Loyola University Chicago Loyola eCommons

Dissertations Theses and Dissertations

1997

Physiological Role of the Short Form of the Histidine Protein Kinase, Cheas, in Bacterial Chemotaxis

Barry Patrick McNamara Loyola University Chicago

Follow this and additional works at: https://ecommons.luc.edu/luc_diss

Part of the Microbiology Commons

Recommended Citation McNamara, Barry Patrick, "Physiological Role of the Short Form of the Histidine Protein Kinase, Cheas, in Bacterial Chemotaxis" (1997). Dissertations. 3422. https://ecommons.luc.edu/luc_diss/3422

This Dissertation is brought to you for free and open access by the Theses and Dissertations at Loyola eCommons. It has been accepted for inclusion in Dissertations by an authorized administrator of Loyola eCommons. For more information, please contact [email protected].

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 3.0 License. Copyright © 1997 Barry Patrick McNamara LOYOLA UNIVERSITY CHICAGO

PHYSIOLOGICAL ROLE OF THE SHORT FORM OF THE HISTIDINE

PROTEIN KINASE, CheA8 , IN BACTERIAL CHEMOTAXIS

A DISSERTATION SUBMITTED TO

THE FACULTY OF THE GRADUATE SCHOOL

IN CANDIDACY FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

DEPARTMENT OF MICROBIOLOGY AND IMMUNOLOGY

BY

BARRY PATRICK MCNAMARA

CHICAGO, ILLINOIS

MAY 1997 Copyright by Barry P. McNamara, 1997 All rights reserved.

ii ACKNOWLEDGMENTS

The culmination of this work would not have been possible were it not for the encouragement and generous support provided to me by my parents Justina and Eugene McNamara during the course of all my studies. I also wish to recognize my immediate co-workers, with whom I have had the honor and privilege of working with for the past several years. Specifically, I thank them for their help in navigating me through the computer world, for their generosity in sharing materials, and for their spirited guidance and technical support.

Without question, I must express my gratitude to each of my committee members, Drs. Thomas Gallagher, John Lopes, Mike Fasullo, Charles Lange for their critical evaluation of my experimental data at each step of the way. and to

Dr. David Hecht for the leadership he provided during each committee meeting.

I especially thank my director, Dr. Alan J. Wolfe, for his persistence and belief in my abilities in light of the occasional setback.

Finally, I will and do thank God for the special opportunity He has given me in being able to take but a small peek into the wonderment of rii s awesome creation and, especially, for the greatest creation and blessing of all, my new wife Ulrike, of whom I do adore.

iii TABLE OF CONTENTS

ACKNOWLEDGMENTS ...... iii

LIST OF ILLUSTRATIONS ...... vii

LIST OF TABLES ...... ix

Chapter

I. INTRODUCTION ...... 1

II. LITERATURE REVIEW ...... 4

2.1. Introduction ...... 4 2.2. Chemosensing and behavioral responses of motile bacteria . 5 2.3. Bacterial chemoreceptors ...... 7 2.3.1. Structure and ligand interactions of chemoreceptors 9 .2. Transmembrane signaling ...... 11 .3. Cellular distribution of chemoreceptors ...... 12 2.4. Biochemistry of the chemotaxis signaling pathway ...... 13 2. 5. Adaptation responses ...... 17 2.6. Genetics of the chemotaxis signaling pathway ...... 19 2. 7. products of the cheA locus ...... 21

Ill. MATERIALS AND EXPERIMENTAL METHODS ...... 25

3. 1. Chemicals and biological reagents...... 25 3.2. Construction of plasmid-borne cheA mutant al lei es . _ . . . . . 25 3.3. Bacterial strains ...... _ . . . . . 28 3.3.1. Isolation of cheA and truncation mutants . 29 .2. Isolation of chromosomal cheA start(S) muian1s . . 32 3.4. Bacterial media ...... _ . . . . . 35 3.5. Growth conditions ...... _ ..... 36 3.5.1. Measurement of microbial growth ...... _ ..... 36 .2. Proteus mirabilis swarmer cell isolation ... _ ..... 36 3.6. Genetic manipulations ...... _ . . . . . 37 3.6.1. Transformation of competent E. coli ...... _ . . . . . 37 .2. Generalized transduction ...... _ . . . . . 38 .3. Replica plating for mutant selection ...... _ . . . . . 39 .4. Complementation analysis ...... _ . . . . . 39 3.7. Methods to assess motility and chemotactic beha\Jior ..... 40

iv 3. 7 .1. Swim plate analysis ...... 41 .2. Anaerobic swim assay conditions ...... 42 .3. Competitive swim assay analysis ...... 42 .4. Swimming speed and tumble frequency determinations ...... 43 3.8. Protein methods ...... 44 3.8.1. CheA overexpression and purification ...... 44 .2. SOS-PAGE and western blot analysis ...... 45 .3. CheA in vitro autophosphorylation assays ...... 46 .4. Generation of affinity purified anti-E. coJ; CheA polyclonal antibodies ...... 47 3.9. Molecular biology methods ...... _ ..... 48 3.9.1. Plasmid and chromosomal DNA preparation ..... 48 .2. DNA cloning and restriction endonuclease analysis ...... _ . . . . . 49 .3. Site-directed mutagenesis ...... _ . . . . . 50 .4. Polymerase chain reaction ...... _ ..... 51 .5. DNA hybridizations ...... 52 .6. DNA sequence analysis ...... 53

IV. EXPERIMENTAL RESULTS ...... _ . . . . . 55

4.1. In vivo complementation of CheA C-terminal truncation mutants by CheAs ...... _ . . . 55 4.2. CheAs overexpression studies ...... _ ...... 58 4.3. In vitro transducer-mediated transphosphorylation

by CheA5 ...... _ ... 62 4.4. Chemotactic behavior among clinical isolates of Enterobacteriaceae ...... _ . . . . _ . . . 63 4.5. CheA expression in the Enterobacteriaceae ... _ ...... 67 4.6. CheZ expression in the Enterobacteriaceae ... _ .... _ ... 73 4. 7. Analysis of the 5' region of enteric cheA . _ . . . . _ . . . 73 4.8. Behavioral studies with strains of E. coli that differ only in

their ability to synthesize CheA5 ...... •.... _ ....•... 80 4.8.1. Design and construction of cheA start(S~ muianis . 81 .2. CheA expression from cheA start(S) mutants .... 83 4.9. Chemotactic behavior of cheA start(S) mutant strains ..... 85

4.10. CheA5 complementation analysis ...... _ ...... 89 4. 11. Competitive swim assay analysis ...... _ ...... 92 4.12. In vitro phosphorylation of Met-98 CheA._ variants ...... 94 4.13. Relative expression levels of CheA._ and C heAs . _ ...... 96

V. DISCUSSION ...... - . _ ..... 101

v 5.1. Chemotaxis-associated activities exhibited by CheA$. . . . 101

5.2. Coexpression of CheAt_ and CheA5 by members of the Enterobacteriaceae...... 105 5.3. Putative physiological role of CheA5 in chemotaxis ...... 109 5.4. Putative signaling checkpoints affected by CheAs ...... 113

APPENDIX

A List of common abbreviations used in this study ...... 117

WORKS CITED ...... 118

VITA ...... 132

vi LIST OF ILLUSTRATIONS

Figure Page

1. Membrane topology and domain organization of MCP molecules . . . . . 1O

2. Circuitry of the chemotactic signal transduction pathway ...... 16

3. Genomic organization of the mocha and meche operons in E. coli . . . . 20

4. Organization of structural domains and associated functions shared among both forms of CheA ...... 22

5. CheAs complementation studies with strains that express various CheA truncation mutants ...... 56

6. Restoration of swimming ability by CheA8 to cells expressing CheA616K(Am) ...... 57

7. Effect of CheA8 concentration on the swimming ability of cells expressing CheA616K(Am) ...... 60

8. Effect of CheA8 overexpression on swimming ability in a wild-type strain ...... 61

9. Swim colony morphologies exhibited by various clinical isolates of Enterobacteriaceae ...... 65

1O. CheA expression among motile Enterobacteriaceae ...... 70

11. CheA expression in isolates of Klebsiella ssp. and Yersin;a s sp...... 72

12. CheZ expression in motile and non-motile enteric bacteria ...... 74

13. Comparison of nucleotide sequences spanning the '3' motB-5' cfleA region of the mocha operon ...... 77

vii 14. Predicted amino acid sequence alignments of the amino-terminal portion of CheA ...... 79

15. Schematic of the mocha operon of E. coli and mutant cheA start(S) alleles used in this study ...... 82

16. lmmunoblot analysis on recombinant strains of E. coli harboring cheA start(S) alleles ...... 84

17. Swimming ability of cells that carry indicated cheA start(S) mutations under aerobic (A) or anaerobic (B) assay conditions ...... 86

18. Influence oh-serine upon the chemotactic ability of cells wild-1ype for chemotaxis when grown under anaerobic conditions ...... 88

19. CheAs induction as a function of IPTG inducer concentration ...... 90

20. Effect of CheAs concentration on the swimming ability of cells harboring cheA start(S) mutations ...... 91

21. Growth characteristics of M98L mutant strains grown in tryptone broth at 31EC ...... 95

22. Purified CheA variant proteins isolated from strain BL21(DE3) transformants harboring CheA expression vectors ...... 97

23. Relative in vitro phosphorylation activities of HisCTag purified CheAL variants ...... 98

24. In vivo steady state levels of CheAt_ and CheAs as a func1ion oi the growth phase of E. coli grown in tryptone broth ...... 100

viii LIST OF TABLES

Table Page

1. Components of the chemotaxis signal transduction system ...... 19

2. Plasmids used in complementation studies ...... 26

3. Plasmids used in strain construction ...... 28

4. Bacterial strains used in this study ...... 33

5. Plasmids used for CheA overexpression and purification ...... 45

6. Oligonucleotide primers used in this study ...... 52

7. Primary clinical isolates of Enterobacteriaceae used in this study .... 64

8. Competitive swim assay analysis ...... 93

9. Tumble frequency and swimming speed determinations ...... 94

ix CHAPTER I

INTRODUCTION

Escherichia coli and Salmonella typhimurium use transmembrane

chemoreceptors to sense chemical gradients in their environment as they swim through it. Changes in the level of beneficial or harmful chemicals result in

altered rotation of flagellar appendages that leads to the net migration of cells in a favorable direction. This behavior, termed bacterial chemotaxis, requires four cytoplasmic proteins, CheA, CheW, CheY, and CheZ that relay chemosensory ·

information from chemoreceptors to components of the flagellar motor via a series of intracellular protein phosphorylation reactions, much like the signaling pathways in eukaryotes. Two other proteins, CheR and CheB, catalyze chemoreceptor methylation and demethylation, respectively, that leads to attenuation of receptor signaling.

The cheA locus of E. coli and S. typhimurium encodes two forms of the histidine protein kinase CheA. The long form, CheAt_, plays an essen1ial role in chemotactic signaling by first autophosphorylating at a histidine residue (His-48) and then donating its phosphate group to either CheY or CheB. The short form,

CheA8 , is produced from the same reading frame by translation starting at the codon corresponding to Met-98 of CheAt_. Thus, CheA8 is identical in primary sequence to CheAt_, except that it lacks the N-terminal portion containing His-48 2 and thus cannot support chemotaxis to mutants lacking Che1'_.

In contrast to the well defined roles of each of the other chemo1axis signaling proteins, the specific role of CheAs in chemotactic signal transduction has remained elusive. CheAs, however, exhibits many signaling-related properties consistent with its shared structural organization with CheAa_. (i)

CheAs can transphosphorylate kinase-deficient Che1'_ mutants in vitro and complements CheA kinase-deficient mutations in vivo. (ii) Genetic and immunological studies indicate that CheAs interacts with other chemotaxis components such as Che1'_, CheW, and chemoreceptors, which comprise a ternary signaling complex. (iii) CheAs, but not Che1'_, binds to and accelerates the dephosphorylating activity of CheZ in vitro. (iv) CheA5 and Che1'_ are produced in approximately equal amounts in vivo. Although it exhibits these activities, CheA5 is not essential for chemotaxis under standard laboratory conditions since cells that lack its expression remain chemotactic.

The goal of my investigations was to identify the physiological role played by CheA5 in E. coli chemotaxis. Despite being non-essential for chemotaxis under standard laboratory conditions, I hypothesized that CheAs participates in tactic responses to environmental stimuli under experimental condi1ions not yet examined. To test my hypothesis, I sought to answer two fundamental questions: (i) what other signaling-related activities does CheA5 exhibit; and (ii) under what experimental conditions do cells require CheA5 for op1imal chemotaxis. 3

In this dissertation, I provide experimental evidence that is consistent for

a role played by CheA5 in bacterial chemotaxis. Specifically, I show that: (i) in vivo CheAs can restore chemotactic ability to E.coli cells expressing a CheA_ variant that lacks a domain required to regulate CheA_ autophosphorylation in response to environmental signals through the coupling protein CheW and chemoreceptors. (ii) motile members of the Enterobacteriaceae family that are pathogenic or commensal in the lower gastrointestinal tract of vertebrates coexpress CheA_, CheA5 , and CheZ; and (iii) cells of E. coli that lack CheA5 migrate to L-serine at a slightly faster rate compared to those that express it.

Consequently, when both cell types are matched against each other in a chemotaxis assay those that lack CheAs are at a competitive advantage with respect to their ability to track and respond to L-serine. These behavioral differences are more pronounced under anaerobic conditions. On the basis of these observations, I conclude that: (i) CheAs can perform many tasks of chemotactic signal transduction except that of autophosphorylation; (ii) CheA5 may play a role in chemotactic signal transduction among motile enteric bacteria in general; and (iii) CheAs appears to play an inhibitory role in chemotaxis, which may be relevant to cells responding to environmental stimuli under conditions not typically encountered in the laboratory. CHAPTER II

LITERATURE REVIEW

2.1. INTRODUCTION

Motility and tactic responses to a wide variety of environmental stimuli play a predominant role in the lifestyles of many members of the microbial world.

Such behavior not only enables both eukaryotic and prokaryotic microorganisms to respond to changes in environmental conditions, but also al lows them to seek out sites of microbial attachment during the establishment of many different symbiotic relationships (e.g., Rhizobium attaching to legumous plants) as well as allowing the interplay between cells during multicellular differentiation (e.g., aggregation and swarming of myxobacteria during fruiting body de1Jelopment).

Since the synthesis and operation of the machinery that propels these organisms through their environment demands a significant investment in cellular resources and energy, it certainly suggests that motility and taxis impart a substantial selective advantage under certain conditions.

The aim of this literature review is to familiarize the reader with the basic features of bacterial chemotaxis that enables motile cells to migrate toward beneficial compounds and away from harmful ones. For two reasons, I specifically review the fundamental concepts of E. coli motility and chemotaxis.

4 5

First, it was the experimental system that I used in defining a role for CheA8 in chemotaxis. Second, E. coli chemotaxis represents one of the more well understood systems that has provided researchers with insights into stimulus­ response behavior at the molecular level (reviewed in 8, 130). I begin with a brief discussion on how these bacteria propel themselves and how they sense their chemical environment. I will follow with an in depth view of the signal transduction pathway that mediates chemotactic responses with an emphasis on the biochemical nature of chemotactic signaling, the role played by each component and their interactions, the genetics of the system, and finally a look at the unique features of the cheA locus whose gene products were the central focus of this study.

2.2. CHEMOSENSING AND BEHAVIORAL RESPONSES OF MOTILE

BACTERIA

E. coli swims through its environment by rotating long semi-rigid flagellar filaments that function much like a ship's propeller. Individual cells possess five to ten flagella, each connected to a flagellar motor, that are distributed randomly over the cell surface in a pattern known as peritrichous flagellation. Although each motor rotates independently, flagellar filaments are brought together by hydrodynamic forces to form a single bundle that turns in unison at the rear of a moving cell. The rotation of each flagellar motor can assume two directions: counterclockwise (CCW) rotation leads to formation of the flagellar bundle that 6

propels the cell forward in a relatively "smooth" path; and clockwise (CW)

rotation which results in the bundle flying apart causing the cell to execute an

erratic behavior called a tumble (121). When the majority of filaments return to

CCW rotation the bundle reforms and the cell swims smoothly again, usually in a

new direction as a result of reorientation of the cell by Brownian motion.

Unstimulated swimming consists of smooth runs of approximately 1 s fol lowed by

intermittent tumbles of 0.1 s (14). The result is a three dimensional random walk

that permits cells to find new food sources.

E. coli is attracted to various simple sugars, amino acids, and dipeptides

and are repelled by fatty acids, alcohols, and some heavy metals (reviewed in

86). It also exhibits behavioral responses to a variety of other environmental stimuli, including those to changes in oxygen tension, temperature, pH, light

intensity and wavelength, and osmolarity. Attractant and repellent compounds are sensed directly by means of their cognate binding to specific chemoreceptor

proteins and not through their beneficial or harmful effects upon cellular metabolism (3). Thus, while some metabolites fail to attract bacteria, some non­ metabolite analogs do (52). Tactic responses to chemoattractants are profoundly sensitive since a single bacterium can detect roughly a one one­ thousandth fold change in chemoeffector concentration in the micromolar range

(119).

The small size and rapid movement of bacteria essentially preclude sensing strategies based on spatial comparisons of chemical stimuli (16). 7

Instead, cells determine their heading in chemical gradients by measuring

changes in concentration as a function of time. Since E. coli swims at speeds of

10-20 body lengths per second (77), comparing current chemoreceptor occupancy with occupancy during the previous few seconds enables cells to make measurements over distances of several body lengths (119). When a cell senses that it is swimming up an attractant gradient or down a repellent gradient it reduces the likelihood of a tumble (14). Since CW rotation corresponds to tumbly behavior and CCW to running (70), runs up a gradient are extended because transition of flagellar rotation from CCW to CW is suppressed. What once was a random walk has now become a purposeful biased random walk.

Chemosensory perception, therefore, enables a cell to modulate its swimming behavior by linking changes in the concentrations of various chemoeffector molecules to the rotational bias of each flagellar motor.

2.3. BACTERIAL CHEMORECEPTORS

E. coli detects chemotactic stimuli with chemoreceptors known as transducers. Four different transducers have so far been identified in E. coli

(31 ). The first three, which also are shared with the closely related enteric bacterium Salmonella typhimurium, are those that mediate responses to serine

(Tsr), aspartate and maltose (Tar), ribose and galactose (Trg). The fourth, which is unique to E.coli, mediates responses to a variety of dipeptides (Tap).

Some chemoeffectors bind directly to their cognate receptor, i.e., aspartate to 8

Tsr, while others bind to their receptors via a periplasmic binding protein, e.g.,

the binding of galactose to Trg via the galactose binding protein (GBP) (50).

Transducers that are specific for the sugars maltose, galactose, and

ribose detect external levels of chemoeffector concentrations and do not

facilitate their transport into the cell. Carbohydrate attractants such as mannitol,

mannose, and glucitol, however, are sensed by a very different mechanism.

Unlike conventional chemoreception, these sugars are transported into the cell

by the phosphoenol-pyruvate(PEP)-dependent carbohydrate phospho­

transferase system (PTS) and somehow are sensed as chemoeffectors during

their uptake and concomitant phosphorylation (71 ). During transport of PTS

carbohydrates, phosphoryl groups are transferred through two common

intermediates, enzyme I (El) and phosphohistidine carrier protein (HPr), and

then to sugar-specific enzymes II (Ell) (104). This phospho-relay event is

hypothesized to generate an intracellular signal that suppresses CW flagellar

rotation, which leads to extended runs that propels cells toward higher substrate

concentrations (79). Recently, Rowsell and coworkers demonstrated that the

chemotaxis signaling proteins CheA, CheW, and CheY are required for

chemotaxis to sugars of the PTS system ( 107).

Finally, transducers can also mediate responses to changes in other environmental parameters. Tsr, for example, functions as a thermoreceptor (83), with temperatures up to 37EC acting as an attractant stimulus. High concentrat­

ions of serine blocks thermosensing, presumably by locking Tsr in a 9

conformation that no longer allows the receptor to respond to changes in

temperature (84). Tsr also mediates a repellent response to weak organic acids,

w~ich are detected because they lower the cytoplasmic pH (64).

2.3.1. STRUCTURE AND LIGAND INTERACTIONS OF CHEMORECEPTORS

Transducers are integral membrane proteins of about 550 amino acids

that span the inner membrane of bacterial cells. Each transducer has a short

amino-terminal portion located within the cytoplasm; followed by a hydrophobic transmembrane sequence, TM1; a periplasmic input or sensory domain; a

second transmembrane sequence, TM2; and a cytoplasmic output or signaling

domain that includes two flanking a-helical methylation segments, K1 and R 1

(Fig. 1) (67). Milligan and Koshland used cross-linking studies with the transducer Tar from S. typhimurium to establish that transducers exist as homodimers, both in the absence and presence of ligand (92). Although protomeric exchange can occur within a given population of transducers, heterodimers between different receptors do not appear to form.

The periplasmic sensory domain of S. typhimurium Tar transducer, which has been analyzed by X-ray crystallographic methods (90), binds a single aspartate molecule at the dimer interface at one of two non-overlapping symmetric binding sites. Binding at one site causes the other to become unavailable. Similarly, the E. coli Tsr binds one serine molecule per dimer (72).

The cytoplasmically located C-terminal signaling domain exhibits a high degree 10

sensory domain periplasm transmembrane cytoplasmic segments membrane -----4. cytoplasm

methylation segments ~-....!!.!..-' ---~

signaling domain

Figure 1. Membrane topology and domain organization of MCP molecules. TM represents transmembrane spanning segments. Bullets represent acceptor sites of MCP methylation in the methylation segments, RI and Kl. 11 of similarity among all transducers identified thus far. Even those from the

archaebacterium Halobacterium salinarium retain -35% amino acid identity to a central portion of the signaling domain of E. coli and S. typhimurium transducers

(143). Genetic studies support the hypothesis that this domain functions as the site of interaction between transducers and the chemotaxis proteins, CheA and

Chew (75).

Chemoreceptors undergo reversible methylation during chemotactic responses and, thus, were first known as methyl-accepting chemotaxis proteins or MCPs (66). Transducers typically have four or five methylation sites per protomer: three of these are in K1 and one or two are in R1 (Fig. 1) (62). The methylation sites of E. coli Tar receptor have been identified as glutamate residues although several glutamine residues, which are encoded at the DNA leve1, are later converted to glutamate residues by post-translational deamination (63). Covalent modification of transducers serves to attenuate chemoreceptor signaling, a process termed adaptation (see below).

2.3.2. TRANSMEMBRANE SIGNALING

Chemoreceptors share a number of features with prototypical eukaryotic receptors. They both possess an extracytoplasmic domain that binds a specific ligand that controls the activity of an intracellular signaling domain, both form dimers, and both bind a single ligand. Bacterial transducers, however, exist as dimers under all sensory states examined thus far. The mechanism of 12 chemoreceptor signaling, therefore, does not appear to involve ligand-induced dimerization, which typifies eukaryotic receptor signaling. The work of Falke and coworkers (39) supports this contention. They introduced disulfide bonds between subunits of the Tar receptor and found that they did not interfere with chemotactic signaling (39).

Alternatively, chemoreceptor signaling could involve propagation of a conformational change across the inner membrane. These events could occur by two distinct mechanisms: (i) an inter-subunit mechanism where a change in the relative position of the ligand-binding domains of transducers leads to a change in the relative positions of their cytoplasmic domains and (ii) an intra­ subunit mechanism whereby conformational changes occur independently within each subunit of the dimer. Since experimental evidence exists that is consistent with both mechanisms, transmembrane signaling by chemoreceptors may involve elements of both intra- and inter-subunit movement (67, 80, 118).

2.3.3. CELLULAR DISTRIBUTION OF CHEMORECEPTORS

One of the more intriguing observations made recently in the field of chemotaxis is that chemoreceptor molecules are not randomly distributed but are clustered, often at either pole of the E. coli cell (82). The receptor-associated signaling proteins, CheA and CheW, also localize in clusters that are located predominately at the cell poles (82). In cells lacking expression of all chemoreceptor classes, CheA and CheW were distributed throughout the 13

cytoplasm. Conversely, in cells lacking CheA or CheW, receptor clustering and

polar localization were substantially diminished. Such observations suggest that

interaction of all three signaling components is essential to establish or maintain

a polar clustered distribution of chemoreceptors.

