawa, K. and Tajima, S. (1988)

Isolation and enzymological characterization of infected and uninfected

protoplasts from root nodules of Glycine max. Physiologia Plantarum 73: 327-334.

Kouchi, H., Fukai, K. and Kihara, A. (1991) Metabolism of glutamate and

aspartate in bacteroids isolated from soybean root nodules. Journal of General

Microbiology 137: 2901-2910.

Page 281: The reasonable man attempts to adapt ... - reading.ac.uk

270

Kouchi, H. and Fukai, K. (1988) Uptake and metabolism of aspartate and

glutamate by soybean nodule bacteroids. In Nitrogen Fixation: Hundred Years

After. Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds). New York: Gustav

Fischer, pp. 561

Kouchi, H. and Nakaji, K. (1985) Utilization and metabolism of

photoassimilated 13C in soybean roots and nodules. Soil Science and Plant

Nutrition 31: 323-334.

Kouchi, H. and Yoneyama, T. (1986) Metabolism of [13C]-labelled

photosynthate in plant cytosol and bacteroids of root nodules of Glycine max.

Physiologia Plantarum 68: 238-244.

Kramer, R. (1994) Secretion of amino acids by bacteria - physiology and

mechanism. FEMS Microbiology Reviews 13: 75-93.

Kretovich, V.L., Karyakina, T.I., Sidelnikova, L.I., Shaposhnikov, G.L. and

Kaloshina, G.S. (1986) Nitrogen fixation and biosynthesis of aspartic acid and

alanine by bacteroids of Rhizobium lupini on various carbon sources. Doklady

Akademii Nauk SSSR 291: 1008-1011.

Kurz, W.G.W. and LaRue, T.A. (1977) Citric acid cycle enzymes and

nitrogenase in nodules of Pisum sativa. Canadian Journal of Microbiology 23:

1197-1200.

Kustu, S.G., McFarland, N.C., Hui, S.P., Esmon, B. and Ames, G.F.-L. (1979)

Nitrogen control in Salmonella typhimurium: Co-regulation of synthesis of

glutamine synthetase and amino acid transport systems. Journal of Bacteriology

138: 218-234.

Page 282: The reasonable man attempts to adapt ... - reading.ac.uk

271

Kustu, S.G. and Ames, G.F.-L. (1974) The histidine-binding protein J, a

histidine transport component, has two different functional sites. Journal of

Biological Chemistry 249: 6976-6983.

Kyte, J. and Doolittle, R.F. (1982) A simple method for displaying the

hydropathic character of a protein. Journal of Molecular Biology 157: 105-132.

Labes, M., Puhler, A. and Simon, R. (1990) A new family of RSF1010-derived

expression and lac-fusion broad-host-range vectors for Gram-negative bacteria.

Gene 89: 37-46.

Laemmli, U.K. (1970) Cleavage of structural proteins during the assembly of

the head of bacteriophage T4. Nature 227: 680-685.

Lafontaine, P.J., LaFrenière, C. and Antoun, H. (1989) Some properties of

carbohydrate and C4-dicarboxylic acid utilization negative mutants of

Rhizobium leguminosarum biovar phaseoli strain P121. Plant Science 120: 195-201.

Landick, R., Oxender, D.L. and Ames, G.F.-L. (1985) Bacterial amino acid

transport systems. In The Enzymes of Biological Membranes, Volume 3. Martonosi,

A.N. (ed). New York: Plenum Press, pp. 577-615.

Langley, D. and Guest, J.R. (1978) Biochemical genetics of the α-keto acid

dehydrogenase complexes of Escherichia coli K12: genetic characterization and

regulatory properties of deletion mutants. Journal of General Microbiology 106:

103-117.

LaNoue, K.F. and Tischler, M.E. (1974) Electrogenic characteristics of the

mitochondrial glutamate-aspartate antiporter. Journal of Biological Chemistry

249: 7522-7528.

Page 283: The reasonable man attempts to adapt ... - reading.ac.uk

272

Le Rudulier, D., Gloux, K. and Riou, N. (1991) Identification of an osmotically

induced periplasmic glycine betaine-binding protein from Rhizobium meliloti.

Biochimica et Biophysica Acta 1061: 197-205.

Lewis, T.A., Gonzalez, R. and Botsford, J.L. (1990) Rhizobium meliloti glutamate

synthase: cloning and initial characterization of the glt locus. Journal of

Bacteriology 172: 2413-2420.

Libbenga, K.R. and Harkes, P.A.A. (1973) Initial proliferation of cortical cells in

the formation of root nodules in Pisum sativum. Planta 114: 17-28.

Lowry, O.H., Rosebrough, N.J., Farr, A.L. and Randall, R.J. (1951) Protein

measurement with the Folin phenol reagent. Journal of Biological Chemistry 193:

265-275.

MacFarlane, S.A. and Merrick, M. (1985) The nucleotide sequence of the

nitrogen regulation gene (ntrB) and the glnA-ntrBC intergenic region of

Klebsiella pneumoniae. Nucleic Acids Research 13: 7591-7606.

MacFarlane, S.A. and Merrick, M.J. (1987) Analysis of the Klebsiella pneumoniae

ntrB gene by site-directed in vitro mutagenesis. Molecular Microbiology 1: 133-

142.

Mao, B., Pear, M.R., McCammon, J.A. and Quiocho, F.A. (1982) Hinge bending

in L-arabinose-binding protein: the Venus's flytrap model. Journal of Biological

Chemistry 257: 1131-1133.

Martin, G.B., Chapman, K.A. and Chelm, B.K. (1988) Role of the

Bradyrhizobium japonicum ntrC gene product in differential regulation of the

glutamine synthetase II gene (glnII). Journal of Bacteriology 170: 5452-5459.

