22.02.2013 Views

POSTERS - BLAST X - University of Utah

POSTERS - BLAST X - University of Utah

POSTERS - BLAST X - University of Utah

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>BLAST</strong> X MEETING<br />

CAMINO REAL SUMIYA HOTEL<br />

CUERNAVACA, MEXICO<br />

JANUARY 18-23, 2009<br />

Meeting Chairperson:<br />

Dr. David Zusman –<strong>University</strong> <strong>of</strong> California, Berkeley, CA<br />

Meeting Vice-Chairperson:<br />

Dr. Robert Bourret – <strong>University</strong> <strong>of</strong> North Carolina, Chapel Hill, NC<br />

Local Arrangements Chairperson<br />

Dr. Georges Dreyfus - Universidad Nacional Autonoma de Mexico, Mexico City<br />

Program Committee:<br />

Dr. John S. Parkinson (Chairperson) – <strong>University</strong> <strong>of</strong> <strong>Utah</strong>, Salt Lake City, UT<br />

Dr. Joe Falke – <strong>University</strong> <strong>of</strong> Colorado, Boulder, CO<br />

Robert Kadner and Robert Macnab Awards Selection Committee:<br />

Dr. Gladys Alexandre – <strong>University</strong> <strong>of</strong> Tennessee, Knoxville, TN<br />

Dr. Gerald Hazelbauer (Chairperson) – <strong>University</strong> <strong>of</strong> Missouri, Columbia, MO<br />

Dr. Ruth Silversmith - <strong>University</strong> <strong>of</strong> North Carolina, Chapel Hill, NC<br />

Dr. Alan Wolfe – Loyola <strong>University</strong>, Maywood, IL<br />

Meeting Review Committee:<br />

Dr. Phillip Aldridge – Newcastle <strong>University</strong>, Newcastle upon Tyne, UK<br />

Dr. Brian Crane (Chairperson) – Cornell <strong>University</strong>, Ithaca, NY<br />

Dr. John Kirby – <strong>University</strong> <strong>of</strong> Iowa, Iowa City, IA<br />

Dr. Birgit Scharf – Virginia Tech <strong>University</strong>, Blacksburg, VA<br />

Board <strong>of</strong> Directors – <strong>BLAST</strong>, Inc.:<br />

Dr. Joe Falke – <strong>University</strong> <strong>of</strong> Colorado, Boulder, CO<br />

Dr. Michael Manson – Texas A&M <strong>University</strong>, College Station, TX<br />

Dr. Philip Matsumura (Chairperson) – <strong>University</strong> <strong>of</strong> Illinois at Chicago, IL<br />

Dr. John S. Parkinson – <strong>University</strong> <strong>of</strong> <strong>Utah</strong>, Salt Lake City, UT<br />

Administrative Assistants:<br />

Ms. Tarra Bollinger – Molecular Biology Consortium, Chicago, IL<br />

Ms. Peggy O’Neill – Molecular Biology Consortium, Chicago, IL<br />

ii


<strong>BLAST</strong> X MEETING SCHEDULE<br />

TIME EVENT LOCATION<br />

Sunday, January 18, 2009<br />

5:00 pm Poster room available for poster setup Orquideas Room<br />

7:00 pm – 8:00 pm Dinner Garden<br />

9:00 pm – 11:00 pm Welcome Reception Crisantemos Room<br />

Monday, January 19, 2009<br />

7:30 am – 8:30 am Breakfast Garden<br />

8:45 am – 8:50 am Welcome -- Local Arrangements Chair (G. Dreyfus) Crisantemos Room<br />

8:50 am – 8:55 am Welcome/ Announcements - Meeting Chair (D. Zusman) Crisantemos Room<br />

9:00 am – 12:00 pm Meeting Session – “Two-Component Systems” Crisantemos Room<br />

10:15 am – 10:30 am C<strong>of</strong>fee Break<br />

12:00 pm – 1:30 pm Lunch Garden<br />

2:00 pm – 4:00 pm Poster Session – even numbered posters Orquideas Room<br />

6:00 pm – 7:30 pm Dinner Garden<br />

7:30 pm – 10:00 pm Meeting Session – “Gliding Motility” Crisantemos Room<br />

8:30 pm – 8:45 pm C<strong>of</strong>fee Break<br />

Tuesday, January 20, 2009<br />

7:30 am – 8:30 am Breakfast Garden<br />

9:00 am – 12:00 pm Meeting Session – “Flagella” Crisantemos Room<br />

10:15 am – 10:30 am C<strong>of</strong>fee Break<br />

12:00 pm – 1:30 pm Lunch Garden<br />

2:00 pm – 4:00 pm Poster Session – odd numbered posters Orquideas Room<br />

6:00 pm – 7:30 pm Dinner Garden<br />

7:30 pm – 10:00 pm Meeting Session – “Regulation” Crisantemos Room<br />

8:30 pm – 8:45 pm C<strong>of</strong>fee Break<br />

Wednesday, January 21, 2009<br />

7:30 am – 8:30 am Breakfast Garden<br />

9:00 am – 12:00 pm Meeting Session – “Chemoreceptors” Crisantemos Room<br />

10:15 am – 10:30 am C<strong>of</strong>fee Break<br />

12:00 pm – 1:30 pm Lunch Garden<br />

1:30 pm Tour bus to Taxco leaves from Hotel Hotel entrance<br />

1:45 pm Tour buses to Xochicalco leave from Hotel Hotel entrance<br />

Thursday, January 22, 2009<br />

7:30 am – 8:30 am Breakfast Garden<br />

9:00 am – 12:00 pm Meeting Session – “Bi<strong>of</strong>ilms and Host Interactions” Crisantemos Room<br />

10:15 am – 10:30 am C<strong>of</strong>fee Break<br />

12:00 pm – 1:30 pm Lunch Garden<br />

6:00 pm – 7:30 pm Dinner Garden<br />

7:30 pm – 7:45 pm Robert Kadner and Robert Macnab Awards Presentation Crisantemos Room<br />

7:45 pm – 10:00 pm Meeting Session – “Signaling and Behavior” Crisantemos Room<br />

8:30 pm – 8:45 pm C<strong>of</strong>fee Break<br />

10:00 pm – 12:00 am Reception Crisantemos Room<br />

Friday, January 23, 2009<br />

7:30 am – 8:30 am Breakfast Garden<br />

Shuttle buses leave for Mexico City Airport<br />

iii


<strong>BLAST</strong> X PROGRAM<br />

January 19, 2009 Two-Component Systems<br />

Monday Morning – 9:00 am – 12:00 pm Chair – Georges Dreyfus<br />

ABSTRACT<br />

PRESENTER TITLE<br />

PAGE NO.<br />

Bourret, Bob<br />

A structural investigation <strong>of</strong> response regulator<br />

autodephosphorylation<br />

Intramolecular autophosphorylation <strong>of</strong> the Escherichia coli ArcB<br />

2<br />

Peña-Sandoval, G. sensor kinase 3<br />

Porter, Steven<br />

BREAK<br />

A bifunctional kinase-phosphatase in bacterial chemotaxis<br />

Message passing: Protein structure assembly from sequence data for<br />

4<br />

Szurmant, Hendrik two-component signaling proteins<br />

Integrated control <strong>of</strong> Caulobacter cell envelope physiology by a hybrid<br />

5<br />

Crosson, Sean two-component/ECF sigma factor signaling network<br />

The role <strong>of</strong> signal transduction in cell wall metabolism in Bacillus<br />

6<br />

Bisicchia, Paola subtilis<br />

The WalK/WalR essential signal transduction pathway and cell wall<br />

7<br />

Dubrac, Sarah homeostasis in Staphylococcus aureus 8<br />

January 19, 2009 Gliding Motility<br />

Monday Evening – 7:30 pm – 9:30 pm Chair – Lotte Søgaard Andersen<br />

Bulyha, Iryna<br />

Dynamic assembly and disassembly <strong>of</strong> the Type IV molecular<br />

machine<br />

A “four component” signal transduction system regulates<br />

9<br />

Higgs, Penelope developmental progression in Myxococcus xanthus<br />

Independence and interdependence <strong>of</strong> Dif and Frz chemosensory<br />

10<br />

Yang, Zhaomin<br />

BREAK<br />

pathways in Myxococcus xanthus chemotaxis 11<br />

Berleman, James Predataxis behavior in Myxococcus xanthus 12<br />

Mauriello, Emilia Dynamic localization <strong>of</strong> FrzCD in Myxococcus xanthus<br />

Identifying novel bacterial cytoskeletal elements and cytoskeletal<br />

13<br />

Gitai, Zemer interactors through high-throughput imaging 14<br />

January 20, 2009 Flagella<br />

Tuesday Morning – 9:00 am – 12:00 pm Chair – Robert Belas<br />

Rao, Christopher<br />

The role <strong>of</strong> positive feedback in controlling flagella assembly<br />

dynamics<br />

Crystal structure <strong>of</strong> FliT, a bacterial flagellar export chaperone for the<br />

15<br />

Imada, Katsumi filament cap protein Hap2 (FliD)<br />

Structural insight into active flagellar motor formation through the<br />

16<br />

Kojima, Seiji periplasmic region <strong>of</strong> MotB 17<br />

Thormann, Kai<br />

BREAK<br />

Stator selection in Shewanella oneidensis 18<br />

Gauthier, Mathieu Taking control <strong>of</strong> the bacterial flagellar motor<br />

Experimental evidence for conformational spread in the bacterial<br />

19<br />

Branch, Richard switch complex<br />

Do individual bacterial flagellar motors use hysteresis to maintain a<br />

20<br />

Reuven, Peter robust output in a noisy environment? – an experimental study 21<br />

Tu, Yuhai Dynamics <strong>of</strong> the bacterial flagellar motor with multiple stators 22<br />

iv


January 20, 2009 Regulation<br />

Tuesday Evening – 7:30 pm – 9:30 pm Chair – Linda Kenney<br />

ABSTRACT<br />

PRESENTER TITLE<br />

PAGE NO.<br />

Kaserer, Alla<br />

Partridge, John<br />

Reimann, Sylvia<br />

BREAK<br />

Tran, Hoa<br />

Martínez, Luary<br />

Lee, Yi-Ying<br />

Effect <strong>of</strong> osmolytes on regulating the activities <strong>of</strong> the SSK1<br />

response regulator from Saccharomyces cerevisiae 23<br />

Regulation <strong>of</strong> Escherichia coli motility by the nitric oxide sensitive<br />

transcriptional repressor NsrR 24<br />

Synthetic lethality uncovers a novel link between the MalT and<br />

OmpR regulons 25<br />

A chemotaxis-like signaling pathway regulates the expression <strong>of</strong><br />

extracellular materials in Geobacter sulfurreducens 26<br />

The two-component regulatory system BarA/SirA is at the top <strong>of</strong> a<br />

multi-factorial regulatory cascade controlling the expression <strong>of</strong> the<br />

SPI-1 and SPI-2 virulence regulons in Salmonella 27<br />

Interaction <strong>of</strong> the transcriptional regulatory complex, FlhDC, with its<br />

target DNA 28<br />

January 21, 2009 Chemoreceptors<br />

Wednesday Morning – 9:00 am – 12:00 pm Chair – Sandy Parkinson<br />

Glekas, George A novel amino acid binding structure in bacterial chemotaxis<br />

Structure and function <strong>of</strong> the Helicobacter pylori chemoreceptor<br />

29<br />

Remington, James TlpB 30<br />

Wright, Gus The TM2-HAMP connection<br />

Structure, assembly and conformational changes in<br />

chemoreceptors studied in intact bacterial cells using Cryo-<br />

31<br />

Khursigara, Cezar<br />

BREAK<br />

electron tomography<br />

Discrete signal-on and -<strong>of</strong>f conformations in the Aer HAMP<br />

32<br />

Watts, Kylie<br />

domain<br />

Investigating the structure <strong>of</strong> ternary complex <strong>of</strong> histidine kinase<br />

CheA, coupling protein CheW, and chemoreceptor by pulsed<br />

33<br />

Bhatnagar, Jaya dipolar ESR spectroscopy 34<br />

Erbse, Annette The chemotactic core signalling complex is ultrastable 35<br />

Briegel, Ariane Electron cryotomography <strong>of</strong> bacterial chemotaxis arrays 36<br />

v


January 22, 2009 Bi<strong>of</strong>ilms & Host Interactions<br />

Thursday Morning – 9:00 am – 12:00 pm Chair – Alan Wolfe<br />

ABSTRACT<br />

PRESENTER TITLE<br />

PAGE NO.<br />

Belas, Robert<br />

Two regulatory proteins control the swim-or-stick switch in<br />

Roseobacters<br />

Deletion analysis <strong>of</strong> RcsC reveals a novel signaling-pathway<br />

37<br />

Oropeza, Ricardo controlling bi<strong>of</strong>ilm formation in Escherichia coli 38<br />

Vlamakis, Hera Regulation <strong>of</strong> cell fate in Bacillus subtilis bi<strong>of</strong>ilms<br />

Protein misfolding done right: The biogenesis <strong>of</strong> bacterial amyloid<br />

39<br />

Chapman, Matt<br />

BREAK<br />

fibers 40<br />

Martinez del Campo, Rhodobacter sphaeroides, a bacterium with two flagellar systems<br />

Ana<br />

and multiple chemotaxis gene homologs<br />

Motility, chemotaxis and virulence <strong>of</strong> Borrelia burgdorferi, the<br />

41<br />

Motaleb, MD<br />

lyme disease spirochete<br />

Pleiotropic phenotypes <strong>of</strong> a Yersinia enterocolicia flhD mutant<br />

42<br />

Prüß, Birgit<br />

include reduced lethality in a chicken embryo model<br />

Regulation <strong>of</strong> motility by quorum sensing in Sinorhizobium meliloti<br />

43<br />

Gonzalez, Juan and its role in symbiosis establishment 44<br />

January 22, 2009 Signaling & Behavior<br />

Thursday Evening – 7:30 pm – 9:30 pm Chair – Judy Armitage<br />

Goldberg, Shalom Engineered single- and multi-cell chemotaxis in E. coli 45<br />

Photo-energy conversion and sensory transduction <strong>of</strong> microbial 46<br />

Jung, Kwang-Hwan rhodopsins in photosynthetic microbes<br />

Function <strong>of</strong> multiple chemotaxis-like pathways in mediating<br />

changes in motility patterns and cellular morphology in<br />

47<br />

Bible, Amber<br />

BREAK<br />

Azospirillum brasilense<br />

Probing adaptation kinetics in vivo by fluorescence resonance<br />

48<br />

Shimizu, Thomas energy transfer<br />

Neumann, Silke Minor receptor signalling in E. coli 49<br />

A systems biology approach to understanding how Bacillus<br />

50<br />

Chastanet, Arnaud makes up its mind<br />

vi


<strong>POSTERS</strong> - <strong>BLAST</strong> X<br />

Poster # Lab Presenter Title Page #<br />

1 Phillip<br />

Aldridge<br />

Aldridge, Christine The characterisation <strong>of</strong> the dynamics <strong>of</strong> the<br />

FliT:FliD:FlhD4C2 interaction and its role in<br />

regulating flagellar assembly<br />

52<br />

Aldridge, Phillip Subunit feedback control <strong>of</strong> flagellar filament<br />

assembly in Caulobacter crescentus<br />

53<br />

Alexandre, Gladys Function <strong>of</strong> unique domains <strong>of</strong> CheA1 from A.<br />

Brasilense in regulating multiple cellular behaviors<br />

54<br />

Brown, Mostyn How does the Rhodobacter sphaeroides flagellar<br />

motor stop – Using a clutch or a brake?<br />

55<br />

Delalez, Nicolas Jacques Dynamics <strong>of</strong> the flagellar motor protein FliM 56<br />

2 Phillip<br />

Aldridge<br />

3 Gladys<br />

Alexandre<br />

4 Judith<br />

Armitage<br />

5 Judith<br />

Armitage<br />

6 Judith<br />

Roberts, Mark A J Using control theory to elucidate connectivity in R. 57<br />

Armitage<br />

Sphaeroides chemotaxis<br />

7 Howard<br />

Berg<br />

Yuan, Junhua Behavior <strong>of</strong> the flagellar rotary motor near zero load 58<br />

8 David<br />

Bolam<br />

Diaz-Mireles, Edith Nutrient sensing by a human gut symbiont 59<br />

9 Edmundo De la Cruz, Miguel Ángel EnvZ-OmpR and CpxA-CpxR regulate ompS1 by 60<br />

Calva<br />

differential promoter expression<br />

10 Edmundo<br />

Calva<br />

Gallego, Ana The LeuO regulon in Salmonella 60<br />

11 Nyles<br />

Miller, Kelly Ann The complex hook basal body structure <strong>of</strong> the Lyme 61<br />

Charon<br />

disease spirochete Borrelia burgdorferi<br />

12 Brian<br />

Airola Airola, Michael How do PAS and HAMP domains communicate? 62<br />

Crane<br />

Insights from Aer2, a heme based sensor for<br />

aerotaxis<br />

13 Brian<br />

Pollard, Abiola M Structure <strong>of</strong> soluble chemoreceptor suggests a 63<br />

Crane<br />

mechanism for propagating conformational signals<br />

14 Georges<br />

Castillo, David Jaime Functional analysis <strong>of</strong> a large non-conserved region 64<br />

Dreyfus<br />

<strong>of</strong> FlgK (HAP1) from Rhodobacter sphaeroides<br />

15 Georges<br />

Dreyfus<br />

16 Thierry<br />

Emonet<br />

17 Joseph<br />

Falke<br />

18 Dimitris<br />

Georgellis<br />

19 Georgellis,<br />

Dimitris<br />

20 Juan<br />

Gonzalez<br />

21 Bertha<br />

González-Pedrajo<br />

22 Rasika<br />

Harshey<br />

23 Rasika<br />

Harshey<br />

De la Mora, Javier The flagellar muramidase from the photosynthetic<br />

bacterium Rhodobacter sphaeroides<br />

Alexander, Roger Modeling scaffold phosphorylation as an adaptation<br />

mechanism in bacterial chemotaxis<br />

Swain, Kalin Testing the Yin-Yang model <strong>of</strong> signal transduction<br />

in a bacterial chemoreceptor cytoplasmic domain<br />

Alvarez, Adrian Fernando Cytochrome d but not cytochrome o rescues the<br />

toluidine blue growth sensitivity <strong>of</strong> arc mutants in E.<br />

Coli<br />

González, Ricardo Searching the physiological signal(s) that regulate<br />

the activity <strong>of</strong> the sensor kinase BarA<br />

Gurich, Nataliya The role <strong>of</strong> quorum sensing in the control <strong>of</strong> motility<br />

and plant invasion by Sinorhizobium meliloti<br />

García-Gómez, Elizabeth Characterization <strong>of</strong> etga, a muramidase associated<br />

with the type III secretion system <strong>of</strong><br />

enteropathogenic Escherichia coli<br />

71<br />

Lee, Jae-Min FlhE: a periplasmic chaperone <strong>of</strong> flagellin? 72<br />

Nieto, Vincent Michael The cyclic-di-GMP receptor protein YcgR localizes<br />

to the flagellar basal body and changes motor bias<br />

in Salmonella<br />

vii<br />

65<br />

66<br />

67<br />

68<br />

69<br />

70<br />

73


<strong>POSTERS</strong> - <strong>BLAST</strong> X<br />

Poster # Lab Presenter Title Page #<br />

24 Michio<br />

Hizukuri, Yohei Analysis <strong>of</strong> the peptidoglycan-binding domain <strong>of</strong> the 75<br />

Homma<br />

flagellar stator protein MotB using systematic<br />

mutagenesis and chimeric protein in Escherichia<br />

coli<br />

25 Michio<br />

Koike, Masafumi Attempt to purify the hook-basal body with C-ring 76<br />

Homma<br />

from the Na+-driven flagellar motor<br />

26 Scott<br />

Kline, Kimberly Mechanism for sortase localization and role in 77<br />

Hultgren<br />

efficient pilus assembly in Enterococcus faecalis<br />

27 Akihiko<br />

Fukuoka, Hajime Visualization <strong>of</strong> exchange <strong>of</strong> rotor component in 78<br />

Ishijima<br />

functioning bacterial flagellar motor<br />

28 Akihiko<br />

Inoue, Yuichi Torque response <strong>of</strong> the sodium-driven chimeric 79<br />

Ishijima<br />

flagellar motor in E.coli induced by reversible<br />

temperature change<br />

29 Christine<br />

Josenhans, Christine The Helicobacter pylori flagellar anti-sigma factor 80<br />

Josenhans<br />

FlgM remains bacteria-associated and interacts<br />

with FlhAc 30 Ikuro<br />

Inaba, Takehiko The localization patterns <strong>of</strong> all histidine kinases in 81<br />

Kawagishi<br />

Escherichia coli cell<br />

31 Ikuro<br />

Nishiyama, So-ichiro Thermosensing function <strong>of</strong> Aer, a redox sensor <strong>of</strong> 82<br />

Kawagishi<br />

E. Coli<br />

32 Kenji<br />

Oosawa<br />

Hayashi, Fumio ATPase activity <strong>of</strong> T3SS specific ATPase InvC 83<br />

33 Kenji<br />

Hayashi, Fumio Characterizations <strong>of</strong> the pseudorevertants from 84<br />

Oosawa Salmonella typhimurium strain SJW1655 and<br />

SJW1660 with the R- and the L-type straight<br />

flagellar filaments<br />

34 Kenji<br />

Hayashi, Fumio Raman optical activity and vibrational circular 85<br />

Oosawa<br />

dichroism <strong>of</strong> flagellar filaments <strong>of</strong> Salmonella<br />

35 Cancelled<br />

Poster<br />

86<br />

36 John<br />

Willett, Jonathan CrdC negatively regulates CheW3 and CheA3 87<br />

Kirby<br />

interaction during signal transduction in<br />

Myxococcus xanthus<br />

37 Jun<br />

Liu, Jun Molecular architecture <strong>of</strong> intact flagellar motor 88<br />

Liu<br />

revealed by Cryo-Electron tomography<br />

38 Janine<br />

Maddock<br />

Dobkowski, Jason Mining the E. Coli GFP fusion collection 89<br />

39 Michael<br />

Adase, Christopher Understanding the fundamental elements <strong>of</strong><br />

90<br />

Manson<br />

signaling in the Tar chemoreceptor<br />

40 Michael<br />

Manson<br />

Crowder, Rachel Leann Linking the TM2 to HAMP—a tough nut to crack? 91<br />

41 Michael<br />

Seely, Andrew Electrostatic effects on signaling mutations in the C- 92<br />

Manson<br />

terminal region <strong>of</strong> the Escherichia coli aspartate<br />

chemoreceptor<br />

42 Philip<br />

Lee, Yi-Ying Interaction <strong>of</strong> the transcriptional regulatory complex, 93<br />

Matsumura<br />

FlhDC, with its target DNA<br />

43 Jonathan<br />

McMurry, Jonathan Application <strong>of</strong> biolayer interferometry to<br />

94<br />

McMurry<br />

understanding interactions among Salmonella<br />

enterica flagellar export apparatus proteins<br />

44 Paul<br />

Simons, Julie A Cross-Species Comparison <strong>of</strong> Chemotactic 95<br />

Milewski<br />

Behavior<br />

45 Makoto<br />

Miyata<br />

Nakane, Daisuke Tethered Mycoplasma 96<br />

viii


<strong>POSTERS</strong> - <strong>BLAST</strong> X<br />

Poster # Lab Presenter Title Page #<br />

46 Makoto<br />

Miyata<br />

47 Tarek<br />

Msadek<br />

48 Tarek<br />

Msadek<br />

49 Keiichi<br />

Namba<br />

50 Keiichi<br />

Namba<br />

51 Keiichi<br />

Namba<br />

52 Keiichi<br />

Namba<br />

53 Keiichi<br />

Namba<br />

Nonaka, Takahiro Molecular shapes <strong>of</strong> Gli123 and Gli521 involved in<br />

gliding motility <strong>of</strong> Mycoplasma mobile<br />

Delauné, Aurelia The essential nature <strong>of</strong> the WalK/WalR signal<br />

transduction pathway is linked to cell wall hydrolase<br />

activity in Staphylococcus aureus<br />

Falord, Mélanie The GraS/GraR two-component system and<br />

dermaseptin resistance in Staphylococcus aureus<br />

Che, Yong-Suk Characterization <strong>of</strong> suppressors <strong>of</strong> the MotB(D33E)<br />

mutation, a putative proton-binding residue <strong>of</strong> the<br />

bacterial flagellar motor<br />

Ibuki, Tatsuya Structure <strong>of</strong> FliJ, a cytoplasmic component <strong>of</strong> the<br />

flagellar type III; protein export apparatus <strong>of</strong><br />

Salmonella<br />

Makino, Fumiaki CryoEM structure <strong>of</strong> the hook-filament junction <strong>of</strong><br />

Salmonella<br />

Nakamura, Shuichi Effect <strong>of</strong> intracellular pH on the torque-speed<br />

relationship <strong>of</strong> bacterial proton-driven flagellar<br />

motor<br />

Yoshimura, Shinsuke Fluorescence imaging <strong>of</strong> assembly and<br />

disassembly <strong>of</strong> the bacterial flagellar protein export<br />

ATPase FliL to the flagellar basal body<br />

Ottemann, Karen M Analysis <strong>of</strong> Helicobacter pylori lacking all four<br />

54 Karen<br />

Ottemann chemoreceptors<br />

55 Rebecca<br />

Parales, Rebecca E. Chemotaxis to pyrimidines and identification <strong>of</strong> a<br />

Parales<br />

cytosine chemoreceptor in Pseudomonas putida<br />

56 Chankyu<br />

Park<br />

57 Simon<br />

Rainville<br />

58 Tracy<br />

Raivio<br />

59 Christopher<br />

Rao<br />

60 Christopher<br />

Rao<br />

61 Kathleen<br />

Ryan<br />

62 Mark<br />

Sansom<br />

63 Florian<br />

Schubot<br />

64 Victoria<br />

Shingler<br />

65 Victor<br />

Sourjik<br />

66 Ann<br />

Stock<br />

67 Claudia<br />

Studdert<br />

Lee, Changhan Reactive aldehydes and motility in Escherichia coli<br />

K12<br />

107<br />

Rainville, Simon Taking control <strong>of</strong> the bacterial flagellar motor 108<br />

Malpica, Roxana Characterization <strong>of</strong> the periplasmic domain <strong>of</strong> the<br />

sensor kinase CpxA: Role <strong>of</strong> conserved residues in<br />

CpxA activity<br />

Saini, Supreet Characterization <strong>of</strong> FliZ as an activator <strong>of</strong> flagellar<br />

genes in Salmonella enterica serovar typhimurium<br />

Wu, Kang Localization <strong>of</strong> the chemotaxis proteins in Bacillus<br />

subtilis<br />

Ryan, Kathleen R RcdA structure and function in regulated CtrA<br />

proteolysis<br />

Hall, Benjamin Modelling MCP signalling mechanisms with highthroughput<br />

simulation <strong>of</strong> Tar TM2<br />

Schubot, Florian D. Structural evidence suggests that antiactivator<br />

ExsD from Pseudomonas aeruginosa is a DNA<br />

binding protein<br />

Herrera Seitz, María Karina A novel PAS-GGDEF-EAL protein involved in<br />

regulation <strong>of</strong> motility in Pseudomonas putida<br />

Sommer, Erik In vivo study <strong>of</strong> the two-component signalling<br />

network in E. coli<br />

Han, Hua Gene regulation by Escherichia coli response<br />

regulator PhoB<br />

Massazza, Diego Ariel New reporter residues <strong>of</strong> trimer formation by<br />

Escherichia coli MCPs<br />

ix<br />

97<br />

98<br />

99<br />

100<br />

101<br />

102<br />

103<br />

104<br />

105<br />

106<br />

109<br />

110<br />

111<br />

112<br />

113<br />

114<br />

115<br />

116<br />

117<br />

118


<strong>POSTERS</strong> - <strong>BLAST</strong> X<br />

Poster # Lab Presenter Title Page #<br />

68 Motohide<br />

Kenri, Tsuyoshi Systematic detection <strong>of</strong> proteins that localize at the 119<br />

Takahashi attachment attachment organelle regions <strong>of</strong> Mycoplasma<br />

Mycoplasma<br />

pneumoniae by fluorescent-protein tagging<br />

69 Barry<br />

Campbell, Asharie Johnson Mapping the signal transduction pathway within the 120<br />

Taylor<br />

PAS domain <strong>of</strong> the Aer receptor<br />

70 Gunnar<br />

Draheim, Roger Modulating two-component signal output with 121<br />

von Heijne<br />

protein-membrane interactions<br />

71 Cancelled<br />

Poster<br />

122<br />

72 Douglas<br />

Copeland, Matthew Francis A mechanical and genetic study <strong>of</strong> Escherichia coli 123<br />

Weibel<br />

swarming motility<br />

73 Douglas<br />

Muralimohan, Abishek Probing hydrodynamic interactrions between 124<br />

Weibel<br />

swimming bacteria using micr<strong>of</strong>ulidics<br />

74 Zhaomin<br />

Black, Wesley P. Examination <strong>of</strong> phosphorylation in the Dif<br />

125<br />

Yang<br />

chemotaxis-like system in Myxococcus xanthus<br />

75 Igor g<br />

(Zhulin) Jouline<br />

Cantwell, , Brian Evolution <strong>of</strong> chemotaxis proteins p on a micro scale 126<br />

76 Igor<br />

(Zhulin) Jouline<br />

Shanafield, Harold Evolution <strong>of</strong> signal transduction in a bacterial genus 127<br />

77 David<br />

Nan, Beiyan The chemosensory receptor FrzCD interacts with 128<br />

Zusman<br />

two A-motility proteins, AglZ and AgmU<br />

78 Dimitris<br />

Barba Ostria, Carlos Arturo Phenotypic characterization <strong>of</strong> all single mutants <strong>of</strong> 129<br />

Georgellis<br />

two component system proteins in Neurospora<br />

crassa<br />

79 Makoto<br />

Miyata<br />

Miyata, Makoto AFM study <strong>of</strong> Mycoplasma mobile's gliding motility 129b<br />

x


SPEAKER ABSTRACTS<br />

1


<strong>BLAST</strong> X Mon. Morning Session<br />

A STRUCTURAL INVESTIGATION OF RESPONSE REGULATOR<br />

AUTODEPHOSPHORYLATION<br />

Yael Pazy 1 , Amy C. Wollish 1 , Stephanie A. Thomas 1 , Peter J. Miller 2 , Edward J. Collins 1,2 ,<br />

Robert B. Bourret 1 , and Ruth E. Silversmith 1<br />

Departments <strong>of</strong> Microbiology & Immunology 1 and Biochemistry & Biophysics 2 , <strong>University</strong> <strong>of</strong><br />

North Carolina, Chapel Hill, NC 27599<br />

Assorted two-component regulatory systems are used to control a wide variety <strong>of</strong><br />

biological processes, which occur over a broad range <strong>of</strong> time scales. To appropriately<br />

synchronize the adaptive responses implemented by response regulators with the<br />

environmental stimuli detected by sensor kinases, the kinetics <strong>of</strong> biochemical signaling<br />

reactions must be at least as fast as the biological process that they regulate. The fraction <strong>of</strong><br />

response regulators in the phosphorylated state is determined by the net result <strong>of</strong><br />

phosphorylation and dephosphorylation. Autodephosphorylation rates reported for various<br />

wildtype response regulators span a range <strong>of</strong> at least 40,000x, consistent with timescales<br />

ranging from about one second to one day. Furthermore, the range <strong>of</strong> rates observed suggests<br />

that autodephosphorylation rates are an important contributor to setting the overall timescales <strong>of</strong><br />

two-component signaling systems.<br />

We previously found that changing the amino acids at the variable active site positions<br />

corresponding to residues 59 and 89 <strong>of</strong> Escherichia coli CheY could alter response regulator<br />

autodephosphorylation rates about 100x (1). Thus, the particular amino acids found at positions<br />

'59' and '89' in response regulators can account for two orders <strong>of</strong> magnitude in<br />

autodephosphorylation rate, but other factors to account for an additional two to three orders <strong>of</strong><br />

magnitude in reaction rate must also exist and remain to be identified. To begin to characterize<br />

the structural basis <strong>of</strong> response regulator autodephosphorylation rate, we determined highresolution<br />

X-ray crystal structures for five CheY mutants that bear amino acid substitutions at<br />

positions 14, 59, and 89 and consequently exhibit autodephosphorylation rates six to 40 times<br />

slower than wildtype CheY. Each structure was determined in the presence <strong>of</strong> the phosphoryl<br />

group analog BeF3 - and thus represents the starting point <strong>of</strong> the autodephosphorylation<br />

reaction. Comparison <strong>of</strong> mutant and wildtype CheY structures, which are matched at all but<br />

three residues and yet support different reaction rates, can potentially provide insight into the<br />

mechanistic basis by which positions '59' and '89' influence autodephosphorylation rate.<br />

Similarly, comparison <strong>of</strong> each mutant CheY structure with the structure <strong>of</strong> a wildtype response<br />

regulator that is matched at eight (three variable and five conserved) active site residues and<br />

yet catalyzes autodephosphorylation at rates 10-80x slower than the CheY mutants can<br />

potentially suggest candidates for additional factors that might contribute to<br />

autodephosphorylation rate.<br />

Response regulator autodephosphorylation likely involves an inline attack on the<br />

phosphoryl group by a nucleophilic water molecule. The structural comparisons outlined above<br />

clearly suggest that autodephosphorylation rate is influenced by the extent to which amino acid<br />

sidechains at positions '59' or '89' sterically occlude access to the phosphoryl group. Additional<br />

factors potentially affecting autodephosphorylation rate will also be discussed.<br />

Reference<br />

1. Thomas, S.A., Brewster, J.A., & Bourret, R.B. (2008) Two nonconserved active site residues<br />

affect response regulator phosphoryl group stability. Molecular Microbiology 69, 453-465.<br />

2


<strong>BLAST</strong> X Mon. Morning Session<br />

INTRAMOLECULAR AUTOPHOSPHORYLATION OF THE ESCHERICHIA COLI ArcB<br />

SENSOR KINASE<br />

Gabriela R. Peña-Sandoval, Luis A. Nuñez Oreza and Dimitris Georgellis<br />

Departamento de Genética Molecular, Instituto de Fisiología Celular, Universidad Nacional<br />

Autónoma de México, 04510 México City, México.<br />

The Arc two-component system is a complex signal transduction system that plays a key<br />

role in regulating energy metabolism at the level <strong>of</strong> transcription in bacteria. This system<br />

comprises the ArcB protein, a tripartite membrane-associated sensor kinase, and the ArcA<br />

protein, a typical response regulator. Under anoxic growth conditions, ArcB autophosphorylates<br />

and transphosphorylates ArcA, which in turn represses or activates the expression <strong>of</strong> its target<br />

operons. Under aerobic conditions, the kinase activity <strong>of</strong> ArcB is silenced by the oxidation <strong>of</strong> two<br />

cytosol-located redox-active cysteine residues that participate in intermolecular disulfide bond<br />

formation, a reaction in which the quinones provide the source <strong>of</strong> oxidative power.<br />

Here we present results demonstrating that the putative leucine-zipper near the second<br />

transmembrane segment <strong>of</strong> ArcB is functional and necessary for proper ArcB signaling.<br />

Moreover, we provide data demonstrating that in contrast to the proposed model <strong>of</strong><br />

intermolecular autophosphorylation, ArcB autophosphorylation is an intra-molecular reaction.<br />

3


<strong>BLAST</strong> X Mon. Morning Session<br />

A BIFUNCTIONAL KINASE-PHOSPHATASE IN BACTERIAL CHEMOTAXIS<br />

Steven L. Porter, Mark A.J. Roberts, Cerys S. Manning and Judith P. Armitage 1<br />

Oxford Centre for Integrative Systems Biology (OCISB), Department <strong>of</strong> Biochemistry, <strong>University</strong><br />

<strong>of</strong> Oxford, South Parks Road, Oxford OX1 3QU<br />

Phosphorylation based signaling pathways employ dephosphorylation mechanisms for<br />

signal termination. Histidine to aspartate phosphosignaling in the two-component system<br />

controlling bacterial chemotaxis has been studied extensively. Rhodobacter sphaeroides has a<br />

complex chemosensory pathway with multiple homologues <strong>of</strong> the Escherichia coli<br />

chemosensory proteins, although it lacks homologues <strong>of</strong> known signal terminating CheY-P<br />

phosphatases such as CheZ, CheC, FliY or CheX. Here we demonstrate that an unusual CheA<br />

homologue, CheA3, is not only a phosphodonor for the principal CheY protein, CheY6, but is<br />

also is a specific phosphatase for CheY6-P. This phosphatase activity accelerates CheY6-P<br />

dephosphorylation to a rate that is comparable with the measured stimulus response time <strong>of</strong> ~1<br />

s. CheA3 possesses only two <strong>of</strong> the five domains found in classical CheAs, the Hpt (P1) and<br />

regulatory (P5) domains, which are joined by a novel 794 amino acid sequence that is required<br />

for phosphatase activity. The P1 domain <strong>of</strong> CheA3 is phosphorylated by CheA4 and it<br />

subsequently acts as a phosphodonor for the response regulators. A CheA3 mutant protein<br />

deleted for the 794 amino acid region lacked phosphatase activity, retained phosphotransfer<br />

function but did not support chemotaxis, suggesting that the phosphatase activity may be<br />

required for chemotaxis. Using a nested deletion approach we show that a 200 amino acid<br />

segment <strong>of</strong> CheA3 is required for phosphatase activity. The phosphatase activity <strong>of</strong> previously<br />

identified non-hybrid histidine protein kinases depends upon the dimerization and histidine<br />

phosphorylation (DHp) domains. CheA3, however, lacks a DHp domain, suggesting that CheA3<br />

is a novel phosphatase.<br />

4


<strong>BLAST</strong> X Mon. Morning Session<br />

MESSAGE PASSING: PROTEIN STRUCTURE ASSEMBLY FROM SEQUENCE DATA FOR<br />

TWO-COMPONENT SIGNALING PROTEINS<br />

Hendrik Szurmant 1 , Martin Weigt 2,3 , Robert White 1,3 , Terry Hwa 3 and James A. Hoch 1<br />

1<br />

Division <strong>of</strong> Cellular Biology, Department <strong>of</strong> Molecular and Experimental Medicine, The Scripps<br />

Research Institute, La Jolla, CA 92037<br />

2<br />

Institute for Scientific Interchange, I-10133 Torino, Italy<br />

3<br />

Center for Theoretical Biological Physics and Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> California at<br />

San Diego, La Jolla, CA 92093<br />

The crystal structure <strong>of</strong> the Bacillus subtilis response regulator Spo0F in complex with<br />

the histidine kinase structural homologue Spo0B defined the active site <strong>of</strong> phosphotransfer and<br />

the spatial interactions <strong>of</strong> two-component systems <strong>of</strong> microbes and plants. The limited<br />

bioinformatic data available at the time was sufficient to deduce and understand the molecular<br />

basis for recognition specificity between histidine kinases and response regulators (J. A. Hoch<br />

and K. I. Varughese. 2001. J. Bacteriol. 183:4941-4949). Today, the availability <strong>of</strong> large protein<br />

databases generated from sequences <strong>of</strong> hundreds <strong>of</strong> bacterial genomes enables more<br />

sophisticated statistical approaches to extract interacting and specificity determining positions<br />

between proteins from protein databases. The goal <strong>of</strong> such studies is to identify protein<br />

interaction surfaces from sequencing data alone, without previous structural knowledge, i.e. cocrystal<br />

data. A number <strong>of</strong> co-variance based approaches producing nearly identical results have<br />

been applied to the highly amplified two-component systems as a means to verify mathematical<br />

data with structural knowledge <strong>of</strong> sensor kinase/response regulator interaction; including one<br />

presented by us at the <strong>BLAST</strong> IX in 2007 (R. A. White et al. 2007. Methods Enzymol. 422:75-<br />

101). While producing potential specificity determining position information, these methods have<br />

an important shortcoming in predicting spatial proximity. They cannot distinguish between<br />

directly correlated (interacting) and indirectly correlated (non-interacting) residue positions. To<br />

address this issue we developed a novel method that combines co-variance analysis with global<br />

inference analysis, adopted from use in statistical physics.<br />

When applied to a set <strong>of</strong> over 2500 representatives <strong>of</strong> the bacterial two-component<br />

signal transduction system, the combination <strong>of</strong> covariance with global inference methods<br />

successfully and robustly identified residue pairs that are proximal in space without resorting to<br />

ad hoc tuning parameters, both for hetero-interactions between sensor kinase (SK) and<br />

response regulator (RR) proteins and for homo-interactions between RR proteins. The<br />

spectacular success <strong>of</strong> this approach illustrates the effectiveness <strong>of</strong> this combination approach<br />

in identifying direct interaction positions based on sequence information alone. We expect this<br />

method to be applicable for predicting interaction surfaces between proteins present in only one<br />

copy per genome as the number <strong>of</strong> sequenced genomes continues to expand, and for<br />

assembling multi-protein structures.<br />

5


<strong>BLAST</strong> X Mon. Morning Session<br />

INTEGRATED CONTROL OF CAULOBACTER CELL ENVELOPE PHYSIOLOGY BY A<br />

HYBRID TWO-COMPONENT/ECF SIGMA FACTOR SIGNALING NETWORK<br />

Robert Foreman, Erin Purcell, Aretha Fiebig, Dan Siegal-Gaskins & Sean Crosson<br />

Department <strong>of</strong> Biochemistry and Molecular Biology, <strong>University</strong> <strong>of</strong> Chicago, 929 E. 57th St.,<br />

Chicago, IL 60637<br />

We present evidence that Caulobacter crescentus encodes a regulatory network that<br />

integrates information about two different signals, visible light and oxidative/osmotic stress, to<br />

regulate the cell envelope and cell adhesion. In this hybrid signaling system, light signals via<br />

the LovK histidine kinase and oxidative/osmotic stress signals via the ECF sigma factor σ T are<br />

integrated to regulate cell envelope physiology. Caulobacter LovK, exhibits light-controlled<br />

autokinase activity and forms a two-component signaling system with the single-domain<br />

receiver protein, LovR. We have shown that the LovK/LovR system can function to modulate<br />

cell adhesion in response to blue light. LovK/LovR is a negative regulator <strong>of</strong> σ T , an envelope<br />

stress sigma factor that is critical for cell survival under osmotic and oxidative stress. σ T , in turn,<br />

is a positive transcriptional regulator <strong>of</strong> the lovK/lovR two-component system. This feedbackregulated<br />

signaling network can serve as a model to probe how bacterial cells integrate and<br />

coordinate their responses to multiple environmental queues.<br />

6


<strong>BLAST</strong> X Mon. Morning Session<br />

THE ROLE OF SIGNAL TRANSDUCTION IN CELL WALL METABOLISM IN BACILLUS<br />

SUBTILIS<br />

Paola Bisicchia, David, Noone, Efthimia Lioliou and Kevin M Devine.<br />

Smurfit Institute <strong>of</strong> Genetics, Trinity College Dublin, Dublin 2. Ireland.<br />

The cell wall <strong>of</strong> Gram positive bacteria is an extracellular structure, physically removed<br />

from the biosynthesis <strong>of</strong> the precursors used in its synthesis. Peptidoglycan and teichoic acid<br />

precursors are synthesized within the cytoplasm and transported across the cytoplasmic<br />

membrane where they are incorporated into the cell wall during growth and cell division. The<br />

spatial separation <strong>of</strong> these processes implies bidirectional signaling between the cell wall and<br />

cytoplasmic compartments, that recent work has begun to elucidate. We have shown that the<br />

essential YycFG two-component signal transduction system <strong>of</strong> Bacillus subtilis controls cell wall<br />

metabolism – during exponential growth, it activates expression <strong>of</strong> the YocH, YvcE and LytE<br />

autolysins and represses expression <strong>of</strong> YoeB (IseA), an inhibitor <strong>of</strong> autolysin activity, and YjeA a<br />

peptidoglycan deacetylase whose activity on peptidoglycan modulates its susceptibility to<br />

autolysin digestion (Howell et al., 2003; Bisicchia et al., 2007; Salzberg and Helmann, 2007;<br />

Yamomoto et al., 2008). Thus we propose that YycG senses some aspect(s) <strong>of</strong> the cell wall<br />

externally, perhaps the Lipid II intermediate, and transduces this information into the cytoplasm<br />

so that the cell wall synthetic activities in these two compartments are coordinated (Dubrac et<br />

al., 2008). We have also demonstrated a close connection between YycFG and the PhoPR<br />

two-component system that controls one <strong>of</strong> the phosphate limitation responses in B. subtilis<br />

(Hulett, 2002). YycFG and PhoPR are closely related phylogenetically - hybrid YycF’-‘PhoP and<br />

PhoP’-‘YycF response regulators are functional and there are similarities in the YycF and PhoP<br />

DNA binding sequences. We have also shown (i) that while YycG can phosphorylate only its<br />

cognate response regulator YycF, PhoR can phosphorylate both PhoP and YycF and (ii) that<br />

cells depleted for YycFG cannot mount a normal PhoPR-mediated phosphate limitation<br />

response. From these observations, we postulated that the roles <strong>of</strong> YycFG and PhoPR might<br />

be linked during cell wall metabolism and phosphate limitation.<br />

In this talk we will present the results <strong>of</strong> further analysis on the relationships between the<br />

YycFG and PhoPR two-component systems and their roles in cell wall metabolism during<br />

growth and phosphate limitation.<br />

References:<br />

Bisicchia et al., (2007) Molecular Microbiology 65: 180-200.<br />

Dubrac et al., (2008) Molecular Microbiology In Press<br />

Howell et al., (2003) Molecular Microbiology 49: 1639-1655.<br />

Howell et al., (2006) Molecular Microbiology 59:1199-1215.<br />

Hulett, (2002) The Pho regulon. in ‘B. subtilis and is closest relatives’ ASM Press.<br />

Salzberg and Helmann (2007) J. Bacteriology 189: 4671-4680.<br />

Yamamoto et al., (2008) Molecular Microbiology 70: 168-182.<br />

7


<strong>BLAST</strong> X Mon. Morning Session<br />

THE WalK/WalR ESSENTIAL SIGNAL TRANSDUCTION PATHWAY AND CELL WALL<br />

HOMEOSTASIS IN STAPHYLOCOCCUS AUREUS<br />

Sarah Dubrac 1 , Aurélia Delauné 1 , Olivier Poupel 1 , Adeline Mallet 2 , Tarek Msadek 1<br />

Biology <strong>of</strong> Gram-positive Pathogens 1 , Department <strong>of</strong> Microbiology, Plateforme de Microscopie<br />

Ultrastructurale 2 , Imagopole, Institut Pasteur, 25 rue du Dr. Roux, 75015 Paris, France<br />

The highly conserved WalK/WalR (aka YycG/YycF) two-component system is specific to<br />

low G+C % Gram-positive bacteria. While this system is essential for cell viability, both the<br />

nature <strong>of</strong> its regulon and its physiological role remained mostly uncharacterized. We have<br />

recently shown that the S. aureus WalKR system positively controls autolytic activity, in<br />

particular that <strong>of</strong> the major autolysins, AtlA, LytM and Sle1, and identified at least ten genes<br />

belonging to the WalKR regulon that are known or thought to be involved in S. aureus cell wall<br />

degradation. While none <strong>of</strong> these genes appears to be essential, we have shown that their<br />

global regulation by the WalKR system results in a drastic down regulation <strong>of</strong> cell wall dynamics,<br />

with a complete arrest <strong>of</strong> both cell wall biosynthesis and turn over under WalKR depletion. As a<br />

consequence <strong>of</strong> these molecular disorders TEM observations revealed that the cell wall <strong>of</strong><br />

WalKR depleted cells was significantly thicker and division septa were abnormally distributed.<br />

Recent advances presented here have shown that this global regulation is directly linked to<br />

WalKR essentiality. While the walRK genes are essential, the WalKR system is inducible since<br />

it is assumed that the WalR response regulator is only active when phosphorylated. While the<br />

activation signal is still unknown, several recent results suggest that it could be related to cell<br />

wall homeostasis.<br />

As cell wall metabolism is a major parameter influencing virulence and particularly the<br />

innate immune response, we are now interested in characterizing the impact <strong>of</strong> WalKR on<br />

S. aureus virulence. Beyond the regulation <strong>of</strong> genes involved in cell wall metabolism, we have<br />

also shown that the WalKR system activates expression <strong>of</strong> at least two genes involved in<br />

interactions with the extracellular host matrix and influences the capacity <strong>of</strong> S. aureus to adhere<br />

to the host matrix.<br />

References:<br />

1. Dubrac, S., I. G. Boneca, O. Poupel, and T. Msadek. 2007. New insights into the<br />

WalK/WalR (YycG/YycF) essential signal transduction pathway reveal a major role in<br />

controlling cell wall metabolism and bi<strong>of</strong>ilm formation in Staphylococcus aureus. J.<br />

Bacteriol. 189:8257-69.<br />

2. Dubrac, S., and T. Msadek. 2008. Tearing down the wall: peptidoglycan metabolism<br />

and the WalK/WalR (YycG/YycF) essential two-component system. Adv. Exp. Med. Biol.<br />

631:214-28.<br />

3. Dubrac, S., P. Bisicchia, K. M. Devine and T. Msadek. 2008. A matter <strong>of</strong> life and<br />

death: cell wall homeostasis and the WalKR (YycGF) essential signal transduction<br />

pathway. Mol. Microbiol. 70: (in press)<br />

8


<strong>BLAST</strong> X Mon. Evening Session<br />

DYNAMIC ASSEMBLY AND DISASSEMBLY OF THE TYPE IV MOLECULAR MACHINE<br />

Iryna Bulyha and Lotte Søgaard-Andersen<br />

Max Planck Institute for Terrestrial Microbiology, D-35043 Marburg, Germany<br />

