Download CHARACTERIZATION OF THE ROLE OF PSEUDOMONAS

Document related concepts

Hedgehog signaling pathway wikipedia , lookup

Flagellum wikipedia , lookup

Magnesium transporter wikipedia , lookup

Protein wikipedia , lookup

Cytokinesis wikipedia , lookup

Intrinsically disordered proteins wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Protein moonlighting wikipedia , lookup

Endomembrane system wikipedia , lookup

Signal transduction wikipedia , lookup

Type three secretion system wikipedia , lookup

Trimeric autotransporter adhesin wikipedia , lookup

List of types of proteins wikipedia , lookup

Transcript
CHARACTERIZATION OF THE ROLE OF PSEUDOMONAS SYRINGAE TYPE III
EFFECTOR HopAF1 IN VIRULENCE
Erica Jania Washington
A dissertation submitted to the faculty of the University of North Carolina at Chapel Hill in
partial fulfillment of the requirements for the degree of Doctor of Philosophy in the
Department of Biology.
Chapel Hill
2013
Approved by:
Jeff Dangl
Miriam Braunstein
Joseph Kieber
Jason Reed
Matt Wolfgang
© 2013
Erica Jania Washington
ALL RIGHTS RESERVED
ii ABSTRACT
Erica Jania Washington: Characterization of the role of Pseudomonas syringae type III
effector HopAF1 in virulence
(Under the direction of Jeff Dangl)
Many plant pathogens, including Pseudomonas syringae, encode the type III
secretion system for translocating effector proteins into the host during infection. Strains of
P. syringae which are not capable of delivering the type III effectors are nonpathogenic.
Therefore, the functions of type III effectors are essential for disease. Although several type
III effectors have been demonstrated to block components of the plant defense response,
the functions of most type III effectors are unknown. Our lab is interested in the type III
effector HopAF1, a type III effector that is present in eleven of the nineteen sequenced
strains of P. syringae and other plant pathogens. Although the presence of HopAF1 in
multiple strains of P. syringae suggests that it plays an important role in virulence, no
function has yet been associated with HopAF1. We generated a tertiary structural prediction
for HopAF1, which suggests that HopAF1 is structurally related to bacterial deamidases.
Deamidation, the irreversible substitution of an amide group with a carboxylate group, is the
mechanism by which several bacterial virulence factors manipulate the activity of a specific
substrate. To identify a potential target of HopAF1 activity, we employed a yeast two-hybrid
screen and identified Arabidopsis methylthioadenosine nucleosidase (MTN1) as a putative
target of HopAF1. This interaction, which we extended to include Arabidopsis MTN2, was
confirmed in planta using coimmunoprecipitations and bimolecular fluorescence
complementation assays. MTNs are enzymes in the Yang cycle, a cycle essential for high
iii levels of ethylene biosynthesis. Ethylene is a key plant hormone for plant developmental
processes, such as senescence and flowering. Ethylene is also induced during PTI.
Therefore, we hypothesized that HopAF1 inhibits PTI by manipulating MTNs and levels of
ethylene in plants. To this end, we used gas chromatography to measure ethylene
biosynthesis in plants treated with bacterially-delivered HopAF1. We determined that
HopAF1 inhibits ethylene biosynthesis in a manner dependent on putative catalytic residues.
Additionally, Yang cycle mutants mtn1 mtn2 and mtk are more susceptible to weak
pathogens. This data is consistent with the idea that HopAF1 targets a novel component of
the plant immune system, the Yang cycle.
iv This work is dedicated to my parents.
v ACKNOWLEDGEMENTS
I am grateful to many colleagues, friends, and family members for the support and
encouragement that have made this dissertation possible. It would be impossible to
adequately thank each individual; therefore I would like to specifically thank a few individuals
for their contributions to my graduate school career.
First, I would like to thank my advisor, Jeff Dangl, for his ongoing support and
enthusiasm. Jeff taught me how to fearlessly follow science where it takes you and in doing
so, always provided me with the resources necessary to help my project progress. In
addition, Jeff took the time to teach me how to think scientifically and critically. I also strongly
believe that without Jeff’s support during a critical time, I would not be able to write this
dissertation. Finally, I would like to thank Jeff for providing me a home in the Dangl-Grant
lab.
During my time in the Dangl-Grant lab I have worked with so many wonderful
individuals that taught me about science and life. I would like to thank Sarah Grant for the
time she invested in supporting me and my research. I would also like to thank Petra Epple
for her guidance, Terry Law for her friendship, Marc Nishimura for mentoring me when I first
joined the lab and Nuria Sanchez Coll for being a role model and a friend. I would also like
to thank Vera Bonardi, Tatiana Mucyn, and Derek Lundberg for many good times over the
years.
Additionally, I would like to thank the scientific community at UNC. I have received
great support from across the campus from professors and members of core facilities. I
specifically would like to thank Tony Perdue of the UNC Biology Microscopy Facility. In
vi addition to spending many hours at the microscope with me, Tony has always been
generous with his encouragement and support. Thank you to Ashutosh Tripathy of the UNC
Macromolecular Interactions Facility, Brenda Temple of the UNC Center for Structural
Biology, Lee Graves and John Sondek.
Thank you to my committee for your patience as I developed as a scientist. Thanks
to Miriam Braunstein and Matt Wolfgang for teaching me bacteriology in my first year at
UNC and for continuing to support me while on my committee. I would also like to Thank you
to Jason Reed for being thoughtful and constructive during committee meetings and journal
clubs. I would also like to thank Joe Kieber. I rotated in the Kieber lab and always felt like an
honorary member of the Kieber lab. The Kieber lab doors remained open to me when I
needed to use the gas chromatography machine to do some of my most important
experiments. Thank you to Gyeong Mee Yoon and Smadar Harpaz-Saad for being so
patient and kind while teaching me how to perform these experiments.
I would also like to thank Sharon Milgram, who recruited me to UNC through the
Interdisciplinary Biomedical Sciences program. Sharon was really invested in helping me
succeed as a person and a scientist. I particularly want to thank her for support in regards to
my participation in the Cell and Molecular Biology training grant program.
I would not be the person or the scientist I am today without my best friend and my
love, David Hubert. I want to thank David for his unconditional love and understanding.
Finally, I would like to thank my family. First, I want to thank my big brother, Derek,
for giving me brotherly advice from across the country and reminding me of what it means to
lead a well-balanced and successful life. My parents have given me the best parental
support imaginable during graduate school; in the form of expensive house repairs, cars and
every nonperishable good they could stock my home with. My parents have always taught
me to be confident and determined. But most important to my survival in graduate school,
they taught me the value of hard work.
vii TABLE OF CONTENTS
LIST OF TABLES ................................................................................................................... x LIST OF FIGURES ................................................................................................................ xi LIST OF ABBREVIATIONS.................................................................................................. xiii Chapter 1 Introduction........................................................................................................ 20 A two-tiered model of the plant innate immune system. ............................................ 21 The phytohormone ethylene plays multiple roles in the plant defense response. ..... 26 Successful pathogens suppress PTI /ETI activation and signaling. .......................... 28 The type III secretion system and its effectors are the main virulence
determinants of Pseudomonas syringae. .................................................................. 28 References ................................................................................................................ 39 Chapter 2 Bacterial virulence factors modulate eukaryotic host cell signaling
systems via deamidation.................................................................................................... 50 Preface ...................................................................................................................... 50 Abstract ..................................................................................................................... 50 Introduction................................................................................................................ 51 Cytotoxic Necrotizing Factor Toxins .......................................................................... 53 Burkholderia Lethal Factor 1 inhibits activity of translation factor eIF4A. .................. 56 Vibrio parahaemolyticus type III effector VopC deamidates small GTPases. ........... 57 Pasteurella multocida toxin ....................................................................................... 58 Cycle Inhibiting Factors ............................................................................................. 61 OspI inhibits host immune responses by deamidating an E2conjugating enzyme. ................................................................................................. 67 Figures ...................................................................................................................... 72 viii References ................................................................................................................ 81 Chapter 3 P. syringae type III effector HopAF1 suppresses plant immunity by
targeting the methionine recycling pathway .................................................................... 90 Preface ...................................................................................................................... 90 Abstract ..................................................................................................................... 91 Introduction................................................................................................................ 92 Results ...................................................................................................................... 95 Discussion ............................................................................................................... 105 Materials and Methods ............................................................................................ 109 Figures .................................................................................................................... 118 References .............................................................................................................. 135 Chapter 4 Conclusions and Future Directions ............................................................... 142 Using a shotgun approach and mass spectrometry to determine if HopAF1
modulates MTN function via deamidation. .............................................................. 143 Determine whether a functional deamidation site variant of MTN1 is impervious
to HopAF1 activity in planta.................................................................................... 145 Generate a HopAF1/MTN1 co-crystal to determine the molecular mechanism
behind HopAF1 inhibition of MTN activity. .............................................................. 147 Figures .................................................................................................................... 150 References .............................................................................................................. 155 ix LIST OF TABLES
Table 1.1 P. syringae type III effectors: defense suppression, localization,
activities and targets ............................................................................................................. 38
Table 2.1 Bacterial virulence factors that use deamidation to modify host proteins. ........... 80
x LIST OF FIGURES
Figure 1.1 P. syringae type III effectors target multiple host cellular components. .............. 37
Figure 2.1 A schematic representation of enzymatic deamidation in proteins. .................... 72
Figure 2.2 The catalytic domains of diverse deamidases. ................................................... 73
Figure 2.3 Crystal structure of the C-terminal region of Pasteurella multocida toxin. .......... 74
Figure 2.4 PMT deamidates and activates heterotrimeric Gα proteins. ............................... 75
Figure 2.5 Model of Cif inactivation of CRL activity and ubiquitination................................. 76
Figure 2.6 Crystal structures of cell cycle-inhibiting factors. ................................................ 77
Figure 2.7 Crystal structure of the CifYp deamidase/NEDD8 complex. ................................. 78
Figure 2.8 Crystal structures of unbound OspI and the OspI/Ubc13 complex. .................... 79
Figure 3.1 The predicted structure of HopAF1 exhibits structural homology to CheD, a
bacterial deamidase. ........................................................................................................... 118
Figure 3.2 Alignment of HopAF1 sequences from P. syringae pathovars and other plant
pathogens. .......................................................................................................................... 119
Figure 3.3 HopAF1 transgenic lines are susceptible to weak pathogens. .......................... 120
Figure 3.4 Arabidopsis transgenic lines expressing wild-type HopAF1 inhibit
flg22-mediated protection against P. syringae pv. tomato DC3000. ................................... 121
Figure 3.5 HR triggered by AvrXv3 is dependent on putative catalytic residues................ 122
Figure 3.6 Establishing a system to study localization of transiently
expressed HopAF1 in N. benthamiana. .............................................................................. 123
Figure 3.7 HopAF1 is targeted to the plasma membrane via acylation.............................. 124
Figure 3.8 The acylation-minus variant of HopAF1 co-localizes with free YFP,
a soluble protein. ................................................................................................................. 125
Figure 3.9 The Yang cycle contributes to ethylene biosynthesis, which is
suppressed by Pto DC3000. ............................................................................................... 126
Figure 3.10 HopAF1 associates with Arabidopsis MTN proteins at the
plasma membrane .............................................................................................................. 127
Figure 3.11 Establishing P. fluorescens as a tool to determine the effect of HopAF1 on
PAMP-induced ethylene accumulation. .............................................................................. 128
xi Figure 3.12 HopAF1 inhibits PAMP-induced ethylene biosynthesis. ................................. 129
Figure 3.13 Pfo-1 is unable to induce high levels of ethylene biosynthesis
in Yang cycle mutants, mtn1 mtn2 and mtk. ....................................................................... 130
Figure 3.14 HopAF1 domain architecture and circular dichroism spectra of HopAF1,
HopAF1H186A, and MTN1. .................................................................................................... 131
Figure 3.15 HopAF1 inhibits MTN1 activity in vitro in a manner dependent on catalytic
activity. ................................................................................................................................ 132
Figure 3.16 The Yang cycle is required for PTI in Arabidopsis. ......................................... 133
Figure 3.17 Proposed model of HopAF1 suppression of plant innate immunity. ............... 134
Figure 4.1 Alignment and structural localization of MTN1 conserved
putative deamidase target residues. ................................................................................... 150
Figure 4.2 MTN1 and MTN2 variants display loss-of-function phenotypes
when residues N113 and N194 are ‘deamidated’. .............................................................. 151
Figure 4.3 Circular dichroism spectra of MTN1 ‘deamidated’ variants does not
vary from wild-type MTN1. .................................................................................................. 152
Figure 4.4 Purification of wild-type MTN1. ......................................................................... 153
Figure 4.5 Purification of ∆130HopAF1. ............................................................................. 154
xii LIST OF ABBREVIATIONS
A/E
attaching and effacing
ABA
abscisic acid
ACS
aminocyclopropane carboxylate synthase
ADP
adenosine diphosphate
ADR1
activated disease resistance 1
AMP
adenosine monophosphate
APX
ascorbate peroxidase
ARF-GEF
ADP-ribosylation factor guanine nucleotide exchange factor
ATP
adenosine triphosphate
avr
avirulence
BAK1
bri1-associated receptor kinase 1
BCL-10
B-cell lymphoma 10
BiFC
bimolecular fluorescence complementation
BIK1
Botrytis-induced kinase 1
BKK1
BAK1-like kinase 1
BLF1
Burkholderia Lethal Factor 1
BRI1
brassinosteroid insensitive 1
C
caryboxyl
CaCo2
human colon carcinoma cell line
cAMP
cyclic adenosine monophosphate
CARD
caspase recruitment domain
CATERPILLER
CARD transcription enhancer, purine binding, pyrine, lots of leucine
repeats
CBL3
calcineurin B-like protein 3
CC
coiled-coil
xiii CD
circular dichroism
Cdc42
cell division control protein 42
CDK
cyclin dependent kinase
cDNA
complementary DNA
Cdt1
cdc10-dependent transcript 1
CeBIP
chitin elicitor binding protein
CERK1
chitin elicitor receptor kinase 1
cfu
colony forming units
CHBP
Cif homolog Burkholderia pseudomallei
CHPA
Cif homolog Photorhabdus asymbiotica
CHPL
Cif homolog Photorhabdus luminescens
CHYP
Cif homolog Yersinia pseudotuberculosis
Cif
cell inhibiting factor
CNF1
cytotoxic necrotizing factor 1
CNF2
cytotoxic necrotizing factor 2
CNF3
cytotoxic necrotizing factor 3
CNFY
cytotoxic necrotizing factor isolated from Yersinia
Col-0
Arabidopsis ecotype Col-0
COP9
constitutive photomorphogenesis
CRL
cullin ring ligases
CSN
COP9 signalosome
cYFP
carboxyl terminus of YFP
DAG
diacylglycerol
DNA
deoxyribonucleic acid
DNT
dermonecrotic toxin
DTT
dithiothreitol
xiv EDTA
ethylenediamine tetraacetic acid
EFR
elongation factor Tu receptor
EF-Tu
elongation factor Tu
EHEC
enterohemorrhagic E. coli
eIF4A
eukaryotic initiation factor 4A
EIL1
ethylene insensitive 3-like 1
EIL2
ethylene insensitive 3-like 2
EIN2
ethylene insensitive 2
EIN3
ethylene insensitive 3
elf18
18 amino acid peptide derived from elongation factor Tu
EPEC
enteropathogenic E. coli
ERF
ethylene response factor
ETI
effector-triggered immunity
ETS
effector-triggered susceptibility
Fla 216
Florida 216
Fla 7060
Florida 7060
flg22
22 amino acid peptide derived from flagellin
FLS2
flagellin sensing 2
GDP
guanosine diphosphate
GmHID1
2-hydroxyisoflavanone dehydratase from Glycine Max
GMP
guanosine monophosphate
GRP7
glycine-rich RNA binding protein 7
GRP8
glycine-rich RNA binding protein 8
GR-RBP
glycine-rich RNA binding protein
GTP
guanosine triphosphate
HA
hemaglutinin
xv HeLa
cell line derived from Henrietta Lacks
Hop
Hrp-dependent outer protein
HR
hypersensitivity response
hrc
HR and conserved genes
hrcC
HR and conserved protein C; structural type III secretion system
protein
hrp
HR and pathogenicity genes
HrpL
HR and pathogenicity protein L; alternative sigma factor
Hsp70
heat shock protein 70
Hsp90
heat shock protein 90
IPTG
isopropyl β-D-1-thiogalactopyranoside
JA
jasmonic acid
JAK/STAT
Janus kinase/ signal transducer and activator of transcription
kb
kilobase
LEE
locus of enterocyte effacement
LPS
lipopolysaccharide
LRR
leucine-rich repeat
LysM
lysin motif
MALTI
mucosa-associated-lymphoid-tissue-lymphoma-translocation gene 1
MAMP
microbe-associate molecular pattern
MAPK
mitogen-activated kinase
MBP
maltose binding protein
MES
2-(4-morpholino)ethanesulfonic acid
MTA
methylthioadenosine
MTK
methylthioribose kinase
MTN1
methylthioadenosine nucleosidase 1
MTN2
methylthioadenosine nucleosidase 2
xvi N
amino
NADPH
nicotinamide adenine dinucleotide phosphate
NB
nucleotide-binding
NEDD8
neural precursor cell expressed, developmentally down-regulated 8
NLR
nucleotide-binding domain and leucine-rich repeat-containing
NOD
nucleotide-binding oligomerization domain
nYFP
amino terminus of YFP
OD
optical density
ORF
open reading frame
PBL
PBS1-like protein
PBS1
AvrPphB susceptible 1
PDB
protein databank
PEPR1
Pep1 receptor 1
Pfo-1
Pseudomonas fluorescens
PI3K
phosphoinositide-3-kinase
PLC2
phospholipase C 2
PLCβ
phospholipase C β
PMT
Pasteurella multocida toxin
PR
pathogenesis-related
PRR
pattern recognition receptor
PsbQ
oxygen-evolving enhancer protein 3
PTI
PAMP-triggered immunity
Pto DC3000
Pseudomonas syringae pv. tomato DC3000
pv
pathovar
Rac
Ras-related C3 botulinum toxin substrate 1
RAR1
required for Mla-12 resistance
xvii RbohD
respiratory burst oxidase homolog D
Rbx
RING box protein 1
RhoA
Ras homology A
RIN4
RPM1-interacting protein 4
RING
really interesting new gene 1
RIPK
RPM1-induced protein kinase 1
RLK
receptor-like kinase
RLP
receptor-like protein
rmsd
root-mean-square deviation
RNA
ribonucleic acid
ROS
reactive oxygen species
rpm
revolutions per minute
RPM1
resistance to Pseudomonas syringae pv. maculicola 1
RPS2
resistance to Pseudomonas syringae pv. syringae 2
RPS5
resistance to Pseudomonas syringae pv. syringae 5
SA
salicylic acid
SID2
salicylic acid induction deficient 2
STAND
signal transduction ATPases with numerous domains
T3E
type III effector
T3SS
type III secretion system
TAB1
TAK1-binding protein 1
TAB2
TAK1-binding protein 2
TAK1
TGFβ-activated kinase 1
TEV
tobacco etch virus
TIR
Toll, interleukin-1 receptor
TRAF6
tumor-necrosis factor receptor-associated factor 6
xviii TTSS1
type III secretion system 1
TTSS2
type III secretion system 2
UBC13
ubiquitin-conjugating enzyme 13
UBL
ubiquitin-like protein
UEV1A
ubiquitin-conjugating enzyme variant 1a
WB
western blot
Xcv
Xanthomonas campestris pv. vesicatoria
YFP
yellow fluorescent protein
ZED1
HopZ-ETI deficient
xix Chapter 1
Introduction
Plants constantly encounter pathogens such as insects, oomycetes, viruses, fungi,
and bacteria in their natural habitats. However, many of these plant-microbe interactions do
not result in disease. Plant species may be resistant to microbes that are successful
pathogens on other plant species. This is called non-host resistance, and can be thought of
as reflecting situations where the microbe in question is not evolutionarily adapted to life on
that plant species (Senthil-Kumar and Mysore, 2013). Many mechanisms of non-host
resistance involve physical barriers that prevent the pathogen from accessing nutrients in
the plant apoplastic fluid. The waxy cuticle layer, the epidermis, and the plant cell wall are
structural barriers common in many plants. Bacterial pathogens may access the apoplastic
fluid through the stomatal openings in the plant leaf surface (Senthil-Kumar and Mysore,
2013). However, another mechanism of non-host resistance is the closure of stomata in the
presence of microbes (Melotto et al., 2006). Moreover, the presence of non-host microbes
also induces reinforcements in the cell wall in the form of callose, lignin, or suberin near
sites of infection (Senthil-Kumar and Mysore, 2013).