The functional significance of receptor clustering in E. coli, however, is

interesting because bacteria would have no functional significance in

chemosensing since they are too small to compare differences in concentrations

of chemoeffector molecules at the opposite ends of a cylindrical cell. In contrast,

large eukaryotic cells, which do deploy their sensory receptors in patches, do so

presumably in order to make spatial discriminations. The functional significance

of polar clustering of chemoreceptors was recently investigated by Berg and

Turner (17) who observed that the sensory system of E.coli does not favor one

cell end over the other. Thus demonstrating that chemoreceptor clusters at any

one pole do not serve to guide bacteria towards chemical gradients.

2.4. BIOCHEMISTRY OF THE CHEMOTAXIS SIGNALING PATHWAY

The signal transduction pathway that mediates bacterial chemotaxis in E. coli and S. typhimurium links changes in the concentration of chemoeffector molecules to altered swimming behavior via a phosphorelay system (128, 54). It is composed of seven cytoplasmic proteins that transmit sensory information from transducers to switch components of flagellar motors. Three proteins central to this pathway belong to a large family of signaling proteins termed two- 14 component regulatory systems that mediate a variety of adaptive responses in all known prokaryotes and in some lower eukaryotes. This family of proteins includes a kinase sensor, which autophosphorylates in response to some environmental cue, and a response regulator, which removes the phoshate from the phosphorylated kinase sensor (reviewed in 4, 28). The autophosphorylating kinase sensor of chemotaxis is CheA and the two response regulators are CheB and CheY.

Transducers by themselves exhibit no known catalytic activity. Rather, they function to modulate the activity of the receptor-bound CheA kinase (26).

CheA distributes information by passing its phosphoryl group to CheB and CheY

(55, 141 ). Phospho-activated CheB functions to modify the methylation state of transducers (78). Phospho-CheY functions to increase the probability of CW flagellar rotation by interacting with the flagellar motor switch ( 11, 105, 106).

Thus, by controlling the flux of phosphate through the system, transducer signaling not only triggers immediate changes in flagellar rotation via the CheA­

Che Y excitation pathway, but also sets in motion the attenuation of its own signaling via the adaptation pathway, which involves CheA and CheB (Fig. 2).

CheA is central to chemotaxis: it integrates sensory information simultaneously from each class of transducers. Found in vitro as a dimer, CheA forms stable complexes with two CheW monomers and a single transducer dimer (42, 88). These ternary complexes serve as the central processing unit of chemotaxis by regulating CheA autophosphorylation in response to receptor 15 occupancy. The CheW protein is necessary for CheA-transducer coupling, however, its mechanistic role in receptor-modulated control of CheA activity remains obscure. All three proteins must be assembled together for sufficient

CheA activation to occur (25, 43, 96). Neither CheW nor transducers alone have a significant effect on CheA activity. Mutants which lack either CheW or transducers cannot generate CW flagellar rotation despite wild-type levels of

CheA and CheY (25). CheA kinase activity is more than 100-fold greater in ternary complexes relative to unbound CheA (26). However, binding of ligand to the ternary complex greatly diminishes CheA autokinase activity, reduces phospho-CheY levels, and results in an increased probability of CCW flagellar rotation (25, 26, 96). Thus, cells encounter less frequent tumbles or changes in direction as they swim towards attractants.

The chemotaxis sensory pathway is unique in several respects. First, unlike most sensory response systems where ligand-bound receptors generate a positive intracellular signal, ligand-bound chemoreceptors generate a negative signal, i.e., they inhibit CheA kinase activity that results in lower levels of phospho-CheY . Second, CheB and CheY are unlike the majority of two­ component response regulators that typically function as regulatory DNA binding proteins that influence gene expression. CheB exhibits catalytic functions that modify the signaling properties of transducers and CheY influences the rotational bias of flagellar motors. 16

·. :® I ' • • o,...... ~' .. .. ~, ,. CH . 3

Figure 2. Circuitry of the chemotactic signal transduction pathway. A, B, R, W, Y, Z represent the chemotaxis proteins CheA, CheB, CheR, CheW, CheY, and CheZ, respectively. MCP represents transmembrane chemoreceptor proteins. A./As represent the long and short forms of CheA, respectively. Solid, long dashed, and short dashed arrows represent dephosphorylation, phosphorylation, and methylation/demethylation reactions, respectively. 17

2.5. ADAPTATION RESPONSES

The ability of bacteria to sense spatial gradients by a temporal

mechanism requires memory on a very short time scale. E. coli possess two feedback mechanisms that work to cancel out recent signal inputs so that cells can be poised to respond to new changes in chemoeffector concentrations (Fig.

2). One involves a biochemical strategy common to both eukaryotic and prokaryotic sensory adaptation: alteration of receptor signaling states by reversible covalent modification. The other involves reducing phospho-CheY levels by a novel dephosphorylation reaction.

The chemotaxis protein, CheR, exhibits methyl transferase activity and functions to catalyze the transfer of methyl groups from S-adenosylmethionine

(AdoMet) to side-chain carboxyl groups of glutamate residues of chemoreceptor molecules ( 124). In contrast, CheB exhibits methylesterase activity and functions to hydrolyze glutamyl methylester bonds (127). This reaction results in the production of methanol and the regeneration of unmodified glutamate residues. CheB also catalyzes irreversible deamidation of glutamine to glutamate (63).

The competing activities of CheR and CheB determine the methylation state of each transducer molecule. Since CheR activity is constitutive, increased

CheB activity, a result of its phosphorylation by CheA, produces transient fluctuations in the methylation level of transducers. In a static chemical environment, approximately half of the sites are modified ( 109). Attractant 18 binding causes a conformational change in transducers that leads to an increase in receptor methylation by exposing glutamyl side chains to CheR and sequestering methyl glutamyl groups from CheB (114). Increased methylation levels result in desensitization of ternary complexes such that higher concentrations of attractant would be required to shift the signaling state back to an active signaling complex. Simply stated, methylation of chemoreceptors tends to counteract the affects of attractant binding.

Adaptation also occurs at the level of CheY. Phospho-CheY has a very short intrinsic half-life in vitro of approximately 10 s (56, 141). Although its rapid turnover contributes to adaptation responses, this mechanism alone does not account for the extraordinary sensitivity of the chemosensory system. Block and co-workers (24) estimated that as few as four ligand bound transducers out of the several thousand that are found per cell are capable of eliciting a chemotactic response within 0.1 s. Attempts to explain this phenomenon have resulted in the identification of an auxiliary protein CheZ, a protein whose function is to accelerate the dephosphorylation of phospho-CheY. Several lines of evidence have established that CheZ, does in fact, control the in vivo levels of phospho-CheY. Blat and Eisenbach demonstrated that soluble but not switch­ bound phospho-CheY binds to CheZ and that upon its dephosphorylation is released from CheZ (20). CheZ, a dimeric protein, undergoes further oligomerization upon interaction with phospho-CheY (23). This oligomerization likely results in the activation of CheZ dephosphorylating activity (22). These 19

observations suggest that increased phospho-CheY levels may be counter-

balanced by activation of Chez.

2.6. GENETICS OF THE CHEMOTAXIS SIGNALING PATHWAY

Components of the chemotactic signal transduction pathway in E. coli and

S. typhimurium were identified initially by analyzing the behavior of numerous strains that harbored mutations in che genes. These studies, conducted many

TABLE 1. Components of the chemotaxis signal transduction system Protein M, [kDa] Mutant phenotype• Biochemical activity MCPb -55 blind to specific ligands chemoreceptor CheA 78 smooth histidine kinase CheA8 69 ? histidine kinase CheW 18 smooth couples CheA to MCPs CheY 14 smooth binds to and alters the rotation of flagellar motors CheZ 24 tumble stimulates P-CheY dephosphorylation CheR 32 smooth methyltransferase CheB 36 tumble methylesterase/amidase • Phenotypic classes based on swimming behavior of cells in liquid media. b Wild-type swimming behavior, but blind to any one of the cognate ligands for each type of MCP. Mutants defective in both tarltsr run incessantly. years ago, nonetheless provide insightful information. Initially, genes whose products are involved in chemotaxis, were described by several groups (9, 100,

101, 122). Their genetic studies revealed the existence of six genes cheA, cheB, cheR, cheW, cha Y, and cheZ that were required for normal chemotaxis.

Cells harboring null mutations in any one gene can be divided into one of two phenotypic classes based on their swimming behavior (Table 1). The first class 20

mocha operon mRNA meche operon mRNA

Figure 3. Genomic organization of the mocha and meche operons in E. coli. P represents native promoter sequences.

describe mutants that exhibit a smooth or run phenotype; a result of the inability to tumble because their flagellar motors predominantly turn CCW.

Representatives of this class are those cells harboring mutations in the genes cheA, cheW, che Y, and cheR. Mutations in the first three genes eliminate any response to repellents, which would normally cause CW rotation. Strains with null mutations in cheR are unable to methylate their MCPs and, therefore, exhibit smooth swimming behavior until they encounter a repellent, which results in incessant tumbly behavior. The second class describes mutations that confer a tumbly phenotype as the result of a predominantly CW-bias. Representatives of this class harbor null mutations in the genes cheB and cheZ; cheB mutants, that are unable to demethylate their MCPs, are also defective in receptor adaptation.

Both cheB and cheZ mutants can respond with CCW flagellar rotation to large jumps in attractant concentration.

The six che genes are organized into tandem operons at 42 min on the E. coli chromosome {Fig. 3) (10). The upstream mocha operon {for motility and chemotaxis) initiates with the transcription of two genes whose products form membrane channels that serve to convert electrochemical ion potential into 21 mechanical rotation of flagellar motors (34, 19) and terminates with the genes cheA and cheW. The downstream meche operon (for methylation and chemotaxis) initiates with the genes tar and tap that encode chemoreceptors followed by the chemotaxis genes cheR, cheB, che Y, and cheZ.

2.7. GENE PRODUCTS OF THE cheA LOCUS

The cheA genes of E. coli and S. typhimurium encode two similar CheA proteins, CheA.._ (78 kDa) and CheA8 (69 kDa), as the result of translation from two distinct in-frame initiation sites (65, 123, 126). As a result, both proteins share identical primary sequence except that CheAs lacks the first 97 amino acid residues (Fig. 4). Since both forms of CheA are synthesized in approximately equal amounts from wild-type cells of E. coli ( 113, 134) neither protein can be considered a minor player in chemotactic signaling.

CheA.._, catalyzes the transfer of the y-phosphoryl group of ATP to the amino-terminal histidine residue, His-48 (55,56). CheA8 , which lacks His-48, does not autophosphorylate (54). His-48 appears to be the only site of CheA autophosphorylation since proteins which lack this residue either naturally (the short form, CheA8 , lacks the region containing His-48) or by mutagenesis

(CheA.._H48Q), do not autophosphorylate in vitro and do not support chemotaxis in vivo (98, 133). Inter-allelic complementation studies indicate that at least some CheA functions can be mediated in trans within heterodimers. Swanson et al. (133) demonstrated that proteins which lack His-48 transphosphorylate the 22

kinase-deficient CheA variant, CheA470GK. Similaiiy, Wolfe and Stewart (137)

demonstrated that CheA5 can transphosphorylate Che.A.47CGK in vitro and can

support chemotaxis by cells that express this mutant protein in vivo.

Sequence analysis, genetic studies, and protease cleavage patterns all

indicate that CheA is organized into distinct functional domains (Fig. 4) (103).

The centrally located kinase domain of CheA contains amino acid sequences

which are thought to be responsible for the binding and catalysis of ATP. Two

domains, P1 and P2, located at the amino-terminal end, are separated by a

stretch of amino acids that resemble charge-rich linkers found in multi domain

regulatory proteins (140). P1 contains the site of autophosphorylation, His-48,

and is required for phosphotransfer to CheY and CheB (132). P2 likely assists

in the interaction between the phosphorylation site in P1 and amino acid

sequences located in CheY (95, 132). Two domains, M and C, located at the

carboxy-terminus, appear to play important roles in enabling CheA to receive

sensory information from transducers. Several reports present evidence that

support this contention. First, missense mutations located throughout the M

receptor r phosphotransfer ..., ~ autokinase -.j ~coupling~

.:;

CheAs

Figure 4. Organization of structural domains and associated functions shared among both forms of CheA. Wavy lines adjacent to the P2 domain represent predicted flexible linker regions. 23 domain suppress, in an allele-specific manner, certain CW-biased transducer mutations, indicating this domain of CheA may physically interact with transducers (75). Second, mutations in the C domain phenotypically suppress, in a non-allelic manner, both CW- and CCW-biased transducer mutations, suggesting that this domain may regulate CheA kinase activity by influencing its ability to interact with CheW (103). Third, certain CheA variants lacking portions of their carboxy-terminus autophosphorylate in an unregulated fashion, i.e., their kinase activity is not affected in response to environmental signals transmitted through reconstituted ternary complexes (30).

Whereas CheAi_ plays an essential role in chemotactic signaling, CheA5 seemingly does not since cells that express CheAi_, but not CheAg, perform chemotaxis under standard laboratory conditions (113). Although the specific role played by CheAs in chemotaxis is currently not known, it does exhibit many signaling-related properties consistent with its shared structural organization with CheAi_. In vitro, CheAg mediates receptor modulated transphosphorylation of mutant CheAi_ proteins deficient in kinase activity or lacking carboxy-terminal segments required for ternary complex formation (133, 137, 138). In vivo, CheA5 restores chemotactic ability to cells that express either kinase-deficient or truncated CheAi_ mutant proteins (137, 138).

In addition to its interaction with CheAi_, presumably as a CheA5-CheA_ heterodimer (137), CheAg interacts physically with other components of the chemotactic signal transduction pathway, e.g., CheW, CheY, and CheZ (42, 51, 24

88). Ternary complexes consisting of chemoreceptors, CheW, and

CheAt/CheA8 heterodimers respond in vitro to the signaling states of ligand bound transducers (133). These complexes have a higher affinity for CheY as compared to either CheA_/CheAs or CheW alone (88). Based upon these observations, CheAs could play a role in regulating CheA_ phosphorylation in a stimulatory or inhibitory capacity.

Additionally, CheA8 , but not CheA_, forms complexes with CheZ in vitro.

By immunoprecipitation with either anti-CheA or anti-CheZ antibodies, the ratio of subunits within the complex is estimated to be 1O to 30 CheZ molecules per

CheAs molecules (51). Recently, Wang and Matsumura (134) demonstrated that although approximately equal amounts of CheA_ and CheA8 were present in whole cell lysates from wild-type cells, only CheA8 co-precipitates with CheZ; an observation which is consistent with the formation of CheA8 /CheZ complexes in vivo. These complexes, reconstituted in vitro, exhibit greater dephosphorylating activity on phospho-CheY than does CheZ alone. Overproduction of CheA5 in wild-type cells increases CCW bias of flagellar motors; an effect that is dependent on the presence of CheZ (134). Collectively these data suggest that

CheA8 may serve some role in chemotaxis. The focus of my investigations was to identify such a role. CHAPTER Ill

MATERIALS AND EXPERIMENTAL METHODS

3.1. CHEMICALS AND BIOLOGICAL REAGENTS

Chemical reagents were purchased from Fisher Scientific (Pittsburgh,

Pa.), Boehringer Mannheim (Indianapolis, Ind.), and Sigma Chemical Company

(St. Louis, Mo.). Antibiotics also were purchased from Sigma Chemical

Company. Radio labeled materials were purchased from Amersham (Arlington

Heights, Ill.) and the bicinchoninic acid (BCA) protein assay reagent was obtained from Pierce Biochemicals (Rockford, Ill.). Modifying enzymes and restriction endonucleases were purchased from Promega Corp. (Madison, Wis.),

New England Biolabs (Beverly, Mass.) or Gibco BRL (Gaithersburg, Md.).

3.2. CONSTRUCTION OF PLASMID-BORNE cheA MUTANT ALLELES

Plasmids used in genetic complementation studies and strain construction are listed in Tables 2 and 3, respectively. While plasmids pMPC3, pAR1 .cheA, and pCS31 encode CheA_ and CheAs, those of pAF1, pAR1 .cheA, and pCS531 b encode only CheAs. Plasmids pMPC3, pAF1, and pAF1 .cheA5 G470K, which encodes a kinase-deficient variant of CheA5 (98), permit regulatable expression from the Serratia marcescens tryptophan promoter, P1rp.

25 26

TABLE 2. Plasmids used in complementation studies Plasmid Relevant cheA alleles• Derivation, construction, or reference pAF1 Ptrp cheA5 CheA5 only (137) pAF1 .cheA5 G470K Ptrp cheA501 (98) pAR1 ilcheA 2.1-kb BamHI deletion of cheA pAR1 .cheA P* cheA Encodes CheAt/CheA5 (42) pAR1.cheA5 Pmc cheA5 Encodes CheA5 only (137) pAJW7 ilcheA pDV4 derivative (125) pAJW103 Ptrp(cheA-cheW) Encodes CheAt_/CheA5 , CheW pCS31 P.,. cheA CheAt_, CheA5 (18) pCS521b P.,. cheA521 CheAs (18) pMPC3 P1rn cheA Encodes CheA /CheAs • Indicates linkage of specific cheA alleles to particular regulatable promoters.

Similarly, plasmids pAR1 .cheA and pAR1 .cheA5 permitted regulatable express- ion from the IPTG-inducible promoter, Pmc linked to /ac/0 , the gene which encodes the repressor protein. Likewise, plasmids pCS31 and pCS521b permitted regulatable expression from the E. coli B arabinose-inducible promoter araPBAD linked to araC that encodes the arabinose activator-repressor protein.

Plasmids pAR1 .cheA and pAR1 .cheA5 were predominantly used in complementation studies to confirm chromosomally encoded cheA mutations that conferred non-chemotactic (Che·) behavior. The vector control plasmid, pAR1, was constructed by excising a 2.1-kb BamHI fragment containing the entire cheA gene from pAR1 .cheA and, thus, no cheA gene products are expressed.

Plasmid pAR1 .cheA503/(Am), kindly provided by R. Stewart (University of

Maryland, College Park, Md.), encodes truncated CheA proteins CheAt_5031(Am) and CheAg5031(Am) that lack both carboxy-terminal domains, M and C. 27

Similarly, plasmids pAR1 .cheA542L(Am), pAR1 .cheA572G(Am), and

pAR1 .cheA616K(Am), also provided by R. Stewart, encode both CheA proteins that lack portions of their C-terminus.

pJN001, a plasmid generated by subcloning a 2.1-kb BamHI fragment containing the entire wild-type cheA gene from pAR1 .cheA into the unique

BamHI site of pAlter-1 (Promega Corp.), was constructed for the purpose of mutagenizing the cheA start(S) AUG initiation codon. Using standard oligonucleotide-directed mutagenesis, an EcoRV site first was introduced between the Shine-Dalgarno sequence and the AUG initiation codon of start(S).

This restriction site was used to track each mutation during various in vitro manipulations. The resultant plasmid, pJN004, was subsequently mutagenized to generate the two pairs of cheA start(S) alleles used in this study (See Fig.

158 of section 4.8.1. Design and construction of strains containing cheA start(S) mutations, chapter IV). The first pair of alleles, cheAM98V(St and cheAM98V(S)", were constructed by altering the cheA start(S) AUG codon to

GUG or GUC, respectively, both of which results in the replacement of methionine-to-valine at position 98 in CheAt.. Similarly, those of the second pair, cheAM98L(St and cheAM98L(S)", were constructed by altering the cheA start(S) AUG to UUG or CUC, respectively, both of which results in the replacement of methionine-to-leucine at position 98 in CheAt.. Consequently, the first member of each pair encodes CheA5 , in addition to mutant CheAt., while the other member does not. The cheAM981 allele (pHS7), which encodes 28

TABLE 3. Plasmids used in strain construction Plasmid Relevant cheA alleles Description, construction or source• pAR1 .cheA503/(Am) CheA5031(Am)b pAR1 .cheA572G(Am) CheA572G(Am)b pAR1 .cheA616K(Am) CheA616K(Am)b pJN001 cheA 2.1-kb BamHI insert in pAlter-1 pJN004 cheAEcoRV EcoRV site in cheA pJNOOS cheAM98V(St CheA_M98V and CheA5 pJN006 cheAM98V(S)" CheA_M98Vonly pJN007 cheAM98L(St CheA_M98L and CheAs pJN008 cheAM98L(S)" CheA_M98L only pJN009 cheA BamHI fragment from pJN004 in pUC19 pJN0010 AcheA::Kmr Kan cassette from pUC4K into pJN009 •Where listed, description indicates encoded cheA gene product. b Truncations of both CheAi_ and CheAg.

CheA_M981 but not CheAs, was kindly provided by J. 5. Parkinson (University of

Utah, Salt Lake City, Utah).

The AcheA:: Kmr allele was constructed by subcloning the 2. 1-kb BamH I fragment derived from pJN004 into the BamHI site of pUC19. A 1-kb EcoRV-Nrul fragment was excised and replaced with a Hincll fragment from pUC4K

(Pharmacia Biotech, Piscataway, N.J.) containing the kanamycin-resistance marker by blunt end ligation.

3.3. BACTERIAL STRAINS

Bacteria used in this study and their relevant genotypes are listed in

Table 4. With the exception of strain ATCC 25922 (AJW676), a clinical isolate derived from a wound abscess obtained from the American Type Culture 29

Collection (Rockville, Md.), all E.coli strains are derivatives of E.coli K-12.

RP437, wild-type for chemotaxis, was kindly provided by J.S. Parkinson. S. typhimurium, laboratory strain LT-2, also wild-type for chemotaxis was kindly provided by Kelly Hughes (University of Washington, Seattle, Wash.). Primary clinical isolates of enteric bacteria were kindly provided by Roberta Carey

(Clinical Microbiology Laboratory, Loyola University Chicago Hospital). These isolates, recovered from a variety of human infectious material, were identified by conventional biochemical tests performed routinely in the clinical microbiology laboratory. Wild-type laboratory strains of Erwinia sp. were graciously provided by David Coplin (Ohio state University, Columbus, Ohio).

Strain RP1515 [cheA169Y(Am)] contains an amber mutation between start(L) and start(S). Thus, cells of this strain express only CheA5 as a result of premature termination of CheA.. translation (123).

3.3.1. Isolation of cheA deletion and truncation mutants

Strain AJW1071 was constructed by transducing the recA::cml allele from strain MH6- into the cheA deletion strain, RP9535 (138). The non-motile, non-chemotactic strain, AJW399, was designed for the purposes of cloning and identifying portions of cheA sequences from enteric species other than E. coli. It was constructed by cotransducing the motA-tar deletion and transposon insertion zec::Tn10 from strain AJW398 into AJW390; a highly motile, chemotactic derivative of JM107. Resultant tetracycline-resistant transductants 30 were screened for their inability to swarm in semisolid agar. This deletion was confirmed by complementation analysis with a set of /..che22 derivatives that carry missense mutations in each of the chemotaxis genes of interest (101 ).

To construct recombinant strains of E.coli bearing premature translation stop codons within the 3' region of cheA, a marker-rescue (also known as allelic replacement) strategy was employed (46). This strategy facilitates the transfer of plasmid-borne mutations into the E. coli chromosome and, thus, preserves single copy expression of any one cheA mutant allele from its native promoter.

This method involved three steps: i) integration of ColE 1-based plasmids into the E.coli chromosome via homologous recombination; ii) resolution of plasmid sequences from the host chromosome; and iii) selection and screening of recombinant strains. The first two steps were achieved by transforming the

Che• host strain, CP366, containing a polA(Ts) allele marked by zig::Tn10 (99).

In this strain, ColE1-based replicons are unable to replicate extrachromosomally

(35). In the first step, plasmids containing the ampicillin resistance gene (bla}, pAR1 .cheA5031(Am), pAR1 .cheA572G(Am), and pAR1 .cheA616K(Am) were transformed into CP366. By incubating transformants at the non-permissive temperature of 42°C on plate media containing ampicillin, ampr colonies termed plasmid integrates were identified. In the second step, resolution of integrated plasmid sequences was carried out by growing selected isolates at 37°C in liquid broth without ampicillin for more than 10 generations. In the last step, ampicillin-sensitive segregates, carrying mutant cheA sequences, were identified 31 replica plating. Since cells expressing truncated CheA proteins should not to exhibit wild-type chemotactic behavior, segregates now containing mutant cheA genes were identified and screened for their inability to swarm in semisolid agar.

Generalized transduction, involving the use of phage P1 kc, was used to transfer each mutant allele out of its po/A(Ts) host into the PolA+ Rec+ Che+ recipient strain RP437. The cheA locus exhibits 20% linkage to eda, a gene whose product enables cells to utilize galacturonic acid as a sole carbon source.

Transductants were, therefore, selected on minimal media containing galacturonic acid and subsequently screened for non-chemotactic behavior.