Myxococcus xanthus harbors two gliding motility systems, A and S. The S(ocial)-system<br />

depends on type IV pili (T4P) and is generally active only when the cells are within contact<br />

distance <strong>of</strong> each other. T4P undergo cycles <strong>of</strong> assembly and retraction. The force for S-motility<br />

is generated by retraction <strong>of</strong> T4P. T4P are localized in a unipolar pattern and are present only at<br />

the leading pole <strong>of</strong> the rod-shaped cells. M. xanthus cells periodically undergo cellular reversals<br />

in which the old leading pole (which harbors T4P) becomes the new lagging pole (which does<br />

not harbor T4P). These observations suggest that in parallel with a cellular reversal, the pole at<br />

which T4P assemble switches. The molecular mechanisms regulating the T4P<br />

extension/retraction cycle and underlying T4P pole switching remain unknown.<br />

To investigate these mechanisms, we focused on the cellular localization <strong>of</strong> five highly<br />

conserved T4P biogenesis proteins (PilB, PilT, PilM, PilC and PilQ), which are present in all<br />

known T4P systems. PilB and PilT are cytoplasmic proteins and members <strong>of</strong> the secretion<br />

ATPase superfamily <strong>of</strong> proteins; PilB is required for assembly <strong>of</strong> T4P, while PilT is necessary for<br />

T4P retraction. PilM shows similarity to MreB/FtsA and is indispensable for T4P assembly. PilC<br />

is an inner membrane protein and is necessary for T4P assembly. The PilQ secretin forms a<br />

gated oligomeric channel for the pilus in the outer membrane. Using immuno-fluorescence<br />

microscopy, and time-lapse fluorescence microscopy with functional YFP-fusions, we show that<br />

PilQ, PilC and PilM are localized in clusters at both cell poles. These clusters have equal<br />

intensities and they do not oscillate between the two poles during reversals. The analysis <strong>of</strong> PilB<br />

and PilT localization revealed that both proteins are localized in polar clusters. PilB is<br />

predominantly localized in a cluster at the leading pole and PilT is predominantly localized in a<br />

cluster at the lagging cell pole. This localization is dynamic and the two proteins oscillate<br />

between the poles during reversals. These observations show that T4P function depends on<br />

two sets <strong>of</strong> proteins: one set is statically localized at both cell poles, and the other set is<br />

dynamically localized. Based on these findings we suggest that during cellular reversals, the<br />

T4P machinery is disassembled at the old leading pole and reassembled at the new leading<br />

pole.<br />

Moreover, we will present the data which suggest that the T4P assembly/retraction cycle<br />

relies on a PilB/PilT competition-based mechanism. According to this model, PilB at the leading<br />

cell pole stimulates T4P assembly, and the occasional accumulation <strong>of</strong> PilT at the leading cell<br />

pole results in retraction. Therefore, the two dynamically localized ATPases determine whether<br />

assembly or disassembly <strong>of</strong> pilus takes place.<br />

9


<strong>BLAST</strong> X Mon. Evening Session<br />

A "FOUR COMPONENT" SIGNAL TRANSDUCTION SYSTEM REGULATES<br />

DEVELOPMENTAL PROGRESSION IN MYXOCOCCUS XANTHUS<br />

Sakthimala Jagadeesan, Bongsoo Lee, and Penelope I. Higgs<br />

Department <strong>of</strong> Ecophysiology, Max Planck Institute for Terrestrial Microbiology, D35043<br />

Marburg, Germany<br />

Myxococcus xanthus responds to starvation by entering a multicellular developmental<br />

program in which 10 5 cells first aggregate into mounds and then within these mounds,<br />

differentiate into environmentally resistant spores. Under standard laboratory conditions,<br />

formation <strong>of</strong> spores within the mounds (fruiting bodies) takes approximately 72 hours. We have<br />

previously demonstrated that progression through the developmental program appears to be<br />

regulated by an atypical two component signal (TCS) transduction system consisting <strong>of</strong> four<br />

TCS homologs (RedC, RedD, RedE, and RedF). While RedC appears to be a typical<br />

membrane bound histidine kinase, RedD consists solely <strong>of</strong> two receiver domains. RedE is a<br />

soluble histidine kinase-like protein, and RedF is a single receiver domain response regulator.<br />

Based on a combination <strong>of</strong> genetic and biochemical analyses, we propose a model for how<br />

these four Red proteins function together to regulate progression through the developmental<br />

program. Our data suggests that development is repressed when the RedC histidine kinase<br />

phosphorylates RedF, a single domain response regulator. Developmental repression is<br />

relieved when, in response to an unknown signal(s), RedC is instead induced to phosphorylate<br />

the response regulator RedD. Surprisingly, the phosphoryl group is then transferred from RedD<br />

to the histidine kinase-like protein, RedE. RedE is then likely made accessible to RedF-P,<br />

whereupon it removes RedF’s phosphoryl group. We present the data that supports this model.<br />

Furthermore, we will address how progression through the developmental program is modulated<br />

by the Red system.<br />

10


<strong>BLAST</strong> X Mon. Evening Session<br />

INDEPENDENCE AND INTERDEPENDENCE OF Dif AND Frz CHEMOSENSORY<br />

PATHWAYS IN MYXOCOCCUS XANTHUS CHEMOTAXIS<br />

Qian Xu, Wesley P. Black, Christena L. Cadieux and Zhaomin Yang<br />

Biological Sciences, Virginia Tech, Blacksburg, Virginia 24061-0910, USA<br />

Dif and Frz, two Myxococcus xanthus chemosensory pathways, are required in<br />

phosphatidylethanolamine (PE) chemotaxis for excitation and adaptation, respectively. DifA and<br />

FrzCD, the homologs <strong>of</strong> methyl-accepting chemoreceptors in the two pathways, were examined<br />

for methylation in the context <strong>of</strong> chemotaxis and inter-pathway interactions. Evidence indicates<br />

that DifA may not undergo methylation but signals transmitting through DifA do modulate FrzCD<br />

methylation. Results also revealed that M. xanthus possesses Dif-dependent and Difindependent<br />

PE sensing mechanisms. Previous studies showed that FrzCD methylation is<br />

decreased by negative chemostimuli but increased by attractants such as PE. Results here<br />

demonstrate that the Dif-dependent sensory mechanism suppresses the increase in FrzCD<br />

methylation in attractant response and elevates FrzCD methylation upon negative stimulation. In<br />

other words, FrzCD methylation is governed by opposing forces from Dif-dependent and Difindependent<br />

sensing mechanisms. We propose that the Dif-independent but Frz-dependent PE<br />

sensing leads to increases in FrzCD methylation and subsequent adaptation, while the Difdependent<br />

PE signaling suppresses or diminishes the increase in FrzCD methylation to<br />

decelerate or delay adaptation. We contend that these antagonistic interactions are crucial for<br />

effective chemotaxis in this gliding bacterium to ensure that adaptation does not occur too<br />

quickly relative to the slow speed <strong>of</strong> M. xanthus movement.<br />

11


<strong>BLAST</strong> X Mon. Evening Session<br />

PREDATAXIS BEHAVIOR IN MYXOCOCCUS XANTHUS<br />

Jeb Berleman, Jodie Scott, Tatiana Chumley, and John R. Kirby *<br />

<strong>University</strong> <strong>of</strong> Iowa, Department <strong>of</strong> Microbiology, Iowa City, IA 52246, USA<br />

Spatial organization <strong>of</strong> cells is important for both multicellular development and tactic<br />

responses to a changing environment. We find that the slow-moving, gliding bacterium,<br />

Myxococcus xanthus, utilizes a Che-like pathway to regulate multicellular rippling during<br />

predation <strong>of</strong> other microbial species. Tracking <strong>of</strong> GFP-labeled cells indicates directed<br />

movement <strong>of</strong> M. xanthus cells during the formation <strong>of</strong> rippling wave structures. Quantitative<br />

analysis <strong>of</strong> rippling indicates that ripple wavelength is adaptable and dependent on prey cell<br />

availability. Methylation <strong>of</strong> the receptor, FrzCD, is required for this adaptation: a frzF<br />

methyltransferase mutant is unable to construct ripples, whereas a frzG methylesterase mutant<br />

forms numerous, tightly packed ripples. Both the frzF and frzG mutant strains are defective in<br />

directing cell movement through prey colonies. These data indicate that the transition to an<br />

organized multicellular state during predation in M. xanthus relies on the tactic behavior <strong>of</strong><br />

individual cells, mediated by a Che-like signal transduction pathway. Predataxis behavior differs<br />

from chemotaxis behavior in that it seems to depend heavily on tactile-stimulation, as opposed<br />

to chemical-stimulation.<br />

12


<strong>BLAST</strong> X Mon. Evening Session<br />

DYNAMIC LOCALIZATION OF FrzCD IN MYXOCOCCUS XANTHUS<br />

Emilia M.F. Mauriello 1 , David P. Astling 1 , Oleksii Sliusarenko 2 and David R. Zusman 1<br />

<strong>University</strong> <strong>of</strong> California, Department <strong>of</strong> Molecular and Cell Biology, Berkeley, CA 94720, USA 1 ;<br />

Yale <strong>University</strong>, Department <strong>of</strong> Molecular, Cell and Developmental Biology New Haven, CT<br />

06520, USA 2<br />

Directional motility in the gliding bacterium Myxococcus xanthus requires controlled cell<br />

reversals mediated by the Frz chemosensory system. FrzCD, a cytoplasmic chemoreceptor,<br />

does not form membrane bound polar clusters typical for most bacteria, but rather cytoplasmic<br />

clusters that are helically arranged and span the cell length. This unusual localization is<br />

maintained in the absence <strong>of</strong> the CheA homologs FrzE or CheA4, and the CheW homologs<br />

FrzA or FrzB. In contrast, MCPs lose their respective polar or cytoplasmic localization in<br />

Escherichia coli and Rhodobacter spheroides strains lacking CheA and/or CheW (1, 2). The<br />

distribution <strong>of</strong> FrzCD in living cells was found to be dynamic: FrzCD was localized in clusters<br />

that continuously changed their size, number, and position. The number <strong>of</strong> FrzCD clusters was<br />

correlated with cellular reversal frequency: fewer clusters were observed in hypo-reversing<br />

mutants and additional clusters observed in hyper-reversing mutants. When moving cells made<br />

side-to-side contacts, FrzCD clusters in adjacent cells showed transient alignments. These<br />

events were frequently followed by one <strong>of</strong> the interacting cells reversing. These observations<br />

suggest that FrzCD detects signals from a cell-contact sensitive signaling system and then relocalizes<br />

as it directs reversals to distributed motility engines.<br />

References:<br />

1. Maddock, J.R., and Shapiro, L. (1993) Science 259: 1717-1723.<br />

2. Wadhams, G.H., Martin, A.C., Warren, A.V., and Armitage, J.P. (2005) Mol Microbiol 58(3):<br />

895-902.<br />

3. Bustamante, V.H., Martínez-Flores, I., Vlamakis H.C., and Zusman, D. (2004) Mol Microbiol<br />

58(5): 1501-1513.<br />

13


<strong>BLAST</strong> X Mon. Evening Session<br />

IDENTIFYING NOVEL BACTERIAL CYTOSKELETAL ELEMENTS AND CYTOSKELETAL<br />

INTERACTORS THROUGH HIGH-THROUGHPUT IMAGING<br />

John Werner, Michael Ingerson-Mahar, Jonathan Guberman, and Zemer Gitai<br />

Princeton <strong>University</strong>, Department <strong>of</strong> Molecular Biology, LTL-355 Washington Rd., Princeton, NJ<br />

08540<br />

Bacterial cytoskeletal proteins polymerize into filamentous structures that represent key<br />

regulators <strong>of</strong> a wide array <strong>of</strong> cellular processes, including cell shape determination, cell division,<br />

and cell polarity. While bacterial homologs <strong>of</strong> all three major eukaryotic cytoskeletal families<br />

have already been described, the upstream regulators <strong>of</strong> bacterial cytoskeletal assembly and<br />

downstream effectors <strong>of</strong> cytoskeletal function remain poorly understood. In addition, recent<br />

studies have suggested that additional filament-forming proteins remain uncharacterized.<br />

All known cytoskeletal elements in both bacteria and eukaryotes have distinct nonuniform<br />

subcellular distributions. Focusing on the asymmetric bacterium, Caulobacter<br />

crescentus, we thus employed a directed high-throughput imaging approach to identify both<br />

novel bacterial cytoskeletal proteins and proteins that act upstream or downstream <strong>of</strong> the<br />

previously-characterized cytoskeletons. We developed a pipeline <strong>of</strong> high-throughput methods<br />

for generating fluorescent protein fusions, expressing them in Caulobacter, imaging their<br />

subcellular distribution at high resolution, and quantitating the resulting imaging data. By using<br />

this pipeline to analyze over 2,800 Caulobacter proteins as both N-and C-terminal mCherry<br />

fusions, we identified a set <strong>of</strong> ~300 localized Caulobacter proteins. This set included the three<br />

known Caulobacter cytoskeletons (the MreB actin homolog, FtsZ tubulin homolog, and<br />

Crescentin intermediate-filament). We also identified a novel protein that localizes to a tight line<br />

that hugs a short region <strong>of</strong> the inner curvature <strong>of</strong> Caulobacter cells. This protein may polymerize<br />

on its own, as it is capable <strong>of</strong> forming linear structures when expressed in heterologous systems<br />

such as E. coli or S. pombe. Preliminary studies suggests that this previously-uncharacterized<br />

cytoskeletal element plays a role in cell shape determination and may exhibit crosstalk with<br />

other cytoskeletal proteins.<br />

To identify upstream regulators <strong>of</strong> cytoskeletal assembly, we modified our pipeline to<br />

allow us to assay the effects <strong>of</strong> overexpressing genes <strong>of</strong> interest on the localization patterns <strong>of</strong><br />

MreB, FtsZ, and Crescentin. A pilot screen <strong>of</strong> ~200 conserved proteins with no known function<br />

identified four proteins that perturbed FtsZ, one that perturbed MreB, one that perturbed both<br />

MreB and FtsZ, and one that perturbed Crescentin. The functions and mechanisms <strong>of</strong> action <strong>of</strong><br />

these candidate cytoskeletal regulators are currently being examined. Finally, to identify<br />

downstream cytoskeletal effectors, we imaged our library <strong>of</strong> localized Caulobacter proteins in<br />

the presence <strong>of</strong> the MreB-delocalizing compound, A22. We found a large number <strong>of</strong> proteins<br />

that are either delocalized or mislocalized by A22 and are currently determining the cellular<br />

functions <strong>of</strong> these proteins as well as the nature <strong>of</strong> their direct or indirect associations with<br />

MreB.<br />

14


<strong>BLAST</strong> X Tue. Morning Session<br />

THE ROLE OF POSITIVE FEEDBACK IN CONTROLLING FLAGELLA ASSEMBLY<br />

DYNAMICS<br />

Supreet Saini 1 , Christy Aldridge 2 , Jonathan Brown 2 , Philip Aldridge 2 , Christopher Rao 1<br />

1<br />

Department <strong>of</strong> Chemical and Biomolecular Engineering, <strong>University</strong> <strong>of</strong> Illinois, Urbana, IL 61801,<br />

United States<br />

2<br />

Institute for Cell and Molecular Biosciences, Newcastle <strong>University</strong>, Framlington Place,<br />

Newcastle upon Tyne NE2 4HH, United Kingdom<br />

Flagellar assembly in Salmonella enterica serovar Typhimurium (S. typhimurium)<br />

proceeds in a sequential manner, starting from the base <strong>of</strong> the flagella and concluding at the<br />

filament tip. A key regulatory step in the assembly process is the σ 28 -FlgM checkpoint, which<br />

prevents the activation <strong>of</strong> σ 28 -dependent Pclass3 promoters prior to completion <strong>of</strong> the hook basal<br />

body. This regulatory checkpoint is typically assumed to involve a binary decision process:<br />

either proceed with Pclass3 gene expression or not, depending on the state <strong>of</strong> assembly.<br />

However, mathematical modeling suggests that this binary checkpoint may in fact result in more<br />

subtle, rheostat-like control. σ 28 is involved in a positive feedback loop, as ιτ positively regulates<br />

its own expression along with the expression <strong>of</strong> FliZ, an activator <strong>of</strong> Pclass2 gene expression. In<br />

addition, the ability <strong>of</strong> σ 28 to regulate gene expression depends on the concentration <strong>of</strong> FlgM in<br />

the cell. This suggests that the response <strong>of</strong> the σ 28 positive feedback loop is tuned by late<br />

protein secretion. In fact, our modeling and experimental results suggests that σ 28 and FlgM are<br />

not only involved in establishing the binary checkpoint between Pclass2 and Pclass3 gene<br />

expression but are also involved in fine tuning the relative timing <strong>of</strong> expression. In particular, the<br />

positive feedback loop involving σ 28 and FliZ establishes the delay between Pclass2 and Pclass3<br />

gene expression.<br />

In this talk, we will discuss our recent work investigating the positive feedback loops<br />

involving σ 28 and FliZ. We have recently shown that FliZ is an FlhD4C2-dependent activator <strong>of</strong><br />

Pclass2 gene expression. In addition, our results indicate the FliZ speeds up the induction <strong>of</strong> Pclass2<br />

genes in a secretion-dependent manner. With regards to σ 28 , we have found that autoregulation<br />

controls the relative timing <strong>of</strong> gene expression. In cells lacking FlgM, both Pclass2 and Pclass3<br />

genes are induced at the same times. Conversely, when the rate <strong>of</strong> FlgM secretion is reduced,<br />

the delay between Pclass2 and Pclass3 gene expression is exaggerated. These results suggest that<br />

timing is responsive to the rate <strong>of</strong> late protein secretion. Moreover, mathematical modeling<br />

predicts that this control is due to autoregulation by σ 28 . To test this model, we have rewired the<br />

flagellar gene circuit by replacing the native PfliA promoter with Pclass1/Pclass2/Pclass3 promoters.<br />

Consistent with the model predictions, these promoter replacement experiments show that<br />

autoregulation plays a key role in enforcing the timing <strong>of</strong> flagellar gene expression. Last, we also<br />

investigated gene expression dynamics at single-cell resolution, and our preliminary results<br />

suggest that the dynamics may exhibit bistability, consistent with control involving feedback.<br />

Collectively, our results suggest that the regulation <strong>of</strong> flagellar gene expression is complex and<br />

involves multiple layers <strong>of</strong> control.<br />

15


<strong>BLAST</strong> X Tue. Morning Session<br />

CRYSTAL STRUCTURE OF FLIT, A BACTERIAL FLAGELLAR EXPORT CHAPERONE FOR<br />

THE FILAMENT CAP PROTREN HAP2 (FliD)<br />

Katsumi Imada, Tohru Minamino, Miki Kinoshita and Keiichi Namba<br />

Graduate School <strong>of</strong> Frontier Biosciences, Osaka <strong>University</strong>, 1-3 Yamadaoka, Suita, Osaka 565-<br />

0871, Japan<br />

The bacterial flagellum is a filamentous organelle responsible for motility. Since the<br />

flagellum extends from the cytoplasm to the cell exterior, the external component proteins have<br />

to be exported from the cytoplasm. The protein subunits are exported by the flagellar specific<br />

export apparatus, which is a member <strong>of</strong> the type III secretion system. The export apparatus is<br />

believed to be located within the C-ring <strong>of</strong> the flagellar basal body and consists <strong>of</strong> at least six<br />

integral membrane proteins (FlhA, FlhB, FliO, FliP, FliQ, FliR) and three soluble proteins (FliH,<br />

FliI, FliJ). In addition to these proteins, other cytoplasmic proteins (FlgN, FliA, FliS, FliT) act as<br />

substrate-specific chaperons that facilitate the export <strong>of</strong> their substrates.<br />

FliT is a flagellar export chaperone for FliD (HAP2), which forms a capping complex at<br />

the distal end <strong>of</strong> the flagellar filament and promotes incorporation <strong>of</strong> flagellin subunits into the<br />

growing filament, and prevents FliD from premature aggregation in the cytoplasm. FliT is not<br />

only involved in protein export but also in regulation <strong>of</strong> flagellar gene expression. FliT negatively<br />

regulates transcription <strong>of</strong> the flagellar class 2 operons by binding to FlhD4C2 complex, which is a<br />

transcriptional activator.<br />

We have determined a crystal structure <strong>of</strong> FliT at 3.2 Å resolution. The structure and<br />

following genetic and biochemical studies have revealed that the C-terminal region <strong>of</strong> FliT<br />

regulates its interactions with other flagellar proteins. We will discuss the molecular mechanisms<br />

<strong>of</strong> protein export and gene expression based on the FliT structure.<br />

16


<strong>BLAST</strong> X Tue. Morning Session<br />

STRUCTURAL INSIGHT INTO ACTIVE FLAGELLAR MOTOR FORMATION THROUGH THE<br />

PERIPLASMIC REGION OF MOTB<br />

Seiji Kojima 1 , Mayuko Sakuma 1 , Yuki Sudo 1 , Chojiro Kojima 2 , Tohru Minamino 3 , Keiichi<br />

Namba 3 , Michio Homma 1 and Katsumi Imada 3*<br />

1 Division <strong>of</strong> Biological Science, Graduate School <strong>of</strong> Science, Nagoya <strong>University</strong>, Chikusa-Ku,<br />

Nagoya 464-8602, Japan; 2 Laboratory <strong>of</strong> Biophysics, Graduate School <strong>of</strong> Biological Sciences,<br />

Nara Institute <strong>of</strong> Science and Technology (NAIST), 8916-5 Takayama, Ikoma, Nara, 630-0192<br />

Japan; 3 Graduate School <strong>of</strong> Frontier Biosciences, Osaka <strong>University</strong>, 1-3 Yamadaoka, Suita,<br />

Osaka 565-0871, Japan<br />

Bacterial flagellar motor is a supramolecular nano-machine powered by the<br />

transmembrane gradient <strong>of</strong> protons or sodium ions, and spins flagellar filaments to drive cell<br />

motility. Motor torque is generated by the rotor-stator interaction that is coupled with ion flow<br />

through the ion-channel in the stator unit, which is composed <strong>of</strong> four MotA and two MotB<br />

subunits. About ten stators are assembled around the perimeter <strong>of</strong> the rotor and anchored to<br />

peptidoglycan layer by the peptidoglycan-binding (PGB) domain <strong>of</strong> MotB. The rotor-stator<br />

assembly is not a rigid complex, so each stator unit can dynamically be exchanged in the<br />

functional motor, and is activated only when assembled around the rotor. However, the<br />

mechanisms <strong>of</strong> the stator assembly and the activation <strong>of</strong> the proton flow remain unclear. Here,<br />

we report the crystal structure <strong>of</strong> a C-terminal fragment <strong>of</strong> MotB (MotBC), which includes the<br />

PGB domain and covers the whole periplasmic region essential for cell motility (PEM), at 1.75 Å<br />

resolution. The structure, and subsequent mutational and biochemical analyses indicate that<br />

dimer formation by the PGB domains is required for motility through the regulation <strong>of</strong> the<br />

arrangement <strong>of</strong> the transmembrane segment. Moreover, we show that large structural changes<br />

in the N-terminal helices should be coupled with both peptidoglycan binding and activation <strong>of</strong><br />

the stator. This work provides novel structural insight into the dynamic behavior <strong>of</strong> the ionchannel<br />

complex regulated by the periplasmic domain, and the activation mechanism <strong>of</strong> the<br />

complex coupled with completion <strong>of</strong> the assembly where it works.<br />

17


<strong>BLAST</strong> X Tue. Morning Session<br />

STATOR SELECTION IN SHEWANELLA ONEIDENSIS MR-1<br />

Anja Paulick, Andrea Koerdt, Kai M. Thormann<br />

Department <strong>of</strong> Ecophysiology, MPI für terrestrische Mikrobiologie, Karl-von-Frisch-Strasse, D-<br />

35043 Marburg, Germany<br />

The Gram-negative metal-ion reducing bacterium Shewanella oneidensis MR-1 is motile<br />

by means <strong>of</strong> a single polar flagellum. We identified two potential stator systems, PomAB and<br />

MotAB, each sufficient as a force generator to drive flagellar rotation. Physiological studies<br />

demonstrate that PomAB is sodium-dependent while MotAB is powered by proton motive force.<br />

Homology comparisons strongly indicate that the MotAB system has been acquired by<br />

horizontal gene transfer, probably as a consequence <strong>of</strong> long-term adaptation from a marine to a<br />

low-sodium freshwater environment. As in S. oneidensis MR-1, a number <strong>of</strong> bacterial species<br />

possess more than one stator system to power a single flagellar system but it is yet unclear,<br />

how selection <strong>of</strong> the stators is achieved.<br />

Expression analysis at the single cell level showed that both stator systems <strong>of</strong> S.<br />

oneidensis MR-1 are expressed simultaneously, and functional fusions <strong>of</strong> PomB and MotB to<br />

mCherry revealed that both stator systems are present in the cell at the same time. While the<br />

Pom system is efficiently localizing to the flagellated cell pole under all conditions, the Mot stator<br />

is located in the cell membrane and only found at the cell pole at high abundance in media with<br />

low sodium content. At low sodium, both stator systems are localizing to the flagellated cell pole<br />

in the majority <strong>of</strong> the cell population, thus indicating that under such conditions a hybrid motor<br />

may be formed. We conclude that stator selection occurs at the level <strong>of</strong> protein localization by<br />

alterations in the localization efficiency in response to sodium levels.<br />

In Vibrio species, two additional proteins, MotX and MotY, are involved in stator<br />

recruitment and sodium-dependent swimming. We therefore analyzed whether S. oneidensis<br />

MR-1 orthologs to MotX and MotY play a role in stator selection. Mutant and localization<br />

analyses demonstrated that both proteins are required for function <strong>of</strong> the Pom as well as the<br />

Mot stator system. As opposed to the Vibrio system, in S. oneidensis MR-1, MotX and MotY are<br />

not required for stator recruitment and also do not play a role in stator selection in response to<br />

sodium conditions.<br />

18


<strong>BLAST</strong> X Tue. Morning Session<br />

TAKING CONTROL OF THE BACTERIAL FLAGELLAR MOTOR<br />

Mathieu Gauthier, Dany Truchon and Simon Rainville<br />

Department <strong>of</strong> Physics, Engineering Physics and Optics and Centre d'optique photonique et<br />

laser, Laval <strong>University</strong>, Québec, Québec, CANADA<br />

The bacterial flagellar motor is a fairly complex machine, requiring 40-50 genes for its<br />

expression, assembly and control. Furthermore, it is embedded in the multiple layers <strong>of</strong> the<br />

bacterial membrane. That explains why, unlike many other molecular motors, it has not yet been<br />

studied in vitro. As spectacular studies <strong>of</strong> linear motors (like kinesin, myosin and dynein) have<br />

clearly demonstrated, an in vitro system provides the essential control over experimental<br />

parameters to achieve the precise study <strong>of</strong> the motor’s physical and chemical characteristics.<br />

Here, we report significant progress towards the development <strong>of</strong> a unique in vitro system to<br />

study quantitatively the bacterial flagellar motor.<br />

Our system consists <strong>of</strong> a filamentous Escherichia coli bacterium partly introduced inside<br />

a micropipette. Femtosecond laser pulses (60 fs and ~ 15 nJ/pulse) are then tightly-focused on<br />

the part <strong>of</strong> the bacterium that is located inside the micropipette. This vaporizes a<br />

submicrometer-sized hole in the wall <strong>of</strong> the bacterium, thereby granting us access to the inside<br />

<strong>of</strong> the cell and the control over the proton-motive force that powers the motor. Using a patchclamp<br />

amplifier, we applied an external voltage between the inside and the outside <strong>of</strong> the<br />

micropipette. If the hole in the bacterium is open, that voltage should translate into a membrane<br />

potential powering the motors outside <strong>of</strong> the micropipette. As we varied the applied potential,<br />

variations in the motor's rotation speed were observed. For these preliminary results, the<br />

rotation speed was observed directly using video microscopy <strong>of</strong> fluorescently labeled filaments.<br />

That system opens numerous possibilities to study the flagellar motor and other membrane<br />

components.<br />

19


<strong>BLAST</strong> X Tue. Morning Session<br />

EXPERIMENTAL EVIDENCE FOR CONFORMATIONAL SPREAD IN THE BACTERIAL<br />

SWITCH COMPLEX<br />

Richard W. Branch 1 , Fan Bai 1 , Dan V. Nicolau 2 , Teuta Pilizota 1 , Bradley Steel 1 , Philip K.<br />

Maini 2 , Richard M. Berry 1<br />

1 Clarendon Laboratory, Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Oxford, Parks Road, Oxford OX1<br />

3PU, UK. 2 Centre for Mathematical Biology, Mathematical Institute, <strong>University</strong> <strong>of</strong> Oxford,<br />

St.Giles Road, Oxford OX1 3LB, UK.<br />

The bacterial switch complex in E. coli controls the direction <strong>of</strong> rotation <strong>of</strong> the bacterial<br />

flagellar motor between clockwise and counterclockwise modes. The complex takes the form <strong>of</strong><br />

a ring composed <strong>of</strong> about 110 FliN, 34 FliM and 26 FliG protein subunits. Regulation is through<br />

binding <strong>of</strong> the signaling molecule CheY-P to FliM. FliG interfaces with the torque-generating<br />

stator units <strong>of</strong> the motor. The precise mechanism by which the complex executes a switch is<br />

unclear.<br />

The complex displays the ultrasensitive nature typical <strong>of</strong> allosteric proteins, with a steep<br />

sigmoidal relationship existing between [CheY-P] and motor rotational bias. Allosteric regulation<br />

<strong>of</strong> proteins has classically been understood in terms <strong>of</strong> the Monod-Wyman-Changeux (MWC) or<br />

Koshland-Nemethy-Filmer (KNF) models. However, it is unrealistic that MWC-type concerted<br />

transitions could be responsible for quaternary conformational changes <strong>of</strong> such a large complex,<br />

and cooperative binding studies in vitro and in vivo have precluded a KNF-type induced-fit<br />

mechanism.<br />

The MWC and KNF models are recognized as limiting cases <strong>of</strong> a general allosteric<br />

scheme that has recently been described in a model <strong>of</strong> conformational spread. The model has<br />

been shown to be capable <strong>of</strong> reproducing motor switching kinetics. A directly observable<br />

consequence <strong>of</strong> conformational spread in the switch complex would be the variation <strong>of</strong> motor<br />

speed associated with the conformational spread <strong>of</strong> ring subunit state. In particular, the duration<br />

<strong>of</strong> switch events should be finite and broadly distributed due to the diffusive random walk <strong>of</strong><br />

conformational spread, and incomplete switches should be observed due to incomplete growth<br />

and shrinkage <strong>of</strong> subunit state domains.<br />

We have used high-resolution back-focal-plane interferometry <strong>of</strong> polystyrene beads<br />

attached to truncated WT E. coli flagella to resolve intermediate states <strong>of</strong> the motor predicted by<br />

conformational spread, and demonstrate detailed quantitative agreement between our<br />

measurements and conformational spread simulations. Individual switch events are not<br />

instantaneous, but follow a broad distribution <strong>of</strong> switch times with mean ~ 20 ms. The shortest<br />

switch events are observed to last less than 1ms, while the longest require over 100ms and take<br />

several revolutions to complete. Intervals between switches are exponentially distributed at all<br />

values <strong>of</strong> bias. Incomplete switches reaching a range <strong>of</strong> intermediate speeds are observed. The<br />

events are Poisson distributed in time with a bias-dependent frequency.<br />

20


<strong>BLAST</strong> X Tue. Morning Session<br />

DO INDIVIDUAL BACTERIAL FLAGELLAR MOTORS USE HYSTERESIS TO MAINTAIN A<br />

ROBUST OUTPUT IN A NOISY ENVIRONMENT? – AN EXPERIMENTAL STUDY<br />

Peter Reuven 1 , Oleg Krichevsky 2 , Michael Eisenbach 1<br />

1 Department <strong>of</strong> Biological Chemistry, Weizmann Institute <strong>of</strong> Science, 76 100, Israel<br />

2 Department <strong>of</strong> Physics, Ben-Gurion <strong>University</strong>, Beer Sheva, 84 105, Israel<br />

It is known that despite imperfections <strong>of</strong> intracellular environment, flagellar motor outputs<br />

are robust against stochastic fluctuations <strong>of</strong> CheY signal. Indirect evidence from our lab<br />

suggests that, to maintain stability, the motor complex might damp out fluctuations in the<br />

intracellular level <strong>of</strong> CheY by having a hysteresis feature - two different thresholds for switching.<br />

In this case, hysteresis means that the default-state motor switches at a higher<br />

threshold from counterclockwise to clockwise state compared to a lower threshold when<br />

switching back. Such behavior will produce hysteretic loop in input/output characteristics <strong>of</strong><br />

flagellar motor.<br />

Our aim is to pin-point the level at which the noise-filtering (via hysteresis) occurs in the<br />

chemotactic network. We are studying flagellar rotation <strong>of</strong> single cells as a function <strong>of</strong> their<br />

intracellular CheY-P concentration, changing the concentration up and down in order to cover<br />

both - counterclockwise to clockwise and clockwise to counterclockwise - switching routes.<br />

Since it is impossible to distinguish CheY from CheY-P in vivo, one has to work under<br />

conditions that maintain CheY constantly fully phosphorylated. The challenge was how to<br />

effectively decrease the concentration <strong>of</strong> the phosphorylated signal.<br />

Knowing that we cannot play with the level <strong>of</strong> phosphorylation we opt to play with the<br />

level <strong>of</strong> the CheY-P protein instead. While to increase protein concentration is a routine task, to<br />

decrease it is less obvious. We cloned a bi-modal, inducible plasmid expressing a CheY fused<br />

to yellow fluorescent protein (YFP) and ssrA degradation tag. YFP is used for quantifying<br />

the signal whereas the degradation tag makes it possible to shorten the lifetime <strong>of</strong> CheY-YFP.<br />

The core assumption <strong>of</strong> our approach is that heat-shock-induced proteases accelerate the<br />

degradation <strong>of</strong> the ssrA-targeted CheY-YFP protein. We have verified this assumption<br />

experimentally.<br />

We monitor the degradation <strong>of</strong> CheY-YFP by a decrease in fluorescence intensity and<br />

this decrease is correlated with the change <strong>of</strong> the direction <strong>of</strong> motor rotation. Input/output<br />

characteristics <strong>of</strong> individual flagellar motors is build from these correlations.<br />

Advancing this study will, hopefully, enable us to deeper understand how the<br />

mechanisms <strong>of</strong> intracellular interactions affect the logic <strong>of</strong> cell's behavior.<br />

21


<strong>BLAST</strong> X Tue. Morning Session<br />

DYNAMICS OF THE BACTERIAL FLAGELLAR MOTOR WITH MULTIPLE STATORS<br />

Giovanni Meacci and Yuhai Tu<br />

IBM T. J. Watson Research Center<br />

The bacterial flagellar motor drives the rotation <strong>of</strong> flagellar filaments and enables many<br />

species <strong>of</strong> bacteria to swim. Torque is generated by interaction <strong>of</strong> stator units, anchored to the<br />

peptidoglycan cell wall, with the rotor. Recent experiments [Yuan, J. & Berg, H. C. (2008) PNAS<br />

105, 1182-1185] show that near zero load the speed <strong>of</strong> the motor is independent <strong>of</strong> the number<br />

<strong>of</strong> stators. Here, we introduce a mathematical model <strong>of</strong> the motor dynamics that explains this<br />

behavior based on a general assumption that the stepping rate <strong>of</strong> a stator depends on the<br />

torque exerted by the stator on the rotor. We find that the motor dynamics can be characterized<br />

by two time scales: the moving-time interval for the mechanical rotation <strong>of</strong> the rotor and the<br />

waiting-time interval determined by the chemical transitions <strong>of</strong> the stators. We show that these<br />

two time scales depend differently on the load, and that their crossover provides the<br />

microscopic explanation for the existence <strong>of</strong> two regimes in the torque-speed curves observed<br />

experimentally. We also analyze the speed fluctuation for a single motor using our model. We<br />

show that the motion is smoothed by having more stator units. However, the mechanism for<br />

such fluctuation reduction is different depending on the load. We predict that the speed<br />

fluctuation is determined by the number <strong>of</strong> steps per revolution only at low load and is controlled<br />

by external noise for high load. Our model can be generalized to study other molecular motor<br />

systems with multiple power-generating units.<br />

22


<strong>BLAST</strong> X Tue. Evening Session<br />

EFFECT OF OSMOLYTES ON REGULATING THE ACTIVITIES OF THE SSK1 RESPONSE<br />

REGULATOR FROM SACCHAROMYCES CEREVISIAE<br />

Alla O. Kaserer, Paul F. Cook and Ann H. West<br />

Department <strong>of</strong> Chemistry and Biochemistry, The <strong>University</strong> <strong>of</strong> Oklahoma, 620 Parrington Oval,<br />

Norman, OK 73019<br />

The multi-step His-Asp phosphorelay system in Saccharomyces cerevisiae allows cells<br />

to adapt to osmotic, oxidative and other environmental stresses. The pathway consists <strong>of</strong> a<br />

hybrid histidine kinase SLN1, a histidine-containing phosphotransfer (HPt) protein YPD1 and<br />

two response regulator proteins, SSK1 and SKN7. Under non-osmotic stress conditions, the<br />

SLN1 kinase is active and phosphoryl groups are shuttled to SSK1 via YPD1. We have<br />

previously demonstrated that YPD1 stabilizes the phosphorylated form <strong>of</strong> SSK1. The cellular<br />

response to hyperosmotic stress involves rapid efflux <strong>of</strong> water and change in intracellular ion<br />

and osmolyte concentration. It is our hypothesis that these changes may affect rates <strong>of</strong><br />

phosphotransfer within the SLN1-YPD1-SSK1 phosphorelay system and the phosphorylated<br />

lifetime <strong>of</strong> response regulators. Therefore, we examined the effect <strong>of</strong> different solute<br />

concentrations on dephosphorylation <strong>of</strong> SSK1 and phosphotransfer rates within the<br />

phosphorelay system using half-life studies and rapid quench kinetics, respectively. These<br />

studies provide new insight and <strong>of</strong>fer a better understanding <strong>of</strong> how this His-Asp multi-step<br />

phosphorelay is environmentally regulated.<br />

23


<strong>BLAST</strong> X Tue. Evening Session<br />

REGULATION OF Escherichia coli MOTILITY BY THE NITRIC OXIDE SENSITIVE<br />

TRANSCRIPTIONAL REPRESSOR NsrR<br />

Jonathan D. Partridge and Stephen Spiro<br />

Department <strong>of</strong> Molecular and Cell Biology, The <strong>University</strong> <strong>of</strong> Texas at Dallas, 800 W. Campbell<br />

Road, Richardson, Texas 75080, USA<br />

There is circumstantial evidence implicating the water-soluble free radical nitric oxide<br />

(NO) as a regulator <strong>of</strong> motility, chemotaxis and bi<strong>of</strong>ilm development. Heme-containing NObinding<br />

domains <strong>of</strong> methyl accepting chemotaxis proteins from Clostridium botulinum and<br />

Thermoanaerobacter tengcongensis have been characterized, though the prediction that these<br />

proteins mediate taxis towards or away from NO has not been tested. In transcriptomics<br />

experiments, the expression <strong>of</strong> some motility genes has been observed to be perturbed by<br />

exposure <strong>of</strong> cultures to sources <strong>of</strong> NO or nitrosative stress (imposed by S-nitrosothiols),<br />

although both positive and negative responses have been reported, and the regulators involved<br />

were not identified. In the non-pathogenic organism Nitrosomonas europaea, NO stimulates<br />

bi<strong>of</strong>ilm formation. In Pseudomonas aeruginosa and Staphylococcus aureus, there is evidence<br />

that NO inhibits bi<strong>of</strong>ilm formation, or stimulates dispersal, and NO stimulates swimming and<br />

swarming motility in P. aeruginosa. In no case has a molecular mechanism been described<br />

which accounts for the effects <strong>of</strong> NO on bi<strong>of</strong>ilm development or motility.<br />

NO is made in bacteria either as a by-product <strong>of</strong> nitrite reduction to ammonia, or as an<br />

intermediate <strong>of</strong> denitrification, and is made by the inducible NO synthase <strong>of</strong> host phagocytes.<br />

Thus pathogenic bacteria can be exposed both to endogenously-generated NO, and to the NO<br />

made by host cells. In Escerichia coli, the regulatory proteins NorR and NsrR mediate adaptive<br />

responses to NO, by controlling the expression <strong>of</strong> genes encoding enzymes that reduce or<br />

oxidize NO to less toxic species. The key NO detoxifying activities are the flavohemoglobin<br />

(encoded by the hmp gene) and the flavorubredoxin (encoded by norVW), the expression <strong>of</strong><br />

which is regulated by NsrR and NorR, respectively. As far as is known, the norVW promoter is<br />

the sole target for regulation by NorR, while NsrR appears to control a large regulon <strong>of</strong> genes<br />

and operons.<br />

The extent <strong>of</strong> the NsrR regulon <strong>of</strong> E. coli has been assessed computationally, and by a<br />

transcriptomics analysis <strong>of</strong> a strain in which NsrR was titrated by the presence <strong>of</strong> multiple copies<br />

<strong>of</strong> a cloned NsrR binding site. We believe that neither approach has provided a comprehensive<br />

inventory <strong>of</strong> all <strong>of</strong> the genes regulated by NsrR. Therefore, we used chromatin<br />

immunoprecipitation and microarray analysis (ChIP-chip) to identify NsrR binding sites in the E.<br />

coli genome. Surprisingly, we found NsrR binding sites associated with the promoter regions <strong>of</strong><br />

three transcription units (mqsR-ygiT, fliAZY and fliLMNOPQR) containing genes with wellestablished<br />

or suspected roles in motility and/or bi<strong>of</strong>ilm development. We have confirmed that<br />

the fliA and fliL promoters are subject to regulation by NsrR and NO, and have identified an<br />

NsrR binding site in the fliA promoter. We have shown that NsrR is a negative regulator <strong>of</strong><br />

motility in both K12 and UPEC strains <strong>of</strong> E. coli. These results provide, for the first time, a<br />

molecular mechanism by which NO might control bacterial motility.<br />

24


<strong>BLAST</strong> X Tue. Evening Session<br />

SYNTHETIC LETHALITY UNCOVERS A NOVEL LINK BETWEEN THE MalT AND OmpR<br />

REGULONS<br />

Sylvia A. Reimann and Alan J. Wolfe<br />

Loyola <strong>University</strong> Chicago, Maywood, Il<br />

Synthetic lethality (SL) is a genetic term for the inviability <strong>of</strong> a double mutant combination<br />

<strong>of</strong> two fully viable single mutants. SL is commonly interpreted as redundancy at an essential<br />

metabolic step. However, a second, less well-known, class <strong>of</strong> SL exists: the so-called “defectdamage-repair”<br />

(DDR) cycles that link apparently unrelated metabolic pathways (Ting et al.,<br />

2008).<br />

We have isolated an SL mutant <strong>of</strong> the form ompR SL(ompR), where SL(ompR) refers to<br />

a mutation that causes death when present in a cell that lacks the two-component response<br />

regulator OmpR. This global regulator is required for proper assembly <strong>of</strong> the cell envelope and<br />

we reasoned that the nature <strong>of</strong> the SL(ompR) mutation would provide further insight into the role<br />

<strong>of</strong> OmpR and its regulon.<br />

Using a combined genetic/genomic approach, we mapped the SL(ompR) mutation to<br />

malT, which encodes the transcription factor MalT. Because overexpression <strong>of</strong> the MalT<br />

inhibitor MalK did not rescue growth, we proposed that the malT mutation leads to a<br />

constitutively active form (MalT c ). To test this hypothesis, we deleted the MalT-dependent lamB,<br />

which encodes an outer membrane porin, and found that this deletion suppressed the SL <strong>of</strong> the<br />

ompR malT c mutant.<br />

Since LamB and OmpR do not perform redundant functions, we propose that the<br />

observed SL is <strong>of</strong> the DDR variety, as follows: the defect (MalT c activity) leads to damage<br />

(constitutive expression <strong>of</strong> LamB) that is repaired by one or more members <strong>of</strong> the OmpR<br />

regulon. To further understand the role <strong>of</strong> OmpR, we are currently seeking OmpR regulon<br />

members that can suppress the SL <strong>of</strong> the ompR malT c mutant.<br />

25


<strong>BLAST</strong> X Tue. Evening Session<br />

A CHEMOTAXIS-LIKE SIGNALING PATHWAY REGULATES THE EXPRESSION OF<br />

EXTRACELLULAR MATERIALS IN GEOBACTER SULFURREDUCENS<br />

Hoa T. Tran 1 , Derek R. Lovley 2 , and Robert M. Weis 1<br />

Departments <strong>of</strong> Chemistry 1 and Microbiology 2 , <strong>University</strong> <strong>of</strong> Massachusetts, Amherst, MA<br />

01003, USA.<br />

The chemotaxis pathway that regulates cell motility toward chemical attractants is wellstudied<br />

in Escherichia coli. By contrast, multiple chemotaxis clusters have been found in many<br />

other bacteria, and evidence is accumulating that these cells use chemotaxis-like pathways to<br />

regulate diverse cellular functions. The genome <strong>of</strong> Geobacter sulfurreducens, a δ-<br />

Proteobacterium found predominantly in the Fe(III) reducing zone <strong>of</strong> sedimentary environment<br />

contains ~70 chemotaxis gene homologs arranged in 6 major clusters. Cluster 5 (Che5) has a<br />

complete set <strong>of</strong> chemotaxis homologs, including the kinase cheA (1 copy), cheW (2 copies),<br />

cheR (1), cheB (1), cheY (3) and four other non-che genes. There are 34 chemoreceptor<br />

homologs in the genome, but none is found in the Che5 cluster. Che5-type clusters have been<br />

identified in the genomes <strong>of</strong> several δ-Proteobacteria, yet their functions are not known. Here,<br />

we report that G. sulfurreducens Che5 cluster regulates gene expression, and in particular the<br />

synthesis <strong>of</strong> extracellular material that is abundant in OmcS and OmcZ, two c-type cytochromes<br />

bound to the outer membrane. OmcS and OmcZ are essential for cell growth in insoluble<br />

electron acceptors and for effective electricity production on electrodes. Deletion mutants <strong>of</strong> the<br />

homologs <strong>of</strong> cheW (gsu2218, gsu2220), cheA (gsu2222) and cheR (gsu2215) increased OmcS<br />

production and decreased OmcZ. In contrast, deletion mutants <strong>of</strong> cheB (gsu2214), one <strong>of</strong> the<br />

three cheYs (gsu2223) and a non-che gene (gsu2216) decreased OmcS and increased OmcZ<br />

production. The chemoreceptors that signal through these Che5 proteins are hypothesized to<br />

belong to a single MA class. Evidence that supports this idea is based on the phenotypes <strong>of</strong><br />

deletion mutants in two mcp genes (gsu1704, and gsu2372), which are similar the cheA mutant.<br />

Wherever the functional parallels can be drawn, the function <strong>of</strong> the homologues in Geobacter<br />

Che5 signaling pathway is similar to their counterpart in the E. coli chemotaxis pathway. In<br />

addition to changes in OmcS and OmcZ expression, the cheA, cheR and cheW (gsu2220)<br />

mutants promote cell aggregation and overproduce non-PilA filamentous material. Moreover,<br />

microarray data <strong>of</strong> the cheR mutant, and quantitative RT-PCR data from the other che mutants<br />

indicate that the Che5 cluster alter the expression <strong>of</strong> ~175 genes. A substantial fraction <strong>of</strong><br />

these are predicted to contain export signals that will result in an to extracellular location. Taken<br />

together, these data demonstrate that G. sulfurreducens Che5 cluster, with one class <strong>of</strong><br />

chemoreceptors, regulates the extracellular matrix material biosynthesis.<br />

This research was supported by the U.S. Department <strong>of</strong> Energy Office <strong>of</strong> Science (BER)<br />

under the Cooperative Agreement No. DE-FC02-02ER63446.<br />

26


<strong>BLAST</strong> X Tue. Evening Session<br />

THE TWO-COMPONENT REGULATORY SYSTEM BarA/SirA IS AT THE TOP OF A MULTI-<br />

FACTORIAL REGULATORY CASCADE CONTROLLING THE EXPRESSION OF THE SPI-1<br />

AND SPI-2 VIRULENCE REGULONS IN SALMONELLA<br />

Luary C. Martínez, José L. Puente and Víctor H. Bustamante<br />

Departamento de Microbiología Molecular, Instituto de Biotecnología, Universidad Nacional<br />

Autónoma de México. Cuernavaca, Mor., México.<br />

Horizontal gene transfer <strong>of</strong> pathogenicity islands has been a major event in the evolution<br />

<strong>of</strong> pathogenic bacteria. Integration <strong>of</strong> regulatory networks to control the expression <strong>of</strong> the gained<br />

genes has represented another essential step in this process. Salmonella Pathogenicity Islands<br />

1 and 2 (SPI-1 and SPI-2) are required at different phases during Salmonella infection in<br />

humans and animals. A positive regulatory cascade comprising the SPI-1-encoded regulators<br />

HilD, HilA and InvF induces expression <strong>of</strong> the SPI-1 regulon in response to conditions<br />

resembling the intestinal environment, such as growth in Luria-Bertani (LB) rich medium. Two<br />

global two-component regulatory systems, OmpR/EnvZ and PhoP/PhoQ, control the expression<br />

<strong>of</strong> the SsrA/B two-component system encoded within SPI-2, which specifically induces the<br />

expression <strong>of</strong> the SPI-2 regulon genes in response to conditions resembling the intracellular<br />

environment, mimicked in vitro by growth at low concentrations <strong>of</strong> phosphate and magnesium.<br />