Once microbes reach the apoplastic fluid, antimicrobials prevent further colonization.
Non-host Arabidopsis apoplastic fluid extracts can restrict the growth of P. syringae
pathovars (Senthil-Kumar and Mysore, 2013). However, P. syringae pathovars that are
capable of causing disease in Arabidopsis are not inhibited by the apoplastic extracts (Fan
et al., 2011). Some antimicrobials are constitutively produced, while others are induced by
non-host microbes.
A two-tiered model of the plant innate immune system.
Adapted microbes successful at surpassing the passive non-host immunity detailed
above must overcome an active two-tiered plant immune system (Jones and Dangl, 2006).
The first stage is PTI or PAMP-triggered immunity. As a key virulence strategy, pathogens
have evolved effector molecules, such as type III effector (T3E) proteins in P. syringae and
RXLR effectors in oomycetes, that target PTI, resulting in disease and effector-triggered
susceptibility (ETS) (Jones and Dangl, 2006) (Table 1.1). Plants evolved the second tier of
the immunity termed ETI for effector-triggered immunity. In ETI, plants use NLR (nucleotidebinding domain and leucine-rich repeat-containing) proteins to detect effectors, often
triggering a rapid and localized cell death response called HR.
PAMP-triggered immunity
The first line of active defense begins with transmembrane pattern recognition
receptors (PRRs) recognizing conserved molecular patterns from pathogens called
pathogen-associated molecular patterns or PAMPs (Boller and Felix, 2009). Since these
microbial signatures are not limited to pathogenic microbes, they are perhaps more aptly
referred to as microbial-associated molecular patterns or MAMPs. For the purposes of this
introduction, we will use the terms PAMPs and PTI. PAMPs are often highly conserved
across a wide range of microbes, fulfill a function essential to the pathogen’s lifecycle, and
yet are not present in the host (Boller and Felix, 2009). Common bacterial PAMPs are
lipopolysaccharide (LPS), peptidoglycan, elongation factor Tu (EF-Tu), and flagellin (Boller
and Felix, 2009). Common fungal PAMPs include chitin and ergosterol, both derived from
cell walls of fungi (Boller and Felix, 2009). A 22 amino acid peptide derived from a
conserved region of the amino terminus of bacterial flagellin (flg22) and the first 18 amino
acids of EF-Tu (elf18) are derivatives of flagellin and EF-Tu that are sufficient for recognition
by host cell surface receptors (Felix et al., 1999; Kunze et al., 2004).
21 PAMP recognition occurs via pattern recognition receptors (PRRs) that are
analogous to Toll-like receptors in animals. PRRs are either receptor-like kinases (RLKs) or
receptor-like proteins (RLPs). RLKs contain an N-terminal extracellular domain, a single
transmembrane domain, and a C-terminal kinase domain. RLPs contain an N-terminal
extracellular domain, a single transmembrane domain, and a C-terminal short cytoplasmic
tail. Some well-studied PRRs include FLS2, EFR, and CERK1 (Monaghan and Zipfel, 2012).
FLS2 was the first PRR to be discovered and as a result, much of what we know about
PAMP recognition stems from FLS2 studies (Gomez-Gomez and Boller, 2000). FLS2 is the
RLK that recognizes flg22 (Gomez-Gomez and Boller, 2000). The C-terminal kinase domain
of FLS2 is a Ser/Thr kinase, except that its catalytic loop contains a CD sequence instead of
RD. For these reason, FLS2 is a non-RD kinase, which appears to be a hallmark of the
kinase domain of immune system related PRRs (Dardick and Ronald, 2006). EFR is a
Brassicaceae-specific RLK PRR that recognizes elf18 (Zipfel et al., 2006). CERK1, a RLK
that recognizes fungal chitin and bacterial peptidoglycan in Arabidopsis, has three
extracellular LysM domains for carbohydrate binding, along with the transmembrane domain
and the C-terminal kinase domain (Miya et al., 2007; Wan et al., 2008). In rice, chitin
recognition requires an additional RLP PRR called CeBiP (Kaku et al., 2006). CeBiP
interacts with rice CERK1 and this interaction is enhanced in the presence of chitin (Shimizu
et al., 2010).
PAMP recognition and downstream signal transduction involve oligomerization of
PRRs. BAK1 (BRI1-associated receptor kinase 1), a positive regulator of PTI signaling forms
ligand-induced complexes with FLS2 and EFR (Chinchilla et al., 2007; Heese et al., 2007;
Roux et al., 2011). This leads to the recruitment of several other RLKs, such as BIK1 and
BKK1, to form heterocomplexes with EFR and FLS2 (Roux et al., 2011). BAK1 has been
implicated in complex formation with multiple PRRs, such as PEPR1, an RLK that
recognizes endogenous peptides, after induction with their cognate MAMPs (Schulze et al.,
22 2010; Yamaguchi et al., 2006) Heterocomplex formation between the BAK1 and PRRs leads
to trans-phosphorylation and signal transduction (Schulze et al., 2010). Schulze et al.,
demonstrated that triggering downstream flg22-mediated responses requires BAK1 with a
functional kinase domain (Schulze et al., 2010). BAK1 is critical to PTI-related signaling
because bak1 mutants have severely reduced responses to flg22 and are completely
insensitive to elf18 (Chinchilla et al., 2007). BAK1 is also required for responses triggered by
other PAMPs such as LPS and peptidoglycan (Shan et al., 2000). However, not all PTI
signaling requires BAK1. For example, instead of interacting with BAK1, CERK1 undergoes
ligand-induced homo-dimerization leading to activation of downstream immune responses
(Liu et al., 2012).
Perception of PAMPs triggers PTI (Jones and Dangl, 2006). PTI is a critical
component of the two-tiered immune system because plants unable to trigger this response
are more susceptible to infection. For example, Arabidopsis and N. benthamiana plants
encoding a non-functional fls2 are more susceptible to bacterial infection (de Torres et al.,
2006; Hann and Rathjen, 2007; Li et al., 2005; Zipfel et al., 2004). Conversely, ectopic
expression of EFR in non-Brassicaceae plants Nicotiana benthamiana and tomato increases
their resistance to a range of bacterial pathogens (Lacombe et al., 2010).
The timing of PTI responses after PAMP recognition ranges from seconds to hours.
Early PTI responses include an influx in ions such as H+, K+, Cl- and Ca2+. Ca2+ activates
calcium-dependent protein kinases and RbohD (Ma and Berkowitz, 2007). RbohD is a
NADPH oxidase that contributes to reactive oxygen species (ROS) burst in Arabidopsis
(Ogasawara et al., 2008). This ROS burst occurs approximately 2 minutes after PAMP
recognition. ROS burst contributes to PTI by activating defense gene expression,
strengthening the cell wall, and initiating cell death (Torres et al., 2006). PTI responses also
include the activation of mitogen-activated kinases (MAPKs), which in turns leads to
activation of the defense-related WRKY transcription factors (Asai et al., 2002). This is only
23 a fraction of the massive reprogramming of gene expression that occurs after PAMP
recognition (Zipfel et al., 2006; Zipfel et al., 2004). Later PTI responses include the induction
of ethylene biosynthesis and callose deposition. Callose deposition, an easily measured
output and classic PTI response, consists of an accumulation of β-1,3-glucan polymers in
the plant cell wall near the sight of infection to form physical barriers against further
pathogen attack. We will address the role of ethylene further in the next section of this
introduction.
Effector-triggered immunity
The second line of defense is the gene-for-gene recognition of pathogen effectors by
plant disease resistance gene products. This line of defense evolved due to the ability of
virulent pathogens to inhibit PTI and thus render the plant susceptible to infection. This
condition is termed effector-triggered susceptibility or ETS (Jones and Dangl, 2006). Plants
evolved disease resistance (R) proteins to combat ETS. Disease resistance proteins are
often referred to as NB-LRR, or NLR, proteins due to their domain architecture, which
consists of a central nucleotide-binding (NB) domain and highly polymorphic C-terminal LRR
region. The N-terminal domains of NLRs commonly consist of either a TIR (Toll, interleukin1 receptor) domain or a CC (coiled-coil) domain. There are approximately 150 NLRs in the
Arabidopsis Col-0 genome (Meyers et al., 2003).
NLRs can detect effectors from a wide variety of plant pathogens including fungi,
oomycetes, viruses, and bacteria. NLR proteins may detect effectors directly or indirectly.
For example, AvrPita, an effector protein from Magnaporthe grisea, interacts directly with the
NLR Pi-ta from rice in yeast and in vitro (Jia et al., 2000). The indirect detection of an
effector by a NLR is more common. The ‘guard hypothesis’ describes the ability of a NLR to
detect a biochemical modification of a host protein by a pathogen effector (Dangl and Jones,
2001; Van der Biezen and Jones, 1998). In doing so, the host takes advantage of the
pathogen’s virulence strategy to initiate a successful immune response. The guard
24 hypothesis implies that multiple effectors modify the same host target in different ways and
that the host target can guarded by multiple NLRs (Jones and Dangl, 2006). This is because
ETI is a driving force of the host-pathogen arms race and is consistent with the finding that
pathogen effectors converge on a subset of host proteins (Mukhtar et al., 2011). A classic
example of the guard hypothesis that fulfills these tenets is the targeting of the Arabidopsis
protein RIN4 (RPM1-Interacting Protein 4) by multiple P. syringae T3Es. AvrRpt2-mediated
cleavage of RIN4 and phosphorylation of RIN4 in the presence of AvrRpm1 and AvrB,
trigger the activation of NLRs RPS2 and RPM1, respectively (Axtell and Staskawicz, 2003;
Mackey et al., 2003; Mackey et al., 2002).
The molecular mechanisms that govern NLR activation and signaling are not fully
understood and there are few generalizable themes describing NLR activation (Bonardi et
al., 2012; Bonardi and Dangl, 2012; Eitas and Dangl, 2010). This is likely due to their rapid
evolution in the arms race and the ability to detect virulence effectors from a diverse range
of plant pathogens. Plant NLRs are similar to CATERPILLER/NOD/LRR proteins and
STAND ATPases (Leipe et al., 2004; Ting and Davis, 2005). STAND ATPases are
molecular switches that cycle between an ADP-bound off state and an ATP-bound on state.
Similarly, NLRs contain a critical catalytic sequence, called the P loop, in their nucleotidebinding domain. The P loop domain is not required for the function of all NLRs. For example,
NLR activation, basal resistance to virulent pathogens, and salicylic acid accumulation
mediated by the ADR1 family of NLRs does not require P loop function (Bonardi et al.,
2011). Proper activation of NLRs requires a conformational change to release the
intramolecular interactions. Additionally, homotypic or heterotypic interactions or the
formation of large complexes may be required for NLR activation (Bonardi and Dangl, 2012).
ETI is a more rapid and higher amplitude form of PTI. The PTI outputs described
above such as ion influx, ROS burst, MAPK cascade signaling, and reprogramming of gene
expression also occur during ETI. However, a rapid and localized cell death at the site of
25 infection called the hypersensitivity response or HR often occurs once ETI is triggered
(Jones and Dangl, 2006). Although the exact mechanism by which ETI restricts pathogen
growth is unknown, mutation of R genes or in genes encoding chaperones required for NLR
stability results in an increase in susceptibility to pathogens.
The phytohormone ethylene plays multiple roles in the plant defense response.