Page 284: The reasonable man attempts to adapt ... - reading.ac.uk

273

Masters, P.S. and Hong, J.-S. (1981) Genetics of the glutamine transport system

in Escherichia coli. Journal of Bacteriology 147: 805-819.

May, G., Faatz, E., Villarejo, M. and Bremer, E. (1986) Binding protein

dependent transport of glycine betaine and its osmotic regulation in Escherichia

coli K12. Molecular and General Genetics 205: 225-233.

McDermott, T.R., Griffith, S.M., Vance, C.P. and Graham, P.H. (1989) Carbon

metabolism in Bradyrhizobium japonicum bacteroids. FEMS Microbiology Letters

63: 327-340.

McDermott, T.R. and Kahn, M.L. (1992) Cloning and mutagenesis of the

Rhizobium meliloti isocitrate dehydrogenase gene. Journal of Bacteriology 174:

4790-4797.

McKay, I.A., Dilworth, M.J. and Glenn, A.R. (1988) C4-dicarboxylate

metabolism in free-living and bacteroid forms of Rhizobium leguminosarum

MNF3841. Journal of General Microbiology 134: 1433-1440.

McKay, I.A., Dilworth, M.J. and Glenn, A.R. (1989) Carbon catabolism in

continuous cultures and bacteroids of Rhizobium leguminosarum MNF3841.

Archives of Microbiology 152: 606-610.

McRae, D.G., Joy, K. and Miller, R.W. (1989a) Amino acid metabolism in

Rhizobium meliloti bacteroids isolated from alfalfa nodules. Plant Physiology

Supplement 89: 180

McRae, D.G., Miller, R.W., Berndt, W.B. and Joy, K. (1989b) Transport of C4-

dicarboxylates and amino acids by Rhizobium meliloti bacteroids. Molecular

Plant-Microbe Interactions 2: 273-278.

Page 285: The reasonable man attempts to adapt ... - reading.ac.uk

274

Meijer, A.J. and Van Dam, K. (1974) The metabolic significance of anion

transport in mitochondria. Biochimica et Biophysica Acta 346: 213-244.

Meissner, P.S., Sisk, W.P. and Berman, M.L. (1987) Bacteriophage λ cloning

system for the construction of directional cDNA libraries. Proceedings of the

National Academy of Sciences of the United States of America 84: 4171-4175.

Merrick, M.J. (1988) Regulation of nitrogen assimilation by bacteria. In The

Nitrogen and Sulphur Cycles. Cole, J.A. and Ferguson, S. (eds). Cambridge:

Cambridge University Press, pp. 331-361.

Miles, J.S. and Guest, J.R. (1987) Molecular genetic aspects of the citric acid

cycle of Escherichia coli. Biochemical Society Symposium 54: 45-65.

Miller, D.M., Olson, J.S., Pflugrath, J.W. and Quiocho, F.A. (1983) Rates of

ligand binding to periplasmic proteins involved in bacterial transport and

chemotaxis. Journal of Biological Chemistry 258: 13665-13672.

Miller, J.H. (1972) Experiments in molecular genetics. New York: Cold Spring

Harbor.

Miller, K. and Maloy, S. (1990) DNA sequence of the putP gene from Salmonella

typhimurium and predicted structure of proline permease. Nucleic Acids Research

18: 3057

Miller, R.W., McRae, D.G., Al-Jobore, A. and Berndt, W.B. (1988) Respiration

supported nitrogenase activity of isolated Rhizobium meliloti bacteroids. Journal

of Cellular Biochemistry 38: 35-49.

Miller, R.W., McRae, D.G. and Joy, K. (1991) Glutamate and γ-aminobutyrate

metabolism in isolated Rhizobium meliloti bacteroids. Molecular Plant-Microbe

Interactions 4: 37-45.

Page 286: The reasonable man attempts to adapt ... - reading.ac.uk

275

Mimmack, M.L., Gallagher, M.P., Hyde, S.C., Pearce, S.R., Booth, I.R. and

Higgins, C.F. (1989) Energy-coupling to periplasmic binding protein-

dependent transport systems: Stoichiometry of ATP hydrolysis during

transport. Proceedings of the National Academy of Sciences of the United States of

America 86: 8257-8261.

Mimura, C.S., Admon, A., Hurt, K.A. and Ames, G.F.-L. (1990) The nucleotide-

binding site of HisP, a membrane protein of the histidine permease.

Identification of amino acid residues photoaffinity labelled by 8-azido ATP.

Journal of Biological Chemistry 265: 19535-19542.

Mimura, C.S., Holbrook, S.R. and Ames, G.F.-L. (1991) Structural model of the

nucleotide binding conserved component of periplasmic permeases.

Proceedings of the National Academy of Sciences of the United States of America 88:

84-88.

Murphy, E., Coll, K.E., Viale, R.O., Tischler, M.E. and Williamson, J.R. (1979)

Kinetics and regulation of the glutamate-aspartate translocator in rat liver

mitochondria. Journal of Biological Chemistry 254: 8369-8376.

Nelson, S.O. and Postma, P.W. (1984) Interactions in vivo between IIIGlc of the

phosphoenolpyruvate: sugar phosphotransferase system and the glycerol and

maltose uptake systems of Salmonella typhimurium. European Journal of

Biochemistry 139: 29-34.

Newcomb, W. (1981) Nodule morphogenesis. In International Review of

Cytology, Supplement 13. Bourne, G.H. and Danielli, J.F. (eds). New York:

Academic Press, pp. 246-298.

Nicholls, D.J., Sundaram, T.K., Atkinson, T. and Minton, N.P. (1990) Cloning

and nucleotide sequences of the mdh and sucD genes from Thermus aquaticus B.