Interestingly, we have recently shown that HilD also induces expression <strong>of</strong> the SsrA/B system,<br />

and thus <strong>of</strong> the SPI-2 regulon, at late stationary phase in LB cultures, indicating that SPI-2<br />

expression is also controlled by a SPI-1/SPI-2 cross-talk mechanism.<br />

The results presented here, together with previous reports, better define the complex<br />

and multi-factorial regulatory cascade that controls SPI-1 and SPI-2 expression through HilD.<br />

We show that the global two-component system BarA/SirA activates the transcription <strong>of</strong> two<br />

small RNA molecules, csrB and csrC. These molecules counteract the negative effect exerted<br />

by the CsrA RNA binding protein on hilD mRNA stability, ensuring the synthesis <strong>of</strong> the<br />

appropriate concentration <strong>of</strong> HilD required for the expression <strong>of</strong> HilA and SsrA/B, the central<br />

positive regulators <strong>of</strong> the SPI-1 and SPI-2 regulons, respectively. Furthermore, we<br />

demonstrated that growth conditions affecting SPI-1 expression in LB (e.g. low salt, acidic pH or<br />

temperatures below 37°C), similarly repress the expression <strong>of</strong> the SPI-2 regulon. However,<br />

while acidic pH seems to negatively regulate the regulatory cascade by affecting the activity <strong>of</strong><br />

the BarA sensor kinase, growth at low salt concentration or at low temperature seems to directly<br />

repress hilD expression through an as yet unidentified mechanism.<br />

27


<strong>BLAST</strong> X Tue. Evening Session<br />

INTERACTION OF THE TRANSCRIPTIONAL REGULATORY COMPLEX, FlhDC, WITH ITS<br />

TARGET DNA<br />

Yi-Ying Lee, and Philip Matsumura<br />

Department <strong>of</strong> Microbiology and Immunology, College <strong>of</strong> Medicine, <strong>University</strong> <strong>of</strong> Illinois at<br />

Chicago, 835 S. Wolcott Ave., M/ C 790, Chicago, Illinois 60612-7344<br />

The bacterial flagellum is the structure that allows bacteria to move and respond to<br />

nutritional and chemical signals in their environment. It is a complex suborganelle and the<br />

transcriptional regulation <strong>of</strong> the 40 plus structural genes is organized in a highly regulated<br />

cascade. At the top <strong>of</strong> the hierarchy is the master operon which codes for FlhD and FlhC. These<br />

two positive transcriptional regulators form a unique heteroheximeric complex which binds<br />

upstream <strong>of</strong> the -35 region and requires sigma 70 for transcription. This complex has an<br />

unusually large ‘footprint’ <strong>of</strong> 48 base pair and bends the DNA 110 degrees. We have proposed<br />

that the DNA bind on the circumference <strong>of</strong> this toroid shaped FlhDC complex. Although we have<br />

determined the sequence 3 footprints on FlhDC regulated promoters, it is not possible to<br />

determine a consensus binding site in these 3 sequences. In this study, we have determined<br />

which bases are important for DNA binding and activity for FlhDC regulated promoter activity.<br />

First, we have divided the FlhDC footprint in the fliA promoter into five segments and found that<br />

two <strong>of</strong> the segments or 40% <strong>of</strong> the footprint were not required for binding. The remaining 30<br />

base pairs were divided into 3-5 base segments and randomly mutagenized and screened for<br />

the ability to bind and activate the fliA promoter. Analysis <strong>of</strong> these data suggests a consensus <strong>of</strong><br />

12 A, 15 A, 34 T, 36 A, 37 T, 44 A, 45 T in FlhD4C2 footprint fragment were important for activity. Five <strong>of</strong><br />

these bases demonstrated high specificity. Finally, this consensus was tested and found to be<br />

important in other FlhDC regulated promoter regions.<br />

28


<strong>BLAST</strong> X Wed. Morning Session<br />

A NOVEL AMINO ACID BINDING STRUCTURE IN BACTERIAL CHEMOTAXIS<br />

George D. Glekas and George W. Ordal<br />

Department <strong>of</strong> Biochemistry, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign<br />

409 MSB, Urbana, IL 61801<br />

Simple flagellated bacteria, such as Bacillus subtilis, possess the ability to sense their<br />

environment and move to more favorable conditions, where there are, for instance, more<br />

nutrients like amino acids, by the process <strong>of</strong> chemotaxis. The initial stage is binding <strong>of</strong> an<br />

amino acid by receptors on the outside <strong>of</strong> the cell. This binding causes conformational changes<br />

that affects the activity <strong>of</strong> enzymes on the inside <strong>of</strong> the cell and alters the movement <strong>of</strong> the<br />

bacteria. The main paradigm for understanding these events is the bacterium Escherichia coli.<br />

However, we have discovered that, in fact, the mechanism used by E. coli is not used by most<br />

bacteria and that the mechanism used by the distantly related bacterium B. subtilis is more likely<br />

to be the general mechanism. Unlike in E. coli, where binding <strong>of</strong> attractant causes a shift <strong>of</strong> the<br />

receptor polypeptide that goes from the outside <strong>of</strong> the cell to the cell interior toward the cell<br />

interior, attractant causes rotational movement <strong>of</strong> the receptors without any comparable interior<br />

shifting<br />

We seek to understand how this rotational movement occurs in the asparagine receptor<br />

McpB. To do this, we have discovered that the most likely conformation <strong>of</strong> the exterior part <strong>of</strong><br />

the receptor is far different from that in the E. coli receptor. Molecular and homology modeling<br />

<strong>of</strong> the McpB sensing domain has led to a structural model that reveals a dual PAS domain<br />

structure similar to the crystal structure <strong>of</strong> the LuxQ sensor. PAS domains are known to be<br />

conserved structures capable <strong>of</strong> binding a great many “small” molecules. Further mutagenetic<br />

analyses <strong>of</strong> putative asparagine-binding residues have not only confirmed the validity <strong>of</strong> the<br />

structural model, they have revealed certain residues that greatly effect chemo-attractant<br />

binding. Using both in vivo chemotactic assays and in vitro isothermal titration calorimetry<br />

performed on purified mutant receptor exterior regions, three residues, all in the upper PAS<br />

domain, have been shown to lower the affinity <strong>of</strong> the McpB receptor for asparagine. Mutations<br />

in the lower PAS domain show no such effect. Further structural and mutagenetic studies have<br />

shown a similar dual PAS architecture in the B. subtilis proline receptor McpC, with similar<br />

residues responsible for amino acid binding. Extensive homology modeling shows that eight <strong>of</strong><br />

the ten B. subtilis chemoreceptors have PAS domains in their sensing domains.<br />

We are now in the process <strong>of</strong> modeling the consequences <strong>of</strong> binding attractant at these<br />

residues on the expected structure <strong>of</strong> the receptor.<br />

29


<strong>BLAST</strong> X Wed. Morning Session<br />

STRUCTURE AND FUNCTION OF THE HELICOBACTER PYLORI CHEMORECEPTOR TlpB<br />

John Goers, Nathan Henderson, S. James Remington, Karen Ottemann * and Karen Guillemin<br />

Institute <strong>of</strong> Molecular Biology, <strong>University</strong> <strong>of</strong> Oregon, Eugene OR, 97403<br />

* Environmental Toxicology, UC Santa Cruz, Santa Cruz, CA 95064<br />

We have shown that the Helicobacter pylori chemoreceptor TlpB is required for<br />

chemotaxis away from acid and from the quorum-sensing molecule autoinducer-2. Here we<br />

report the determination <strong>of</strong> an atomic resolution crystal structure for the periplasmic domain <strong>of</strong><br />

TlpB. The structure reveals a PAS domain inserted into a pair <strong>of</strong> antiparallel helices and shows<br />

unexpected structural homology to the LuxQ receptor from Vibrio harveyi. Furthermore, the PAS<br />

domain tightly binds urea, suggesting that the TlpB receptor may be associated with detection <strong>of</strong><br />

urea. We present mutational and physical evidence for interaction <strong>of</strong> TlpB with urea and its role<br />

in chemotaxis.<br />

30


<strong>BLAST</strong> X Wed. Morning Session<br />

THE TM2-HAMP CONNECTION<br />

Gus A. Wright, Rachel L. Crowder and Michael D. Manson<br />

Department <strong>of</strong> Biology, Texas A&M <strong>University</strong>, College Station, TX 77843<br />

The HAMP domain—a conserved protein motif in most Histidine kinases, Adenylate<br />

cyclases, Methyl-accepting chemotaxis proteins, and Phosphatases—typically conforms to a<br />

helix–connector-helix domain architecture. Recently, an NMR solution structure was determined<br />

for the Af1503 HAMP from the archaebacteria Archeoglobus fulgidis (Hulko, et al, Cell 126: 929-<br />

940, 2006). This structure revealed that the HAMP domain is a parallel four-helix bundle with a<br />

knob-on-knob packing. Similar structures <strong>of</strong> HAMP domains are proposed to exist in all 5<br />

chemoreceptors <strong>of</strong> Escherichia coli, including the aspartate chemoreceptor Tar. The HAMP<br />

domain <strong>of</strong> Tar receives information from TM2 and regulates the transmission <strong>of</strong> the signal to the<br />

kinase signaling domain. Mutations were introduced into the TM2-HAMP connector region <strong>of</strong><br />

Tar to determine if manipulating the input signal into the HAMP affects the output signal to the<br />

signaling domain. MLLT residues between R214 and P219 were deleted (-4 through -1), and<br />

additional LLT tandem repeats were added up to 8 residues after P219 (+1 through +8).<br />

Aspartate sensitivity, rotational bias, mean reversal frequency, and in vivo methylation were<br />

measured for these mutants. The results suggest that adding helical turns in this region<br />

destabilizes, or “relaxes”, the HAMP/signaling domains; while, removing helical turns from this<br />

region stabilizes, or “tightens”, the HAMP/signaling domains. In order to determine if increasing<br />

flexibility <strong>of</strong> the TM2/HAMP connector affects the output signal, a second set <strong>of</strong> mutants were<br />

constructed to replace the MLLT region with four tandem glycine residues (4G). Glycine<br />

residues were then subtracted (-4G through -1G) and added (+1G through +5G). Data collected<br />

from this experiment suggests that increasing the flexibility <strong>of</strong> this region dampens the input<br />

signal from TM2, in addition to destabilizing the HAMP/signaling domains.<br />

31


<strong>BLAST</strong> X Wed. Morning Session<br />

STRUCTURE, ASSEMBLY AND CONFORMATIONAL CHANGES IN CHEMORECEPTORS<br />

STUDIED IN INTACT BACTERIAL CELLS USING CRYO-ELECTRON TOMOGRAPHY<br />

Cezar M. Khursigara, Xiongwu Wu*, Peijun Zhang, Jon Lefman, Mario J. Borgnia, Yuhai Tu ‡ ,<br />

Jacqueline Milne and Sriram Subramaniam<br />

National Cancer Institute and *National Heart, Lung and Blood Institute, NIH, Bethesda, MD<br />

20892.<br />

‡ T. J. Watson Research Center, IBM, Yorktown Heights, NY 10598<br />

Bacteria respond to changes in their chemical environment by activating an assembly <strong>of</strong><br />

proteins that collectively represent the bacterial chemotaxis apparatus. In Gram-negative<br />

bacteria the core-signaling unit <strong>of</strong> the chemotaxis machinery is a ternary complex composed <strong>of</strong><br />

chemoreceptors, CheA and CheW that localize primarily to the poles <strong>of</strong> the cell and form<br />

extended arrays. Using cryo-electron tomography, we describe and compare the architecture,<br />

localization and spatial relationship between macromolecular complexes involved in chemotaxis<br />

signaling and cellular motility in three different Gram-negative bacteria. In addition, by<br />

combining the tomographic analysis with 3D averaging methods we demonstrate that trimeric<br />

chemoreceptors in E. coli display two distinct conformations that differ principally in<br />

arrangement <strong>of</strong> the HAMP domains within each trimer.<br />

32


<strong>BLAST</strong> X Wed. Morning Session<br />

DISCRETE SIGNAL-ON AND -OFF CONFORMATIONS IN THE AER HAMP DOMAIN<br />

Kylie J. Watts, Mark S. Johnson and Barry L. Taylor<br />

Dept. Microbiology and Mol. Genetics, Loma Linda <strong>University</strong>, Loma Linda, CA, USA<br />

The PAS-FAD sensor <strong>of</strong> the aerotaxis receptor, Aer, signals through HAMP and<br />

signaling domains that are similar to these domains in other chemoreceptors. Our previous<br />

crosslinking studies showed that the AS-1 and AS-2 helices <strong>of</strong> the Aer-HAMP domain might<br />

form a four-helix bundle similar to the Af1503 and Tar HAMP domains. In this study, AS-1<br />

residues were crosslinked to AS-2′ residues in di-cys Aer mutants (using 13 proximal and 4<br />

distal di-cys pairs). The results confirmed a parallel four-helix HAMP bundle for Aer, but one in<br />

which AS-2 is rotated compared to the orientation <strong>of</strong> AS-2 in Af1503 or Tar.<br />

We extended our HAMP crosslinking studies to probe for structural differences between<br />

the signal-on (CW) and signal-<strong>of</strong>f (CCW) states. In our previous crosslinking studies, we used<br />

the oxidant copper phenanthroline, which maintains Aer in the signal-<strong>of</strong>f state. In order to<br />

generate snapshots <strong>of</strong> the signal-on state, CW lesions such as PAS-N85S were engineered into<br />

the Aer di-cys and single-cys mutants. When the AS-1 to AS-2′ di-cys crosslinking experiments<br />

were repeated in mutants containing N85S, several di-cys pairs showed significant increases in<br />

dimer formation rates. These di-cys pairs were located at the distal end <strong>of</strong> the HAMP four-helix<br />

bundle. In contrast, no significant crosslinking changes were observed at the proximal end <strong>of</strong><br />

the four-helix bundle. The data supports a model in which the distal ends <strong>of</strong> the HAMP helices<br />

move closer together during signal transduction. This could be due to an inward lateral<br />

movement <strong>of</strong> the helices, and may include some element <strong>of</strong> rotation. However, the entire Aer-<br />

HAMP domain does not appear to rotate as has been proposed for Af1503.<br />

In Aer, HAMP lesions that lock the receptor in the signal-on (CW) state cluster at the<br />

distal end <strong>of</strong> a HAMP four-helix bundle, indicating a possible site for PAS-HAMP interactions<br />

during signal transduction. We used PEG-maleimide to determine in vivo the solvent-accessible<br />

surface <strong>of</strong> the HAMP and proximal signaling domains (residues 206-275). Solvent accessibility<br />

was restricted for most AS-2, but not AS-1 or connector, residues in Aer. This indicates that AS-<br />

2 residues that are exposed to solvent in Af1503 and Tar, and are predicted to be exposed in an<br />

Aer-HAMP model, are buried in vivo in Aer. We are currently investigating whether these<br />

residues are buried in a PAS-HAMP contact domain. We are also probing the surface <strong>of</strong> the<br />

HAMP domain in the signal-on state (with N85S) to determine whether there are differences in<br />

accessibility between the two signaling states.<br />

33


<strong>BLAST</strong> X Wed. Morning Session<br />

INVESTIGATING THE STRUCTURE OF TERNARY COMPLEX OF HISTIDINE KINASE CheA,<br />

COUPLING PROTEIN CheW, AND CHEMORECEPTOR BY PULSED DIPOLAR ESR<br />

SPECTROSCOPY<br />

Jaya Bhatnagar, Peter P. Borbat, Jack H. Freed, Brian R. Crane<br />

Cornell <strong>University</strong>, B 150 Caldwell Hall, Ithaca, NY 14853<br />

A central question in understanding the mechanism <strong>of</strong> chemotaxis involves the nature <strong>of</strong><br />

interactions between histidine kinase CheA, adaptor protein CheW and receptors. Pulsed<br />

dipolar ESR spectroscopy (PDS) has developed as a valuable technique for structural<br />

characterization <strong>of</strong> protein complexes. PDS provides long-range distance information between<br />

spin-labeled residues in the proteins. A set <strong>of</strong> distance measurements can be subsequently<br />

used to model the assembly structure <strong>of</strong> the whole complex. In our previous work, we<br />

demonstrated the success <strong>of</strong> this approach in predicting the structure <strong>of</strong> complex <strong>of</strong> CheW with<br />

CheA. We have now applied PDS to the ternary complex formed by CheA, CheW and soluble<br />

chemoreceptor fragments. Our results indicate changes in distance distributions from spinlabeled<br />

sites on P4, P5 domains and CheW in the presence <strong>of</strong> unlabeled receptor. Dipolar<br />

signals between spin-labeled receptor and CheA∆289 (domains P3, P4 and P5 together) or<br />

CheW provide important insights about the relative position and orientation <strong>of</strong> the three<br />

components with respect to each other. Based on this data we have developed a structural<br />

model <strong>of</strong> the ternary complex and the conformational changes CheA undergoes upon binding to<br />

receptor.<br />

34


<strong>BLAST</strong> X Wed. Morning Session<br />

THE CHEMOTACTIC CORE SIGNALLING COMPLEX IS ULTRASTABLE<br />

Annette H. Erbse and Joseph J. Falke<br />

Department <strong>of</strong> Chemistry and Biochemistry, <strong>University</strong> <strong>of</strong> Colorado, Boulder, Campus Box 215,<br />

Boulder, CO 80309<br />

The chemotatic core complex, composed <strong>of</strong> the transmembrane receptor, the histidinekinase<br />

CheA and the coupling protein CheW, is the central building block <strong>of</strong> the extensive,<br />

highly cooperative polar signaling clusters in bacteria involved in sensing chemical gradients.<br />

Recent studies have shown that the receptors are organized hexagonal arrays <strong>of</strong> trimers-<strong>of</strong>dimers.<br />

But it is still unclear how these arrays are formed and stabilized, how CheA and CheW<br />

are incorporated into single core complexes and how the complexes are interconnected to build<br />

the cooperative signaling network. Here we focus on the stability <strong>of</strong> the core complex. We show<br />

that the isolated, membrane-bound core complex is stable for at least 24 hours, both when it is<br />

assembled in vivo and in vitro. All three components are needed to achieve this ultra-stability,<br />

which is dependent on electrostatic interactions. By contrast, the stability is independent <strong>of</strong><br />

ligand binding, receptor methylation or kinase activity. We propose that the assembly <strong>of</strong> the<br />

signaling clusters is cooperative, such that interactions between CheA , CheW and the receptor<br />

trimers-<strong>of</strong>-dimers are needed not only for receptor regulated kinase activity <strong>of</strong> individual core<br />

complexes, but also to position neighbouring complexes during the formation <strong>of</strong> a multi-linked<br />

network. The resulting ultra-stable network is the foundation for the ordered organization and<br />

the hypersensitivity <strong>of</strong> the signaling patches.<br />

35


<strong>BLAST</strong> X Wed. Morning Session<br />

ELECTRON CRYOTOMOGRAPHY OF BACTERIAL CHEMOTAXIS ARRAYS<br />

Ariane Briegel 1 , H. Jane Ding 1 , Zhuo Li 1 , John Werner 2 , Zemer Gitai 2 , D. Prabha Dias 1 ,<br />

Rasmus B. Jensen 3 , Elitza Tocheva 1 and Grant Jensen 1<br />

1 California Institute <strong>of</strong> Technology, CA<br />

2 Princeton <strong>University</strong>, NJ<br />

3 <strong>University</strong> <strong>of</strong> Roskilde, Denmark<br />

Motile prokaryotes are able to sense and to respond to ambient conditions through a<br />

process known as chemotaxis. Attractants and repellents bind to the sensing domain <strong>of</strong> methylaccepting<br />

chemotaxis proteins (MCPs), thereby regulating the activity <strong>of</strong> the histidine kinase<br />

CheA. Together with the linking protein CheW, CheA is located at the distal tip <strong>of</strong> the<br />

cytoplasmic signaling domain <strong>of</strong> the MCPs. If activated, CheA phophorylates CheY (CheY-P),<br />

which in turn controls the direction <strong>of</strong> flagellar rotation. Together with CheA and a linking protein<br />

CheW, the MCPs form extended chemotaxis arrays at the cell poles.<br />

Electron cryotomograhy (ECT) makes it possible to visualize chemoreceptor clusters in<br />

prokaryotes in vivo at macromolecular resolution (4-8 nm). While high-resolution structures <strong>of</strong><br />

the individual chemotaxis proteins are available, their arrangement and position in the arrays<br />

remain unclear. Understanding this "mesoscale" architecture <strong>of</strong> the clusters is critical, however,<br />

since it is vital to the arrays' cooperative signal amplification and regulation. In order to<br />

unambiguously identify the chemotaxis arrays inside cells, we have correlated ECT with<br />

fluorescent light microscopy (FLM), using slightly fixed and immobilized Caulobacter crescentus<br />

cells with a fusion <strong>of</strong> the red-fluorescent protein, mCherry, to the C-terminus <strong>of</strong> the<br />

chemoreceptor (McpA). After plunge freezing, we imaged the same cells by ECT. In<br />

combination with ECT <strong>of</strong> near-native wild-type and mutant cells, we used the correlated FLM<br />

and ECT approach to identify the chemotactic array, its location and its in-vivo structure. We<br />

demonstrate that in wild-type Caulobacter crescentus cells preserved in a near-native state, the<br />

chemoreceptors are hexagonally packed with a lattice spacing <strong>of</strong> 12 nm, just a few tens <strong>of</strong><br />

nanometers away from the flagellar motor that they control. The arrays were always found on<br />

the concave side <strong>of</strong> the cell, further demonstrating that Caulobacter cells maintain dorsal/ventral<br />

as well as anterior/posterior asymmetry. Placing the known crystal structure <strong>of</strong> a trimer <strong>of</strong><br />

receptor dimers at each vertex <strong>of</strong> the lattice accounts well for the density, supporting an array<br />

composition unlike the published models for Escherichia coli [1] or Thermotoga maritima [2].<br />

We are now in the process <strong>of</strong> comparing the chemotaxis arrays <strong>of</strong> a wide range <strong>of</strong> bacteria to<br />

determine the similarities and differences <strong>of</strong> these macromolecular assemblies at the<br />

‘mesoscale’ level.<br />

1. Shimizu T. S. et al, Nat Cell Biol 11 (2000) 792.<br />

2. Park, S.-Y. et al, Nat Struct Mol Biol 13 (2006) 400.<br />

36


<strong>BLAST</strong> X Thurs. Morning Session<br />

TWO REGULATORY PROTEINS CONTROL THE SWIM-OR-STICK SWITCH IN<br />

ROSEOBACTERS<br />

Robert Belas<br />

<strong>University</strong> <strong>of</strong> Maryland Biotechnology Institute, Center <strong>of</strong> Marine Biotechnology, 701 East Pratt<br />

St., Baltimore, MD 21202<br />

Members <strong>of</strong> the Roseobacter clade <strong>of</strong> α-Proteobacteria are among the most abundant<br />

and ecologically relevant marine bacteria. One <strong>of</strong> the most salient features <strong>of</strong> the roseobacters<br />

from aspects <strong>of</strong> marine ecology is their ability to enter into close physical and physiological<br />

relationships with “red tide” phytoplankton such as din<strong>of</strong>lagellates. For example, Silicibacter sp.<br />

TM1040, our model roseobacter, forms a symbiosis with the din<strong>of</strong>lagellate Pfiesteria piscicida,<br />

such that the din<strong>of</strong>lagellate cannot live without TM1040. Aiding TM1040 in development <strong>of</strong> the<br />

symbiosis is a biphasic swim-or-stick lifestyle wherein a genetic regulatory circuit controls<br />

whether the bacteria are motile and chemotactic or sessile and develop a bi<strong>of</strong>ilm. Bacterial<br />

swimming and chemotaxis behavior are initial, essential steps in establishment <strong>of</strong> the symbiosis.<br />

Once near the host surface, motility and flagellar synthesis are downregulated, while bi<strong>of</strong>ilm<br />

formation and synthesis <strong>of</strong> an antibiotic are upregulated. The abilities to swim using flagella and<br />

to form a bi<strong>of</strong>ilm via adhesins have been demonstrated to be important traits for both pathogenic<br />

and symbiotic bacteria. While it is generally agreed that motility and bi<strong>of</strong>ilm development are<br />

mutually exclusive, the molecular mechanisms that underlie the lifestyle switch remain virtually<br />

unknown for most bacterial species. We have used genetic screens to search for mutants<br />

defective in either the motile or the sessile phenotype, and have discovered many new genes<br />

including two previously unknown and novel regulatory proteins, FlaC and FlaD that are<br />

envisaged to act together with cyclic dimeric GMP to play important roles in the swim-or-stick<br />

switch. FlaC is predicted to function as a response regulator protein, with homology to a protein<br />

<strong>of</strong> Caulobacter crescentus known to be important for cell envelope function. FlaC - cells are<br />

skewed towards the motile phase, e.g., their populations have a greater percentage <strong>of</strong> motile<br />

cells and fewer rosettes, and have defects in antibiotic synthesis and bi<strong>of</strong>ilm formation. Thus,<br />

FlaC determines whether the switch is in the swim or stick position. FlaD is predicted to be a<br />

MarR-type DNA-binding protein. Mutations in flaD result in nonmotile cells that synthesize but<br />

cannot rotate their flagella, i.e., they produce paralyzed flagella. We hypothesize that FlaD is<br />

involved in the function <strong>of</strong> the flagellar motor, either by (1) acting directly to control transcription<br />

<strong>of</strong> the class IV fliL operon or (2) acting indirectly to control transcription or activity <strong>of</strong> a protein<br />

that acts as a ‘clutch’ to engage or disengage the flagellar motor. The implications <strong>of</strong> FlaC and<br />

FlaD activities in the swim-or-stick strategy and their impact on the symbiosis will be discussed.<br />

37


<strong>BLAST</strong> X Thurs. Morning Session<br />

DELETION ANALYSIS OF RcsC REVEALS A NOVEL SIGNALING-PATHWAY<br />

CONTROLLING BIOFILM FORMATION IN ESCHERICHIA COLI<br />

Ricardo Oropeza, Rosalva Salgado, Ismael Hernandez-Lucas and Edmundo Calva<br />

Departmento de Microbiologia Molecular, Instituto de Biotecnologia, Universidad Nacional<br />

Autonoma de Mexico. Av. Universidad 2001. Col. Chamilpa. Cuernavaca, Morelos. Mexico, CP<br />

62210<br />

RcsC is a hybrid histidine kinase that forms part <strong>of</strong> a phosphorelay signal transduction<br />

pathway with RcsD and RcsB. Besides the typical domains <strong>of</strong> a sensor kinase, i.e. the<br />

periplasmic (P), linker (L), dimerization and H-containing (A), and ATP- binding (B), RcsC<br />

possesses a receiver domain (D) at the carboxy-terminal domain.<br />

In order to study the role played by each <strong>of</strong> the RcsC domains, four plasmids containing<br />

several <strong>of</strong> these domains were constructed (i.e. PLAB, LAB, AB and ABD) and transformed in<br />

Escherichia coli wild type. Different amounts <strong>of</strong> bi<strong>of</strong>ilm were produced, assessed by crystal<br />

violet staining, depending on the RcsC domains expressed by the plasmid.<br />

E. coli transformed with the plasmid expressing the ABD subdomains produced the<br />

highest amount <strong>of</strong> bi<strong>of</strong>ilm, while the lowest amount <strong>of</strong> bi<strong>of</strong>ilm was produced under the control <strong>of</strong><br />

the PLAB expressing plasmid. This phenotype was observed in the same ratio when the<br />

plasmids were transformed in a ΔrcsCDB strain.<br />

Several mutants on genes involved in bi<strong>of</strong>ilm formation were transformed with this set <strong>of</strong><br />

plasmids. Bi<strong>of</strong>ilm formation was abolished in the pgaABCD and nhaR backgrounds but not in<br />

the csrB and uvrY backgrounds. Our results suggest the existence <strong>of</strong> a signaling pathway<br />

depending <strong>of</strong> RcsC but independent <strong>of</strong> RcsD and RcsB, activating bi<strong>of</strong>ilm formation by the<br />

pgaABCD operon.<br />

38


<strong>BLAST</strong> X Thurs. Morning Session<br />

REGULATION OF CELL FATE IN BACILLUS SUBTILIS BIOFILMS<br />

H. C. Vlamakis 1* , C. Aguilar 1* , R. Losick 2 , and R. Kolter 1<br />

1 Harvard Medical School, Boston, MA, 2 Harvard <strong>University</strong>, Cambridge, MA.<br />

*These authors contributed equally to this work.<br />

Many microbial populations differentiate from free-living planktonic cells into surfaceassociated<br />

multicellular communities known as bi<strong>of</strong>ilms. Within a bi<strong>of</strong>ilm, motile Bacillus subtilis<br />

cells differentiate into non-motile chains <strong>of</strong> cells that form parallel bundles held together by an<br />

extracellular matrix. These bundles eventually produce aerial structures that serve as<br />

preferential sites for sporulation. By analyzing strains harboring multiple cell-type specific<br />

promoter fusions we can visualize the spatial anatomy <strong>of</strong> at least three physiologically distinct<br />

cell populations within mature bi<strong>of</strong>ilms. Motile, matrix-producing, and sporulating cells localize to<br />

distinct regions within the bi<strong>of</strong>ilm and the localization and percentage <strong>of</strong> each cell type is<br />

dynamic. Mutants unable to produce extracellular matrix form unstructured bi<strong>of</strong>ilms that are<br />

deficient in sporulation. This suggests that in architecturally complex bi<strong>of</strong>ilms, spore formation<br />

is coupled to the production <strong>of</strong> extracellular matrix. The coupling <strong>of</strong> matrix production and<br />

sporulation could be explained by the phosphorylation state <strong>of</strong> the master transcriptional<br />

regulator Spo0A. Spo0A is phosphorylated both directly and through a phosphorelay by at least<br />

five different histidine kinase proteins. When cells have low levels <strong>of</strong> Spo0A-P, matrix genes<br />

are expressed; however, at higher levels <strong>of</strong> Spo0A-P, sporulation commences. We have found<br />

that a deletion <strong>of</strong> kinD, a gene encoding one <strong>of</strong> the kinases that feed into the Spo0A<br />

phosphorelay, is sufficient to restore sporulation to matrix-deficient mutants. We hypothesize<br />

that KinD is not acting as a kinase under these conditions, but rather functions as a<br />

phosphatase to delay sporulation until matrix (or a matrix-encased signal) is sensed.<br />

39


<strong>BLAST</strong> X Thurs. Morning Session<br />

PROTEIN MISFOLDING DONE RIGHT: THE BIOGENESIS OF BACTERIAL AMYLOID<br />

FIBERS<br />

Xuan Wang, Neal Hammer and Matt Chapman<br />

<strong>University</strong> <strong>of</strong> Michigan, Department <strong>of</strong> Molecular, Cellular and Developmental Biology, Ann<br />

Arbor, MI, 48109<br />

Many Enterobacteriaceae spp., including E. coli, produce surface-localized amyloid<br />

fibers called curli. Curli fibers are associated with bi<strong>of</strong>ilm formation, host cell adhesion and<br />

invasion, and immune system activation. Unlike disease-associated amyloid formation, curli<br />

biogenesis is a directed and highly regulated process. The major curli subunit protein, CsgA,<br />

polymerizes into amyloid after interacting the CsgB nucleator protein. CsgB presents an<br />

amyloid-like template to CsgA on the cell surface that initiates fiber formation. CsgA has five<br />

imperfect repeating units (R1-R5) that are each predicted to form strand-loop-strand structures.<br />

Asn and Gln residues in R1 and R5 were found to be required for efficient amyloid formation<br />

and for interaction with the CsgB nucleator protein. Furthermore, the polymerization <strong>of</strong> CsgA<br />

was tempered by the presence <strong>of</strong> conserved aspartic residues in R2, R3 and R4. When these<br />

aspartic acid residues were changed to alanine (CsgA*), polymerization was significantly faster<br />

in vitro. Even more remarkable was the observation that CsgA* assembled into an amyloid fiber<br />

in vivo in the absence <strong>of</strong> CsgB. The ability <strong>of</strong> CsgA* to polymerize into amyloid more efficiently,<br />

and in the absence <strong>of</strong> CsgB, was not without consequences. Cells expressing CsgA* grew more<br />

slowly when compared to cells expressing wild type CsgA. This analysis suggests that aspartic<br />

acid residues can potently inhibit functional amyloid formation. CsgA has apparently evolved to<br />

efficiently assemble into an amyloid in vivo only in the presence <strong>of</strong> CsgB. This suggests an<br />

elegant mechanism to control amyloid formation by regulating the temporal and spatial<br />

interactions between CsgA and CsgB.<br />

40


<strong>BLAST</strong> X Thurs. Morning Session<br />

RHODOBACTER SPHAEROIDES, A BACTERIUM WITH TWO FLAGELLAR SYSTEMS AND<br />

MULTIPLE CHEMOTAXIS GENE HOMOLOGS<br />

Ana Martínez del Campo 1 , Sebastian Poggio 2 , Teresa Ballado 1 , Aurora Osorio 2 , Javier de la<br />

Mora 1 , Laura Camarena 2 and Georges Dreyfus 1<br />

Instituto de Fisiología Celular 1 , Instituto de Investigaciones Biomédicas 2 , Universidad Nacional<br />

Autónoma de México, 04510 México DF, México.<br />

Rhodobacter sphaeroides has two flagellar systems (fla1 and fla2). One <strong>of</strong> these<br />

systems has been shown to be functional and is required for the synthesis <strong>of</strong> the wellcharacterized<br />

single subpolar flagellum (fla1), while the other was found only after the genome<br />

sequence <strong>of</strong> this bacterium was completed (fla2). In this work we found that the second flagellar<br />

system <strong>of</strong> R. sphaeroides can be expressed and encodes a functional flagellum. This second<br />

flagellar system produces polar flagella that are required for swimming. Phylogenic analysis<br />

suggests that the flagellar system that was initially characterized, was in fact, acquired by<br />

horizontal transfer from a γ-proteobacterium, while the second flagellar system contains the<br />

native genes.<br />

In addition to having two flagellar systems, this photosynthetic bacterium posses several<br />

reiterated chemotactic genes (2 cheB, 3 cheR, 4 cheA and cheW and 6 cheY), which are<br />

encoded in three operons (cheOp1, cheOp2 and cheOp3). In spite <strong>of</strong> this, only some <strong>of</strong> the<br />

gene copies are required when the cell is swimming with the fla1 flagellum. The presence <strong>of</strong> a<br />

second functional flagellum (fla2) suggests that some <strong>of</strong> these genes could be involved in its<br />

tactic control. To test this hypothesis we proceeded to individually mutate each cheY gene. We<br />

show evidence that CheY1, CheY2 and CheY5 control de chemotactic behavior mediated by<br />

fla2 flagella. Additionally, we identified that open reading frame RSP6099 encodes the fla2 FliM<br />

protein. Furthermore CheY1, CheY2 and CheY5 are located within cheOp1, which is not<br />

essential for chemotaxis mediated by the fla1 system. This raises the question: What is the role<br />

<strong>of</strong> cheOp1?<br />

41


<strong>BLAST</strong> X Thurs. Morning Session<br />

MOTILITY, CHEMOTAXIS AND VIRULENCE OF BORRELIA BURGDORFERI, THE LYME<br />

DISEASE SPIROCHETE<br />

M. A. Motaleb 1 , P. Stewart 2 . A. Bestor 2 , P. Rosa 2 and N. Charon 3<br />

1 Dept <strong>of</strong> Microbiology & Immunology, East Carolina <strong>University</strong>, Greenville, NC<br />

2 Human Bacterial Pathogenesis, NIH, RML, Hamilton, MT<br />

3 Dept <strong>of</strong> Microbiology, Immunology & cell Biology, West Virginia <strong>University</strong>, Morgantown, WV<br />

Borrelia burgdorferi is the causative agent <strong>of</strong> Lyme disease. It is the most prevalent<br />

arthropod borne infection in the United States with 27,444 reported cases on 2007. The<br />

disease is a multiple-systemic disorder with various clinical manifestations including erythema<br />

migrans rash, arthritis, cardiac, musculoskeletal and neurological manifestations.<br />

B. burgdorferi exists in nature in an enzootic cycle. Ixodes scapularis ticks (commonly<br />

known as deer ticks) acquire the infection when they feed on an infected host, mainly rodents.<br />

During subsequent tick feeding, which lasts for several days, B. burgdorferi migrate from the tick<br />

midgut, pass through the salivary glands, and are then transmitted to the mammal through the<br />

saliva. B. burgdorferi is highly invasive. After being deposited in the skin following a tick bite, the<br />

spirochetes can invade many tissues including the joints, heart, and nervous system.<br />

Motility and chemotaxis are critical for bacterial survival and adaptation in diverse<br />

environmental conditions. In several species <strong>of</strong> bacteria, motility and chemotaxis have been<br />

shown to be associated with the disease process. Results obtained using B. burgdorferi with<br />

mutations in key motility and chemotaxis genes also indicate that these activities are required<br />

for the pathogenesis <strong>of</strong> Lyme disease. These studies could lead to the development <strong>of</strong> a novel<br />

pharmacological agent to treat/prevent Lyme disease.<br />

42


<strong>BLAST</strong> X Thurs. Morning Session<br />

PLEIOTROPIC PHENOTYPES OF A YERSINIA ENTEROCOLICIA FLHD MUTANT INCLUDE<br />

REDUCED LETHALITY IN A CHICKEN EMBRYO MODEL<br />

Birgit M. Prüß 1 , Megan K.T. Ramsett 1 , Nathan J. Carr 1 , Jyoti G. Iyer 1 , Penelope S. Gibbs 1 ,<br />

Philip Matsumura 2 , and Shelley M. Horne 1<br />

1<br />

Veterinary & Microbiological Sciences Department, North Dakota State <strong>University</strong>, Fargo ND<br />

58108<br />

2<br />

Department <strong>of</strong> Microbiology and Immunology, <strong>University</strong> <strong>of</strong> Illinois at Chicago, Chicago IL<br />

60612<br />

The goal <strong>of</strong> this study was to correlate phenotypes with gene regulation by several<br />

flagellar regulators in Yersinia enterocolitica. FlhD/FlhC was initially described as a flagella<br />

transcriptional activator and later recognized as a global regulator in several enteric bacteria (1,<br />

2). In Y. enterocolitica, FlhD/FlhC positively affected the expression levels <strong>of</strong> genes <strong>of</strong> histidine<br />

degradation and pyrimidine biosynthesis, while repressing the urease genes (1). A second<br />

protein that is involved in the regulation <strong>of</strong> flagellar genes, the sigma factor FliA, exhibited a<br />

negative effect upon the expression levels <strong>of</strong> seven plasmid-encoded virulence genes (3). In<br />

addition, eight flagellar operons were regulated by FliA. Among the differences to Escherichia<br />

coli were a 10 fold regulation <strong>of</strong> fliZ expression by FliA and a lack <strong>of</strong> FliA regulation <strong>of</strong> the flgM<br />

operon.<br />

Phenotypes relating to FlhD/FlhC and FliA gene regulation were investigated. These<br />

phenotypes included growth on carbon and nitrogen sources, and virulence (4). Growth was<br />

determined with Phenotype MicroArrays (Biolog). Compared to the wild-type strain, flhD and fliA<br />

mutants exhibited increased growth on purines as carbon sources and decreased growth on<br />

pyrimidines and histidine as nitrogen sources. Several dipeptides provided differential growth<br />

conditions between the wild-type strain and both mutants. Gene regulation was determined for<br />

the dpp (dipeptide transport) and opp (oligopeptide transport) genes and was found to correlate<br />

with the observed phenotypes. Phenotypes relating to virulence were determined with the<br />

chicken embryo lethality assay that was previously established and used for E. coli strains (5).<br />

Relative to the wild-type strain, the flhD mutant caused a reduced lethality in this assay, while<br />

the fliA mutant caused lethality similar to the wild-type. Mutants were able to colonize infected<br />

embryo organs at levels that were comparable to the wild-type. In addition, a mutant in flhB,<br />

encoding one component <strong>of</strong> the flagellar type III secretion system also caused a reduced<br />

embryo lethality. Since genes <strong>of</strong> the type III secretion system are regulated by FlhD/FlhC and<br />

not by FliA, we believe that the lethality phenotype <strong>of</strong> the flhD mutant is due to regulation <strong>of</strong> the<br />

type III secretion genes.<br />

1. V. Kapatral et al., Microbiol. 150, 2289 (2004).<br />

2. B. M. Prüß et al., J. Bacteriol. 185, 534 (2003).<br />

3. S. M. Horne, B. M. Prüß, Arch. Microbiol. 185, 115 (2006).<br />

4. M. K. Townsend et al., BMC Microbiol. 8, 12 (2008).<br />

5. P. S. Gibbs, J. J. Maurer, L. K. Nolan, R. E. Wooley, Avian Dis. 47, 370 (2003).<br />

43


<strong>BLAST</strong> X Thurs. Morning Session<br />

REGULATION OF MOTILITY BY QUORUM SENSING IN SINORHIZOBIUM MELILOTI AND<br />

ITS ROLE IN SYMBIOSIS ESTABLISHMENT<br />

Juan E. González*, Nataliya Gurich, Jennifer L. Morris, Konrad Mueller, and Arati V. Patankar<br />

Department <strong>of</strong> Molecular and Cell Biology, <strong>University</strong> <strong>of</strong> Texas at Dallas, Richardson, Texas,<br />

USA<br />

Quorum sensing is a mechanism widely used by bacteria to coordinate their behavior in<br />

response to a particular cell population density. Signal molecules, termed autoinducers, are<br />

produced by bacteria, and at a high population density, accumulate in the environment. Once a<br />

threshold level <strong>of</strong> autoinducer is reached, they bind to their cognate transcriptional regulators<br />

and activate or repress expression <strong>of</strong> target genes, thereby preparing the bacteria for behaviors<br />

associated with high cell density, such as interacting with eukaryotic hosts.<br />

In Sinorhizobium meliloti, this mechanism is utilized to appropriately modulate gene<br />

expression and permit the establishment <strong>of</strong> a nitrogen-fixing symbiosis with its host plant<br />

Medicago sativa. S. meliloti possesses a quorum-sensing system composed <strong>of</strong> two<br />

transcriptional regulators, SinR and ExpR, and the SinR-controlled autoinducer synthase SinI,<br />

which is responsible for the biosynthesis <strong>of</strong> the signal molecule in the form <strong>of</strong> an N-acyl<br />

homoserine lactone (AHL). These AHLs, in conjunction with the ExpR regulator, control a<br />

variety <strong>of</strong> downstream genes. The concentration <strong>of</strong> AHLs varies with changes in population<br />

density. As a result, expression <strong>of</strong> quorum-sensing-dependent genes may exhibit different<br />

patterns during various stages <strong>of</strong> bacterial growth. Work in our laboratory has shown that the S.<br />

meliloti ExpR/Sin quorum-sensing system regulates over 200 genes, including those involved in<br />

exopolysaccharide synthesis, motility and chemotaxis, metal transport, and other metabolic<br />

functions, thereby playing an important role during plant-bacteria interactions.<br />

Inoculation <strong>of</strong> plants with a sinI-deficient strain results in a delay in invasion as well as a<br />

significant reduction in the total number <strong>of</strong> nodules per plant when compared to the wild type,<br />

resulting in plant development deficiencies. Concurrently, expression <strong>of</strong> most <strong>of</strong> the motility and<br />

chemotaxis genes in the sinI mutant fail to be down-regulated by quorum sensing at high cell<br />

population density. Microarray and real-time PCR analyses revealed that the ExpR/Sin system<br />

adjusts the expression <strong>of</strong> the transcriptional regulators VisN/VisR and Rem, which in turn<br />

modulate downstream motility genes in a population-density-dependent manner to decrease<br />

motility. Recently we have shown that mutating flagellar production in a sinI mutant restores<br />

bacterial competency for symbiosis establishment to wild type levels, suggesting that the<br />

elimination <strong>of</strong> flagella during the invasion process is crucial. Therefore, down-regulation <strong>of</strong><br />

motility and chemotaxis by the ExpR/Sin quorum-sensing system plays an essential role in<br />

successful plant invasion by S. meliloti.<br />

44


<strong>BLAST</strong> X Thurs. Evening Session<br />

ENGINEERED SINGLE- AND MULTI-CELL CHEMOTAXIS IN E. COLI<br />

Shalom D. Goldberg 1 , Paige Derr 2 , William F. DeGrado 1 , and Mark Goulian 2,3<br />

1 Department <strong>of</strong> Biochemistry and Biophysics, <strong>University</strong> <strong>of</strong> Pennsylvania School <strong>of</strong> Medicine,<br />

Philadelphia, PA, 19104<br />

2 Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Pennsylvania, 209 South 33rd Street, Philadelphia, PA<br />

19104<br />

3 Department <strong>of</strong> Biology, <strong>University</strong> <strong>of</strong> Pennsylvania, 433 S <strong>University</strong> Avenue, Philadelphia, PA<br />

19104<br />

We have engineered the chemotaxis system <strong>of</strong> E. coli to enable responses to molecules<br />

that are not attractants for wild-type cells. The system depends on an artificially introduced<br />

enzymatic activity that converts the target molecule into a ligand for an E. coli chemoreceptor,<br />

thereby allowing the cells to respond to the new attractant. Two systems, designed to respond<br />

to asparagine and to phenylacetyl glycine respectively, showed robust chemotactic responses.<br />

In addition, their behavior in a mixed population was suggestive <strong>of</strong> a “hitchhiker” effect in which<br />

cells producing the ligand can induce chemotaxis <strong>of</strong> neighboring cells lacking the enzymatic<br />

activity. This behavior was exploited to design a complex system <strong>of</strong> two strains that are mutually<br />

interdependent for their activity, which functions as a simple microbial consortium.<br />

45


<strong>BLAST</strong> X Thurs. Evening Session<br />

PHOTO-ENERGY CONVERSION AND SENSORY TRANSDUCTION OF MICROBIAL<br />

RHODOPSINS IN PHOTOSYNTHETIC MICROBES<br />

So Young Kim, Keon Ah Lee, Ah Reum Choi, Song-I Han, and Kwang-Hwan Jung<br />

Department <strong>of</strong> Life Science and Interdisciplinary Program <strong>of</strong> Integrated Biotechnology, Sogang<br />

Univeristy, Seoul 121-742, Korea (kjung@sogang.ac.kr)<br />

Microbial rhodopsins, seven transmembrane proteins which contain all-trans/13 cis<br />

retinal as a chromophore, have been known for three decades and extensively studies in<br />

extreme halophiles. Photosynthetic microbes possess lots <strong>of</strong> photoactive proteins including<br />

chlorophyll-based pigments, phytochromes, phototropin-related blue light receptors, and<br />

cryptochromes. Surprisingly, recent genome sequencing projects discovered additional<br />

photoactive receptors, retinal-based rhodopsins, in cyanobacterial and algal genera. Analysis <strong>of</strong><br />

the Anabaena and Chlamyrhodopsin revealed that they have sensory functions, which based on<br />

our work with haloarchaeal rhodopsins, may use a variety <strong>of</strong> signaling mechanisms. Anabaena<br />

rhodopsin is interacted with a tetramer <strong>of</strong> 14kDa soluble transducer (ASRT) and one <strong>of</strong> their<br />

putative functions is a global regulation <strong>of</strong> phycobilin protein. The Anabaena rhodopsin shows a<br />

visible light-absorbing pigment (540-550nm) and it has mixed photochemical reaction <strong>of</strong> all trans<br />

and 13 cis form <strong>of</strong> retinal in ground state. Two Chlamydomonas rhodopsins are involved in<br />

phototaxis and photophobic responses based on electrical measurements by RNAi experiment.<br />

The rhodopsins from Gloeobacter violaceus and Acetabularia acetabulum is light-driven proton<br />

pump coexisted with photosynthetic machinery. The genes were functionally expressed in<br />

Escherichia coli and bound all-trans retinal to form a pigment in the presence <strong>of</strong> N- and Cterminal<br />

MISTIC sequences. Gloeobacter and Acetabularia rhodopsin I showed a light-driven<br />

proton pumping activity similar to proteorhodopsin.<br />

46


<strong>BLAST</strong> X Thurs. Evening Session<br />

FUNCTION OF MULTIPLE CHEMOTAXIS-LIKE PATHWAYS IN MEDIATING CHANGES IN<br />

MOTILITY PATTERNS AND CELLULAR MORPHOLOGY IN AZOSPIRILLUM BRASILENSE<br />

Amber N. Bible 1 , Zhihong Xie 1 , Matthew Russell 1 and Gladys Alexandre 1,2<br />

1 Department <strong>of</strong> Biochemistry, Cellular and Molecular Biology and 2 Department <strong>of</strong> Microbiology,<br />

The <strong>University</strong> <strong>of</strong> Tennessee, Knoxville, TN 37996<br />

Molecular details on bacterial chemotaxis have been derived from studies <strong>of</strong> model<br />

organisms such as Escherichia coli and Bacillus subtilis which genome encode for a single<br />

chemotaxis pathway that functions to modulate changes in motility patterns. Comparative<br />

genomics analysis indicates that the genome <strong>of</strong> many bacteria possess multiple chemotaxis-like<br />

(Che) pathways. A. brasilense is a plant-associated bacterium that can differentiate in at least<br />

four different cell types (swimmer, swarmer, aggregated and cyst cells). Transition from one cell<br />

type to the other depends on the environmental (especially nutritional) conditions. One <strong>of</strong> the 4<br />

Che-like pathways (Che1) encoded within the genome <strong>of</strong> the alphaproteobacterium A.<br />

brasilense was recently shown to regulate changes in motility patterns, cell-to-cell aggregation<br />

concomitant with changes in cell length (Bible et al., 2008). We will present evidence for the role<br />

<strong>of</strong> two Che pathways and several chemoreceptors in controlling the ability <strong>of</strong> cells to modulate<br />

multiple cellular responses, including cell length, that suggest that cross-regulation between<br />

parallel chemotaxis pathways may function to coordinate and integrate a set <strong>of</strong> cellular<br />

functions. Experimental evidence suggests that proteins that function in the<br />

methylation/demethylation <strong>of</strong> chemoreceptors may have a critical role in this cross-regulation.<br />