There are three main hormones in plants that regulate the plant defense response
against abiotic and biotic stresses: salicylic acid (SA), jasmonic acid (JA), and ethylene. The
simplest model of how these hormones function states that ethylene and JA mediate
resistance to necrotrophs and herbivorous insects, while SA mediates resistance to
biotrophs (Robert-Seilaniantz et al., 2007). Although there are a few exceptions, the
ethylene and JA pathways are synergistic to each other and antagonistic to SA (Ecker and
Davis, 1987; Penninckx et al., 1998; Schenk et al., 2000). There is cross-talk between these
pathways to allow the plant to fine tune its defense response against the multiple invaders it
may encounter.
In recent years, there have been numerous examples of plant pathogens that hijack
specific hormone-regulated signaling pathways, supporting the hypothesis that
phytohormones are important components of the plant defense response. A well-studied
example is coronatine, a toxin from P. syringae pv. tomato DC3000 (hereafter Pto DC3000).
Coronatine triggers a hyper-induction of ethylene- and JA-mediated responses. This results
in the suppression of SA-dependent responses that are responsible for resistance against
biotrophs like P. syringae, leading to enhanced bacterial growth (Cui et al., 2005; LaurieBerry et al., 2006; Schroeder and Hilbi, 2008; Zhao et al., 2003). The P. syringae T3E HopI1
suppresses SA accumulation (Jelenska et al., 2007). Additionally, AvrPtoB stimulates ABA
(abscisic acid) biosynthesis and ABA responses that also antagonize SA biosynthesis and
SA-mediated defenses (de Torres-Zabala et al., 2007). AvrRpt2 alters the auxin physiology
to promote host susceptibility (Chen et al., 2007). XopD, a T3E from Xanthomonas
26 campestris pv. vesicatoria, (Xcv) was recently found to target a tomato ethylene response
factor (ERF) called SIERF4. SIERF4 mutants display reduced ethylene levels and an
increase in susceptibility to Xcv (Kim et al., 2013), supporting the hypothesis that ethylene is
involved in plant defense.
Ethylene is a gaseous plant hormone that is required for normal plant development
processes such as seed germination, leaf senescence, and leaf and petal abscission
(Kieber, 1997). As mentioned above, ethylene biosynthesis is induced during the plant
defense response. This occurs after recognition of PAMPs and the induction of the MAPK
signaling cascade. Phosphorylation stabilizes ACS (aminocyclopropane carboxylate
synthase) proteins and thus increases the biosynthesis of ethylene (Liu and Zhang, 2004).
There is a growing body of work focused on elucidating the role of ethylene in
defense against pathogens. Ethylene up-regulates the expression of pathogenesis-related
(PR) proteins and increases the accumulation of hydroxyproline-rich, cell-wall strengthening
proteins (Ecker and Davis, 1987; Penninckx et al., 1998). Ethylene also increases SAinduced expression of PR-1. This suggests that in addition to functioning synergistically with
JA, ethylene can also support SA-dependent defense responses (Lawton et al., 1994).
Furthermore, ethylene directly up-regulates the expression of FLS2. FLS2 transcription is
directly controlled by the binding of ERFs EIN3/EIL1/EIL2 to the FLS2 promoter and in
ethylene insensitive mutants (ein2), both FLS2 expression and FLS2 protein accumulation
are suppressed. As a result, ein2 plants are more susceptible to weakened strains of P.
syringae (Boutrot et al., 2010; Mersmann et al., 2010).
However, there is also data supporting the idea that pathogens up-regulate ethylene
as a part of their virulence strategy. For example, P. syringae T3Es AvrPto and AvrPtoB
promote increases in the biosynthesis of ethylene to promote disease symptoms in tomato
(Cohn and Martin, 2005). ET-responsive transcription factors EIN3 and EIL1 repress PAMPinduced genes, including the SA biosynthesis gene SID2 and the suppression of SA
27 promotes susceptibility to P. syringae (Chen et al., 2009). Furthermore, ethylene perception
is also necessary for manifestation of disease symptoms, following Pto DC3000 infection in
Arabidopsis (Bent et al., 1992). In conclusion, the role of ethylene in plant defense signaling
pathways is complex and needs more attention. Further experimentation is required for
complete understanding of how all phytohormones regulate defense pathways and how
pathogens promote disease by disrupting hormone homeostasis.
Successful pathogens suppress PTI /ETI activation and signaling.
Well-adapted pathogens have evolved methods to breach the multiple lines of plant
defense described above. Pathogens employ effectors to block components of PTI, ETI, or
both. Fungi and oomycetes encode hundreds of effectors as part of their virulence
strategies. However, little is known about their function and mechanisms by which they
thwart plant immune systems. For that reason, in this section, we will mainly focus on T3Es
from bacterial phytopathogens, mostly from Pseudomonas syringae. Bacterial pathogen
strains can each inject approximately 20 to 30 virulence effectors into the host cell during
infections, and strains across one widespread species, Pseudomonas syringae deploy
approximately 60 protein families, many of which have highly variable members (Baltrus et
al., 2011) (Figure 1.1; Table 1.1).
The type III secretion system and its effectors are the main virulence determinants of
Pseudomonas syringae.
Before detailing how T3Es target various defense machinery in the plant cell to
suppress immune system input signaling and outputs responses, it is important to digress to
first introduce the type III secretion system. This will be followed by a detailed discussion of
T3E functions, largely in the context of the plant pathogen P. syringae.
The type III secretion system
In order to cause disease, pathogenic bacteria interact with their host in a variety of
ways. These include using various secretion systems, toxins, and adhesions to exploit or
28 disrupt normal host processes and defense responses. One widespread and critical
bacterial virulence mechanism is the type III secretion system (T3SS), which consists of a
multi-ring base located in the bacterial inner and outer membranes and a protruding needlelike appendage. This needle-like pilus is called the Hrp pilus in phytopathogenic bacteria.
The Hrp pilus is used to deliver T3Es into the host cell to cause disease. The T3SS is
conserved across a broad range of Gram-negative pathogens including animal pathogens
such as Salmonella, Yersinia pestis, and Escherichia coli and phytopathogens
Pseudomonas syringae, Ralstonia solanacearum and various Xanthomonas species
(Chatterjee et al., 2013; Deslandes and Rivas, 2012; Raymond et al., 2013).
Arabidopsis-P. syringae model system
P. syringae causes disease in agronomically important crops such as tomatoes,
tobacco, and beans. P. syringae also infects Arabidopsis thaliana, a genetically tractable
model plant system (Whalen et al., 1991). Since P. syringae was also considered to be
genetically tractable, this established a new plant-pathogen system for studying molecular
mechanisms utilized by pathogens to causes disease progression and the subsequent host
defense responses.
P. syringae can survive in a surprisingly large number of environmental reservoirs
such as rivers, snow, and clouds (Morris et al., 2007; Morris et al., 2010; Morris et al., 2008).
Additionally, P. syringae can infect a wide range of plant hosts. The life cycle of P. syringae
begins with epiphytic growth on the surfaces of the aerial parts of plants, such as leaves and
flowers. Some strains of P. syringae, such as P. syringae pv. syringae B728A, can multiply
to high titers as an epiphyte, while others, such as Pto DC3000, must enter the apoplast to
multiply significantly (Feil et al., 2005). Given the proper environmental conditions, such as
during high humidity or heavy rain, P. syringae can enter the apoplastic region inside of
leaves via stomatal openings. In the apoplast, transcription of the T3SS genes, which are
encoded by the tightly linked hrp (hypersensitivity response and pathogenicity) and hrc
29 (hypersensitivity response and conserved) genes, is up-regulated. Although the specific in
planta signal that is responsible for up-regulating the hrp and hrc genes is unknown, Huynh
et al., demonstrated that fructose, sucrose, and mannitol upregulate AvrB transcription in
defined media (Huynh et al., 1989). Although P. syringae remains extracellular, the T3SS
translocates T3Es into the host cell to suppress components of PTI and ETI and contribute
to pathogenicity (Grant et al., 2006). P. syringae type III secretion system mutants are
nonpathogenic; therefore the type III secretion system and its effectors are required for the
pathogenicity of P. syringae (Lindgren et al., 1986). Furthermore, Lindgren et al., suggested
that these same effectors are responsible for HR triggered in non-host plants (Lindgren et
al., 1986).
Bacterial genomics, next generation sequencing, and definition of T3E families
The effort to identify T3E families in all strains of P. syringae is ongoing and has
involved many approaches over the years. Scoring robust effector loss-of-function
phenotypes using transposon mutagenesis screens is difficult because, due to functional
redundancy of sequence un-related T3Es, there is often only a weak, or even undetectable,
contribution of any single effector to virulence. Therefore, identification of T3Es has relied
heavily on bioinformatics-based methods. Research on the molecular mechanisms of P.
syringae T3Es has focused on the three ‘gold standard’ P. syringae strains that were
genome sequenced first; Pto DC3000, P. syringae pv. syringae B728A, and P. syringae pv.
phaseolicola 1448A (Buell et al., 2003; Feil et al., 2005; Joardar et al., 2005).
Type III effectors have several conserved features that allow for pattern-based
genome searches. Studies were performed that searched the promoters of type III effectors
for the hrp-box, a cis-acting element transcriptionally up-regulated by the alternative sigma
factor HrpL. Chang et al., published the results of a differential fluorescence induction
screen that identified type III effectors in Pto DC3000 and P. syringae pv. phaseolicola
30 1448A (Chang et al., 2005). This was followed by validation of the translocation of the
candidate effector proteins (Chang et al., 2005).
Next generation DNA sequencing has made it possible to quickly generate data on
multiple, diverse P. syringae genomes. Using a combination of Illumina and 454 platforms,
Baltrus et al., generated draft genome sequences for 14 previously unsequenced,
phylogenetically diverse strains of P. syringae (Baltrus et al., 2011). The search for type III
effectors among these genomes led to the identification of eight new type III effectors
families, resulting in a total of approximately 60 T3E families in this diverse collection of P.
syringae strains that expands the five distinct phylogroups (Baltrus et al., 2011).
Effectors that target PTI
Multiple effectors target RLKs at the plasma membrane and downstream
components of PTI signaling pathways. AvrPto and AvrPtoB share no primary or structural
homology, but they are redundant P. syringae T3Es that suppress the function of RLKs.
AvrPtoB has an N-terminal kinase domain that is sufficient to suppress flg22-induced PTI
(Gohre et al., 2008). The C-terminus of AvrPtoB is an E3 ubiquitin ligase domain that can
ubiquitinate and degrade RLKs FLS2 and CERK1 (Gimenez-Ibanez et al., 2009; Gohre et
al., 2008). AvrPto can bind and inhibit the activity of FLS2 and EFR (Xiang et al., 2008).
Although, it has been reported that AvrPto and AvrPtoB can inhibit BAK1 function, there is
conflicting evidence for this interaction (Shan et al., 2008; Xiang et al., 2011). AvrPphB
cleaves kinases, such as BIK1, that function downstream of PAMP recognition and integrate
signals from multiple PRRs to block PTI responses (Zhang et al., 2010). Interestingly,
AvrPphB cleavage of PBS1 triggers RPS5-mediated ETI in Arabidopsis inbreds that carry
the functional RPS5 NLR receptor gene (Ade et al., 2007; Shao et al., 2003). AvrAC, a T3E
from Xanthomonas campestris pv. campestris also inhibits PTI by targeting BIK1 and RIPK
(Feng et al., 2012). HopF2 and HopAI1 suppress PTI by inhibiting MAPK cascades. HopF2
ADP-ribosylates MKK5 and inhibits its kinase activity (Wang et al., 2010). HopAI1 uses
31 phosphothreonine lysase activity to inhibit the function of MPK3 and MPK6 (Zhang et al.,
2007). Similar to HopF2, disruption of PTI by targeting MAPK cascades triggers activation of
a NLR and ETI (Zhang et al., 2012).
Using effectors as probes of plant cell biology and plant defense has led to the
identification of components of the plant immune system that extend beyond PRRs and
NLRs. HopM1 targets vesicle trafficking by inhibiting the function of an Arabidopsis ARFGEF protein MIN7. HopM1 degradation of MIN7 is proteasome-dependent (Nomura et al.,
2006). Interestingly, when ETI is triggered by AvrRpt2, AvrPphB, and HopA1, HopM1mediated destabilization of MIN7 is inhibited (Nomura et al., 2011). This demonstrates a
novel situation in which ETI contributes to plant defense by preventing bacterial suppression
of a key component of PTI (Nomura et al., 2011). HopU1 is a P. syringae type III effector
mono-ADP-ribosyltransferase that targets Arabidopsis glycine-rich RNA binding proteins
(GR-RBPs) GRP7 and GRP8 (Fu et al., 2007). HopU1 negatively affects the ability of GRP7
to bind RNA (Jeong et al., 2011). Specifically, HopU1 blocks the interaction between GRP7
and FLS2 and EFR transcripts (Nicaise et al., 2013), leading to a reduction in FLS2
accumulation after infection with a virulent P. syringae strain expressing HopU1 (Nicaise et
al., 2013). This is consistent with HopU1 inhibiting PAMP responses triggered by flg22, such
as callose deposition (Fu et al., 2007). T3Es HopZ1a and HopZ1b have also been
implicated in PTI suppression. HopZ1a and HopZ1b degrade the soybean protein GmHID1
(2-hydroxyisoflavanone dehydratase), a protein involved in the biosynthesis of antimicrobial
phytoalexins (Zhou et al., 2011). Soybean plants with GmHID1 silenced are more
susceptible to infection (Zhou et al., 2011). HopZ1a also interacts with and acetylates
tubulin to inhibit formation of stable cytoskeletons (Lee et al., 2012). HopZ1a expression
blocks callose formation (Lee et al., 2012). Additionally, HopI1 localizes to the chloroplast
where it inhibits Hsp70 function, which leads to a decrease in SA and suppression of PTI
(Jelenska et al., 2010; Jelenska et al., 2007). Overexpression of HopN1 leads to a decrease
32 in PAMP-induced ROS burst and callose deposition (Rodriguez-Herva et al., 2012). Finally,
HopG1 also inhibits callose and ROS burst, although the molecular mechanism is currently
unknown (Block et al., 2010).
Effectors that block ETI
The first evidence of type III effectors blocking HR came from studies in the 1990s in
which Xanthomonas campestris pv. campestris hrp mutants induced a programmed cell
death in the vascular system of crucifers (Jackson et al., 1999). This suggested that there
are type III effectors in the hrp regulon that are blocking this HR-like response. More recent
studies have identified P. syringae type III effectors that are involved in suppressing ETI. For
example, Jackson et al. demonstrated that VirPphA, AvrPphC, and AvrPphF from P.
syringae pv. phaseolicola were able to inhibit NLR activation (Jackson et al., 1999). AvrPtoB
suppresses ETI-associated programmed cell death in tomato plants via ubiquitin ligase
activity in its C-terminus (Abramovitch et al., 2003). HopF2 targets RIN4 and blocks ETI
(Wilton et al., 2010). AvrAC can also block ETI (Feng et al., 2012). Furthermore, Jamir et al.,
showed that effectors could inhibit cell death triggered in yeast, suggesting those effectors
target programmed cell death pathways conserved in eukaryotes (Jamir et al., 2004).
Subcellular localizations and enzymatic functions of P. syringae type III effectors
As described above, T3Es can inhibit PTI and ETI and thus contribute to virulence.
Given the diversity of targets noted above for particular T3Es, it is not surprising to learn that
they function in various subcellular locations. Thus, determining the role of any T3E in
bacterial virulence requires multiple approaches and gathering pertinent forms of data, such
as the subcellular localization of the effector, the enzymatic function, and a potential host
target.
Determining the subcellular localization of a T3E in a plant cell can help determine
which host pathways or host defense mechanisms are targeted by the effectors. AvrRpm1
and AvrB were the first T3Es found to be targeted to the plasma membrane via the
33 eukaryotic-specific posttranslational modification of fatty acid acylation at the N-terminus
(Nimchuk et al., 2000). At the plasma membrane that AvrRpm1 and AvrB trigger the
phosphorylation of RIN4, leading to the elicitation of RPM1-mediated ETI. Myristoylation
sites of AvrRpm1 and AvrB are required to trigger ETI (Nimchuk et al., 2000). Since then
multiple type III effectors such have been shown to localize to the plasma membrane via
acylation and their functions are dependent on their proper targeting (Dowen et al., 2009;
Lewis et al., 2008; Robert-Seilaniantz et al., 2006; Shan et al., 2000).