The resultant transductants were made recA::cml by using strain MH6-, which was kindly provided by A. Binney (Purdue University, Lafayette, Ind.) as the source of donor DNA. Each allele was confirmed by complementation analysis and by direct-cycle sequencing of amplified sequences spanning the 3' portion of cheA.

A description of the resultant strains used in identifying chemotaxis­ associated activities of CheAs are as follows. Cells of strain AJW377, which carry the allele cheA503/(Am), express the truncated proteins CheAt_5031(Am) and CheAs5031(Am) that lack the entire Mand C region reported to interact with

MCPs and CheW, respectively. Similarly, cells of strain AJW425 and AJW430, which carry the alleles cheA572G(Am) and cheA616K(Am), respectively encode truncated CheAt_ and CheAs lack sequentially smaller portions of the M domain and presumably all of the C domain. 32

3.3.2. Isolation of chromosomal cheA start(S) mutants

Strains carrying cheA start(S) mutations were constructed by transferring plasmid-borne cheA mutations from the ColE1-type replicons, pJN004-pJN008, into the E. coli chromosome of strain AJW484, which was originally constructed by transferring the plasmid-borne L1cheA:: Kmr allele from pJN010 into the chromosome of strain CP366. Cells of strain AJW484 exhibit non-chemotactic behavior in semisolid agar as the result of no cheA expression. Similarly, a Che· derivative of the E.coli clinical isolate (strain AJW751) was constructed by transducing the L1cheA:: Kmr allele from strain AJW484 into E. coli ATCC 25922.

Recombinant strains of AJW484 possessing cheA start(S) mutations were screened for their ability to carry out chemotaxis in semisolid agar.

To avoid the possibility of introducing any pleiotropic effects associated with the po/A(Ts) allele, po/A+ was transferred out of strain RP437 into each recombinant strain by generalized transduction. Two flanking genetic markers, zig::Tn10 and rha+, a gene whose product enables cells to utilize L-rhamnose as the sole carbon source, were chosen to select for and screen resultant transductants, respectively. Transductants were selected on minimal media containing L-rhamnose as the sole carbon source and subsequently screened for tetracycline sensitivity by growth in media supplemented with chlortetracycline hydrochloride (50 µg/ml) and fusaric acid (12 µg/ml) (85).

A description of the resultant strains used in identifying the conditions under which cells require CheA5 for chemotaxis are as follows. Cells of strain 33

TABLE 4. Bacterial strains used in this stud~ Strain Relevant genot~E!e Source E.coli AJW377 RP437 cheA503/(Am) recA::cml eda+ (137) AJW382 cheAw.t This study AJW390 Che+ recA+ derivative of JM107 (89) AJW394 BL21 (DE3) pBPM11 This study AJW398 RP437 A(motA-tar)2211 zec::Tn10 eda· (89) AJW399 AJW390 A(motA-tar)2211 zec::Tn10 (89) AJW425 RP437 cheA572G(Am) recA::cml eda+ (137) AJW430 RP437 cheA616K(Am) recA::cml eda+ (137) AJW484 CP366 AcheA::Kmr (892) AJW532 cheAEccRv thi-1 thr-1(Am) leuB6 his-4 rha+ This study AJW533 cheAM98V(St po/A+ rha+ This study AJW534 cheAM98V(S)- po/A+ rha+ This study AJW535 cheAM98L(St po/A+ rha+ This study AJW536 cheAM98L(S)- po/A+ rha+ This study AJW676 ATCC25922 clinical isolate ATcca AJW751 AJW676 AcheA::Kmr This study AJW767 BL21 (DE3) pDE2 This study AJW779 BL21 (DE3) pBPM8 This study AJW780 BL21 (DE3) pBPM10 This study AJW1319 AJW535 tht* This study AJW1320 AJW536 /eu+ This study AJW1335 AJW535 /eu+ This study AJW1336 AJW536 thr+ This study AJW1071 RP437 A(cheA)1643 eda+ recA::cml (138)

BL21(DE3) F- ompT hsdS8(r8• M8") gal dcm (DE3) (131 )b CP366 po/A(Ts) zig::Tn10 thi-1 thr-1(Am) leuB6 his-4 (99) rha DHSa F'/endA1 hsdR17(rK-mK+) supE44 thi-1 (139) recA1 gyrA (Nair) re/A1 A(lac/ZYA-argF)U169 deoR( 80dlacA(lacZ)M15) HCB543 RP437 recA sr/::Tn10 (68) JM107 F' traD36 lacfl A(lacZ)M15 proA+a+1 (142) e14"(McrA") A(lac-proAB) thi gyrA96 (Nair) endA1 hsdR17(rK-mK+) re/A1 supE44 K-12 Wild-type ATcca MH6- recA::cm/ A Binney RP437 thr-1(Am) leuB6 his-4 metF-159(Am) eda-50 (102) rpsL 136 thi-1 lacY1 ara-14 mtl-1 xyl-5 tonA31 tsx-78 RP1515 RP437 cheA169Y(Am) (123) 34

TABLE 4. - continued Strain Relevant genotype Source E.coli RP1616 RP437 .1(cheZ)6725 (73) RP9535 RP437 AcheA1643 ecJa• (74)

S. typhimurium LT-2 Wild-type (K. Hughes )d

Erwinia ssp. E. amylovora Wild-type D. Coplin8 E. carotovora Wild-type D. Coplin8 E. herbicola Wild-type D. Coplin8 E. stewartii Che· Mot· species D. Coplin8 •American Type Culture Collection, Rockville, Md. b Novagen, Inc., Madison, Wis. c Purdue University, West Lafayette, Ind. d University of Washington, Seattle, Wash. •Ohio State University, Columbus, Ohio

AJW532, which carry the genetically silent EcoRV site in cheA, express both wild-type CheAt_ and CheA5 . Cells which carry the alleles cheAM98V(St (strain

AJW533) and cheAM98L(St (strain AJW535) express CheAt_M98V and

Che~M98L, respectively, and retain synthesis of wild-type CheA5 . Cells which carry the alleles cheAM98V(S)" (strain AJW534) and cheAM98L(S)" (strain

AJW536) also express CheAt_M98Vand Che~M98L, respectively, but do not express CheAg. Each chromosomally encoded cheA allele was verified by immunoblotting and direct-cycle sequencing of the region spanning start(S).

The genes, leuB6 and thr-1(Am), were chosen as selectable markers in order to differentiate between cells that make CheA5 from those that do not.

Revertants prototrophic for L-threonine or L-leucine metabolism, Thr• and Leu•, respectively, were selected by plating 1-5 x 107 cells of each member of the 35

M98l pair onto M63 plate media, which lacked the appropriate amino acid.

Thus, cells that express CheAs (strain AJW535) could be differentiated from

those that did not (strain AJW536) based on their requirement of exogenous L­

threonine or L-leucine for growth.

3.4. BACTERIAL MEDIA

Cells of all laboratory strains and clinical isolates, except those of Erwinia

ssp., were grown in tryptone broth (TB) containing 1% Bacto-tryptone and 0.5%

NaCl; or in Luria-Bertani broth (LB) composed of TB and 0.5% Bacto-yeast extract; or in M63 minimal media containing 1X M63 salts (30 g/L KH2P04, 70 g/l K2HP04, 20 g/l (NH4}iS04, 5 mg/l FeS04), 100 µg/ml thiamine, 1mM

MgS04, 0.2% o-galacturonic acid as the sole carbon and energy source, and the amino acids required for growth: L-threonine, L-leucine, L-histidine, and L­ methionine (20µg/ml). Cells of Erwinia ssp. were grown in PCG media containing 1% peptone, 0.1 % casamino acids, and 0.2% o-glucose. All plate media consisted of 1.5% Bacto-agar. Media was supplemented with the antibiotics (ampicillin, 100 µg/ml; carbenicillin, 100 µg/ml; kanamycin sulfate 50

µg/ml; chloramphenicol, 40 µg/ml; tetracycline hydrochloride, 15 µg/ml) when necessary. Ampicillin, carbenicillin, and kanamycin were suspended in water at

1000 X concentrations, sterilized through 0.2 µM filters, and stored at -20EC.

Chlorampheni~I and tetracycline were suspended in 95% (vol/vol) EtOH at

1000 X concentrations, sterilized through 0.2 µM PTFE hydrophobic filters 36

(VWR, Chicago, Ill.), and stored at -20°C.

Prescribed amounts of isopropyl-J3-o-thiogalactopyranoside (IPTG) were added for induction of cheA genes linked to P tac promoters. Stock solutions of

IPTG, at 1OOmM, were prepared in water, sterilized through 0.2 µM filters, and aliquoted for storage at -20°C.

3.5. GROWTH CONDITIONS

3.5.1. Measurement of microbial growth

Bacterial growth was monitored spectrophotometrically with a Bausch &

Lomb Spectronic 20 at a wavelength of 600 nm (optical density (00600]). Single colonies were inoculated into 3-ml of LB and incubated overnight (ON) at either

31°C or 37°C. Pre-inoculum cultures of early exponential phase cells were prepared by diluting ON cultures 1 :200 into 5-ml of fresh LB media. Upon

reaching an 00600 of 0.1-0.2, cultures were diluted 1: 1O into 50-ml of prewarmed, LB media in a 500-ml Erlenmeyer flasks. Aeration levels were controlled by adjusting the rate of shaking (225-275 rpm) on a rotary shaker.

3.5.2. Proteus mirabilis swarmer cell isolation

Hyperflagellated swarmer cells of P. mirabilis were prepared by spotting

5-µI aliquots of cultures, pre-grown in LB, onto the center of LB plates.

Following incubation at 37°C for 10 h (the time required for swarming cells to cover most of the surface), cells were harvested by swirling 10-ml of sterile 37 phosphate-buffered saline (PBS; 20mM sodium phosphate [pH 7.2), 100mM

NaCl) over the surface of each plate as described by Belas et al. (12).

3.6. GENETIC MANIPULATIONS

3.6.1. Transformation of competent E. coli

To make cells of E.coli strains competent, they were grown in 5-ml cultures of Y-b media (2% Bacto-typtone, 0.5% Bacto-yeast extract, 0.5%

MgS04, adjusted to pH 7.6 with KOH) at 37°C under aeration until 00600= 0.3.

Cells were diluted 1:20 into fresh prewarmed Y-b media grown under identical

conditions until 00600=0.48, chilled for 5 min on ice, and centrifuged (5 min, 4°C, at 3K rpm in a Beckman SS-34 sorval rotor). Supernatants were decanted, cell pellets were resuspended in ice-cold TBF1 solution (30 mM potassium acetate,

100 mM KCI, 1O mM CaCl2, 50 mM MnCl2, and 15% [VN] glycerol, adjusted to pH 5.8 with acetic acid), and chilled on ice for 5 min. Cells again were collected by cenrtifugation, supernatants decanted, cell pellets resuspended in ice-cold

TBF2 (10 mM MOPS, 75 mM CaCl2, 10 mM KCI, and 15% (vol/vol) glycerol, adjusted to pH 6.5 with KOH), and chilled on ice for 15 min.

Transformation of cells proceeded by gently mixing a 50-µI aliquot of thawed competent cells and 1-3 µI of plasmid DNA (1-10 ng), prepared as described in section 3.9.1. Plasmid and chromosome DNA preparation, into chilled microcentrifuge tubes. Contents were incubated on ice for 30 min, followed by a heat-shock pulse at 37°C for 60 s, and placed on ice for 2 min. 38

SOC media (2% Bacto-tryptone, 0.5% Bacto-yeast extract, 1O mM NaCl, 2.5 mM

MgCl2, 10 mM MgS04, and 20 mM o-glucose) was added to transformed cells,

incubated at 37°C for 1 h, plated onto selective media, and were incubated at

37°C overnight.

3.6.2. Generalized transduction

Generalized transduction, involving the use of P1kc (a clear plaque variant) was used in all genetic transfers (120). Donor phage lysates were prepared by inoculating single colonies into 5-ml of LB grown at 37°C ON. 50-µI of ON cultures were inoculated into two tubes containing fresh LB media (5-ml, each supplemented with 0.2% o-glucose and 5 mM CaCl2); to one tube no phage was added (control); and to the other 100-µI of P1 phage lysate was added. Growth at 37°C continued until lysis was complete. Lysates were clarified by the addition of 100-µI chloroform followed by centrifugation at 3000x g for 15 min.

Recipients were prepared by diluting ON cultures into fresh LB media

(supplemented with 0.1% o-glucose and 10 mM CaCl2) with an initial 00600 =

0.02-0.04. These cultures were aerated at 37°C until they reached an

00600=0.4-0.6. Cells were then infected with 0-, 10-, or 100-µI donor lysate and incubated with no aeration for 20 min at 37°C. To prevent phage readsorption,

200-µI of 1 M sodium citrate was added, the cells centrifuged, and the resultant pellets resuspended in 500-µI LB supplemented with 250 mM sodium citrate. 39

The cells were then incubated at 37°C for 1 h, centrifuged, resuspended in 100-

µI of 1 M sodium citrate, and plated onto appropriate selection media.

Transductions involving donor and/or recipient strains which carried the po/A(ts) allele were grown at the permissive temperature of 31°C.

3.6.3. Replica plating for mutant selection

Replica plating, which permits the simultaneous transfer of large numbers of individual colonies from master plates, was used to screen for ampicillin­ sensitive segregates that resulted from allelic replacements in a po/A(ts) background. Individual colonies were grown on master plates by appropriately diluting overnight LB cultures in sterile PBS such that 100-200 colonies per plate eould be screened. Colonies about 3 mm in diameter were first transferred from the master plate to velveteen cloth, and from the cloth, transferred to LB plates containing 100 µI/ml ampicillin and then to complete media LB plates lacking ampicillin. By noting the marked orientation of each plate, ampicillin-sensitive segregates were identified by those which grew on LB plate media alone. All incubations were performed ON at 30°C.

3.6.4. Complementation analysis

Intra-allelic complementation was used initially to confirm the identity of those cheA alleles that encoded various truncated forms of CheA. This was carried out by testing non-chemotactic cells of strain AJW377, AJW425, and 40

AJW430 for their ability to generate wild-type chemotactic behavior when transformed with plasmids carrying the allele cheA (pMPC3) but not alleles cheA503/(Am) [pAR1cheA503/(Am)], cheA572G(Am) [pAR1cheA572G(Am)] or cheA616K(Am) [pAR1cheA616K(Am)], respectively. Complementation studies also were used to assess the ability of CheAs to restore chemotactic behavior in cells that lacked its expression or those that expressed various truncated forms of CheA.

To confirm the large motA-tar deletion in the nonchemotactic strain,

AJW399, interallelic complementation and recombination studies were performed using a set of 'A.che22 derivatives that carry missense mutations in each of the chemotaxis genes of interest (101 ). Lysogens of 'A.che22 and its derivatives were identified by testing cells from plaque centers for immunity to homoimmune /... phages and sensitivity to heteroimmune /... phages. Lysogens were then scored for chemotactic ability in semisolid agar plates.

3. 7. METHODS TO ASSESS MOTILITY AND CHEMOTACTIC BEHAVIOR

Wild-type cells of E.coli, when grown between 30°C to 37°C, are considerably more motile when they are in early- to mid-exponential phase (2,

6). Chemotactic behavior and/or che gene expression from cells of all enteric bacteria were, therefore, monitored from cultures grown in TB at 31°C until they

reached mid-exponential phase (00600 - 0.3). Motility and chemotactic behavior of cultures also were assessed by microscopic inspection with a Nikon bright 41 field microscope.

3. 7 .1. Swim plate analysis

Swim plate analysis was used to qualitatively and quantitatively measure chemotactic ability. When chemotactic cells of E.coli are inoculated near the center of a petri dish containing TB semisolid agar, they migrate radially outward to form 2 or 3 concentric bands. These bands represent subpopulations of cells responding to spatial gradients of certain metabolizable chemoattractants present in tryptone (a casein hydrolysate). This behavior requires that cells transport, metabolize, and assimilate these chemoattractants.

The outer band represents cells that consume L-serine and most of the oxygen, the second band are those that consume L-aspartate, and when present the third band are those that consume all of the L-threonine ( 136). Microscopically, cells swimming in semisolid agar behave as if they were negotiating a three­ dimensional maze.

TB semisolid agar, otherwise known as swim plate media, consisted of TB and 0.2-0.3% Bacto-agar. A 5-µI aliquot of a given culture grown to mid­

5 6 exponential phase (00600=0.3; 5 X 10 to 5 X 10 cells) was spotted onto the surface of swim plates followed by incubation at 31°C in a humid environment (a closed incubator containing a dish of water). Generally migration rates were determined by measuring the displacement of the outer band (representing taxis to L-serine) as a function of time from at least 3 plates per strain per experiment. 42

Swim plates were photographed with Polariod type 665 positive-negative film by placing plates on a light box illuminated obliquely from below and viewed against a dark background.

3.7.2. Anaerobic swim assay conditions

Cultures and swim plates grown anaerobically were handled in a model

1029 glove box anaerobic chamber (Forma Scientific, Marietta, Ohio) maintained at 35°C. The atmospheric environment was comprised of 85% N2, 10% H2, and

5% C02 (Airco Medical Gases, Carol Stream, Ill.) and was monitored with a

Model 110 gas analyzer (Coy Laboratory Products, Torrence, Calif.). Tryptone swim plates, supplemented with 0.5% KN03 (45), were prepared and allowed to solidify under ambient air, and then placed in the chamber for at least 16 h to facilitate the exchange of gases. Ten-µI aliquots of ON cultures grown anaerobically (-1 X 105 cells) were spotted onto the surface of reduced, i.e., environments that are relatively free of oxygen, swim plate media.

3.7.3. Competitive swim assay analysis

To determine whether CheA5 confers a competitive advantage or disadvantage to cells with respect to swimming ability , I mixed equal numbers of

6 cells (- 5 X 10 ) that synthesize CheAt_M98L and CheA5 (allele cheAM9BL(St) with those that synthesize only CheAt_M98L (allele cheAM9BL(S)") and spotted them in triplicate onto the surface of TB swim plates, which were incubated 43 aerobically at 31°C or anaerobically at 35°C. During the next several hours, a band of cells, which represents cells responding to a gradient of L-serine as a result of its consumption, migrated outward. Just before this band reach the opposite edge of the petri dish, I removed randomly three 10 µM samples from this band and respotted them onto fresh plates. These passages were repeated twice more. I removed randomly three samples per plate from the migrating band of cells when they reached the opposite edge of the third swim plate. To determine the number of cells of each respective population, i.e., those that synthesize CheA5 and those that do not, I subjected these samples to serial dilutions in PBS and plated those dilutions onto M63 media supplemented with o­ glucose and L-leucine or L-threonine. All plates were also supplemented with L­ methionine and L-histidine. This difference in amino acid supplementation permitted me to distinguish cells of both classes because I paired a Thr+ prototroph of strain AJW535 (strain AJW1319; allele cheAM98L(St) with a Leu+ prototroph of strain AJW536 (strain AJW1320; allele cheAM98L(S)"). To rule out the possibility that differences in amino acid supplementation would effect migration rates, a Leu+ prototroph of strain AJW535 (strain 1335; allele cheAM98L(St) was paired with a Thr+ prototroph of strain AJW536 (strain

AJW1336; allele cheAM98L(S)").

3.7.4. Swimming speed and tumble frequency determinations

Cells were grown and observed as described (6, 7), except that they were 44

grown at 31°C in TB and were maintained at 31°C during video measurements

on a thermally-controlled microscope stage (144). All observations were made

4.75 h after inoculation (mid-exponential phase). Swimming speed and cell

tumble frequency were determined as described (6).

3.8. PROTEIN METHODS

3.8.1. CheA overexpression and purification

Plasmids designed to overexpress CheA variant proteins are listed in

Table 5. Each corresponding cheA allele was subcloned by excising the 2.1-kb

Ndel-BamHI fragment from the corresponding pAR1-based plasmid to the 4.7-kb

Ndel-BamHI fragment of the pET-14b (Novagene, Madison, Wis.). This vector

facilitates the overexpression and rapid single step affinity purification of N­

terminal His•Tag protein fusions. Resultant plasmids were transformed into the

host strain [BL21 (DE3)] that contains a chromosomal copy of the T7 RNA

polymerase gene under the control of the IPTG inducible lacUV5 promoter (131 ).

Addition of IPTG to growing cultures of BL21 (DE3) transformants results in

efficient transcription of cheA.

Cultures of each transformant were grown initially in 3-ml LB media

supplemented with 100 µg/ml carbenicillin. Upon reaching an 00600 of ~ 0.2,

cultures were back diluted into 50-ml of fresh LB media containing 100 µg/ml

carbenicillin and aerated at 37°C under until they reached an 00600 of ~0.6. An aliquot, representing the uninduced control, was removed before addition of 45

TABLE 5. Plasmids used for CheA overexpression and purification Plasmid• Relevant cheA alleles Derivation, construction, or reference pBPM8b cheAM98V(S)­ pET-14b His•CheAi_M98V pBPM10b cheAM98L(S)" pET-14b His•CheAi_M98L pBPM11b cheAM98/(S)­ pET-14b His•CheAi_M981

pDE1b cheA pET-14b His•CheAi_ + CheA5 pDE2 cheAs pET-14b His•CheA5 pHS7 cheAM98/(S)" CheA, M981 only (113) • pDEI encodes His•Tag w.t. CheAL and w.t. CheAs.

IPTG (0.4 mM final cone.). Three h post-induction, cells were chilled on ice and

harvested by centrifugation at 2000 X g for 20 min at 4°C. Whole cell lysates

were prepared by sonicating cell pellets resuspended in 20 mM Tris-HCI (pH

7.9), 5 mM imidazole, 0.5 M NaCl. Single step purification of N-terminal His•Tag

2 CheA fusion proteins was achieved by adsorbing resultant lysates over Ni +

columns and were eluted and handled as described in the 6th edition of the pET

System Manual (Novagen [58]). Transformants of BL21 (DE3) expressing CheA5 were grown at 30°C in LB and induced with 0.1 mM IPTG for only 2 h.

3.8.2. SOS-PAGE and western blot analysis

SDS-polyacrylamide gel electrophoresis (SOS-PAGE) was performed with a discontinuous gel system as described (69) using a SE 600 protein gel apparatus (Hoefer Scientific Instruments, San Francisco, Calif.). Whole cell

lysates of total protein were prepared by sonicating resuspended cultures in 0.1 vol of phosphate-buffered saline (pH 7.2). Soluble and insoluble fractions of cellular proteins were prepared by centrifuging sonicates at 2, 000 X g for 30 min 46 at 4 °C. Protein concentration of total or cellular fractions were determined with a bicinchoninic acid protein assay reagent kit (Pierce, Rockford, Ill.). Cellular protein (40-100 µg per well) were suspended in 2X SOS-PAGE sample-load buffer and were loaded onto 10% or 15% polyacrylamide gels containing 0.1 %

SOS. Resolved proteins were detected either by Coomassie blue staining or by western blotting. For western blots, resolved proteins were electrophoretically transferred overnight to 0.45 µm nitrocellulose filters using a Trans-Blot Cell apparatus (Bio-Rad Laboratories, Richmond, Calif.). Filters were blocked with

5% (wUvol) BLOTTO in TBS-T (20 mM Tris·HCI [pH 7.6], 137 mM NaCl, and

0.05% to 0.2% Tween 20) for 1 hat room temperature before being subjected to sequential incubations (1 h each at room temp.) with affinity-purified polyclonal rabbit anti-E. coli CheA antibodies (generated by standard procedures [48]) or purified CheZ at 1: 1500 dilutions. Immobilized immune complexes were probed with goat anti-rabbit immunoglobulin G (heavy- and light-chain specific) alkaline phosphatase conjugated antibodies, diluted at 1: 15, 000. All washes between incubations were carried out with TBS-T (0.05 % to 0.2 % Tween 20). Color development was achieved with nitroblue tetrazolium and 5-bromo-4-chloro-3- indoylphosphate as described (48).

3.8.3. CheA in vitro autophosphorylation assays

To assess the in vitro autophosphorylation ability of wild-type CheAL, wild­ type CheAs, and Che~ variants, reaction mixtures containing 1O µM of each 47

CheA protein alone were incubated in TKMO buffer (50 mM Tris-HCI [pH 8.0], 50 mM KCI, 5 mM MgCl2, 0.5 mM dithiothreitol, 1O % glycerol}, and 50 µM (y-

32P]ATP (specific activity typically -5000 cpm/pmol) (Hess et al., 1991 ). At the conclusion of a 15 min labeling period at room temperature (RT}, reactions were terminated by the addition of 2X SOS-PAGE sample-load buffer. Samples were subjected to SOS-PAGE (10%), fixed in 45% (vol/vol) methanol, 7% (vol/vol) glacial acetic acid, 10% (vol/vol) glycerol, and 0.5% (vol/vol) dimethyl sulfoxide, dried for 2 h, and visualized by autoradiography. Total CPMs were determined by with a Betascope 603 blot analyzer (Betagen, Mountainview, Calif.).