The implications in the lifestyle <strong>of</strong> this bacterium will also be discussed in lights <strong>of</strong> recent<br />

experimental evidence obtained.<br />

Bible, A. N., Stephens, B. B., Ortega, D. R., Xie, Z. and G. Alexandre (2008) Function <strong>of</strong> a<br />

chemotaxis-like signal transduction pathway in modulating motility, cell clumping, and cell length<br />

in the alphaproteobacterium Azospirillum brasilense. J Bacteriol 190: 6365-6375.<br />

47


<strong>BLAST</strong> X Thurs. Evening Session<br />

PROBING ADAPTATION KINETICS IN VIVO BY FLUORESCENCE RESONANCE ENERGY<br />

TRANSFER<br />

Thomas S. Shimizu 1 , Yuhai Tu 2 and Howard C. Berg 1<br />

1 Department <strong>of</strong> Molecular & Cellular Biology, Harvard <strong>University</strong>, Cambridge, MA 02138.<br />

2 T. J. Watson Research Center, IBM, Yorktown Heights, NY 10598.<br />

Bacteria sense spatial gradients by taking time derivatives <strong>of</strong> ligand concentrations<br />

measured during runs <strong>of</strong> a random walk 1 . The remarkable sensitivity to shallow gradients in<br />

Escherichia coli has been explained mainly by cooperativity between receptors and<br />

ultrasensitivity <strong>of</strong> the flagellar motor. We have revisited the experimental findings <strong>of</strong> Block,<br />

Segall and Berg 2 , where the chemotactic response <strong>of</strong> tethered cells to time-varying stimuli were<br />

characterized quantitatively. A simple theoretical model 3 that combines robust adaptation 4 with<br />

an allosteric model <strong>of</strong> receptor cooperativity 5-7 can explain the general features <strong>of</strong> responses to<br />

temporal ramps and oscillatory stimuli.<br />

A notable feature <strong>of</strong> this model is that the steady-state amplitude <strong>of</strong> responses to<br />

exponential ramps do not depend on the degree <strong>of</strong> receptor cooperativity (the parameter N <strong>of</strong><br />

an MWC-type allosteric model 5 ). The time required to reach this steady state, however, depends<br />

inversely on N, so cooperativity speeds up computation <strong>of</strong> the derivative signal, but does not<br />

determine its amplitude. The latter is instead determined by the adaptation kinetics, and this<br />

relation allows us to infer quantitative characteristics <strong>of</strong> adaptation in vivo from measured rampresponse<br />

data.<br />

Here we present novel experiments in which the chemotactic responses <strong>of</strong> E. coli<br />

populations during time-varying stimuli are monitored by fluorescence resonance energy<br />

transfer 8 (FRET). This approach is far more efficient than the earlier experiments <strong>of</strong> Block et al. 2 ,<br />

in which the chemotactic responses <strong>of</strong> individual cells were characterized through the stochastic<br />

output <strong>of</strong> the motor. We find that the sensitivity <strong>of</strong> E. coli to gradients depends strongly on<br />

temperature, and using our model framework, we analyze how ultrasensitvity in the adaptation<br />

system 9 contributes to gradient sensitivity in vivo.<br />

REFERENCES<br />

1. Berg, H. C. & Brown, D. A. Chemotaxis in Escherichia coli analysed by three-dimensional<br />

tracking. Nature 239, 500-4 (1972).<br />

2. Block, S. M., Segall, J. E. & Berg, H. C. Adaptation kinetics in bacterial chemotaxis. J Bacteriol<br />

154, 312-23 (1983).<br />

3. Tu, Y., Shimizu, T. S. & Berg, H. C. Modeling the chemotactic response <strong>of</strong> Escherichia coli to<br />

time-varying stimuli. Proc Natl Acad Sci U S A 105, 14855-60 (2008).<br />

4. Barkai, N. & Leibler, S. Robustness in simple biochemical networks. Nature 387, 913-7 (1997).<br />

5. Monod, J., Wyman, J. & Changeux, J. P. On the Nature <strong>of</strong> Allosteric Transitions: A Plausible<br />

Model. J Mol Biol 12, 88-118 (1965).<br />

6. Mello, B. A. & Tu, Y. An allosteric model for heterogeneous receptor complexes: Understanding<br />

bacterial chemotaxis responses to multiple stimuli. Proc Natl Acad Sci U S A 102, 17354-9<br />

(2005).<br />

7. Keymer, J. E., Endres, R. G., Skoge, M., Meir, Y. & Wingreen, N. S. Chemosensing in<br />

Escherichia coli: two regimes <strong>of</strong> two-state receptors. Proc Natl Acad Sci U S A 103, 1786-91<br />

(2006).<br />

8. Sourjik, V., Vaknin, A., Shimizu, T. S. & Berg, H. C. In Vivo Measurement by FRET <strong>of</strong> Pathway<br />

Activity in Bacterial Chemotaxis. Methods Enzymol 423, 363-91 (2007).<br />

9. Emonet, T. & Cluzel, P. Relationship between cellular response and behavioral variability in<br />

bacterial chemotaxis. Proc Natl Acad Sci U S A 105, 3304-9 (2008).<br />

48


<strong>BLAST</strong> X Thurs. Evening Session<br />

MINOR RECEPTOR SIGNALLING IN E. COLI<br />

Silke Neumann, Ned Wingreen* and Victor Sourjik<br />

Ruprecht-Karls-Universität Heidelberg, Zentrum für molekulare Biologie Heidelberg (ZMBH), Im<br />

Neuenheimer Feld 282, 69120 Heidelberg, Germany<br />

* Department <strong>of</strong> Molecular Biology - Princeton <strong>University</strong><br />

Ligand recognition in the chemotaxis pathway <strong>of</strong> E. coli proceeds through binding <strong>of</strong><br />

ligands to transmembrane receptors, either directly or indirectly through periplasmic binding<br />

proteins. E. coli has five types <strong>of</strong> receptors, with two high-abundance (or major) receptors – Tsr<br />

for serine and Tar for aspartate and maltose – and three low-abundance (or minor) receptors –<br />

Tap for dipeptides, Trg for ribose, galactose and glucose, and Aer for redox potential. Together<br />

with the histidine kinase CheA, receptors form chemosensory complexes which in turn are<br />

organized in tight clusters where receptors <strong>of</strong> different ligand specificities are intermixed. Signal<br />

processing is thought to occur within these receptor clusters through allosteric interactions<br />

between receptor dimers. To compare signal processing by minor and major receptors, we<br />

systematically investigated responses mediated by Trg and Tap, and by Tar and Tsr in respect<br />

to response sensitivity, relation between receptor occupancy and kinase inactivation, dynamic<br />

range <strong>of</strong> the response, adaptation time to a range <strong>of</strong> stimuli, as well as integration <strong>of</strong> signals that<br />

are sensed by different receptors using an in vivo FRET-based kinase assay. Our experimental<br />

analysis shows that signals are amplified and integrated differently by the two receptor<br />

populations, but in both cases signal processing can be quantitatively explained by the same<br />

allosteric model.<br />

49


<strong>BLAST</strong> X Thurs. Evening Session<br />

A SYSTEMS BIOLOGY APPROACH TO UNDERSTANDING HOW BACILLUS MAKES UP ITS<br />

MIND<br />

Arnaud Chastanet 1 , Guocheng Yuan 2 , Thomas M. Norman 1 , Jun Liu 3 and Richard Losick 1<br />

1 Molecular and Cellular Biology Department, Harvard <strong>University</strong>,<br />

2 Department <strong>of</strong> Biostatistics and Computational Biology, Harvard School <strong>of</strong> Public Health,<br />

3 Department <strong>of</strong> Statistics, Harvard <strong>University</strong>.<br />

Understanding how cells make decisions and differentiate are key biological questions.<br />

Mechanisms underlying such behaviors integrate multiple environmental signals in intricate<br />

networks in order to appropriately respond to the situations. The sporulation process that takes<br />

place in Bacillus subtilis under adverse conditions perfectly exemplifies this kind <strong>of</strong> question. For<br />

this, numerous signals and control systems are integrated at the level <strong>of</strong> a ”decider” protein<br />

called Spo0A, a transcriptional regulator belonging to the two-component systems family. The<br />

decision to sporulate is taken during the first two hours after optimal sporulation conditions have<br />

been reached. During this time, Spo0A accumulates slowly reaching a high level at hour two. It<br />

has been previously shown that while some <strong>of</strong> its targets are activated at low concentration,<br />

thus early on, others are switched on later, when the maximal quantity <strong>of</strong> the regulator has been<br />

achieved. Interestingly, even in optimal conditions, only a fraction <strong>of</strong> the population will finally<br />

decide to sporulate, a phenomenon described as bistability.<br />

We are attempting to understand how this two-stage activation <strong>of</strong> Spo0A is achieved<br />

through an interdisciplinary approach combining the methods <strong>of</strong> genetics and mathematics. The<br />

time resolved picture <strong>of</strong> the regulatory process we have obtained has revealed a multiple step<br />

process involving successive switches. First, Spo0A activity is rising during log phase, activating<br />

some switches. Then under conditions <strong>of</strong> nutrient limitation, Spo0A is further activated to a mid<br />

and variable extent throughout the population. During this period, “low-threshold” genes are<br />

turned ON. In an ultimate step, a bistable switch allows Spo0A to be activated to a high level but<br />

only in a portion <strong>of</strong> the population. These cells express high threshold genes and proceed to<br />

sporulate.<br />

50


POSTER ABSTRACTS<br />

51


<strong>BLAST</strong> X ______ Poster #1<br />

THE CHARACTERISATION OF THE DYNAMICS OF THE FliT:FliD:FlhD4C2 INTERACTION<br />

AND ITS ROLE IN REGULATING FLAGELLAR ASSEMBLY<br />

C. Aldridge 1,2 , K. Poonchareon 1,2 , S. Saini 3 , A. Soloyva 2 , M. Banfield 4 , T. Minamino 5,6 , C. V.<br />

Rao 3 , and P. D. Aldridge 1,2<br />

1: Centre for Bacterial Cell Biology, Newcastle <strong>University</strong>, Framlington Place, Newcastle upon<br />

Tyne, United Kingdom, NE2 4HH. 2: Institute for Cell and Molecular Biosciences, Newcastle<br />

<strong>University</strong>, Framlington Place, Newcastle upon Tyne, United Kingdom, NE2 4HH. 3: Department<br />

<strong>of</strong> Chemical and Biomolecular Engineering, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, Urbana,<br />

Illinois, United States, 61801. 4: Dept. <strong>of</strong> Biological Chemistry, John Innes Centre, Norwich,<br />

NR4 7UH UK. 5: Graduate School <strong>of</strong> Frontier Biosciences, Osaka <strong>University</strong>, 1-3 Yamadaoka,<br />

Suita, Osaka 565-0871, Japan. 6: Dynamic NanoMachine Project, ICORP, JST, 1-3<br />

Yamadaoka, Suita, Osaka 565-0871, Japan.<br />

Each bacterial cell <strong>of</strong> Salmonella enterica serovar Typhimurium produces a discrete<br />

number <strong>of</strong> complete flagella. Flagellar gene expression in Salmonella is negatively regulated by<br />

the Type 3 secretion chaperone FliT. FliT is known to interact with the flagellar structural subunit<br />

FliD and the master transcriptional regulator FlhD4C2. In this regulatory circuit FliD is proposed<br />

to act as an anti-regulator - a regulatory role similar to that observed for FlgM inhibition <strong>of</strong> s 28<br />

activity. We were interested in determining the kinetics <strong>of</strong> the FliT:FliD:FlhD4C2 regulatory circuit<br />

and how they influence flagellar assembly. We have shown that the FliT:FliD interaction is a 1:1<br />

ratio while in solution FliT is a dimer. Surface plasma resonance (SPR) and analytical<br />

ultracentrifugation (AUC) analysis <strong>of</strong> the dynamics <strong>of</strong> the FliT:FlhD4C2 interaction showed that it<br />

was very different from FliT:FliD. A FliT:FlhD4C2 complex is only observed when excess FliT is<br />

added to FlhD4C2. AUC analysis identified that a stable intermediate <strong>of</strong> a combination <strong>of</strong> FlhD,<br />

FlhC and/or FliT exists. Our data suggests that FliT acts to dissociate the FlhD4C2 complex and<br />

that this is a transient interaction allowing for the FlhD4C2 complex to reform. This suggests that<br />

a limiting factor in the regulation <strong>of</strong> FlhD4C2 activity is the concentration <strong>of</strong> free FliT. To test this<br />

model, we have investigated the affect <strong>of</strong> overexpressing FliT on flagellar gene expression and<br />

basal body assembly.<br />

52


<strong>BLAST</strong> X ______ Poster #2<br />

SUBUNIT FEEDBACK CONTROL OF FLAGELLAR FILAMENT ASSEMBLY IN<br />

CAULOBACTER CRESCENTUS<br />

Phillip D. Aldridge 1,2§ Alexandra Faulds-Pain 1,2 , Christine Aldridge 1,2 , Giulia Grimaldi 1,2 ,<br />

Christopher Birchall 1,2 , Shuichi Nakamura 3 , Tomoko Miyata 3 , Joe Gray 4 , Guanglai Li 5 , Jay<br />

Tang 5 , Keiichi Namba 3,6 , Tohru Minamino 3,6<br />

1: Centre for Bacterial Cell Biology, Newcastle <strong>University</strong>, Framlington Place, Newcastle upon<br />

Tyne, United Kingdom, NE2 4HH 2: Institute for Cell and Molecular Biosciences, Newcastle<br />

<strong>University</strong>, Framlington Place, Newcastle upon Tyne, United Kingdom, NE2 4HH 3: Graduate<br />

School <strong>of</strong> Frontier Biosciences, Osaka <strong>University</strong>, 1-3 Yamadaoka, Suita, Osaka 565-0871,<br />

Japan. 4: Pinnacle, Newcastle <strong>University</strong>, Framlington Place, Newcastle upon Tyne, United<br />

Kingdom, NE2 4HH 5: Physics Department, Brown <strong>University</strong>, 184 Hope Street, Providence, RI<br />

02912, USA 6: Dynamic NanoMachine Project, ICORP, JST, 1-3 Yamadaoka, Suita, Osaka<br />

565-0871, Japan.<br />

Bacterial flagellar filaments play key roles in surface attachment and host-bacterial<br />

interactions as well as motility. Approximately 40% <strong>of</strong> annotated flagellar systems (n = 278)<br />

utilise multiple flagellin variants to assemble their flagellar filaments. Here we have investigated<br />

the ability <strong>of</strong> the model flagellar system <strong>of</strong> Caulobacter crescentus to assemble its flagellar<br />

filament from six flagellins: FljJ, FljK, FljL, FljM, FljN and FljO. A flagellin gene mutant collection<br />

<strong>of</strong> multiple gene deletion combinations, exhibited a range <strong>of</strong> Mot phenotypes from impaired<br />

motility (Mot -/+ ) to motile (Mot + ). Further characterisation <strong>of</strong> the mutant collection showed: 1)<br />

there is no strict requirement for all six flagellins to assemble a filament exhibiting wild type<br />

characteristics; 2) a correlation between slower swimming speeds and shorter filament lengths<br />

in all ∆fljK ∆fljM Mot + mutants; and 3) the flagellins FljM – FljO are less stable than FljJ – FljL.<br />

Our data suggests that the flagellins FljJ, FljK and FljL play both a regulatory and structural role<br />

during filament assembly. In contrast, the flagellins FljM to FljO possess only a structural role.<br />

We propose the model that the observed order <strong>of</strong> multiple flagellin incorporation in C.<br />

crescentus, and plausibly other flagellar systems, is a result <strong>of</strong> the system coupling flagellin<br />

synthesis to filament assembly.<br />

53


<strong>BLAST</strong> X ______ Poster #3<br />

FUNCTION OF UNIQUE DOMAINS OF CheA1 FROM A. BRASILENSE IN REGULATING<br />

MULTIPLE CELLULAR BEHAVIORS<br />

Amber N. Bible and Gladys Alexandre<br />

Department <strong>of</strong> Biochemistry, Cellular, and Molecular Biology at the <strong>University</strong> <strong>of</strong> Tennessee,<br />

Knoxville<br />

The alpha-proteobacterium Azospirillum brasilense contains four different chemotaxis<br />

operons. One <strong>of</strong> the 4 Che-like pathways (Che1) encoded within the genome <strong>of</strong> the<br />

alphaproteobacterium A. brasilense was recently shown to regulate changes in motility patterns,<br />

cell-to-cell aggregation (clumping) concomitant with changes in cell length (Bible et al., 2008).<br />

Mutations affecting cheA1 decrease chemotaxis, cell length, but lead to an increase in<br />

clumping relative to the wild type A. brasilense. Interestingly, CheA1 is a hybrid histidine kinase<br />

with two P5-like domains and a REC-like receiver domain at its C-terminus. In addition, the Nterminus<br />

<strong>of</strong> CheA1 comprises a highly conserved polytopic domain <strong>of</strong> unknown function,<br />

suggesting that CheA1 may be a membrane-bound protein. Experimental data indicate that<br />

CheA1 is produced as a membrane-bound protein that localizes to the cell pole, similar to other<br />

Che1 proteins. Noticeably, the polytopic N-terminal domain <strong>of</strong> CheA1 is not required for the<br />

localization <strong>of</strong> CheA1 neither to the cell pole, nor for changes in cell length or cell-to-cell<br />

aggregation but it is essential for wild-type chemotaxis and aerotaxis behaviors. Data obtained<br />

from a combination <strong>of</strong> in-frame deletions <strong>of</strong> the C-terminal domains and site-specific<br />

mutagenesis approaches with behavioral assays suggest that the second P5-like domain <strong>of</strong><br />

CheA1 has a role in modulating changes in cell length. A working model for the functions <strong>of</strong> the<br />

N- and C-terminal domains <strong>of</strong> CheA1 in modulating chemotaxis, aerotaxis, clumping and cell<br />

length changes will be presented in lights <strong>of</strong> recent experimental data.<br />

Bible, A. N., Stephens, B. B., Ortega, D. R., Xie, Z. and G. Alexandre (2008) Function <strong>of</strong> a<br />

chemotaxis-like signal transduction pathway in modulating motility, cell clumping, and cell length<br />

in the alphaproteobacterium Azospirillum brasilense. J Bacteriol 190: 6365-6375.<br />

54


<strong>BLAST</strong> X ___ Poster #4<br />

HOW DOES THE RHODOBACTER SPHAEROIDES FLAGELLAR MOTOR STOP – USING A<br />

CLUTCH OR A BRAKE?<br />

M. Brown, T. Pilizota, M. Leake, R. Berry, J. Armitage<br />

<strong>University</strong> <strong>of</strong> Oxford, Oxford, UK<br />

Unlike most species with bidirectional motors, Rhodobacter sphaeroides employs a<br />

unidirectional stop-start flagellar motor, where stops are analogous to tumbles. By controlling<br />

when the motor stops, cells can accumulate in areas favourable for their survival.<br />

We asked the question, how do the CheYs stop motor rotation in R. sphaeroides; by<br />

causing the torque-generating units to disengage from the rotor, allowing free rotational<br />

movement, or by jamming the rotor, locking it in a particular configuration?<br />

These hypothesises were tested by applying external force (viscous flow or optical<br />

tweezers) to chemotactically stopped motors.<br />

We found that the motor is stopped with a brake mechanism and that approximately 3-4<br />

times more torque acts on the motor when stopped than when it rotates. Furthermore, by<br />

monitoring the position <strong>of</strong> sub-micron beads attached to flagella stubs we discovered that stops<br />

can only occur at a number <strong>of</strong> discrete angles. Analysis is underway to determine the number<br />

<strong>of</strong> steps per revolution.<br />

55


<strong>BLAST</strong> X ______ Poster #5<br />

DYNAMICS OF THE FLAGELLAR MOTOR PROTEIN FliM<br />

Nicolas J. Delalez<br />

Microbiology Unit, Biochemistry Dept, <strong>University</strong> <strong>of</strong> Oxford, South Parks Road, Oxford OX1<br />

3QU, UK<br />

Escherichia coli swims by rotating 4-6 long helical filaments to propel the cell through its<br />

environment. The flagellar motor that rotates the filament is bidirectional and composed <strong>of</strong> two<br />

parts: the rotor and the stator, the stator being the fixed component against which the rotor<br />

spins. The stator is composed <strong>of</strong> units <strong>of</strong> two integral membrane proteins, MotA and MotB, the<br />

stoichiometry <strong>of</strong> each unit being (MotA4:MotB2). The rotor is composed <strong>of</strong> multiple rings,<br />

including the C-ring which is localized at the base <strong>of</strong> the motor and is the switch complex that<br />

gives the bidirectionality to the E. coli motor. The C-ring comprises FliG (~ 25 copies), FliM (~34<br />

copies) and FliN (> 100 copies).<br />

By using a TIRF microscope, and expressing GFP-MotB from the genome <strong>of</strong> E. coli in<br />

place <strong>of</strong> the wild-type gene, it has recently been possible to monitor the protein stoichiometry,<br />

dynamics and turnover <strong>of</strong> this stator component with single-molecule precision in functioning<br />

bacterial flagellar motors. Repeating these experiments with different flagellar motor proteins will<br />

greatly enhance our understanding <strong>of</strong> the mechanism <strong>of</strong> this motor.<br />

In particular, one <strong>of</strong> the main questions for our understanding is the mechanism <strong>of</strong> the<br />

switching process. The C-terminus <strong>of</strong> FliM, a component <strong>of</strong> the C-ring, has therefore been<br />

tagged with Ypet and expressed from the genome. Data will be presented on the construction <strong>of</strong><br />

this mutant as well as data on its dynamics, turnover and stoichoimetry, and the influence <strong>of</strong> the<br />

response regulator CheY on these features. The consequences <strong>of</strong> these results and future work<br />

will be discussed.<br />

56


<strong>BLAST</strong> X Poster #6<br />

USING CONTROL THEORY TO ELUCIDATE CONNECTIVITY IN R. SPHAEROIDES<br />

CHEMOTAXIS<br />

Mark A. J. Roberts 1 , Elias August 2 , Judith P. Armitage 1 and Antonis Papachristodoulou 3<br />

1<br />

Department <strong>of</strong> Biochemistry, <strong>University</strong> <strong>of</strong> Oxford, South Parks Road, Oxford, OX1 3QU, UK<br />

2<br />

Control Group, Department <strong>of</strong> Engineering Science, Oxford <strong>University</strong>, Parks Road, Oxford,<br />

OX1 3PJ, UK<br />

3<br />

Oxford Centre for Integrative Systems Biology, Department <strong>of</strong> Biochemistry, South Parks Road,<br />

Oxford, OX1 3QU, UK<br />

With an increasing number <strong>of</strong> sequenced bacterial genomes it becomes evident that the<br />

chemotactic sensory mechanism <strong>of</strong> bacteria is more complex than E. coli. In this poster we<br />

describe how ideas from engineering control theory can be used to develop a novel approach<br />

for designing experiments in order to elucidate the biochemical network structure <strong>of</strong> signalling<br />

pathways in general. The goal is to develop a systematic approach for finding the best<br />

experiment that will delineate the network structure.<br />

We then apply this method to the chemotaxis pathway <strong>of</strong> R. sphaeroides, which has<br />

multiple homologues <strong>of</strong> the E. coli proteins. To achieve this we are constructing, in silico,<br />

various possible models <strong>of</strong> R. sphaeroides chemotaxis that can explain experimental<br />

observations. These models include the different possible interactions for the CheB and CheY<br />

proteins. Applying results from optimal control theory, we determined the best input (ligand)<br />

pr<strong>of</strong>ile that gives an output which would allow us to discriminate best between the proposed<br />

models, aiming to invalidate some <strong>of</strong> them. This input ligand pr<strong>of</strong>ile is then administered to R.<br />

sphaeroides in a flow cell and the response is measured using a tethered cell assay. We have<br />

also developed methods to determine the best initial conditions to discriminate between the<br />

models, based on the limitations <strong>of</strong> what can be implemented biochemically, and these were<br />

then also tested in a tethered cell assay.<br />

We used the experimental results from these designed tethered cell experiments to<br />

invalidate some <strong>of</strong> the proposed network structures and hence suggest a probable network<br />

connectivity for the multiple CheY and CheB proteins within R. sphaeroides.<br />

This is an exciting approach to determine network structures in a fast and efficient<br />

manner and can be applied to a wide range <strong>of</strong> signalling pathways as well as potentially<br />

allowing chemotaxis pathways in other species using published genomes to generate the<br />

necessary models.<br />

57


<strong>BLAST</strong> X Poster #7<br />

BEHAVIOR OF THE FLAGELLAR ROTARY MOTOR NEAR ZERO LOAD<br />

Junhua Yuan and Howard Berg<br />

Department <strong>of</strong> Molecular and Cellular Biology, Harvard <strong>University</strong>, Cambridge, MA 02138<br />

The physiology <strong>of</strong> the flagellar rotary motor has been studied extensively in the regime <strong>of</strong><br />

relatively high load and low speed. Here, we describe an assay that allows systematic study <strong>of</strong><br />

the motor near zero load. Sixty-nanometer-diameter gold spheres were attached to hooks <strong>of</strong><br />

cells lacking flagellar filaments, and light scattered from a sphere was monitored at the image<br />

plane <strong>of</strong> a microscope through a small pinhole. Resurrection experiments were carried out near<br />

zero load. Paralyzed motors <strong>of</strong> cells carrying a motA point mutation were resurrected at 23°C by<br />

expression <strong>of</strong> wild-type MotA, and speeds jumped from zero to a maximum value <strong>of</strong> about 300<br />

Hz in one step. The temperature dependence <strong>of</strong> the speed near zero load also was studied and<br />

showed a high activation enthalpy comparable to that observed previously in electrorotation<br />

experiments. The assay has been modified so that both the speed and the direction <strong>of</strong> rotation<br />

can be monitored near zero load. Switching properties <strong>of</strong> the flagellar motor near zero load are<br />

under investigation.<br />

58


<strong>BLAST</strong> X ______ Poster #8<br />

NUTRIENT SENSING BY A HUMAN GUT SYMBIONT<br />

Hongjun Zheng 1 , Susan Firbank 1 , Eric Martens 2 , Edith Diaz 1 , Rick Lewis 1 , Jeff Gordon 2 and<br />

David Bolam 1<br />

1. Institute for Cell and Molecular Biosciences, Newcastle <strong>University</strong>, The Medical School,<br />

Framlington Place, Newcastle upon Tyne NE2 4HH, UK<br />

2. Center for Genome Sciences, Washington <strong>University</strong>, St Louis, USA.<br />

The gut microbiota play an important role in human health and nutrition. Bacteroides<br />

thetaiotaomicron (Bt) is a dominant gut symbiont whose main sources <strong>of</strong> carbon and energy are<br />

dietary and host polysaccharides, reflected in the presence on the genome <strong>of</strong> a large number <strong>of</strong><br />

genes involved in the sensing, acquisition and processing <strong>of</strong> complex carbohydrates. One <strong>of</strong> the<br />

most striking features <strong>of</strong> the Bt genome are 33 genes encoding novel hybrid two component<br />

systems (HTCS) that contain all <strong>of</strong> the domains <strong>of</strong> a classical TCS in a single polypeptide.<br />

Custom Genechip arrays reveal that when Bt is grown in the presence <strong>of</strong> inulin, a β-1,2 linked<br />

fructose storage polysaccharide, a specific locus composed <strong>of</strong> eight genes predicted to be<br />

involved in polysaccharide binding, uptake and degradation is activated. The closest regulatory<br />

protein to this locus is an HTCS, BT1754. Genetic, biochemical and structural studies reveal<br />

that BT1754 is the sensor that controls fructan utilisation and the identity <strong>of</strong> the signalling<br />

molecule itself, as well as the mechanism <strong>of</strong> signal perception and insights into signal<br />

transduction across the cytoplasmic membrane.<br />

59


<strong>BLAST</strong> X Poster #9<br />

THE LeuO REGULON IN SALMONELLA<br />

Edmundo Calva, Víctor H. Bustamante, Miguel Ángel De la Cruz, Marcos Fernández-Mora,<br />

Mario Alberto Flores-Valdez, Ana Lucía Gallego-Hernández, Carmen Guadarrama, Ismael<br />

Hernández-Lucas, Omar Ortega, Alejandra Vázquez, and Ricardo Oropeza.<br />

Instituto de Biotecnología UNAM, Cuernavaca, Morelos, Mexico.<br />

LeuO is a LysR-type transcriptional regulator, which antagonizes the repressing effect by<br />

the H-NS and StpA nucleoid proteins on the Salmonella enterica serovar Typhi ompS1<br />

quiescent porin gene, allowing the transcriptional activation by OmpR among other regulators.<br />

LeuO is also a positive regulator for the OmpS2 quiescent porin, for a quiescent putative<br />

periplasmic arylsulfate sulfotransferasa (AssT), and for an open reading frame <strong>of</strong> unknown<br />

function, STY3068. It is a negative regulator for the OmpX porin, a thiol peroxidase (Tpx), and<br />

for STY1978. These proteins were identified by a whole-genome proteomic analysis and the<br />

regulatory effects were confirmed on transcriptional gene reporter fusions. STY 3068 appears to<br />

be part <strong>of</strong> an operon (5’-STY3070-STY3064-3’), as transcriptional regulation occurs upstream,<br />

at STY 3070. The assT gene forms a LeuO-dependent operon with STY3371 and STY3372,<br />

which are paralogues <strong>of</strong> the dsbA and dsbB genes, respectively. This operon (5’-assT-<br />

STY3371-STY3372-3’) is repressed by OmpR.<br />

H-NS represses the expression <strong>of</strong> ompS1, assT and STY3070 and, on the other hand,<br />

activates the expression <strong>of</strong> tpx and STY1978. Expression <strong>of</strong> ompX was independent <strong>of</strong> H-NS.<br />

Interestingly, ompS2 is negatively regulated by H-NS in Escherichia coli and in S. Typhimurium,<br />

although in S. Typhi other negative regulators are involved aside from H-NS. LeuO specifically<br />

binds to the 5´ intergenic regions <strong>of</strong> ompS1, ompS2, assT, STY3070, ompX, and tpx. LeuO did<br />

not bind to the promoter region <strong>of</strong> STY1978, suggesting that it is regulating in an indirect<br />

manner.<br />

Hence, LeuO regulates a cadre <strong>of</strong> genes which appear to be mostly involved in the<br />

response to stress and in virulence.<br />

60


<strong>BLAST</strong> X Poster #10<br />

EnvZ-OmpR AND CpxA-CpxR REGULATE ompS1 BY DIFFERENTIAL PROMOTER<br />

EXPRESSION<br />

De la Cruz, M.A, Flores-Valdez, M.A., and Calva, E.<br />

Departamento de Microbiología Molecular, Instituto de Biotecnología UNAM Av. Universidad<br />

2001 Col. Chamilpa C.P. 62210 Cuernavaca, Morelos México<br />

The Salmonella enterica ompS1 gene encodes a quiescent porin that belongs to the<br />

OmpC/OmpF family. We recently reported that LeuO, OmpR, H-NS and StpA are regulators <strong>of</strong><br />

ompS1 expression. We have now detailed the mechanism <strong>of</strong> OmpR regulation. In vivo,<br />

phosphorylated OmpR (OmpR-P) showed to be determinant for the regulation <strong>of</strong> both ompS1<br />

promoters: OmpR-P activated the P1 promoter and repressed the P2 promoter, in an EnvZdependent<br />

manner. In vitro, OmpR-P had a seventy two-fold higher binding-affinity to the<br />

ompS1 promoter region than OmpR, being 2 to 20-fold higher than to the major porin-encoding<br />

ompC and ompF genes, respectively. In addition to EnvZ-OmpR, we found that CpxA-CpxR<br />

positively regulated ompS1 expression. CpxR, together OmpR, was necessary for the activation<br />

<strong>of</strong> the P1 promoter. CpxR also activated the P2 promoter, but only in the absence <strong>of</strong> OmpR. Our<br />

model involves LeuO as an antagonist that relieves H-NS and StpA-mediated silencing, allowing<br />

the binding <strong>of</strong> OmpR and CpxR. High osmolarity and acid pH, or both, stimulated ompS1<br />

expression in an OmpR and CpxR-independent manner, yet dependent on DNA supercoiling.<br />

61


<strong>BLAST</strong> X Poster #11<br />

THE COMPLEX HOOK BASAL BODY STRUCTURE OF THE LYME DISEASE SPIROCHETE<br />

BORRELIA BURGDORFERI<br />

K. Miller 1 , R. Duda 2 , R. Hendrix 2 , M. Motaleb 1,3 , and N.W. Charon 1<br />

1<br />

West Virginia <strong>University</strong> Health Sciences Center, Microbiology and Immunology Department,<br />

Box 9177, Room 2077, Morgantown, WV 26506<br />

2<br />

<strong>University</strong> <strong>of</strong> Pittsburgh, Department <strong>of</strong> Biological Sciences, 4249 Fifth Avenue, 357A,<br />

Crawford Hall, Pittsburgh, PA 15260<br />

3<br />

East Carolina <strong>University</strong>, Department <strong>of</strong> Microbiology and Immunology, Greenville, NC 27834<br />

FlgE is the structural protein <strong>of</strong> the flagellar hook <strong>of</strong> B. burgdorferi. Previous western blot<br />

analysis using polyclonal FlgE antibody indicated that the majority <strong>of</strong> the hook migrated as a<br />

high molecular weight complex in SDS-PAGE gels instead <strong>of</strong> at the position <strong>of</strong> monomeric FlgE<br />

(46 kDa). Both the high molecular complex and a small amount <strong>of</strong> the monomer were present in<br />

wild-type and complemented cells but absent in flgE mutants. These data suggest that FlgE<br />

may be cross-linked into oligomers in a manner similar to that found in bacteriophage capsid<br />

proteins and in some bacterial pili. High molecular weight FlgE complexes have been observed<br />

in other spirochete species (Treponema phagedenis and Treponema denticola), suggesting that<br />

this property is conserved. We are interested in determining whether or not FlgE is cross-linked,<br />

and what amino acids are involved in the cross-linking.<br />

The hook-basal body complexes were purified using a refined version <strong>of</strong> methods<br />

previously reported (Sal, et al, 2008). Following purification <strong>of</strong> periplasmic flagella that contained<br />

attached hook-basal body complexes, the filament portions were depolymerized by acidic<br />

conditions or the use <strong>of</strong> urea. Hook-basal body complexes were then isolated by pelleting using<br />

an ultracentrifuge recovered in a neutralizing buffer. Electron microscopy <strong>of</strong> these preparations<br />

showed that acidic conditions resulted in more intact hook-basal body complexes than urea<br />

treatment, although the high molecular weight forms <strong>of</strong> FlgE were found using both methods. In<br />

order to determine results from each experiment, the experimental samples were run on 10%<br />

SDS-PAGE gels. We are using proteolytic finger printing and peptide mapping methods to<br />

compare the high molecular weight complexes <strong>of</strong> FlgE with the monomeric form (histidinetagged<br />

recombinant). Hook-basal body complexes and recombinant FlgE were treated with<br />

proteases or hydroxylamine at various temperatures (37, 45, and 65°C) and run in SDS-PAGE<br />

gels. Western blotting was used to assess the stability <strong>of</strong> the FlgE complex as well as to<br />

determine the size <strong>of</strong> (antigenic) peptide fragments resulting from proteolytic digestion. Digest<br />

patterns were compared between recombinant FlgE and the isolated hook-basal body<br />

complexes.<br />

Refining the hook-basal body isolation procedure resulted in improved hook-basal body<br />

preparations. The FlgE complex is stable under harsh conditions (both urea and acid treatment).<br />

Differences and similarities in hydroxylamine digest patterns between recombinant FlgE and<br />

purified hook-basal body complexes supports the hypothesis that FlgE is cross-linked.<br />

62


<strong>BLAST</strong> X Poster #12<br />

HOW DO PAS AND HAMP DOMAINS COMMUNICATE? INSIGHTS FROM Aer2, A HEME<br />

BASED SENSOR FOR AEROTAXIS<br />

Michael Airola (1), Joanne Widom (1), Kylie Watts (2), Brian Crane (1)<br />

(1) Dept. <strong>of</strong> Chemistry and Chemical Biology, Cornell <strong>University</strong>, Ithaca, NY 14850<br />

(2) Dept. <strong>of</strong> Biochemistry and Microbiology, Loma Linda <strong>University</strong>, Loma Linda, CA 92350<br />

Thousands <strong>of</strong> PAS and HAMP domains have been identified yet how these common<br />

signal transduction domains function and communicate is still poorly understood. Structural<br />

studies <strong>of</strong> the HAMP domain have been especially difficult due to it <strong>of</strong>ten being contained within<br />

integral membrane proteins. Recently Aer2, from P. aeruginosa, has been identified as a<br />

soluble aerotaxis receptor that binds oxygen directly through a heme moiety located in the PAS<br />

domain. Using a combination <strong>of</strong> spectroscopic and structural techniques we sought to determine<br />

if oxygen binding induces large-scale conformational changes. Initial results point to an<br />

interaction between the PAS and HAMP domains <strong>of</strong> Aer2 giving insight to how these<br />

widespread signaling modules may function.<br />

63


<strong>BLAST</strong> X Poster #13<br />

STRUCTURE OF SOLUBLE CHEMORECEPTOR SUGGESTS A MECHANISM FOR<br />

PROPAGATING CONFORMATIONAL SIGNALS<br />

Abiola Pollard and Brian Crane<br />

Cornell <strong>University</strong>, Chemistry Research Bldg. G63, Ithaca, NY 14853<br />

Transmembrane chemoreceptors, also known as methyl-accepting chemotaxis proteins<br />

(MCP’s), translate extracellular signals into intracellular responses in the bacterial chemotaxis<br />

system. MCP ligand binding domains control the activity <strong>of</strong> the CheA kinase, situated ~200 Å<br />

away, across the cytoplasmic membrane. The 2.15 Å resolution crystal structure <strong>of</strong> a T.<br />

maritima soluble receptor (Tm14) reveals distortions in its dimeric four-helix bundle that provide<br />

insight into the conformational states available to MCP’s for propagating signals. A bulge in one<br />

helix generates asymmetry between subunits that displaces the kinase-interacting tip, which<br />

resides over 100 Å away. The maximum bundle distortion maps to the adaptational region <strong>of</strong><br />

transmembrane MCP’s where reversible methylation <strong>of</strong> acidic residues tunes receptor activity.<br />

Minor alterations in coiled-coil packing geometry translates the bulge distortion to a >25 Å<br />

movement <strong>of</strong> the tip. The Tm14 structure discloses how alterations in local helical structure,<br />

which could be induced by changes in methylation state and/or by conformational signals from<br />

membrane proximal regions, can reposition a remote domain that regulates the CheA kinase.<br />

64


<strong>BLAST</strong> X Poster #14<br />

FUNCTIONAL ANALYSIS OF A LARGE NON-CONSERVED REGION OF FlgK (HAP1) FROM<br />

RHODOBACTER SPHAEROIDES<br />

Castillo, D.J., Ballado, T., Camarena, L., Dreyfus, G.<br />

Departamento de Genética Molecular, Instituto de Fisiología Celular, Universidad Nacional<br />

Autónoma de México, Apdo. Postal 70-243 Cd. Universitaria, 04510, Mexico, D.F., Mexico.<br />

The single subpolar flagellum <strong>of</strong> Rhodobacter sphaeroides shows an enlarged hookfilamentjunction.<br />

One <strong>of</strong> the two proteins that compose this section <strong>of</strong> the filament is HAP1Rs<br />

(FlgKRs) it contains a central non-conserved region <strong>of</strong> 860 amino acids that makes this protein<br />

about three times larger than its homologue in Salmonella enterica serovar Typhimurium. We<br />

investigated the role <strong>of</strong> this central portion <strong>of</strong> the unusually large HAP1 protein <strong>of</strong> R.<br />

sphaeroides by monitoring the effects <strong>of</strong> serial deletions in flgKRs, the gene encoding HAP1Rs,<br />

on swimming and swarming. Two deletion mutants did not assemble functional flagella, two<br />

were paralyzed and five exhibited reduced free-swimming speeds. Some mutants produced<br />

unusual swarming patterns on s<strong>of</strong>t agar without or with Ficoll 400. A segment <strong>of</strong> approximately<br />

200-aa <strong>of</strong> the central region <strong>of</strong> HAP1Rs that aligns with the variable region <strong>of</strong> the flagellin<br />

sequence from other γ- and β-proteobacteria was also found. Therefore, it is possible that the<br />

origin <strong>of</strong> this large central domain <strong>of</strong> HAP1Rs could be associated with an event <strong>of</strong> horizontal<br />

transfer and subsequent duplications and/or insertions.<br />

65


<strong>BLAST</strong> X Poster #15<br />

THE FLAGELLAR MURAMIDASE FROM THE PHOTOSYNTHETIC BACTERIUM<br />

RHODOBACTER SPHAEROIDES<br />

Javier de la Mora 1 , Teresa Ballado 1 , Aldo E. García-Guerrero 1 , Laura Camarena 2 and Georges<br />

Dreyfus 1 Instituto de Fisiología Celular 1 , Instituto de Investigaciones Biomédicas 2 , Universidad<br />

Nacional Autónoma de México, 04510, México City, México.<br />

The biogenesis <strong>of</strong> flagellum from R. sphaeroides is a tightly regulated process and<br />

requires the expression <strong>of</strong> more than 50 genes in a strict hierarchical manner (1). The flagellar<br />

rod is composed by four proteins, FlgB, FlgC, FlgF and FlgG. Besides these structural<br />

components several more proteins are required for rod assembly among these FlgJ from<br />

Salmonella enterica, has been postulated to be a dual function protein: the N-terminal half could<br />

function as a scaffold or cap essential for rod assembly and the C-terminal half may function as<br />

a muramidase that degrades the peptidoglycan layer to facilitate rod penetration (2, 3).We have<br />

previously reported that the FlgJ protein <strong>of</strong> R. sphaeroides lacks the C-terminal muramidase<br />

domain and that mutations in this protein renders a Fla - phenotype (4).The absence <strong>of</strong> the<br />

muramidase domain in this protein, suggests that other polypeptide must accomplish this<br />

function. In this work we describe the characterization <strong>of</strong> an open reading frame (orf) RSP0072,<br />

which is located within the flgG operon in R. sphaeroides. The amino acid sequence analysis <strong>of</strong><br />

this gene product showed the presence <strong>of</strong> a soluble lytic transglycosylase domain. The deletion<br />

<strong>of</strong> the N-terminal region (112 aa) <strong>of</strong> the product <strong>of</strong> RSP0072 renders a non-motile phenotype as<br />

determined by swarm assays in s<strong>of</strong>t agar. Electron micrographs revealed the lack <strong>of</strong> flagellum in<br />

the mutant cells. The purified wild-type protein showed lytic activity on extracts <strong>of</strong> M.<br />

lysodeikticus. Interaction assays suggests that the protein encoded by RSP0072 interacts with<br />

the flagellar rod-scaffolding protein FlgJ. We propose that the product <strong>of</strong> RSP0072 is a flagellar<br />

muramidase that is exported to the periplasm via the Sec pathway where it interacts with FlgJ to<br />

open a gap in the peptidoglycan layer for the subsequent penetration <strong>of</strong> the nascent flagellar<br />

structure (5).<br />

References:<br />

1) Poggio S., et al., 2005, Mol. Microbiol., 58:969-83.<br />

2) Nambu T., et al., 1999, J. Bacteriol., 181:1555-1561.<br />

3) Hirano T., et al., 2001, J. Mol. Biol., 312:359-369.<br />

4) Gonzalez-Pedrajo B., et al., 2002, Biochim. Biophys. Acta, 1579:55-63.<br />

5) De la Mora J., et al., 2007, J. Bacteriol., 189:7998-04.<br />

66


<strong>BLAST</strong> X Poster #16<br />

MODELING SCAFFOLD PHOSPHORYLATION AS AN ADAPTATION MECHANISM IN<br />

BACTERIAL CHEMOTAXIS<br />

Roger Alexander, Adam Bildersee, and Thierry Emonet<br />

Department <strong>of</strong> Molecular, Cellular, and Developmental Biology, Kline Biology Tower 1054, Yale<br />

<strong>University</strong>, New Haven, CT 06520<br />

Bacterial chemotaxis is one <strong>of</strong> the best-understood signaling networks in biology.<br />

Chemotaxis in E. coli is a model system for thoroughly understanding a behavioral phenotype at<br />

the molecular level. Chemotaxis has been studied experimentally in a variety <strong>of</strong> other species,<br />

and modeled in B. subtilis, but no dynamical models <strong>of</strong> chemotaxis network architectures from<br />

other species have yet been built. A key feature <strong>of</strong> the chemotaxis signal transduction network<br />

is adaptation: after a transient response to a step change in input stimulus, its activity returns to<br />

its pre-stimulus steady state level, allowing the system to respond to higher stimulus<br />

concentrations without saturation. In E. coli, adaptation is mediated through two enzymes, CheR<br />

and CheB, that methylate and demethylate the sensory receptors. In this work we consider<br />

scaffold phosphorylation as an adaptation mechanism complementary to receptor methylation.<br />

The receptors form a sensory array in the membrane at the cell pole; the scaffold protein<br />

couples the receptors to the kinase CheA which they activate. In all current dynamical models <strong>of</strong><br />

chemotaxis, the receptors, scaffolds, and kinases are treated as a single species, a receptorkinase<br />

complex. This is understandable, because in E. coli, the scaffold CheW is a passive<br />

mediator <strong>of</strong> the receptor-kinase interaction. However, in other organisms, the scaffold CheV<br />

has not only a CheW scaffold domain, but also a receiver domain that is the target <strong>of</strong><br />

phosphorylation by the kinase. Bacillus subtilis has both CheW and CheV scaffolds, and<br />

phosphorylation <strong>of</strong> CheV is known to be necessary for adaptation. In B. subtilis, phospho-CheV<br />

decouples receptor from kinase, so it actively mediates their interaction. The evolutionary<br />

distance between B. subtilis and E. coli is wide, and there are other differences between their<br />

chemotaxis network architectures that confound an understanding <strong>of</strong> the specific role <strong>of</strong> scaffold<br />

phosphorylation. Therefore we choose to focus on a close relative <strong>of</strong> E. coli, Salmonella<br />

enterica serovar typhimurium. The chemotaxis network architecture in Salmonella is almost<br />

identical to that in E. coli, except that Salmonella has both CheW and CheV scaffolds. A mutant<br />

Salmonella that lacks methylation-based adaptation is able to adapt partially, presumably<br />

through its scaffold phosphorylation mechanism. We have built a dynamical model <strong>of</strong> the<br />

chemotaxis network in E. coli that explicitly represents receptor-kinase coupling by the scaffold.<br />

We have extended that model to include scaffold phosphorylation in Salmonella. Our model <strong>of</strong> a<br />

Salmonella mutant that lacks methylation exhibits partial adaptation, consistent with<br />

experimental results. This work is the first step in a long-range program to explore how<br />

evolutionary changes in chemotaxis network architecture affects the dynamics and function <strong>of</strong><br />

the network. Evolution <strong>of</strong> dynamic signal transduction networks is an important emerging<br />

research area in systems biology.<br />

67


<strong>BLAST</strong> X Poster #17<br />

TESTING THE YIN-YANG MODEL OF SIGNAL TRANSDUCTION IN A BACTERIAL<br />

CHEMORECEPTOR CYTOPLASMIC DOMAIN<br />

Ka Lin E. Swain and Joseph J. Falke<br />

Department <strong>of</strong> Chemistry and Biochemistry, <strong>University</strong> <strong>of</strong> Colorado, Boulder, Campus Box 215,<br />

Boulder, CO, 80309<br />

The bacterial transmembrane aspartate receptor (Tar) <strong>of</strong> E. coli and S. typhimurium<br />

chemotaxis is a homodimer that assembles to form larger oligomers, most likely a trimer-<strong>of</strong><br />

dimmers. The homodimer can be divided into three modules: (i) the transmembrane signaling<br />

module comprised <strong>of</strong> the periplasmic ligand binding domain and the transmembrane helices, (ii)<br />

the cytoplasmic HAMP domain which serves as a signal conversion module, and (iii) the<br />

cytoplasmic kinase control module possessing the adaptation sites and a protein interaction<br />

region that binds the CheA kinase. The kinase control module is a 4-helix bundle essential for<br />

transmitting the integrated signal output <strong>of</strong> the HAMP domain and the adaptation sites to the<br />

bound histidine kinase CheA. Considerable evidence indicates that structural rearrangments <strong>of</strong><br />

the subunit-subunit interface within the homodimeric 4-helix bundle are important in signaling.<br />

This study further probes the mechanism <strong>of</strong> signal transduction in the kinase control module <strong>of</strong><br />

the S. Typhimurium aspartate receptor. The approach mutates the conserved “knob” residues<br />

<strong>of</strong> “sockets” that stabilize helix-helix contacts to alanines, in order to destabilize the packing <strong>of</strong><br />

adjacent helices in the 4- helix bundle. Knob mutations that lock the receptor in “on” and “<strong>of</strong>f”<br />

signaling states are identified by their opposite effects on CheA kinase and receptor methylation<br />

rates. The results suggest a novel “Yin-Yang” mechanism in which the helix packing states <strong>of</strong><br />

the adaptation region (I) and the protein interaction region (II) have opposing effects on receptor<br />

on-<strong>of</strong>f switching: stable helix packing in region I and unstable packing in region II drive the<br />

receptor into the kinase-activating on-state, while unstable packing in region I and stable<br />

packing in region II favor the kinase-inactivating <strong>of</strong>f-state.<br />

68


<strong>BLAST</strong> X Poster #18<br />

CYTOCHROME d BUT NOT CYTOCHROME o RESCUES THE TOLUIDINE BLUE GROWTH<br />

SENSITIVITY OF arc MUTANTS IN E. COLI<br />

Adrián F. Alvarez, Roxana Malpica, Martha Contreras, Edgardo Escamilla, Everardo Ramírez<br />

and Dimitris Georgellis.<br />

Departamento de Genética Molecular, Instituto de Fisiología Celular, Universidad Nacional<br />