The chloroplasts, mitochondria, and nucleus are all organelles targeted by T3Es.
HopI1 is targeted to the chloroplasts via a noncanonical targeting sequence (Jelenska et al.,
2007). HopI1 alters thylakoid structure, suppresses SA accumulation and SA-related
defenses and interacts with Hsp70 (Jelenska et al., 2010; Jelenska et al., 2007). HopN1
also is targeted to the chloroplast, where it degrades a photosystem II protein, PsbQ
(Rodriguez-Herva et al., 2012). HopG1 is targeted to the mitochondria and although its
specific protein target it currently unknown, HopG1 causes a reduction in PAMP-induced
callose deposition and ROS burst (Block et al., 2010). Although there are no published
examples of P. syringae T3Es being targeted directly to the nucleus, the AvrBs3 family of
T3Es from Xanthomonas spp. are targeted to the nucleus, where they act as transcriptional
activators to directly regulate host gene expression (Boch and Bonas, 2010).
Along with specific localizations, many T3Es have specific enzymatic activities used
to disrupt host processes and suppress PTI/ETI. Unfortunately, these diverse biochemical
functions have only recently started to be deciphered. Several type III effectors have been
shown to possess cysteine protease activity. AvrRpt2 cleaves RIN4, which activates the
RIN4-associated NB-LRR receptor, RPS2 (Kim et al., 2005). AvrPphB is a cysteine protease
that cleaves PBS1, which activates RPS5 (Shao et al., 2003). HopN1 is also a cysteine
protease that cleaves PsbQ (Lopez-Solanilla et al., 2004). There are other identified
enzymatic functions for T3Es. HopAO1 is a protein tyrosine phosphatase (Espinosa et al.,
34 2003). HopU1 is an ADP ribosyltransferase (Fu et al., 2007). AvrAC is a uridylyl transferase
from Xanthomonas campestris pv. campestris (Feng et al., 2012). HopX1 is predicted to be
a transglutaminase (Nimchuk et al., 2007). While the transglutaminase function has not
been shown, the putative catalytic residues are required for HopX1-triggered ETI in beans
(Nimchuk et al., 2007).
Assigning enzymatic functions to T3Es is challenging. T3Es have relatively low
conservation at the primary amino acid sequence level to proteins of known biochemical
function. This suggests that convergent evolution into conserved folds to modify host
signaling pathways. Solving crystal structures of T3Es is way to help determine a conserved
structure associated with a function and/or specific residues required for function (Boutemy
et al., 2011; Jeong et al., 2011; Lee et al., 2004; Singer et al., 2004). In lieu of an actual
crystal structure, generating tertiary structure predictions can also provide some information
about the function of a T3E. AvrRpm1 is predicted to have a fold similar to the catalytic
domain of poly-ADP-ribosyl polymerase and the central domains of HopQ1 is predicted to
have homology to nucleoside hydrolase enzymes (Cherkis et al., 2012; Li et al., 2013).
Additionally, SseI from Salmonella is predicted to have deamidase function based on
homology to the bacterial deamidase PMT. However, despite many efforts, SseI has not
been shown to have deamidation activity (Bhaskaran and Stebbins, 2012). Deamidation is
the irreversible conversion of glutamine and asparagines to glutamic acid and aspartic acid,
respectively. Since deamidation is emerging as a posttranslational modification utilized by
multiple bacterial toxins to modify host signaling pathways, we will explore bacterial
deamidases further in Chapter 2 (Washington et al., 2013).
Although the enzymatic activities of multiple T3E proteins have been discovered, the
link between enzymatic activity, the function of each effector’s host cellular target protein,
and the mechanism of suppression of both ETI and PTI often remain unclear. Unraveling
how T3Es use enzymatic activity to disrupt the function of host proteins will reveal details of
35 the molecular mechanisms by which these host proteins function normally. For some highly
conserved host proteins this may have implications beyond pathology that expand into the
realms of host cell signaling pathways and cell biology.
36 Figure 1.1. P. syringae type III effectors target multiple host cellular components.
Adapted from Xin and He, 2013. Bacterial PAMPs, such as flagellin, trigger PTI. P. syringae
injects type III effector proteins into the plant cell, via the type III secretion system, to block
components of PTI and ETI. AvrPto and AvrPtoB inhibit the activity of PRRs at the plasma
membrane through inhibition of PRR kinase activity or degradation, respectively. AvrPphB
cleaves BIK1, PBS1, and PBS1-like proteins (PBLs). HopF2 ADP-ribosylates MKK5 and
inhibits its activity. HopF2 may also inhibit PTI at the site of the plasma membrane by
blocking BIK1 phosphorylation. HopAI1 also targets the MAPK signaling cascade by using
phosphothreonine lyase activity to inhibit the function of MPK3 and MPK6. HopU1 is a
mono-ADP-ribosyltransferase that targets glycine-rich RNA-binding proteins (GR-RBPs).
HopU1 prevents GRP7 from interacting with FLS2 and EFR transcripts, leading to a
reduction in FLS2 transcript and protein accumulation. HopM1 targets vesicle trafficking by
inhibiting the function of the ARF-GEF protein MIN7. HopZ1a has been reported to have
multiple roles. HopZ1a degrades HID1, a protein involved in biosynthesis of antimicrobial
phytoalexins in soybean. HopZ1a also acetylates tubulin. Recently, HopZ1a has been
reported to target a pseudokinase, ZED1, which triggers ZAR1-mediated recognition in
Arabidopsis. Multiple effectors target RIN4. AvrRpt2 cleaves RIN4, triggering RPS2mediated ETI. HopF2 blocks ETI triggered by AvrRpt2 by interacting with RIN4.
Phosphorylation of RIN4 in the presence of AvrRpm1 and AvrB is mediated by RIPK. AvrPto
and AvrPtoB can degrade RIN4. Finally, AvrB targets RIN4 in complex with
MPK4/RAR1/Hsp90 to alter jasmonate signaling. HopI1is targeted to the chloroplast, where
it alters thylakoid structure, suppresses SA acculumation, and SA-related defenses. HopN1
is also targeted to the chloroplast. HopN1 degrades photosystem II protein, PsbQ. Although
HopG1 localizes to the mitochondria, its direct target and function remain unknown.
37 Table 1.1. P. syringae type III effectors: defense suppression, localization, activities and targets
Effector
AvrB
Host target(s)
Biochemical Activity
Subcellular
RIN4/RAR1/MPK4
?
Plasma membrane
PBS1/BIK1/PBLs
Cysteine protease
Plasma membrane
Pto/Fen/EFR/FLS2/
BAK1?
?
Plasma membrane
Pto/Fen/BAK1/FLS2/RLKs/BAK1?
E3 ubiquitin ligase
?
AvrRpm1
AvrRpt2
RIN4
Plasma membrane
RIN4
Poly-ADPib
lt
f
?
Cysteine protease
?
HopAI1
HopAO1
HopI1
MPK3/MPK6
?
Phosphothreonine lyase
Tyrosine phosphatase
?
?
Hsp70
J domain
Chloroplast
AvrPphB
AvrPto
AvrPtoB
38
HopF2
HopG1
HopM1
HopN1
HopU1
HopX1
HopZ1a/b
HopZ2
HopZ3
RIN4/MKK5
?
AtMIN7
PsbQ
GRP7, GRP8
?
GmHID1; ZED1
?
?
Mono-ADPribosyltransferase
?
?
Cysteine protease
Mono-ADPib
lt
f
Transglutaminase?
Acetyltransferase?/cysteine
protease?
Acetyltransferase?/cysteine
protease?
Acetyltransferase?/cysteine
protease?
Plasma membrane
Mitochondria
Vesicles
Chloroplast
Nucleo-cytoplasmic
?
Plasma membrane
Microtubules (a)
Plasma membrane
?
Key References
(Cui et al., 2010; Mackey et al., 2002;
Nimchuk et al., 2000)
(Ade et al., 2007; Dowen et al., 2009; Shao
et al., 2003; Zhang et al., 2010)
(Shan et al., 2000; Xiang et al., 2011; Xiang
et al., 2008)
(Gimenez-Ibanez et al., 2009; Gohre et al.,
2008; Shan et al., 2000; Xiang et al., 2011;
Xiang et al., 2008)
(Cherkis et al., 2012; Nimchuk et al., 2000)
(Axtell et al., 2003; Axtell et al., 2001; Axtell
and Staskawicz, 2003)
(Zhang et al., 2007)
(Espinosa et al., 2003)
(Espinosa et al., 2003; Jelenska et al.,
2010; Jelenska et al., 2007)
(Robert-Seilaniantz et al., 2006; Wang et
al., 2010; Wilton et al., 2010)
(Block et al., 2010)
(Nomura et al., 2006; Nomura et al., 2011)
(Rodriguez-Herva et al., 2012)
(Fu et al., 2007; Jeong et al., 2011)
(Nimchuk et al., 2007)
(Lee et al., 2012; Lewis et al., 2008; Lewis
et al., 2013; Zhou et al., 2011)
(Lee et al., 2012; Lewis et al., 2008; Zhou et
al., 2011)
(Lee et al., 2012; Lewis et al., 2008; Zhou et
al., 2011)
REFERENCES
Abramovitch, R.B., Kim, Y.J., Chen, S., Dickman, M.B., and Martin, G.B. (2003).
Pseudomonas type III effector AvrPtoB induces plant disease susceptibility by
inhibition of host programmed cell death. The EMBO journal 22, 60-69.
Ade, J., DeYoung, B.J., Golstein, C., and Innes, R.W. (2007). Indirect activation of a plant
nucleotide binding site-leucine-rich repeat protein by a bacterial protease.
Proceedings of the National Academy of Sciences of the United States of America
104, 2531-2536.
Asai, T., Tena, G., Plotnikova, J., Willmann, M.R., Chiu, W.L., Gomez-Gomez, L., Boller, T.,
Ausubel, F.M., and Sheen, J. (2002). MAP kinase signalling cascade in Arabidopsis
innate immunity. Nature 415, 977-983.
Axtell, M.J., Chisholm, S.T., Dahlbeck, D., and Staskawicz, B.J. (2003). Genetic and
molecular evidence that the Pseudomonas syringae type III effector protein AvrRpt2
is a cysteine protease. Molecular microbiology 49, 1537-1546.
Axtell, M.J., McNellis, T.W., Mudgett, M.B., Hsu, C.S., and Staskawicz, B.J. (2001).
Mutational analysis of the Arabidopsis RPS2 disease resistance gene and the
corresponding pseudomonas syringae avrRpt2 avirulence gene. Molecular plantmicrobe interactions : MPMI 14, 181-188.
Axtell, M.J., and Staskawicz, B.J. (2003). Initiation of RPS2-specified disease resistance in
Arabidopsis is coupled to the AvrRpt2-directed elimination of RIN4. Cell 112, 369377.
Baltrus, D.A., Nishimura, M.T., Romanchuk, A., Chang, J.H., Mukhtar, M.S., Cherkis, K.,
Roach, J., Grant, S.R., Jones, C.D., and Dangl, J.L. (2011). Dynamic evolution of
pathogenicity revealed by sequencing and comparative genomics of 19
Pseudomonas syringae isolates. PLoS pathogens 7, e1002132.
Bent, A.F., Innes, R.W., Ecker, J.R., and Staskawicz, B.J. (1992). Disease development in
ethylene-insensitive Arabidopsis thaliana infected with virulent and avirulent
Pseudomonas and Xanthomonas pathogens. Molecular plant-microbe interactions :
MPMI 5, 372-378.
Bhaskaran, S.S., and Stebbins, C.E. (2012). Structure of the catalytic domain of the
Salmonella virulence factor SseI. Acta crystallographica Section D, Biological
crystallography 68, 1613-1621.
Block, A., Guo, M., Li, G., Elowsky, C., Clemente, T.E., and Alfano, J.R. (2010). The
Pseudomonas syringae type III effector HopG1 targets mitochondria, alters plant
development and suppresses plant innate immunity. Cellular microbiology 12, 318330.
Boch, J., and Bonas, U. (2010). Xanthomonas AvrBs3 family-type III effectors: discovery
and function. Annual review of phytopathology 48, 419-436.
39 Boller, T., and Felix, G. (2009). A renaissance of elicitors: perception of microbe-associated
molecular patterns and danger signals by pattern-recognition receptors. Annual
review of plant biology 60, 379-406.
Bonardi, V., Cherkis, K., Nishimura, M.T., and Dangl, J.L. (2012). A new eye on NLR
proteins: focused on clarity or diffused by complexity? Current opinion in immunology
24, 41-50.
Bonardi, V., and Dangl, J.L. (2012). How complex are intracellular immune receptor
signaling complexes? Frontiers in plant science 3, 237.
Bonardi, V., Tang, S., Stallmann, A., Roberts, M., Cherkis, K., and Dangl, J.L. (2011).
Expanded functions for a family of plant intracellular immune receptors beyond
specific recognition of pathogen effectors. Proceedings of the National Academy of
Sciences of the United States of America 108, 16463-16468.
Boutemy, L.S., King, S.R., Win, J., Hughes, R.K., Clarke, T.A., Blumenschein, T.M.,
Kamoun, S., and Banfield, M.J. (2011). Structures of Phytophthora RXLR effector
proteins: a conserved but adaptable fold underpins functional diversity. The Journal
of biological chemistry 286, 35834-35842.
Boutrot, F., Segonzac, C., Chang, K.N., Qiao, H., Ecker, J.R., Zipfel, C., and Rathjen, J.P.
(2010). Direct transcriptional control of the Arabidopsis immune receptor FLS2 by the
ethylene-dependent transcription factors EIN3 and EIL1. Proceedings of the National
Academy of Sciences of the United States of America 107, 14502-14507.
Buell, C.R., Joardar, V., Lindeberg, M., Selengut, J., Paulsen, I.T., Gwinn, M.L., Dodson,
R.J., Deboy, R.T., Durkin, A.S., Kolonay, J.F., et al. (2003). The complete genome
sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv.
tomato DC3000. Proceedings of the National Academy of Sciences of the United
States of America 100, 10181-10186.
Chang, J.H., Urbach, J.M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R.,
Ausubel, F.M., and Dangl, J.L. (2005). A high-throughput, near-saturating screen for
type III effector genes from Pseudomonas syringae. Proceedings of the National
Academy of Sciences of the United States of America 102, 2549-2554.
Chatterjee, S., Chaudhury, S., McShan, A.C., Kaur, K., and De Guzman, R.N. (2013).
Structure and biophysics of type III secretion in bacteria. Biochemistry 52, 25082517.
Chen, H., Xue, L., Chintamanani, S., Germain, H., Lin, H., Cui, H., Cai, R., Zuo, J., Tang, X.,
Li, X., et al. (2009). ETHYLENE INSENSITIVE3 and ETHYLENE INSENSITIVE3LIKE1 repress SALICYLIC ACID INDUCTION DEFICIENT2 expression to negatively
regulate plant innate immunity in Arabidopsis. The Plant cell 21, 2527-2540.
Chen, Z., Agnew, J.L., Cohen, J.D., He, P., Shan, L., Sheen, J., and Kunkel, B.N. (2007).
Pseudomonas syringae type III effector AvrRpt2 alters Arabidopsis thaliana auxin
physiology. Proceedings of the National Academy of Sciences of the United States of
America 104, 20131-20136.
40 Cherkis, K.A., Temple, B.R., Chung, E.H., Sondek, J., and Dangl, J.L. (2012). AvrRpm1
missense mutations weakly activate RPS2-mediated immune response in
Arabidopsis thaliana. PloS one 7, e42633.
Chinchilla, D., Zipfel, C., Robatzek, S., Kemmerling, B., Nurnberger, T., Jones, J.D., Felix,
G., and Boller, T. (2007). A flagellin-induced complex of the receptor FLS2 and BAK1
initiates plant defence. Nature 448, 497-500.