3.8.4. Generation of affinity purified anti-E. coli CheA polyclonal antibodies

High-titer antiserum recognizing both forms of E.coli CheA was generated by immunizing a female New Zealand White rabbit (approx. 5 Kg) with 50 µg

(emulsified in Freund's incomplete adjuvant) per injection of His•Tag purified

CheA Immunizations were carried out as described (47). Preimmune serum was collected before the first immunization. The rabbit was injected subcutaneously at 5 separate sites in the dorsal-lumbar region. The rabbit was subsequently boosted 4 weeks later. Two weeks after boosting, 35 ml of whole blood was taken via an ear puncture. Serum was collected by allowing whole blood to coagulate at 37°C for 60 min, then overnight storage at 4°C, and finally centrifuged at 1O,OOOx g for 1O min at 4 °C. Specificity and titer of anti-E coli

CheA antibodies in the serum was determined by enzyme-linked immunosorbant 48

assay (ELISA) as described (48).

Polyclonal anti-E coli CheA antibodies were affinity purified with an

Aminolink™ Kit (Pierce, Rockford, Ill). Immobilization of purified His•Tag CheA

to Aminolink columns was carried out according to the manufacturer's protocol.

An approximate 64.7% binding efficiency was achieved as determined indirectly

by comparing the protein concentration of the original solution to that in the

effluent using BCA protein analysis. Affinity purification was carried out accord­

ing to the manufacturer's protocol. Antibody titer was determined by ELISA

3.9. MOLECULAR BIOLOGY METHODS

3. 9.1. Plasmid and chromosomal DNA preparation

Large and small scale preparations of double-stranded ( ds) plasmid DNA were done by the alkaline lysis method (60) using either the Wizard Mini-prep

purification system (Promega) or the plasmid Midi kit by QIAGEN®. Rapid

isolation of chromosomal DNA was based on the principle of salting-out proteins

as described (91 ). Cells grown in 50-ml LB were harvested, washed 1X in PBS,

resuspended in SET buffer (75 mM NaCl, 25 mM EDTA, 20 mM Tris-HCI [pH

7.5]), and lysed in the presence of 1 mg/ml lysozyme. Preparations were

incubated at 37°C for 0.5 h, and then incubated at 55°C for 2 h treated with 0.1 vol of 10% SDS and 1 mg/ml proteinase K. Cellular proteins were precipitated by the addition of 0.33 vol of 6 M NaCl and 1 vol chloroform followed by incubation for 1 h at RT on a rotating vertical wheel. The aqueous phase 49 containing DNA was collected by centrifugation at 3,200 Xg for 1O min. Plasmid and chromosomal were washed 1X in 70% ethanol, dried, and dissolved in TE buffer (10 mM Tris-HCI [pH 8.0], 1 mM EDTA). DNA concentrations, estimated by UV spectrophotometric determinations, were calculated by the formula: concentration (mg/ml) = Asj20.

3.9.2. DNA cloning and restriction endonuclease analysis

Restriction endonuclease digestions of chromosomal of plasmid DNA were performed in accordance with the manufacturer's specifications and as described (110). Digested DNA fragments were electrophoretically separated on agarose gels 0.8%-1.8% (wt/vol) in 1X TAE buffer (40 mM Tris-acetate, 1 mM

EDTA). Fragments < 500 bp were purified with the Wizard PCR Prep

Purification System (Promega), otherwise they were purified by treating agarose pieces with 2.5 vol of 6 M NaCl followed by incubation at 50°C for 15 min, and isolation of DNA with the Wizard Mini-prep purification system (Promega).

Cohesive-end ligation of DNA fragments was performed as described by (111 ).

Vector arms of plasmid DNAs were usually dephosphorylated at 5' termini using calf intestinal alkaline phosphatase (Promega). PCR generated DNA fragments were ligated into pGEM-T according to the manufacturer's protocol (Promega).

Competent cells of DH5a. were routinely transformed with ligated plasmid DNA 50

3.9.3. Site-directed mutagenesis

Site-directed mutagenesis reactions using the Altered SitesTM in vitro

Mutagenesis System (Promega), were performed in accordance with the manufacturer's specifications. Single-stranded DNA (ssDNA) templates for these reactions were isolated from Tcr transformants of strain JM109 bearing pAlter-1 plasmids. Specifically, single colonies were grown at 37°C in 5-ml of

TYP broth (1.6% Bacto-tryptone, 0.5% Bacto-yeast extract, 0.5% NaCl, 0.25%

K2HP04) with aeration. After 30 min., cultures were infected with the helper phage R408 at a multiplicity of - 10, grown an additional 6 h, and harvested by centrifugation. Single-stranded DNA was isolated from supernatants by use of the Wizard™ M13 DNA Purification System (Promega).

Mutagenesis reactions were performed by annealing 0.25 pmol phosphorylated ampicillin repair oligonucleotides and 0.25 pmol of phosphory­ lated cheA mutagenic oligonucleotides (Table 6) to 1.25 pmol ssDNA (pJN004)

templates in annealing buffer (20 mM Tris-HCI [pH 7.5), 10 mM MgCl2, 50 mM

NaCl). Incubations proceeded at 70°C for 5 min followed by slow cooling to RT for 30 min. After annealing of primers, synthesis and extension reactions proceeded in extension buffer (10 mM Tris-HCI [pH 7.5), 10 mM ATP, 2 mM

OTT, 0.5 mM dNTPs) with 2 U T4 ligase and 10 U T4 DNA polymerase.

Reactions proceeded for 1.5 h at 37°C. Competent cells of BMH71-18 mutS were transformed with mutagenized plasmid DNA. Transformants were incubated ON at 37°C in LB containing ampicillin. Competent cells of DH5a 51 were subsequently transformed with ds-plasmids isolated from transformants of

BMH71-18 mutS. Apr colonies were grown at 37°C in 5-ml LB supplemented with ampicillin ON. ds-DNA plasmids were isolated for restriction endonuclease mapping and sequence analysis.

3. 9.4. Polymerase chain reaction

PCR primers (Table 6) used to amplify the 5' region of cheA genes from motile enteric bacteria were designed on the basis of alignments of nucleotide sequences encompassing the 3' region of motB from E. coli, S. typhimurium and

Bacillus subtilis and from nucleotide sequences encompassing the 5' region of cheA derived from E. coli, S. typhimurium, Caulobacter crescentus and

Pseudomonas aeruginosa. Forward primer F1 .1 is complementary to a highly conserved sequence located within the 3' region of the motB gene; the reverse primer R2 is complementary to a highly conserved sequence located within the cheA gene. Each PCR reaction, performed in a 100-µI volume, contained chromosomal DNA template (-100 ng), each primer (0.1 µM), deoxynucleoside

triphosphates (0.2 mM each), MgCl2 (2.0 mM), PCR buffer (20 mM TrisHCI and

50 mM KCI), and 2.0 U Taq DNA polymerase (Gibco-BRL). PCR reactions were performed in an Omn-E Thermal Cycler (Hybaid, Middlesex, UK) beginning with a 5 min denaturation step at 94°C and a 3 min "hot start" addition of Taq polymerase at 80°C, followed by 30 cycles of denaturation for 1 min at 94°C, annealing for 1 min at 52°C and extension for 1 min at 72°C. A final 5 min 52

Table 6. Oligonucleotide primers used in this study Primer 5' to 3' nucleotide sequence Use• CS.1 AAACGAAGGATATCATGCAAG cheA EcoRV mut. CS.2 AAACGAAGGATATCGTGCAAGAACAGC cheAM98V(St mut. CS.3 AAACGAAGGATATCGTCCAAGAACAGC cheAM98V(S)" mut. CS.4 AAACGAAGGATATCTTGCAAGAACAGC cheAM98L(St mut. CS.5 AAACGAAGGATATCCTCCAAGAACAGC cheAM98L(S)" mut. CS.6 GGCTTCAGCGTTTTGCAGGA EcoRV seq. F1.1 CCTGATGATGCCGTCAACCG motB forward PCR R2 CTGGGTGATAACCAGCTCGCC cheA reverse PCR RSA GTATTCAAAGCTGGCGGCATCCGG cheA start(S) seq. 904(3) ATTCCATAACAGCATTCAGC cheA5031(Am) seq. RW.1 CACCGCCGCACCGGTGAGTAAAAAGG cheA572G(Am) and cheA616K(Am) seg. • Denotes whether oligonucleotides served as primers used with associated cheA alleles for mutagenesis (mut.), sequencing (seq.), or PCR amplification. extension step at 72°C was employed.

3.9.5. DNA hybridizations

To clone and identify portions of cheA from enteric species other than E. coli competent cells of AJW399 were transformed with recombinant pGEM-T plasmids and plated onto LB agar plates supplemented with ampicillin.

Transformants bearing cheA nucleotide sequences were identified by colony lift hybridization using the BioTrace NT Binding Matrices protocol (Gelman

Sciences, Ann Arbor, Mich.). All filters were incubated for approximately 16 h at

45°C and probed with a 516 base pair Ndel-San fragment obtained from the 5' region of the E. coli cheA gene present in the plasmid pAR1 .cheA (137).

Southern blot hybridizations were performed first by digesting chromosomal DNA with appropriate restriction enzymes, separating the resultant 53

fragments on 0.8% (wt/vol) agarose gels, followed by blotting onto 0.45 µm

nitrocellulose filters via the alkaline transfer method (112). Hybridizations

proceeded at 55°C for - 16 h with ds-DNA probes (100 pmol) labeled by

incubation with [a-32P]ATP by using the Ready-to-Go DNA labeling kit

(Pharmacia). Membranes were subsequently washed twice with 2X SSC (0.3 M

NaCl, 0.03 M sodium citrate)-0.1% (wt/vol) SOS for 10 min and twice with 0.1X

SSC-0.1 % SOS for 5 min. Blots were exposed to X-ray film at -70°C for at least

24 h.

3.9.6. DNA sequence analysis

Sequence determinations of plasmid-borne cheA alleles were performed

with Sequenase, version 2.0 (U.S. Biochemicals, Cleveland, Ohio) using the

chain termination method of Sanger et al. (1977). Chromosomally encoded

cheA alleles were determined by sequencing ds-DNAs using the Circumvent™

Thermal Cycle Dideoxy DNA Sequencing Kit or with an ABI Prism 377

automated DNA sequencer with a Taq-DyeDeoxy Terminator Cycle Sequencing

kit developed by Applied Biosystems Inc. (Perkin-Elmer Corp., Foster City,

Calif.) .. In four separate reactions, chromosomal DNA templates (0.01 pmol) were combined with appropriate 5' end-labeled DNA sequencing primers (0.6

pmol) in the presence of 1X Circumvent™ Sequencing Buffer, 0.2% (vol/vol)

Triton X-100, 3-µI each of four different Circumvent™ Deoxy/Dideoxy

Sequencing Mixes, and 2 U of VentR fJ ( exo·) DNA polymerase. Reactions were 54 performed in an Omn-E Thermal Cycler (Hybaid, Middlesex, UK) with 20 cycles of denaturation for 20 sat 95°C, annealing for 20 sat 55°C and extension for 20 sat 72°C.

All sequencing reactions were subjected to electrophoresis in 6% Long

Ranger™ gels (J. T. Baker, Inc., Phillipsburg, N.J.) containing 1.2X TBE (10.7 mM Tris-HCI, 10.7 mM boric acid, 0.24 mM EDTA}, with 0.6X TBE running buffer and a running temp of 47°C. Gels were dried for 1.5 hand exposed to X-ray film for ON exposure.

Alignments of nucleotide and amino acid sequences were generated by using the University of Wisconsin Genetics Computer Group PileUp program

(version 8. 1) with a gap weight of 3. 0 and a gap length weight of 0. 1. CHAPTER IV

EXPERIMENTAL RESULTS

4.1. IN VIVO COMPLEMENTATION OF CheA C-TERMINAL TRUNCATION

MUTANTS BY CheA5

To assess whether the C-terminal portion of CheA5 , like that of CheAL, receives chemosensory information through its association with ternary complexes, we asked if CheA5 could pass chemosensory information from transducers to CheY in vitro and, thus, support chemotaxis in strains missing portions of their C-termini. For each strain, we introduced a premature termination codon at specific sites along the 3' region of cheA. These strains, which expressed CheA proteins that were missing portions of their C-termini, were non-chemotactic. Initially, we tested whether CheA5 can restore chemotactic ability to cells that express these truncated CheA variants. We first transformed competent cells of strains AJW377[cheA503/(Am)],

AJW425[cheA572G(Am)], and AJW430 [cheA616K(Am)] with plasmids that expressed either CheAL and CheA5 (pMPC3), CheA5 alone (pAF1), or the CheA5 kinase-deficient variant CheA5G470K (pAF1 cheA 8 G470K). Cells of the resultant transformants were inoculated at the center of tryptone swim plates (0.2% agar), incubated at 31°C, and the displacements of the outermost edges of their

55 56

2.0 ,,.-...., su '-" ell ...... ;::::s 1.5 "'O ~ 1.0

AJW437 AJW377 AJW425 AJW430 !icheA cheA503I(am) cheA572G(am) cheA616K(am)

Fig. 5. CheA5 complementation studies with strains that express various CheA truncation mutants. Plots indicate the displacement of the outermost edges of swim colonies produced by cells of strains AJW437(!lcheA), AJW377[cheA503/(Am)], AJW425[cheA572G(Am)], and AJW430

[cheA616K(Am)] transformed with pAF1 .cheA8 G470K (blank bar), pAF1 (hatched bar), and pMPC3 (stippled bar) that direct expression of CheA5G470K, CheA5 , or CheAL and CheA5 , respectively. Radial displacements were measured after 10 h at 31 °C. Error bars represent the standard errors of the mean displacements for 3 separate determinations. 57

3.5

3.0 CheAL + CheAS (pMPC3) • CheA 8 ) I (pAF 1cheA 8 2.5 •/ I E I 0 2.0 ,.. en ~ ::l I m1 "O cu 1.5 I• .rJ 0::: I .ti I /. (pAF1 cheA G470K) 1 .0 • • 8 ,.. ~ .-.-·-·__...--. ./ .--.-.-.-·.- 0.5 ..•

0.0 '--~-'-~--'~~....._~__._~~...._~_._~~...... _~_, 0 5 10 15 20

Time (h)

Fig. 6. Restoration of swimming ability by CheA5 to cells expressing CheA616K(Am). Plots indicate the time course of displacements (radius) of the outermost edges of swim colonies produced by cells of strain AJW430

[cheA616K(Am)] transformed with pMPC3, pAF1, and pAF1 .cheA8 G470K that direct the expression of CheAL and CheA5 , CheA8 , and CheA8 G470K, respectively. Swim plates were prepared and handled as described in section 3.7.1. The standard errors of the mean displacements for 3 independent experiments are shown only when in excess of 0.05 cm. 58 migration were measured.

Transformants of all strains that expressed CheAL and CheA5 from pMPC3 formed two concentric bands in semisolid agar with the outer and inner bands, formed by responses to and consumption of L-serine and L-aspartate.

Whereas cells of strain AJW430[cheA616K(Am)] that expressed CheA5 alone

(pAF1) formed two concentric bands, the outermost of which migrated about 1.2 cm after 10 h incubation, those of strains AJW437[~cheA],

AJW377[cheA503/(Am)], and AJW425[cheA572G(Am)] formed dense irregular swim colonies whose displacements were indistinguishable from those produced by cells that expressed the CheA8 kinase-deficient variant CheA5 G470K (Fig. 5).

Such colonies are characteristic of non-chemotactic cells which cannot tumble.

Cells of strain AJW430 that expressed both CheA proteins (pMPC3) formed an outermost band that migrated at a rate (0.43 cm/h) faster than that formed by cells that expressed CheA5 alone (0.25 cm/h) (Fig. 6). Microscopic inspection of free-swimming cells grown at 31°C in TB until mid-exponential phase revealed that cells of both transformants tumbled occasionally; cells of strain AJW430 which expressed CheA5G470K did not.

4.2. CheA5 OVEREXPRESSION STUDIES

To verify the observations that CheA5 complements allele cheA616K(Am) but not alleles cheA503/(Am) and cheA572G(Am), we transformed competent cells of their respective parental strains with plasmids pCS521 b and 59 pAR1 .cheA8 . These plasmids permit regulatable expression of CheA5 , as measured by western blot analysis (Fig. 19A), under the control of the arabinose­

and IPTG- inducible promoters araP8Ao and tacP, respectively. Cells of the resultant transformants were inoculated at the center of tryptone swim plates

(0.2% agar) supplemented with appropriate amounts of inducer and After incubating the plates at 31°C for 12 h, the displacement of the outermost edges were measured and compared (Fig. 7A). The formation of swim colonies by cells of strain AJW430 [cheA616K(Am)] carrying either plasmid resulted in a response dependent upon the level of induction. Cells induced by either 25 µM

IPTG or between 0 and 50 µM arabinose formed the most rapidly migrating colonies, which resembled those formed by wild-type cells (Fig. 7C). Induction by smaller concentrations of IPTG or by larger concentrations of either IPTG or arabinose yielded colonies that were less well-formed and which migrated less rapidly. For example, cells induced by either 0 or 500 µM IPTG formed small, dense swim colonies that resembled those produced by non-chemotactic cells

(Fig. 78 and D, respectively). In contrast, at all inducer concentrations tested, transformants of strains AJW377 [cheA503/(Am)] and AJW425 [cheA572G(Am)] formed dense irregular colonies indistinguishable in morphology and migration rate from that shown in Fig. 70. Thus, CheA5 can complement defects in chemotaxis associated with allele cheA616K(Am), but not with alleles cheA503/(Am) and cheA572G(Am), in a dose-dependent manner.

Since overexpression of CheA5 inhibited chemotaxis in the 60

A D A:RABINIJZH! • fPTC - -

a --~~~--~~~~~~~~~~~~~~~~ D ECNJ ZlllO INDUCER (µ,M) B c D

Fig. 7. Effect of CheA5 concentration on the swimming ability of cells expressing CheA616K(Am). Displacement of the outermost edges of swim colonies produced by cells of strain AJW430[cheA616K(Am)] transformed with pCS521 b and pAR 1.cheA8 , which express CheA8 from promoters inducible by arabinose and IPTG, respectively, plotted as a function of inducer concentration (A). Radial displacements were measured after 12 h incubation at 31°C. The standard errors of the mean displacements were less than 0.1 cm. (B to D) Photographs of swim colonies produced by cells of strain AJW430 transformed with pAR1 .cheA8 induced with IPTG at concentrations of 0 (B), 25 (C), and 500 (D) µM. 61

3.5 None (pAJW7) ------·

3.0

\ E 2.5 0 \ CheAL + CheA8 (pCS31) -(/) :::J \ "O cu \ c::: 2.0 \

~ CheA8 (pCS521 b) 1.5 ' ·------·

1.0 '--~~~~...... _~~~~__.__~~~~--'-~~~~--' 0 250 500 750 1000

Arabinose (µM)

Fig. 8. Effect of CheA5 overexpression on swimming ability in a wild-type strain. Displacement of the outermost edges of swimming colonies by wild­ type cells transformed with plasmids that express neither CheA protein(•},

CheAL and CheA5 (•),or CheA5 alone(•) are plotted as a function of arabinose concentration. Radial displacements were measured after 11.3 h at 30°C. The standard errors of the mean displacements for duplicate swim colonies were less than 0.1 cm. 62 cheA616K(Am) background (strain AJW430), we tested whether such over­ expression could also inhibit chemotaxis in the wild-type strain HCB543. This strain was transformed with plasmids that expressed both CheAL and CheA5

(pCS31) or CheA5 alone (pCS521b) from the arabinose-inducible promoter araPsAo· As a control, we also transformed strain HCB543 with a plasmid

(pAJW7) that expressed neither cheA gene product. Cells of the resultant transformants were inoculated at the center of tryptone swim plates supplemented with indicated amounts of arabinose and the displacement of the outermost edge of each swim colony were measured (Fig. 8). The displacement by cells transformed with the control plasmid (pAJW7) was unaffected by the presence of arabinose. In contrast, the displacements by cells transformed with the other two plasmids decreased as the concentration of arabinose increased to

250 µM. However, cells that overexpressed CheAL and CheA5 (pCS31) were significantly less sensitive to the arabinose concentration than were cells that overexpressed CheA5 alone (pCS521b).

4.3. IN VITRO TRANSDUCER-MEDIATED TRANSPHOSPHORYLATION BY

CheA5

In collaboration with R. C. Stewart (University of Maryland, College Park,

Md.), a series of in vitro experiments were designed to test whether transducers and CheW regulate CheA5-mediated transphosphorylation of CheA C-terminal truncation mutants in a ligand dependent manner. He performed in vitro 63 receptor-coupled assays by adding mixtures of CheA8 and certain C-terminal

CheAL truncation mutants in the presence of the Tsr transducer and CheW. Two methods were used to monitor these in vitro reactions: (i) direct measurement of the rate by which phosphorylated CheAL appears and (ii) indirect measurement of CheAL kinase activity by monitoring the rate by which phosphorylated CheY appears.

Even though CheA8 was unable to restore any observable chemotaxis to cells of the CheAL 5031(Am) mutant, the coupling of Tsr to CheA8 by CheW greatly stimulated its ability to mediate transphosphorylation of CheAL5031(Am) in a manner similar to that of wild-type CheAL. Such stimulation decreased in the presence of the Tsr-specific chemoattractant L-serine. Whether coupled to Tsr through CheW or not, neither CheA8 alone nor CheAL5031(Am) alone auto­ phosphorylated. In contrast to its effect upon CheA8-mediated transphosphory­ lation of CheAL 5031(Am), coupling of CheA8 to Tsr through CheW did not stimulate CheA8-mediated transphosphorylation of CheAL572G(Am) and, thus, likely explains why CheA8 does not restore chemotaxis to cells of the

CheAL 572G(Am) mutant.

4.4. CHEMOTACTIC BEHAVIOR AMONG CLINICAL ISOLATES OF

ENTEROBACTERIACEAE

Since E. coli and S. typhimurium coexpress both forms of CheA, I hypothesized that other motile species of Enterobacteriaceae would likewise 64

TABLE 7. Primary clinical isolates of Enterobacteriaceae used in this study Strain Relevant featuresa Normal florab Escherichia coif opportunistic pathogen, Mor yes Salmonella typhimurium primary enteric pathogen, Mor no Shigella flexneri primary enteric pathogen, Mor no Klebsiella pneumoniae opportunistic pathogen, Mof yes oxytoca opportunistic pathogen, Mof yes Enterobacter aerogenes opportunistic pathogen, Mot+ yes cloacae opportunistic pathogen, Mor yes Serratia marcescens opportunistic pathogen, Mot+ yes Citrobacter freundii opportunistic pathogen, Mor yes Morganella morganii opportunistic pathogen, Mot+ yes Proteus mirabi/is opportunistic pathogen, Mord yes vulgaris opportunistic pathogen, Mot+d yes Yersinia enteroco/itica primary enteric pathogen, Mor no pestis primary enteric pathogen, Mof no a Mot+ denotes motility behavior by virtue of peritrichous flagella. b Normal flora of the human lower gastrointestinal tract (40). c ATCC25922 (American Type Culture Collection, Rockville, Md). d Isolates of Proteus ssp. display both typical swimming motility in liquid media and swarm motility on surfaces by virtue of hyper flagellated cells. coexpress both forms of CheA. To initiate these studies, I obtained primary clinical isolates of common bacterial enteric pathogens (Table 7). These isolates, which were recovered from a variety of human infectious material, were identified by conventional biochemical tests performed routinely in the clinical microbiology laboratory.

Since all members of the Enterobacteriaceae family grow readily under aerobic and anaerobic conditions, I examined the motility and chemotactic behavior of each isolate by swim plate analysis under both growth conditions

(Fig. 9). Although isolates of E. coli, S. typhimurium, E. cloacae, P. mirabilis, S. 65

Aerobic Anaerobic

Fig. 9. Swim colony morphologies exhibited by various clinical isolates of Enterobacteriaceae. Cells of each isolate were inoculated near the center of tryptone swarm plates containing 0.2% agar and incubated 6-10 h under aerobic or anaerobic conditions as indicated. 66 marcescens, and C. freundii exhibited chemotactic behavior, resultant swim colonies of each isolate varied in morphology and size. Although the swim colonies of each isolate typically formed dense outer rings, I observed some variability with respect to the formation of an inner ring. No inner bands emerged from anaerobically grown swim colonies except those of P. mirabilis, which formed the most prominent inner band that appeared to be composed of several distinct migrating fronts. Cells of the plant related genus Erwinia (Table 4) also displayed chemotactic ability under both conditions, although their migration rates were marginally slower under anaerobic conditions (data not shown). As expected, cells of S. f/exnerii, K. pneumoniae, and K. oxytoca were non-motile under both conditions tested (data not shown).