Autónoma de México, 04510 México D.F., México<br />

The Arc (anoxic redox control) two-component signal transduction system, consisting <strong>of</strong><br />

the ArcB sensor kinase and the ArcA response regulator, allows adaptive responses <strong>of</strong><br />

Escherichia coli to changes <strong>of</strong> O2 availability. Under anaerobic conditions, the ArcB sensor<br />

kinase autophosphorylates and then transphosphorylates ArcA, which in this form acts as a<br />

global transcriptional regulator that controls the expression <strong>of</strong> many operons involved in<br />

respiratory or fermentative metabolism. Under aerobic conditions, the system is inactivated<br />

through the inhibition <strong>of</strong> the kinase activity <strong>of</strong> ArcB by the quinone electron carriers, and the<br />

subsequent ArcA dephosphorylation. The arcA gene was previously known as the dye gene<br />

because null mutants were growth sensitive to the photosensitizer redox dyes toluidine blue and<br />

methylene blue, a phenotype whose molecular basis still remains elusive. Toluidine blue is a<br />

redox dye that, in presence <strong>of</strong> light, allows the production <strong>of</strong> reactive oxygen species (ROS). In<br />

this study we report that the toluidine blue O effect on the arc mutants is light independent, and<br />

only observed during aerobic growth conditions. Moreover, seventeen suppressor mutants with<br />

restored growth were generated and analyzed. Thirteen <strong>of</strong> those possessed insertion elements<br />

(IS) upstream the cydAB operon, rendering its expression ArcA independent. Finally, it was<br />

found that, in contrast to cythocrome d, cythocrome o was not able to confer toluidine blue<br />

resistance to arc mutants, thereby representing an intriguing difference between the two<br />

terminal oxidases.<br />

69


<strong>BLAST</strong> X Poster #19<br />

SEARCHING THE PHYSIOLOGICAL SIGNAL(S) THAT REGULATE THE ACTIVITY OF THE<br />

SENSOR KINASE BarA<br />

González-Chávez, R., Rodríguez-Rangel, C., Georgellis, D.<br />

Departamento de Genética Molecular, Instituto de Fisiología Celular, UNAM, México, D.F.<br />

04510.Tel. 56-22-57-38. dimitris@ifc.unam.mx<br />

The BarA/UvrY two-component system <strong>of</strong> Escherichia coli comprises the BarA protein<br />

as the histidine sensor kinase, and UvrY as the cognate response regulator. The E. coli<br />

BarA/UvrY two-component system (TCS) and its homologues in other gram-negative bacteria,<br />

such as BarA/SirA <strong>of</strong> Salmonella, ExpS/ExpA <strong>of</strong> Erwinia, VarS/VarA <strong>of</strong> Vibrio, and GacS/GacA<br />

<strong>of</strong> Pseudomonas species, have been shown to positively control expression <strong>of</strong> the noncoding<br />

CsrB and CsrC RNAs in E. coli. These small regulatory RNAs together with the 6.8 kD CsrA<br />

protein constitute the Csr (carbon storage regulation) system, a post-transcriptional regulatory<br />

system that has pr<strong>of</strong>ound effects on central carbon metabolism, motility and multicellular<br />

behavior <strong>of</strong> E. coli.<br />

No physiological signals able to regulate the BarA sensor kinase and thereby control the<br />

UvrY response regulator have so far been identified for any <strong>of</strong> the orthologous TCSs. However,<br />

in a recent study it was demonstrated that pH lower than 5.5 provides an environment that does<br />

not allow activation <strong>of</strong> the BarA/UvrY signaling pathway, providing an important physiological<br />

tool for further experimentation in this direction.<br />

Here, we present experiments aiming at identifying the environmental signal(s) to which<br />

BarA respond. Our results indicate that short fatty acids, such as formate and acetate, act as<br />

direct signals for BarA. The implications <strong>of</strong> our findings on the overall physiology <strong>of</strong> the cell are<br />

be discussed.<br />

70


<strong>BLAST</strong> X Poster #20<br />

THE ROLE OF QUORUM SENSING IN THE CONTROL OF MOTILITY AND PLANT INVASION<br />

BY SINORHIZOBIUM MELILOTI<br />

Nataliya Gurich* and Juan E. González<br />

Department <strong>of</strong> Molecular and Cell Biology, <strong>University</strong> <strong>of</strong> Texas at Dallas, Richardson, Texas,<br />

USA<br />

Quorum sensing, a population density dependent regulation <strong>of</strong> gene expression, is used<br />

by various bacteria to establish symbiotic or pathogenic bacterium-host associations. This<br />

mechanism requires the production <strong>of</strong> signaling molecules called autoinducers. At high cell<br />

population densities, autoinducer concentration in the surrounding area reaches a threshold<br />

level, which leads to activation <strong>of</strong> specific transcriptional regulators and the control <strong>of</strong> numerous<br />

phenotypes.<br />

Sinorhizobium meliloti is a gram-negative soil bacterium that can form a nitrogen-fixing<br />

symbiotic association with its host Medicago sativa. The quorum sensing system in S. meliloti is<br />

comprised <strong>of</strong> a transcriptional regulator, SinR and an autoinducer synthase, SinI, which<br />

specifies production <strong>of</strong> N-acyl homoserine lactone (AHL) signaling molecules. In conjunction<br />

with AHL, an additional regulator, ExpR, controls expression <strong>of</strong> several hundred genes. Recent<br />

microarray studies in our laboratory described the control <strong>of</strong> motility and chemotaxis in S.<br />

meliloti by quorum sensing through the regulators VisN/VisR and Rem in a population-densitydependent<br />

manner. The wild type strain is motile during the early log phase <strong>of</strong> growth but shuts<br />

down flagella synthesis genes during the mid- and late log phases. In contrast, the sinI mutant<br />

remains motile throughout all phases <strong>of</strong> growth, and at the late log phase, 35 motility and<br />

chemotaxis genes are upregulated when compared to wild type.<br />

This inability to repress flagella production by the sinI mutant leads to severe invasion<br />

deficiency. Detailed analysis <strong>of</strong> microarray results suggested that the inability <strong>of</strong> the sinI mutant<br />

to shut down flagella synthesis might be detrimental to successful plant invasion. Mutation <strong>of</strong><br />

flagella synthesis in the sinI strain restored invasion efficiency to wild type levels. Therefore,<br />

control <strong>of</strong> motility and chemotaxis by the ExpR/Sin quorum sensing system plays an important<br />

role in plant invasion and may provide a competitive edge for strains possessing it.<br />

71


<strong>BLAST</strong> X Poster #21<br />

CHARACTERIZATION OF EtgA, A MURAMIDASE ASSOCIATED WITH THE TYPE III<br />

SECRETION SYSTEM OF ENTEROPATHOGENIC ESCHERICHIA COLI<br />

Elizabeth García-Gómez, Norma Espinosa, Bertha González-Pedrajo<br />

Departamento de Genética Molecular, Instituto de Fisiología Celular, Universidad Nacional<br />

Autónoma de México. Ap. Postal 70-243. Ciudad Universitaria, México D. F., 04510.<br />

Tel. (5255) 56-22-59-65 egarcia@ifc.unam.mx, bpedrajo@ifc.unam.mx<br />

Enteropathogenic E. coli (EPEC) is a human pathogen that produces severe diarrhea<br />

and death among children from developing countries. EPEC infection is characterized by the<br />

formation <strong>of</strong> an attaching and effacing histopathology (A/E lesion) that consists <strong>of</strong> gut epithelial<br />

microvilli destruction, intimate adherence <strong>of</strong> the bacterium to the enterocyte and the<br />

development <strong>of</strong> an actin-rich pedestal-like structure beneath the adherent non-invasive bacteria.<br />

EPEC utilizes a type three secretion system (T3SS) to translocate effector proteins<br />

directly from the bacterial cytoplasm into the host cell cytosol, subverting diverse enterocytic cell<br />

signaling pathways and producing drastic cytoskeletal reorganization. EPEC genes required for<br />

the assembly <strong>of</strong> the T3SS and A/E lesion development are contained on a 35.6 kb pathogenicity<br />

island known as the locus <strong>of</strong> enterocyte effacement (LEE).<br />

The T3SS or injectisome is a macromolecular protein complex that needs to span the<br />

periplasmic space for its assembly; however, the peptidoglycan cell wall constitutes a barrier<br />

that allows the free passage <strong>of</strong> only small proteins. To overcome this obstacle, lytic<br />

transglycosylases (LTs) have been identified as specialized enzymes associated with different<br />

transport systems. It has been proposed that during T3SS biogenesis a LT facilitates a<br />

temporally and spatially controlled opening <strong>of</strong> the peptidoglycan layer.<br />

In this study, we have identified and characterized a LEE open reading frame with<br />

unknown function, rorf3 (renamed etgA), which encodes a protein with a lytic transglycosylase<br />

domain and a signal sequence. A truncated version <strong>of</strong> EtgA lacking its signal sequence<br />

(EtgAns) was purified by nickel chromatography as an N-terminal His-tagged recombinant<br />

protein. Lytic activity was determined by zymograms. An invariantly conserved glutamate<br />

residue at position 42 was replaced by alanine through site-directed mutagenesis, and its<br />

enzymatic activity was evaluated. The effect <strong>of</strong> over-producing EtgA, EtgAns and EtgAE42A<br />

over EPEC growth and secretion was evaluated. Furthermore, the subcellular localization <strong>of</strong> the<br />

protein was determined. Additionally, an etgA null mutant strain was generated to evaluate the<br />

role <strong>of</strong> EtgA in T3SS assembly.<br />

Our results show that EtgA is a T3SS associated muralytic enzyme. The highly<br />

conserved glutamate residue (catalytic residue) is essential for EtgA function. EPEC growth rate<br />

was affected when EtgA was overproduced, but not with EtgAns or EtgAE42A. Our data<br />

demonstrate that the protein is secreted by the Sec pathway and that it disrupts the<br />

peptidoglycan layer, causing bacterial lysis. Periplasmic localization <strong>of</strong> EtgAns was determined.<br />

Finally, we show that EtgA is essential for efficient protein secretion.<br />

72


<strong>BLAST</strong> X Poster #22<br />

flhE: A PERIPLASMIC CHAPERONE OF FLAGELLIN?<br />

Jae-Min Lee* and Rasika Harshey<br />

Section <strong>of</strong> Molecular Genetics and Microbiology, <strong>University</strong> <strong>of</strong> Texas at Austin, Texas 78712<br />

flhE is the last gene in the flhBAE flagellar operon whose first two members encode<br />

components <strong>of</strong> the Type III secretion pathway in Salmonella typhimurium. The role <strong>of</strong> flhE is still<br />

a mystery. Absence <strong>of</strong> flhE does not affect swimming motility, but reduces swarming motility. In<br />

this study, we show that FlhE is a periplasmic protein, and have localized it within the flagellar<br />

basal body. We have found a new ‘green’ phenotype associated with a flhE deletion mutant,<br />

which we are using as a tool to understand flhE function. The ‘green’ phenotype refers to green<br />

colony color when bacteria are plated on aniline blue-alizarin yellow pH indicator plates. The<br />

green color should be indicative <strong>of</strong> a lowered pH. However, independent pH measurements<br />

failed to confirm pH differences between wild-type and flhE strains. Curiously, mutations that<br />

prevented assembly but not secretion <strong>of</strong> flagellar filament subunits (flgK, fliD), eliminated the<br />

green color. Immunoblotting and immunostaining assays showed that there was more ‘free’<br />

flagellin in flhE mutant supernatants than in wild-type bacteria isolated from swarm plates.<br />

Based on these and other experiments, we hypothesize that FlhE is a periplasmic chaperone for<br />

the flagellar filament subunits. We propose that improperly folded filaments interact with dyes on<br />

the pH indicator plates giving rise to the ‘green’ phenotype, and are more easily broken when<br />

subjected to higher viscous forces on the flagellar filament during swarming motility.<br />

73


<strong>BLAST</strong> X Poster #23<br />

THE Cyclic-di-GMP RECEPTOR PROTEIN YcgR LOCALIZES TO THE FLAGELLAR BASAL<br />

BODY AND CHANGES MOTOR BIAS IN SALMONELLA<br />

Vincent Nieto* and Rasika Harshey<br />

Section <strong>of</strong> Molecular Genetics and Microbiology, <strong>University</strong> <strong>of</strong> Texas at Austin, Texas 78712<br />

Cyclic-di-GMP (cdG) is a bacterial second messenger that plays a central role in the<br />

transition between motile and sessile states. In Salmonella enterica, absence <strong>of</strong> the cdG<br />

phosphodiesterase YhjH is known to elevate cdG levels, inhibit motility, and promote bi<strong>of</strong>ilm<br />

formation. YcgR, a cdG -binding protein, is required for this phenomenon. We show in this study<br />

that the absence <strong>of</strong> YhjH inhibits chemotaxis, and that this inhibition is more pronounced when<br />

YcgR is overexpressed from a plasmid. In liquid media, the bacteria were predominantly smooth<br />

swimming. Tethering experiments showed a pronounced shift to CCW rotation with a significant<br />

slowing <strong>of</strong> motor rotation. Inhibition <strong>of</strong> chemotaxis was therefore at the level <strong>of</strong> either production,<br />

or activity <strong>of</strong> the chemotaxis response regulator CheY~P, which acts at the switch to change the<br />

default CCW bias to CW. Expression <strong>of</strong> a GFP-tagged cdG-binding protein YcgR in a yhjH<br />

mutant background, resulted in punctate fluorescent dots along the cell membrane. Presence <strong>of</strong><br />

the puncta was dependent on the presence <strong>of</strong> flagellar basal bodies, but not on presence <strong>of</strong><br />

chemotaxis signaling components, suggesting that YcgR inhibits chemotaxis by acting at the<br />

switch. Inhibition <strong>of</strong> chemotaxis may represent a novel strategy to prepare for a sedentary<br />

existence by disfavoring migration away from a substrate on which a bi<strong>of</strong>ilm is to be formed.<br />

74


<strong>BLAST</strong> X Poster #24<br />

ANALYSIS OF THE PEPTIDOGLYCAN-BINDING DOMAIN OF THE FLAGELLAR STATOR<br />

PROTEIN MotB USING SYSTEMATIC MUTAGENESIS AND CHIMERIC PROTEIN IN<br />

ESCHERICHIA COLI<br />

Yohei Hizukuri 1 , John Frederick Morton 1 , Toshiharu Yakushi 1, 2 , Seiji Kojima 1 and Michio<br />

Homma 1<br />

1<br />

Division <strong>of</strong> Biological Science, Graduate School <strong>of</strong> Science, Nagoya <strong>University</strong>, Furo-Cho,<br />

Chikusa-Ku, Nagoya 464-8602, Japan<br />

2<br />

Present address: Applied Molecular Bioscience, Graduate School <strong>of</strong> Medicine, Yamaguchi<br />

<strong>University</strong>, Yamaguchi 753-8515, Japan<br />

The bacterial flagellar stator proteins, MotA and MotB, form a complex and are thought<br />

to be anchored to the peptidoglycan (PG) layer by the C-terminal conserved peptidoglycanbinding<br />

(PGB) motif <strong>of</strong> MotB. To clarify the role <strong>of</strong> the C-terminal PGB region <strong>of</strong> MotB, we<br />

performed systematic cysteine mutagenesis <strong>of</strong> this region in the Escherichia coli MotB.<br />

Furthermore, we constructed three chimeric MotB proteins whose PGB regions were replaced<br />

with corresponding regions <strong>of</strong> other PGB proteins, E. coli Pal (peptidoglycan-associated<br />

lipoprotein), PomB (Vibrio MotB homolog) or MotY (Vibrio flagellar T-ring protein). Although<br />

these chimeric proteins did not complement the motB mutant, we were able to isolate two<br />

independent motile revertants from cells producing the MotB-Pal chimera. One is the F172V<br />

mutation in the Pal region <strong>of</strong> the chimera, and this mutation did not affect the ability to bind to<br />

PG when introduced into native Pal. The other is the P159L mutation in the MotB region, and<br />

Pro159 mutation in native MotB has been reported to affect a spatial positioning <strong>of</strong> the<br />

MotA/MotB stator complex relative to the motor switch complex. We speculated that the PGB<br />

region <strong>of</strong> MotB-Pal chimera was properly aligned by the mutations and the stator and rotor could<br />

interact properly. We tried to interpret phenotype <strong>of</strong> MotB Cys mutants by using the crystal<br />

structure <strong>of</strong> the E. coli Pal, and found that MotB mutants that affected motility were nearly<br />

overlapped with the predicted PG-binding residues <strong>of</strong> Pal. Our results indicate that the PGB<br />

regions from functionally distinct proteins, MotB and Pal, works similarly to anchor the stator<br />

complex.<br />

75


<strong>BLAST</strong> X Poster #25<br />

ATTEMPT TO PURIFY THE HOOK-BASAL BODY WITH C-RING FROM THE Na + -DRIVEN<br />

FLAGELLAR MOTOR<br />

Masafumi Koike, Seiji Kojima, Michio Homma<br />

Division <strong>of</strong> Biological Science, Graduate School <strong>of</strong> Science, Nagoya <strong>University</strong>, Chikusa-ku,<br />

Nagoya 464-8602, Japan.<br />

The flagellum is the biggest architecture in the bacterial organ. At the base <strong>of</strong> each<br />

flagellum, there is a rotary motor embedded in the membrane whose energy source is the<br />

electrochemical gradient <strong>of</strong> the specific ion across the membrane. E. coli and Salmonella have<br />

the H + driven motor, and Vibrio alginolyticus has Na + driven type. The motor is composed <strong>of</strong> the<br />

rotor and the stator, and torque is generated by the interactions between them. The hook-basal<br />

body (HBB) is the rotor part <strong>of</strong> the motor composed <strong>of</strong> rod, LP-, MS-ring and hook. C-ring is<br />

mounted on the cytoplasmic side <strong>of</strong> the MS-ring <strong>of</strong> HBB, and believed to be involved in torque<br />

generation in rotor side <strong>of</strong> the motor. It is composed <strong>of</strong> three proteins; FliM, FliN, and FliG. FliG<br />

is the most closely participated in the torque generation among them. To elucidate the<br />

mechanism <strong>of</strong> torque generation, we undertake the isolation <strong>of</strong> C-ring from the Na + -driven polar<br />

flagellum <strong>of</strong> V. alginolyticus to investigate rotor-stator interactions. We applied the C-ring<br />

isolation method established for Salmonella, but that resulted in the HBB without C-ring. It<br />

indicates that solubilization by Triton X-100 was harsh for C-ring <strong>of</strong> Vibrio. Suitable detergents<br />

were screened and we found that CHAPS could solubilize HBB with FliG attached. Currently<br />

immunoelectron microscopic observation is ongoing to directly detect FliG attached on the MSring.<br />

Also, we are searching more favorable conditions to isolate remaining C-ring proteins, FliM<br />

and FliN.<br />

76


<strong>BLAST</strong> X Poster #26<br />

MECHANISM FOR SORTASE LOCALIZATION AND ROLE IN EFFICIENT PILUS ASSEMBLY<br />

IN ENTEROCOCCUS FAECALIS<br />

Kimberly A. Kline*† Andrew L. Kau*, Swaine L. Chen*, Birgitta Henriques-Normark†, Staffan<br />

Normark†, Michael G. Caparon* and Scott J. Hultgren*<br />

*Department <strong>of</strong> Molecular Microbiology, Washington <strong>University</strong> School <strong>of</strong> Medicine, St. Louis,<br />

Missouri 63110; †Department <strong>of</strong> Microbiology,Tumor and Cell Biology, Karolinska Institutet,<br />

Solna, Sweden<br />

Pathogenic streptococci and enterococci primarily rely on the conserved secretory (Sec)<br />

pathway for the translocation and secretion <strong>of</strong> virulence factors out <strong>of</strong> the cell. Since many <strong>of</strong><br />

these secreted Gram positive virulence factors are subsequently attached to the bacterial cell<br />

surface via sortase enzymes, we sought to investigate the spatial relationship between<br />

secretion and cell wall attachment in Enterococcus faecalis. We discovered that Sortase A<br />

(SrtA) and Sortase C (SrtC) are co-localized with SecA at single foci in enterococcus. The SrtAprocessed<br />

substrate aggregation substance accumulated in single foci implying a single site <strong>of</strong><br />

secretion for these proteins. Furthermore, we show by electron and immun<strong>of</strong>luorescent<br />

microscopy and immunoblot analysis that in the absence <strong>of</strong> the pilus polymerizing SrtC, pilin<br />

subunits also accumulate in single foci. Proteins that localized to single foci in E. faecalis were<br />

found to share a positively charged domain flanking a transmembrane helix. Mutation or<br />

deletion <strong>of</strong> this domain in SrtC abolished both its retention at single foci and its function in<br />

efficient pilus assembly. We conclude that a positively charged domain adjacent to a<br />

transmembrane helix can act as a localization retention signal for the focal<br />

compartmentalization <strong>of</strong> membrane proteins. Finally, we have examined the localization <strong>of</strong> the<br />

secretion and sorting machinery and substrates throughout the bacterial cell cycle and present a<br />

model for spatial and temporal organization <strong>of</strong> the molecular components leading to efficient<br />

pilus biogenesis in E. faecalis.<br />

77


<strong>BLAST</strong> X Poster #27<br />

VISUALIZATION OF EXCHANGE OF ROTOR COMPONENT IN FUNCTIONING BACTERIAL<br />

FLAGELLAR MOTOR<br />

Hajime Fukuoka, Shun Terasawa, Yuichi Inoue, Akihiko Ishijima<br />

IMRAM, Tohoku Univ., 2-1-1 Katahira, Aoba-ku, Sendai, 980-5877, Japan<br />

Bacterial flagellum is a supramolecular complex penetrating the bacterial cell envelope,<br />

including the cytoplasmic and the outer membranes. Bacterial flagellum consists <strong>of</strong> a basal body<br />

(rotary motor), a helical filament (propeller), and a hook (universal joint). The flagellar motor is<br />

driven by the electrochemical potential <strong>of</strong> H +<br />

or Na + , and the interaction between stator and a<br />

rotor <strong>of</strong> flagellar motor is thought to generate the motor rotation. Stator parts are exchanged in a<br />

functioning motor and the assembly <strong>of</strong> stator to the motor requires coupling ion for the motor<br />

rotation. In this study, we focused on protein dynamics <strong>of</strong> rotor in functioning bacterial flagellar<br />

motors. We constructed GFP-fusions <strong>of</strong> rotor components, and investigated whether rotor<br />

components are exchanged in a functioning motor by Fluorescent Recovery After<br />

Photobleaching (FRAP) experiment for a single motor.<br />

We constructed the expression systems <strong>of</strong> GFP-FliG, FliM-GFP, and GFP-FliN as GFPfusions<br />

<strong>of</strong> rotor components. In tethered cells producing each GFP-fusions, we observed the<br />

localization <strong>of</strong> fluorescent spot at the rotational center. Each GFP fusion was probably<br />

incorporated into flagellar motor as a rotor component. To observe the exchange <strong>of</strong> rotor<br />

components, we carried out FRAP experiment using evanescent illumination to the motor<br />

located at rotational center <strong>of</strong> the tethered cell. When fluorescent spot <strong>of</strong> FliM-GFP or GFP-FliN<br />

localized at rotational center was photobleached, the fluorescence at the rotational center<br />

recovered with the passage <strong>of</strong> time. On the other hand, the recovery <strong>of</strong> fluorescence was not<br />

observed in cells producing GFP-FliG. These results indicate that some rotor components, FliN<br />

and FliM at least, assemble to the motor even after the functional motor is constructed.<br />

Probably, in functioning motor, some components <strong>of</strong> flagellar structure are exchanged<br />

dynamically. We would like to clarify turnover rates <strong>of</strong> rotor components until the annual<br />

meeting.<br />

78


<strong>BLAST</strong> X Poster #28<br />

TORQUE RESPONSE OF THE SODIUM-DRIVEN CHIMERIC FLAGELLAR MOTOR IN<br />

E.COLI INDUCED BY REVERSIBLE TEMPERATURE CHANGE<br />

Yuichi Inoue (1), Kuniaki Takeda (2), Hajime Fukuoka (1), Hiroto Takahashi (1) and Akihiko<br />

Ishijima (1).<br />

1: IMRAM, Tohoku <strong>University</strong>; 2: Graduate School <strong>of</strong> Life Sciences, Tohoku <strong>University</strong>, Katahira<br />

2-1-1, Aoba-ku, Sendai 980-8577, Japan<br />

Mechanical step <strong>of</strong> motor proteins is a key to understand the molecular mechanism <strong>of</strong><br />

energy transduction from chemical energy into mechanical work. We have demonstrated the 26<br />

steps in a rotation <strong>of</strong> Na + -driven chimeric flagellar motor (Sowa et al., 2005) and the torquespeed<br />

relationship (Inoue et al., 2008) using back-focal-plane interferometry. However, present<br />

time resolution is not enough to understand molecular mechanism <strong>of</strong> the steps <strong>of</strong> motor proteins<br />

including not only flagellar motors but also linear motors as skeletal myosin. One possible option<br />

would be to slow down the temperature-dependent processes in a chemo-mechanical cycle by<br />

reducing temperature.<br />

We report the cooling experiment with a water-cooling chip in a simple and reversible<br />

way. A small chip <strong>of</strong> ~18mm*18mm*2mm was made with glass cover slides and plastic tubes to<br />

lead cooling water. This chip was made direct contact on a sample chamber and placed in the<br />

light axis <strong>of</strong> the microscope. By changing temperature <strong>of</strong> the cooling flow, sample temperature<br />

from 23 to -8 degree Celsius was monitored using thermocouple. Temperature change could be<br />

applied repeatedly in a time constant < 60 sec.<br />

Simultaneous measurement <strong>of</strong> the motor speed and sample temperature showed the<br />

speed change as reported for low temperature (Chen and Berg, 2000). With increasing<br />

temperature with hot water, however, sudden drop <strong>of</strong> speed was measured over ~40 deg C.<br />

When the temperature returned back to room temperature, the speed was restored mostly in<br />

several minutes. The drop and recovery <strong>of</strong> the speed were coincided with stepwise change in<br />

the generated torque.<br />

These results suggest that our cooling system is useful not only for the cooling<br />

experiment for improving time resolution but also for the heating experiment to understand heat<br />

resistance which might be related to stator dynamics.<br />

79


<strong>BLAST</strong> X Poster #29<br />

THE HELICOBACTER PYLORI FLAGELLAR ANTI-SIGMA FACTOR FlgM REMAINS<br />

BACTERIA-ASSOCIATED AND INTERACTS WITH FlhAC<br />

Melanie Rust 1 , Sophie Borchert 1 , Eugenia Gripp 1 , Sarah Kühne 1 , Eike Niehus 1 , Sebastian<br />

Suerbaum 1 , Kelly T. Hughes 2 , Christine Josenhans 1<br />

1 Department for Med. Microbiology and Hospital Epidemiology, Hannover Medical School ,Carl-<br />

Neuberg-Strasse 1, D-30625 Hannover, Germany, 2 Department <strong>of</strong> Biology, <strong>University</strong> <strong>of</strong> <strong>Utah</strong>,<br />

USA<br />

Helicobacter pylori persistently colonize in the human gastric mucus with the help <strong>of</strong> their<br />

polar flagellar organelles. A particular feature <strong>of</strong> flagella in most Helicobacter species including<br />

H. pylori is the presence <strong>of</strong> a membraneous flagellar sheath. Previously, flagellar regulators<br />

FliA, RpoN and the functional master regulator FlhA were characterized. H. pylori also<br />

possesses an anti-sigma factor FlgM, which has an unusually short N-terminus, yet is functional<br />

in Salmonella. FlgM in H. pylori fulfills a similar conserved function as an antagonist and binding<br />

partner to FliA. However, FlgM <strong>of</strong> H. pylori is unusual, since it lacks an N-terminal domain<br />

present in other FlgM homologs, e.g. FlgM <strong>of</strong> Salmonella, whose function is intimately coupled<br />

to its secretion through the flagellar type III secretion system.<br />

The aim <strong>of</strong> the present study was to characterize in more detail the localization and<br />

potential for secretion <strong>of</strong> the short H. pylori FlgM in the presence <strong>of</strong> a flagellar sheath.<br />

Furthermore we investigated its interaction with other flagellar proteins in the basal body, which<br />

may be required for its function in flagellar regulation. FlgM was expressed in flhA mutants and<br />

was differentially localized in bacterial fractions <strong>of</strong> flhA mutant bacteria in comparison to wild<br />

type bacteria. H. pylori FlgM was only released into the medium in very minor amounts in wild<br />

type bacteria, where the bulk amount <strong>of</strong> the protein was retained in the cytoplasm, and some<br />

FlgM was detected in the flagellar fraction. FlgM-GFP (green fluorescent protein) and FlgM-V5<br />

translational fusions were generated and expressed in H. pylori. FlgM displayed a<br />

predominantly polar distribution in microscopy. Evidence was gathered that it is able to interact<br />

with the C-terminal domain <strong>of</strong> the flagellar basal body protein FlhA. We conclude that in H. pylori<br />

and possibly in other closely related bacteria, which also possess a truncated FlgM, secretion<br />

may not be paramount for the regulatory function <strong>of</strong> FlgM and that protein interactions at the<br />

flagellar basal body, in particular with FlhAC, may determine the turnover and localization <strong>of</strong><br />

functional FlgM in H. pylori.<br />

We gratefully acknowledge the German Research Council, grant Jo344/2-2, for financial<br />

support.<br />

80


<strong>BLAST</strong> X Poster #30<br />

LOCALIZATION PATTERNS OF THE HISTIDINE KINASES IN AN ESCHERICHIA COLI CELL<br />

Takehiko Inaba 1 , Satomi Banno 2 , Hiroyuki Sawaki 2 , Akiko Yamakawa 2 , Masayuki Yoshimoto 3<br />

Michio Homma 3 and Ikuro Kawagishi 1,4<br />

1: Res. Cen. Micro-Nano Tech., Hosei Univ.; 2: Natl. Inst. Infect. Dis.; 3: Div. <strong>of</strong> Biol. Science,<br />

Nagoya Univ., 4: Dept. Frontier Biosci., Hosei Univ.<br />

E.coli has 30 histidine kinases (HKs). In response to a specific environmental stimulus,<br />

each HK controls the activity <strong>of</strong> the downstream response regulator (RR) by modulating its<br />

activities <strong>of</strong> autophosphorylation on the specific His residue and transfer <strong>of</strong> the phosphoryl<br />

group to the specific Asp residue <strong>of</strong> RR (Generally referred to as the "two-component"<br />

regulatory system or the His-Asp phosphorelay system). The activated form <strong>of</strong> RR regulates<br />

specific cell functions (in most cases, gene expression). Almost all HKs resemble MCPs<br />

(methyl-accepting chemotaxis proteins) in that they have two transmembrane regions and form<br />

homodimers. CheA and NtrB are cytosolic proteins. CheA forms a chemosensory complex with<br />

the trasmembrane chemoreceptors (MCPs) and the adaptor CheW. This complex localizes to<br />

the cell pole and forms a huge cluster, which plays a critical role in signal amplification.<br />

Do the other HKs also localize to particular regions <strong>of</strong> the membrane? To answer this<br />

question, we constructed HK-GFP fusion proteins and observed the localization via<br />

fluorescence microscopy. As mentioned above, cytosolic CheA-GFP localized to the cell pole,<br />

whereas the other cytosolic HK NtrB-GFP was diffused throughout the cytoplasm. Although<br />

many <strong>of</strong> the transmembrane HK-GFPs were diffused evenly in the cytoplasmic membrane,<br />

some HK-GFPs showed characteristic localization patterns. In particular, the GFP fusions to<br />

the anaerobic sensors TorS and ArcB localized to the cell pole. Most <strong>of</strong> HKs that localized to<br />

the pole are <strong>of</strong> hybrid type, i.e. they have receiver and HPt domains, which are involved in multistep<br />

phophorelay. A series <strong>of</strong> deletion from TorS-GFP revealed that none <strong>of</strong> these domains nor<br />

the transmitter domain was required for polar localization. Nevertheless, polar localization <strong>of</strong><br />

such hybrid HKs might provide a basis <strong>of</strong> signal integration and crosstalk.<br />

81


<strong>BLAST</strong> X Poster #31<br />

THERMOSENSING FUNCTION OF Aer, A REDOX SENSOR OF E. COLI<br />

So-ichiro Nishiyama 1 , Shinji Ohno 2 , Noriko Ohta 3 , Akihiko Ishijima 4 1, 5<br />

and Ikuro Kawagishi<br />

1<br />

Department <strong>of</strong> Frontier Bioscience, Faculty <strong>of</strong> Bioscience and Applied Chemistry, Hosei<br />

<strong>University</strong><br />

2<br />

Department <strong>of</strong> Material Chemistry, Faculty <strong>of</strong> Engineering, Hosei <strong>University</strong><br />

3<br />

Division <strong>of</strong> Biological Science, Graduate School <strong>of</strong> Science, Nagoya <strong>University</strong><br />

4<br />

Institute <strong>of</strong> Multidisciplinary, Research for Advanced Materials, Tohoku <strong>University</strong><br />

5<br />

Department <strong>of</strong> Frontier Bioscience, Faculty <strong>of</strong> Engineering, Hosei <strong>University</strong><br />

Some motile bacteria can sense temperature and move to temperatures best-suited to<br />

growth. This behavior, called thermotaxis, has been extensively studied in Escherichia coli.<br />

Our early studies revealed that E. coli thermotaxis is mediated by chemoreceptors that also<br />

sense amino acids or sugars: Tsr (serine), Tar (aspartate and maltose) and Trg (ribose and<br />

galactose) function as warm sensors, producing counter-clockwise or clockwise flagellar rotation<br />

signals upon temperature upshift and downshift, respectively. Tap (dipeptides, pyrimidines)<br />

functions as a cold sensor, producing CW and CCW signals upon temperature increases and<br />

decreases, respectively. Unique among these temperature sensors, Tar switches from a warm<br />

sensor to a cold sensor after adaptation to its ligand, aspartate. Intensive studies <strong>of</strong> Tar<br />

revealed that the receptor’s transmembrane and methylation domains play important roles in<br />

thermotactic responses, but what part <strong>of</strong> the chemoreceptor molecule actually senses<br />

temperature, remains unknown.<br />

In this study, we found that the aerotaxis transducer Aer also has temperature-sensing<br />

ability. An otherwise receptor-less strain expressing only aer showed extremely smooth<br />

swimming and did not show any thermoresponse. However, after imposing a CW rotational<br />

bias by adding the general repellent, glycerol (up to 10% w/v) or by co-expressing a cytoplasmic<br />

fragment <strong>of</strong> Tar that does not mediate a thermoresponse, the cells showed CCW and CW<br />

responses to temperature upshifts and downshifts, respectively. These results suggest that Aer<br />

functions as a warm sensor, even though, unlike the Tar, Tap, Tsr, and Trg chemoreceptors,<br />

Aer does not have a periplasmic ligand-binding domain. Thus, temperature sensing by E. coli<br />

chemoreceptors may be a general attribute <strong>of</strong> their highly-conserved cytoplasmic signaling<br />

domain (or their less conserved transmembrane domain).<br />

82


<strong>BLAST</strong> X Poster #32<br />

ATPase ACTIVITY OF T3SS SPECIFIC ATPase InvC<br />

Fumio Hayashi, Eri Inobe, Kenji Oosawa<br />

Department <strong>of</strong> Chemistry and Chemical Biology, Graduate school <strong>of</strong> Engineering, Gunma<br />

<strong>University</strong>, 1-5-1 Tenjin, Kiryu, Gunma, 376-8515, Japan<br />

The type III secretion systems (T3SSs) are widely used by Gram-nagative bacteria, and<br />

there are essential systems <strong>of</strong> many bacterial pathogenic to humans, animals, and plants. The<br />

systems are anchored to the bacterial envelope by a multi-ring, and a needle-like extracellular<br />

structure facilitates the translocation <strong>of</strong> the virulence proteins (effectors) to a host cell from the<br />

bacterial cytosol. An ATPase that is believed to be in close association with the basal body is<br />

involved in the protein translocation.<br />

The specific ATPase <strong>of</strong> T3SSs in Salmonella enterica serovar Typhimurium is InvC. We<br />

purified InvC, determined the low kcat value <strong>of</strong> InvC-ATPase activity and the high Km value for<br />

ATP, and found 2~3-fold stimulation <strong>of</strong> InvC-ATPase activity in the presence <strong>of</strong> phospholipid.<br />

To examine the possibility that InvC-ATPase activity is further stimulated by the interaction with<br />

an effector or an effecter-specific chaperon, we purified SicP (chaperon) and SptP (effector) and<br />

measured InvC-ATPase activity in the presence <strong>of</strong> SicP or SptP. SicP- or SptP-dependent<br />

stimulation <strong>of</strong> InvC-ATPase activity was not detected yet.<br />

83


<strong>BLAST</strong> X Poster #33<br />

CHARACTERIZATIONS OF THE PSEUDOREVERTANTS FROM SALMONELLA<br />

TYPHIMURIUM STRAIN SJW1655 AND SJW1660 WITH THE R- AND THE L-TYPE<br />

STRAIGHT FLAGELLAR FILAMENTS<br />

Hidetoshi Tomaru, Fumio Hayashi, Eiji Furukawa, Shigeru Yamaguchi & Kenji Oosawa<br />

Department <strong>of</strong> Chemistry and Chemical Biology, Graduate School <strong>of</strong> Engineering, Gunma<br />

<strong>University</strong>, 1-5-1 Tenjin, Kiryu, Gunma 376-8515, JAPAN<br />

A cell <strong>of</strong> Salmonella typhimurium swims by rotating its flagellar filaments. A wild-type<br />

cell carries a left-handed helical filament that is called a normal filament. On the other hand,<br />

various mutants carrying different helical shapes were isolated. There are curly, coiled, semicoiled<br />

and two kinds, L- and R-types, <strong>of</strong> straight filaments known. Transformations <strong>of</strong> filament<br />

helical shape among normal, semi-coiled and curly were observed during tumbling <strong>of</strong> the cell.<br />

To investigate the transformation mechanism <strong>of</strong> flagellar filaments, we isolated 95 revertants,<br />

which recovered their swarming abilities, from Salmonella typhimurium strain SJW1655 with the<br />

R-type straight flagellar filament. The numbers <strong>of</strong> the pseudorevertants carrying another<br />

mutation site were 64 and the second mutation sites were 4. Similarly, we isolated 106<br />

revertants (including 101 pseudorevertants) from SJW1660 (the L-type straight flagellar<br />

filament). The numbers <strong>of</strong> the second mutation sites were 19. We also measured the swimming<br />

speeds <strong>of</strong> the pseudorevertants and observed the flagellar shapes isolated from the<br />

pseudorevertants. We will discuss, in this meeting, the differences in the distribution <strong>of</strong> the<br />

second mutation sites, the swimming speeds, and the flagellar shapes <strong>of</strong> the pseudorevertants<br />

from between SJW1655 and SJW1660.<br />

84


<strong>BLAST</strong> X Poster #34<br />

RAMAN OPTICAL ACTIVITY AND VIBRATIONAL CIRCULAR DICHROISM OF FLAGELLAR<br />

FILAMENTS OF SALMONELLA<br />

Tomonori Uchiyama 1 , Fumio Hayashi 1 , Masashi Sonoyama 2 , Yoshiaki Hamada 3 , Rina K.<br />

Dukor 4 , Laurence A. Nafie 4, 5 , Kenji Oosawa 1 (kenji@nms.gunma-u.ac.jp)<br />

1<br />

Department <strong>of</strong> Nano-Material Systems, Gunma <strong>University</strong>, Kiryu, Japan<br />

2<br />

Department <strong>of</strong> Applied Physics, Nagoya <strong>University</strong>, Nagoya, Japan<br />

3<br />

The Open <strong>University</strong> <strong>of</strong> Japan, Chiba, Japan<br />

4<br />

BioTools Inc., Jupiter, FL 33458, USA<br />

5<br />

Department <strong>of</strong> Chemistry, Syracuse <strong>University</strong>, Syracuse, NY 13244-4100, USA<br />

The flagellar filament <strong>of</strong> Salmonella is an assembly <strong>of</strong> a single protein, flagellin. There<br />

are different helical and straight shape filaments from wild type and mutants. The difference<br />

between two types <strong>of</strong> straight filaments, L- and R-type straight, is the inclination <strong>of</strong> the flagellin<br />

monomer arrangements. The structures <strong>of</strong> the straight filaments were analyzed by X-ray<br />

diffraction and electron microscope, while it is difficult to analyze the structure <strong>of</strong> helical flagellar<br />

filaments by these methods, due to deficiency in the symmetric property <strong>of</strong> the molecular<br />

assembly. On the other hand, Raman optical activity (ROA) and vibrational circular dichroism<br />

(VCD) spectroscopies are available new techniques for studying structure and dynamics <strong>of</strong><br />

chiral molecules and the solution structure <strong>of</strong> biomacromolecules. In the present study, these<br />

methods are employed for investigating the structural characteristics <strong>of</strong> bacterial flagellar<br />

filaments.<br />

In the ROA spectra, intensive peaks were observed only from the L-type filaments, while<br />

no significant signals were observed from flagellin monomer and other shapes <strong>of</strong> the filaments<br />

(R-type straight and normal helical) in our measurement conditions. The intensive signals<br />

disappeared or were weakened when the L-type straight filaments were shortened.<br />

In the VCD spectra, intensive peaks in the amide I region were observed from the L-type<br />

filaments. Whereas peaks observed from the R-type filaments were different from those from Ltype<br />

and only weak peaks were observed in the normal filaments and flagellin monomer.<br />

From these results, it is thought that these spectra reflect the differences <strong>of</strong> structure and the<br />

physicochemical properties <strong>of</strong> flagellin subunits in three types <strong>of</strong> filaments.<br />

85


<strong>BLAST</strong> X Poster #35<br />

Poster Cancelled<br />

86


<strong>BLAST</strong> X Poster #36<br />

CrdC NEGATIVELY REGULATES CheW3 AND CheA3 INTERACTION DURING SIGNAL<br />

TRANSDUCTION IN MYXOCOCCUS XANTHUS<br />

Jonathan Willett, Susanne Mueller, John Kirby<br />

<strong>University</strong> <strong>of</strong> Iowa, Department <strong>of</strong> Microbiology, 3-403 Bowen Science Building<br />

51 Newton Road, Iowa City, IA 52242<br />

Previous work on the Che3 system <strong>of</strong> Myxococcus xanthus has led to the discovery that<br />

a chemosensory signal-transduction system affects developmentally regulated gene expression.<br />

The Che3 system contains homologs <strong>of</strong> the prototypical chemotaxis proteins such as CheA,<br />

CheW, CheB, CheR, and MCPs but lacks a CheY homolog. The output <strong>of</strong> the system consists<br />

<strong>of</strong> a NtrC transcriptional activator termed CrdA. There are several other unique proteins<br />

comprising the Che3 pathway besides CrdA, with one <strong>of</strong> the more interesting being a protein<br />

termed CrdC. CrdC is contained in the same transcriptional unit as CheW3. Preliminary data<br />

has shown that CrdC interacts strongly CheW3 in yeast-two hybrid experiments. More<br />

interestingly, the presence <strong>of</strong> CrdC in a yeast-three hybrid experiment has been shown to inhibit<br />

the interaction between CheA3 and CheW3. We hypothesize that CrdC thereby presents a<br />

unique mechanism for regulation <strong>of</strong> signal transduction through the Che3 pathway by<br />

decoupling CheA3 and CheW3.<br />

87


<strong>BLAST</strong> X Poster #37<br />

MOLECULAR ARCHITECTURE OF INTACT FLAGELLAR MOTOR REVEALED BY CRYO-<br />

ELECTRON TOMOGRAPHY<br />

Jun Liu 1 , Tao Lin 1 , Douglas J. Botkin 1 , Erin McCrum 1 , Hanspeter Winkler 2 , Steven J. Norris 1<br />

1<br />

Department <strong>of</strong> Pathology and Laboratory Medicine, <strong>University</strong> <strong>of</strong> Texas Medical School at<br />

Houston, Houston, TX 77225-0708, USA.<br />

2<br />

Institute <strong>of</strong> Molecular Biophysics, Florida State <strong>University</strong>, Tallahassee, FL, 32306-4380, USA<br />

Motility is <strong>of</strong>ten important for virulence <strong>of</strong> bacterial pathogens, and the flagellum is the<br />

main organelle for motility in bacteria. Bacterial flagella are helical propellers turned by the<br />

flagellar motor, a remarkable nano-machine embedded in the bacterial cell envelope. Powered<br />

by the proton gradient across the cytoplasmic membrane, the motor converts electrochemical<br />

energy into torque through an interaction between a rotating, cylindrical basal body at the end <strong>of</strong><br />

the flagellar filament and the stator, a surrounding protein assembly embedded in the<br />

cytoplasmic membrane. Of the 50 genes needed to build a functional flagellum, at least 25<br />

produce proteins essential for flagellar assembly. Although structural studies have revealed the<br />

stunning complexity <strong>of</strong> the basal body, flagellar assembly and rotation remain poorly understood<br />

at the molecular level, mainly because <strong>of</strong> the lack <strong>of</strong> structural information about the membranebound<br />

stators and the torque-generating mechanism in particular. Here, we present the<br />

structures <strong>of</strong> infectious wild-type and mutant Borrelia burgd<strong>of</strong>eri organisms and their flagella<br />

motors in situ using high throughput Cryo-Electron Tomography (Cryo-ET). By averaging the 3-<br />

D images <strong>of</strong> ~1280 flagellar motors, we obtained a ~3 nm resolution model <strong>of</strong> the combined<br />

stator and rotor structure in its cellular environment. We have also been able to identify<br />

distinctive structural changes resulting from the mutation <strong>of</strong> a flagellar gene. This is direct<br />

mapping <strong>of</strong> a single genetic code change into the 3-D structure <strong>of</strong> a functioning molecular<br />

machine in situ. Our results provide new insights into the motor structure and the molecular<br />

basis for bacterial motility.<br />

88


<strong>BLAST</strong> X Poster #38<br />

MINING THE E. COLI GFP FUSION COLLECTION<br />

Jason Dobkowski, Aleksandra Sikora, John Brooks, Richard Zielke and Janine Maddock<br />

Department <strong>of</strong> MCDB, <strong>University</strong> <strong>of</strong> Michigan, 830 N. <strong>University</strong>, Ann Arbor, MI 48109<br />

Use <strong>of</strong> green fluorescent protein (GFP) has greatly increased the ability to visualize<br />

protein localization within bacterial cells. A library <strong>of</strong> GFP fusions, expressed from an inducible<br />

promoter, was created with each protein from every open reading frame in the Escherichia coli<br />

genome (Kitagawa et al, 2005). The addition <strong>of</strong> the 26.9 kDa GFP, however, can lead to<br />

protein misfolding and aggregation and the formation <strong>of</strong> polar inclusion bodies. These polar<br />

inclusion bodies are <strong>of</strong>ten mistaken for bona fide polar localization <strong>of</strong> the protein fusion. Using a<br />

purification and screening approach, we have sorted polarly localized GFP fusions into those<br />

that form inclusion bodies and those that are likely real polar proteins. Those that form inclusion<br />

bodies are being used to identify interacting partners that co-aggregate with the misfolded<br />

protein. Proteins not previously not known to be polar are being verified using<br />

immun<strong>of</strong>luorescence microscopy and the localization <strong>of</strong> known interacting partners determined.<br />

We are particularly interested in the spatial organization <strong>of</strong> membrane bound histidine kinases<br />

involved in sensing environmental stress. Our current line <strong>of</strong> focus is to determine whether the<br />

activation <strong>of</strong> the signal transduction cascade has any influence <strong>of</strong> their polar localization.<br />

89


<strong>BLAST</strong> X Poster #39<br />

UNDERSTANDING THE FUNDAMENTAL ELEMENTS OF SIGNALING IN THE Tar<br />

CHEMORECEPTOR<br />

Christopher Adase, Michael Manson<br />

Texas A&M <strong>University</strong>, 3258 TAMU, BSBE Room 303<br />

The Tar chemoreceptor <strong>of</strong> Escherichia coli has two different attractants: L-aspartate,<br />

which binds directly to the receptor; and maltose, which interacts with the receptor indirectly<br />

though maltose-binding protein (MBP). Tar also senses Ni 2+ and Co 2+ as repellents. The Tsr<br />

chemoreceptor interacts directly with L-serine as an attractant and also senses repellents such<br />

as indole, acetate, and benzoate in an unknown manner. Tar-Tsr and Tsr-Tar chimeras have<br />

been created using the endogenous Nde1 site found at the end <strong>of</strong> the region encoding their<br />

respective HAMP domains. The Tar-Tsr hybrid could sense Ni 2+ as a repellent, whereas the Tsr-<br />

Tar hybrid could not. Thus, Ni 2+ probably interacts, directly or indirectly, with an area within the<br />

first 256 amino acids <strong>of</strong> Tar. Using knockouts <strong>of</strong> nikA, nikB, and nikC, we tested whether<br />

periplasmic NikA is necessary and sufficient for Ni 2+ taxis, as has been reported, and whether<br />

Ni 2+ must be taken up into the cytoplasm. I have also created additional Tar-Tsr chimeras to<br />

determine exactly what portion <strong>of</strong> Tar is involved in Ni 2+ sensing. Repellent taxis was assayed<br />

using novel micr<strong>of</strong>luidic techniques developed by the Jayaraman lab in the Department <strong>of</strong><br />

Chemical Engineering at Texas A&M.<br />

90


<strong>BLAST</strong> X Poster #40<br />

LINKING THE TM2 TO HAMP—A TOUGH NUT TO CRACK?<br />

Rachel L. Crowder, Gus A. Wright, and Michael D. Manson<br />

Department <strong>of</strong> Biology, Texas A&M <strong>University</strong>, College Station, TX 77843<br />