Cohn, J.R., and Martin, G.B. (2005). Pseudomonas syringae pv. tomato type III effectors
AvrPto and AvrPtoB promote ethylene-dependent cell death in tomato. The Plant
journal : for cell and molecular biology 44, 139-154.
Cui, H., Wang, Y., Xue, L., Chu, J., Yan, C., Fu, J., Chen, M., Innes, R.W., and Zhou, J.M.
(2010). Pseudomonas syringae effector protein AvrB perturbs Arabidopsis hormone
signaling by activating MAP kinase 4. Cell host & microbe 7, 164-175.
Cui, J., Bahrami, A.K., Pringle, E.G., Hernandez-Guzman, G., Bender, C.L., Pierce, N.E.,
and Ausubel, F.M. (2005). Pseudomonas syringae manipulates systemic plant
defenses against pathogens and herbivores. Proceedings of the National Academy
of Sciences of the United States of America 102, 1791-1796.
Dangl, J.L., and Jones, J.D. (2001). Plant pathogens and integrated defence responses to
infection. Nature 411, 826-833.
Dardick, C., and Ronald, P. (2006). Plant and animal pathogen recognition receptors signal
through non-RD kinases. PLoS pathogens 2, e2.
de Torres-Zabala, M., Truman, W., Bennett, M.H., Lafforgue, G., Mansfield, J.W., Rodriguez
Egea, P., Bogre, L., and Grant, M. (2007). Pseudomonas syringae pv. tomato hijacks
the Arabidopsis abscisic acid signalling pathway to cause disease. The EMBO
journal 26, 1434-1443.
de Torres, M., Mansfield, J.W., Grabov, N., Brown, I.R., Ammouneh, H., Tsiamis, G.,
Forsyth, A., Robatzek, S., Grant, M., and Boch, J. (2006). Pseudomonas syringae
effector AvrPtoB suppresses basal defence in Arabidopsis. The Plant journal : for cell
and molecular biology 47, 368-382.
Deslandes, L., and Rivas, S. (2012). Catch me if you can: bacterial effectors and plant
targets. Trends in plant science 17, 644-655.
Dowen, R.H., Engel, J.L., Shao, F., Ecker, J.R., and Dixon, J.E. (2009). A family of bacterial
cysteine protease type III effectors utilizes acylation-dependent and -independent
strategies to localize to plasma membranes. The Journal of biological chemistry 284,
15867-15879.
Ecker, J.R., and Davis, R.W. (1987). Plant defense genes are regulated by ethylene.
Proceedings of the National Academy of Sciences of the United States of America
84, 5202-5206.
Eitas, T.K., and Dangl, J.L. (2010). NB-LRR proteins: pairs, pieces, perception, partners,
and pathways. Current opinion in plant biology 13, 472-477.
41 Espinosa, A., Guo, M., Tam, V.C., Fu, Z.Q., and Alfano, J.R. (2003). The Pseudomonas
syringae type III-secreted protein HopPtoD2 possesses protein tyrosine phosphatase
activity and suppresses programmed cell death in plants. Molecular microbiology 49,
377-387.
Fan, J., Crooks, C., Creissen, G., Hill, L., Fairhurst, S., Doerner, P., and Lamb, C. (2011).
Pseudomonas sax genes overcome aliphatic isothiocyanate-mediated non-host
resistance in Arabidopsis. Science 331, 1185-1188.
Feil, H., Feil, W.S., Chain, P., Larimer, F., DiBartolo, G., Copeland, A., Lykidis, A., Trong, S.,
Nolan, M., Goltsman, E., et al. (2005). Comparison of the complete genome
sequences of Pseudomonas syringae pv. syringae B728a and pv. tomato DC3000.
Proceedings of the National Academy of Sciences of the United States of America
102, 11064-11069.
Felix, G., Duran, J.D., Volko, S., and Boller, T. (1999). Plants have a sensitive perception
system for the most conserved domain of bacterial flagellin. The Plant journal : for
cell and molecular biology 18, 265-276.
Feng, F., Yang, F., Rong, W., Wu, X., Zhang, J., Chen, S., He, C., and Zhou, J.M. (2012). A
Xanthomonas uridine 5'-monophosphate transferase inhibits plant immune kinases.
Nature 485, 114-118.
Fu, Z.Q., Guo, M., Jeong, B.R., Tian, F., Elthon, T.E., Cerny, R.L., Staiger, D., and Alfano,
J.R. (2007). A type III effector ADP-ribosylates RNA-binding proteins and quells plant
immunity. Nature 447, 284-288.
Gimenez-Ibanez, S., Hann, D.R., Ntoukakis, V., Petutschnig, E., Lipka, V., and Rathjen, J.P.
(2009). AvrPtoB targets the LysM receptor kinase CERK1 to promote bacterial
virulence on plants. Current biology : CB 19, 423-429.
Gohre, V., Spallek, T., Haweker, H., Mersmann, S., Mentzel, T., Boller, T., de Torres, M.,
Mansfield, J.W., and Robatzek, S. (2008). Plant pattern-recognition receptor FLS2 is
directed for degradation by the bacterial ubiquitin ligase AvrPtoB. Current biology :
CB 18, 1824-1832.
Gomez-Gomez, L., and Boller, T. (2000). FLS2: an LRR receptor-like kinase involved in the
perception of the bacterial elicitor flagellin in Arabidopsis. Molecular cell 5, 10031011.
Grant, S.R., Fisher, E.J., Chang, J.H., Mole, B.M., and Dangl, J.L. (2006). Subterfuge and
manipulation: type III effector proteins of phytopathogenic bacteria. Annual review of
microbiology 60, 425-449.
Hann, D.R., and Rathjen, J.P. (2007). Early events in the pathogenicity of Pseudomonas
syringae on Nicotiana benthamiana. The Plant journal : for cell and molecular biology
49, 607-618.
Heese, A., Hann, D.R., Gimenez-Ibanez, S., Jones, A.M., He, K., Li, J., Schroeder, J.I.,
Peck, S.C., and Rathjen, J.P. (2007). The receptor-like kinase SERK3/BAK1 is a
42 central regulator of innate immunity in plants. Proceedings of the National Academy
of Sciences of the United States of America 104, 12217-12222.
Huynh, T.V., Dahlbeck, D., and Staskawicz, B.J. (1989). Bacterial blight of soybean:
regulation of a pathogen gene determining host cultivar specificity. Science 245,
1374-1377.
Jackson, R.W., Athanassopoulos, E., Tsiamis, G., Mansfield, J.W., Sesma, A., Arnold, D.L.,
Gibbon, M.J., Murillo, J., Taylor, J.D., and Vivian, A. (1999). Identification of a
pathogenicity island, which contains genes for virulence and avirulence, on a large
native plasmid in the bean pathogen Pseudomonas syringae pathovar phaseolicola.
Proceedings of the National Academy of Sciences of the United States of America
96, 10875-10880.
Jamir, Y., Guo, M., Oh, H.S., Petnicki-Ocwieja, T., Chen, S., Tang, X., Dickman, M.B.,
Collmer, A., and Alfano, J.R. (2004). Identification of Pseudomonas syringae type III
effectors that can suppress programmed cell death in plants and yeast. The Plant
journal : for cell and molecular biology 37, 554-565.
Jelenska, J., van Hal, J.A., and Greenberg, J.T. (2010). Pseudomonas syringae hijacks
plant stress chaperone machinery for virulence. Proceedings of the National
Academy of Sciences of the United States of America 107, 13177-13182.
Jelenska, J., Yao, N., Vinatzer, B.A., Wright, C.M., Brodsky, J.L., and Greenberg, J.T.
(2007). A J domain virulence effector of Pseudomonas syringae remodels host
chloroplasts and suppresses defenses. Current biology : CB 17, 499-508.
Jeong, B.R., Lin, Y., Joe, A., Guo, M., Korneli, C., Yang, H., Wang, P., Yu, M., Cerny, R.L.,
Staiger, D., et al. (2011). Structure function analysis of an ADP-ribosyltransferase
type III effector and its RNA-binding target in plant immunity. The Journal of
biological chemistry 286, 43272-43281.
Jia, Y., McAdams, S.A., Bryan, G.T., Hershey, H.P., and Valent, B. (2000). Direct interaction
of resistance gene and avirulence gene products confers rice blast resistance. The
EMBO journal 19, 4004-4014.
Joardar, V., Lindeberg, M., Jackson, R.W., Selengut, J., Dodson, R., Brinkac, L.M.,
Daugherty, S.C., Deboy, R., Durkin, A.S., Giglio, M.G., et al. (2005). Whole-genome
sequence analysis of Pseudomonas syringae pv. phaseolicola 1448A reveals
divergence among pathovars in genes involved in virulence and transposition.
Journal of bacteriology 187, 6488-6498.
Jones, J.D., and Dangl, J.L. (2006). The plant immune system. Nature 444, 323-329.
Kaku, H., Nishizawa, Y., Ishii-Minami, N., Akimoto-Tomiyama, C., Dohmae, N., Takio, K.,
Minami, E., and Shibuya, N. (2006). Plant cells recognize chitin fragments for
defense signaling through a plasma membrane receptor. Proceedings of the National
Academy of Sciences of the United States of America 103, 11086-11091.
Kieber, J.J. (1997). The ethylene response pathway in Arabidopsis. Annual review of plant
physiology and plant molecular biology 48, 277-296.
43 Kim, H.S., Desveaux, D., Singer, A.U., Patel, P., Sondek, J., and Dangl, J.L. (2005). The
Pseudomonas syringae effector AvrRpt2 cleaves its C-terminally acylated target,
RIN4, from Arabidopsis membranes to block RPM1 activation. Proceedings of the
National Academy of Sciences of the United States of America 102, 6496-6501.
Kim, J.G., Stork, W., and Mudgett, M.B. (2013). Xanthomonas type III effector XopD
desumoylates tomato transcription factor SlERF4 to suppress ethylene responses
and promote pathogen growth. Cell host & microbe 13, 143-154.
Kunze, G., Zipfel, C., Robatzek, S., Niehaus, K., Boller, T., and Felix, G. (2004). The N
terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis
plants. The Plant cell 16, 3496-3507.
Lacombe, S., Rougon-Cardoso, A., Sherwood, E., Peeters, N., Dahlbeck, D., van Esse,
H.P., Smoker, M., Rallapalli, G., Thomma, B.P., Staskawicz, B., et al. (2010).
Interfamily transfer of a plant pattern-recognition receptor confers broad-spectrum
bacterial resistance. Nature biotechnology 28, 365-369.
Laurie-Berry, N., Joardar, V., Street, I.H., and Kunkel, B.N. (2006). The Arabidopsis thaliana
JASMONATE INSENSITIVE 1 gene is required for suppression of salicylic aciddependent defenses during infection by Pseudomonas syringae. Molecular plantmicrobe interactions : MPMI 19, 789-800.
Lawton, K.A., Potter, S.L., Uknes, S., and Ryals, J. (1994). Acquired Resistance Signal
Transduction in Arabidopsis Is Ethylene Independent. The Plant cell 6, 581-588.
Lee, A.H., Hurley, B., Felsensteiner, C., Yea, C., Ckurshumova, W., Bartetzko, V., Wang,
P.W., Quach, V., Lewis, J.D., Liu, Y.C., et al. (2012). A bacterial acetyltransferase
destroys plant microtubule networks and blocks secretion. PLoS pathogens 8,
e1002523.
Lee, C.C., Wood, M.D., Ng, K., Andersen, C.B., Liu, Y., Luginbuhl, P., Spraggon, G., and
Katagiri, F. (2004). Crystal structure of the type III effector AvrB from Pseudomonas
syringae. Structure 12, 487-494.
Leipe, D.D., Koonin, E.V., and Aravind, L. (2004). STAND, a class of P-loop NTPases
including animal and plant regulators of programmed cell death: multiple, complex
domain architectures, unusual phyletic patterns, and evolution by horizontal gene
transfer. Journal of molecular biology 343, 1-28.
Lewis, J.D., Abada, W., Ma, W., Guttman, D.S., and Desveaux, D. (2008). The HopZ family
of Pseudomonas syringae type III effectors require myristoylation for virulence and
avirulence functions in Arabidopsis thaliana. Journal of bacteriology 190, 2880-2891.
Lewis, J.D., Lee, A.H., Hassan, J.A., Wan, J., Hurley, B., Jhingree, J.R., Wang, P.W., Lo, T.,
Youn, J.Y., Guttman, D.S., et al. (2013). The Arabidopsis ZED1 pseudokinase is
required for ZAR1-mediated immunity induced by the Pseudomonas syringae type III
effector HopZ1a. Proceedings of the National Academy of Sciences of the United
States of America.
44 Lewis, J.D., Wu, R., Guttman, D.S., and Desveaux, D. (2010). Allele-specific virulence
attenuation of the Pseudomonas syringae HopZ1a type III effector via the
Arabidopsis ZAR1 resistance protein. PLoS genetics 6, e1000894.
Li, W., Chiang, Y.H., and Coaker, G. (2013). The HopQ1 effector's nucleoside hydrolase-like
domain is required for bacterial virulence in arabidopsis and tomato, but not host
recognition in tobacco. PloS one 8, e59684.
Li, X., Lin, H., Zhang, W., Zou, Y., Zhang, J., Tang, X., and Zhou, J.M. (2005). Flagellin
induces innate immunity in nonhost interactions that is suppressed by Pseudomonas
syringae effectors. Proceedings of the National Academy of Sciences of the United
States of America 102, 12990-12995.
Lindgren, P.B., Peet, R.C., and Panopoulos, N.J. (1986). Gene cluster of Pseudomonas
syringae pv. "phaseolicola" controls pathogenicity of bean plants and hypersensitivity
of nonhost plants. Journal of bacteriology 168, 512-522.
Liu, T., Liu, Z., Song, C., Hu, Y., Han, Z., She, J., Fan, F., Wang, J., Jin, C., Chang, J., et al.
(2012). Chitin-induced dimerization activates a plant immune receptor. Science 336,
1160-1164.
Liu, Y., and Zhang, S. (2004). Phosphorylation of 1-aminocyclopropane-1-carboxylic acid
synthase by MPK6, a stress-responsive mitogen-activated protein kinase, induces
ethylene biosynthesis in Arabidopsis. The Plant cell 16, 3386-3399.
Lopez-Solanilla, E., Bronstein, P.A., Schneider, A.R., and Collmer, A. (2004). HopPtoN is a
Pseudomonas syringae Hrp (type III secretion system) cysteine protease effector
that suppresses pathogen-induced necrosis associated with both compatible and
incompatible plant interactions. Molecular microbiology 54, 353-365.
Ma, W., and Berkowitz, G.A. (2007). The grateful dead: calcium and cell death in plant
innate immunity. Cellular microbiology 9, 2571-2585.
Mackey, D., Belkhadir, Y., Alonso, J.M., Ecker, J.R., and Dangl, J.L. (2003). Arabidopsis
RIN4 is a target of the type III virulence effector AvrRpt2 and modulates RPS2mediated resistance. Cell 112, 379-389.
Mackey, D., Holt, B.F., 3rd, Wiig, A., and Dangl, J.L. (2002). RIN4 interacts with
Pseudomonas syringae type III effector molecules and is required for RPM1mediated resistance in Arabidopsis. Cell 108, 743-754.
Melotto, M., Underwood, W., Koczan, J., Nomura, K., and He, S.Y. (2006). Plant stomata
function in innate immunity against bacterial invasion. Cell 126, 969-980.
Mersmann, S., Bourdais, G., Rietz, S., and Robatzek, S. (2010). Ethylene signaling
regulates accumulation of the FLS2 receptor and is required for the oxidative burst
contributing to plant immunity. Plant physiology 154, 391-400.
Meyers, B.C., Kozik, A., Griego, A., Kuang, H., and Michelmore, R.W. (2003). Genome-wide
analysis of NBS-LRR-encoding genes in Arabidopsis. The Plant cell 15, 809-834.