Chemotactic cells of E. coli typically form 2 concentric bands when placed on tryptone swim plate media. The outer and inner bands represent subpopulations of cells responding to spatial gradients of L-serine and L­ aspartate, respectively. Since chemotactic cells of S. typhimurium, E. cloacae and S. marcescens form bands similar to those formed by E. coli, I tested whether these bands represent tactic responses to L-serine and L-aspartate, respectively. I reasoned that if a migrating band from a separate swim colony were to fuse with either of those formed by E. coli, this would indicate that cells of each population were responding to and consuming the same chemoattractant. When cells of E. coli and either S. typhimurium, E. cloacae or

S. marcescens were spotted separately and allowed to migrate in tryptone swim 67

plate media, both the outer and inner bands of each swim colony fused with

those formed by E. coli (data not shown). As a control, when cells of E. coli were

spotted separately and allowed to migrate into each other, a similar pattern was

observed (data not shown). The results of these band fusion studies are

consistent with the hypothesis that chemotactic cells of several enteric species

track and respond to L-serine and L-aspartate.

4.5. CheA EXPRESSION IN THE ENTEROBACTERIACEAE

I assessed CheA expression from each isolate of enteric bacteria by

immunoblot analysis using affinity-purified polyclonal rabbit antibodies that were

raised against His• Tag purified CheAL. To demonstrate the specificity of this

anti-E. coli CheA antibody, I performed immunoblot analysis on lysates from a variety of chemotactic and non-chemotactic strains of E. coli. Wild-type cells of

E. coli (strains ATCC23716 [K-12], ATCC25922 and RP437) yielded three

prominent immunoreactive proteins (Fig. 1OA). In each case, the two larger

proteins corresponded to CheAL (apparent MW - 73- to 74-kDa) and CheA8

(apparent MW - 62- to 64-kDa). In contrast, the smallest protein (apparent MW

- 49-kDa) was not derived from cheA since cells deleted for cheA (strain

AJW1071) synthesized neither CheAL nor CheA8 yet expressed this cross­

reactive protein.

Cells that possess the cheA169Y(Am) allele (strain RP1515) contain an

introduced UAG termination codon located between the translation initiation sites 68 of CheAL and CheA5 . As expected, these cells synthesized the 62-kDa CheA protein but not the 73-kDa CheA protein. In contrast, cells that possess the cheA98ML allele (strain AJW536), which contains a mutation that interrupts the

CheA5 translation initiation site in cheA, synthesized the 73-kDa CheA protein but not the 62-kDa protein. Whereas the inability to complete CheAL translation resulted in reduced expression of CheA5 , the inability to express CheA5 exerted no effect upon the expression of CheAL.

lmmunoblot analyses were performed on lysates of clinical isolates of S. typhimurium, S. marcescens, E. cloacae and C. freundii, which were grown under similar conditions to those for E. coli. Each clinical isolate exhibiting chemotaxis synthesized two prominent immunoreactive proteins similar in size to

E. coli CheAL and CheA5 . One motile, but non-chemotactic, isolate of C. freundii synthesized only the smaller immunoreactive protein (Fig. 1OB). Chemotactic isolates of E. aerogenes and M. morganii also synthesized both immunoreactive proteins (data not shown). Interestingly, cells of Yersinia enteroco/itica, which exhibit motility only at the permissive temperature of 25°C and not at 37°C, expressed both immunoreactive proteins in a thermal regulated fashion (Fig.

118). Isolates of the non-motile species K. oxytoca, K. pneumoniae, S. f/exneri and Yersinia pestis synthesized neither CheA protein (Fig. 11A).

As with E. coli, I noted that the two immunoreactive products varied somewhat in their relative molecular weight even between isolates of the same species. Interestingly, the apparent differences observed between the two 69

Fig. 10. CheA expression among motile Enterobacteriaceae. lmmunoblot analysis with anti-E. coli CheA antibodies after SOS-PAGE analysis (10% polyacrylamide). (A) Lysates were prepared from E. coli laboratory strains (K-12, RP437, RP1515, AJW536, and AJW1071) and a human derived clinical isolate of E. coli (ATCC25922). Cells were grown with aeration in tryptone broth (TB) (1 % [wt/vol] tryptone, 0.5% [wt/vol] NaCl) at 31°C to an optical density at 600 nm of -0.3 before being harvested. (B) Lysates were prepared from wild-type or ticheZ laboratory strains of E.coli (RP437 and RP1616, respectively), from the wild-type laboratory S. typhimurium strain LT-2, and from motile, chemotactic clinical isolates of S. typhimurium, S. marcescens, E. cloacae, or C. freundii. The first isolate of C. freundii was motile but not chemotactic. Cells were grown as described for panel A. (C) Lysates were prepared from E. coli deleted or wild­ type for cheA (strains AJW1071 and RP437, respectively), from plant pathogens of the genus Erwinia or from clinical isolates of human pathogens of the genus Proteus. Cells of Proteus ssp. were handled as described for panel A, except that differentiated swarmer cells of P. mirabilis (SW) were grown on LB agar media at 37°C until they covered most of the agar surface. Cells of Erwinia ssp. were grown as described in panel A, except that PCG medium (1 % [wt/vol] peptone, 0.1 % [wt/vol) Casamino acids, 0.2% [vol/vol] o-glucose) was substituted for TB. In each panel, the arrows at the left of the blot point to the position of E. coli CheAL and CheA5 proteins and at right of the blot the molecular mass markers are labeled in kilodaltons. 70

t\' A .....~ ! - 97.4

t - 68

-43 B

97.4 CheA __.. ___ _ L ,- ,,,,,,, --- 68 CheA8 ..... --

- 43 c

, , ~ - ' - 68 ·--~~ :::.: - 68

-·~, -43 -43 71 immunoreactive proteins synthesized by E. coli (- 9-kDa), E. cloacae (- 8-kDa) and C. freundii (- 7-kDa) strains was greater than the apparent differences between the same two proteins synthesized by S. typhimurium (- 5.5-kDa) and s. marcescens (- 5.5-kDa) (Fig. 10A, B).

To determine whether coexpression of CheAL and CheA5 was a feature shared among more diverse members of the family Enterobacteriaceae, I performed immunoblot analyses on cell lysates from a variety of plant pathogens belonging to the genus Erwinia and from opportunistic human pathogens of the genus Proteus (Fig. 1OC). Like E. coli strain RP437, P. mirabilis and P. vulgaris co-expressed two immunoreactive bands. Furthermore, differentiated swarmer cells of P. mirabilis harvested from the surface of LB plate media synthesized significantly more amounts of both proteins than did undifferentiated swimmer cells harvested from LB liquid cultures. In contrast, E. amylovora, E. carotovora and E. herbicola appeared to synthesize a single immunoreactive band of about

72 to 74-kDa. Since these cells are chemotactic, it is likely that this band corresponds to CheAL. Cells of the non-motile E. stewartii synthesized no detectable CheA protein.

Based on immunoblot analysis, non-motile cells of K. pneumoniae seemingly do not synthesize any detectible level of CheA. I performed Southern blot analysis with chromosomal DNA derived from K. pneumoniae to confirm this finding at the genetic level. No hybridization was detected between chromosomal DNA of any size from K. pneumoniae and a heterologous DNA 72 A

97.4-

a•~ CheAL

68- ~·~CheA s

43-

B Y. enterocolitica Y. pestis 2s 0 c 37°C 25°C 37°c

-68

-43 Fig. 11. CheA expression in isolates of klebsiella ssp. and Yersinia ssp. lmmunoblot analysis with anti­ E. coli CheA antibodies after SOS-PAGE (10% polyacrylamide). (A) Lysates were prepared from non­ motile clinical isolates of Klebsiella oxytoca, Klebsiella pneumoniae, and laboratory strains of E. coli deleted or wild-type for cheA (RP1616 and RP437, respectively). (B) Lysates were prepared from isolates of entero­ pathogenic Y. enterocolitica and Y. pestis. Cells were grown as described in panel A of Fig. 10, except that cells of Yersiniae ssp. were grown at 25°C or 37°C.

Position of E. coli CheAL and CheA8 are indicated by arrows in panel A. Molecular mass markers are in kilodaltons. 73 probe derived from a portion of the E. coli cheA gene that corresponds to the highly conserved central kinase region of CheA. As expected, this probe hybridized to a chromosomal DNA fragment of the expected size from E. coli

(data not shown).

4.6. CheZ EXPRESSION IN THE ENTEROBACTERIACEAE

Because CheA5 has been shown to interact physically with CheZ, I suspected that CheZ is a signaling protein shared by most, if not all, motile enteric bacteria. To test this hypothesis, I performed immunoblot analysis using affinity purified anti-E. coli CheZ antibodies on lysates of several enteric species.

Like E. coli and S. typhimurium, chemotactic clinical isolates of S. marcescens and E. cloacae synthesized an immunoreactive protein of molecular mass that approximates that of E. coli CheZ (~ 24 kDa). Cells of P. mirabilis, P. vulgaris and the chemotactic Erwinia ssp. also expressed this protein (Fig. 12A, B).

4.7. ANALYSIS OF THE 5' REGION OF ENTERIC cheA GENES

I examined nucleotide sequences of the 5' portion of cheA genes derived from several enteric isolates to identify tandem translation start sites that would direct the synthesis of CheAL and CheA5 . I designed two primers that would amplify a region of chromosomal DNA that contains cheA start(L) and start(S) from both E. coli and S. typhimurium. Since one primer must bind to sequences upstream of cheA, I designed an oligonucleotide (forward primer F1 .1) that 74

A

- 29 Chez ,..

B

29 Chez - lllH

Fig. 12. CheZ expression in motile and non-motile enteric bacteria. lmmunoblot analysis with anti-E. coli CheZ antibodies after SOS-PAGE (15% polyacrylamide). (A) Lysates were prepared from laboratory strains of E.coli wild-type or deleted for cheZ (RP437 and RP1616, respectively), from the wild-type laboratory S. typhimurium strain LT-2, or from motile clinical isolates of S. typhimurium, S. marcescens, and E. cloacae. Cells were grown as described for Fig. 1OA. (B) Lysates prepared from wild­ type E. coli (RP437), from plant pathogens of the genus Erwinia, or from clinical isolates of the human pathogens of the genus Proteus. Cells were grown as described in Fig. 1OC. Note cells of E. stewartii are non­ motile. Molecular mass markers are labeled in kilodaltons to the right of each blot. 75

would prime within the 3' region of motB, the gene located immediately upstream

of cheA in the mocha operon of E. coli and S. typhimurium (Fig. 15A). To

identify the most conserved sequences within this region, I aligned the motB

sequences from E. coli, S. typhimurium, and B. subtilis (93). To ensure that the

amplified products would contain start(S) sequences, I designed a reverse

primer (R1) complementary to a sequence that encodes a portion of the highly

conserved CheA kinase domain located about 550-bp 3' to start(S) from E. coli.

For this purpose, I aligned cheA sequences from E. coli, S. typhimurium, B.

subtilis, Caulobacter crescentus (5), and Pseudomonas aeruginosa (97).

As predicted, these primers amplified a product of about 1.0-kb from

chromosomal DNA isolated from wild-type E. coli cells but not from cells deleted for cheA (strain HCB137). Amplified products of about 1.0-kb were also

obtained from S. typhimurium, S. marcescens, P. mirabilis and one isolate of E.

cloacae chromosomal DNA. In contrast, smaller amplified products, which

ranged from about 0.6-kb to about 0.9-kb, were obtained from E. herbicola, P.

vulgaris, C. freundii and a second isolate of E. cloacae DNA (data not shown).

Subsequently, we cloned the 1.0-kb products obtained from S.

marcescens and one of the E. cloacae isolates as well as a 0.9-kb product from the second E. cloacae isolate into the plasmid pGEM-T for DNA sequence

analysis (Fig. 13). The first 414 nucleotides of cheA from E.coli, S. typhimurium,

E. cloacae and S. marcescens shared considerable identity. E. cloacae and S.

marcescens were most similar (87% identity), whereas E. coli and S. 76

Fig. 13. Comparison of nucleotide sequences spanning the '3' motB-5' cheA region of the mocha operon. E. coli and S. typhimurium sequences have been reported previously (65, 125, 126, 129). AS. typhimurium clinical isolate sequenced within this region was identical to the sequence published for LT-2. E. cloacae sequence was derived from two independent clinical isolates. S. marcescens sequence was derived from two independent PCR products of a single isolate. Putative Shine-Dalgarno sequences (119A) are underlined, and verified or putative translation codons are enclosed in boxes. The UGA that encodes the termination codon of motB is boldfaced and resides five base pairs upstream of the GUG that encodes start(L) of cheA. B ... A A' E.coli . . .gc·gccc cggaaaa.acc t:gaggttgca ccacaggtca gtgttc:ccac aatgccatca gccgaaccg£! 66 B .• t:.yphimurivm . . .-t-tat -ac--c-cg- --gc-c.oii--c --g-c--ca- -c--a------tc----aa- --g------B' .• e.Icaceae . a--att -aa--c-g-- --g--cc-tt: --ttc--ctg cc-----a-- -t:c----e-- ---a-t---- s. marseacens . ·• . .a--att -aa--c-g-- -·,g------e --ttc--ctg oec-----.a-- -t:c- - - -c-- ---a-t----

B' ~ ___.., B. co.l1 ggtg•~a.g~a.gcatgga______.,.. ... ta.taagcga.t t:tttatcaga cattttt.tga tga.agcggac: gaact:.gt:.t:.gg 136 S. typhimuri um - ---t:------t------t--- --gc------B. C'loac-ea:e --·---t------t.:------c------c------c::------t:------s.nwrseecens -----t------·------t------c------a------c------t:------E-coli ct.gacatgga gcagca.ttt.g ctggttttgc agcoggaagc gccaga.tgcc: ga.acaattg.lill atgeoat:.ctt 20.S s. eyphimurium ----t------c------a----g t------gt- e--g--e--t ------c--- -c-----t-- B. c.l oa oeae -c--t------a--c:--- --c-~----g t.---o--·--- a--g---t-- --g--gc-c------______.... B .nMtrsescena -c--t------.a--o------a----g t---t----- a--g--ct- - - -s--src-c-

£.coll tcgsrsrctgcc ca.ctcgatca aaggaggggc aggaactttt ggc:t:tcagcg ttt:tgcagga aaccacgcat 276 S. t:yph.tmur.tum ---c--g--g -----c--t- ----c ..·-c-- g--c-.:.a--c ------eta --c------·------E. c.l oaaeae c--t--9--g --t--c--t:- ----c--a-- c-----9--- --a--t-c-a -cc------g------$. marse.scenB" c--t--g--g --t--t--t- ----c--ci;-- c-----g--- -·-a- -t-c-c -cc------g------

St As-+ c- R. C'O.li ctgatggaaa acctget:::cga tga.agcc:aga cgaggtga~oa.actcaa caccgaca.tt atca.at:ctgt 34G B. t:yphimuri um t-a------g------c-- --c-----a- ----g----- t------c---- B. c.loae'eae t:-a------t-- -a- - --.:lie- - --c------g----- t------c---- S.l!Nitra:escens t-at----91- c------g-- -a.--lt-tc-- --c-----c- ----g----- t------ttc---- .. c• E.coli ------t:tttggaaac: gaaggacat.c ~~aagaac agctc..-gacgc t::.tataaacag t:agcaagagc eggatgocgc 416 S. t:yph.:l'.muriWlt ------a--t--t: -----g------c------a-c --cg------·----·- E. c1oacea'!' ...... c--a--t--t -----g--·------·---""'"" c------agt ---gc------t--c------_,.,...... ______s.marsescens ~----·----- c-ta--t:--t ------t::- c---oc-tgc - ~ -gc- - -a.- -a-----t--

B.col:l. cagcttcg.a.t tatat:.ctgcc aggcct:tgcg: tca.actggca tt:.agaagcga aaggcga.aao gccatcogca 486 s. t:yph i niu:r i um a-~------a ------a -t--gc-a-- c--91:.----g c:-g--g------C!a-gc:-g--g E. c.loaceae ------t--a --c------a -t--90---- c- -g- -c- -g. c-g-----c:------ggt tg-c9-a.--c s. ma.rsescens t:.--t--t:.--- _ _; __ -t--t.- -a--tc-t-- c-----c--c c-c-----c::------ggt:. -t---ttc-c

E.coli 9t:gacrcc9.a.t t:a.agt.gt:gg t.; Lgccaa.a.agl;. gae1ccgcaag 526 s. typhimurium --cgtaga'...a ccgcg-cac- aag-gcgg<:g atc-a-g--- 526 B.aloaceae ---gtt:.-otg ccgoaaaac- ga9-gttgtg --ta--gtt- 52~ s.marsescen8 ---gtg-cc- -tt.t--c::c::cc gaaa////// II/I/I/Ill 51() -....J -....J 78

Fig. 14. Predicted amino acid sequence alignments of the amino-terminal portion of CheA. Comparison of the P1 domain of CheA from E.coli (103), S. typhimurium (126), E. cloacae and S. marcescens. Residues that are not identical in three out of the four species are bold-faced. Slashes represent non­ aligned residues. CheA1 * 1 E.cloacae VSMDISDFYQ TFFDEADELL ADMEQHLLDL VPEAPDSEQL NAIFRAAHSI KGGAGTFGFT S.marsescens VSMDISDFYQ TFFDEADELL ADMEQHLLDL VPEAPDSEQL NAIFRAAHSI KGGAGTFGFT S.typhimurium VSMDISDFYQ TFFDEADELL ADMEQHLLDL VPESPDAEQL NAIFRAAHSI KGGAGTFGFT E.coli VSMDISDFYQ TFFDEADELL ADMEQHLLVL QPEAPDAEQL NAIFRAAHSI KGGAGTFGFS

A St A Ee s s 61 E.cloacae ILQETTHLME NLLDKARRGE MQLNTDIINL FLETKDIMQE QLDAYKSSAE PDAASFEYIC S.marsescens LLQETTHLLE TLLDKSRRGD MQLNTDIIFL FLETIDIMQE QLDSYPCSAE PDAASFDYIC S.typhimurium ILQETTHLME NLLDEARRGE MQLNTDIINL FLETKDIMQE QLDAYKNSEE PDAASFEYIC E.coli VLQETTHLME NLLDEARRGE MQLNTDIINL FLETKDIMQE QLDAYKQSQE PDAASFDYIC

121 E.cloacae NALRQLALEA KGEVAAAVVP //AAKLSVVD S.marsescens QALRQLALEA KGEVSFPVVP FFAPK///// S. typhimurium NALRQLALEA KGETTPAVVE TAALSAAIQE E.coli QALRQLALEA KGETPSAV// ///TRLSVVA 80 marcescens were least similar (75% identity). All four species shared several features associated with translation. These included: (i) the ribosome binding site (RBS) centered 8-bp upstream of the GUG that encodes the translation initiation codon for CheAL in E. coli and S. typhimurium; (ii) the RBS centered 7- bp upstream of the AUG that encodes the translation initiation codon for CheA8 in E. coli (65); (iii) the RBS centered 5-bp upstream of the AUG that is purported to encode the translation initiation codon for CheA8 in S. typhimurium (126); and

(iv) a set of three inverted repeats, two that lie 5' of the CheAL translation initiation codon and one that flanks the CheA8 translation initiation codon reported for E. coli (65). The two overlapping repeats located 5' of the translation initiation codon of CheAL (A/A', B/B') exhibited greater variability between species than did the repeat that overlaps the initiation site for E. coli

CheA8 (C/C').

In E.coli and S. typhimurium, this 414-bp region encodes the phosphorylation domain P1 of CheA . Throughout this domain, the predicted amino acid sequences of CheAL from all four species shared 86% to 94% identity (Fig. 14). Beyond this region, CheAL from these species shared little or no identity or similarity.

4.8. BEHAVIORAL STUDIES WITH STRAINS OF E.coli THAT DIFFER ONLY

IN THEIR ABILITY TO SYNTHESIZE CheA5

To test the hypothesis that CheA8 plays a role in chemotactic responses 81 under some experimental condition not yet examined, I constructed isogenic strains of E. coli that differed only in their ability to express CheA8 . In comparing the chemotactic behavior of each isogenic strain under several experimental conditions, I found that cells lacking CheA8 possessed a competitive advantage compared to those that retained its synthesis with respect to their ability to track and respond to L-serine. Anaerobic conditions enhanced this advantage.

4.8.1. Design and construction of strains containing cheA start(S)

mutations

Mutant cheA alleles designed to eliminate CheA8 synthesis were made by altering the start(S) initiation codon. However, since start(S) also encodes Met-

98 within CheAL, these alterations resulted in changes in the coding sequence for CheAL. Such a constraint introduces two potential variables which could affect chemotactic behavior: (i) loss of CheA8 expression, and (ii) aberrant activity associated with an amino acid replacement at position 98 within CheAL.

To circumvent this constraint, we designed two pairs of start(S) codon mutations that differ only in their ability to encode CheA8 expression (Fig. 158).

While the first pair of alleles cheAM98V(St and cheAM98V(S)- (constructed by altering the AUG start(S) codon to GUG or GUC, respectively) encode methionine-to-valine replacements at position 98 in CheAL; that of the second pair cheAM98L(St and cheAM98L(S)" (constructed by altering the AUG start(S) codon to UUG or CUC, respectively) encode methionine-to-leucine 82 A mocha Operon mRNA Fl:b_ motA motB cheA cheW tar • • • ~1 • • • • • • • Start(S~ ~ CheA8 (557 aa) • • • Start(L) •. CheAL (654 aa) • • • • • 94 96 98 100 I I I I ... ACG AAG GAC ATC dI!i CAA GAA CAG ... S.D. B

Strain cheA mutations Gene product(s)

RP437 ... ACG AAG GAC ATC ATG CAA GAA CAG ... CheAL + CheAs AJW532 ... ACG AAG GAT ATC ATG CAA GAA CAG ... CheAL + CheAs AJW533 ... ACG AAG GAT ATC GTG CAA GAA CAG ... CheALM98V + CheAs AJW534 ... ACG AAG GAT ATC GTC CAA GAA CAG ... CheALM98V AJW535 ... ACG AAG GAT ATC TTG CAA GAA CAG ... CheALM98L + CheAs AJW536 ... ACG AAG GAT ATC CTC CAA GAA CAG ... CheALM98L S.D. EcoRV Fig. 15. Schematic of the mocha operon of E. coli and mutant cheA start(S) alleles used in this study. (A) Start(L) and start(S) represent the translation initiation sites for CheAL and CheA5 , respectively. Start(S) Shine­ Dalgarno sequence is underlined and labeled S.D. The ATG that encodes start(S) is italicized and underlined. The numbers above the nucleotide sequence correspond to codons within the CheAL coding region. (B) Each cheA allele, its expected gene product(s), and the strain that harbors it are listed. The 98th codon of CheAL is boxed. The start(S) S.D. sequence and the EcoRV restriction site, which marks each mutant start(S) allele, are underlined. The C-T mutation that creates the EcoRV site is italicized and underlined. 83 replacements. Consequently, the first member of each pair expresses wild-type

CheA8 , while the second member does not.

4.8.2. CheA expression from cheA start(S) mutants

To ensure that my cheA start(S) constructs were correct, I performed immunoblot analysis on cell lysates of each recombinant strain with affinity­ purified polyclonal anti-CheA antibodies (Fig. 16A). Cells of strain AJW532, like that of the wild-type control, expressed similar ratios of wild-type CheAL to CheA8 suggesting that the genetically silent EcoRV site in cheA has no effect upon the synthesis of either form of CheA. Cells of each of the remaining mutants expressed similar, if not identical, steady state levels of their respective CheAL variant, CheALM98V and CheALM98L, respectively. Notably, both cheAM98V(St and cheAM98L(St mutants expressed significantly lower steady state levels of wild-type CheA8 (Fig. 16A and. B); a result that is consistent with the lower efficiency of translation initiation from the infrequently use start codons,

GUG and UUG. In contrast, both cheAM98V(St and cheAM98L(St mutants failed to express any detectable levels of CheA8 (Fig. 16A and B). Furthermore, no other prominent immunoreactive bands were observed from lysates of any strain besides the ~ 40-kDa cross-reactive band. Collectively, these observations indicated that at the level of cheA expression, each pair of mutant strains differed only in their ability to synthesize CheA8 . 84

A

I I II ._.. ~_. -··-~·-· ~ CheAL ~CheA8

43- --,,.