The HAMP domain is a widely conserved motif found in transmembrane-signaling<br />

proteins in prokaryotes and lower eukaryotes. It consists <strong>of</strong> a pair <strong>of</strong> amphipathic helices joined<br />

by a flexible linker. Recently, the solution structure <strong>of</strong> the Archeoglobus fulgidis Af1503 HAMP<br />

domain was determined using NMR (Hulko et al, Cell 126: 929-940, 2006). The domain forms a<br />

parallel four-helix bundle that packs in a non-canonical knob-on-knob conformation. Several<br />

models have been proposed to explain how the four helix bundle transmits the downward piston<br />

movement <strong>of</strong> transmembrane 2 (TM2) <strong>of</strong> E. coli chemoreceptors into the signaling domain to<br />

inhibit CheA kinase activity. The connection between TM2 and the HAMP domain is likely to be<br />

important for transducing the input signal from TM2 to HAMP. We hypothesized that increasing<br />

the flexibility <strong>of</strong> the connector should attenuate the output signal. To test this idea, residues<br />

between Met-215 Thr-218 <strong>of</strong> the E. coli aspartate receptor Tar were replaced with four Gly<br />

residues. Gly residues were then deleted (-4G through -1G) and added (+1G through +5G).<br />

Aspartate chemotaxis, rotational bias <strong>of</strong> tethered cells, mean reversal frequencies <strong>of</strong> tethered<br />

cells, and in vivo methylation levels were measured. These experiments suggest that increasing<br />

flexibility between TM2 and HAMP strongly biases the receptor to the “<strong>of</strong>f” (CCW-signaling)<br />

state and decreases aspartate mediated signal transmission between TM2 and HAMP.<br />

91


<strong>BLAST</strong> X Poster #41<br />

ELECTROSTATIC EFFECTS ON SIGNALING MUTATIONS IN THE C-TERMINAL REGION<br />

OF THE ESCHERICHIA COLI ASPARTATE CHEMORECEPTOR<br />

Andrew L. Seely, Run-Zhi Lai, and Michael D. Manson, Department <strong>of</strong> Biology, Texas A&M<br />

<strong>University</strong>, College Station, TX 77843<br />

The Escherichia coli aspartate chemoreceptor (Tar) responds to attractant by modulating<br />

the rotational bias <strong>of</strong> the flagellar motor. Previous studies measured the effects on Tar signaling<br />

when positive residues in its extreme C-terminal region were either neutralized or changed to<br />

negative residues. While this region is important as a linker to the NWETF pentapeptide where<br />

methylation and demethylation enzymes bind, it also affects transmembrane signaling. An<br />

R505A substitution decreased the receptor’s ability to respond to aspartate by 40%, and an<br />

R505E charge reversal completely abolished stimulation by aspartate. It was also noted that<br />

R505E may interact with D273 to destabilize the “on” signaling state (Lai, et al, Advanced<br />

Publication in Biochemistry, 2008). We hypothesize that secondary mutations within the<br />

background <strong>of</strong> the R505 mutations, and an additional R514A/E mutant, can potentially rescue or<br />

exacerbate the ability <strong>of</strong> mutated Tar to respond to aspartate. Specific mutations will include<br />

D273N/R, D263R, R228A/E, R505D and R509E. We will test our hypothesis in vivo through<br />

chemotaxis swarm assays, receptor methylation assays, and tethered cell assays. Our<br />

hypothesis will also be tested in vitro through CheA-Kinase assays.<br />

92


<strong>BLAST</strong> X Poster #42<br />

INTERACTION OF THE TRANSCRIPTIONAL REGULATORY COMPLEX, FlhDC, WITH ITS<br />

TARGET DNA<br />

Yi-Ying Lee, and Philip Matsumura<br />

Department <strong>of</strong> Microbiology and Immunology, College <strong>of</strong> Medicine, <strong>University</strong> <strong>of</strong> Illinois at<br />

Chicago, 835 S. Wolcott Ave., M/ C 790, Chicago, Illinois 60612-7344<br />

The bacterial flagellum is the structure that allows bacteria to move and respond to<br />

nutritional and chemical signals in their environment. It is a complex suborganelle and the<br />

transcriptional regulation <strong>of</strong> the 40 plus structural genes is organized in a highly regulated<br />

cascade. At the top <strong>of</strong> the hierarchy is the master operon which codes for FlhD and FlhC. These<br />

two positive transcriptional regulators form a unique heteroheximeric complex which binds<br />

upstream <strong>of</strong> the -35 region and requires sigma 70 for transcription. This complex has an<br />

unusually large ‘footprint’ <strong>of</strong> 48 base pair and bends the DNA 110 degrees. We have proposed<br />

that the DNA bind on the circumference <strong>of</strong> this toroid shaped FlhDC complex. Although we have<br />

determined the sequence 3 footprints on FlhDC regulated promoters, it is not possible to<br />

determine a consensus binding site in these 3 sequences. In this study, we have determined<br />

which bases are important for DNA binding and activity for FlhDC regulated promoter activity.<br />

First, we have divided the FlhDC footprint in the fliA promoter into five segments and found that<br />

two <strong>of</strong> the segments or 40% <strong>of</strong> the footprint were not required for binding. The remaining 30<br />

base pairs were divided into 3-5 base segments and randomly mutagenized and screened for<br />

the ability to bind and activate the fliA promoter. Analysis <strong>of</strong> these data suggests a consensus <strong>of</strong><br />

12 A, 15 A, 34 T, 36 A, 37 T, 44 A, 45 T in FlhD4C2 footprint fragment were important for activity. Five <strong>of</strong><br />

these bases demonstrated high specificity. Finally, this consensus was tested and found to be<br />

important in other FlhDC regulated promoter regions.<br />

93


<strong>BLAST</strong> X Poster #43<br />

APPLICATION OF BIOLAYER INTERFEROMETRY TO UNDERSTANDING INTERACTIONS<br />

AMONG SALMONELLA ENTERICA FLAGELLAR EXPORT APPARATUS PROTEINS<br />

Jonathan L. McMurry<br />

Department <strong>of</strong> Chemistry & Biochemistry, Kennesaw State <strong>University</strong>, Kennesaw, GA<br />

Biolayer interferometry (BLI) is an emerging optical biosensing technology for the<br />

analysis <strong>of</strong> dynamic biomolecular interactions. Measurements are made <strong>of</strong> changes in the<br />

interference pattern <strong>of</strong> white light reflected <strong>of</strong>f <strong>of</strong> two surfaces, one <strong>of</strong> which possesses a layer<br />

<strong>of</strong> immobilized protein (or other molecule) and the other an internal reference. Binding <strong>of</strong><br />

second protein to the immobilized one results in a change in distance between the two surfaces<br />

and thus a shift in wavelength <strong>of</strong> the interference pattern. Kinetic and affinity constants can thus<br />

be determined in real-time, without attainment <strong>of</strong> equilibrium, utilizing small, recoverable<br />

quantities <strong>of</strong> label-free biomolecules. Using a commercial BLI biosensor, interactions among<br />

proteins involved in flagellar export in Salmonella enterica were investigated. Among<br />

interactions measured were those between two proteins involved in specificity switching, FliK<br />

and FlhB. FliK binding to wild-type FlhB and two variants (N269A and P270A) surprisingly<br />

evidenced similar kinetic pr<strong>of</strong>iles with submicromolar KDs resulting from fast-on and fast-<strong>of</strong>f rate<br />

constants. Other interactions between export proteins, e.g., FliH-FliI, FliH-FliJ, demonstrated<br />

stable, low kd binding, as expected. Additional efforts include screening for interactions too<br />

weak to allow for copurification or other traditional binding assays, attempts to demonstrate<br />

interactions between export proteins and substrates and analysis <strong>of</strong> more-than-pairwise<br />

interactions. Further application <strong>of</strong> BLI to the investigation <strong>of</strong> flagellar export protein dynamics<br />

will allow for better understanding <strong>of</strong> the mechanism <strong>of</strong> export.<br />

94


<strong>BLAST</strong> X Poster #44<br />

A CROSS-SPECIES COMPARISION OF CHEMOTACTIC BEHAVIOR<br />

Julie Simons and Paul Milewski<br />

<strong>University</strong> <strong>of</strong> Wisconsin—Madison, Department <strong>of</strong> Mathematics, 480 Lincoln Dr., Madison, WI<br />

53706<br />

Understanding the population-level behavior <strong>of</strong> bacteria is <strong>of</strong> importance not only in<br />

exploring how simple organisms can perform complex behavior, but also to be able to optimize<br />

their potential for bioremediation and other uses. We are interested in modeling the<br />

chemotactic behavior <strong>of</strong> Rhodobacter sphaeroides and the better-understood Escherichia coli<br />

using a partial differential equation model known as the Keller-Segel model and experimental<br />

data, with the aim <strong>of</strong> being able to make a cross-species comparison. Swarm-plate experiments<br />

with uniform concentrations <strong>of</strong> the chemoattractant L-aspartate were performed for both bacteria<br />

over several concentrations <strong>of</strong> aspartate. Separately, growth experiments in liquid cultures were<br />

undertaken to quantify differences in aspartate concentration dependent growth between the<br />

two species. From the data we are able to determine parameter values for a Keller-Segel model<br />

and thus quantify differences between non-chemotactic diffusive behavior, growth effects, and<br />

chemotactic behavior. This quantitative modeling work comparing population-level behavior <strong>of</strong><br />

these bacteria allows one to deduce metabolic function parameters in agar, which are not<br />

possible to find experimentally and not incorporated in many previous models. We find that a<br />

significant proportion <strong>of</strong> the E. coli wild-type population appears to be non-chemotactic whereas<br />

the R. sphaeroides wild-type population appears to be primarily chemotactic, something not<br />

explored in other studies. Our parameters also indicate a joint saturation <strong>of</strong> growth and<br />

chemotaxis, which we postulate is a common evolutionary result. These findings provide a<br />

platform from which to explore incorporating cell-level knowledge into macro-scale behavioral<br />

models and the effects <strong>of</strong> heterogeneity <strong>of</strong> populations.<br />

95


<strong>BLAST</strong> X Poster #45<br />

TETHERED MYCOPLASMA<br />

Daisuke Nakane and Makoto Miyata<br />

Department <strong>of</strong> Biology, Graduate School <strong>of</strong> Science, Osaka City <strong>University</strong><br />

Mycoplasma mobile is a pathogenic flask-shaped bacterium 0.8 micron long. They bind<br />

to solid surfaces by “legs” sticking out from the base <strong>of</strong> membrane protrusion, “neck”, and glide<br />

by a unique mechanism. Recently, we proposed a working model, power stroke model, where<br />

the cells are propelled by many legs repeatedly catching and releasing the solid surface, driven<br />

by the energy <strong>of</strong> ATP hydrolysis. Here, to detect the movement <strong>of</strong> legs, we reduced the number<br />

<strong>of</strong> working legs and amplified the leg movement.<br />

When the cells were treated by 0.1% Tween 60, they were elongated from 0.8 micron to<br />

2.0-4.0 micron in 15 min with the extension <strong>of</strong> cytoskeletal structures. A previous study showed<br />

that the direct binding target <strong>of</strong> mycoplasma is a sialic acid, and the addition <strong>of</strong> free sialic acids<br />

dissociated gliding cells from the solid surface. Here, in the presence <strong>of</strong> 0.25-1 mM <strong>of</strong> sialic<br />

acid, the elongated cells pivoted widely, plausibly resulting from the reduction the leg number.<br />

The pivoting ceased when strong light was applied in the presence <strong>of</strong> a fluorescence dye,<br />

suggesting that the pivoting is caused by the gliding legs. Azide and the antibody against the<br />

gliding machinery affected the distribution <strong>of</strong> pivoting angle differently, although both reduce the<br />

gliding speed. On the basis <strong>of</strong> these results, the movements <strong>of</strong> legs will be discussed.<br />

96


<strong>BLAST</strong> X Poster #46<br />

MOLECULAR SHAPES OF Gli123 AND Gli521 INVOLVED IN GLIDING MOTILITY OF<br />

MYCOPLASMA MOBILE<br />

Takahiro Nonaka, Jun Adan-Kubo, Makoto Miyata<br />

Department <strong>of</strong> Biology, Graduate School <strong>of</strong> Science, Osaka City <strong>University</strong><br />

3-3-138 Sugimoto Sumiyoshi-ku, Osaka-shi, JAPAN<br />

Mycoplasma mobile has no flagella or pili, and its genome contains no genes related to<br />

known bacterial motility. However, M. mobile binds to solid surfaces and glides smoothly and<br />

continuously, with a unique mechanism. M. mobile cells form a membrane protrusion at the<br />

leading pole. Three huge proteins, Gli123 (123 kDa), Gli349 (349 kDa), and Gli521 (521 kDa),<br />

localize at neck, the base <strong>of</strong> protrusion, and form the gliding machinery. These proteins are<br />

suggested to have the roles <strong>of</strong> scaffold for other gliding proteins, glass binding, and force<br />

transmission, respectively. In this study, we purified Gli123 and Gli521 proteins from M. mobile<br />

cells by biochemical procedures, rotary shadowed, and observed their molecular shapes by<br />

transmission electron microscopy. The Gli123 molecule shaped asymmetrical oval, 31 nm long<br />

and 16 nm wide. The Gli521 molecule consisted <strong>of</strong> three flexible arms about 130 nm long, each<br />

<strong>of</strong> them had spherical part at the distal end. This shape is reminiscent <strong>of</strong> clathrin-triskelion<br />

which is widely distributed in eukaryotic cells. Clathrin molecules self-assemble, control the<br />

membrane shape and form a vesicle. The characteristic molecular shape <strong>of</strong> Gli521 may<br />

suggest that this protein forms a two dimensional sheet like clathrin, and plays a critical role to<br />

form the membrane protrusion in a M. mobile cell.<br />

97


<strong>BLAST</strong> X Poster #47<br />

THE ESSENTIAL NATURE OF THE WALK/WALR SIGNAL TRANSDUCTION PATHWAY IS<br />

LINKED TO CELL WALL HYDROLASE ACTIVITY IN STAPHYLOCOCCUS AUREUS<br />

Aurélia Delauné 1 , Olivier Poupel 1 , Adeline Mallet 2 , Sarah Dubrac 1 , and Tarek Msadek 1<br />

Biology <strong>of</strong> Gram-positive Pathogens 1 , Department <strong>of</strong> Microbiology, Plateforme de Microscopie<br />

Ultrastructurale 2 , Imagopole, Institut Pasteur, 25 rue du Dr. Roux, 75015 Paris, France<br />

Staphylococcus aureus, a major Gram-positive human pathogen, is a leading cause <strong>of</strong><br />

both nosocomial and community infections due to its considerable capacity for adaptation. One<br />

<strong>of</strong> the principal mechanisms involved in this process are the so-called two-component systems,<br />

bacterial signal transduction pathways with a sensor histidine kinase that is autophosphorylated<br />

in response to specific environmental stimuli and then transfers the phosphoryl group to its<br />

cognate response regulator, which consequently regulates target gene expression. The WalKR<br />

two-component system is well conserved and specific to low G+C Gram-positive bacteria,<br />

including Bacillus subtilis, Staphylococcus aureus and Streptococcus pneumoniae. This system<br />

has been shown to be essential for cell viability.<br />

We have recently demonstrated that the WalKR system positively controls autolytic<br />

activity, and identified ten genes belonging to the WalKR regulon that appear to be involved in<br />

S. aureus cell wall degradation. Reasoning that the essential nature <strong>of</strong> this signaling pathway<br />

may be related to its role as a master regulatory system for cell wall metabolism, we tested<br />

whether uncoupling autolysin gene expression from WalKR-dependent regulation could<br />

compensate for the essential nature <strong>of</strong> the system. Several genes from the WalKR regulon were<br />

expressed from a cadmium chloride-inducible promoter in a WalKR-independent manner.<br />

Candidate genes were identified whose expression allowed cells to grow in the absence <strong>of</strong><br />

WalKR.<br />

98


<strong>BLAST</strong> X Poster #48<br />

THE GraS/GraR TWO-COMPONENT SYSTEM AND DERMASEPTIN RESISTANCE IN<br />

STAPHYLOCOCCUS AUREUS<br />

Mélanie Falord 1 , Pierre Joanne 2 , Chahrazade El Amri 2 , and Tarek Msadek 1<br />

Biology <strong>of</strong> Gram-positive Pathogens 1 , Department <strong>of</strong> Microbiology, Institut Pasteur, 25 rue du<br />

Dr. Roux, 75015 Paris, France<br />

Peptidome de la peau des amphibiens 2 , FRE 2852, CNRS, Université Pierre et Marie Curie, 2<br />

place Jussieu, 75251 Paris 75005, France<br />

Staphylococcus aureus, a major Gram-positive human pathogen, is a leading cause <strong>of</strong><br />

both nosocomial and community infections due to its considerable capacity for adaptation.<br />

S. aureus is able to resist Cationic Anti-Microbial Peptides (CAMPs) by increasing its positive<br />

cell surface charges through D-alanylation <strong>of</strong> wall teichoic acids and lysylination <strong>of</strong><br />

phospholipids, leading to electrostatic repulsion <strong>of</strong> CAMPs. Synthesis <strong>of</strong> the major enzymes<br />

involved in these mechanisms (DltA, MprF) is positively controlled by the GraS/GraR twocomponent<br />

system. Two-component systems are bacterial signal transduction pathways with a<br />

sensor histidine kinase that is autophosphorylated in response to specific environmental stimuli<br />

and then transfers the phosphoryl group to its cognate response regulator which consequently<br />

regulates target gene expression. In S. aureus, GraS is involved in CAMP sensing, promoting<br />

bacterial resistance through GraR.<br />

Here, we have shown that a ∆graRS mutant in S. aureus is sensitive to Colistin and<br />

Dermaseptins. Moreover, we demonstrated that the graRS genes are part <strong>of</strong> a three-gene<br />

operon also containing graX, a gene with unknown function but essential to the system, and<br />

defined the operon transcriptional start site. To define the mechanism by which GraS, GraR and<br />

GraX confer CAMP resistance to S. aureus, the three proteins were overexpressed and purified<br />

to test their in vitro interactions. A potential GraR binding site upstream from the vraFG operon,<br />

known to be controlled by GraS/GraR, was identified both by in silico analysis and lacZ<br />

transcriptional fusion experiments.<br />

99


<strong>BLAST</strong> X Poster #49<br />

CHARACTERIZATION OF SUPPRESSORS OF THE MotB(D33E) MUTATION, A PUTATIVE<br />

PROTON-BINDING RESIDUE OF THE BACTERIAL FLAGELLAR MOTOR<br />

Yong-Suk Che, Shuichi Nakamura, Yusuke Morimoto, Nobunori Kami-ike, Keiichi Namba and<br />

Tohru Minamino<br />

Graduate School <strong>of</strong> Frontier Biosciences, Osaka <strong>University</strong>, 1-3 Yamadaoka, Suita, Osaka<br />

565-0871, Japan<br />

MotA and MotB form the stator <strong>of</strong> the proton-driven bacterial flagellar motor, which<br />

conducts protons and couples proton flow to motor rotation. Asp-33 <strong>of</strong> Salmonella Typhimurium<br />

MotB, which is a putative proton-binding site, is critical for torque generation. However, how<br />

does the protonation <strong>of</strong> Asp could drive the conformational changes requiring for torque<br />

generation is largely unknown.<br />

Here, we carried out genetic and motility analysis <strong>of</strong> a slow motile motB(D33E) mutant<br />

and its pseudorevertants. First, we confirmed that MotB(D33E) forms the complex with MotA in<br />

the cytoplasmic membrane. Then, we isolated suppressor mutants from the motB(D33E) mutant<br />

and identified the suppressor mutation sites. Next, we characterized the torque-speed<br />

relationship <strong>of</strong> the flagellar motors <strong>of</strong> wild-type, motB(D33E) mutant and its suppressors by the<br />

bead assays. As a result, we found that while the wild-type motor torque was almost constant<br />

over a wide range <strong>of</strong> rotation rate, the MotB(D33E) mutation caused ≈40% reduction in nearstall<br />

torque and a sharp decline in the torque-speed curve with an apparent maximal rotation<br />

rate <strong>of</strong> ≈20 Hz. Furthermore, we also found that the second-site mutations could recover the<br />

near-stall torque but not the sharp decline <strong>of</strong> torque-speed curve and the maximum rotation<br />

rate. These results together suggested that MotB(D33E) mutation reduced both protonconducting<br />

activity and torque generation step involving the stator-rotor interactions coupled<br />

with protonation/deprotonation <strong>of</strong> Glu-33 and the second-site mutations could recover the torque<br />

generation step but not the proton-conducting activity.<br />

Recently, to measure the proton-conducting activity <strong>of</strong> the motB(D33E) mutant and its<br />

pseudorevertants, we developed a novel system to monitor intracellular pH <strong>of</strong> cells<br />

overexpressing MotA/MotB mutant proteins utilizing pH-sensitive GFP (pHluorin). Experimental<br />

results will also be discussed.<br />

100


<strong>BLAST</strong> X Poster #50<br />

STRUCTURE OF FliJ, A CYTOPLASMIC COMPONENT OF THE FLAGELLAR TYPE III<br />

PROTEIN EXPORT APPARATUS OF SALMONELLA<br />

Tatsuya Ibuki 1, 2 , Masafumi Shimada 1, 2 , Tohru Minamino 1, 2 , Katsumi Imada 1, 2 and Keiichi<br />

1, 2<br />

Namba<br />

Grad. Sch. <strong>of</strong> Frontier Biosci., Osaka Univ. 1 , Dynamic NanoMachine Project, ICORP, JST 2<br />

The flagellum is a motile organelle composed <strong>of</strong> the basal body rings and the tubular<br />

axial structure. The axial component proteins synthesized in the cytoplasm are transferred into<br />

the central channel <strong>of</strong> the flagellum by the flagellar type III protein-export apparatus for selfassembly<br />

at the growing end. The apparatus is composed <strong>of</strong> six transmembrane proteins (FlhA,<br />

FlhB, FliO, FliP, FliQ, FliR) and three soluble components (FliH, FliI, FliJ). FliJ is an essential<br />

component for protein export. Although FliJ is thought to be a general chaperone, its function is<br />

still unclear. Here we report a crystal structure <strong>of</strong> FliJ. We obtained hexagonal bi-pyramid<br />

crystals <strong>of</strong> FliJ with N-terminal extra three residues, and determined the structure at 2.1-Å<br />

resolution using anomalous diffraction data from a mercury derivative collected at SPring-8<br />

BL41XU.<br />

FliJ consists <strong>of</strong> two α-helices, one is a 13-turn helix and the other a 21-turn helix, which<br />

form a coiled-coil structure. FliJ has a remarkable structural similarity to the γ subunit <strong>of</strong> F0F1-<br />

ATPsynthase. The other soluble components FliH and FliI are also known to have similarity to<br />

other components <strong>of</strong> F0F1-ATPsynthase. The structure <strong>of</strong> FliI closely resembles those <strong>of</strong> the α/β<br />

subunits, and the amino-acid sequence <strong>of</strong> FliH has two regions that are similar to those <strong>of</strong> the b<br />

and δ subunits, respectively.<br />

We will discuss details <strong>of</strong> the structure and possible functional mechanism <strong>of</strong> FliJ, and<br />

similarity between the flagellar export apparatus and F0F1-ATPsynthase.<br />

101


<strong>BLAST</strong> X Poster #51<br />

CRYOEM STRUCTURE OF THE HOOK-FILAMENT JUNCTION OF SALMONELLA<br />

Fumiaki Makino (1), Takayuki Kato (1), Keiichi Namba (1)<br />

1: Grad. Sch. <strong>of</strong> Frontier Biosci., Osaka Univ.<br />

The bacterial flagellum is a biological nanomachine for the locomotion <strong>of</strong> bacteria. The<br />

flagellum consists <strong>of</strong> three functional parts: the basal body as a rotary motor, the filaments as a<br />

helical propeller, and the hook as a universal joint that connects the above-mentioned two parts.<br />

The filament is a tubular structure made <strong>of</strong> a single protein, FliC. It transforms into various<br />

helical forms in response to mechanical perturbation by reversal <strong>of</strong> motor rotation. The hook is<br />

also a tubular structure made <strong>of</strong> a single protein, FlgE, but is more flexible than the filament,<br />

allowing it to transmit motor torque to the filament regardless <strong>of</strong> its orientation. The hookfilament<br />

junction made <strong>of</strong> two proteins, FlgK and FlgL, connects these two mechanically distinct<br />

structures. It works as a mechanical adapter, allowing the two parts to go through dynamic<br />

conformational changes independently. To understand the adapter mechanism, we have<br />

analyzed the structure <strong>of</strong> the hook-filament junction by electron cryomicroscopy (cryoEM). We<br />

needed a Salmonella strain that produces short filaments for efficient data collection. We used a<br />

strain, MMC1660, which produces short filament due to a significantly low efficiency <strong>of</strong> FliC<br />

expression controlled by the addition <strong>of</strong> tetracycline. We established a procedure to purify the<br />

hook-basal body with short filament, collected cryoEM images and carried out image analysis.<br />

We will show the cryoEM structure <strong>of</strong> the hook-filament junction complex.<br />

102


<strong>BLAST</strong> X Poster #52<br />

EFFECT OF INTRACELLULAR pH ON THE TORQUE-SPEED RELATIONSHIP OF<br />

BACTERIAL PROTON-DRIVEN FLAGELLAR MOTOR<br />

Shuichi Nakamura 1,2 , Nobunori Kami-ike 2 , Jun-ichi Yokota 1,2 , Seishi Kudo 3 ,Tohru Minamino 1,2 ,<br />

and Keiichi Namba 1,2<br />

1 Graduate School <strong>of</strong> Frontier Bioscience, Osaka <strong>University</strong> 1-3 Yamadaoka, Suita, Osaka 565-<br />

0871, Japan, 2 Dynamic NanoMachine Project, ICORP, JST, 1-3 Yamadaoka, Suita, Osaka 565-<br />

0871, Japan, 3 Faculty <strong>of</strong> Engineering, Toin <strong>University</strong> <strong>of</strong> Yokohama, 1614 Kurogane-cho, Aoba,<br />

Yokohama 225-8502, Japan<br />

The flagellar motor <strong>of</strong> Salmonella is a rotary nanomachine driven by proton motive force.<br />

It has been shown that an increase in the intracellular proton concentration abolishes the motor<br />

function, suggesting that the stator complex has intracellular proton-binding sites by which<br />

intracellular protons kinetically interfere with the rate <strong>of</strong> proton translocation for the motor<br />

rotation. In this study, to understand the coupling mechanism <strong>of</strong> proton flux with torque<br />

generation, we have investigated the effect <strong>of</strong> intracellular pH on the rotation rate <strong>of</strong> the flagellar<br />

motor. Intracellular pH was manipulated by adding potassium benzoate, which crosses the<br />

cytoplasmic membrane in the neutral form and equilibrates the intra- and extracellular pH<br />

without changing proton motive force. We showed that the rotation rates at low loads sharply<br />

decreased as the external pH decreased in the presence <strong>of</strong> benzoate, suggesting that a high<br />

intracellular proton concentration lowered the proton translocation rate. Also, computer<br />

simulation by a simple kinetic model suggested that the decrease in intracellular pH interferes<br />

with proton dissociation from an intracellular proton-binding site <strong>of</strong> the stator. We conclude that<br />

the intracellular pH is a critical factor that determines the flagellar rotation rate and that proton<br />

dissociation from the motor is a rate-limiting step in the mechanochemical reaction cycle.<br />

103


<strong>BLAST</strong> X Poster #53<br />

FLUORESCENCE IMAGING OF ASSEMBLY AND DISASSEMBLY OF THE BACTERIAL<br />

FLAGELLAR PROTEIN EXPORT ATPase FliI TO THE FLAGELLAR BASAL BODY<br />

Shinsuke Yoshimura, Tohru Minamino, Keiichi Namba<br />

Grad. Schl. <strong>of</strong> Frontier Biosci., Osaka Univ.<br />

For construction <strong>of</strong> the bacterial flagellum, most <strong>of</strong> the flagellar proteins are translocated<br />

into the central channel <strong>of</strong> the growing structure by the flagellar protein export apparatus. FliI<br />

ATPase forms a complex with its regulator FliH and facilitates the initial entry <strong>of</strong> export<br />

substrates into the export gate made <strong>of</strong> six membrane proteins. The FliH/FliI complex also binds<br />

to a C ring protein, FliN, through the FliH-FliN interaction for efficient export. However, it<br />

remains unclear how these reactions proceed within the cell.<br />

In this study, we constructed FliI-CFP and FliI-YFP fusion proteins and analyzed their<br />

localization by fluorescence microscopy. A few bright spots within each cell suggested that<br />

many <strong>of</strong> them are bound to the C ring, and breaching experiments showed their rapid assembly<br />

and disassembly. We confirmed that both FliH and the C ring are required for the localization,<br />

but faint spots observed even in the absence <strong>of</strong> the C ring suggested binding <strong>of</strong> the FliI hexamer<br />

to the gate. FliI-YFP formed a complex with FliH∆1 missing residues 2-10 but the complex did<br />

not show the localization. FliH∆1 interacted with neither FliN nor the gate-forming proteins.<br />

Alanine-scanning mutagenesis <strong>of</strong> FliH revealed that only two residues, Trp-7 and Trp-10, are<br />

responsible for these interactions. Taken all together, hydrophobic interactions <strong>of</strong> two Trp<br />

residues <strong>of</strong> FliH with other export components seem to drive the cycling reaction <strong>of</strong><br />

assembly/disassembly <strong>of</strong> FliI for efficient export.<br />

104


<strong>BLAST</strong> X Poster #54<br />

ANALYSIS OF HELICOBACTER PYLORI LACKING ALL FOUR CHEMORECEPTORS<br />

Karen M. Ottemann<br />

UC Santa Cruz<br />

Helicobacter pylori is an epsilon proteobacter that uses chemotaxis and motility to infect<br />

human stomachs and cause ulcer disease. Based on genome sequencing and annotation, H.<br />

pylori is predicted to have four chemoreceptors. Here we describe the construction and<br />

characterization <strong>of</strong> mutants lacking one, two, three, or all four chemoreceptors in strain mG27,<br />

and a comparison <strong>of</strong> chemoreceptor expression across many H. pylori strains. We find that one<br />

chemoreceptor is sufficient to confer s<strong>of</strong>t-agar migration in our standard Brucella Broth/Fetal<br />

Bovine Serum s<strong>of</strong>t agar. This finding suggests that ligands for each chemoreceptor are found in<br />

this milieu. We also find that mutants lacking some chemoreceptor have an altered stopping<br />

frequency, as noted by Schweinitzer et al (J Bac 2008 190:3244). Mutants lacking all four<br />

chemoreceptors are non-chemotactic and completely smooth swimming, supporting that the<br />

four MA-domain proteins are the only chemoreceptors in this system. Our work thus shows<br />

experimentally that H. pylori possesses four chemoreceptors, and sets the stage for dissecting<br />

what each chemoreceptor senses.<br />

105


<strong>BLAST</strong> X Poster #55<br />

CHEMOTAXIS TO PYRIMIDINES AND IDENTIFICATION OF A CYTOSINE<br />

CHEMORECEPTOR IN PSEUDOMONAS PUTIDA<br />

X. Liu, P.L. Wood, J.V. Parales, and R.E. Parales<br />

Department <strong>of</strong> Microbiology, <strong>University</strong> <strong>of</strong> California, Davis, CA, 95616<br />

Pseudomonads are motile bacteria that are widespread in nature and are known for their<br />

catabolic versatility. These organisms have a conserved chemotaxis system that is homologous<br />

to that present in Escherichia coli. Unlike E. coli, however, which has only one set <strong>of</strong> chemotaxis<br />

(che) genes in a single gene cluster, Pseudomonas species have multiple che gene<br />

homologues organized in several unlinked gene clusters. In addition, genome sequence<br />

analyses have revealed that Pseudomonas genomes encode numerous MCP-like proteins,<br />

suggesting that these organisms can also sense and respond to a wide range <strong>of</strong> chemicals and<br />

environmental conditions. For example, Pseudomonas aeruginosa PAO1 has 26 MCP-like<br />

genes and Pseudomonas putida F1 has 27. We have been studying the chemotactic responses<br />

<strong>of</strong> P. putida strains to a variety <strong>of</strong> carbon and nitrogen sources and initiated a study in which we<br />

have individually deleted each <strong>of</strong> the 27 MCP-like genes in P. putida F1 and tested for mutant<br />

taxis phenotypes. We use both qualitative and quantitative chemotaxis assays to measure the<br />

responses, including a new quantitative capillary assay carried out in 96-well plates. In this<br />

study we demonstrated that P. putida strains F1 and PRS2000 are attracted to cytosine, but not<br />

thymine or uracil. The chemotactic response to cytosine was constitutively expressed under all<br />

tested growth conditions. In contrast, P. aeruginosa PAO1 was not chemotactic to any <strong>of</strong> these<br />

pyrimidines. Chemotaxis assays with a mutant strain <strong>of</strong> P. putida F1 in which the putative<br />

methyl-accepting chemotaxis protein-encoding gene Pput_0623 was deleted revealed that this<br />

gene (designated mcpC) encodes a chemoreceptor for positive chemotaxis to cytosine.<br />

Complementation <strong>of</strong> the F1ΔmcpC mutant with the wild-type gene restored chemotaxis to<br />

cytosine. In addition, introduction <strong>of</strong> this gene into P. aeruginosa PAO1 conferred the ability to<br />

respond to cytosine. To our knowledge, this is the first report <strong>of</strong> a chemoreceptor for cytosine.<br />

106


<strong>BLAST</strong> X Poster #56<br />

REACTIVE ALDEHYDES AND MOTILITY IN ESCHERICHIA COLI K12<br />

Changhan Lee, Junghoon Lee, Jongchul Shin, Insook Kim, Kwanghee Baek, and Chankyu<br />

Park<br />

Department <strong>of</strong> Biological Sciences, KAIST, Daejon, Korea<br />

Graduate School <strong>of</strong> Biotechnology, Kyung Hee <strong>University</strong>, Yongin, Korea<br />

The short chain carbohydrates such as glyoxal and methylglyoxal are generated in vivo<br />

from various sugars either by an oxidative stress or by a cellular metabolism, which are believed<br />

to be removed by the glutathione-dependent glyoxalase system. We isolated a number <strong>of</strong> E. coli<br />

mutants conferring resistance or sensitivity to these aldehydes and found that some <strong>of</strong> them<br />

affect motility. In the case <strong>of</strong> glyoxalse I mutant (gloA) that is deficient in the removal <strong>of</strong><br />

aldehyde using glutathione, significant reductions in the free-swimming as well as in swarming<br />

behaviors were observed. When we introduced the flhDC-lacZ fusion contained in phage<br />

Lambda into the wild-type and gloA strains, expression <strong>of</strong> LacZ was considerably reduced in the<br />

gloA mutant compared to that <strong>of</strong> wild type, suggesting that a decrease in flagellar expression is<br />

responsible for the impaired motility. This is further confirmed by the reduced expression <strong>of</strong><br />

flagellin as detected by anti-flagellin antiserum. We observed similar effects <strong>of</strong> other glyoxalrelated<br />

genes on motility, suggesting a possibility that an intracellular level <strong>of</strong> aldehyde is<br />

somehow associated with flagellar gene expression. Cellular redox change due to an aldehyde,<br />

affecting motility, will be discussed.<br />

107


<strong>BLAST</strong> X Poster #57<br />

TAKING CONTROL OF THE BACTERIAL FLAGELLAR MOTOR<br />

Mathieu Gauthier, Dany Truchon, Alexandre Bastien, Simon Rainville<br />

Laval <strong>University</strong>, Physics department and COPL, Pavillon Optique Photonique, 2139, 2375, rue<br />

de la Terrasse, Quebec, Quebec, G1V 0A6, Canada<br />

The bacterial flagellar motor is a fairly complex machine, requiring 40-50 genes for its<br />

expression, assembly and control. Furthermore, it is embedded in the multiple layers <strong>of</strong> the<br />

bacterial membrane. That explains why, unlike many other molecular motors, it has not yet been<br />

studied in vitro. As spectacular studies <strong>of</strong> linear motors (like kinesin, myosin and dynein) have<br />

clearly demonstrated, an in vitro system provides the essential control over experimental<br />

parameters to achieve the precise study <strong>of</strong> the motor’s physical and chemical characteristics.<br />

Here, we report significant progress towards the development <strong>of</strong> a unique in vitro system to<br />

study quantitatively the bacterial flagellar motor.<br />

Our system consists <strong>of</strong> a filamentous Escherichia coli bacterium partly introduced inside<br />

a micropipette. Femtosecond laser pulses (60 fs and ~ 15 nJ/pulse) are then tightly-focused on<br />

the part <strong>of</strong> the bacterium that is located inside the micropipette. This vaporizes a<br />

submicrometer-sized hole in the wall <strong>of</strong> the bacterium, thereby granting us access to the inside<br />

<strong>of</strong> the cell and the control over the proton-motive force that powers the motor. Using a patchclamp<br />

amplifier, we applied an external voltage between the inside and the outside <strong>of</strong> the<br />

micropipette. If the hole in the bacterium is open, that voltage should translate into a membrane<br />

potential powering the motors outside <strong>of</strong> the micropipette. As we varied the applied potential,<br />

variations in the motor's rotation speed were observed. For these preliminary results, the<br />

rotation speed was observed directly using video microscopy <strong>of</strong> fluorescently labeled filaments.<br />

That system opens numerous possibilities to study the flagellar motor and other membrane<br />

components.<br />

108


<strong>BLAST</strong> X Poster #58<br />

CHARACTERIZATION OF THE PERIPLASMIC DOMAIN OF THE SENSOR KINASE CpxA:<br />

ROLE OF CONSERVED RESIDUES IN CpxA ACTIVITY<br />

Malpica R and Raivio TL. Department <strong>of</strong> Biological Sciences, <strong>University</strong> <strong>of</strong> Alberta, CW405<br />

Biological Sciences Building, Edmonton, Alberta, Canada T6G 2E9.<br />

The Cpx two-component signal transduction system plays a major role in the<br />

conservation <strong>of</strong> cell envelope integrity, as well in the regulation <strong>of</strong> surface structure assembly,<br />

cellular attachment and pathogenesis in Escherichia coli. This system is comprised <strong>of</strong> the innermembrane<br />

sensor kinase CpxA and the cytosolic response regulator CpxR. Envelope stress<br />

caused by misfolded and mislocalized proteins activates the Cpx pathway, which also responds<br />

to alkaline pH and overexpression <strong>of</strong> the outer membrane lipoprotein NlpE. In the presence <strong>of</strong><br />

these activating cues, CpxA autophosphorylates at a conserved His residue and then<br />

transphosphorylates CpxR (CpxR-P). In turn, CpxR-P regulates the expression <strong>of</strong> several genes<br />

such as periplasmic protein folding and degrading factors involved in envelope protein<br />

manteinance under adverse conditions. It is known that in the absence <strong>of</strong> stress, the periplasmic<br />

protein CpxP, whose expression is positively controlled by this pathway, inhibits the activation <strong>of</strong><br />

the Cpx system. The periplasmic domain <strong>of</strong> CpxA (CpxA-pd), which contains 133 residues, has<br />

been proposed as the signal reception site and therefore as the regulatory element <strong>of</strong> the kinase<br />

and phosphatase activities <strong>of</strong> CpxA.<br />

To explore the role <strong>of</strong> CpxA-pd components in signal sensing and CpxA activity, we<br />

generated mutants that carry substitution mutations in highly conserved residues <strong>of</strong> this domain.<br />

The phenotypes <strong>of</strong> these mutants were evaluated, using alkaline pH or NlpE as inducing cues<br />

and the activatable cpxP´-lacZ fusion as a reporter <strong>of</strong> Cpx pathway activity. Remarkably, the<br />

substitution <strong>of</strong> Phe108, Gly130 and Pro164 by Ala lead to a constitutively active phenotype.<br />

Also, mutants in other residues displayed a significant increase on pathway activity in the<br />

absence <strong>of</strong> the inducing cues. Thus, we have identified relevant CpxA-pd elements for both<br />

signalling and the overall enzymatic activity <strong>of</strong> CpxA.<br />

109


<strong>BLAST</strong> X Poster #59<br />

CHARACTERIZATION OF FliZ AS AN ACTIVATOR OF FLAGELLAR GENES IN<br />

SALMONELLA ENTERICA SEROVAR TYPHIMURIUM<br />

Supreet Saini 1 , Phillip Aldridge 2,3 , and Christopher Rao 1<br />

Department <strong>of</strong> Chemical and Biomolecular Engineering, <strong>University</strong> <strong>of</strong> Illinois at Urbana-<br />

Champaign, Urbana, Illinois 61801, 1<br />

Centre for Bacterial Cell Biology, Newcastle <strong>University</strong>, Framlington Place, Newcastle upon<br />

Tyne NE2 4HH, United Kingdom, 2<br />

Institute for Cell and Molecular Biosciences, Newcastle <strong>University</strong>, Framlington Place,<br />

Newcastle upon Tyne NE2 4HH, United Kingdom 3<br />

Flagellar assembly proceeds in a sequential manner, beginning at the base and<br />

concluding with the filament. A critical aspect <strong>of</strong> assembly is that gene expression is coupled to<br />

assembly. When cells transition from a nonflagellated to a flagellated state, gene expression is<br />

sequential, reflecting the manner in which the flagellum is made. A key mechanism for<br />

establishing this temporal hierarchy is the σ 28 -FlgM checkpoint, which couples the expression <strong>of</strong><br />

late flagellar (Pclass3) genes to the completion <strong>of</strong> the hook-basal body. In this work, we<br />

investigated the role <strong>of</strong> FliZ in coupling middle flagellar (Pclass2) gene expression to assembly<br />

in Salmonella enterica serovar Typhimurium (Salmonella). We also demonstrate that significant<br />

cross talk exists between different secretion systems in Salmonella with FliZ as the cross talk<br />

element. Specifically, we show that FliZ is also an activator <strong>of</strong> Salmonella Pathogenicity Island-1<br />

(SPI1) encoded virulence genes and the fim, lpf, std, saf, bcf, stb, stc, stf, and sth loci encoded<br />

fimbriae in Salmonella. We demonstrate that FliZ is an FlhD4C2-dependent activator <strong>of</strong><br />

Pclass2/middle gene expression. Our results suggest that FliZ regulates the concentration <strong>of</strong><br />

FlhD4C2 posttranslationally which leads to faster induction <strong>of</strong> Pclass2/middle genes. We also<br />

demonstrate that FliZ functions independently <strong>of</strong> the flagellum-specific sigma factor σ 28 and the<br />

filament-cap chaperone/ FlhD4C2 inhibitor FliT. We correlate our gene expression experiments<br />

by discussing the role <strong>of</strong> FliZ in flagellar biosynthesis during swimming and swarming motility.<br />

FliZ was found to effect SPI1 activity levels in a HilD dependent posttranslational manner and its<br />

effect on type I fimbriae gene expression was found to be FimZ dependent.<br />

110


<strong>BLAST</strong> X Poster #60<br />

LOCALIZATION OF THE CHEMOTAXIS PROTEINS IN BACILLUS SUBTILIS<br />

Kang Wu and Christopher Rao<br />

<strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign<br />

The Bacillus subtilis chemotaxis pathway utilizes three proteins – CheC, CheD, and<br />

CheV - not found in Escherichia coli. These proteins are thought to be involved in two<br />

methylation-independent adaptation systems not found in E. coli. While the functions <strong>of</strong> these<br />

proteins have been characterized to some degree, the details concerning their interactions<br />

within the polar signaling complex are still unknown. To better understand these interactions, we<br />

are investigating the factors influencing the localization <strong>of</strong> the B. subtilis chemotaxis proteins<br />

using immun<strong>of</strong>luorescence and fluorescent protein fusions. Of significance, we have<br />

constructed functional fluorescent protein fusions to CheC and CheD and also explored their colocalization<br />

using two-color immun<strong>of</strong>luorescence. Using these approaches, we have quantified<br />

the localization <strong>of</strong> these proteins in a number <strong>of</strong> different mutants. This poster will discuss some<br />

<strong>of</strong> our preliminary results regarding the factors influencing localization <strong>of</strong> the chemotaxis<br />

proteins in B. subtilis.<br />

111


<strong>BLAST</strong> X Poster #61<br />

RcdA STRUCTURE AND FUNCTION IN REGULATED CtrA PROTEOLYSIS<br />

James A. Taylor*, Jeremy D. Wilbur † , Kathleen R. Ryan*<br />

* Department <strong>of</strong> Plant & Microbial Biology, <strong>University</strong> <strong>of</strong> California, Berkeley, Berkeley, CA<br />

94720-3102<br />

†<br />

Graduate Group in Biophysics, <strong>University</strong> <strong>of</strong> California, San Francisco, San Francisco, CA<br />

94158<br />

The Caulobacter crescentus master cell cycle regulator CtrA must undergo cyclic<br />

activation and deactivation to drive orderly progression through the division cycle. CtrA is<br />

essential for viability and directly activates or represses the transcription <strong>of</strong> ~100 genes.<br />

However, it also blocks the initiation <strong>of</strong> DNA replication, so CtrA activity must be eliminated from<br />

the cell before chromosome replication can occur. To this end, CtrA is rapidly degraded<br />

specifically at the G1-S cell cycle transition by the ClpXP protease. Regulated CtrA proteolysis<br />

in vivo requires two other factors, the single-domain response regulator CpdR and RcdA, a<br />

conserved protein <strong>of</strong> unknown function. Although each <strong>of</strong> these proteins is cytoplasmic, they all<br />

colocalize at one pole <strong>of</strong> the cell during CtrA degradation.<br />

We are investigating the role <strong>of</strong> RcdA in CtrA proteolysis. RcdA was proposed to act as<br />

an adaptor bridging CtrA and ClpXP to promote CtrA degradation. However, RcdA is not<br />

necessary for CtrA proteolysis by ClpXP in vitro and does not change the rate <strong>of</strong> CtrA<br />

degradation. We have therefore taken a structure-function approach to learn how RcdA<br />

contributes to CtrA proteolysis in vivo.<br />

Using X-ray crystallography, we have found that RcdA is a dimer, and each monomer<br />

consists <strong>of</strong> a three-helix bundle. Peptides at the N- and C-termini and a peptide linking helices<br />

two and three are disordered in the crystal structure. We have created point mutations and<br />

deletions to alter conserved surface features <strong>of</strong> RcdA. We expressed these proteins in a<br />

Caulobacter ∆rcdA mutant to identify regions <strong>of</strong> RcdA necessary for regulated CtrA degradation<br />

and for the polar localization <strong>of</strong> CtrA and RcdA itself.<br />

We have identified three types <strong>of</strong> rcdA mutants: 1) those that permit CtrA degradation<br />

and protein localization, 2) those that cannot support either CtrA degradation or protein<br />

localization, and 3), those that allow CtrA degradation without polar accumulation <strong>of</strong> either RcdA<br />

or CtrA. Surprisingly, RcdA does not need to be stably located at the cell pole to promote CtrA<br />

proteolysis. These results suggest that RcdA can act via transient interaction with other<br />

degradation proteins at the pole, or that RcdA’s function can be performed anywhere in the cell.<br />

For example, RcdA could inhibit an unknown negative regulator <strong>of</strong> CtrA proteolysis. We are<br />

examining the rcdA mutants in further detail and are screening for additional proteins that<br />

regulate CtrA degradation in Caulobacter.<br />

112


<strong>BLAST</strong> X Poster #62<br />

MODELLING MCP SIGNALLING MECHANISMS WITH HIGH-THROUGHPUT SIMULATION<br />

OF Tar TM2<br />

Benjamin A Hall, Mark SP Sansom<br />

Department <strong>of</strong> Biochemistry, <strong>University</strong> <strong>of</strong> Oxford, South Parks Road, OX1 3QU<br />

Transmembrane helices play a multiple vital roles in cell function, including intercell<br />

signalling processes, channel gating and active transport. As such, there exists a considerable<br />

amount <strong>of</strong> data on the biological function and structural properties <strong>of</strong> naturally occurring helices<br />

and their mutants. Simulation studies can provide insight into the dynamics and behaviour <strong>of</strong><br />

biomolecular systems in a variety <strong>of</strong> environments, but however such analyses are<br />

computationally expensive and typically difficult to automate. Coarse grain simulations are<br />

becoming an increasingly popular tool for understanding the properties <strong>of</strong> biological systems,<br />

overcoming canonical limits <strong>of</strong> atomistic simulations such as timescale or system size. Both<br />

such techniques involve several manual steps, including system build, simulation set up and<br />

analysis. Here we present Sidekick, a piece <strong>of</strong> s<strong>of</strong>tware which automates these processes to<br />

allow for the set up <strong>of</strong> massive numbers <strong>of</strong> coarse grain simulations on the basis <strong>of</strong> a small set<br />

<strong>of</strong> input sequences, or a single sequence and a scanning mutation. We demonstrate the use <strong>of</strong><br />

this s<strong>of</strong>tware to approach the mechanism <strong>of</strong> signalling in the MCPs, based on existing<br />

mutational data for TM2 from different MCPs and organisms. By observing the change in<br />

positions and orientations <strong>of</strong> the helix with different mutations, we propose that a piston model<br />

dominates the signalling event, though there may be a lesser role for rotation <strong>of</strong> the helix.<br />

113


<strong>BLAST</strong> X Poster #63<br />

STRUCTURAL EVIDENCE SUGGESTS THAT ANTIACTIVATOR ExsD FROM<br />

PSEUDOMONAS AERUGINOSA IS A DNA BINDING PROTEIN<br />

Robert C. Bernhards, Xing Jing, Nancy J. Vogelaar, Howard Robinson#, Florian D.<br />