FEMS Microbiology Letters 70: 7-14.

Page 287: The reasonable man attempts to adapt ... - reading.ac.uk

276

Ninfa, A.J. and Magasanik, B. (1986) Covalent modification of the glnG

product NRI by the glnL product NRII regulates transcription of the glnAG

operon in Escherichia coli. Proceedings of the National Academy of Sciences of the

United States of America 83: 5909-5913.

Nishiyama, M., Horinouchi, S. and Beppu, T. (1991) Characterization of an

operon encoding succinyl-CoA synthetase and malate dehydrogenase from

Thermus flavus AT-62 and its expression in Escherichia coli. Molecular and General

Genetics 226: 1-9.

Nobile, S. and Deshusses, J. (1988) Detection of ligand-protein binding by

direct electrophoresis of the complex. Journal of Chromatography 449: 331-336.

Noel, D., Nikaido, K. and Ames, G.F.-L. (1979) A single amino acid

substitution in a histidine-transport protein drastically alters its mobility in

sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Biochemistry 18:

4159-4165.

Nohno, T., Saito, T. and Hong, J. (1986) Cloning and complete nucleotide

sequence of the Escherichia coli glutamine permease operon (glnHPQ). Molecular

and General Genetics 205: 260-269.

O'Gara, F., Manian, S. and Meade, J. (1984) Isolation of an Asm- mutant of

Rhizobium japonicum defective in symbiotic N2-fixation. FEMS Microbiology

Letters 24: 241-245.

O'Gara, F. and Shanmugam, K.T. (1976) Control of symbiotic nitrogen fixation

in Rhizobia regulation: Export of fixed N2 as NH4+. Biochimica et Biophysica Acta

437: 313-321.

Page 288: The reasonable man attempts to adapt ... - reading.ac.uk

277

O'Hara, G.W., Riley, I.T., Glenn, A.R. and Dilworth, M.J. (1985) The

ammonium permease of Rhizobium leguminosarum MNF3841. Journal of General

Microbiology 131: 757-764.

Ogilvie-Villa, S., DeBusk, R.M. and DeBusk, A.G. (1981) Characterization of 2-

aminoisobutyric acid transport in Neurospora cassa: A general amino acid

permease-specific substrate. Journal of Bacteriology 147: 944-948.

Oh, B.-H., Pandit, J., Kang, C.-H., Nikaido, K., Gokcen, S., Ames, G.F.-L. and

Kim, S.-H. (1993) Three-dimensional structures of the periplasmic

lysine/arginine/ornithine-binding protein with and without a ligand. Journal of

Biological Chemistry 268: 11348-11355.

Ohyama, T. and Kumazawa, K. (1980) Nitrogen assimilation in soybean

nodules. II. 15N2 assimilation in bacteroid and cytosol fractions of soybean

nodules. Soil Science and Plant Nutrition 26: 205-213.

Osburne, M.S. and Signer, E.R. (1980) Ammonium assimilation in Rhizobium

meliloti. Journal of Bacteriology 143: 1234-1240.

Ouyang, L.J. and Day, D.A. (1992) Transport properties of symbiosomes

isolated from siratro nodules. Plant Physiology and Biochemistry 30: 613-623.

Overduin, P., Boos, W. and Tommassen, J. (1988) Nucleotide sequence of the

ugp genes of Escherichia coli K-12: homology to the maltose system. Molecular

Microbiology 2: 767-775.

Oxender, D.L. (1972) Membrane Transport. Annual Review of Biochemistry 41:

777-814.

Pall, M.L. (1969) Amino acid transport in Neurospora crassa. I. Properties of

two amino acid transport systems. Biochimica et Biophysica Acta 173: 113-129.

Page 289: The reasonable man attempts to adapt ... - reading.ac.uk

278

Patriarca, E.J., Chiurazzi, M., Manco, G., Riccio, A., Lamberti, A., Depaolis, A.,

Rossi, M., Defez, R. and Iaccarino, M. (1992) Activation of the Rhizobium

leguminosarum glnII gene by NtrC is dependent on upstream DNA sequences.

Molecular and General Genetics 234: 337-345.

Patriarca, E.J., Riccio, A., Tate, R., Colonna-Romano, S., Iaccarino, M. and Defez,

R. (1993) The ntrBC genes of Rhizobium leguminosarum are part of a complex

operon subject to negative regulation. Molecular Microbiology 9: 569-577.

Payne, G., Spudich, E.N. and Ames, G.F.-L. (1985) A mutational hot-spot in the

hisM gene of the histidine transport operon in Salmonella typhimurium is due to

deletion of repeated sequences and results in an altered specificity of transport.

Molecular and General Genetics 200: 493-496.

Pearce, S.R., Mimmack, M.L., Gallagher, M.P., Gileadi, U., Hyde, S.C. and

Higgins, C.F. (1992) Membrane topology of the integral membrane

components, OppB and OppC, of the oligopeptide permease of Salmonella

typhimurium. Molecular Microbiology 6: 47-57.

Perego, M., Higgins, C.F., Pearce, S.R., Gallagher, M.P. and Hoch, J.A. (1991)

The oligopeptide transport system of Bacillus subtilis plays a role in the initiation

of sporulation. Molecular Microbiology 5: 173-185.

Petronilli, V. and Ames, G.F.-L. (1991) Binding protein-independent histidine

permease mutants: uncoupling of ATP hydrolysis from transmembrane

signaling. Journal of Biological Chemistry 266: 16293-16296.

Pflugrath, J.W. and Quiocho, F.A. (1988) The 2A resolution structure of the

sulfate-binding protein involved in active transport in Salmonella typhimurium.