B ~· ~·

97.4-

68-

Fig. 16. lmmunoblot analysis on recombinant strains of E. coli harboring cheA start(S) alleles. Lysates prepared from aerobically grown cells (A), or anaerobically grown cells (B}, of strains wild-type or deleted for cheA (RP437 and AJW1071, respectively) or those harboring their respective start(S) alleles. Whole cell proteins were transferred to nitrocellulose membranes and probed with anti-E. coli CheA antibodies after SOS-PAGE (10% polyacrylamide). Molecular mass standards in kilodaltons are listed to the left of each blot. 85

4.9. CHEMOTACTIC BEHAVIOR OF cheA start(S) MUTANT STRAINS

Although the EcoRV mutation, used to track mutant cheA alleles during genetic manipulations, had no effect upon the synthesis of either form of CheA, we were concerned that it may exert an effect upon chemotactic behavior.

However, when directly compared, the migration rates of cells of strain AJW532 were identical to that of the isogenic wild-type strain, AJW382 (data not shown), suggesting that the EcoRV mutation has no impact upon chemotactic behavior under these conditions. Thus, strain AJW532 was used as the wild-type control in all subsequent behavioral experiments.

On tryptone swim plates, colonies of each recombinant strain resembled those of the chemotactically wild-type strain AJW532: they typically formed inner and outer bands. Although their band morphologies were similar, cells of the cheAM98V(St mutant migrated at similar rates to those of the cheAM98V(S)­ mutant both of which were approximately 0.43 cm/h (Fig. 17A). Similarly, cells of the cheAM98L(St mutant migrated at identical rates of 0.56 cm/h to those of the cheAM98L(Stmutant. Notably, cells of strains AJW533 and AJW534, which harbored a valine substitution at Met-98, generally had lower migration rates compared to those that had a leucine at Met-98 (strains AJW535 and AJW536)

(Fig. 17A).

To test whether CheA8 participates in chemotactic responses under non­ standard conditions, I performed swim assays under different growth temperatures (27°C to 39°C), or by altering the pH (6.0 to 8.0), osmolarity (0.5% 86

A 4

3

E -(.)

0 0 4 6 8

4 B

3

E -(.) ~ 2 'U (IJ 0:::

0 6 8 10 12 14

Time (h)

Fig. 17. Swimming ability of cells that carry cheA start(S) mutations under (A) aerobic or (B) anaerobic conditions. Plots of 3 separate experiments indicate the time course of displacements of the outermost edges of swim colonies produced by cells that express w.t. CheAL and CheA8 (D; strain AJW532), CheALM98L with or without w.t. CheA8 (+and .t.; strains AJW535 and AJW536, respectively) and CheALM98V with or without w.t. CheA8 (.~and 'V; strains AJW533 and AJW534, respectively). The standard errors of the mean displacements are shown only for those values greater than 0.05 cm. 87 to 3.5% NaCl), or viscosity (0.2%)to 0.45% agar) of tryptone swim plates. Under all conditions tested, however, the migration rates of each isogenic pair were similar, if not, identical (data not shown). Cells of both strains expressing

CheAL M98V, regardless of CheA5 expression, continued to migrate at rates somewhat slower than those of both strains expressing CheAL M98L.

Since all members of the Enterobacteriaceae family are facultative anaerobes, I suspected that CheA5 might play a role in chemotaxis when cells encounter anaerobic environments. On tryptone swim plates that lacked oxygen, cells of both CheAL M98V mutants, regardless of CheA5 expression, continued to exhibit similar, if not, identical migration rates (Fig. 178). These results parallel those of our previous aerobic studies. In contrast, cells of the CheALM98L mutant that express CheA5 migrated at 0.29 cm/h compared to the 0.34 cm/h rate exhibited by those of the CheAL M98L mutant that lack CheA5 , which represents a rate difference of 85% (Fig. 178). Since these differences could reflect an impairment in growth, we examined the growth profile of each mutant in liquid tryptone broth. Under both aerobic and anaerobic conditions, the growth rates of all five strains that carried cheA Start(S) mutations were similar, if not identical, to that of wild-type strain AJW382 (data not shown).

Swim colonies of each recombinant strain grown on tryptone swim plates that lacked oxygen resembled those of the chemotactically wild-type strain

AJW532, however, they typically formed only an outer band, which under aerobic growth conditions, represents cells responding to L-serine. To identify 88

1.

1.0 • 0- ....0 0-- 1,.' .. ""'""° .. .. rl1 .e o.s "q "O ~ ~ "O ' ' ~ -~ 0.6 ' "O ' J,,. ~ a. .. "O ...... ~= 0.4 .. 00 .. .. D • • ~Asp . . . "D 0.2 • {] • Ser

0.0 10-5 104 10-3 10-2 10° Concentration (M)

Fig. 18. Influence of L-serine upon chemotactic ability of cells wild-type for chemotaxis when grown under anaerobic conditions. Plots of two separate experiments indicate the displacement of the outermost edges of swim colonies produced by cells of strain AJW532(cheAEcoRv). Cells, inoculated near the center of tryptone swarm plates containing 0.2% agar and various concen-trations of either L-serine (open squares) or L-aspartate (open circles). Radial displacements were measured after 12 h incubation at 35 °C when grown under anaerobic conditions. Measurements were standardized to the mean displacement observed for swim colonies incubated in the absence of L-serine and L-aspartate. The standard error of the mean displacements for duplicate swim colonies were <0.04 cm. 89

the attractant sensed by cells residing in the outer band of colonies that formed anaerobically, I challenged cells of strain AJW532 to a range of concentrations of

L-serine or L-aspartate. Since a dose-dependent reduction in the size of swim colonies was observed only with increasing amounts of L-serine (Fig. 18), we conclude this band also represents cells responding to L-serine.

4.10. CheA5 COMPLEMENTATION ANALYSIS

Cells of strain AJW533 (allele cheAM98V(St) and AJW535 (allele cheAM98L(St) synthesize significantly lower amounts of CheA8 compared to those of AJW532 (cheAEcoRv) (Fig. 16A, B). Therefore, I may have under­ estimated the effect exerted by CheA8 upon migration rates. To test this possibility, competent cells of AJW534 and AJW536 that do not express CheA8 were transformed with pAR1 .cheA5 , a plasmid that permits control of CheA8 synthesis by means of IPTG induction from tandem tac promoters. Cells of the resultant transformants were spotted onto TB swim plates containing a range of

IPTG concentrations and incubated either under aerobic or anaerobic conditions.

Simultaneously, cells of each transformant were grown aerobically or

anaerobically in TB, harvested at an 00600 -0.3, and subjected to immunoblot analysis.

Under aerobic conditions, cells of the M98L mutant that synthesized

CheA8 migrated as rapidly as those that did not (Fig. 20A) even when the levels of CheA8 exceeded those of CheAL (e.g., those induced by 25 µM IPTG; 90

A ID ID B ff') ff') tr) tr) ,_,~ ,_,~ < +pARl +pARl.ckAs < +pARl +pARl.~S' 0 100 0 25 5 10 25 50 100 0 100 0 2.5 5 10 2550100 [IPTQI 97.4------~_,._ ... Che;\, ----rfl!l/ll/# ~

43-

Fig. 19. CheA5 induction as a function of IPTG inducer concentration. lmmunoblot analysis were performed on whole cell lysates prepared from cells of strain AJW536[cheAM98L(SrJ or those harboring pAR1 .cheA8 or pAR1 after SOS-PAGE (10% polyacrylamide) and probed with anti-CheA antibodies. Cells were grown under aerobic (A) or anaerobic (B) conditions in the absence or presence of IPTG at the concentrations indicated. Molecular mass standards in kilodaltons are to the left of the blot. 91 A 3.o ...... !II" - "' ..,..., ~ ~ >lj: - Ill - '!'ii ~ ,.,, - "' --...... : ~ ,,...... 2.5 ...... ~ .. - ......

--- -·- --- ...... - '"' ,, 2.0 ------... " --E ' "'· u ' '-" 1.5 ' "' "'::s ' :.a ' ' Cl:S ...... ~ 1.0 ...... -cheA98MV+pARJ ...... - - cheA 98MV+ pAR 1. cheAs 0.5 • •· cheA98ML+pARJ -· cheA98ML+pARJ.cheAs 0.0 0 2.5 5 10 25 50 100 B 2.5 ...... " 2.0 ...... "" E ' ""- --u . - '-" 1.5 ' -. -.. _ "'::s :.a . ""' Cl:S ____ .... ______-- ..... _ ~ 1.0 "'x.....,._

Fig. 20. Effect of CheA5 concentration on the swimming ability of cells harboring cheA start(S) mutations. Plots of two separate experiments represent radial displacement of the outermost edges of swim colonies of cells that synthesize either CheAL M98V or CheAL M98L but not CheA5 (strains AJW534 and AJW536, respectively) transformed with the plasmid pAR1 .cheA8 or with the vector control pAR1 as a function of inducer concentration. Radial displacements were measured after 12 h. Tyrptone plates containing 0.2% agar and the designated concentration of IPTG were incubated under aerobic conditions at 31°C (A) or under anaerobic conditions at 35°C (B). The standard error of the mean displacements of quadruplicate swim clonies were less than 0.05 cm. 92

Fig.198). Under anaerobic conditions, however, the migration of cells that

synthesized CheA5 was progressively worse compared to those that did not (Fig.

208) even when the levels of CheA5 approximated those of CheAL (e.g., those

induced by 2.5 µM IPTG; Fig. 198). This inhibitory effect of CheA5 upon swimming migration under anaerobic conditions was not as dramatic in cells of the M98V mutant. For example, cells of the pAR1 .cheA8 transformant when

grown anaerobically in the presence of 10 µM IPTG migrated as rapidly as those that did not (Fig. 208) even when the level of CheA5 synthesis exceeded those

of CheAL (data not shown). Like cells of the M98L mutant when grown

aerobically, those of the M98V mutant that synthesized CheA5 migrated as

rapidly as those that did not (Fig. 20A) even when the levels of CheA5 exceeded those of CheAL (data not shown). The results of these studies and those of single copy expression of CheA5 demonstrate that under anaerobic conditions the inhibitory effect of CheA5 upon swim migration in TB swim plates resides

predominantly in cells of the M98L mutant but not in cells of the M98V mutant.

4.11. COMPETITIVE SWIM ASSAY ANALYSIS

Since CheA5 exerted an effect upon the behavior of free swimming cells that was not detected in swim assays performed under anaerobic conditions, I

attempted to enhance this effect by directly competing cells that synthesize

CheA5 against those that do not. Specifically, I mixed equal numbers of cells

that synthesized both CheALM98L and CheA5 (cheAM98L(St) with those that 93

Table 8. Competitive swim assay analysisa Selectable Mean CFU X106 (±SE)b Strain cheA allele marker Aerobic growth Anaerobic growth AJW1319 cheAM98L(St Thr+ 2 ± 15 3 ± 0.9 AJW1320 cheAM98L(St Leu+ 104 ± 4 339 ± 1 AJW1335 cheAM98L(St Leu+ 38 ± 6 Noc AJW1336 cheAM98L(St Thr+ 82 ± 8 Noc a For each strain, data was compiled from a total of 9 samples recovered from the leading edge of the outermost band of three independent swim colonies. b CFU, colony forming units identified by growth on M63 media lacking either L-leucine or L­ threonine; and SE, standard error of the mean. c ND, not determined. synthesized only CheALM98L (cheAM98L(Sn and spotted these mixtures onto the surface of TB swim plates. By plating out cell populations that were extracted from the leading edge of swim colonies grown aaerobically or anaerobically onto selective plate media, I was able to differentiate and quantify both cell types.

Cells that did not synthesize CheA5 migrated in the leading edge of the L- serine band more so than those cells which synthesized both CheAL M98L and

CheA5 (Table 8). Thus, cells that do not synthesize CheA5 appear to out compete those that do with regards to their ability to track and respond to L- serine. This observation was true regardless of the prototrophic marker used to differentiate the two cell types. It was also true regardless of oxygen concentration: the absence of oxygen appeared to enhance this effect several- fold (Table 8).

The ability of cells to form a migrating band in TB swim plate media involves not only chemotaxis, but also depends on their ability to transport and metabolize chemoattractant compounds, to swim, and their resultant growth. 94

TABLE 9. Tumble frequency and swimming speed determinationsa nb Tumble frequency Swimming speed 1 1 Strain (tumbles s- ; mean± SEt (µM s- ; mean± SEt AJW1319 1800 0.482±0.010 21.1±0.14 AJW1320 1800 0.472 ± 0.010 20.9 ± 0.14 AJW536 900 0.479 ± 0.015 21.4 ± 0.20 AJW535 900 0.469 ± 0.015 21.5 ± 0.18 CP366 600 0.513 ± 0.018 21.1±0.25 •Data were compared statistically by analysis of variance using SAS software package (SAS Institute, Cary, N. C.) General Linear Models (GLM) Procedure as described (Amsler, 1996; Amsler et al., 1993).

Differences in Tumble frequency (F4,5995 = 1.14, P < 0.3367) or Mean swimming speed (F4,5995 = 1.89, P < 0 .1099 are not considered to be significant. h Number of individual cells measured. 0 SE, standard error of the mean.

Thus, these observations could have resulted from some defect in one of these processes. However, this seems unlikely. All four strains grew in TB at virtually identical rates (Fig. 21) and migrated in TB swim plate media at rates similar to those of their respective parents (data not shown). Similarly, cells of each population swam in motility buffer at similar speeds, and in the absence of any attractants or repellents, exhibited similar tumble frequencies (Table 9).

4.12. In vitro phosphorylation of Met-98 CheAL variants

To test whether amino acid replacements made at Met-98 in CheAL exert an effect upon CheAL function, I compared the relative in vitro kinase activity of each CheAL variant to that of wild-type CheAL. His•Tag fusions of CheAL, CheA8 , and each CheAL variant, CheALM98V, CheALM98L, and CheALM981, a variant used by Sanatinia et al. (1995), were first overproduced and purified to near homogeneity as judged by SOS-PAGE, and confirmed as cheA gene products by immunoblotting with anti-CheA antibodies (Fig. 22A, B). 95

A 1.00

0 0

0. 01 .__.___.__._....._.___.___.___,__...__.~__.__~...__. 0 2 4 6 8 10 12 14

B 1.00

0 0

0.01 ----~~_._~~~~~~~~~ 0 2 4 6 8 10 12 14 Time (h)

Fig. 21. Growth characteristics of M98L mutant strains grown in tryptone broth at 31°C. Plots of three independent experiments represent cell densities of strains AJW1319[(StThr+], (o); AJW1320[(S)-Leu+], (.t.); AJW1335[(StLeu+], (•);and AJW1336[(S)Thr+] as function of time after culture initiation. The standard errors of each 00600 are smaller than the size of each symbol. 96

After a 20 min labeling period at room temperature, kinase activity of each variant was determined by measuring the amount of CheA autophosphorylation

in the presence of [y- 32 P]ATP. His•Tag wild-type CheAL, which contains the site of phosphorylation (His-48), was phosphorylated. In contrast, His•Tag CheA5 , which lacks the first 97 N-terminal amino acids and, thus, does not contain His-

48, did not exhibit any detectible level of phosphorylation (data not shown).

Among those tested, His• Tag wild-type CheAL exhibited the highest level of

phosphorylation (Fig. 23). Relative to that of wild-type CheAL, His•Tag variants of CheAL M98L CheAL M98V, and CheAL M981 exhibited phosphorylation levels of

61%, 41%, and 13%, respectively (Fig. 23).

4.13. RELATIVE EXPRESSION LEVELS OF CheAL AND CheA5

My studies indicate that several species of Enterobacteriaceae share certain features of translation that could be involved in synthesis of either cheA gene product (Fig. 13). For example, any one of the three putative stem loop structures in and around start(L) and start(S) may be regulate the translation of

CheAL and CheA5 , respectively. To test this prediction, I examined whether the

relative ratios of CheAL and CheA5 synthesis would be altered as cells experienced changes in metabolic growth. Specifically, I examined synthesis of

CheAL and CheA5 as a function of growth phases and as function of aerobic versus anaerobic growth. Cells of strain AJW532 were grown aerobically or anaerobically in TB and were harvested at several time points that represent 97 A

CheAL~ .. -68 CheA8 • -43

-29

B

-68

-43

-29 Fig. 22. Purified CheA variant proteins isolated from strain BL21 (DE3) transformants harboring CheA expression vectors. CheAL encoded by plasmid pAR1 .cheA; CheA5 encoded by plasmid pAR1 .cheA8 ; CheALM981 encoded by plasmid pBPM11; CheALM98V encoded by plasmid pBPM8; and CheALM98L encoded by plasmid pBPM10. (A) SOS-PAGE (10% polyacrylamide) stained with Coomassie blue. (B) lmmunoblot analysis using anti-CheA antibodies following SOS-PAGE (10% polyacrylamide). Molecular mass markers in kilodaltons are shown to the right in each panel. 98

70 CheALM981 60 - CheALM98V ~ CheALM98L ~ a_ () 50 .....m 0 I- 40

::R0 "'C ()) 30 -~ m E I- 20 z0 10

0

Fig. 23. Relative in vitro phosphorylation activities of His•Tag purified CheAL variants. Phosphorylation levels of each variant, represented as % total CPM, were normalized such that the phosphorylation level of w.t. CheAL was 100. Each level of phosphorylation is the average of three determinations and is shown along with the standard deviation. 99 growth between early-exponential and post-exponential phases.

The steady state levels of CheAL and CheA5 were examined by immunoblot analysis. Seemingly the relative ratios of CheAL:CheA5 appeared to be altered with respect to growth phase. Maximal levels of both CheA proteins appeared to occur during mid-exponential growth (e.g., cells harvested between

4.5 and 5 h growth; Fig. 24A). In contrast, lower steady state levels of CheA5 compared to CheAL were observed at times corresponding to early-exponential and late-exponential growth phases (Fig. 24A). A similar pattern in steady state levels of both CheA proteins were also observed for cells grown under both conditions (Fig. 24A and B). Maximal steady state levels of both CheA proteins appears to occur during mid-exponential growth aerobically between 3.5 and 5 hours post-inoculation and anaerobically between 3.5 and 4.5 hours post­ inoculation. In contrast, cells in early-exponential and late-exponential to stationary growth exhibit higher steady state levels of CheAL relative to those of

CheA5 . 100

A Post-inoculation (hr)

2.0 2.5 2.8 3.1 3.5 4.0 4.5 5 6 7 8 9 10 24 97.4-

68-

43-

B Post-inoculation (hr)

2.0 2.5 2.8 3.1 3.5 4.0 4.5 5 6 7 8 9 10 24 97.4- 68-. -- 43-

Fig. 24. In vivo steady state levels of CheAL and CheA5 as a function of the growth phase of E. coli grown in tryptone broth. lmmunoblot analysis were performed on lysates of cells of strain AJW532[cheAEcoRv1 grown at 35 °C under aerobic (A) or anaerobic (B) conditions and were harvested at indicated times after inoculation. Blots were probed with anti-CheA antibodies after SOS-PAGE (10% polyacrylamide). Molecular mass markers in kilodaltons are indicated to the left of each panel. CHAPTERV

DISCUSSION

In this study, I provide evidence for a physiological role in bacterial chemotaxis for the short form of the CheA histidine kinase, CheA5 . These findings are:

(i) The C-terminal portion of CheA5 , like that of the corresponding region found in CheA._, receives chemosensory information from chemoreceptors.

(ii) All motile, chemotactic members of Enterobacteriaceae that were examined coexpress CheA._, CheA5 , and CheZ. The only exception were those of Erwinia ssp. that appeared to only express CheA._ and CheZ. Furthermore, sequence analysis revealed the identities of two in-frame translation start sites within the 5' portion of cheA loci from S. marcescens and E. cloacae, which resemble those previously identified in E. coli and S. typhimurium.

(iii) Cells that do not synthesize CheA5 out compete those that do express

CheAs with respect to their ability to track and respond to L-serine. This difference is more pronounced under anaerobic conditions.

5.1. CHEMOTAXIS-ASSOCIATED ACTIVITIES EXHIBITED BY CheA5

Although unable to autophosphorylate, CheA5 exhibits kinase activity that

101 102 is consistent with its shared structural domain organization with CheAt_ (Fig. 4).

Recently, Swanson et al. (133) demonstrated that the C-terminal portion of

CheAt_ regulates autophosphorylation in response to environmental signals through CheW and receptor. To determine whether the C-terminal region of

CheAs, like that of CheAt_, can receive chemosensory information for its own regulation, we examined in vivo and in vitro the ability of CheAs to complement

CheAt_ proteins truncated to various extents along their C-terminus. In vivo,

CheA5 restored chemotactic ability to cells that expressed the variant,

CheAt_616K(Am), which lacked a 38 amino acid segment of its C-terminal region, but not to those that expressed the variants CheAt_ 5031(Am) and

CheAt_572G(Am), which lacked progressively larger portions of their C-termini

(Fig. 5). In vitro, CheA5 transphosphorylated the kinase deficient variant

CheAt_ 5031(Am), albeit at significantly lower rates when compared to that of wild­ type CheAt_. In contrast, CheAs did not transphosphorylate the variant

CheAt_572G(Am) despite its capacity to autophosphorylate.

These results are consistent with the hypothesis that the C-terminal portion of CheAs regulates its own activity as a function of its ability to interact with CheW and the receptor. The C-terminal region of CheA, which contains the

Mand C domains, is thought to interact with transducers and CheW (30, 43, 75).

Deletion of these domains renders the kinase activity of CheAt_ insensitive to the presence of transducers and CheW (30). Thus, as part of CheAt_ or CheA5 , this region likely functions in a regulatory capacity. Alternatively, it may also 103 participate in the proper orientation or functioning of associated regulatory regions located elsewhere in CheA.

In vitro, CheAs mediates receptor modulated transphosphorylation of

CheAi_5031(Am). In vivo, however, it does not restore chemotactic ability to cells that express CheAi_5031(Am). These contradictory observations could be explained by the relatively lower initial rates of CheA5-mediated transphosphory­ lation of CheAi_5031(Am) and the concomitant lower production of phospho-CheY seen in vitro as compared to that of wild-type CheAi_. Presumably, the low level of CheAs-mediated transphosphorylation of CheAi_ 5031(Am) in vitro does not support the level of CheY phosphorylation in vivo that is required by cells to carry out chemotactic behavior in tryptone swim plates. Evidence exists that support this hypothesis. Low levels of CheAs-mediated transphosphory-lation of other certain kinase-deficient CheAi_ variants in vitro do not support chemotactic ability to cells that express these variants in vivo (38). Perhaps upon closer examination these cells would exhibit some level of chemotactic behavior as a result of CheAs transphosphorylation activity in vivo. The rotational behavior of flagellar motors, which can easily be measured by tethering cells to a coverslip via their flagella (73), may reflect a near to wild-type rotational bias, i.e., a more

CW bias, by cells that express both CheAi_ 5031(Am) and CheA8 compared to those that only express CheAi_5031(Am).

The lack of restoration of swimming ability by CheA8 in cells that expressed the variant CheAi_572G(Am) likely results from the presence of an 104

inhibitory fragment of the CheAt_ C-terminal regulatory domain. The dominant

negative influence of expressing cheA572G(Am) in cells that also express the

wild-type cheA allele supports this contention (Fig. 5). This may also explain why, in vitro, CheAs was unable to transphosphorylate another variant

CheAt_542L(Am), which also contains a portion of its C-terminal domain. Such

an inhibitory fragment of CheAt_ could either (i) interfere with the interaction of

CheAs with CheW and/or transducers or (ii) prevent the formation of heterodimers involving CheAs and the truncated versions of CheA.