45 Miya, A., Albert, P., Shinya, T., Desaki, Y., Ichimura, K., Shirasu, K., Narusaka, Y.,
Kawakami, N., Kaku, H., and Shibuya, N. (2007). CERK1, a LysM receptor kinase, is
essential for chitin elicitor signaling in Arabidopsis. Proceedings of the National
Academy of Sciences of the United States of America 104, 19613-19618.
Monaghan, J., and Zipfel, C. (2012). Plant pattern recognition receptor complexes at the
plasma membrane. Current opinion in plant biology 15, 349-357.
Morris, C.E., Kinkel, L.L., Xiao, K., Prior, P., and Sands, D.C. (2007). Surprising niche for the
plant pathogen Pseudomonas syringae. Infection, genetics and evolution : journal of
molecular epidemiology and evolutionary genetics in infectious diseases 7, 84-92.
Morris, C.E., Sands, D.C., Vanneste, J.L., Montarry, J., Oakley, B., Guilbaud, C., and Glaux,
C. (2010). Inferring the evolutionary history of the plant pathogen Pseudomonas
syringae from its biogeography in headwaters of rivers in North America, Europe,
and New Zealand. mBio 1.
Morris, C.E., Sands, D.C., Vinatzer, B.A., Glaux, C., Guilbaud, C., Buffiere, A., Yan, S.,
Dominguez, H., and Thompson, B.M. (2008). The life history of the plant pathogen
Pseudomonas syringae is linked to the water cycle. The ISME journal 2, 321-334.
Mukhtar, M.S., Carvunis, A.R., Dreze, M., Epple, P., Steinbrenner, J., Moore, J., Tasan, M.,
Galli, M., Hao, T., Nishimura, M.T., et al. (2011). Independently evolved virulence
effectors converge onto hubs in a plant immune system network. Science 333, 596601.
Nicaise, V., Joe, A., Jeong, B.R., Korneli, C., Boutrot, F., Westedt, I., Staiger, D., Alfano,
J.R., and Zipfel, C. (2013). Pseudomonas HopU1 modulates plant immune receptor
levels by blocking the interaction of their mRNAs with GRP7. The EMBO journal 32,
701-712.
Nimchuk, Z., Marois, E., Kjemtrup, S., Leister, R.T., Katagiri, F., and Dangl, J.L. (2000).
Eukaryotic fatty acylation drives plasma membrane targeting and enhances function
of several type III effector proteins from Pseudomonas syringae. Cell 101, 353-363.
Nimchuk, Z.L., Fisher, E.J., Desveaux, D., Chang, J.H., and Dangl, J.L. (2007). The HopX
(AvrPphE) family of Pseudomonas syringae type III effectors require a catalytic triad
and a novel N-terminal domain for function. Molecular plant-microbe interactions :
MPMI 20, 346-357.
Nomura, K., Debroy, S., Lee, Y.H., Pumplin, N., Jones, J., and He, S.Y. (2006). A bacterial
virulence protein suppresses host innate immunity to cause plant disease. Science
313, 220-223.
Nomura, K., Mecey, C., Lee, Y.N., Imboden, L.A., Chang, J.H., and He, S.Y. (2011).
Effector-triggered immunity blocks pathogen degradation of an immunity-associated
vesicle traffic regulator in Arabidopsis. Proceedings of the National Academy of
Sciences of the United States of America 108, 10774-10779.
Ogasawara, Y., Kaya, H., Hiraoka, G., Yumoto, F., Kimura, S., Kadota, Y., Hishinuma, H.,
Senzaki, E., Yamagoe, S., Nagata, K., et al. (2008). Synergistic activation of the
46 Arabidopsis NADPH oxidase AtrbohD by Ca2+ and phosphorylation. The Journal of
biological chemistry 283, 8885-8892.
Penninckx, I.A., Thomma, B.P., Buchala, A., Metraux, J.P., and Broekaert, W.F. (1998).
Concomitant activation of jasmonate and ethylene response pathways is required for
induction of a plant defensin gene in Arabidopsis. The Plant cell 10, 2103-2113.
Raymond, B., Young, J.C., Pallett, M., Endres, R.G., Clements, A., and Frankel, G. (2013).
Subversion of trafficking, apoptosis, and innate immunity by type III secretion system
effectors. Trends in microbiology 21, 430-441.
Robert-Seilaniantz, A., Navarro, L., Bari, R., and Jones, J.D. (2007). Pathological hormone
imbalances. Current opinion in plant biology 10, 372-379.
Robert-Seilaniantz, A., Shan, L., Zhou, J.M., and Tang, X. (2006). The Pseudomonas
syringae pv. tomato DC3000 type III effector HopF2 has a putative myristoylation site
required for its avirulence and virulence functions. Molecular plant-microbe
interactions : MPMI 19, 130-138.
Rodriguez-Herva, J.J., Gonzalez-Melendi, P., Cuartas-Lanza, R., Antunez-Lamas, M., RioAlvarez, I., Li, Z., Lopez-Torrejon, G., Diaz, I., Del Pozo, J.C., Chakravarthy, S., et al.
(2012). A bacterial cysteine protease effector protein interferes with photosynthesis
to suppress plant innate immune responses. Cellular microbiology 14, 669-681.
Roux, M., Schwessinger, B., Albrecht, C., Chinchilla, D., Jones, A., Holton, N., Malinovsky,
F.G., Tor, M., de Vries, S., and Zipfel, C. (2011). The Arabidopsis leucine-rich repeat
receptor-like kinases BAK1/SERK3 and BKK1/SERK4 are required for innate
immunity to hemibiotrophic and biotrophic pathogens. The Plant cell 23, 2440-2455.
Schenk, P.M., Kazan, K., Wilson, I., Anderson, J.P., Richmond, T., Somerville, S.C., and
Manners, J.M. (2000). Coordinated plant defense responses in Arabidopsis revealed
by microarray analysis. Proceedings of the National Academy of Sciences of the
United States of America 97, 11655-11660.
Schroeder, G.N., and Hilbi, H. (2008). Molecular pathogenesis of Shigella spp.: controlling
host cell signaling, invasion, and death by type III secretion. Clinical microbiology
reviews 21, 134-156.
Schulze, B., Mentzel, T., Jehle, A.K., Mueller, K., Beeler, S., Boller, T., Felix, G., and
Chinchilla, D. (2010). Rapid heteromerization and phosphorylation of ligand-activated
plant transmembrane receptors and their associated kinase BAK1. The Journal of
biological chemistry 285, 9444-9451.
Senthil-Kumar, M., and Mysore, K.S. (2013). Nonhost resistance against bacterial
pathogens: retrospectives and prospects. Annual review of phytopathology 51, 407427.
Shan, L., He, P., Li, J., Heese, A., Peck, S.C., Nurnberger, T., Martin, G.B., and Sheen, J.
(2008). Bacterial effectors target the common signaling partner BAK1 to disrupt
multiple MAMP receptor-signaling complexes and impede plant immunity. Cell host &
microbe 4, 17-27.
47 Shan, L., Thara, V.K., Martin, G.B., Zhou, J.M., and Tang, X. (2000). The pseudomonas
AvrPto protein is differentially recognized by tomato and tobacco and is localized to
the plant plasma membrane. The Plant cell 12, 2323-2338.
Shao, F., Golstein, C., Ade, J., Stoutemyer, M., Dixon, J.E., and Innes, R.W. (2003).
Cleavage of Arabidopsis PBS1 by a bacterial type III effector. Science 301, 12301233.
Shimizu, T., Nakano, T., Takamizawa, D., Desaki, Y., Ishii-Minami, N., Nishizawa, Y.,
Minami, E., Okada, K., Yamane, H., Kaku, H., et al. (2010). Two LysM receptor
molecules, CEBiP and OsCERK1, cooperatively regulate chitin elicitor signaling in
rice. The Plant journal : for cell and molecular biology 64, 204-214.
Singer, A.U., Desveaux, D., Betts, L., Chang, J.H., Nimchuk, Z., Grant, S.R., Dangl, J.L.,
and Sondek, J. (2004). Crystal structures of the type III effector protein AvrPphF and
its chaperone reveal residues required for plant pathogenesis. Structure 12, 16691681.
Ting, J.P., and Davis, B.K. (2005). CATERPILLER: a novel gene family important in
immunity, cell death, and diseases. Annual review of immunology 23, 387-414.
Torres, M.A., Jones, J.D., and Dangl, J.L. (2006). Reactive oxygen species signaling in
response to pathogens. Plant physiology 141, 373-378.
Van der Biezen, E.A., and Jones, J.D. (1998). Plant disease-resistance proteins and the
gene-for-gene concept. Trends in biochemical sciences 23, 454-456.
Wan, J., Zhang, X.C., Neece, D., Ramonell, K.M., Clough, S., Kim, S.Y., Stacey, M.G., and
Stacey, G. (2008). A LysM receptor-like kinase plays a critical role in chitin signaling
and fungal resistance in Arabidopsis. The Plant cell 20, 471-481.
Wang, Y., Li, J., Hou, S., Wang, X., Li, Y., Ren, D., Chen, S., Tang, X., and Zhou, J.M.
(2010). A Pseudomonas syringae ADP-ribosyltransferase inhibits Arabidopsis
mitogen-activated protein kinase kinases. The Plant cell 22, 2033-2044.
Washington, E.J., Banfield, M.J., and Dangl, J.L. (2013). What a Difference a Dalton Makes:
Bacterial Virulence Factors Modulate Eukaryotic Host Cell Signaling Systems via
Deamidation. Microbiology and molecular biology reviews : MMBR 77, 527-539.
Whalen, M.C., Innes, R.W., Bent, A.F., and Staskawicz, B.J. (1991). Identification of
Pseudomonas syringae pathogens of Arabidopsis and a bacterial locus determining
avirulence on both Arabidopsis and soybean. The Plant cell 3, 49-59.
Wilton, M., Subramaniam, R., Elmore, J., Felsensteiner, C., Coaker, G., and Desveaux, D.
(2010). The type III effector HopF2Pto targets Arabidopsis RIN4 protein to promote
Pseudomonas syringae virulence. Proceedings of the National Academy of Sciences
of the United States of America 107, 2349-2354.
Xiang, T., Zong, N., Zhang, J., Chen, J., Chen, M., and Zhou, J.M. (2011). BAK1 is not a
target of the Pseudomonas syringae effector AvrPto. Molecular plant-microbe
interactions : MPMI 24, 100-107.
48 Xiang, T., Zong, N., Zou, Y., Wu, Y., Zhang, J., Xing, W., Li, Y., Tang, X., Zhu, L., Chai, J.,
et al. (2008). Pseudomonas syringae effector AvrPto blocks innate immunity by
targeting receptor kinases. Current biology : CB 18, 74-80.
Yamaguchi, Y., Pearce, G., and Ryan, C.A. (2006). The cell surface leucine-rich repeat
receptor for AtPep1, an endogenous peptide elicitor in Arabidopsis, is functional in
transgenic tobacco cells. Proceedings of the National Academy of Sciences of the
United States of America 103, 10104-10109.
Zhang, J., Li, W., Xiang, T., Liu, Z., Laluk, K., Ding, X., Zou, Y., Gao, M., Zhang, X., Chen,
S., et al. (2010). Receptor-like cytoplasmic kinases integrate signaling from multiple
plant immune receptors and are targeted by a Pseudomonas syringae effector. Cell
host & microbe 7, 290-301.
Zhang, J., Shao, F., Li, Y., Cui, H., Chen, L., Li, H., Zou, Y., Long, C., Lan, L., Chai, J., et al.
(2007). A Pseudomonas syringae effector inactivates MAPKs to suppress PAMPinduced immunity in plants. Cell host & microbe 1, 175-185.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., Ba, H., Zhou, J., and Zhang, Y.
(2012). Disruption of PAMP-induced MAP kinase cascade by a Pseudomonas
syringae effector activates plant immunity mediated by the NB-LRR protein SUMM2.
Cell host & microbe 11, 253-263.
Zhao, Y., Thilmony, R., Bender, C.L., Schaller, A., He, S.Y., and Howe, G.A. (2003).
Virulence systems of Pseudomonas syringae pv. tomato promote bacterial speck
disease in tomato by targeting the jasmonate signaling pathway. The Plant journal :
for cell and molecular biology 36, 485-499.
Zhou, H., Lin, J., Johnson, A., Morgan, R.L., Zhong, W., and Ma, W. (2011). Pseudomonas
syringae type III effector HopZ1 targets a host enzyme to suppress isoflavone
biosynthesis and promote infection in soybean. Cell host & microbe 9, 177-186.
Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J.D., Boller, T., and Felix, G. (2006).
Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts
Agrobacterium-mediated transformation. Cell 125, 749-760.
Zipfel, C., Robatzek, S., Navarro, L., Oakeley, E.J., Jones, J.D., Felix, G., and Boller, T.
(2004). Bacterial disease resistance in Arabidopsis through flagellin perception.
Nature 428, 764-767.
49 Chapter 2
Bacterial virulence factors modulate eukaryotic host cell signaling systems via
deamidation
Preface
The following chapter was published in September 2013 in Microbiology and
Molecular Biology Reviews under the title “What a difference a Dalton makes: Bacterial
virulence factors modulate eukaryotic host cell signaling systems via deamidation” (2013,
77, 527-539). I wrote this manuscript while planning and completing experiments for the
work described in Chapter 3. I received direction and edits from our collaborator, Mark
Banfield, from the John Innes Centre in Norwich, UK and my thesis advisor, Jeff Dangl.
Abstract
Pathogenic bacteria commonly deploy enzymes to promote virulence. These
enzymes can modulate the function of host cell targets. While the actions of some enzymes
can be very obvious (e.g. digesting plant cell walls), others have more subtle activities.
Depending on the lifestyle of the bacteria these subtle modifications can be crucially
important for pathogenesis. In particular, if bacteria rely on a living host, subtle mechanisms
to alter host cellular function are likely to dominate. Several bacterial virulence factors have
evolved to use enzymatic deamidation as a subtle post-translational mechanism to modify
the function of host protein targets. Deamidation is the irreversible conversion of amino
acids glutamine and asparagine to glutamic acid and aspartic acid, respectively.
Interestingly, all currently characterized bacterial deamidases affect the function of the target
protein by modifying a single glutamine residue in the sequence. Deamidation of target host
proteins can disrupt host signaling and downstream processes by either activating or
inactivating the target. Despite the subtlety of this modification, it has been shown to cause
dramatic, context-dependent, effects on host cells. Several crystal structures of bacterial
deamidases have been solved. All are members of the papain-like superfamily and display a
cysteine-based catalytic triad. However, these proteins form distinct structural subfamilies
and feature combinations of modular domains of various functions. Based on the diversity of
pathogens that use deamidation as a mechanism to promote virulence, and the recent
identification of multiple deamidases, it is clear that this enzymatic activity is emerging as an
important and widespread feature in bacterial pathogenesis.
Introduction
Many bacterial pathogens use diverse suites of virulence factors to contribute to
pathogenicity. These virulence factors include toxins and type III effectors, proteins injected
into host cells via specialized type III secretion systems. Effectors often modify eukaryotic
host target proteins with posttranslational modifications that alter normal cellular function.
Commonly described posttranslational modifications utilized by effectors include
ubiquitination, acetylation, and AMPylation (Cui and Shao, 2011; Ribet and Cossart, 2010;
Yarbrough and Orth, 2009). Recently, enzymatic deamidation has emerged as a common
posttranslational modification utilized by a broad range of bacterial pathogens of both plants
and animals to alter the function of host proteins. Deamidation is the substitution of an
amide group with a carboxylate group (Figure 2.1). Therefore, it converts glutamines and
asparagines to glutamic acid and aspartic acid, respectively. This irreversible amino acid
conversion results in an increase of approximately 1 Dalton in the mass of the target protein,
an increase in the negative charge of the target protein, and the release of ammonia.
Nonspecific deamidation can occur spontaneously as proteins age and are degraded
(Robinson and Rudd, 1974). In contrast, specific enzymatic deamidation can regulate
normal cellular functions, such as chemotaxis and protein turnover in prokaryotes, or to
51 disrupt eukaryotic host cell function during infection (Chao et al., 2006; Striebel et al., 2009).