Journal of Molecular Biology 200: 163-180.

Page 290: The reasonable man attempts to adapt ... - reading.ac.uk

279

Platt, T. and Bear, D.G. (1983) Role of RNA polymerase, rho factor, and

ribosomes in transcription termination. In Gene function in prokaryotes.

Beckwith, J., Davies, J. and Gallant, J.A. (eds). New York: Cold Spring Harbor,

pp. 123-161.

Poole, P.S., Franklin, M., Glenn, A.R. and Dilworth, M.J. (1985) The transport

of L-glutamate by Rhizobium leguminosarum involves a common amino acid

carrier. Journal of General Microbiology 131: 1441-1448.

Poole, P.S., Dilworth, M.J. and Glenn, A.R... (1987) Ammonia is the preferred

nitrogen source in several rhizobia. Journal of General Microbiology 133: 1707-

1712.

Poole, P.S., Blyth, A., Reid, C.J. and Walters, K. (1994a) Myo-inositol catabolism

and catabolite regulation in Rhizobium leguminosarum bv viciae. Microbiology 140:

2787-2795.

Poole, P.S., Schofield, N.A., Reid, C.J., Drew, E.M. and Walshaw, D.L. (1994b)

Identification of chromosomal genes located downstream of dctD that affect the

requirement for calcium and the lipopolysaccharide layer of Rhizobium

leguminosarum. Microbiology 140: 2797-2809.

Poolman, B. and Konings, W.N. (1993) Secondary solute transport in bacteria.

Biochimica et Biophysica Acta 1183: 5-39.

Povolo, S., Tombolini, R., Morea, A., Anderson, A.J., Casella, S. and Nuti, M.P.

(1994) Isolation and characterization of mutants of Rhizobium meliloti unable to

synthesize poly-beta-hydroxybutyrate. Canadian Journal of Microbiology 40: 823-

829.

Page 291: The reasonable man attempts to adapt ... - reading.ac.uk

280

Prossnitz, E., Nikaido, K., Ulbrich, S. and Ames, G.F.-L. (1988) Formaldehyde

and photoactivatable cross-linking of the periplasmic binding protein to a

membrane component of the histidine transport system of Salmonella

typhimurium. Journal of Biological Chemistry 234: 17917-17920.

Prossnitz, E., Gee, A. and Ames, G.F.-L. (1989) Reconstitution of the histidine

periplasmic transport system in membrane vesicles. Energy coupling and

interaction between the binding protein and the membrane complex. Journal of

Biological Chemistry 264: 5006-5014.

Putnoky, P., Kiss, G.B., Ott, I. and Kondorosi, A. (1983) Tn5 carries a

streptomycin resistance determinant downstream from the kanamycin

resistance gene. Molecular and General Genetics 191: 288-294.

Quandt, J. and Hynes, M.F. (1993) Versatile suicide vectors which allow direct

selection for gene replacement in Gram-negative bacteria. Gene 127: 15-21.

Quiocho, F.A. (1986) Carbohydrate-binding proteins: tertiary stuctures and

protein-sugar interactions. Annual Review of Biochemistry 55: 287-316.

Quiocho, F.A. (1990) Atomic structures of periplasmic binding proteins and

the high-affinity active transport systems in bacteria. Philosophical Transactions

of the Royal Society of London Series B-Bioligical Sciences 326: 341-351.

Rao, E.Y.T., Rao, T.K. and DeBusk, A.G. (1975) Isolation and characterization

of a mutant of Neurospora crassa deficient in general amino acid permease

activity. Biochimica et Biophysica Acta 413: 45-51.

Rastogi, V.K. and Watson, R.J. (1991) Aspartate aminotransferase activity is

required for aspartate catabolism and symbiotic nitrogen fixation in Rhizobium

meliloti. Journal of Bacteriology 173: 2879-2887.

Page 292: The reasonable man attempts to adapt ... - reading.ac.uk

281

Reeves, H.C., Rabin, R., Wegener, W.S. and Ajl, S.J. (1971) Assays of enzymes

of the tricarboxylic acid and glyoxylate cycles. Methods in Microbiology 6A: 425-

462.

Reibach, P.H. and Streeter, J.G. (1983) Metabolism of 14C-labeled

photosynthate and distribution of enzymes of glucose metabolism in soybean

nodules. Plant Physiology 72: 634-640.

Reibach, P.H. and Streeter, J.G. (1984) Evaluation of active versus passive

uptake of metabolites by Rhizobium japonicum bacteroids. Journal of Bacteriology

159: 47-52.

Reid CJ (1995) The regulation of the DCT system in Rhizobium leguminosarum

biovar viciae. PhD Thesis. University of Reading.

Reitzer, L.J. and Magasanik, B. (1983) Isolation of the nitrogen assimilation

regulator, NRI, the product of the glnG gene of Escherichia coli. Proceedings of the

National Academy of Sciences of the United States of America 80: 5554-5558.

Reitzer, L.J. and Magasanik, B. (1985) Expression of glnA in Escherichia coli is

regulated at tandem promoters. Proceedings of the National Academy of Sciences of

the United States of America 82: 1979-1983.

Reitzer, L.J. and Magasanik, B. (1987) Ammonia assimilation and the

biosynthesis of glutamine, glutamate, aspartate, L-alanine and D-alanine. In

Escherichia coli and Salmonella typhimurium cellular and molecular biology.

Neidhart, F.C. (ed). Washington DC: American Society for Microbiology, pp.

302-318.

Reitzer, L.J. and Magasanik, B. (1986) Transcription of glnA in E. coli is

stimulated by activator bound to sites far from the promoter. Cell 45: 785-792.