Our results indicate that the dimeric form of CheA plays an integral role in signal transduction in bacterial chemotaxis. Since restoration of swimming ability by CheAs in vivo was observed only in cells that expressed

CheA._616K(Am), a variant which contains all but the last 38 amino acids of

CheA, it indicates that transducer-mediated regulation of CheA kinase activity requires the presence and/or participation of two intact C-terminal regulatory domains within a CheA dimer. This model, which was proposed originally by

Swanson et al. (133), predicts that the two C-terminal domains provided by

CheAs/CheAt_616K(Am) heterodimers are sufficient in transmitting environmental signals through dimeric transducers and CheW. It also predicts that the last 38 amino acids of CheA are dispensable for regulatory function. 105

5.2. COEXPRESSION OF Che~ AND CheA5 BY MEMBERS OF THE ENTEROBACTERIACEAE

All motile, chemotactic members of the Enterobacteriaceae family that

were examined coexpressed CheAi._, CheAs. and CheZ. Several lines of

evidence support this contention. (i) The two prominent immunoreactive

proteins expressed from cells of S. typhimurium, S. marcescens, E. cloacae, C.

freundii, P. mirabilis and other chemotactic isolates were similar in size to E. coli

CheAi._ and CheAs (Figs. 10A, B, C, and 11A and B). (ii) Similarly, the one

prominent immunoreactive band expressed from all motile, chemotactic cells

examined were similar in size to E. coli CheZ (Figs. 12A and B). (iii) Cells of the

non-motile species K. oxytoca, K. pneumoniae, S. flexneri, and Y. pestis

synthesized neither anti-E. coli CheA immunoreactive protein (Figs. 11A and B).

(iii) Analysis of nucleotide sequences of the cheA loci from isolates of S.

marcescens and E. cloacae revealed the presence of in-frame translation

initiation sites similar to those observed in the cheA loci of E. coli and S.

typhimurium (Fig. 13).

Two other interesting observations are also consistent with these findings.

First, cells of Y. enterocolitica grown at 25°C, but not by those grown at 37°C, co-express two immunoreactive proteins of similar size to E. coli CheAi._ and

CheAs. (Fig. 11 B). This observation is reminiscent of the coordinate thermal regulation of three flagellin gene products, which were shown to be synthesized only during growth at 28°C (61 ). Second, differentiated swarmer cells of P. 106

mirabilis appeared to synthesize higher levels of both immunoreactive proteins in

comparison to broth-grown undifferentiated swimmer cells {Fig. 1OC). This

observation is similar to that of flaA gene expression, a gene that encodes the

subunit of the flagellar filament, whose transcription in swarmer cells of P.

mirabilis was found to be eight-fold higher during swarmer cell differentiation

(13). To my knowledge, this was a new observation of coordinate regulation of a

chemotaxis gene product from cells that appear to undergo a dramatic switch in

genetic programs and, thus, was reported recently (89).

In contrast to the observed pattern of cheA gene expression among all

chemotactic bacteria examined, cells of the enteric species E. amylovora, E.

carotovora, and E. herbicola, which are commonly associated with plant

diseases, synthesized a single immunoreactive protein of similar size to E. coli

CheAi._ {Fig. 1 OC). Such findings indicate that cells of these species may lack

the necessary translational elements that would direct CheA5 synthesis. Since I

was unable to amplify and clone the 5' portion of cheA from any one of these

species, this explanation remains pure speculation. Alternatively, this apparent

lack of CheAs expression could be a attributed to the growth conditions used to

cultivate these cells for western blot analysis. Perhaps under some other growth condition{s}, i.e., different media formulation or growth temperature, co­ expression of both CheA proteins would be observed.

The apparent difference in the relative molecular weights between the

CheAi._ and CheAs proteins of E.coli (- 9-kDa) is greater than those of S. 107 typhimurium (- 5.5-kDa) (Fig. 10A). The initiation site proposed for S. typhimurium CheAs (Stock et al., 1988) does not correspond to the initiation site reported for E. coli (Fig. 13) (65, 113). My observations support this hypothesis.

The positioning of start(S) relative to start(L) appears to be unique in each species. Seemingly, S. typhimurium uses the upstream site while E. coli uses the downstream one. On the basis of immunoblot analysis, I conclude that S. marcescens uses the upstream site since the size difference between the same two proteins was - 5.5-kDa. Those of E. cloacae and C. freundii use the downstream since the size difference between these same two proteins was - 8- kDa and - 7-kDa, respectively. Although the apparent lack of a Shine-Dalgarno sequence explains why the downstream site would not initiate translation in S. marcescens, there exists no obvious features that can explain why E. coli and E. cloacae seemingly use the downstream site while S. typhimurium preferentially uses the upstream one.

To reconcile the observation that E.coli does not require CheA5 for chemotaxis under standard laboratory conditions with the fact that it synthesizes

Che,\_ and CheAs in approximately equivalent amounts (Fig. 16) (113, 134),

Sanatinia et al. suggested that CheAs might be an evolutionary relic ( 113).

Although, the argument goes that CheA5 no longer possesses a physiological role, if continues to be expressed because of constraints placed on the nucleotide sequence by the amino acid sequence in and around Met-98 of

Che,\_. In light of my observation that enteric bacteria utilize two different 108 internal sequences to initiate translation of CheA8 , it seems unlikely that such constraints exist.

As in E. coli and S. typhimurium, the two CheA proteins and CheZ protein varied in size among the several species of Enterobacteriaceae examined.

Based on my analysis of the corresponding N-terminal amino acid sequence of

CheAt_, I do not find these observations surprising (Fig. 14). For example, both forms of CheA from S. typhimurium are known to contain seventeen more amino acids than do their E. coli homologs; five that reside in L 1, a proposed linker region of CheA that connects the domains P1 and P2, and twelve located within

L2, a linker region that connects P1 with the central transmitter domain (95).

Comparison of the predicted amino acid sequences that comprise the entire P1 domain of CheAt_ from E. coli, S. typhimurium, E. cloacae, and S. marcescens, indicates that this region shares 86% to 94% identity. Beyond, this region, however, they share little or no identity or similarity (Fig. 14). Based on the analysis of this region, I contend that the observed differences in molecular mass of enteric CheA proteins results primarily from variability in the length of their respective domain linkers L 1 and L2.

In addition to the in-frame translation initiation sites that are shared by several enteric species, three inverted repeats that flank or overlap start(L) and start(S) were also found to be shared by E. coli, S. typhimurium, E. cloacae, and

S. marcescens (Fig. 13). These inverted repeats, which are predicted to form secondary stem-loop structures at the mRNA level, have been implicated in 109

modulating the relative levels of CheAi_ and CheA5 (65). Consistent with this

prediction is the report that disruption of the CIC' hairpin, which would occlude

start(S), results in a seven-fold increase in CheAs synthesis (113). Presumably the BIB' stem-loop structure, which would occlude start (L}, serves a similar

purpose in regulating the translation of CheAi_.

5.3. PUTATIVE PHYSIOLOGICAL ROLE OF CheA5 IN CHEMOTAXIS

To determine the physiological role of CheA5 in chemotaxis, I constructed two pairs of strains that differed only in their ability to synthesize CheA5 . To do so, I altered the ATG that encodes the start(S) initiation codon. Because this codon encodes Met-98 of CheAi_, these altered codons resulted in the synthesis of two CheA variants, CheAi_ M98V and CheAi_ M98L. The design of these cheA alleles, however, allowed me to normalize for the effects of amino acid replacements made at Met-98. Cells of each of the resultant mutant (strains

AJW534 and AJW536) that made no detectible CheA5 , nevertheless, migrated at rates that were indistinguishable from cells of each respective strain that synthesized both forms of CheA (strains AJW533 and AJW535; Fig. 17A).

These initial observations are similar to those made by Sanatinia et al. (113), who demonstrated that cells which only expressed a CheAi_M981 mutant migrated at 80% of the rate as compared to wild-type cells. Collectively, these results initially suggest that CheA5 is not required for chemotaxis.

Since these studies were conducted under only one set of laboratory 110 conditions, I hypothesized that CheAs may be involved in tactic responses under conditions of extremes in osmolarity, pH, temperature, or the viscosity of

TB swim media. However, no differences in the relative migration rates of each respective member of both isogenic pairs were observed. In addition, no differences in swarming motility behavior, which represents a specialized form of surface translocation on solid plate media (49), nor in the formation of symmetrical patterns by motile cells of E.coli (32), were observed.

Because CheAs expression appears to be limited to chemotactic members of the Enterobacteriaceae family that inhabit the lower gastrointestinal tract of vertebrates, I suspected that it may play a role in chemotactic signaling under anaerobic conditions. Surprisingly, that role seems to be inhibitory. Under anaerobic conditions, cells that synthesize only CheAt_ M98L migrated more rapidly than do cells that synthesize both CheAt_M98L and CheAs (Fig. 178).

The observed differences in migration rates between these two strains, although statistically significant, represents only a 10% difference. When cells of each strain were placed together on a swim plate, cells that synthesize only

CheAt_M98L out competed those that synthesize both CheA proteins by an approximately ten-fold margin (Table 8). Furthermore, a similar two-to-five fold difference was observed under aerobic conditions (Table 8). Since no differences in growth (Fig. 21 ), motility, and tumble frequency exist between both cell types (Table 9), this competitive advantage by cells lacking CheA5 synthesis evidently represents a significant difference in behavior. 111

These results seemingly present us with a perplexing question: why do cells of E. coli, as well as those of other species of Enterobacteriaceae, synthesize significant amounts of a protein whose presence appears to leave cells at a disadvantage. Perhaps the micro environment of a standard tryptone swim plate does not accurately mimic an ecological niche(s) encountered by E. coli and other members of the Enterobacteriaceae family. One hypothesis is that

CheAs may provide some adaptive value to enteric bacteria when they encounter the predominantly anaerobic environment of the lower GI tract.

To appreciate the challenges encountered by enteric bacteria as they traverse through the alimentary tract of vertebrates one must remember that they first encounter extreme changes in temperature and pH. They must also cope with high concentrations of noxious chemicals such as bile salts, compounds that are known to act as chemorepellents, and many digestive enzymes upon entering the predominantly alkaline environment of the duodenum. E. coli must also compete with a myriad of resident microorganisms for available nutrients.

The terminal ileum of the small intestine appears to be a transitional zone between the sparsely populated or sterile upper small intestine, which includes the jejunum and proximal ileum, and the heavily populated colon where anaerobes outnumber aerobes by a factor of 100-1000 (53). Whereas the upper small intestine contains both aerobic and anaerobic organisms, the colon contains predominantly anaerobic bacteria (116). Clearly these two compartments of the alimentary tract alone represent two vastly different 112 ecological niches.

Several lines of evidence are consistent with this model: (i) motile, chemotactic cells of enteric species are common inhabitants or are enteropathogenic in the lower GI tract of humans and other vertebrates co­ express both forms of CheA; (ii) motile species of Erwinia that rarely are associated with the lower GI tract (117) synthesize little, if any, CheA8 ; and (iii) no one has yet reported coexpression of CheA_ and CheAs from motile species of non-enteric bacteria. Based upon an analysis of cloned sequences of cheA genes derived from B. subtilis (41), Rhizobium meliloti (44), Rhodobacter sphaeroides (135), Halobacterium salinarium (108), Usteria monocytogenes

(36), Caulobacter crescentus (5) and Pseudomonas aeruginosa (97), there appears to be only a single translation initiation site that predictably would direct the synthesis of only one cheA gene product. In addition, only one cheA gene product has been shown to be synthesized in motile cells of B. subtilis and H. salinarium (41, 108).

Alternatively, CheAs may play some role in chemotaxis when cells encounter free-living aquatic or terrestrial environments. One possibility is that

CheAs may participate in tactic responses to changes in oxygen and/or PTS sugar concentration. The chemotactic signaling pathway, which includes CheA,

CheW, and CheY, has recently been implicated in integrating sensory information from the signaling pathways that mediate responses to changes in these environmental conditions (107). 113

5.4. PUTATIVE SIGNALING CHECKPOINTS AFFECTED BY CheA5

CheAs has been shown to interact productively with CheAt_ and, therefore,

coul_d function to enhance or diminish CheAt_ autokinase activity. As an inhibitor,

CheAs could influence the signaling efficiency of chemoreceptor complexes by

titrating out the number of available sites for CheA phosphorylation thereby

lowering the production of phospho-CheY. Based on one current model of

chemoreceptor signaling proposed by Bob Bourret (27), a given ligand binding

site can transduce signals through a single pre-defined cytoplasmic domain.

With equal probability, the mutually exclusive binding of ligand to transducers would turn off the kinase activity of its associated CheA protomer while the other would be unaffected. Ternary complexes that contain CheAt_ homodimers, would be capable of inactivating only one of the two CheA protomers regardless of the orientation of ligand binding and, thus, a reduction in phospho-CheY production would result. Two scenarios exist, however, for complexes that contain

CheAt_/CheAs heterodimers. Reduced phospho-CheY production only occurs if the orientation of ligand binding results in the inactivation of the CheA8 protomer since trans-phosphorylation of the CheAt_ protomer would not occur.

Alternatively, phospho-CheY production would be unaffected if the orientation of ligand binding results in the inactivation of the CheAt_ protomer since its trans­ phosphorylation by the CheAs protomer would proceed unabated. Thus, Iigand binding affects the activity of CheAt_/CheA5 receptor complexes only half of the time compared to its affecting CheAt./CheAt. receptor complexes al I of the time. 114

As a result, cells that express only CheAt_ would be more sensitive responders to its environment. Perhaps some micro environment(s) within the lower GI tract is more accommodating to cells that are capable of regulating the sensitivity of its chemoreceptors.

Alternatively, CheAs could enhance CheAt_ autokinase activity by stabilizing CheAt_. Several examples of in-frame overlapping gene products illustrate this point. The two subunits of the aspartokinase II enzyme of Bacillus subtilis are translated from in-frame overlapping coding sequences (33).

Although the smaller J3 subunit is not essential for the catalytic activity of the larger a. subunit, its association with J3 imparts greater recovery of activity in vitro presumably through its stabilizing effect (94). Recently, Hwang et al. (59) demonstrated that the smaller of the two overlapping gene products CvaA* stabilizes the larger gene product CvaA that encodes the antibacterial peptide toxin Colicin V to enhance its secretion into the medium. The stability of either

CheAt_M98Vor CheAt_M98L, however, appeared to be unaffected by the presence of CheA5 as estimated by immunoblot analysis (Fig. 16).

The relative positions of start(L) and start{S) and their predicted secondary structures suggest that changes in the physiological state of the cell might function to modulate the synthesis of CheA5 relative to CheAt_. Swanson et al. {133) proposed that CheA5 could influence the distribution of available

CheA phosphorylation sites across a finite population of receptor molecules. At low levels of CheA5 , CheAt_ homodimers would predominate where each 115 receptor effectively controls two phosphorylation sites. Conversely, at high levels of CheAs, all CheAt_ molecules would be associated as heterodimers.

Resultant receptors would control one CheA phosphorylation site, whereas those containing CheAs homodimers would be silent. Intermediate CheAs levels could therefore maximize the number of receptors linked to the control of a single CheA phosphorylation site. Such regulation could influence the sensitivity and/or other signaling characteristics of the cell.

Based on these findings, I hypothesized that these structures would modulate the relative levels of CheAt_ and CheAs synthesis as cells undergo a change in either growth rate or physiological state. To test this hypothesis I examined the relative ratio of CheAt_ and CheAs synthesis in cells grown aerobically versus those grown anaerobically. Although their ratios varied as cells proceeded through their growth phase, a similar pattern of CheA synthesis was observed under either growth condition (Fig. 24). These differences could yet reflect a regulatory role for CheAs in chemotaxis as cells undergo changes in their state of growth.

Lastly, CheAs may also influence the signaling activities of CheZ, a downstream signaling component of the chemotactic signal transduction pathway (Fig. 2). Several observations support this contention: (i) CheAs, but not CheAt_, interacts with and increases the activity of CheZ (88, 134); (ii) the

CCW bias in flagellar rotation that is associated with the overproduction of

CheAs is dependent upon the presence of CheZ (134); and (iii) most, if not all, 116

Enterobacteriaceae synthesize CheZ (Fig. 12). With the exception of P. aeruginosa (87), no CheZ homologs have been reported outside the

Enterobacteriaceae family.

These observations are consistent with a role for CheA5 in regulating

CheZ function in vivo. CheZ, a protein whose function is to accelerate the dephosphorylation of phospho-CheY, has recently been implicated in mediating adaptation responses in E. coli (37). Depending on its oligomeric state, CheZ was found to reduce the level of phospho-CheY which in turn increased the probability of CCW flagellar rotation (23). This interaction of CheAs to CheZ may represent an additional feature in regulating the phosphatase activity of

CheZ.

Wang and Matsumura ( 134) recently demonstrated that a reduced form of

CheA5 is capable of binding CheZ in vitro. It appears that, in the oxidized state,

CheAs forms intermolecular disulfide bonds that correlates with its inability to bind CheZ. Could CheAs act as a redox sensor of the chemotactic signal transduction pathway? If so, it could explain why the behavioral differences seen between cells that only synthesize CheAt_M98L and those that synthesized both CheAt_M98L and CheA5 were more pronounced under anaerobic conditions.

The most reasonable place to begin such an investigation would be to look at the effects of mutagenizing any one or combination of the four naturally occurring cysteine residues found in CheAs upon chemotactic ability under aerobic and anaerobic conditions. APPENDIX A

List of common abbreviations used in this study

CFU ...... colony forming units Che+ ...... chemotactic cm ...... centimeter( s) CPM ...... counts per minute CW ...... clockwise flagellar rotation CCW ...... counterclockwise flagellar rotation GBP ...... galactose binding protein h ...... hour(s) IPTG ...... isopropyl-~-o-thiogalactopyranoside kDa ...... kilodalton(s) LB ...... Luria-Bertani broth MCP ...... methyl-accepting chemotaxis protein or chemoreceptor min ...... minute(s) Mot+ ...... motility OD ...... optical density ON ...... overnight RBS ...... ribosome binding site s ...... second( s) S.D...... Shine-Dalgarno nucleotide sequence SOS ...... sodium dodecyl sulfate start(L) . . . . . cheA translation initiation site that directs the synthesis of CheAt_

start(S) . . . . . cheA translation initiation site that directs the synthesis of CheA5 Tap ...... chemoreceptor proteins which recognize various dipeptides Tar ...... chemoreceptor proteins which recognize L-aspartate and maltose Tsr ...... chemoreceptor proteins which recognize L-serine TB ...... Tryptone broth Trg ...... chemoreceptor proteins which recognize ribose and galactose w.t...... wild-type

117 WORKS CITED

1. Adler, J. 1966. Chemotaxis in bacteria. Science 153:708-716.

2. Adler, J., and B. Templeton. 1967. The effect of environmental conditions on the motility of . Journal of General Microbiology 46:175-184.

3. Adler, J. 1969. Chemoreceptors in bacteria. Science 166:1588-1597.

4. Alex, L. A., and M. I. Simon. 1994. Protein histidine kinases and signal transduction in prokaryotes and eukaryotes. Trends in Genetics 10:133- 138.

5. Alley, D. 1995. Personal communication.

6. Amsler, C. D., M. Cho, and P. Matsumura. 1993. Multiple factors underlying the maximum motility of Escherichia coli as cultures enter post­ exponential growth. Journal of Bacteriology 175:6238-6244.

7. Amsler, C. D. 1996. Use of computer-assisted motion analysis for quantitative measurements of swimming behavior in peritrichously flagellated bacteria. Analytical Biochemistry 235:20-25.

8. Armitage, J.P. 1992. Behaviora~ responses in bacteria. Annual Review of Physiology 54:683-714.

9. Armstrong, J. B., and J. Adler. 1969. Complementation of nonchemotactic mutants of Escherichia coli. Genetics 61 :61-66.

10. Armstrong, J. B., J. Adler, and M. M. Dahl. 1969. Location of genes for motility and chemotaxis on the Escherichia coli genetic map. Journal of Bacteriology 97: 156-161.

11. Barak, R., and M. Eisenbach. 1992. Correlation between phosphorylation of the chemotaxis protein CheY and its activity at the flagellar motor. Biochemistry 31:1821-1826.

118 119

12. Belas, R., D. Erskine, and D. Flaherty. 1991. Proteus mirabifis mutants defective in swarmer cell differentiation and multicellular behavior. Journal of Bacteriology 173:6279-6288.

13. Belas, R. 1994. Expression of multiple flagellin-encoding genes of Proteus mirabilis. Journal of Bacteriology 176:7169-7181.

14. Berg, H. C., and D. A. Brown. 1972. Chemotaxis in Escherichia coli analyzed by three-dimensional tracking. Nature (London) 239:500-504.

15. Berg, H. C. 1975. Bacterial behavior. Nature (London) 254:389-392.

16. Berg, H. C. 1988. A physicist looks at bacterial chemotaxis. Cold Spring Harbor Symposium Quantitative Biology 53:1-9.

17. Berg, H. C. and L. Turner. 1995. Cells of Escherichia coli swim either end forward. Proceedings of the National Academy of Sciences U.S.A. 92:477-479.

18. Blair, D. F., and H. C. Berg. 1988. Restoration of torque in defective flagellar motors. Science 242:1678-1681.

19. Blair, D. F., and H. C. Berg. 1990. The MotA protein of E. coli is a proton­ conducting component of the flagellar motor. Cell 60:439-449.

20. Blat, Y., and M. Eisenbach. 1994. Phosphorylation-dependent binding of the chemotaxis signal molecule CheY to its phosphatase, CheZ. Biochemistry 33:9902-906.

21. Blat, Y. and M. Eisenbach. 1996. Conserved C-terminus of the phosphatase CheZ is a binding domain for the chemotactic response regulator CheY. Biochemistry 33:215-226.

22. Blat, Y., and M. Eisenbach. 1996. Mutants with defective phosphatase activity show no phosphorylation-dependent oligomerization of the phosphatase Chez of bacterial chemotaxis. The Journal of Biological Chemistry 271: 1232-1236.

23. Blat, Y., and M. Eisenbach. 1996. Oligomerization of the phosphatase CheZ upon interaction with the phosphorylated form of Che Y, a signal transduction protein involved in bacterial chemotaxis. The Journal of Biological Chemistry 271: 1226-1231. 120

24. Block, S. M., J. E. Segall, and H. C. Berg. 1982. Impulse responses in bacterial chemotaxis. Cell 31:215-216.

25. Borkovich, K. A., N. Kaplan, J. F. Hess, and M. I. Simon. 1989. Transmembrane signal transduction in bacterial chemotaxis involves ligand-dependent activation of phosphate group transfer. Proceedings of the National Academy of Sciences U.S.A. 86:1208-1212.

26. Borkovich, K. A., and M. I. Simon. 1990. The dynamics of protein phosphorylation in bacterial chemotaxis. Cell 63:1339-1348.

27. Bourret, R. B. 1995. Personal communication.

28. Bourret, R. B., J. F. Hess, K. A. Borkovich, A. A. Pakula, and M. I. Simon. 1989. Protein phosphorylation in chemotaxis and two-component regulatory systems of bacteria. The Journal of Biological Chemistry 264:7085-7088.

29. Bourret, R. B., K. A. Borkovich, and M. I. Simon. 1991. Signal transduction pathways involving protein phosphorylation in prokaryotes. Annual Review of Biochemistry 60:401-441.

30. Bourret, R.B., J. Davagnino, and M.I. Simon. 1993. The carboxy­ terminal portion of the CheA kinase mediates regulation of autophosphorylation by transducer and CheW. Journal of Bacteriology 175:2097-2101.

31. Boyd, A., A. Krikos, and M. I. Simon. 1981. Sensory transducers of E. coli are encoded by homologous genes. Cell 26:333-343.

32. Budrene, E. 0., and H. C. Berg. 1991. Complex patterns formed by motile cells of Escherichia coli. Nature (London) 349:630-633.

33. Chen, N.-Y., and H. Paulus. 1988. Mechanism of expression of the overlapping genes of Bacillus subtilis aspartokinase II. The Journal of Biological Chemistry 263:9526-9532.

34. Dean, G. E., R. M. Macnab, J. Stader, P. Matsumura, and C. Burks. 1984. Gene sequence and predicted amino acid sequence of the motA protein, a membrane-associated protein required for flagellar rotation in Escherichia coli. Journal of Bacteriology 159:991-999.

35. Delucia, P, and J. Cairns. 1969. Isolation of an E.coli strain with a 121

mutation affecting DNA polymerase. Nature (London) 224:1164-1166.

36. Dons, L., J.E. Olsen, 0. F. Rasmussen. 1994. Characterization of two putative Usteria monocytogenes genes encoding polypeptides homologous to the sensor protein CheA and the response regulator CheY of chemotaxis. DNA Sequences 4:301-311.

37. Eisenbach, M. 1996. Control of bacterial chemotaxis. Molecular Microbiology 20:903-910.

38. Ellefson, D. D., U. Weber, and A. J. Wolfe. 1997. Genetic analysis of the catalytic domain of the chemotaxis-associated histidine kinase CheA Journal of Bacteriology 179:825-830.