Schubot*<br />

Department <strong>of</strong> Biological Sciences, Virginia Polytechnic Institute & State <strong>University</strong>,<br />

Washington Street, Blacksburg, VA 24060; #Biology Department, Brookhaven National<br />

Laboratory, Upton, NY 11973-5000.<br />

The opportunistic pathogen P.aeruginosa utilizes a type III secretion system (T3SS) to<br />

support acute infections in predisposed individuals. In this bacterium expression <strong>of</strong> all T3SSrelated<br />

genes is dependent on the AraC-type transcriptional activator ExsA. Prior to host<br />

contact, the T3SS is inactive and ExsA is repressed by the antiactivator protein ExsD. The<br />

repression, thought to occur through direct interactions between the two proteins, is relieved<br />

upon opening <strong>of</strong> the type III secretion (T3S) channel when secretion chaperone ExsC<br />

sequesters ExsD. We have solved the crystal structure <strong>of</strong> Δ20ExsD, a protease-resistant<br />

fragment <strong>of</strong> ExsD that lacks only the twenty amino terminal residues <strong>of</strong> the wild type protein at<br />

2.6 Å. Surprisingly the structure revealed similarities between ExsD and the DNA binding<br />

domain <strong>of</strong> transcriptional repressor KorB. A model <strong>of</strong> an ExsD-DNA complex constructed on the<br />

basis <strong>of</strong> this homology produced a realistic complex that is supported by the prevalence <strong>of</strong><br />

conserved residues in the putative DNA binding site and the results <strong>of</strong> differential scanning<br />

fluorimetry studies. Our findings challenge the currently held model that ExsD solely acts<br />

through interactions with ExsA and raise new questions with respect to the underlying<br />

mechanism <strong>of</strong> ExsA regulation.<br />

114


<strong>BLAST</strong> X Poster #64<br />

A NOVEL PAS-GGDEF-EAL PROTEIN INVOLVED IN REGULATION OF MOTILITY IN<br />

PSEUDOMONAS PUTIDA<br />

Herrera Seitz, K 1 and Shingler, V 2<br />

1 2<br />

IIB, FCEyN, Univ. Nac. del Mar del Plata, Molecular Biology Department, Umeå <strong>University</strong>,<br />

Sweden.<br />

Chemotaxis allows motile bacteria to respond to chemical gradients to relocate<br />

themselves near the source <strong>of</strong> nutrients. Soil Pseudomonads are known to be able to grow in a<br />

wide variety <strong>of</strong> carbon sources, including many that are considered environmentally toxic. Unlike<br />

previously characterized chemotactic responses in Pseudomonas strains, taxis <strong>of</strong> P. putida<br />

CF600 and P. putida KT2440 towards methyl-phenols is dependent upon its ability to<br />

metabolize the compound, rather than on a classical ligand-binding chemoreceptor. E. coli<br />

metabolism-dependent taxis responses are mediated by the Aer receptor that is closely related<br />

to chemoreceptors, but which contains a FAD-binding PAS sensory domain. P. putida<br />

possesses three aer-like genes. During analysis <strong>of</strong> the Aer-like receptors <strong>of</strong> P. putida, Aer-1 was<br />

found to be encoded in a dicistonic operon with PP2258, a PAS-GGDEF-EAL domain protein.<br />

Our attention was drawn to PP2258 because a null mutant was found to exhibit a general<br />

motility defect on solid, but not in liquid, media (Sarand et al., 2008).<br />

GGDEF- and EAL-domains are associated with diguanylate cyclase and<br />

phosphodiesterase activities that are involved in turnover <strong>of</strong> the near ubiquitous bacterial<br />

second messenger c-di-GMP. The levels <strong>of</strong> c-di-GMP can modify cells behavior and motility;<br />

therefore we reasoned that PP2258 link to motility via c-di-GMP signaling. As a first approach to<br />

test this idea, the biochemical properties <strong>of</strong> wild type and mutant derivatives <strong>of</strong> PP2258 were<br />

studied using over expression <strong>of</strong> the protein both in E. coli and P. putida.<br />

When PP2258 was over expressed in E. coli or P. putida cells, c-di-GMP levels were<br />

markedly increased compared to those <strong>of</strong> control cells, although accumulation <strong>of</strong> c-di-GMP was<br />

much lower in E. coli than in P. putida extracts. Alanine substitutions <strong>of</strong> the GGDEF domain<br />

associated with c-di-GMP synthesis causes a major decrease c-di-GMP accumulation, while an<br />

alanine substitution in EAL domain associated with c-di-GMP hydrolysis led to a >7-fold<br />

increase in accumulation. Together, these results suggest that PP2258 could be one <strong>of</strong> the rare<br />

examples <strong>of</strong> a dual GGDEF-EAL domain protein where both domains are catalytically active.<br />

References:<br />

Sarand, I., Österberg, S., Holmqvist, S., Holmfeldt, P. Skärfstad, E., Parale, R. E., & Shingler, V.<br />

(2008) Metabolism-dependent taxis towards (methyl)phenols is coupled through the most<br />

abundant <strong>of</strong> three polar localized Aer-like proteins <strong>of</strong> Pseudomonas putida. Environ.<br />

Microbiology. 10:1320-1334<br />

115


<strong>BLAST</strong> X Poster #65<br />

IN VIVO STUDY OF THE TWO-COMPONENT SIGNALLING NETWORK IN E. COLI<br />

Erik Sommer and Victor Sourjik<br />

Zentrum für Molekulare Biologie der Universität Heidelberg (ZMBH), Im Neuenheimer Feld 282,<br />

69120 Heidelberg, Germany<br />

Contact: e.sommer@zmbh.uni-heidelberg.de<br />

Two-component systems are the most widespread sensing systems in prokaryotes and<br />

lower eukaryotes, with multiple members <strong>of</strong> this class being present in one organism. We are<br />

interested in investigating the interconnection among different two-component signalling<br />

pathways in Escherichia coli. To map interactions between the pathways in vivo and to study<br />

relative cellular distribution <strong>of</strong> their proteins, we assay real-time dynamics <strong>of</strong> protein interactions<br />

and their dependencies on stimulation using fluorescence imaging and fluorescence resonance<br />

energy transfer (FRET)- and fluorescence recovery after photobleaching (FRAP)-microscopy.<br />

Additionally, intracellular processing <strong>of</strong> sensed stimuli with regard to amplification, integration<br />

and possible cross-talk between the systems will be investigated. Such analysis will help to<br />

establish an integral picture <strong>of</strong> cell signalling performed by prokaryotic organisms.<br />

116


<strong>BLAST</strong> X Poster #66<br />

GENE REGULATION BY ESCHERICHIA COLI RESPONSE REGULATOR PhoB<br />

Hua Han 1,2 , Timothy R. Mack 1,2 , Rong Gao 1.3 , and Ann M. Stock 1,3<br />

1 Center for Advanced Biotechnology and Medicine, UMDNJ-Robert Wood Johnson Medical<br />

School, Piscataway, New Jersey 08854, USA,<br />

2 Department <strong>of</strong> Biochemistry, Graduate School <strong>of</strong> Biomedical Sciences, UMDNJ-Robert Wood<br />

Johnson Medical School, and 3 Howard Hughes Medical Institute.<br />

Structural analysis <strong>of</strong> the Escherichia coli response regulator transcription factor PhoB<br />

indicates that the protein dimerizes in two different orientations, both <strong>of</strong> which are mediated by<br />

the receiver domain. The two dimers exhibit two-fold rotational symmetry: one involves the α4β5-α5<br />

surface and the other involves the α1/α5 surface. The α4-β5-α5 dimer is observed when<br />

the protein is crystallized in the presence <strong>of</strong> the phosphoryl analog BeF3 - while the α1/α5 dimer<br />

is observed in its absence. From these studies a model <strong>of</strong> the inactive and active states <strong>of</strong><br />

PhoB has been proposed that involves the formation <strong>of</strong> two distinct dimers. In order to gain<br />

further insight into the roles <strong>of</strong> these dimers we have engineered a series <strong>of</strong> mutations in PhoB<br />

intended to selectively perturb each <strong>of</strong> them. Our results indicate that perturbations to the α4β5-α5<br />

surface disrupt phosphorylation-dependent dimerization and DNA binding as well as<br />

PhoB mediated transcriptional activation <strong>of</strong> phoA, while perturbations to the α1/α5 surface do<br />

not. Furthermore, experiments with a GCN4 leucine zipper/PhoB chimera protein indicate that<br />

PhoB is activated through an intermolecular mechanism. Together these results support a<br />

model <strong>of</strong> activation <strong>of</strong> PhoB in which phosphorylation promotes dimerization via the α4-β5-α5<br />

face which in turn enhances DNA binding to a pair <strong>of</strong> direct-repeat half-sites – a model that we<br />

propose to be common for most all OmpR/PhoB transcription factors. These data contrast with<br />

a recent proposal that the α1/α5 dimer corresponds to the active form <strong>of</strong> PhoB, a conclusion<br />

derived from structural analysis <strong>of</strong> constitutively active mutant PhoB proteins (Arribas-<br />

Bosacoma et al., 2007 J. Mol. Biol. 366:626-641). We have also examined the kinetics <strong>of</strong> gene<br />

expression <strong>of</strong> several PhoB-regulated genes under conditions <strong>of</strong> phosphate limitation.<br />

117


<strong>BLAST</strong> X Poster #67<br />

NEW REPORTER RESIDUES OF TRIMER FORMATION BY ESCHERICHIA COLI MCPs<br />

Diego A. Massazza 1 , John S. Parkinson 2 , Claudia A. Studdert 1<br />

1 Instituto de Investigaciones Biológicas, Universidad Nacional de Mar del Plata, Argentina<br />

2 Biology Department, <strong>University</strong> <strong>of</strong> <strong>Utah</strong><br />

It is currently accepted that E. coli chemoreceptors are organized in vivo in trimer-<strong>of</strong>dimer<br />

arrangements. Using receptors bearing a cysteine reporter residue near the trimer axis,<br />

we previously demonstrated efficient formation <strong>of</strong> 2- and 3-subunit crosslinking products upon<br />

treatment <strong>of</strong> intact cells with the trifunctional maleimide reagent TMEA. In this work, we<br />

assessed the in vivo crosslinking behavior <strong>of</strong> receptors bearing cysteine reporters at different<br />

positions along their cytoplasmic domains.<br />

We found that the formation <strong>of</strong> 3-subunit crosslinking products declined with increasing<br />

distance <strong>of</strong> the reporter from the cytoplasmic tip. This result suggests that the in vivo<br />

arrangement <strong>of</strong> the trimer <strong>of</strong> dimers resembles that observed in the crystal structure <strong>of</strong> the Tsr<br />

cytoplasmic domain, in which the dimers contact one another at the tip and splay apart in the<br />

regions that are farther from the tip.<br />

We also found that stimulation with repellents immediately before the TMEA treatment<br />

caused a slight but reproducible increase in crosslinking efficiency. This behavior is consistent<br />

with previous observations suggesting that receptor dimers within the trimer move closer<br />

together after repellent stimuli.<br />

In cells co-expressing different marked receptors, we analyzed the formation <strong>of</strong> mixed<br />

crosslinking products. When the cysteine reporters were located the same distance from the tip,<br />

mixed crosslinking products formed with high efficiency. Mixed products did not form between<br />

reporters at different distances from the tip. This result suggests that there are no significant<br />

vertical displacements within the trimer <strong>of</strong> dimers, nor sufficient flexibility between the receptors<br />

to get crosslinking between reporters at diferent levels in the dimers.<br />

118


<strong>BLAST</strong> X Poster #68<br />

SYSTEMATIC DETECTION OF PROTEINS THAT LOCALIZE AT THE ATTACHMENT<br />

ORGANELLE REGIONS OF MYCOPLASMA PNEUMONIAE BY FLUORESCENT-PROTEIN<br />

TAGGING<br />

Tsuyoshi Kenri 1 , Atsuko Horino 1 , Mayumi Kubota 1 , Yuko Sasaki 1 , Daisuke Nakane 2 and<br />

Makoto Miyata 2<br />

1. Department <strong>of</strong> Bacterial Pathogenesis and Infection Control, National Institute <strong>of</strong> Infectious<br />

Disease, 4-7-1, Gakuen, Musashimurayama, Tokyo, 208-0011, Japan 2. Graduate School <strong>of</strong><br />

Science, Osaka City <strong>University</strong>, Sumiyoshi-ku, Osaka, 558-8585, Japan<br />

Mycoplasma pneumoniae is a cell wall-less bacterium with minimum range <strong>of</strong> genome<br />

size required for self-replication. It is also known as a general cause <strong>of</strong> bronchitis and<br />

pneumonia in humans. Pathogenicity <strong>of</strong> this bacterium is largely depends on its ability to attach<br />

to respiratory epithelial cells (cytadherence). The M. pneumoniae cells attached to cell surface<br />

also exhibit gliding motility. The cytadherence and gliding motility are mediated by a unique cell<br />

terminal structure <strong>of</strong> this bacterium, the attachment organelle, which is a membrane protrusion<br />

supported by internal cytoskeletal structures. Since the attachment organelle duplicates before<br />

cell division, the organelle formation is thought to be coordinated with cell cycles. A number <strong>of</strong><br />

protein components <strong>of</strong> the attachment organelle (cytadherence-related proteins) have been<br />

identified so far (HMW1, HMW2, HMW3, P1, P30, P65, P200, P24, P41, P90(B) and P40(C)).<br />

However, configuration and fine structure <strong>of</strong> these proteins in the organelle are not fully<br />

understood. In addition, the studies <strong>of</strong> Triton X-100 insoluble fraction <strong>of</strong> M. pneumoniae cells,<br />

cross-linking analysis <strong>of</strong> P1 adhesin protein and isolation <strong>of</strong> various gliding mutants are<br />

suggesting a possibility that there are more protein factors that associate with the attachment<br />

organelle components. In this study, to explore whether there are more proteins that localize at<br />

the organelle, we perform systematic fluorescent-protein tagging analysis for M. pneumoniae<br />

ORFs. At present, we have finished the localization analysis <strong>of</strong> 620 ORFs (about 90% <strong>of</strong> total<br />

ORFs). In this analysis, we observed that about 50 ORF products including previously known 9<br />

cytadherence-related proteins localized at the attachment organelle region. The other ORF<br />

products identified are 10 proteins involved in DNA, RNA and protein synthesis, 20 homologs <strong>of</strong><br />

variety <strong>of</strong> enzymes, 2 chaperons and 15 proteins <strong>of</strong> unknown function. Probably, some <strong>of</strong> these<br />

proteins may be the components <strong>of</strong> the attachment organelle and have function in formation <strong>of</strong><br />

the organelle, cytadherence, gliding motility and coordination <strong>of</strong> cell cycle.<br />

119


<strong>BLAST</strong> X Poster #69<br />

MAPPING THE SIGNAL TRANSDUCTION PATHWAY WITHIN THE PAS DOMAIN OF THE<br />

Aer RECEPTOR<br />

Asharie J. Campbell, Kylie J. Watts, Mark S. Johnson, and Barry L. Taylor<br />

Dept. Microbiology and Mol. Genetics, Loma Linda <strong>University</strong>, Loma Linda CA, USA<br />

The E. coli Aer receptor senses redox changes through an FAD-binding PAS domain,<br />

and transmits this redox status to the HAMP and signaling domains. Little is known about the<br />

conformational changes that take place within the PAS domain. In this study, we used errorprone<br />

PCR mutagenesis to find residues critical for PAS FAD-binding, sensing and signal<br />

transduction. We screened 10,000 colonies for function, measured expression for 1,300 clones,<br />

sequenced 400 mutants, and found 84 Aer aerotaxis-defective mutants that had just one amino<br />

acid substitution. Of these, there were 72 substitutions in the PAS domain, 11 in the F1 region<br />

and 1 in the TM region. The swimming behavior <strong>of</strong> the cells expressing these mutant proteins<br />

included those that 1) were locked in a smooth swimming (CCW), "signal-<strong>of</strong>f" state (60/84), 2)<br />

had increased tumbling (CW) frequency (4/84) or 3) were locked in the CW "signal-on" state<br />

(20/84). Approximately half (49/84) <strong>of</strong> the Aer mutants were functionally rescued by Tar. Mutant<br />

proteins (11/84) that expressed at levels less than 30% <strong>of</strong> wild-type Aer showed enhanced<br />

protein degradation rates. These replacements had altered side-chain polarity, mapped to<br />

positions on or near loops and, with one exception, yielded a CCW (signal-<strong>of</strong>f) phenotype. In<br />

contrast, all but one <strong>of</strong> the CW (signal-on) mutants were stable, and clustered in three localized<br />

regions: 1) in the putative FAD-binding cleft, 2) on the rear surface <strong>of</strong> the FAD-binding cleft and<br />

3) at or near a loop in the N-terminal Cap region. When expressed at a 1:1 ratio with wild-type<br />

Aer, three CW-locked mutants were dominant, abolishing wild-type mediated aerotaxis. Most<br />

but not all <strong>of</strong> the FAD-binding lesions resulted in an unstable protein; those that were most<br />

stable were located outside <strong>of</strong> the putative FAD-binding cleft.<br />

The localization <strong>of</strong> gain-<strong>of</strong>-function (CW) lesions to three distinct clusters suggests that<br />

conformational changes in these specific regions mimic the signal-on state <strong>of</strong> the PAS sensor.<br />

If so, signaling within the Aer-PAS sensor would begin in the FAD-binding cleft and propagate<br />

outward to the N-Cap loop. Previously, we found that removing part <strong>of</strong> the N-Cap mimics the<br />

signal-on state. Thus, an attractive model is one where FAD reduction in the PAS domain<br />

initiates a conformational change that propagates to the N-Cap loop, which, in turn, acts as a<br />

hinge around which the N-cap moves, unmasking the signal-on state.<br />

120


<strong>BLAST</strong> X Poster #70<br />

MODULATING TWO-COMPONENT SIGNAL OUTPUT WITH PROTEIN-MEMBRANE<br />

INTERACTIONS<br />

Roger R. Draheim*, Morten H. H. Nørholm, Salomé C. Botelho, Karl Enquist, and Gunnar von<br />

Heijne<br />

Center for Biomembrane Research, Department <strong>of</strong> Biochemistry and Biophysics, The Arrhenius<br />

Laboratories for Natural Sciences, Stockholm, <strong>University</strong>, 10691, Stockholm, Sweden<br />

The most prevalent type <strong>of</strong> environmental sensor in prokaryotic organisms is the twocomponent<br />

system (TCS). A canonical TCS consists <strong>of</strong> a membrane-spanning sensor histidine<br />

kinase (SHK) and a cytoplasmic response regulator (RR). TCSs have been shown to regulate a<br />

diverse array <strong>of</strong> virulence factors, therefore identifying two-component signaling pathways that<br />

lead to enhanced pathogenicity is essential to understanding complex host-pathogen<br />

interactions. The specific hypothesis examined is that SHK-membrane interactions can be<br />

identified and harnessed to modulate SHK signal output in a predictable manner. If individual<br />

SHK signal output could be directly manipulated, then two-component signaling pathways could<br />

be rapidly unraveled in any pathogenic organism <strong>of</strong> interest.<br />

Three discrete steps are proposed to establish a broadly-applicable methodology. The<br />

first will identify interactions between TM2 <strong>of</strong> a well-characterized SHK (e.g., EnvZ) and the cell<br />

membrane using a glycosylation-mapping technique. The second will identify HAMP-membrane<br />

interactions within EnvZ using a translocon-challenge method. Protein-membrane interactions<br />

that are identified will be confirmed using circular dichroism and fluorescent techniques. The<br />

third will harness protein-membrane interactions identified during the first two to directly and<br />

incrementally modulate SHK signal output. This experimentation will determine the feasibility <strong>of</strong><br />

coupling modulated SHK output with transcriptional pr<strong>of</strong>iling to rapidly unravel two-component<br />

signaling pathways in any pathogenic organism <strong>of</strong> interest. A high-throughput approach using<br />

fluorescent transcriptional reporter systems has been established to expedite these steps.<br />

The long-term goal <strong>of</strong> this research is to create a method for rapidly identifying the<br />

signaling pathways that regulate the virulence <strong>of</strong> pathogenic microorganisms. This research will<br />

lead to a better understanding <strong>of</strong> complex host-pathogen interactions and will result in the<br />

detection <strong>of</strong> previously unidentified therapeutic targets.<br />

121


<strong>BLAST</strong> X Poster #71<br />

Poster Cancelled<br />

122


<strong>BLAST</strong> X Poster #72<br />

A MECHANICAL AND GENETIC STUDY OF ESCHERICHIA COLI SWARMING MOTILITY<br />

Matthew Copeland and Douglas B. Weibel<br />

Department <strong>of</strong> Biochemistry, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, WI 53706 USA<br />

Bacterial swarming is a phenotype associated with the motility <strong>of</strong> bacteria across<br />

surfaces in search <strong>of</strong> resources. In this abstract we describe two approaches to understand this<br />

phenotype in Escherichia coli; a ‘mechanistic’ approach to elucidate potential flagella/flagella<br />

interactions between adjacent swarming cells and a genetic investigation to chart the<br />

expression <strong>of</strong> genes unique to the swarming phenotype.<br />

In contrast to swimming motility, swarming bacteria cells migrate cooperatively across<br />

surfaces. The role <strong>of</strong> physical interactions in the coordination <strong>of</strong> swarming motility is unknown. It<br />

has been shown that swarming bacteria align along their long axis and move as multicellular<br />

rafts from which the characteristic dynamic swirling patterns <strong>of</strong> swarming emerge. The<br />

alignment <strong>of</strong> cells may facilitate or accompany the intercellular bundling <strong>of</strong> flagella between<br />

adjacent cells and play a role in the characteristic, coordinated movement observed during<br />

swarming motility. To test this hypothesis, we have created strains <strong>of</strong> E. coli with fluorescent<br />

flagella and are using space- and time-resolved fluorescence resonance energy transfer (FRET)<br />

to measure flagella/flagella interactions.<br />

Iron starvation is known to signal swarmer cell differentiation in Vibrio parahaemolyticus<br />

and the genes for iron uptake and metabolism have been shown to be elevated in swarming<br />

populations <strong>of</strong> Salmonella typhimurium. We have found that iron metabolism and iron<br />

acquisition genes are upregulated in swarming cells <strong>of</strong> E. coli versus planktonic cells. A specific<br />

role for iron in E. coli swarming cells is unknown. Like V. parahaemolyticus, iron starvation may<br />

be a signal for swarmer cell differentiation in E. coli or swarming development and motility may<br />

require intracellular levels <strong>of</strong> iron in excess <strong>of</strong> those necessary for growth under vegetative<br />

conditions. We are exploring the role <strong>of</strong> iron in E. coli swarming using a combination <strong>of</strong> chemical<br />

and gene expression techniques and strains containing iron metabolism gene knockouts.<br />

123


<strong>BLAST</strong> X Poster #73<br />

PROBING HYDRODYNAMIC INTERACTRIONS BETWEEN SWIMMING BACTERIA USING<br />

MICROFULIDICS<br />

Abishek Muralimohan, Douglas B. Weibel<br />

Department <strong>of</strong> Biochemistry, <strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

433 Babcock Drive, Madison, WI 53706, U.S.A.<br />

Populations <strong>of</strong> self propelled motile bacteria have been shown to exhibit collective<br />

swimming behavior, leading to large scale fluid motion and fluid transport 1 . Such collective<br />

behavior arises from hydrodynamic interactions (HI) between the individual bacterial cells. While<br />

computational studies have been performed on HI between pairs <strong>of</strong> swimming cells 2 , there have<br />

been no reports on experimental measurements <strong>of</strong> HI. Here, we describe a PDMS based<br />

micr<strong>of</strong>luidic device that promotes pair wise interactions between swimming bacteria. The device<br />

generates collisions between pairs <strong>of</strong> cells and HI between them is quantified by the length<br />

scale (persistence) over which the two cells swim together as a single unit. We have used this<br />

device to quantify the strength <strong>of</strong> HI between pairs <strong>of</strong> filamentous E. coli hcb437 cells based on<br />

a number <strong>of</strong> parameters including cell length, angle <strong>of</strong> initial contact, trajectory, and velocity.<br />

Our observations indicate two subpopulations based on the persistence length <strong>of</strong> interaction:<br />

long persisters are characterized by low angles <strong>of</strong> contact (9.0 µm); short<br />

persisters are characterized by a wider range <strong>of</strong> contact angles (10° – 60°) and shorter cells (6<br />

µm -9 µm). We believe that our micr<strong>of</strong>luidics-based approach is well suited to the study <strong>of</strong><br />

bacterial interactions at a single cell level. A broader understanding <strong>of</strong> HI between motile<br />

bacteria will shed light on other forms <strong>of</strong> collective motility, including swarming.<br />

1. Underhill P, et. al. (2008) Phys. Rev. Lett. 100: 248101.<br />

2. Ishikawa T, et al. (2007) Biophys. J. 93, 2217.<br />

124


<strong>BLAST</strong> X Poster #74<br />

EXAMINATION OF PHOSPHORYLATION IN THE Dif CHEMOTAXIS-LIKE SYSTEM IN<br />

MYXOCOCCUS XANTHUS<br />

Wesley P. Black and Zhaomin Yang<br />

Department <strong>of</strong> Biological Sciences, Virginia Polytechnic and State <strong>University</strong>, Life Sciences I,<br />

Washington Street, Blacksburg, VA 24061<br />

Myxococcus xanthus is able to move on surfaces using social (S) gliding motility, which<br />

is powered by the retraction <strong>of</strong> type IV pili (Tfp). Exopolysaccharides (EPS), also essential for S<br />

motility, have been proposed to function as the anchor and trigger for Tfp retraction in M.<br />

xanthus. EPS production in M. xanthus is regulated by the Dif chemotaxis-like signal<br />

transduction pathway. DifA, DifC and DifE, homologous to methyl-accepting chemotaxis<br />

proteins (MCPs), CheW and CheA respectively, are positive regulators <strong>of</strong> EPS production. DifD<br />

and DifG, which are respective homologs <strong>of</strong> CheY and CheC, are negative regulators <strong>of</strong> EPS<br />

production. It was demonstrated previously that DifD (CheY-like) does not function downstream<br />

<strong>of</strong> DifE (CheA-like) in the regulation <strong>of</strong> EPS production.<br />

The purpose <strong>of</strong> this study was to examine the phosphorylation <strong>of</strong> heterologously<br />

expressed and purified Dif proteins in vitro. Protein phosphorylation, phosphate transfer and<br />

dephosporylation were monitored using γ- 32 P-ATP. We demonstrate the autophosphorylation <strong>of</strong><br />

DifE, which is an atypical CheA-like kinase with an extra domain. Unlike observations in other<br />

systems, the presence <strong>of</strong> both DifA (MCP-like) and DifC (CheW-like) had inhibitory rather than<br />

stimulatory effects on DifE autophosphorylation. In addition, we show that the phosphate from<br />

DifE-phosphate can be transferred to DifD (CheY-like). Lastly, DifG, which is similar to the CheC<br />

phosphatase, accelerates the dephosphorylation <strong>of</strong> DifD. These results are consistent with a<br />

model where the DifE kinase positively regulates EPS production, while DifD and DifG function<br />

collectively to divert phosphate away from an unidentified downstream phosphorylation<br />

substrate <strong>of</strong> DifE.<br />

125


<strong>BLAST</strong> X Poster #75<br />

EVOLUTION OF CHEMOTAXIS PROTEINS ON A MICRO SCALE<br />

Brian Cantwell and Igor Zhulin<br />

<strong>University</strong> <strong>of</strong> Tennessee, Department <strong>of</strong> Microbiology, Knoxville, TN<br />

Oak Ridge National Laboratory, Joint Institute for Computational Sciences, Oak Ridge, TN<br />

Computational analysis <strong>of</strong> very closely related genomes <strong>of</strong>fers the advantage <strong>of</strong> more<br />

accurate tracing <strong>of</strong> evolutionary events. The chemotaxis system <strong>of</strong> enteric bacteria Escherichia<br />

coli and Salmonella enterica are extremely well studied, and the availability <strong>of</strong> over fifty<br />

sequenced genomes <strong>of</strong> Enterobacteriacea <strong>of</strong>fers an opportunity to determine the<br />

microevolutionary trends in chemotaxis <strong>of</strong> enterics. Similarly, seventeen sequenced genomes<br />

are available for the family Shewanellaceae for which Shewanella oneidensis MR-1 has been<br />

most studied by genetic and biochemical methods. In this work we examine the evolution <strong>of</strong><br />

the chemotaxis systems <strong>of</strong> the Enterobacteriaceae and Shewanellaceae using computational<br />

biology methods. Protein sequences <strong>of</strong> chemotaxis proteins and receptors were extracted from<br />

non-redundant database by matching to domain models and organized into orthologous groups<br />

based on reciprocal <strong>BLAST</strong> hits, genome context, and phylogenetic relationships. All<br />

Enterobacteriacea contain a single set <strong>of</strong> chemotaxis proteins and chemoreceptors orthologous<br />

to the E. coli Tsr, Tap, and Aer proteins. Most enteric species contain numerous additional<br />

chemoreceptors including multiple orthologs <strong>of</strong> the E. coli Trg protein. Shewanella species have<br />

one common set <strong>of</strong> chemotaxis proteins with some species having an additional set <strong>of</strong><br />

chemotaxis proteins. Two chemoreceptors are common among all seventeen Shewanella<br />

species with an additional ten chemoreceptors found in at least fifteen <strong>of</strong> the seventeen<br />

sequenced genomes. To examine the evolutionary trends within chemotaxis proteins, we<br />

compared orthologous proteins by computing pairwise percentage identity and comparing<br />

domains across species, genus, and family lines. As expected, the domains with highest<br />

conservation include the catalytic domains <strong>of</strong> the core chemotaxis proteins as well as the<br />

signaling subdomain <strong>of</strong> the chemoreceptors. The most divergent domains include the P2<br />

domains <strong>of</strong> CheA, sensory domains <strong>of</strong> chemoreceptors, HAMP domains, and the methylation<br />

subdomains <strong>of</strong> the Aer proteins.<br />

126


<strong>BLAST</strong> X Poster #76<br />

EVOLUTION OF SIGNAL TRANSDUCTION IN A BACTERIAL GENUS<br />

Harold Shanafield, Luke E. Ulrich and Igor B. Zhulin<br />

National Institute for Computational Sciences, <strong>University</strong> <strong>of</strong> Tennessee – Oak Ridge National<br />

Laboratory<br />

The recent completion <strong>of</strong> sequencing <strong>of</strong> multiple genomes in the Shewanella genus<br />

provides a unique opportunity to study evolution at a much finer scale than previously possible.<br />

Using a bi-directional best <strong>BLAST</strong> hit approach at the protein domain rather than traditional<br />

whole-protein sequence level we analyzed the evolutionary relationships <strong>of</strong> proteins predicted to<br />

be involved in signal transduction. Based on these relationships, we have determined a core set<br />

<strong>of</strong> 99 proteins across the first 11 sequenced Shewanella genomes that were highly conserved in<br />

both domain architecture and protein sequence. The core included one <strong>of</strong> the several<br />

chemotaxis systems found in Shewanella and several two-component regulatory systems. A<br />

large group <strong>of</strong> orthologous signal transduction proteins across multiple genomes showed some<br />

primary sequence drift and were classified as “significant similarity”, and finally there were<br />

several unique signal transduction proteins in each organism. We also quantified a recent<br />

disproportionate loss <strong>of</strong> signal transduction genes in Shewanella denitrificans OS217 above and<br />

beyond the overall reduction in that organism’s genome size, and an enrichment <strong>of</strong> signal<br />

transduction genes in Shewanella amazonensis SB2B. Possible relationships <strong>of</strong> the observed<br />

changes with the metabolism and environment are discussed.<br />

127


<strong>BLAST</strong> X Poster #77<br />

THE CHEMOSENSORY RECEPTOR FrzCD INTERACTS WITH TWO A-MOTILITY<br />

PROTEINS, AglZ AND AgmU<br />

Beiyan Nan and David R. Zusman<br />

Department <strong>of</strong> Molecular and Cell Biology, <strong>University</strong> <strong>of</strong> California, Berkeley, CA 94720, USA.<br />

The Frz chemosensory system <strong>of</strong> Myxococcus xanthus controls directed motility by regulating<br />

cellular reversals. FrzCD, the Frz system chemoreceptor, plays a central role in the Frz signal<br />

transduction pathway. Recently, evidence was obtained that AglZ, an A-motility protein, and<br />

FrzS, an S-motility protein, are localized in separate complexes that change their positions as<br />

cells move forward and reverse. We were interested in studying how FrzCD might communicate<br />

with these motility complexes. To gain information on these interactions, we have been doing<br />

pull-down experiments using Myxococcus cell extracts and GST-tagged FrzCD as bait. The<br />

GST-FrzCD interacting proteins were identified by mass spectrometry. Five proteins were found<br />

to reproducibly bind to GST-FrzCD besides FrzA, FrzB and FrzE. Two <strong>of</strong> these were A-motility<br />

associated protein, AglZ and AgmU. The interactions were confirmed using formaldehyde crosslinking,<br />

which showed that FrzCD interacts directly with the N-terminal pseudo-receiver domain<br />

<strong>of</strong> AglZ and two TPR clusters <strong>of</strong> AgmU. These results provide preliminary evidence for a direct<br />

role in the control <strong>of</strong> the A-motility system by the FrzCD receptor. The roles <strong>of</strong> the other FrzCDinteracting<br />

proteins remain to be identified.<br />

128


<strong>BLAST</strong> X Poster #78<br />

PHENOTYPIC CHARACTERIZATION OF ALL SINGLE MUTANTS OF TWO COMPONENT<br />

SYSTEM PROTEINS IN NEUROSPORA CRASSA<br />

Barba C., Chavez-Canales M., Salas G., Hernandez C., Sanchez O and Georgellis D.<br />

Department <strong>of</strong> Molecular Genetics, Instituto de Fisiologia Celular, UNAM<br />

Perception and response to environmental stimuli is essential for the growth and survival<br />

<strong>of</strong> all organisms. The sensing and processing <strong>of</strong> these stimuli are carried out by molecular<br />

circuits within the cell, which detect, amplify and integrate them into a specific response. In<br />

prokaryotes these molecular circuits are typically organized by protein pairs, "sensory kinase"<br />

proteins (SK) and "response regulator" proteins (RR) that belong to the large family <strong>of</strong> two<br />

component systems (TCS). This organization implies that each SK activates its cognate RR,<br />

and thus providing specificity to signal propagation and output. However, an interesting variation<br />

<strong>of</strong> TCS architecture is observed in filamentous fungi where multiple SKs appear to use a single<br />

phosphotransfer protein to relay signals to a few RRs. Therefore, the question <strong>of</strong> whether a<br />

specific signal can generate a specific response, or whether the various signals sensed by<br />

individual SKs result in the same response, is raised. To explore this intriguing question we<br />

used Neurospora crassa, which has eleven SKs, one phosphotransfer protein and three RRs,<br />

as our model. Here, by using single mutants <strong>of</strong> all SKs and RRs, we demonstrate that these<br />

signaling cascades are involved in the regulation <strong>of</strong> various developmental processes, and in<br />

responses to environmental conditions, such as osmotic, oxidative and fungicide stress.<br />

129


<strong>BLAST</strong> X ____________ Poster #79<br />

AFM STUDY OF MYCOPLASMA MOBILE’S GLIDING MOTILITY<br />

Charles Lesoil (1) , Hiroshi Sekiguchi (1) , Takahiro Nonaka (2) , Makoto Miyata (2) , Toshiya<br />

Osada (1) , Atsushi Ikai (3)<br />

(1) Department <strong>of</strong> life science, Graduate school <strong>of</strong> Bioscience and Biotechnology, Tokyo<br />

Institute <strong>of</strong> Technology.<br />

(2) Department <strong>of</strong> Biology , Graduate school <strong>of</strong> Science, Osaka City <strong>University</strong><br />

(3) Innovation Research Center, Tokyo Institute <strong>of</strong> Technology<br />

Mycoplasma mobile is a parasitic bacterium that lacks the peptidoglycan layer but still<br />

presents recognizable flask-shape cell morphology. It glides along cell or glass surfaces at an<br />

average speed <strong>of</strong> 2.0 to 4.5 μm/s towards the tapered end <strong>of</strong> the cell called the head, with a<br />

unique mechanism. Recent studies have identified four proteins Gli23, Gli349 and Gli521 that<br />

are involved in this system, and a model for gliding motility has been presented, but more<br />

experimental data are needed to obtain the arrangement and detailed role <strong>of</strong> each protein in this<br />

system.<br />

In this study, we propose a novel approach to investigate the gliding motility system <strong>of</strong><br />

M. mobile using an AFM (Atomic Force Microscope) both as an imaging and a force<br />

measurement device.<br />

AFM Images <strong>of</strong> biotinylated M. mobile cells were obtained through immobilization on a<br />

Streptavidin modified mica surface. Pictures showing the morphology <strong>of</strong> individual Gli349 and<br />

Gli521 molecules in dried and liquid conditions were also obtained and were consistent with<br />

previous electron micrographs <strong>of</strong> the proteins. Investigation <strong>of</strong> the interaction between Gli<br />

molecules and Sialyllactose, the direct binding target in gliding was also conducted using AFM<br />

tips decorated with Gli349 or Gli521 molecules and the results showed a specific interaction<br />

between Gli349 and Sialyllactose, whereas Gli521 did not show any interaction. Nano<br />

indentation <strong>of</strong> living cells was also performed and revealed a great variation in the local stiffness<br />

<strong>of</strong> the cell, consistent with the available information about M. mobile’s cytoskeleton.<br />

129b


<strong>BLAST</strong> X PARTICIPANT LIST<br />

130


Christopher Adase<br />

Texas A&M <strong>University</strong><br />

3258 TAMU<br />

BSBE Room 303<br />

College Station, TX 77843<br />

Phone: (979) 845-1249<br />

xpage@neo.tamu.edu<br />

Michael Airola<br />

Cornell <strong>University</strong><br />

G63 S.T. Olin Laboratory<br />

Chemistry Research Building<br />

Ithaca, NY 14853<br />

Phone: (607) 255-4970<br />

mva5@cornell.edu<br />

Christine Aldridge<br />

Newcastle <strong>University</strong><br />

The Medical School<br />

Framlington Place<br />

Newcastle upon Tyne<br />

NE2 4HH<br />

United Kingdom<br />

Phone: +44 1912227704<br />

Fax: +44 1912227424<br />

christine.aldridge@ncl.ac.uk<br />

Phillip Aldridge<br />

Newcastle <strong>University</strong><br />

The Medical School<br />

Framlington Place<br />

Newcastle upon tyne<br />

NE2 4HH<br />

United Kingdom<br />

Phone: +44 1912227704<br />

Fax: +44 1912227424<br />

p.d.aldridge@ncl.ac.uk<br />

Roger Alexander<br />

Yale <strong>University</strong><br />

Kline Biology Tower 1054<br />

New Haven, CT 06511<br />

Phone: (404) 217-5664<br />

roger.alexander@yale.edu<br />

131<br />

Gladys Alexandre<br />

The <strong>University</strong> <strong>of</strong> Tennessee<br />

M407 Walters Life Sciences<br />

1414 West Cumberland Ave.<br />

Knoxville, TN 37996<br />

Phone: (865) 974-0866<br />

Fax: (865) 974-6306<br />

galexan2@utk.edu<br />

Adrian Alvarez<br />

Instituto de Fisiologia Celular - UNAM<br />

Ciudad Universitaria<br />

México D.F. 04510<br />

México<br />

Phone: +525556225738<br />

aalvarez@ifc.unam.mx<br />

Angel Andrade<br />

UNAM<br />

Circuito interior Cd Universitaria<br />

México City 05100<br />

México<br />

Phone: +5256225965<br />

aandrade@live.com.mx<br />

Judy Armitage<br />

Oxford<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: +44 1865 613293<br />

armitage@bioch.ox.ac.uk<br />

Teresa Ballado<br />

UNAM<br />

Instituto de Fisiologia Celular<br />

Circuito Exterior S/N Cd. Universitaria<br />

México City 04510<br />

México<br />

Phone: (5255) 56225618<br />

Fax: (5255) 56225611<br />

tballado@ifc.unam.mx


Rina Barak<br />

Weizmann Institute <strong>of</strong> Science<br />

Herzel St.<br />

Rehovot 76100<br />

Israel<br />

Phone: 972-8-9342710<br />

Fax: 972-8-9344112<br />

rina.barak@weizmann.ac.il<br />

Carlos Arturo Barba Ostria<br />

Instituto de Fisiologia Celular - UNAM<br />

Ciudad Universitaria<br />

México D.F. 04510<br />

México<br />

Phone: +525556225738<br />

cbarba@ifc.unam.mx<br />

Robert Belas<br />

<strong>University</strong> <strong>of</strong> Maryland<br />

Biotechnology Institute<br />

701 East Pratt Street<br />

Baltimore, MD 21202<br />

Phone: (410) 234-8876<br />

belas@umbi.umd.edu<br />

James Berleman<br />

<strong>University</strong> <strong>of</strong> Iowa<br />

51 Newton Rd.<br />

Iowa City, IA 52242<br />

Phone: (404) 372-4836<br />

berleman@gmail.com<br />

Jaya Bhatnagar<br />

Cornell <strong>University</strong><br />

Chemistry Research Bldg.<br />

Ithaca, NY 14853<br />

Phone: (607) 255-4970<br />

jb394@cornell.edu<br />

132<br />

Amber Bible<br />

<strong>University</strong> <strong>of</strong> Tennessee, Knoxville<br />

M407 Walters Life Sciences<br />

1414 Cumberland Avenue<br />

Knoxville, TN 37996<br />

Phone: (865) 974-2364<br />

abible@utk.edu<br />

Paola Bisicchia<br />

Trinity College<br />

Lincoln Place Gate<br />

Smurfit Institute <strong>of</strong> Genetics<br />

Dublin 2<br />

Ireland<br />

Phone: 003538962447<br />

paolabisicchia@yahoo.it<br />

Wesley Black<br />

Virginia Tech<br />

Life Sciences I<br />

Washington St.<br />

Blacksburg, VA 24061<br />

Phone: (540) 231-9381<br />

weblack@vt.edu<br />

Bob Bourret<br />

<strong>University</strong> <strong>of</strong> North Carolina<br />

Chapel Hill, NC 27599-7290<br />

Phone: (919) 966-2679<br />

Fax: (919) 962-8103<br />

bourret@med.unc.edu<br />

Richard Branch<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

Clarendon Laboratory<br />

Parks Road<br />

Oxford<br />

OX1 3PU<br />

United Kingdom<br />

Phone: +44 01865272357<br />

r.branch1@physics.ox.ac.uk


Ariane Briegel<br />

California Institute <strong>of</strong> Technology<br />

1200 E. California Blvd.<br />

Mail code 114-96<br />

Pasadena, CA 91125<br />

Phone: (626) 395-8848<br />

briegel@caltech.edu<br />

Mostyn Brown<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: +44 01865 275 298<br />

mostyn.brown@bioch.ox.ac.uk<br />

Iryna Bulyha<br />

Max Planck Institute<br />

for Terrestrial Microbiology<br />

Karl-von-Frisch-Strasse, 8<br />

35043 Marburg<br />

Germany<br />

Phone: +496421178222<br />

Fax: +496421178209<br />

bulyha@mpi-marburg.mpg.de<br />

Victor Bustamante<br />

Universidad Nacional Autonoma de México<br />

Av. Universidad 2001 Colonia Chamilpa<br />

Cuernavaca, Morelos 62210<br />

México<br />

Phone: (52) 777 329 16 27<br />

victor@ibt.unam.mx<br />

Edmundo Calva<br />

Instituto de Biotecnología UNAM<br />

Av. Universidad 2001<br />

Cuernavaca 62210<br />

México<br />

Phone: +52-777-329-1645<br />

Fax: +52-777-313-8673<br />

ecalva@ibt.unam.mx<br />

133<br />

Karen Camargo<br />

Universidad Nacional Autonoma de México<br />

C.U. Instituto de Fisiologia Celular<br />

México Distrito Federal 04510<br />

México<br />

Phone: 52 556225618<br />

karen_100490@hotmail.com<br />

Asharie Campbell<br />

Loma Linda <strong>University</strong><br />

11021 Campus Street<br />

Loma Linda, CA 92350<br />

Phone: (909) 558-1000<br />

ajohnson07b@llu.edu<br />

Eva Campodonico<br />

<strong>University</strong> <strong>of</strong> California, Berkeley<br />

31 Koshland Hall<br />

Berkeley, CA 94720<br />

Phone: (510) 643-5457<br />

eva.campodonico@berkeley.edu<br />

Brian Cantwell<br />

<strong>University</strong> <strong>of</strong> Tennessee<br />

WLS 437<br />

Knoxville, TN 37996-0845<br />

Phone: (865) 974-7687<br />

Fax: (865) 974-4007<br />

bcantwe1@utk.edu<br />

C. Britt Carlson<br />

HHMI/UMDNJ<br />

Center for Advanced Biotechnology and<br />

Medicine<br />

679 Hoes Lane, Room 324<br />

Piscataway, NJ 08854<br />

Phone: (732) 235-4206<br />

carlson@cabm.rutgers.edu


David Castillo<br />

Universidad Nacional Autónoma de México<br />

IFC-UNAM México D.F., C.P. 04510<br />

México City 04510<br />

México<br />

Phone: (+52) 55 562<br />

Fax: (+52) 55 56225611<br />

castillo@ifc.unam.mx<br />

Matt Chapman<br />

<strong>University</strong> <strong>of</strong> Michigan<br />

830 North <strong>University</strong><br />

Ann Arbor, MI 48109<br />

Phone: (734) 764-7592<br />

Fax: (734) 647-0884<br />

chapmanm@umich.edu<br />

Arnaud Chastanet<br />

Harvard <strong>University</strong><br />

16 Divinity ave<br />

Biolabs<br />

Cambridge, MA 02138<br />

Phone: (617) 384-7622<br />

Fax: (617) 496-4642<br />

arnaud@mcb.harvard.edu<br />

Yong-Suk Che<br />

Osaka <strong>University</strong><br />

Suita1-3 Yamadaoka<br />

Osaka 565-0871<br />

Japan<br />

Phone: +81-6-6879-4625<br />

yongsuk@fbs.osaka-u.ac.jp<br />

Matthew Copeland<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

473 Biochemistry Addition<br />

433 Babcock Drive<br />

Madison, WI 53706<br />

Phone: (608) 263-2636<br />

mfcopeland@wisc.edu<br />

134<br />

Brian Crane<br />

Cornell <strong>University</strong><br />

Baker Laboratory<br />

Ithaca, NY 14850<br />

Phone: (607) 254-8634<br />

Fax: (607) 255-1248<br />

bc69@cornell.edu<br />

Sean Crosson<br />

<strong>University</strong> <strong>of</strong> Chicago<br />

929 E. 57th St.<br />

GCIS-W138<br />

Chicago, IL 60637<br />

Phone: (773) 834-1926<br />

scrosson@uchicago.edu<br />

Rachel Crowder<br />

Texas A&M <strong>University</strong><br />

3258 TAMU<br />

College Station, TX 77843<br />

Phone: (979) 845-1249<br />

rcrowder@tamu.edu<br />

Rick Dahlquist<br />

Dept <strong>of</strong> Chemistry & Biochemistry<br />

<strong>University</strong> <strong>of</strong> California Santa Barbara<br />

Santa Barbara, CA 93106<br />

Phone: (805) 893-5326<br />

dahlquist@chem.ucsb.edu<br />

Miguel De la Cruz<br />

Instituto de Biotecnología UNAM<br />

Av Universidad 2001 Col Chamilpa<br />

Cuernavaca 62210<br />

México<br />

Phone: 52 777 3 29 16 27<br />

mike@ibt.unam.mx


Javier De la Mora<br />

Universidad Nacional Autonoma de México<br />

C.U., Instituto de Fisiologia Celular<br />

México, Distrito Federal 04510<br />

México<br />

Phone: 52 556225618<br />

fbravo@ifc.unam.mx<br />

Nicolas Delalez<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

Biochemistry Dept, Microbiology Unit,<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: +44 01865613315<br />

nicolas.delalez@bioch.ox.ac.uk<br />

Aurelia Delaune<br />

Institut Pasteur<br />

25 Rue du Dr. Roux<br />

75015 Paris<br />

France<br />

Phone: 33 1 45 68 88 48<br />

Fax: 33 1 45 68 89 38<br />

aureliad@pasteur.fr<br />

Kevin Devine<br />

Trinity College Dublin<br />

Dublin 2<br />

Ireland<br />

Phone: (353)-1-896-1872<br />

Fax: (353)-1-6714968<br />

kdevine@tcd.ie<br />

Miguel Diaz<br />

UNAM<br />

Ciudad Universitaria<br />

Apartado postal 70-243<br />

México 04510<br />

México<br />

Phone: (52) 56225965<br />

madiaz@ifc.unam.mx<br />

135<br />

Edith Diaz-Mireles<br />

Newcastle <strong>University</strong> Medical School<br />

Framlington Place<br />

Newcastle Upon Tyne<br />

NE2 4HH<br />

United Kingdom<br />

Phone: +44 0191-2228947<br />

Fax: +44 0191-2227424<br />

edith.diaz-mireles@ncl.ac.uk<br />

Jason Dobkowski<br />

<strong>University</strong> <strong>of</strong> Michigan<br />

830 N. <strong>University</strong><br />

Ann Arbor, MI 48103<br />

Phone: (734) 647-5677<br />

jdobkow@umich.edu<br />

Roger Draheim<br />

Stockholm <strong>University</strong><br />

DBB - von Heijne Group<br />

Svante Arrhenius vag 12, plan 4<br />

10691 Stockholm<br />

Sweden<br />

Phone: +4686747656<br />

rogerdraheim@gmail.com<br />

Georges Dreyfus<br />

UNAM<br />

Instituto de Fisiologia Celular<br />

Circuito exterior s/n Cd. Universitaria<br />

México City 04510<br />

México<br />

Phone: (5255) 56225618<br />

gdreyfus@ifc.unam.mx<br />

Sarah Dubrac<br />

Institut Pasteur<br />

25 Rue du Dr. Roux<br />

75015 Paris<br />

France<br />

Phone: (33) 1-45-68-88-48<br />

Fax: (33) 1 45 68 89 38<br />

sdubrac@pasteur.fr


Michael Eisenbach<br />

Weizmann Institute <strong>of</strong> Science<br />

PO Box 26<br />

Rehovot 76100<br />

Israel<br />

Phone: 972-8-9343923<br />

Fax: 972-8-9472722<br />

m.eisenbach@weizmann.ac.il<br />

Annette Erbse<br />

UC Boulder<br />

76 Chemistry<br />

Boulder, CO 80309<br />

Phone: (303) 492-3597<br />

erbse@colorado.edu<br />

Joseph Falke<br />

<strong>University</strong> <strong>of</strong> Colorado<br />

UCB 215<br />

Boulder, CO 80309-0215<br />

Phone: (303) 492-3503<br />

falke@colorado.edu<br />

Melanie Falord<br />

Institut Pasteur<br />

25 Rue du Dr. Roux<br />

75015 Paris<br />

France<br />

Phone: 33 1 44 38 94 87<br />

Fax: 33 1 45 68 89 38<br />

mfalord@pasteur.fr<br />

Hajime Fukuoka<br />

Tohoku <strong>University</strong><br />

Sendai Katahira Aoba-Ku<br />

980-8577<br />

Japan<br />

Phone: 81-22-217-5804<br />

Fax: 81-22-217-5804<br />

f-hajime@tagen.tohoku.ac.jp<br />

136<br />

Ana Gallego<br />

Instituto de Biotecnologia, UNAM<br />

Av Universidad 2001<br />

Cuernavaca 62210<br />

México<br />

Phone: 52 777 3291627<br />

analgh@ibt.unam.mx<br />

Elizabeth García-Gómez<br />

Universidad Nacional Autonoma de México,<br />

Instituto de Fisiología Celular<br />

Ciudad Universitaria, México D. F<br />

Ap. Postal 70-243.<br />

México 04510<br />

México<br />

Phone: (55)56225965<br />

egarcia@ifc.unam.mx<br />

Aldo García-Guerrero<br />

Universidad Nacional Autónoma de México<br />

Instituto de Fisiología Celuar<br />

Circuito Exterior s/n Cd. Universitaria<br />

México 04510<br />

México<br />

Phone: (5255) 56225618<br />

aldog.2602@gmail.com<br />

Mathieu Gauthier<br />

Universite Laval<br />

Pavillon Alexandre-Vachon<br />

Quebec QC G1K7P4<br />

Canada<br />

Phone: 418-656-2131<br />

mathieu.gauthier.5@ulaval.ca<br />

Meztlli Gaytán<br />

UNAM<br />

Ap. postal 70-243<br />

Ciudad Universitaria México D.F<br />

México 04510<br />

México<br />

Phone: +52 56225965<br />

ogaytan@ifc.unam.mx


Dimitris Georgellis<br />

UNAM<br />

Circuito exterior S/N<br />

México D.F. 04510<br />

México<br />

Phone: +52 55 5622 5738<br />

dimitris@ifc.unam.mx<br />

Zemer Gitai<br />

Princeton <strong>University</strong><br />

LTL-355<br />

Washington Rd<br />

Princeton, NJ 08540<br />

Phone: (609) 258-9420<br />

zgitai@princeton.edu<br />

George Glekas<br />

<strong>University</strong> <strong>of</strong> Illinois Urbana-Champaign<br />

409 Medical Sciences Building<br />

506 S. Mathews Ave.<br />

Ubana, IL 61801<br />

Phone: (217) 333-0268<br />

glekas@illinois.edu<br />

Shalom Goldberg<br />

<strong>University</strong> <strong>of</strong> Pennsylvania School <strong>of</strong><br />