Page 293: The reasonable man attempts to adapt ... - reading.ac.uk

282

Reyes, M., Treptow, N.A. and Shuman, H.A. (1986) Transport of p-

nitrophenyl-α-maltoside by the maltose transport system of Escherichia coli and

its subsequent hydrolysis by a cytoplasmic α-maltosidase. Journal of Bacteriology

165: 918-922.

Reyes, M. and Shuman, H.A. (1988) Overproduction of the MalK protein

prevents expression of the Escherichia coli mal operon. Journal of Bacteriology 170:

4598-4602.

Reynolds, P.H.S., Boland, M.J. and Farnden, K.J.F. (1981) Enzymes of nitrogen

metabolism in legume nodules: Partial purification and properties of the

aspartate aminotransferases from lupine nodules. Archives of Biochemistry and

Biophysics 209: 524-533.

Richarme, G. and Kepes, A. (1983) Study of binding protein-ligand interactions

by ammonium sulphate-assisted adsorption on cellulose esters filters.

Biochimica et Biophysica Acta 742: 16-24.

Robertson, J.G., Lyttleton, P., Bullivant, S. and Grayston, G.F. (1978)

Membranes in lupin root nodules. I. The role of Golgi bodies in the biogenesis

of infection threads and peribacteroid membranes. Journal of Cell Science 30:

129-149.

Robertson, J.G. and Taylor, M.P. (1973) Acid and alkaline invertases in roots

and nodules of Lupinus angustifolius infected with Rhizobium lupini. Planta 112:

1-6.

Robson, R., Woodley, P. and Jones, R. (1986) Second gene (nifH*) coding for a

nitrogenase iron protein in Azotobacter chroococcum is adjacent to a gene coding

for a ferredoxin-like protein. EMBO Journal 5: 1159-1163.

Page 294: The reasonable man attempts to adapt ... - reading.ac.uk

283

Ronson, C.W., Lyttleton, P. and Robertson, J.G. (1981) C4-dicarboxylate

transport mutants of Rhizobium trifolii form ineffective nodules on Trifolium

repens. Proceedings of the National Academy of Sciences of the United States of

America 78: 4284-4288.

Ronson, C.W. and Primrose, S.B. (1979) Carbohydrate metabolism in Rhizobium

trifolii: identification and symbiotic properties of mutants. Journal of General

Microbiology 112: 77-88.

Rosenberg, M. and Court, D. (1979) Regulatory sequences involved in the

promotion and termination of RNA transcription. Annual Review of Genetics 13:

319-353.

Rosendahl, L., Dilworth, M.J. and Glenn, A.R. (1992) Exchange of metabolites

across the peribacteroid membrane in pea root nodules. Journal of Plant

Physiology 139: 635-638.

Rossi, M., Defez, R., Chiurazzi, M., Lamberti, A., Fuggi, A. and Iaccarino, M...

(1989) Regulation of glutamine synthetase isoenzymes in Rhizobium

leguminosarum biovar viciae. Journal of General Microbiology 135: 629-637.

Rudner, D.Z., LeDeaux, J.R., Ireton, K. and Grossman, A.D. (1991) The spo0K

locus of Bacillus subtilis is homologous to the oligopeptide permease locus and

is required for sporulation and competence. Journal of Bacteriology 173: 1388-

1398.

Ruvkun, G.B. and Ausubel, F.M. (1981) A general method for site-directed

mutagenesis in prokaryotes. Nature 289: 85-88.

Ryan, E., Bodley, F. and Fottrell, P.F. (1972) Purification and characterization of

aspartate aminotransferases from soybean root nodules and Rhizobium

japonicum. Phytochemistry 11: 957-963.

Page 295: The reasonable man attempts to adapt ... - reading.ac.uk

284

Ryan, E. and Fottrell, P.F. (1974) Subcellular localization of enzymes involved

in the assimilation of ammonia by soybean root nodules. Phytochemistry 13:

2647-2652.

Sack, J.S., Saper, M.A. and Quiocho, F.A. (1989) Periplasmic binding protein

structure and function: refined X-ray structures of the

leucine/isoleucine/valine binding protein and its complex with leucine. Journal

of Molecular Biology 206: 171-191.

Salminen, S.O. and Streeter, J.G. (1987a) Uptake and metabolism of

carbohydrates by Bradyrhizobium japonicum bacteroids. Plant Physiology 83: 535-

540.

Salminen, S.O. and Streeter, J.G. (1987b) Involvement of glutamate in the

respiratory metabolism of Bradyrhizobium japonicum bacteroids. Journal of

Bacteriology 169: 495-499.

Salminen, S.O. and Streeter, J.G. (1990) Factors contributing to the

accumulation of glutamate in Bradyrhizobium japonicum bacteroids under

microaerobic conditions. Journal of General Microbiology 136: 2119-2126.

Salminen, S.O. and Streeter, J.G. (1991) Improved analysis of metabolite uptake

by Bradyrhizobium japonicum bacteroids. Canadian Journal of Microbiology 37:

238-243.

Salminen, S.O. and Streeter, J.G. (1992) Labeling of carbon pools in

Bradyrhizobium japonicum and Rhizobium leguminosarum bv. viciae bacteroids

following incubation of intact nodules with 14CO2. Plant Physiology 100: 597-

604.

Sambrook, J., Fritsch, E.F. and Maniatis, T. (1989) Molecular Cloning: A

Laboratory Manual. New York: Cold Spring Harbor Laboratory Press.

Page 296: The reasonable man attempts to adapt ... - reading.ac.uk

285

San-Francisco, M.J.D. and Jacobson, G.R. (1986) Glucose uptake and

phosphorylating activities in two species of slow-growing Rhizobium. FEMS

Microbiology Letters 35: 71-74.