Here, we focus on deamidases that contribute to bacterial virulence.
The topic of this review is the six currently known families of bacterial virulence
factors that use deamidation to modulate host functions during infection (Table 2.1).
Cytotoxic Necrotizing Factors, CNFs, are a family of deamidases from E. coli (CNF1, 2, and
3) and Yersinia pseudotuberculosis (CNFY). The CNFs target a glutamine residue (either
Gln63 or Gln61) in the switch II domain of GTPase proteins that is critical for function (Flatau
et al., 1997; Schmidt et al., 1997). Deamidation of this glutamine leads to constitutive
activation of the target GTPases, resulting in cytoskeletal rearrangements. Reorganization
of the actin cytoskeleton is one mechanism used by invasive bacteria to promote entry into
host cells (Doye et al., 2002; Galan and Zhou, 2000). BLF1 is a toxin from Burkholderia
pseudomallei that is lethal to mice and tissue culture cells (Cruz-Migoni et al., 2011). BLF1
inhibits host protein synthesis via deamidation of eIF4A (Cruz-Migoni et al., 2011). VopC is a
type III effector from Vibrio parahaemolyticus that deamidates and constitutively activates
small GTPases (Zhang et al., 2012). Pasteurella multocida toxin (PMT) is the major
virulence factor of Pasteurella multocida, the causal agent of a variety of mammalian and
avian diseases. PMT constitutively activates various G proteins including Gαq/11, Gαi, and
Gα12/13 via deamidation of a glutamine that is also in the switch II region (Orth et al., 2009).
Cell cycle inhibiting factors, Cifs, are type III effectors and cyclomodulins from multiple
bacterial species. Cifs inhibit ubiquitination pathways by deamidating glutamine 40 of
ubiquitin and the ubiquitin-like protein, NEDD8 (Cui et al., 2010). OspI is a type III effector
protein from Shigella flexneri that dampens host immune responses by deamidating UBC13
and disrupting the TRAF6-mediated signaling pathway (Sanada et al., 2012). We review the
details of each of these six families, specifically with respect to their three-dimensional
structures and the impact that deamidation has on the function of their host target proteins.
52 We conclude that deamidation, as a non-reversible modification, is likely an all or nothing
virulence switch to alter diverse cellular functions across diverse pathosystems.
Cytotoxic Necrotizing Factor Toxins
Cytotoxic Necrotizing Factor toxins deamidate Rho GTPases.
Cytotoxic Necrotizing Factor, CNF1, from E. coli was the first bacterial deamidase
identified (Flatau et al., 1997; Schmidt et al., 1997). CNF1 is expressed in uropathogenic
strains of E. coli (Caprioli A, 1983). Following the discovery of CNF1, a number of
homologues were identified in different strains of E. coli. For example, CNF2 is produced by
enteropathogenic E. coli strains isolated from cows and sheep (Oswald et al., 1989). The
amino acid sequence of CNF2 is 85% identical to CNF1. CNF3 was isolated from
necrotoxigenic E. coli that infects sheep and goats (Orden et al., 2007). CNF3 shares 70%
amino acid identity with CNF1. Interestingly, CNFs are not restricted to E. coli. Yersinia
pseudotuberculosis produces CNFY, which is 61% identical to CNF1 (Lockman et al., 2002).
Further, dermonecrotizing Toxin, DNT, from Bordetella is a virulence factor that also shares
sequence homology with CNF1. However, DNT behaves most efficiently as a
transglutaminase in vitro, active in the presence of polyamines (Masuda et al., 2000;
Schmidt et al., 2001). Therefore, we will not include a detailed discussion of DNT here.
Schmidt et al. and Flatau et al. demonstrated that CNF1 converts Gln63 of the critical
RhoA GTPase switch II region to glutamic acid via enzymatic deamidation (Flatau et al.,
1997; Schmidt et al., 1997). Importantly, neither Gln29 nor Gln52, the other two glutamine
residues in RhoA, were modified by CNF1 (Flatau et al., 1997; Schmidt et al., 1997).
Deamidation of Gln63 triggers constitutive activation of RhoA, functionally mimicking the
GTP-bound state. Furthermore, the RhoA Gln63Glu variant, that mimics the product of the
deamidation reaction, phenocopies the CNF1-treated RhoA on microinjection into Vero cells
(Flatau et al., 1997; Schmidt et al., 1997).
53 RhoA is not the only target of CNFs. CNF deamidases possess different GTPase
substrate specificities. CNF1 can deamidate Rac and Cdc42 in vitro and Cdc42 in intact
HeLa cells (Lerm et al., 1999; Schmidt et al., 1997). CNF2 can modify RhoA and Rac when
co-expressed in E. coli (Sugai et al., 1999). CNF3 can deamidate RhoA, Rac, and Cdc42 in
intact HeLa cells treated with recombinant toxin (Stoll et al., 2009). Although RhoA, Rac,
and Cdc42 serve as substrates for CNFY deamidation in vitro, only RhoA could be
deamidated by CNFY in intact cells (Hoffmann et al., 2004).
Rho GTPases are molecular switches that cycle between an inactive, GDP-bound
state and an active, GTP bound state (Heasman and Ridley, 2008). The active state initiates
a variety of downstream responses, including reorganization of the actin cytoskeleton. Rho
GTPases that are unable to hydrolyze GTP are maintained in a constitutively active state.
Activation of different members of the Rho GTPases family leads to specific effects on the
cytoskeleton. RhoA activation leads to actin reorganization into stress fibers, Rac activation
causes membrane ruffling, and Cdc42 activation causes the formation of filopodia (Nobes
and Hall, 1995; Ridley and Hall, 1992; Ridley et al., 1992). Treatment with CNF1 induces
stress fiber formation, microspikes, membrane ruffles, and multinucleation in intact cells
(Flatau et al., 1997; Lerm et al., 1999; Schmidt et al., 1997). CNF3 and CNFY also trigger
strong formation of actin stress fibers in intact cells (Hoffmann et al., 2004; Stoll et al., 2009).
Surprisingly, there is no evidence of CNF3-triggered membrane ruffling or filopodia
formation. It has been demonstrated that CNF1 is required for the invasion of uropathogenic
E.coli into uroepithelial cell monolayers (Doye et al., 2002). This is consistent with other
bacterial pathogens, such as Salmonella, that attack the host cytoskeletal to facilitate
bacterial entry (Galan and Zhou, 2000).
The C-terminus of CNF1 forms a single compact catalytic domain.
CNFs are modular proteins. The N-terminal domains of these toxins are required for
binding and entry into host cells via receptor-mediated endocytosis (Contamin et al., 2000;
54 Lemichez et al., 1997). The structure of the catalytic C-terminus of CNF1 (residues 7201014) revealed a three-layered α/β/β sandwich, consisting of two mixed 5-stranded β-sheets
that are flanked by two α-helices (Figure 2.2A) (Buetow et al., 2001). The structure of CNF1
identified a novel overall protein fold that contained an arrangement of residues in the active
site reminiscent of the papain-like superfamily of enzymes that include the transglutaminase
blood coagulation factor XIII and GMP synthetase (Buetow and Ghosh, 2003). More
recently, it was determined that CNF1 is also structurally similar to Thermotoga maritima
deamidase protein, CheD, and a bacterial protein of unknown function, YfiH (Chao et al.,
2006). Papain-like superfamily enzymes typically contain a catalytic triad consisting of
histidine, cysteine, and asparagine/aspartic acid. The cysteine and histidine form a thiolateimidazolium pair that is critical for activating the cysteine nucleophile, with the
asparagine/aspartic acid thought to be involved in orienting the histidine. CNF deamidases
contain the invariant cysteine and histidine residues (Schmidt et al., 1998). Interestingly, in
CNF1, valine is the third member of the catalytic triad and this residue is involved in
orienting the histidine through a main chain carbonyl hydrogen bond (Buetow et al., 2001).
Structure-based mutagenesis of conserved residues in the catalytic pocket identified
Asn835 and Ser864 that are also required for CNF1 activity (Figure 2.2A) (Buetow and
Ghosh, 2003). Further, deletion of five individual loop regions surrounding the CNF1
catalytic pocket revealed that loops 8 and 9 are required for deamidation of RhoA (Buetow
and Ghosh, 2003). Most likely Asn835 and Ser864 have indirect roles in catalysis and loops
8 and 9 are important for substrate recognition.
A peptide derived from RhoA residues 59-69 could be deamidated by CNF1, thus
defining the minimal region of RhoA required for CNF1 deamidation (Flatau et al., 2000).
Mutations outside of these key residues did not interfere with the ability of CNF1 to
deamidate RhoA (Flatau et al., 2000). This suggests that CNF1 interacts with a narrow
55 surface of RhoA, and likely other GTPase substrates, and does not require a significant
binding surface for substrate recognition.
Burkholderia Lethal Factor 1 inhibits activity of translation factor eIF4A.
Burkholderia Lethal Factor 1, formerly known as BPSL1549, is a potent toxin from
Burkholderia pseudomallei. Melioidosis is a disease caused by infection with Burkholderia
pseudomallei, a pathogen endemic in areas in Southeast Asia and northern Australia
(Wiersinga et al., 2006). It is associated with severe septicemia and can remain dormant in
the host for decades. The molecular basis for the pathogenicity of B. pseudomallei is poorly
understood.
BLF1 interacts with the human translation factor eIF4A in a human cell lysates (CruzMigoni et al., 2011), suggesting that BLF1 might inhibit translation. This interaction was
confirmed by co-immunoprecipitation. BLF1 reduces endogenous host cell protein synthesis
and triggers increased stress granule formation, which is associated with translational
blocks. In order to determine the outcome of the BLF1/elF4A interaction, FLAG-tagged
eIF4A was purified from human cells expressing BLF1. Subsequent mass spectrometric
analysis of eIF4A revealed deamidation of Gln339 to Glu. To determine how deamidation of
this residue altered eIF4A activity, the Gln339Glu variant was generated. Although this
variant was capable of binding both RNA and ATP, and maintained wild-type levels of
ATPase activity, helicase activity was reduced by 47% (Cruz-Migoni et al., 2011).
The crystal structure of BLF1 revealed an α/β fold comprising a sandwich of two
mixed β-sheets surrounded by loops and α-helices (Figure 2.2B) (Cruz-Migoni et al., 2011).
The β-sheet core of the catalytic pocket is structurally similar to that of the deamidase
domain of CNF1, with an rmsd of 3.9 Angstroms over 170 residues (Cruz-Migoni et al.,
2011). Interestingly, these deamidase domains share only 9% amino acid identity. BLF1
lacks the receptor binding and translocation domains of CNF1, an expected evolutionary
innovation, since the intra-cellular lifestyle of B. pseudomallei renders these domains
56 unnecessary. However, the invariant catalytic residues found in CNF1, cysteine and
histidine, are conserved in the catalytic pocket. The third residue of the catalytic triad in
BLF1 is a threonine. BLF1 deamidation of eIF4A and efficient inhibition of protein translation
is dependent on the catalytic cysteine as mutation of this residue to a serine renders BLF1
essentially non-functional (Cruz-Migoni et al., 2011).
Vibrio parahaemolyticus type III effector VopC deamidates small GTPases.
A recent addition to the CNF family of deamidases is VopC from Vibrio
parahaemolyticus. V. parahaemolyticus is an enteric pathogen that contains two type III
secretion systems, TTSS1 and TTSS2 (Broberg et al., 2011); TTSS2 is encoded within a
pathogenicity island. The V. parahaemolyticus TTSS2 is necessary and sufficient for
invasion of and replication in both HeLa cells and CaCo2 cells. Among the type III effectors
encoded in the T3SS2 pathogenicity island is VopC, which shares 20% amino acid identity
to the catalytic domain of CNF from E. coli. As VopC is a type III effector protein it does not
share the same mechanism for host cell delivery as CNF1. As a result, VopC and CNF1
have different N-terminal domains.
Both the catalytic cysteine and histidine residues of CNF1 are conserved in VopC
(Cys220 and His235). Although a VopC structure has yet to be determined, homology
modeling suggests that the carbonyl oxygen of Leu187 completes the papain-like catalytic
triad. VopC deamidates and constitutively activates small GTPases, Rac and Cdc42, to
facilitate host cell invasion (Zhang et al., 2012). VopC is not able to activate RhoA. However,
as noted above, differences in substrate specificity are common among CNF homologs.
VopC activates Rac by deamidating Gln61 on the critical switch II domain. V.
parahaemolyticus invasion of nonphagocytic HeLa cells is dependent on catalytically active
VopC (Zhang et al., 2012).
57 Pasteurella multocida toxin
PMT is a multidomain toxin, with catalytic activity located at the C-terminus.
Pasteurella multocida toxin, PMT, is a monomeric, 146 kDa enzyme. P. multocida
causes severe diseases in animals and humans. It is associated with atrophic rhinitis in pigs
and dermatonecrosis and respiratory diseases, also known as snuffles, in cattle and rabbits
(Felix et al., 1992; Mushin and Schoenbaum, 1980). In humans, P. multocida can cause
dermonecrosis from bite wounds and bacteremia (Garcia, 1997).
PMT is a multidomain toxin. The N-terminal region of PMT has sequence homology
with the N-termini of CNFs, suggesting that this region is also used for binding to the host
cell surface. In fact, the first 506 N-terminal residues of PMT are sufficient to bind to cells
and to compete with full-length toxin in this assay (Pullinger et al., 2001).
The C-terminal region of PMT contains the catalytic center of the enzyme and is
sufficient for PMT-mediated toxicity when delivered into cells by electroporation (Pullinger et
al., 2001). The crystal structure of the C-terminal region of PMT reveals a multi-domain
Trojan horse-like structure with the feet, head, and body consisting of different domains (C1,
C2 and C3) that possess different functions (Figure 2.3A). The C1 domain, or the feet,
consists of seven α-helices. The N-terminal α-helices of the C1 domain are structurally
similar to the N-terminal domain of Clostridium difficile toxin B (Kamitani et al., 2010). In
toxin B, these helices target the protein to the plasma membrane. Similarly, the first four αhelices of C1 are required to target the catalytic domain of PMT to the plasma membrane
(Kamitani et al., 2010; Kitadokoro et al., 2007). Deletion of the C1 domain causes a
reduction, but not a complete loss, of PMT function, suggesting that localization to the
plasma membrane is necessary for full PMT-induced toxicity (Kamitani et al., 2010).
The function of the C2 domain, the body of the Trojan horse, is currently unknown. It
is the largest section of the PMT C-terminal region and consists of two subdomains. Each
subdomain contains typical α/β folds. Searches using the DALI server revealed possible
58 structural homology between the second subdomain and phosphate binding enzymes
(Kitadokoro et al., 2007). Therefore, the C2 domain may be enzymatic or may be involved in
binding target proteins in the host cell.
The C3 domain is the catalytic deamidase domain of PMT. Residues Cys1165 and
His1205 (which form the thiolate-imidazolium pair) and His1223 are required for PMT
function (Orth et al., 2003; Ward et al., 1998). Comparison with the structures of other
members of the papain-like superfamily reveals Asp1220 of PMT is the third residue of the
catalytic triad. Mutation of any of these residues resulted in a complete loss of PMT function
(Kitadokoro et al., 2007). Interestingly, the function of the catalytic cysteine requires the
reduction of a disulfide bond with Cys1159 (Figure 2.3B) (Kitadokoro et al., 2007).
Reduction of the disulfide bond between Cys1159 and Cys1165 causes a rotation of the
Cys1165 side chain towards the catalytic triad (Figure 2.3C). Therefore, the disulfide bond
has to be reduced for complete deamidase activity. Additionally, a Gln1225 in the catalytic
cleft was shown to be required for deamidase activity. It is postulated that this glutamine
forms an oxyanion hole to stabilize the tetrahedral intermediate, similar to the role of Gln19
in papain (Kitadokoro et al., 2007).
PMT deamidation of heterotrimeric G protein families affects several downstream signaling
events.