39. Falke, J. J., A. F. Demberg, D. A. Sternberg, N. Zalkin, D. L. Milligan, and D. E. Koshland, Jr. 1988. Structure of a bacterial sensory receptor. The Journal of Biological Chemistry 263:14850-14858.

40. Finegold, S. M., V. L. Sutter, and G. E. Mathisen. 1983. Normal indigenous intestinal flora p. 3-31. In: Hentges, D. J. Ed. Human intestinal microflora in health and disease. Academic Press, New York, N.Y.

41. Fuhrer, D. K., and G. W. Ordal. 1991. Bacillus subtilis CheN, a homolog of CheA, the central regulator of chemotaxis in Escherichia coli. Journal of Bacteriology 173:7443-7448.

42. Gegner J. A., and F. W. Dahlquist. 1991. Signal transduction in bacteria: CheW forms a reversible complex with the protein kinase CheA. Proceedings of the National Academy of Sciences U.S.A. 88:750-754.

43. Gegner J.A., D.R. Graham, A. F. Roth, and F. W. Dahlquist. 1992. Assembly of an MCP receptor, CheW, and kinase CheA complex in the bacterial chemotaxis signal transduction pathway. Cell 18:975-982.

44. Greek, M., J. Platzer, V. Sourjik, and R. Schmitt. 1995. Analysis of a chemotaxis operon in Rhizobium meliloti. Molecular Microbiology 15:989- 1000.

45. Green, G.N., and R. B. Gennis. 1983. Isolation and characterization of an Escherichia coli mutant lacking cytochrome d terminal oxidase. Journal of Bacteriology 154:1269-1275.

46. Gutterson, N. I., and D. E. Koshland, Jr. 1983. Replacement and 122

amplification of bacterial genes with sequences altered in vitro. Proceedings of the National Academy of Sciences U.S.A. 80:4894-4898.

47. Harlow, E., and D. Lane. 1988. Immunizations p. 92-119. In Antibodies: a Laboratory Manual. Cold Spring Harbor Laboratories, Cold Spring Harbor, N.Y.

48. Harlow, E., and D. Lane. 1988. lmmunoassays p. 563-566. In Antibodies: a Laboratory Manual. Cold Spring Harbor Laboratories, Cold Spring Harbor, N.Y.

49. Harshey, R. M., and T. Matsuyama. 1994. Dimorphic transition in Escherichia coli and Salmonella typhimurium: surface induced differentiation into hyperflagellated swarmer cells. Proceedings of the National Academy of Sciences U.S.A. 91:8631-8635.

50. Hazelbauer, G. L., and J. Adler. 1971. Role of the galactose binding protein in chemotaxis of Escherichia coli toward galactose. Nature (London) New Biologist 230:101-104.

51. Hazelbauer, G. L., R. Yaghmai, G. G. Burrows, J. W. Baumgartner, D. P. Dutton, and D. G. Morgan. 1990. In biology of the chemotactic response (Society of General Microbiology Symposium, No. 46) Matsumura, P., S. Roman, K. Voltz, and D. McNally, eds, pp. 135-154. Cambridge University Press, Cambridge, Mass.

52. Hedblom, M. L., and J. Adler. 1983. Chemotactic response of Escherichia coli to chemically synthesized amino acids. Journal of Bacteriology 155: 1463-1466.

53. Hentges, D. J. 1993. The anaerobic microflora of the human body. Clinical Infectious Diseases. 16(Suppl 4):S175-S180.

54. Hess, J. F., K. Oosawa, P. Matsumura, and M. I. Simon. 1987. Protein phosphorylation is involved in bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 84:7609-7613.

55. Hess, J. F., R. B. Bourret, and M. I. Simon. 1988. Histidine phosphorylation and phosphoryl group transfer in bacterial chemotaxis. Nature (Lendon) 336:139-143.

56. Hess, J. F., K. Oosawa, N. Kaplan, and M. I. Simon. 1988. Phosphorylation of three proteins in the signaling pathway of bacterial 123

chemotaxis. Cell 53:79-87.

57. Hess, J. F., R. B. Bourret, and M. I. Simon. 1991. Phosphorylation assays for proteins of the two-component regulatory system controlling chemotaxis in Escherichia coli. Methods in Enzymology 200: 188-204.

58. Hoffmann, A., and R. G. Roeder. 1991. Purification of his-tagged proteins in non-denaturing conditions suggests a convenient method for protein interaction studies. Nucleic Acids Research 19:6337-6338.

59. Hwang, J, M. Manuvakhova, and P. C. Tai. 1997. Characterization of in­ frame proteins encoded by cvaA, an essential gene in the Colicin V secretion system: CvaA* stabilizes CvaA to enhance secretion. Journal of Bacteriology 179:689-696.

60. lsh-Horowicz, D., and J. F. Burke. 1981. Rapid and efficient cosmid cloning. Nucleic Acids Research 9:2989-2998.

61. Kapatral, V., and S. A. Minnich. 1995. Co-ordinate, temperature­ sensitive regulation of the three Yersnia enterocolitica flagellin genes. Molecular Microbiology 17:49-56.

62. Kehry, M. R., and F. W. Dahlquist. 1982. The methyl-accepting chemotaxis proteins of Escherichia coli. Identification of the multiple methylation sites on methyl-accepting chemotaxis proteins. The Journal of Biological Chemistry 257: 10378-10386.

63. Kehry, M. R., M. W. Bond, M. W. Hunkapiller, and F. W. Dahlquist. 1983. Enzymatic deamination of methyl-accepting chemotaxis proteins in Escherichia coli catalyzed by the cheB gene product. Proceedings of the National Academy of Sciences U.S.A. 80:3599-3603.

64. Kihara, M., and R. M. Macnab. 1981. Cytoplasmic pH mediates pH taxis and weak-acid repellent taxis of bacteria. Journal of Bacteriology 145:1209-1221.

65. Kofoid, E. C., and J. S. Parkinson. 1991. Tandem translational starts in the cheA locus of Escherichia coli. Journal of Bacteriology 173:2116- 2119.

66. Kort, E. N., M. F. Goy, S. H. Larsen, and J. Adler. 1975. Methylation of a membrane protein involved in bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 72:3939-3943. 124

67. Krikos, A., N. Mutoh, A. Boyd, and M. I. Simon. 1983. Sensory transducers of E. coli are composed of discrete structural and functional domains. Cell 33:615-622.

68. Krikos, A., M. P. Conley, A. Boyd, H. C. Berg, and M. I. Simon. 1985. Chimeric chemosensory transducers of Escherichia coli. Proceedings of the National Academy of Sciences U.S.A. 82:1326-1330.

69. Laemmli, U. K. 1970. Cleavage of structural proteins during assembly of the head of bacteriophage T4. Nature (London) 227:680-685.

70. Larsen, S. H., R. W. Reader, E. N. Kort, W.-W. Tso, and J. Adler. 197 4. Change in direction of flagellar rotation is the basis of the chemotactic response in Escherichia coli. Nature (London) 249:74-77.

71. Lengeler, J. W., A.-M. Auburger, R. Mayer, and A. Pecher. 1981. The phosphoenolpyruvate-dependent carbohydrate: phosphotransferase system enzymes II as chemoreceptors in chemotaxis of Escherichia coli K12. Molecular General Genetics 183:163-170.

72. Lin, L.-N., J. Li, J. F. Brandts, and R. M. Weis. 1994. The serine receptor of bacterial chemotaxis exhibits half-site saturation for serine binding. Biochemistry 33:6564-6570.

73. Liu, J., and J. S. Parkinson. 1989. Role of CheW protein in coupling membrane receptors to the intracellular signaling system of bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 86:8703-8707.

74. Liu, J. 1990. Molecular genetics of the chemotactic signaling pathway in Escherichia coli. Ph.D. thesis. University of Utah, Salt Lake City.

75. Liu, J., and J. S. Parkinson. 1991. Genetic evidence for interaction between the CheW and Tsr proteins during chemoreceptor signaling by Escherichia coli. Journal of Bacteriology 173:4941-4951.

76. Long, D. G., and R. M. Weis. 1992. Oligomerization of the cytoplasmic fragment from the aspartate receptor of Escherichia coli. Biochemistry 31 :9904-9911.

77. Lowe, G., M. Meister, and H. C. Berg. 1987. Rapid rotation offlagellar bundles in swimming bacteria. Nature (London) 325:637-640. 125

78. Lupas, A., and J. Stock. 1989. Phosphorylation of an N-terminal regulatory domain activates the CheB methyltransferase in bacterial chemotaxis. The Journal of Biological Chemistry. 264:17337-17342.

79. Lux, R., K. Jahreis, K. Bettenbrock, J. S. Parkinson, and J. W. Lengeler. 1995. Coupling of the phosphotransferase system and the methyl-accepting chemotaxis protein-dependent chemotaxis signaling pathways of Escherichia coli. Proceedings of the National Academy of Sciences U.S.A. 92:11583-11587.

80. Lynch, B. A., and D. E. Koshland, Jr. 1992. Similarities between the aspartate receptor and the trp repressor of E. coli. Implications for transmembrane signaling. FEBS Letters 307:3-9.

81. Macnab, R. M., and D. E. Koshland. 1972. The gradient-sensing mechanism in bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 69:2509-2512.

82. Maddock, J. R., and L. Shapiro. 1993. Polar localization of the chemoreceptor complex in the Escherichia coli cell. Science 259:1717- 1723.

83. Maeda, K., Y. lmae, J.-1. Shioi, and F. Oosawa. 1976. Effect of temperature on motility and chemotaxis of Escherichia coli. Journal of Bacteriology 127:1039-1046.

84. Maeda, K., and Y. lmae. 1979. Thermosensory transduction in Escherichia coli: inhibition of the thermoresponse by L-serine. Proceedings of the National Academy of Sciences U.S.A. 76:91-95.

85. Maloy, S. R., and W. D. Nunn. 1981. Selection for loss of tetracycline resistance by Escherichia coli. Journal of Bacteriology 145:1110-1112.

86. Manson, M. D. 1992. Bacterial motility and chemotaxis. Advances in Microbial Physiology 33:277-344.

87. Masduki, A., J. Nakamura, T. Ohga, R. Umezaki, J. Kato, H. Ohtake. 1995. Isolation and characterization of chemotaxis mutants and genes of Pseudomonas aeruginosa. Journal of Bacteriology 177:948-952.

88. McNally, D. F, and P. Matsumura. 1991. Bacterial chemotaxis signaling complexes: formation of a CheA/CheW complex enhances autophosphorylation and affinity for CheY. Proceedings of the National 126

Academy of Sciences U.S.A. 88:6269-6273.

89. McNamara, B. P., and A. J. Wolfe. 1997. Coexpression of the long and short forms of CheA, the chemotaxis histidine kinase, by members of the family Enterobacteriaceae. Journal of Bacteriology 179:1813-1818.

90. Milburn, M. V., G. G. Prive, D. L. Milligan, W. G. Scott, J. Yeh, J. Jancarlk, D. E. Koshland, Jr., and S.-H. Kim. 1991. Three-dimensional structures of the ligand-binding domain of the bacterial aspartate receptor with and without ligand. Science 254:1342-1347.

91. Miller, S. A., D. D. Dykes, and H. F. Polesky. 1988. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Research 16:1215.

92. Milligan, D. L., and D. E. Koshland, Jr. 1988. Site-directed cross-linking. Establishment of the dimeric structure of the aspartate receptor of bacterial chemotaxis. The Journal of Biological Chemistry 263:6268-6275.

93. Miral, D. B., V. M. Lustre, and M. J. Chamberlin. 1992. An operon of Bacillus subtilis motility genes transcribed by the sigma-D form of RNA polymerase. Journal of Bacteriology 174:4197 -4204.

94. Moir, D, and H. Paulus. 1977. Properties and subunit structure of aspartokinase II from Bacillus subtilis. The Journal of Biological Chemistry 252:4648-4654.

95. Morrison, T. B., and J. S. Parkinson. 1994. Liberation of an interaction domain from the phosphotransfer region of CheA, a signaling kinase of E. coli. Proceedings of the National Academy of Sciences U. S.A 91: 5485- 5489.

96. Ninfa, E. G., A. Stock, S. Mowbray, and J. Stock. 1991. Reconstitution of the bacterial chemotaxis signal transduction system from purified components. The Journal of Biological Chemistry 266:9764-9770.

97. Ohtake, H. 1995. Personal communication.

98. Oosawa, K. N., J. F. Hess, and M. I. Simon. 1988. Mutants defective in bacterial chemotaxis show modified protein phosphorylation. Cell 53:89- 96.

99. Park, C. P., and G. L. Hazelbauer. 1986. Mutations specifically affecting 127

ligand interaction of the Trg chemosensory transducer. Journal of Bacteriology 167: 101-109.

100. Parkinson, J. S. 1976. cheA, cheB, and cheC genes of Escherichia coli and their role in chemotaxis. Journal of Bacteriology 126:758-770.

101. Parkinson, J. S. 1978. Complementation analysis and deletion mapping of Escherichia coli mutants defective in chemotaxis. Journal of Bacteriology 135:45-53.

102. Parkinson, J. S., and S. E. Houts. 1982. Isolation and behavior of Escherichia coli deletion mutants lacking chemotaxis functions. Journal of Bacteriology 151: 106-113.

103. Parkinson, J. S., and E. C. Kofoid. 1992. Communication modules in bacterial signaling proteins. Annual Review of Genetics 26:71-112.

104. Postma, P. W., J. W. Lengeler, and G. R. Jacobson. 1993. Phosphoenolpyruvate:carbohydrate phosphotransferase systems of bacteria. Microbiological Reviews 57:543-594.

105. Ravid, S., P. Matsumura, and M. Eisenbach. 1986. Restoration of flagellar clockwise rotation in bacterial envelopes by insertion of the chemotaxis protein CheY. Proceedings of the National Academy of Sciences U.S.A. 83:7157-7161.

106. Roman, S. J., M. Meyers, K. Volz, and P. Matsumura. 1992. A chemotactic signaling surface on CheY defined by suppressors of flagellar switch mutations. Journal of Bacteriology 174:6247-6255.

107. Rowsell, E. H., J.M. Smith, A. J. Wolfe, and B. L. Taylor. 1995. CheA, CheW, and CheY are required for chemotaxis to oxygen and sugars of the phosphotransferase system in Escherichia coli. Journal of Bacteriology 177:6011-6014.

108. Rudolph, J., and D. Oesterhelt. 1995. Chemotaxis and phototaxis require a CheA histidine kinase in the archaeon Halobacterium salinarium. The EMBO Journal 14:667-673.

109. Russell, C. B., R. C. Stewart, and F. W. Dahlquist. 1989. Control of transducer methylation levels in Escherichia coli: investigation of components essential for modulation of methylation and demethylation reactions. Journal of Bacteriology 171:3609-3618. 128

110. Sambrook, J.E., E. F. Fritsch, and T. Maniatis. 1989. Enzymes used in Molecular Cloning, p. 5.31. In C. Nolan (2nd ed.}, Molecular Cloning: a laboratory manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.

111. Sambrook, J. E., E. F. Fritsch, and T. Maniatis. 1989. Plasmid Vectors, p. 1.65-1.71. In C. Nolan (2nd ed.), Molecular Cloning: a laboratory manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.

112. Sambrook, J.E., E. F. Fritsch, and T. Maniatis. 1989. Southern Hybridization, for Analysis of Genomic DNA, p. 9.31-9.57. In C. Nolan (2nd ed.}, Molecular Cloning: a laboratory manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.

113. Sanatinia, H., E. C. Kofoid, T. B. Morrison, and J. S. Parkinson. 1995. The smaller of the two overlapping cheA gene products is not essential for chemotaxis in Escherichia coli. Journal of Bacteriology 177:2713-2720.

114. Sanders, D. A., and D. E. Koshland, Jr. 1988. Receptor interactions through phosphorylation and methylation pathways in bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 85:8425-8429.

115. Sanger, F., S. Nicklen, and A. R. Coulson. 1977. DNA sequencing with chain-terminating inhibitors. Proceedings of the National Academy of Sciences U.S.A 74:5463-5467.

116. Schaechter, M., and S. L. Gorbach. 1996. The intestinal ecology of E. coli revisited. American Society for Microbiology News 62:304.

117. Schneierson, S. S., and E. J. Bottone. 1973. Erwinia infections in man. Critical Reviews in Clinical Laboratory Sciences 4:341-355.

118. Scott, W. G., and B. L. Stoddard. 1994. Transmembrane signaling and the aspartate receptor. Current Biology 2:877-887.

119. Segall, J. E., M. D. Manson, and H. C. Berg. 1986. Temporal comparisons in bacterial chemotaxis. Proceedings of the National Academy of Sciences U.S.A. 83:8987-8991.

119a. Shine, J. and L. Dalgamo. 1974. The 3'-terminal sequence of Escherichia coli 16S ribosomal RNA: complementary to nonsense triplets and ribosome binding sites. Proceedings of the National Academy of 129

Sciences U.S.A. 71:1342-1346.

120. Silhavy, T. J., M. L. Berman, and L. W. Enquist. 1984. Experiments with gene fusions. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y.

121. Silverman, M., and M. I. Simon. 1974. Flagellar rotation and the mechanism of bacterial motility. Nature (London) 249:73-74.

122. Silverman, M., and M. I. Simon. 1977. Identification of polypeptides necessary for chemotaxis in Escherichia coli. Journal of Bacteriology 130:1317-1325.

123. Smith, R. A., and J. S. Parkinson. 1980. Overlapping genes at the cheA locus of Escherichia coli. Proceedings of the National Academy of Sciences U.S.A. 77:5370-5374.

124. Springer, W.R., and D. E. Koshland, Jr. 1977. Identification of a protein methyltransferase as the cheR gene product in the bacterial sensing system. Proceedings of the National Academy of Sciences U.S.A. 74:533- 537.

125. Stader, J., P. Matsumura, D. Vacante, G. E. Dean, and R. M. Macnab. 1986. Nucleotide sequence of the Escherichia coli motB gene and site­ limited incorporation of its product into the cytoplasmic membrane. Journal of Bacteriology 166:244-252.

126. Stock A., T. Chen, D. Welsh, and J. Stock. 1988. CheA protein, a central regulator of bacterial chemotaxis, belongs to a family of proteins that control gene expression in response to changing environmental conditions. Proceedings of the National Academy of Sciences U.S.A. 85: 1403-1407.

127. Stock, J., and D. E. Koshi and, Jr. 1978. A protein methylesterase involved in bacterial sensing. Proceedings of the National Academy of Sciences U.S.A. 75:3659-3663.

128. Stock, J., G. S. Lukat, and A. Stock. 1991. Bacterial chemotaxis and the molecular logic of intracellular signal transduction networks. Annual Review of Biophysics and Biophysical Chemistry 20:109-136.

129. Stock, J. 1995. Personal communication.

130. Stock, J. and M. G. Surette. 1996. Chemotaxis. p. 1103-1129. In F. C. 130

Neidhardt, R. Curtiss Ill, J. L. Ingraham, E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M. Riley, M. Schaechter, and H. E. Umbarger (ed), Escherichia coli and Salmonella: cellular and molecular biology, 2nd ed. American Society for Microbiology, Washington, D. C.

131. Studier, F. W., A. H. Rosenberg, J. J. Dunn, and J. W. Dubendorff. 1990. Use of T7 RNA polymerase to direct expression of cloned genes. Methods in Enzymology 185:60-89.

132. Swanson, R. V., S. C. Schuster, and M. I. Simon. 1993. Expression of CheA fragments which define domains encoding kinase, phosphotransfer, and CheY binding activities. Biochemistry 32:7623-7629.

133. Swanson, R. V., R. B. Bourret, and M. I. Simon. 1993. Intermolecular complementation of the kinase activity of CheA. Molecular Microbiology 8:435-441.

134. Wang, H, and P. Matsumura. 1996. Characterization of CheAS/CheZ complex: a specific interaction resulting in enhanced dephosphorylating activity on CheY-phosphate. Molecular Microbiology 19:695-703.

135. Ward, M. J., A. W. Bell, P.A. Hamblin, H. L. Packer, and J. P. Armitage. 1995. Identification of a chemotaxis operon with two cheY genes in Rhodobacter sphaeroides. Molecular Microbiology 17:357-366.

136. Wolfe, A. J., and H. C. Berg. 1989. Migration of bacteria in semisolid agar. Proceedings of the National Academy of Sciences U.S.A. 86:6873- 6977.

137. Wolfe, A. J., and R. C. Stewart. 1993. The short form of the CheA protein restores kinase activity and chemotactic ability to kinase-deficient mutants. Proceedings of the National Academy of Sciences U.S.A. 90:1518-1522.

138. Wolfe, A. J., B. P. McNamara, and R. Stewart. 1994. The short form of CheA couples chemoreception to CheA phosphorylation. Journal of Bacteriology 176: 1878-1885.

139. Woodcock, D. M., P. J. Crowther, J. Doherty, S. Jefferson, E. DeCruz, M. Noyer-Weidner, S. S. Smith, M. Z. Michael, and M. W. Graham. 1989. Quantitative evaluation of Escherichia coli host strains for tolerance to cytosine methylation in plasmid and phage recombination. Nucleic Acids Research. 17:3469-3478. 131

140. Wootton, J. C., and H. M. Drummond. 1989. The Q-linker: a class of interdomain sequences found in bacterial multidomain regulatory proteins. Protein Engineering 2:535-543.

141. Wylie, D., A. Stock, C.-Y. Wong, and J. Stock. 1988. Sensory transduction in bacterial chemotaxis involves phosphotransfer between Che proteins. Biochemical and Biophysical Research Communications 151 :891-896.

142. Yanisch-Perron, C., J. Veira, and J. Messing. 1985. Improved M13 phage cloning vectors and host strains: nucleotide sequences of the M13mp18 and pUC19 vectors. Gene 33:103-119.

143. Yao, V. J., and J. L. Spudich. 1992. Primary structure of an archaebacterial transducer, a methyl-accepting protein associated with sensory rhodopsin I. Proceedings of the National Academy of Sciences U.S.A. 89:11915-11919.

144. Zhu, X., C. D. Amsler, K. Voltz, and P. Matsumura. 1996. Tyrosine 106 of CheY plays an important role in chemotaxis signal transduction in Escherichia coli. Journal of Bacteriology 178:4208-4215. VITA

The author, Barry P. McNamara, was born in Detroit, Michigan to Justina

and Eugene McNamara. He received his secondary education at Plymouth

Salem High School and graduated in 1983. He then attended Eastern Michigan

University in Ypsilanti, Michigan where he received a Bachelor of Science

degree in 1989. During this time, Mr. McNamara focused his studies in the field

of biology with a concentration in Microbiology.

Upon receiving his degree, Mr. McNamara was employed by the

University of Michigan Hospitals where he worked as a Laboratory Technologist

in the Clinical Microbiology Laboratory. During this time he was also employed

at the ANR Pipeline Company where he assisted in developing methods to monitor microbial populations present in gas and oil field produced fluids.

In August 1991, Mr. McNamara began his graduate studies in the

Department of Microbiology and Immunology at Loyola University Chicago. The following August, he joined the research team of Dr. Alan Wolfe where he focused his efforts on studying the signal transduction pathway that mediates bacterial chemotaxis in Escherichia coli. Mr. McNamara was the recipient of a

132 133

Pre-doctoral Minority Supplemental Fellowship awarded by the American

Society for Microbiology in 1992 and a Dissertation Fellowship awarded by the

Ford Foundation in 1995.

Mr. McNamara has recently accepted a position as a Post-Doctoral

Fellow in the Laboratory of Dr. Michael Dannenberg within the Department of

Infectious Diseases at the University of Maryland at Baltimore. He has begun investigating protein secretion from enteropathogenic Escherichia coli and its role in initiating aberrant signal transduction pathways within mammalian host cells during infection. DISSERTATION APPROVAL SHEET

The dissertation submitted by Barry P. McNamara has been read and approved by the following committee:

Alan J. Wolfe, Ph.D., Advisor Associate Professor, Microbiology and Immunology Loyola University Chicago

Charles F. Lange, Ph.D. Professor Emeritus, Microbiology and Immunology Loyola University Chicago

David Hecht, M.D. Associate Professor, Infectious Diseases Loyola University Chicago

Thomas Gallagher, Ph.D. Assistant Professor, Microbiology and Immunology Loyola University Chicago

John Lopes, Ph.D. Associate Professor, Molecular and Cellular Biochemistry Loyola University Chicago

The final copies have been examined by the advisor of the dissertation and the signature which appears below verifies the fact that any necessary changes have been incorporated and that the dissertation is now given final approval by the committee with reference to content and form.

The dissertation is, therefore, accepted in partial fulfillment of the requirements for the degree of Doctor of Philosophy.

Date