Saroso, S., Glenn, A.R. and Dilworth, M.J. (1984) Carbon utilization by free-

living and bacteroid forms of cowpea Rhizobium strain NGR234. Journal of

General Microbiology 130: 1809-1814.

Saroso, S., Dilworth, M.J. and Glenn, A.R. (1986) The use of activities of carbon

catabolic enzymes as a probe for the carbon nutrition of snakebean nodule

bacteroids. Journal of General Microbiology 132: 243-249.

Saurin, W., Köster, W. and Dassa, E. (1994) Bacterial binding protein-

dependent permeases: characterization of distinctive signatures for functionally

related integral cytoplasmic membrane proteins. Molecular Microbiology 12:

993-1004.

Schellenberg, G.D. and Furlong, C.E. (1977) Resolution of the multiplicity of

the glutamate and aspartate transport systems of Escherichia coli. Journal of

Biological Chemistry 252: 9055-9064.

Schmitz, G., Nikaido, K. and Ames, G.F.-L. (1988) Regulation of a transport

operon promoter in Salmonella typhimurium: Identification of sites essential for

nitrogen regulation. Molecular and General Genetics 215: 107-117.

Schneider, E. and Walter, C. (1991) A chimeric nucleotide-binding protein,

encoded by a hisP-malK hybrid gene is functional in maltose transport in

Salmonella typhimurium. Molecular Microbiology 5: 1375-1383.

Page 297: The reasonable man attempts to adapt ... - reading.ac.uk

286

Schroeder, D.D., Allison, A.J. and Buchanan, J.M. (1969) Biosynthesis of the

purines. XXXII. Effect of albizziin and other reagents on the activity of

formylglycinamide ribonucleotide amidotransferase. Journal of Biological

Chemistry 244: 5856-5865.

Senior, P.J. (1971) Poly-β-hydroxybutyrate biosynthesis and the regulation of

glucose metabolism in Azotobacter beijerinckii. The Biochemical Journal 125: 55-66.

Senior, P.J., Beech, G.A., Ritchie, G.A.F. and Dawes, E.A. (1972) The role of

oxygen limitation in the formation of poly-β-hydroxybutyrate during batch and

continuous culture of Azotobacter beijerinckii. Biochemical Journal 128: 1193-1201.

Shatters, R.G., Somerville, J.E. and Kahn, M.L... (1989) Regulation of glutamine

synthetase II activity in Rhizobium meliloti 104A14. Journal of Bacteriology 171:

5087-5094.

Sherwood, M.T. (1970) Improved synthetic medium for the growth of

Rhizobium. Journal of Applied Bacteriology 33: 708-713.

Shuman, H.A. (1982) Active transport of maltose in Escherichia coli K12. Journal

of Biological Chemistry 257: 5455-5461.

Shuman, H.A. and Silhavy, T.J. (1981) Identification of the malK gene product.

Journal of Biological Chemistry 256: 560-562.

Shyamala, V., Baichwal, V., Beall, E. and Ames, G.F.-L. (1991) Structure-

function analysis of the histidine permease and comparison with cystic fibrosis

mutations. Journal of Biological Chemistry 266: 18714-18719.

Silhavy, T.J., Szmelcman, S., Boos, W. and Schwartz, M. (1975) On the

significance of the retention of ligand by protein. Proceedings of the National

Academy of Sciences of the United States of America 72: 2120-2124.

Page 298: The reasonable man attempts to adapt ... - reading.ac.uk

287

Simon, R., Priefer, U. and Pühler, A. (1983) A broad host-range mobilization

system for in vivo genetic engineering: transposon mutagenesis of Gram-

negative bacteria. Biotechnology 1: 784-791.

Simon, R., Quandt, J. and Klipp, W. (1989) New derivatives of transposon Tn5

suitable for mobilization of replicons, generation of operon fusions and

induction of genes in Gram-negative bacteria. Gene 80: 161-169.

Smith, M.W. and Neidhardt, F.C. (1983) 2-Oxoacid dehydrogenase complexes

of Escherichia coli: cellular amounts and patterns of synthesis. Journal of

Bacteriology 156: 81-88.

Somerville, J.E., Shatters, R.G. and Kahn, M.L. (1989) Isolation,

characterization, and complementation of Rhizobium meliloti 104A14 mutants

that lack glutamine synthetase II activity. Journal of Bacteriology 171: 5079-5086.

Spaink, H.P., Okker, R.J.H., Wijffelman, C.A., Pees, E. and Lugtenberg, B.J.J.

(1987) Promoters in the nodulation region of the Rhizobium leguminosarum SYM

plasmid PRL1JI. Plant Molecular Biology 9: 27-39.

Speiser, D.M. and Ames, G.F.-L. (1991) Mutants of the histidine periplasmic

permease of Salmonella typhimurium that allow transport in the absence of

histidine-binding proteins. Journal of Bacteriology 173: 1444-1451.

Stein, W.D. (1990) Channels, Carriers and Pumps. San Diego: Academic.

Stern, M.J., Higgins, C.F. and Ames, G.F.-L. (1984) Isolation and

characterization of lac fusions to two nitrogen-regulated promoters. Molecular

and General Genetics 195: 219-227.

Page 299: The reasonable man attempts to adapt ... - reading.ac.uk

288

Stern, M.J., Prossnitz, E. and Ames, G.F.-L. (1988) Role of the intercistronic

region in post-transcriptional control of gene expression in the histidine

transport operon of Salmonella typhimurium: involvement of REP sequences.

Molecular Microbiology 2: 141-152.