Medicine<br />

422 Curie Blvd.<br />

Philadelphia, PA 19104<br />

Phone: (215) 898-3495<br />

shalom@mail.med.upenn.edu<br />

Juan Gonzalez<br />

<strong>University</strong> <strong>of</strong> Texas at Dallas<br />

RL11, 800 W. Campbell, Rd.<br />

Richardson, TX 75080<br />

Phone: (972) 883-2526<br />

jgonzal@utdallas.edu<br />

137<br />

Ricardo González<br />

UNAM<br />

Instituto de Fisiología Celular<br />

Ciudad Universitaria<br />

México 04510<br />

México<br />

Phone: 52 5556225738<br />

rgonzalez@ifc.unam.mx<br />

Bertha Gonzalez-Pedrajo<br />

National Autonomous <strong>University</strong> <strong>of</strong> México<br />

Apartado Postal 70-243<br />

México 04510<br />

México<br />

Phone: 5255 56225965<br />

Fax: 5255 56225611<br />

bpedrajo@ifc.unam.mx<br />

Nataliya Gurich<br />

<strong>University</strong> <strong>of</strong> Texas at Dallas<br />

RL11, 800 W. Campbell, Rd.<br />

Richardson, TX 75080<br />

Phone: (972) 883-6291<br />

nxg010400@utdallas.edu<br />

Benjamin Hall<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: +441865275380<br />

Fax: +441865275273<br />

benjamin.hall@bioch.ox.ac.uk<br />

Hua Han<br />

UMDNJ/CABM<br />

Rm 326, CABM<br />

679 Hoes Lane<br />

Piscataway, NJ 08854<br />

Phone: (732) 235-4206<br />

han@cabm.rutgers.edu


Rasika Harshey<br />

<strong>University</strong> <strong>of</strong> Texas at Austin<br />

1 <strong>University</strong> Station<br />

A1000<br />

Austin, TX 78712<br />

Phone: (512) 471-6881<br />

Fax: (512) 471-7088<br />

rasika@uts.cc.utexas.edu<br />

Caroline Harwood<br />

<strong>University</strong> <strong>of</strong> Washington<br />

Dept <strong>of</strong> Microbiology- Box 357242<br />

1705 NE Pacific Street<br />

Seattle, WA 98112<br />

Phone: (206) 221-2848<br />

csh5@u.washington.edu<br />

Fumio Hayashi<br />

Gunma <strong>University</strong><br />

Kiryu1-5-1 Tenjin<br />

376-8515<br />

Japan<br />

Phone: +81-277-30-1663<br />

Fax: +81-277-30-1663<br />

hayashi@chem-bio.gunma-u.ac.jp<br />

Gerald Hazelbauer<br />

<strong>University</strong> <strong>of</strong> Missouri<br />

117 Schweitzer Hall<br />

Columbia, MO 65211<br />

Phone: (573) 882-4845<br />

Fax: (573) 882-5635<br />

hazelbauerg@missouri.edu<br />

María Herrera Seitz<br />

Universidad Nacional de Mar del Plata<br />

Funes 3250<br />

Mar del Plata 7600<br />

Argentina<br />

Phone: 54 223 4753030<br />

Fax: 54 223 4753150<br />

khseitz@mdp.edu.ar<br />

138<br />

Penelope Higgs<br />

Max-Planck-Institute for<br />

Terrestrial Microbiology<br />

Karl-von-Frisch Strasse<br />

D35043 Marburg<br />

Germany<br />

Phone: +49 6421 178301<br />

Fax: +49 6421 178309<br />

higgs@mpi-marburg.mpg.de<br />

Yohei Hizukuri<br />

Nagoya <strong>University</strong><br />

NagoyaFuro-Cho, Chikusa-Ku<br />

464-8602<br />

Japan<br />

Phone: 81-52-789-3543<br />

Fax: 81-52-789-3001<br />

4hiz1103@bunshi3.bio.nagoya-u.ac.jp<br />

Shelley Horne<br />

North Dakota State <strong>University</strong><br />

NDSU Vet & MIcro Sci<br />

PO Box 6050 - Dept 7690<br />

Fargo, ND 58108<br />

Phone: (701) 231-6741<br />

Fax: (701) 231-9692<br />

sm.horne@ndsu.edu<br />

Basarab Hosu<br />

Harvard <strong>University</strong><br />

16 Divinity Avenue<br />

Bio Labs Buliding Rom 3068<br />

Cambridge, MA 02138<br />

Phone: (617) 495-6127<br />

Fax: (617) 496-1114<br />

ghosu@mcb.harvard.edu<br />

Colin Hughes<br />

Cambridge <strong>University</strong><br />

Dept <strong>of</strong> Pathology<br />

Cambridge<br />

CB4 3AH<br />

United Kingdom<br />

Phone: 44-1223-338538<br />

ch@mole.bio.cam.ac.uk


Tatsuya Ibuki<br />

Osaka <strong>University</strong><br />

suitayamadaoka 1-3<br />

565-0871<br />

Japan<br />

Phone: 81 06-6879-4625<br />

Fax: 81 06-6879-4652<br />

tatsuya@fbs.osaka-u.ac.jp<br />

Katsumi Imada<br />

Osaka <strong>University</strong><br />

Suita1-3 Yamadaoka<br />

565-0871<br />

Japan<br />

Phone: +81-6-6879-4625<br />

Fax: +81-6-6879-4652<br />

kimada@fbs.osaka-u.ac.jp<br />

Takehiko Inaba<br />

Hosei <strong>University</strong><br />

Koganei Midori-cho 3-11-15<br />

184-0003<br />

Japan<br />

Phone: +81-42-387-7173<br />

takehiko.inaba.13@k.hosei.ac.jp<br />

Yuichi Inoue<br />

Tohoku <strong>University</strong><br />

Sendai Katahira 2-1-1<br />

Aoba-ku, 980-8577<br />

Japan<br />

Phone: 81-22-217-5804<br />

inoue@tagen.tohoku.ac.jp<br />

Christine Josenhans<br />

Medical Uiniversity Hannover<br />

Carl-Neuberg-Strasse 1<br />

30625 Hannover<br />

Germany<br />

Phone: +495115324354<br />

Fax: +495115324354<br />

josenhans.christine@mh-hannover.de<br />

139<br />

Katy Juarez<br />

IBT-UNAM<br />

Av. Universidad 2001, Chamilpa<br />

Cuernavaca 62210<br />

México<br />

Phone: (52) 5556227240<br />

katy@ibt.unam.mx<br />

Jung Kwang-Hwan<br />

Sogang <strong>University</strong><br />

Mapo-Gu, Shinsu-Dong 1 R1203<br />

121-742 Seoul<br />

Republic <strong>of</strong> Korea<br />

Phone: 82-2-705-8795<br />

Fax: 82-2-704-3601<br />

kjung@sogang.ac.kr<br />

Alla Kaserer<br />

<strong>University</strong> <strong>of</strong> Oklahoma<br />

620 Parrington Oval<br />

Norman, OK 73019<br />

Phone: (405) 325-1532<br />

Fax: (405) 325-6111<br />

alla@ou.edu<br />

Linda Kenney<br />

Univ. <strong>of</strong> Illinois-Chicago<br />

835 S. Wolcott M/C 790<br />

Chicago, IL 60612<br />

Phone: (312) 413-0576<br />

Fax: (312) 996-6415<br />

kenneyl@uic.edu<br />

Tsuyoshi Kenri<br />

National Institute <strong>of</strong> Infectious Disease<br />

Musashimurayama4-7-1<br />

Gakuen, 208-0011<br />

Japan<br />

Phone: 81-42-561-0771<br />

Fax: 81-42-565-3315<br />

kenri@nih.go.jp


Cezar Khursigara<br />

National Institutes <strong>of</strong> Health<br />

50 South Drive<br />

Bldg 50/Rm 4306<br />

Bethesda, MD 20892<br />

Phone: (301) 594-2236<br />

khursigarac@mail.nih.gov<br />

John Kirby<br />

<strong>University</strong> <strong>of</strong> Iowa<br />

51 Newton Road<br />

Iowa City, IA 52242<br />

Phone: (319) 335-7818<br />

Fax: (319) 335-9006<br />

john-kirby@uiowa.edu<br />

Kimberly Kline<br />

Washington Univ School <strong>of</strong> Medicine<br />

Department <strong>of</strong> Molecular Microbiology<br />

Box 8230<br />

St. Louis, MO 63110<br />

Phone: (314) 266-0639<br />

k-kline@borcim.wustl.edu<br />

Masafumi Koike<br />

Nagoya <strong>University</strong><br />

Chikusa-ku<br />

Nagoya 464-8602<br />

Japan<br />

Phone: 81 0527892992<br />

Fax: 81 0527893001<br />

4koike@bunshi4.bio.nagoya-u.ac.jp<br />

Seiji Kojima<br />

Nagoya <strong>University</strong><br />

Furo-cho 1, Chikusa-ku<br />

Nagoya 464-8602<br />

Japan<br />

Phone: 81-52-789-2992<br />

Fax: 81-52-789-3001<br />

4seiji@bunshi4.bio.nagoya-u.ac.jp<br />

140<br />

Changhan Lee<br />

KAIST<br />

CA Daejeon 4101, BMRC, 373-1<br />

Gusung-dong, Yusung-gu 90504<br />

Republic <strong>of</strong> Korea<br />

Phone: +82-42-350-2669<br />

changhanlee@hotmail.com<br />

Jae-Min Lee<br />

<strong>University</strong> <strong>of</strong> Texas at Austin<br />

1 <strong>University</strong> Station A5000<br />

Austin, TX 78712-0162<br />

Phone: (512) 471-6799<br />

jaeminlee@mail.utexas.edu<br />

Yi-Ying Lee<br />

<strong>University</strong> <strong>of</strong> Illinoise at Chicago<br />

835 S Wolcott (M/C 790)<br />

Chicago, IL 60612<br />

Phone: (312) 413-0288<br />

yyinglee@uic.edu<br />

Jun Liu<br />

UT Houston Medical School<br />

6431 Fannin, MSB 2.228<br />

Houston, TX 77030<br />

Phone: (713) 500-5342<br />

jun.liu.1@uth.tmc.edu<br />

Janine Maddock<br />

<strong>University</strong> <strong>of</strong> Michigan<br />

830 North <strong>University</strong><br />

Ann Arbor, MI 48109<br />

Phone: (734)-936-8068<br />

maddock@umich.edu


Fumiaki Makino<br />

Osaka <strong>University</strong><br />

Suita1-3 Yamadaoka, Suita<br />

Osaka 565-0871<br />

Japan<br />

Phone: 81-06-6879-4625<br />

h1839@fbs.osaka-u.ac.jp<br />

Roxana Malpica<br />

<strong>University</strong> <strong>of</strong> Alberta<br />

CW405 Biological Sciences Building<br />

Edmonton AB T6G2E9<br />

Canada<br />

Phone: 780-492 4339<br />

malpica@ualberta.ca<br />

Mike Manson<br />

Texas A&M <strong>University</strong><br />

MS 3258<br />

BSBE 303<br />

College Station, TX 77843<br />

Phone: (979) 845-5158<br />

mike@mail.bio.tamu.edu<br />

Luary Martinez<br />

UNAM-Instituto de Biotecnologia<br />

Av. Universidad 2001, Colonia Chamilpa<br />

Cuernavaca 62210<br />

México<br />

Phone: 52-777 329 16 27<br />

luary@ibt.unam.mx<br />

Ana Martinez del Campo<br />

UNAM<br />

Instituto de Fisiologia Celular<br />

Circuito exterior s/n Cd. Universitaria<br />

México City 04510<br />

México<br />

Phone: (5255) 56225618<br />

acampo@ifc.unam.mx<br />

141<br />

Diego Massazza<br />

Universidad Nacional de Mar del Plata<br />

Funes 3250<br />

Mar del Plata 7600<br />

Argentina<br />

Phone: 54 223 4753030<br />

Fax: 54 223 4753150<br />

diegomassazza@hotmail.com<br />

Emilia Mauriello<br />

<strong>University</strong> <strong>of</strong> California, Berkeley<br />

31 Koshland Hall<br />

Berkeley, CA 94720<br />

Phone: (510) 643-5457<br />

emilia-mauriello@berkeley.edu<br />

Jonathan McMurry<br />

Kennesaw State <strong>University</strong><br />

1000 Chastain Rd. MB #1203<br />

Kennesaw, GA 30144<br />

Phone: (770) 499-3238<br />

jmcmurr1@kennesaw.edu<br />

Paul Milewski<br />

<strong>University</strong> <strong>of</strong> Wisconsin<br />

480 Lincoln Dr.<br />

Madison, WI 53706<br />

Phone: (608) 262-3220<br />

milewski@math.wisc.edu<br />

Kelly Miller<br />

West Virginia Univ., Health Sciences Center<br />

Box 9177, Room 2077<br />

Morgantown, WV 26506<br />

Phone: (304) 293-5959<br />

kmille49@mix.wvu.edu


Makoto Miyata<br />

Osaka City <strong>University</strong><br />

3-3-138 Sugimoto Sumiyoshi-ku<br />

Osaka 558-8585<br />

Japan<br />

Phone: +81(6)6605 3157<br />

Fax: +81(6)6605 3158<br />

miyata@sci.osaka-cu.ac.jp<br />

Md Motaleb<br />

East Carolina <strong>University</strong><br />

600 Moye Blvd<br />

BT 116<br />

Greenville, NC 27834<br />

Phone: (252) 744-3129<br />

Fax: (252) 744-3535<br />

motalebm@ecu.edu<br />

Tarek Msadek<br />

Institut Pasteur<br />

25 Rue du Dr. Roux<br />

75015 Paris<br />

France<br />

Phone: 33 1 45 68 88 09<br />

Fax: 33 1 45 68 89 38<br />

tmsadek@pasteur.fr<br />

Abishek Muralimohan<br />

<strong>University</strong> <strong>of</strong> Wisconsin - Madison<br />

433 Babcock Dr.<br />

Madison, WI 53706<br />

Phone: (608) 263-2636<br />

muralimohan@wisc.edu<br />

Shuichi Nakamura<br />

Osaka Univevrsity<br />

Suita1-3,Yamadaoka<br />

Osaka, 565-0871<br />

Japan<br />

Phone: +81-6-6879-4625<br />

naka@fbs.osaka-u.ac.jp<br />

142<br />

Daisuke Nakane<br />

Osaka City <strong>University</strong><br />

3-3-138 Sugimoto Sumiyoshi-ku<br />

Osaka 558-8585<br />

Japan<br />

Phone: +81 6 6605 3157<br />

Fax: +81 6 6605 3158<br />

nakane@sci.osaka-cu.ac.jp<br />

Beiyan Nan<br />

<strong>University</strong> <strong>of</strong> California, Berkeley<br />

Department <strong>of</strong> Molecular and Cell Biology<br />

Berkeley, CA 94720<br />

Phone: (510) 643-5457<br />

nanbeiyan@gmail.com<br />

Silke Neumann<br />

Ruprecht-Karls-Universität Heidelberg<br />

Im Neuenheimer Feld 282<br />

69120 Heidelberg<br />

Germany<br />

Phone: +49 6221546856<br />

s.neumann@zmbh.uni-heidelberg.de<br />

Vincent Nieto<br />

<strong>University</strong> <strong>of</strong> Texas at Austin<br />

2506 Speedway<br />

Austin, TX 78712<br />

Phone: (512) 471-6799<br />

vince.nieto@gmail.com<br />

So-ichiro Nishiyama<br />

Hosei <strong>University</strong><br />

3-7-2 Kajino-cho, Koganei<br />

Koganei 184-8584<br />

Japan<br />

Phone: +81-42-387-7173<br />

Fax: +81-42-387-7002<br />

soichiro.nishiyama.s3@k.hosei.ac.jp


Takahiro Nonaka<br />

Osaka City <strong>University</strong><br />

3-3-138 Sugimoto Sumiyoshi-ku<br />

Osaka 558-8585<br />

Japan<br />

Phone: +81-(0)6-6605-2575<br />

nonaka@sci.osaka-cu.ac.jp<br />

Luis Alberto Núñez Oreza<br />

Instituto de Fisiologia Celular - UNAM<br />

Ciudad Universitaria<br />

México D.F. 04510<br />

México<br />

Phone: +525556225738<br />

lanoreza@hotmail.com<br />

Ricardo Oropeza<br />

Instituto de Biotecnologia<br />

Universidad Nacional Autonoma de México<br />

Av. UNIVERSIDAD # 2001<br />

col. Chamilpa<br />

Cuernavaca 62210<br />

México<br />

Phone: 52-777-329-16-27<br />

oropeza@ibt.unam.mx<br />

Davi Ortega<br />

<strong>University</strong> <strong>of</strong> Tennessee / ORNL<br />

2856 Brock Av<br />

Knoxville, TN 37919<br />

Phone: (865) 384-9507<br />

dortega@utk.edu<br />

Karen Ottemann<br />

UC Santa Cruz<br />

1156 High Street<br />

METX<br />

Santa Cruz, CA 95064<br />

Phone: (831) 459-3482<br />

ottemann@ucsc.edu<br />

143<br />

Rebecca Parales<br />

<strong>University</strong> <strong>of</strong> California, Davis<br />

226 Briggs Hall<br />

1 Shields Ave.<br />

Davis, CA 95616<br />

Phone: (530) 754-5233<br />

Fax: (530) 752-9014<br />

reparales@ucdavis.edu<br />

Chankyu Park<br />

KAIST<br />

95014 CA Daejoen 4101, BMRC, KAIST<br />

373-1, Gusung-dong, Yusung-gu<br />

Republic <strong>of</strong> Korea<br />

Phone: +82-42-350-2669<br />

nasa@kaist.ac.kr<br />

John Parkinson<br />

<strong>University</strong> <strong>of</strong> <strong>Utah</strong><br />

257 South 1400 East<br />

Salt Lake City, UT 84112-0840<br />

Phone: (801) 581-7639<br />

parkinson@biology.utah.edu<br />

Jonathan Partridge<br />

<strong>University</strong> <strong>of</strong> Texas at Dallas<br />

800 W Campbell Road<br />

Richardson, TX 75080<br />

Phone: (972) 883-2519<br />

j.partridge@gmail.com<br />

Gabriela Peña-Sandoval<br />

Universidad Nacional Autonoma de México<br />

Circuito Exterior s/n<br />

Ciudad Universitaria, Copilco<br />

México City 04510<br />

México<br />

Phone: (52) 55 5622 5738<br />

gsandova@ifc.unam.mx


Abiola Pollard<br />

Cornell <strong>University</strong><br />

Chemistry Research Building<br />

G63 S.T. Olin Laboratory<br />

Ithaca, NY 14853<br />

Phone: (607) 255-4970<br />

amp65@cornell.edu<br />

Steven Porter<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: 01865 275298<br />

steven.porter@bioch.ox.ac.uk<br />

Birgit Prüß<br />

North Dakota State <strong>University</strong><br />

1523 Centennial Blvd.<br />

Fargo, ND 58108<br />

Phone: (701) 231-7848<br />

birgit.pruess@ndsu.edu<br />

Simon Rainville<br />

Laval <strong>University</strong><br />

Pavillon d'optique photonique<br />

2375, rue de la Terrasse<br />

Québec QC G1V 0A6<br />

Canada<br />

Phone: (418) 656-2131<br />

Fax: (418) 656-2623<br />

rainville@phy.ulaval.ca<br />

Everardo Ramírez<br />

Universidad Nacional Autonoma de México<br />

Cto. Exterior, Cd. Universitaria.<br />

Coyoacán, D. F.<br />

México, D. F. 04510<br />

México<br />

Phone: 525556225738<br />

evergrinch@yahoo.com.mx<br />

144<br />

Christopher Rao<br />

<strong>University</strong> <strong>of</strong> Illinois<br />

211 Roger Adams Lab<br />

Urbana, IL 61801<br />

Phone: (217) 244-2247<br />

chris@scs.uiuc.edu<br />

Sylvia Reimann<br />

Loyola <strong>University</strong> Chicago<br />

2160 S. First Ave<br />

Maguire Bldg 105, Rm 3822<br />

Maywood, IL 60153<br />

Phone: (708) 216-0845<br />

Fax: (708) 216-9574<br />

sreimann@lumc.edu<br />

S. James Remington<br />

<strong>University</strong> <strong>of</strong> Oregon<br />

Institute <strong>of</strong> Molecular Biology<br />

Eugene, OR 97403<br />

Phone: (541) 346-5190<br />

jremington@uoxray.uoregon.edu<br />

Peter Reuven<br />

Weizmann Institute <strong>of</strong> Science<br />

Ullman Bldg 11-a<br />

Rehovot 76100<br />

Israel<br />

Phone: 00972 8 934 2701<br />

Fax: 00972 8 934 4112<br />

peter.reuven@weizmann.ac.il<br />

Mark Roberts<br />

<strong>University</strong> <strong>of</strong> Oxford<br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: 44 1865 613367<br />

Fax: 44 1865 613338<br />

mark.roberts@bioch.ox.ac.uk


Claudia Rodríguez<br />

UNAM, Instituto de Fisiología Celular<br />

Ciudad Universitaria, Col Copilco<br />

Distrito Federal 04510<br />

México<br />

Phone: 52 01 55 56 22 57 38<br />

Fax: 52 01 55 56 22 56 11<br />

crangel@ifc.unam.mx<br />

Mariana Romo<br />

Instituto de Fisiología Celular, UNAM<br />

Circuito Externo S/N Col Copilco<br />

México City 04510<br />

México<br />

Phone: 52 55 62 25 965<br />

mromoc@ifc.unam.mx<br />

Kathleen Ryan<br />

UC Berkeley<br />

Plant & Microbial Biology<br />

251 Koshland Hall<br />

Berkeley, CA 94720<br />

Phone: (510) 643-9387<br />

Fax: (510) 642-4995<br />

kryan@nature.berkeley.edu<br />

Supreet Saini<br />

<strong>University</strong> <strong>of</strong> Illinois, Urbana Champaign<br />

114 Roger Adams Lab, Box C-3<br />

MC-712, 600 South Mathews Av<br />

Urbana, IL 61801<br />

Phone: (217) 244-7528<br />

saini3@uiuc.edu<br />

Griselda Salas<br />

Universidad Nacional Autónoma de México<br />

Cto. Exterior, Cd. Universitaria<br />

Coyoacán, D. F.<br />

México, D. F. 04510<br />

México<br />

Phone: 525556225738<br />

gsalas@ifc.unam.mx<br />

145<br />

Oscar Sánchez<br />

UNAM<br />

Cto. Ext. S/N Ciudad Universitaria<br />

México, D.F. 04510<br />

México<br />

Phone: 525556225738<br />

Fax: 525556225611<br />

oscars_69_2@hotmail.com<br />

Birgit Scharf<br />

Virginia Polytechnic Institute<br />

and State <strong>University</strong><br />

Life Science I<br />

Washington Street<br />

Blacksburg, VA 24061<br />

Phone: (617) 495-4217<br />

Fax: (617) 496-1114<br />

bscharf@vt.edu<br />

Florian Schubot<br />

Virginia Tech<br />

Life Science I, Room 125<br />

Washington Street<br />

Blacksburg, VA 24061<br />

Phone: (540) 231-2393<br />

fschubot@vt.edu<br />

Andrew Seely<br />

Texas A&M <strong>University</strong><br />

MS 3258 TAMU<br />

BSBE Room 303<br />

College Station, TX 77843<br />

Phone: (979) 845-1249<br />

aseely@mail.bio.tamu.edu<br />

Harold Shanafield<br />

<strong>University</strong> <strong>of</strong> Tennessee<br />

1414 West Cumberland<br />

F437<br />

Knoxville, TN 37996<br />

Phone: (865) 974-7687<br />

hshanafi@utk.edu


Thomas Shimizu<br />

Harvard <strong>University</strong><br />

16 Divinity Ave (BL3063)<br />

Cambridge, MA 02138<br />

Phone: (617) 495-4217<br />

Fax: (617) 496-1114<br />

tshimizu@mcb.harvard.edu<br />

Ruth Silversmith<br />

<strong>University</strong> <strong>of</strong> North Carolina<br />

Room 804 Mary Ellen Jones<br />

Chapel Hill, NC 27599<br />

Phone: (919) 966-2679<br />

silversr@med.unc.edu<br />

Julie Simons<br />

<strong>University</strong> <strong>of</strong> Wisconsin<br />

480 Lincoln Dr.<br />

Madison, WI 53706<br />

Phone: (608) 263-3239<br />

simons@math.wisc.edu<br />

Lotte Søgaard-Andersen<br />

Max Planck Institute<br />

for Terrestrial Microbiology<br />

Karl-von-Frisch Str.<br />

35043 Marburg<br />

Germany<br />

Phone: +49 6421 178 201<br />

Fax: +49 6421 178 209<br />

sogaard@mpi-marburg.mpg.de<br />

Erik Sommer<br />

<strong>University</strong> <strong>of</strong> Heidelberg<br />

Im Neuenheimer Feld 282<br />

69120 Heidelberg<br />

Germany<br />

Phone: +49 6221 546856<br />

e.sommer@zmbh.uni-heidelberg.de<br />

146<br />

Stephen Spiro<br />

<strong>University</strong> <strong>of</strong> Texas at Dallas<br />

800 W Campbell Road<br />

Richardson, TX 75080<br />

Phone: (972) 883-6896<br />

stephen.spiro@utdallas.edu<br />

Diane Stassi<br />

NIH<br />

6701 Rockledge Dr.<br />

Room 3202, MSC 7808<br />

Bethesda, MD 20892<br />

Phone: (301) 435-2514<br />

Fax: (301) 480-0940<br />

stassid@csr.nih.gov<br />

Ann Stock<br />

UMDNJ<br />

Robert Wood Johnson Medical School<br />

Center for Advanced Biotech. and Medicine<br />

679 Hoes Lane<br />

Piscataway, NJ 07976<br />

Phone: (732) 235-4844<br />

Fax: (732) 235-5289<br />

stock@cabm.rutgers.edu<br />

Claudia Studdert<br />

Universidad Nacional de Mar del Plata<br />

Funes 3250<br />

Mar del Plata 7600<br />

Argentina<br />

Phone: 54 223 4753030<br />

Fax: 54 223 4753150<br />

studdert@mdp.edu.ar<br />

Kalin Swain<br />

<strong>University</strong> <strong>of</strong> Colorado, Boulder<br />

Dept <strong>of</strong> Chemistry and Biochemistry<br />

Campus Box 215<br />

Boulder, CO 80309<br />

Phone: (303) 492-3592<br />

Fax: (303) 493-5894<br />

swain@colorado.edu


Hendrik Szurmant<br />

The Scripps Research Institute<br />

10550 N Torrey Pines Rd<br />

MEM-116<br />

La Jolla, CA 92037<br />

Phone: (858) 784-7904<br />

Fax: (858) 784-7966<br />

szurmant@scripps.edu<br />

Barry Taylor<br />

Loma Linda <strong>University</strong><br />

Alumni Hall<br />

Loma Linda, CA 92350<br />

Phone: (909) 558 4881<br />

Fax: (909) 558 4035<br />

bltaylor@llu.edu<br />

Kai Thormann<br />

Max Planck Institute<br />

For Terrestrial Microbiology<br />

Karl-von-Frisch-Strasse<br />

D-35043 Marburg<br />

Germany<br />

Phone: +49 6421 178 302<br />

Fax: +49 6421 178 309<br />

thormann@mpi-marburg.mpg.de<br />

Hoa Tran<br />

<strong>University</strong> <strong>of</strong> Massachusetts<br />

Morrill IV N<br />

639 N.Plesant St.<br />

Amherst, MA MA 01003<br />

Phone: (413) 545-9647<br />

htt@chem.umass.edu<br />

Yuhai Tu<br />

IBM Research<br />

1101 Kichawan Rd./Rt. 134<br />

Yorktown Heights, NY 10598<br />

Phone: (914) 945-2762<br />

yuhai@us.ibm.com<br />

147<br />

Alejandra Vazquez Ramos<br />

Universidad Nacional Autónoma de México<br />

Av. Universidad #2034 Col. Chamilpa<br />

Cuernavaca 62210<br />

México<br />

Phone: 52 777 329 1627<br />

Fax: 52 777 331 38673<br />

avazquez@ibt.unam.mx<br />

Juan-Jesus Vicente Ruiz<br />

<strong>University</strong> <strong>of</strong> California, Berkeley<br />

31 Koshland Hall<br />

Berkeley, CA 94720<br />

Phone: (510) 643-5457<br />

juanjesus.vicente@gmail.com<br />

Hera Vlamakis<br />

Harvard Medical School<br />

200 Longwood Ave.<br />

D1-219<br />

Boston, MA 02115<br />

Phone: (617) 432-4359<br />

hera_vlamakis@hms.harvard.edu<br />

George Wadhams<br />

Oxford <strong>University</strong><br />

South Parks Road<br />

Oxford<br />

OX1 3QU<br />

United Kingdom<br />

Phone: +44 01865 613329<br />

george.wadhams@bioch.ox.ac.uk<br />

Kylie Watts<br />

Loma Linda <strong>University</strong><br />

Dept <strong>of</strong> Microbiology & Mol Genetics<br />

AHBS 120<br />

Loma Linda, CA 92350<br />

Phone: (909) 558-1000<br />

Fax: (909) 558-4035<br />

kwatts@llu.edu


Ann West<br />

<strong>University</strong> <strong>of</strong> Oklahoma<br />

620 Parrington Oval<br />

Norman, OK 73019<br />

Phone: (405) 325-1529<br />

Fax: (405) 325-6111<br />

awest@ou.edu<br />

Jonathan Willett<br />

<strong>University</strong> <strong>of</strong> Iowa<br />

51 Newton Road<br />

Iowa City, IA 52242<br />

Phone: (319) 335-7938<br />

jonathan-willett@uiowa.edu<br />

Alan Wolfe<br />

Loyola <strong>University</strong> Chicago<br />

Maguire Center<br />

2160 South First Avenue<br />

Maywood, IL 60153<br />

Phone: (708) 216-5814<br />

Fax: (708) 216-9574<br />

awolfe@lumc.edu<br />

Gus Wright<br />

Texas A&M <strong>University</strong><br />

MS 3258 BSBE 303<br />

College Station, TX 77843<br />

Phone: (979) 845-1249<br />

gwright@mail.bio.tamu.edu<br />

Kang Wu<br />

<strong>University</strong> <strong>of</strong> Illinois, Urbana Champaign<br />

114 RAL Box C-3 (M/C 712)<br />

600 S Mathews Ave,<br />

Urbana, IL 61801<br />

Phone: (217) 244-7528<br />

kangwu@uiuc.edu<br />

148<br />

Zhaomin Yang<br />

Virginia Tech<br />

103 LS1 (Mailcode: 0910)<br />

Blacksburg, VA 24061<br />

Phone: (540) 231-1350<br />

zmyang@vt.edu<br />

Shinsuke Yoshimura<br />

Osaka <strong>University</strong><br />

Suita 3-#601, Yamadaoka 1-chome<br />

Osaka 565-0871<br />

Japan<br />

Phone: 81-06-6879-4625<br />

Fax: 81-06-6879-4652<br />

yoshimura@fbs.osaka-u.ac.jp<br />

Junhua Yuan<br />

Harvard <strong>University</strong><br />

16 Divinity Ave, BioLabs 3063<br />

Cambridge, MA 02138<br />

Phone: (617) 495-4217<br />

jyuan@mcb.harvard.edu<br />

David Zusman<br />

<strong>University</strong> <strong>of</strong> California<br />

16 Barker Hall #3204<br />

Berkeley, CA 94720-3204<br />

Phone: (510) 642-2293<br />

Fax: (510) 643-6334<br />

zusman@berkeley.edu


<strong>BLAST</strong> STAFF<br />

Tarra Bollinger<br />

Molecular Biology Consortium<br />

835 S. Wolcott (M/C 790)<br />

Chicago, IL 60612<br />

Phone: (312) 996-1216<br />

Fax: (312) 413-2952<br />

tbolli1@uic.edu<br />

149<br />

Peggy O'Neill<br />

Molecular Biology Consortium<br />

835 S. Wolcott (M/C 790)<br />

Chicago, IL 60612<br />

Phone: (312) 996-1216<br />

Fax: (312) 413-2952<br />

oneill@uic.edu


INDEX<br />

150


A<br />

Participant’s Name, Abstract page(s), Contact Information Page – PI Lab<br />

Adase, Christopher, 90, 130 - Manson, Michael<br />

Airola, Michael, 63, 130 - Crane, Brian<br />

Aldridge, Christine, 52, 130 - Aldridge, Phillip<br />

Aldridge, Phillip, 52, 53, 130 - Aldridge, Phillip<br />

Alexander, Roger, 67, 130 - Emonet, Thierry<br />

Alexandre, Gladys, 47, 54, 130 - Alexandre, Gladys<br />

Alvarez, Adrian, 69, 130 - Georgellis, Dimitris<br />

Andrade, Angel, 130 - González-Pedrajo, Bertha<br />

Armitage, Judy, 4, 55, 56, 57, 130 - Armitage, Judy<br />

B<br />

Ballado, Teresa, 130 - Dreyfus, Georges<br />

Barak, Rina, 132 - Eisenbach, Michael<br />

Barba Ostria, Carlos Arturo, 129, 132 - Georgellis, Dimitris<br />

Belas, Robert, 37, 132 - Belas, Robert<br />

Berleman, James, 12, 132 - Kirby, John<br />

Bhatnagar, Jaya, 34, 132 - Crane, Brian<br />

Bible, Amber, 47, 132 - Alexandre, Gladys<br />

Bisicchia, Paola, 7, 132 - Devine, Kevin<br />

Black, Wesley, 125, 132 - Yang, Zhaomin<br />

Bollinger, Tarra, 149 - Matsumura, Philip<br />

Bourret, Bob, 2, 132 - Bourret, Bob<br />

Branch, Richard, 20, 132 - Berry, Richard<br />

Briegel, Ariane, 36, 133 - Jensen, Grant<br />

Brown, Mostyn, 55, 133 - Armitage, Judith<br />

Bulyha, Iryna, 9, 133 - Søgaard-Andersen, Lotte<br />

Bustamante, Victor, 133 - Puente, Jose Luis<br />

C<br />

Calva, Edmundo, 38, 60, 61, 133 - Calva, Edmundo<br />

Camargo, Karen, 133 - Dreyfus, Georges<br />

Campbell, Asharie Johnson, 120, 133 - Taylor, Barry<br />

Campodonico, Eva, 133 - Zusman, David<br />

Cantwell, Brian, 126, 133 - Zhulin, Igor<br />

Carlson, C. Britt, 133 - Stock, Ann<br />

Castillo, David, 65, 134 - Dreyfus, Georges<br />

Chapman, Matt, 40, 134 - Chapman, Matt<br />

Chastanet, Arnaud, 50, 134 - Losick, Rich<br />

Che, Yong-Suk, 100, 134 - Namba, Keiichi<br />

Copeland, Matthew, 123, 134 - Weibel, Douglas<br />

Crane, Brian, 34, 63, 64, 134 - Crane, Brian<br />

Crosson, Sean, 6, 134 - Crosson, Sean<br />

Crowder, Rachel, 91, 134 - Manson, Michael<br />

D<br />

Dahlquist, Rick, 134 - Dahlquist, Rick<br />

151


Participant’s Name, Abstract page(s), Contact Information Page – PI Lab<br />

De la Cruz, Miguel, 61, 134 - Calva, Edmundo<br />

De la Mora, Javier, 66, 135 - Dreyfus, Georges<br />

Delalez, Nicolas, 56, 135 - Armitage, Judith<br />

Delauné, Aurelia, 98, 135 - Msadek, Tarek<br />

Devine, Kevin, 7, 135 - Devine, Kevin<br />

Diaz, Miguel, 135 - González-Pedrajo, Bertha<br />

Diaz-Mireles, Edith, 59, 135 - Bolam, David<br />

Dobkowski, Jason, 9, 135 - Maddock, Janine<br />

Draheim, Roger, 121, 135 - von Heijne, Gunnar<br />

Dreyfus, Georges, 41, 65, 66, 135 - Dreyfus, Georges<br />

Dubrac, Sarah, 8, 135 - Msadek, Tarek<br />

E<br />

Eisenbach, Michael, 21, 136 - Eisenbach, Michael<br />

Erbse, Annette, 35, 136 - Falke, Joseph<br />

F<br />

Falke, Joseph, 35, 68, 136 - Falke, Joseph<br />

Falord, Mélanie, 99, 136 - Msadek, Tarek<br />

Fukuoka, Hajime, 78, 136 - Ishijima, Akihiko<br />

G<br />

Gallego, Ana, 136 - Calva, Edmundo<br />

García-Gómez, Elizabeth, 72, 136 - González-Pedrajo, Bertha<br />

García-Guerrero, Aldo, 136 - Dreyfus, Georges<br />

Gauthier, Mathieu, 19, 136 - Rainville, Simon<br />

Gaytán, Meztlli, 136 - González-Pedrajo, Bertha<br />

Georgellis, Dimitris, 3, 69, 70, 137 - Georgellis, Dimitris<br />

Gitai, Zemer, 14, 137 - Gitai, Zemer<br />

Glekas, George, 29, 137 - Ordal, George<br />

Goldberg, Shalom, 45, 137 - DeGrado, William<br />

Gonzalez, Juan, 44, 71, 137 - Gonzalez, Juan<br />

González, Ricardo, 70, 137 - Georgellis, Dimitris<br />

González-Pedrajo, Bertha, 72, 137 - González-Pedrajo, Bertha<br />

Gurich, Nataliya, 71, 137 - Gonzalez, Juan<br />

H<br />

Hall, Benjamin, 113, 137 - Sansom, Mark<br />

Han, Hua, 117, 137 - Stock, Ann<br />

Harshey, Rasika, 73, 74, 138 - Harshey, Rasika<br />

Harwood, Caroline, 138 - Harwood, Carrie<br />

Hayashi, Fumio, 83, 84, 85, 138 - Oosawa, Kenji<br />

Hazelbauer, Gerald, 138 - Hazelbauer, Gerald<br />

Herrera Seitz, María, 115, 138 - Shingler, Victoria<br />

Higgs, Penelope, 10, 138 - Higgs, Penelope<br />

Hizukuri, Yohei, 75, 138 - Homma, Michio<br />

Horne, Shelley, 138 - Prüß, Birgit<br />

Hosu, Basarab, 138 - Berg, Howard<br />

152


Participant’s Name, Abstract page(s), Contact Information Page – PI Lab<br />

Hughes, Colin, 138 - Hughes, Colin<br />

I<br />

Ibuki, Tatsuya, 101, 139 - Namba, Keiichi<br />

Imada, Katsumi, 16, 139 - Namba, Keiichi<br />

Inaba, Takehiko, 81, 139 - Kawagishi, Ikuro<br />

Inoue, Yuichi, 79, 139 - Ishijima, Akihiko<br />

J<br />

Josenhans, Christine, 80, 139 - Josenhans, Christine<br />

Juarez, Katy, 139 - Juarez, Katy<br />

Jung, Kwang-Hwan, 46, 139 - Jung, Kwang -Hwan<br />

K<br />

Kaserer, Alla, 23, 139 - West, Ann<br />

Kenney, Linda, 139 - Kenney, Linda<br />

Kenri, Tsuyoshi, 119, 139 - Takahashi, Motohide<br />

Khursigara, Cezar, 32, 140 - Subrmaniam, Sriram<br />

Kirby, John, 12, 86, 87, 140 - Kirby, John<br />

Kline, Kimberly, 77, 140 - Hultgren, Scott<br />

Koike, Masafumi, 76, 140 - Homma, Michio<br />

Kojima, Seiji, 17, 140 - Homma, Michio<br />

L<br />

Lee, Changhan, 107, 140 - Park, Chankyu<br />

Lee, Jae-Min, 73, 140 - Harshey, Rasika<br />

Lee, Yi-Ying, 28, 93, 140 - Matsumura, Philip<br />

Liu, Jun, 88, 140 - Liu, Jun<br />

M<br />

Maddock, Janine, 89, 140 - Maddock, Janine<br />

Makino, Fumiaki, 102, 141 - Namba, Keiichi<br />

Malpica, Roxana, 109, 141 - Raivio, Tracy<br />

Manson, Michael, 31, 90, 91, 92, 141 - Manson, Michael<br />

Martínez del Campo, Ana, 41, 141 - Dreyfus, Georges<br />

Martinez, Luary, 27, 141 - Puente, Jose Luis<br />

Massazza, Diego, 118, 141 - Studdert, Claudia<br />

Mauriello, Emilia, 13, 141 - Zusman, David<br />

McMurry, Jonathan, 94, 141 - McMurry, Jonathan<br />

Milewski, Paul, 95, 141 - Milewski, Paul<br />

Miller, Kelly, 62, 141 - Charon, Nyles<br />

Miyata, Makoto, 96, 97, 142 - Miyata, Makoto<br />

Motaleb, Md, 42, 142 - Motaleb, Md<br />

Msadek, Tarek, 8, 98, 99, 142 - Msadek, Tarek<br />

Muralimohan, Abishek, 124, 142 - Weibel, Douglas<br />

N<br />

Nakamura, Shuichi, 103, 142 - Namba, Keiichi<br />

153


Participant’s Name, Abstract page(s), Contact Information Page – PI Lab<br />

Nakane, Daisuke, 96, 142 - Miyata, Makoto<br />

Nan, Beiyan, 128, 142 - Zusman, David<br />

Neumann, Silke, 49, 142 - Sourjik, Victor<br />

Nieto, Vincent, 74, 142 - Harshey, Rasika<br />

Nishiyama, So-ichiro, 82, 142 - Kawagishi, Ikuro<br />

Nonaka, Takahiro, 97, 143 - Miyata, Makoto<br />

Núñez Oreza, Luis Alberto, 143 - Georgellis, Dimitri<br />

O<br />

O'Neill, Peggy, 149 - Matsumura, Philip<br />

Oropeza, Ricardo, 38, 143 - Calva, Edmundo<br />

Ortega, Davi, 143 - Zhulin, Igor<br />

Ottemann, Karen, 105, 143 - Ottemann, Karen<br />

P<br />

Parales, Rebecca, 106, 143 - Parales, Rebecca<br />

Park, Chankyu, 107, 143 - Park, Chankyu<br />

Parkinson, John, 143 - Parkinson, John<br />

Partridge, John, 24, 143 - Spiro, Stephen<br />

Peña-Sandoval, Gabriela, 3, 143 - Georgellis, Dimitris<br />

Pollard, Abiola, 64, 144 - Crane, Brian<br />

Porter, Steven, 4, 144 - Armitage, Judith<br />

Prüß, Birgit, 43, 144 - Prüß, Birgit<br />

R<br />

Rainville, Simon, 19, 108, 144 - Rainville, Simon<br />

Ramírez, Everardo, 144 - Georgellis, Dimitris<br />

Rao, Christopher, 15, 110, 111, 144 - Rao, Christopher<br />

Reimann, Sylvia, 25, 144 - Wolfe, Alan<br />

Remington, James, 30, 144 - Remington, James<br />

Reuven, Peter, 21, 144 - Eisenbach, Michael<br />

Roberts, Mark, 57, 144 - Armitage, Judith<br />

Rodríguez, Claudia, 145 - Georgellis, Dimitris<br />

Romo, Mariana, 145 - González-Pedrajo, Bertha<br />

Ryan, Kathleen, 112, 145 - Ryan, Kathleen<br />

S<br />

Saini, Supreet, 110, 145 - Rao, Christopher<br />

Salas, Griselda, 145 - Georgellis, Dimitris<br />

Sánchez, Oscar, 145 - Georgellis, Dimitris<br />

Scharf, Birgit, 145 - Scharf, Birgit<br />

Schubot, Florian, 114, 145 - Schubot, Florian<br />

Seely, Andrew, 92, 145 - Manson, Michael<br />

Shanafield, Harold, 127, 145 - Zhulin, Igor<br />

Shimizu, Thomas, 48, 146 - Berg, Howard<br />

Silversmith, Ruth, 146 - Bourret, Robert<br />

Simons, Julie, 95, 146 - Milewski, Paul<br />

Søgaard-Andersen, Lotte, 9, 146 - Søgaard-Andersen, Lotte<br />

154


Participant’s Name, Abstract page(s), Contact Information Page – PI Lab<br />

Sommer, Erik, 116, 146 - Sourjik, Victor<br />

Spiro, Stephen, 24, 146 - Spiro, Stephen<br />

Stassi, Diane, 146 - Stassi, Diane<br />

Stock, Ann, 117, 146 - Stock, Ann<br />

Studdert, Claudia, 118, 146 - Studdert, Claudia<br />

Swain, Kalin, 68, 146 - Falke, Joseph<br />

Szurmant, Hendrik, 5, 147 - Szurmant, Hendrik<br />

T<br />

Taylor, Barry, 33, 120, 147 - Taylor, Barry<br />

Thormann, Kai, 18, 147 - Thormann, Kai<br />

Tran, Hoa, 26, 147 - Weis, Robert<br />

Tu, Yuhai, 22, 147 - Tu, Yuhai<br />

V<br />

Vazquez Ramos, Alejandra, 147 - Puente, Jose Luis<br />

Vicente Ruiz, Juan-Jesus, 147 - Zusman, David<br />

Vlamakis, Hera, 39, 147 - Kolter, Roberto<br />

W<br />

Wadhams, George, 147 - Wadhams, George<br />

Watts, Kylie, 33, 147 - Taylor, Barry<br />

West, Ann, 23, 148 - West, Ann<br />

Willett, Jonathan, 87, 148 - Kirby, John<br />

Wolfe, Alan, 25, 148 - Wolfe, Alan<br />

Wright, Gus, 31, 148 - Manson, Michael<br />

Wu, Kang, 111, 148 - Rao, Christopher<br />

Y<br />

Yang, Zhaomin, 11, 125, 148 - Yang, Zhaomin<br />

Yoshimura, Shinsuke, 104, 148 - Namba, Keiichi<br />

Yuan, Junhua, 58, 148 - Berg, Howard<br />

Z<br />

Zusman, David, 13, 128, 148 - Zusman, David<br />

155

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!