Stock, J.B., Ninfa, A.J. and Stock, A.M. (1989) Protein-phosphorylation and

regulation of adaptive responses in bacteria. Microbiological Reviews 53: 450-

490.

Stovall, I. and Cole, M. (1978) Organic acid metabolism by isolated Rhizobium

japonicum bacteroids. Plant Physiology 61: 787-790.

Streeter, J.G. (1987) Carbohydrate, organic acid, and amino acid composition of

bacteroids and cytosol from soybean nodules. Plant Physiology 85: 768-773.

Streeter, J.G. (1989) Estimation of ammonium concentration in the cytosol of

soybean nodules. Plant Physiology 90: 779-782.

Streeter, J.G. and Salminen, S.O. (1990) Periplasmic metabolism of glutamate

and aspartate by intact Bradyrhizobium japonicum bacteroids. Biochimica et

Biophysica Acta 1035: 257-265.

Stripf, R. and Werner, D. (1978) Differentiation of Rhizobium japonicum, II.

Enzymatic activities in bacteroids and plant cytoplasm during the development

of nodules of Glycine max. Zeitshrift für Naturforshung 33c: 373-381.

Suganuma, N. and Yamamoto, Y. (1987) Carbon metabolism related to

nitrogen fixation in soybean root nodules. Soil Science and Plant Nutrition 33:

79-91.

Page 300: The reasonable man attempts to adapt ... - reading.ac.uk

289

Szeto, W.W., Nixon, B.T., Ronson, C.W. and Ausubel, F.M. (1987)

Identification and characterization of the Rhizobium meliloti ntrC gene - R.

meliloti has separate regulatory pathways for activation of nitrogen fixation

genes in free living and symbiotic cells. Journal of Bacteriology 169: 1423-1432.

Ta, T.C., Macdowall, F.D.H., Faris, M.A. and Joy, K.W. (1988) Metabolism of

nitrogen fixed by root nodules of alfalfa (Medicago sativa L.): I. Utilization of

[14C, 15N]glutamate and [14C, 15N]glutamine in the synthesis of γ-aminobutyrate.

Biochemistry and Cell Biology 66: 1342-1348.

Tajima, K., Kimura, I. and Sasahara, H. (1986) Succinate metabolism of isolated

soybean nodule bacteroids at low oxygen. Agricultural and Biological Chemistry

50: 1009-1014.

Tajima, S., Kouzai, K. and Kimura, I. (1988) NAD(P)/NAD(P)H ratio and

energy charge in succinate degrading soybean nodule bacteroids. In Nitrogen

Fixation: Hundred Years After. Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds).

New York: Gustav Fischer, pp. 564

Tajima, S. and Kouchi, H. (1990) Metabolism of C4-dicarboxylates and amino

acids in soybean nodule bacteroids, and their role in energy supply to

nitrogenase. In Nitrogen Fixation: Achievements and Objectives. Gresshoff, P.M.,

Roth, L.E., Stacey, G. and Newton, W.E. (eds). New York: Chapman and Hall,

pp. 309-314.

Tajima, S. and Kouzai, K. (1989) Nucleotide pools in soybean nodule tissues, a

survey of NAD(P)/NAD(P)H ratios and energy change. Plant and Cell

Physiology 30: 589-593.

Talibart, R., Jebbar, M., Gouesbet, G., Himdikabbab, S., Wroblewski, H., Blanco,

C. and Bernard, T. (1994) Osmoadaptation in rhizobia: ectoine-induced salt

tolerance. Journal of Bacteriology 176: 5210-5217.

Page 301: The reasonable man attempts to adapt ... - reading.ac.uk

290

Tempest, D.W. and Neijssel, O.M. (1992) Physiological and energetic aspects of

bacterial metabolite overproduction. FEMS Microbiology Letters 100: 169-176.

Tinoco, I., Borer, P.N., Dengler, B. and Levine, M.D. (1973) Improved

estimation of secodary structure in ribonucleic acids. Nature New Biology 246:

40-41.

Tolner, B., Poolman, B. and Konings, W.N. (1992a) Characterization and

functional expression in Escherichia coli of the sodium/proton/glutamate

symport proteins of Bacillus stearothermophilus and Bacillus caldotenax. Molecular

Microbiology 6: 2845-2856.

Tolner, B., Poolman, B., Wallace, B. and Konings, W.N. (1992b) Revised

nucleotide sequence of the gltP gene, which encodes the proton-glutamate-

aspartate transport protein of Escherichia coli K-12. Journal of Bacteriology 174:

2391-2393.

Tolner, B., Ubbinkkok, T., Poolman, B. and Konings, W.N. (1995)

Characterization of the proton/glutamate symport protein of Bacillus subtilis

and its functional expression in Escherichia coli. Journal of Bacteriology 177: 2863-

2869.

Treptow, N.A. and Shuman, H.A. (1985) Genetic evidence for substrate and

binding protein recognition by the MalF and MalG proteins, cytoplasmic

membrane components of the Escherichia coli maltose transport system. Journal

of Bacteriology 163: 654-660.

Treptow, N.A. and Shuman, H.A. (1988) Allele-specific malE mutations that

restore interactions between maltose-binding protein and the inner-membrane

components of the maltose transport system. Journal of Molecular Biology 202:

809-822.

Page 302: The reasonable man attempts to adapt ... - reading.ac.uk

291

Trinchant, J.C., Birot, A.M. and Rigaud, J. (1981) Oxygen supply and energy-

yielding substrates for nitrogen fixation (acetylene reduction) by bacteroid

preparations. Journal of General Microbiology 125: 159-165.

Trinick, M.J., Dilworth, M.J. and Grounds, M. (1976) Factors affecting the

reduction of acetylene by root nodules of Lupinus species. New Phytologist 77:

359-370.

Tubb, R.S. (1976) Regulation of nitrogen fixation in Rhizobium sp. Applied and

Environmental Microbiology 32: 483-488.

Udvardi, M.K., Salom, C.L. and Day, D.A. (1988) Transport of L-glutamate

across the bacteroid membrane but not the peribacteroid membrane from

soybean root nodules. Molecular Plant-Microbe Interactions 1: 250-254.

Udvardi, M.K., Ou Yang, L.-J., Young, S. and Day, D.A. (1990) Sugar and

amino acid transport across symbiotic membranes from soybean nodules.

Molecular Plant-Microbe Interactions 3: 334-340.

Udvardi, M.K. and Day, D.A. (1988) Metabolite transport across the

peribacteroid membrane from soybean root nodules. In Nitrogen Fixation:

Hundred Years After. Bothe, H., de Bruijn, F.J. and Newton, W.E. (eds). New

York: Gustav Fischer, pp. 534

Ueno-Nishio, S., Mango, S., Reitzer, L.J. and Magasanik, B. (1984) Identification

and regulation of the glnL promoter of the complex glnALG operon of

Escherichia coli. Journal of Bacteriology 160: 379-384.

Valverde, M.A., Diaz, M., Sepulveda, F.V., Gill, D.R., Hyde, S.C. and Higgins,

C.F. (1992) Volume-regulated chloride channels associated with the human

multidrug-resistance P-glycoprotein. Nature 355: 830-833.

Page 303: The reasonable man attempts to adapt ... - reading.ac.uk

292

van Heeswijk, W.C., Rabenberg, M., Westerhoff, H.V. and Kahn, D. (1993) The

genes of the glutamine synthetase adenylylation cascade are not regulated by

nitrogen in Escherichia coli. Molecular Microbiology 9: 443-457.

von Heijne, G. (1986) The distribution of positively charged residues in

bacterial inner membrane proteins correlates with the trans-membrane

topology. EMBO Journal 5: 3021-3027.

von Heijne, G. (1992) Membrane protein structure prediction. Hydrophobicity

analysis and the positive-inside rule. Journal of Molecular Biology 225: 487-494.

Vyas, N.K., Vyas, M.N. and Quiocho, F.A. (1991) Comparison of the

periplasmic receptors for L-arabinose, D-glucose/D-galactose and D-ribose.

Journal of Biological Chemistry 266: 5226-5237.

Walker, J.E., Saraste, M., Runswick, M.J. and Gay, N.J. (1982) Distantly related

sequences in the α- and β-subunits of ATP synthase, myosin, kinases and other

ATP- requiring enzymes and a common nucleotide binding fold. EMBO

Journal 1: 945-951.

Wallace, B., Yang, Y.-J., Hong, J. and Lum, D. (1990) Cloning and sequencing

of a gene encoding a glutamate and aspartate carrier of Escherichia coli K-12.

Journal of Bacteriology 172: 3214-3220.

Waters, J.K., Karr, D.B. and Emerich, D.W. (1985) Malate dehydrogenase from

Rhizobium japonicum 311b-143 bacteroids and Glycine max root nodule

mitochondria. Biochemistry 24: 6479-6486.

Watson, R.J., Chan, Y.K., Wheatcroft, R., Yang, A.F. and Han, S. (1988)

Rhizobium meliloti genes required for C4-dicarboxylate transport and symbiotic

nitrogen fixation are located on a megaplasmid. Journal of Bacteriology 170: 927-

934.

Page 304: The reasonable man attempts to adapt ... - reading.ac.uk

293

Watson, R.J., Rastogi, V.K. and Chan, Y.K. (1993) Aspartate transport in

Rhizobium meliloti. Journal of General Microbiology 139: 1315-1323.

Weiner, J.H. and Heppel, L.A. (1971) A binding protein for glutamine and its

relation to active transport in Escherichia coli. Journal of Biological Chemistry 246:

6933-6941.

Weitzman, P.D.J. (1972) Regulation of α-ketoglutarate dehydrogenase activity

in Acinetobacter. FEBS Letters 22: 323-326.

Werner, D. and Stripf, R. (1978) Differentiation of Rhizobium japonicum, I.

Enzymatic comparison of nitrogenase repressed and derepressed free-living

cells and of bacteroids. Zeitshrift für Naturforshung 33c: 245-252.

Willetts, N.S. and Mount, D.W. (1969) Genetic analysis of recombination-

deficient mutants of Escherichia coli K-12 carrying rec mutations cotransducible

with thyA. Journal of Bacteriology 100: 923-934.

Williamson, J.R. and Cooper, R.H. (1980) Regulation of the citric acid cycle in

mammalian systems. FEBS Letters 117 (Supplement): K73-K85.

Wong, P.P. and Evans, H.J. (1971) Poly-β-hydroxybutyrate utilization by

soybean (Glycine max Merr.) nodules and assessment of its role in maintenance

of nitrogenase activity. Plant Physiology 47: 750-755.

Wood, W.B. (1966) Host specificity of DNA produced by Escherichia coli;

bacterial mutations affecting the restriction and modification of DNA. Journal of

Molecular Biology 16: 118-133.

Yamato, I., Ohsawa, M. and Anraku, Y. (1990) Defective cation-coupling

mutants of Escherichia coli Na+/proline symport carrier. Journal of Biological

Chemistry 265: 2450-2455.

Page 305: The reasonable man attempts to adapt ... - reading.ac.uk

294

Yao, P.Y. and Vincent, J.M. (1969) Host specificity in the root hair "curling

factor" of Rhizobium sp. Australian Journal of Biological Sciences 22: 413-423.


Recommended