Unlike bacterial toxins that have high specificity for their host targets, such as the
CNFs, PMT mediates its virulence effect through activation of multiple heterotrimeric Gprotein families. This leads to a plethora of changes in host signaling that lead to
dysregulation of normal host cell homeostasis. A variety of cell systems and assays have
been used to establish that the pleiotropic effects of PMT in target cells are due to the ability
of PMT to activate multiple α subunits of diverse heterotrimeric G proteins (Hennig et al.,
2008; Lacerda et al., 1996; Orth et al., 2007; Orth et al., 2008; Orth et al., 2005; Preuss et
al., 2009; Staddon et al., 1990; Takasaki et al., 2004; Thomas et al., 2001; Wilson et al.,
59 1997; Zywietz et al., 2001). Heterotrimeric G proteins consist of four major families; Gαs,
Gαi, Gαq/11, and Gα12/13. When PMT and Gαi proteins were co-expressed in E. coli, PMT
modifies Gαi via deamidation (Orth et al., 2009). PMT converts the conserved Gln205 in the
switch II region of Gαi to glutamic acid. PMT-mediated activation of Gαi proteins leads to the
inhibition of adenylyl cyclase and reduction of cAMP accumulation (Figure 2.4) (Orth et al.,
2008).
PMT activation of Gαq/11 proteins results in the stimulation of phospholipase Cβ
(PLCβ) (Figure 2.4). PLCβ catalyzes the hydrolysis of phostidylinositol-4,5-bisphosphate to
inositol-1,4,5-triphosphate and diacylglycerol. It was originally postulated that PMT was
capable of activating Gαq (Orth et al., 2004; Zywietz et al., 2001). However, more recent in
vitro expression assays coupled with a mass spectrometry output demonstrated that PMT is
also able to deamidate Gα11 (Orth et al., 2012). PMT-induced Gα11 activation of PLCβ is
weaker than the Gαq activation of PLCβ (Kamitani et al., 2011). Therefore, although PMT
can deamidate and activate both Gαq and Gα11 in vitro, the relative contributions of these
activities in vivo may differ. Regardless, activation of the Gαq/11 sub-family via PMT results in
an increased release of second messengers inositol-1,4,5-triphosphate and diacylglycerol,
and a consequent increase in calcium mobilization. This leads to multiple calcium-mediated
responses such as activation of protein kinase C and secretion of chloride ions (Hennig et
al., 2008; Staddon et al., 1990; Wilson et al., 1997). PMT also stimulates the JAK/STAT
pathway in a Gαq- dependent manner (Orth et al., 2007).
PMT also activates the small GTPase RhoA, resulting in reorganization of the actin
cytoskeleton and formation of actin stress fibers (Lacerda et al., 1996; Thomas et al., 2001).
PMT induces RhoA indirectly through activation of multiple heterotrimeric Gα protein families
(Figure 2.4). PMT-induced RhoA activation can occur through Gαq (Wilson et al., 1997;
Zywietz et al., 2001) and Gα12/13 (Orth et al., 2005). A novel cyclic peptide that specifically
inhibits Gαq partially blocked PMT-induced RhoA activation (Orth et al., 2005; Takasaki et
60 al., 2004). Additionally, PMT-induced RhoA activation was inhibited in Gα12/13-deficient
HEK293m3 cells (Orth et al., 2005). However, it was not clear whether PMT functioned
through Gα12, Gα13, or both. Recently, mass spectrometric analysis revealed that PMT can
deamidate both Gα12 and Gα13 (Orth et al., 2012).
Finally, because Gα activation leads to the release of Gβγ from the heterotrimeric
complex, PMT also stimulates increased signaling through Gβγ-specific effectors such as
phosphoinositide-3-kinase (Figure 2.4) (Preuss et al., 2009).
Cycle Inhibiting Factors
Members of the cycle inhibiting factor (Cif) family of deamidases are found in many bacterial
pathogens.
Enteropathogenic (EPEC) and enterohemorrhagic (EHEC) E. coli are causal agents
of intestinal diarrhea (Croxen and Finlay, 2010; Kaper et al., 2004). EPEC is the cause of
potentially fatal diarrhea in infants in developing countries. In contrast, highly infectious
EHEC outbreaks occur most often in developed countries, most notably North America,
Japan, and Europe. Although the diseases caused by these two E. coli pathovars differ, they
both cause the formation of attaching and effacing (A/E) lesions (Croxen and Finlay, 2010).
The genes responsible for attaching and effacing lesions, the type III secretion system and
type III effectors, are encoded in the 35 kb pathogenicity island called the locus of
enterocyte effacement (LEE) (McDaniel et al., 1995). However, several effectors are nonLEE encoded (Nle) (Dean and Kenny, 2009). Cif was the first Nle effector identified in EPEC
and EHEC (Marches et al., 2003).
Cif homologs are encoded in multiple bacterial pathogen genomes and are all
associated with horizontally transferred genetic elements (Jubelin et al., 2009; Marches et
al., 2003). Cif homologues from Burkholderia pseudomallei, Yersinia pseudotuberculosis,
Photorhabdus luminescens, and Photorhabdus asymbiotica have been named CHBP,
CHYP, CHPL, and CHPA, respectively. Interestingly, these proteins have low primary amino
61 acid identity compared to Cif from E. coli: 56% for CHYP, 26% for CHBP, 23% for CHPL and
26% for CHPA (Jubelin et al., 2009; Yao et al., 2009).
Deamidation of NEDD8 by Cif leads to disruption of the ubiquitin-proteasome system.
CifEc was shown to interact with NEDD8 in a yeast-2-hybrid system and this
interaction was verified by in vitro pull-down and co-expression assays (Cui et al., 2010;
Jubelin et al., 2010; Morikawa et al., 2010). Cif proteins modify NEDD8 and ubiquitin via
deamidation (Cui et al., 2010). Ubiquitin and ubiquitin-like (UBL) proteins such as NEDD8
control many cellular processes by targeting various proteins for degradation through the
ubiquitin-proteasome system. Ligation of ubiquitin and UBLs, such as NEDD8, to substrate
proteins requires a series of three enzymes; an E1-activating enzyme, an E2-conjugating
enzyme, and an E3 ligase. Cullin-RING ligase (CRLs) are a large superfamily of E3 ubiquitin
ligases (Wei et al., 2008). The core CRL complex consists of cullin proteins, which are the
scaffold proteins of CRLs and are modified by NEDD8, and a small RING protein.
In eukaryotic cells, cullins are continuously neddylated and deneddylated. The
conjugation and removal of NEDD8 is an important cycling mechanism that regulates CRL
activity. The COP9 signalosome (CSN), a conserved multiprotein complex with protease
activity, is responsible for cullin deneddylation (Wei et al., 2008). When cullins are
neddylated they are maintained in an active, open conformation that allows substrate
proteins to interact with the Rbx protein at the cullin C-terminus (Duda et al., 2008). This
flexible protein conformation and interaction leads to the ubiquitination and subsequent
degradation of target proteins, some of which include important cell cycle regulators (Figure
2.5A).
Cifs target residue Gln40 in NEDD8 and ubiquitin with exquisite selectivity (neither
Gln39 nor Gln41 of NEDD8 is a target for deamidation) (Cui et al., 2010; Yao et al., 2012).
To date, recombinant CifEc, CifBp and CifYp have all been shown to deamidate Gln40 of
NEDD8 in vitro (Cui et al., 2010; Yao et al., 2012). Ectopic expression of the Gln40Glu
62 NEDD8 variant in HeLa cells phenocopies either the ectopic expression of Cifs or the effects
of Cif during bacterial infection demonstrating this activity is responsible for the Cif-mediated
cytopathic phenotype.
The deamidation of NEDD8 by Cif results in reduction in the rate of cullin
deneddylation (Boh et al., 2011; Jubelin et al., 2010; Morikawa et al., 2010). The Gln40Glu
mutation in NEDD8 is sufficient to decrease the deneddylation rate by the CSN in the
absence of Cif (Boh et al., 2011). Gln40 is located close to the cullin-NEDD8 conjugation
site. Deamidation of Gln40 in NEDD8 may prevent CSN-mediated deneddylation of CRLs
through interference with the global NEDD8-induced conformational change of cullins (which
may be necessary for CSN recognition (Figure 2.5B) (Boh et al., 2011). In contrast to
previous studies, it was shown in yeast that CSN can remove deamidated NEDD8 from
CRLs more efficiently than wild-type NEDD8 (Toro et al., 2013). This causes an increase in
deneddylated CRLs. Additionally, it is appropriate to mention that deamidation of NEDD8
and subsequent effects on CRL activity have not been reconstituted in vitro. Therefore, the
exact effect of Cif-mediated deamidation of NEDD8 remains unclear.
Cif causes cytopathic phenotypes such as rearrangement of the host cytoskeleton
and the formation of actin stress fibers. However, the hallmark of Cif infection is the
inhibition of cell cycle progression. These multiple cytopathic phenotypes can be directly
attributed to disruption of the ubiquitin-protease system. Depending on the stage of the cells
during infection, Cifs can arrest the cell cycle at either the G1/S transition or the G2/M
transition (Marches et al., 2003; Taieb et al., 2006). G1/S and G2/M transitions require
complexes that consist of a cyclin and a cyclin-dependent kinase (CDK). Inhibiting CRL
activity prevents ubiquitination and degradation of cell cycle regulators such as p21 and p27,
resulting in inactivation of cyclin-dependent kinases and cell cycle arrest (Marches et al.,
2003; Samba-Louaka et al., 2008; Taieb et al., 2006). This triggers an increase in the
inactive and phosphorylated CDKs, leading to cell cycle arrest (Marches et al., 2003).
63 Cifs also inhibit the degradation of a variety of other CRL substrates such as pIκB,
Cdt1, β-catenin, and RhoA (Marches et al., 2003). Therefore, inhibition of CRL activity can
account for the multiple phenotypes observed during infection with bacterial species
expressing Cifs.
Multiple structures reveal that Cif proteins are members of the papain-like superfamily.
The crystal structures of CifEc, CHBP, CHYP, and CHPL have been solved (Figure
2.6). Although the primary amino acid similarity among the homologs is low, the structures
are highly conserved (Crow et al., 2009; Hsu et al., 2008). The disordered N-termini provide
type III secretion signals and are either not included in expressed proteins or are not
observed in the crystal structures. To date, the only crystal structure of CifEc includes
truncation of the first 100 amino acids (Hsu et al., 2008). The overall fold of Cifs comprises a
head-and-tail domain organization (Figure 2.6A). The tail region, also known as the helical
extension domain, is encoded in the N-terminus of the protein. The head domain, also
known as the globular core, is formed by the C-terminus of the protein. The head domain
consists of an anti-parallel β-sheet surrounded by α-helices. This globular domain displays
overall structural homology to the papain-like superfamily. Similar to the other deamidases
covered in this review, the active sites of all Cifs contain an invariant Cys-His thiolateimidazolium pair. The third residue of the catalytic triad in Cifs is a glutamine (Figure 2.6B).
Each of these catalytic residues are required for Cif-induced cytopathic effects, deamidation
of NEDD8 and ubiquitin, and the resulting increase in stability of cell cycle regulators, p21
and p27 (Cui et al., 2010; Hsu et al., 2008; Jubelin et al., 2009; Yao et al., 2009).
An interesting feature observed in the structures of Cifs is the so-called occluding
loop. This loop was first identified in CifEc and is conserved across the Cif family (Crow et al.,
2009; Hsu et al., 2008). The occluding loop partially blocks access to the catalytic site
(substrate binding cleft of papain) and was hypothesized to be important in defining
specificity for Cif substrates (Figure 2.6B; see below).
64 Cif deamidases were the first to be co-crystallized with their substrates.
The crystal structures of CHYP, CHPL, and CHBP have been determined in complex
with NEDD8 (Crow et al., 2012; Yao et al., 2012). Additionally, the co-crystal structure of
CHBP and ubiquitin has also been determined (Yao et al., 2012). The structure of these
complexes reveals an extensive binding interface between Cif proteins and their
ubiquitin/NEDD8 substrates (Figure 2.7A), which involves both the head and the tail
domains of Cifs. Multiple alanine mutations in CHBP in the regions that interface with either
the N-terminal β-hairpin or the residues adjacent to Gln40 in ubiquitin or NEDD8 were
sufficient to abolish deamidation and Cif-induced cell cycle arrest (Yao et al., 2012).
Furthermore, alanine mutations in the corresponding residues in ubiquitin and NEDD8
disrupted both the interaction and deamidation (Yao et al., 2012). Interestingly, all individual
alanine mutations in CHYP (that were tested) were not sufficient to disrupt NEDD8 complex
formation. Only mutations that introduced steric clashes (based on the complex structure)
prevented binding (Yao et al., 2012).
There is very little change in the structure of Cifs when bound to their substrates,
compared to the uncomplexed states (Crow et al., 2012; Crow et al., 2009; Yao et al., 2012).
However, a significant re-orientation of the flexible C-terminal tail of NEDD8 and ubiquitin is
observed on binding by Cifs. Interestingly, it is the occluding loop in Cifs that appears
responsible for forcing this re-orientation during substrate binding. Displacement of the
flexible C-terminal tail is likely important for substrate recognition and orienting Gln40 in the
Cif catalytic pocket (Crow et al., 2012; Yao et al., 2012). Consistent with this, deletion of the
C-terminal domain of either ubiquitin or NEDD8 diminished deamidation by Cifs (Crow et al.,
2009). Therefore, multiple areas within the extensive interface between Cif deamidases and
their substrates are important for activity.
CHBP has a preference for NEDD8 as a substrate (Cui et al., 2010), and CHYP
deamidation of NEDD8 is ten-fold more efficient than deamidation of ubiquitin (Crow et al.,
65 2012). Molecular-dynamics studies revealed that this efficiency difference may be due to
decreased flexibility of NEDD8 in complex with Cif, compared to ubiquitin (Yao et al., 2012).
Residue Glu31 of NEDD8 interacts electrostatically with several residues in CHBP. The
residue at the same position in ubiquitin is glutamine, which is not capable of electrostatic
interactions. To determine whether this residue contributes to the difference in movement
and deamidation efficiency, NEDD8 Glu31Gln and the ubiquitin Gln31Glu variants were
generated. NEDD8 Glu31Gln exhibited increased molecular dynamic motion in simulations
and deamidation efficiency in in vitro assays was comparable to that observed for wild-type
ubiquitin. Conversely, ubiquitin Gln31Glu exhibited decreased molecular-dynamic motion,
bound more tightly to CHBP, and was deamidated at a level similar to wild-type NEDD8
(Yao et al., 2012). Thus, the Gln31 substrate residue may determine the difference in
substrate deamidation efficiency by Cifs.
The data from the Cif/NEDD8 complexed structures also suggests a molecular
mechanism of enzymatic deamidation by Cifs (Figure 2.7B,C). All solved structures of
currently known enzymatic deamidases contain conserved papain-like catalytic residues and
α/β folds similar to those in the papain-like superfamily. As a result, the molecular
mechanism of deamidation is likely to be similar to the protease and transglutaminase
reactions catalyzed by other superfamily members. During these reactions, the catalytic
histidine residue deprotonates the cysteine sulfhydryl group to increase its ability to act as a
nucleophile. Then, this nucleophile attacks the amide substrate, resulting in an acyl-enzyme
intermediate that is resolved by hydrolysis. The catalytic histidine’s imidazole ring is often
coordinated by a hydrogen bond with a third residue and this is ensures optimal orientation
of the thiolate-imidazolium pair (Figure 2.7B,C) (Crow et al., 2012).
66 OspI inhibits host immune responses by deamidating an E2-conjugating enzyme.
Deamidation of Ubc13 by OspI inhibits host inflammatory responses.
Shigella species are routinely found in patients in developing countries that suffer
from severe diarrhea (Schroeder and Hilbi, 2008). The severe symptoms associated with
Shigella infections are due to the invasion and destruction of the colonic and rectal
epithelium and a severe inflammatory reaction (Schroeder and Hilbi, 2008).
S. flexneri infections trigger changes to the host cell membrane that initiate the DAG
(diacylglycerol)-CBM complex-NF-κB signaling pathway. S. flexneri invades the host cell via
macropinocytosis. It is then able to escape from the phagosome and into the cytoplasm
through an unknown mechanism (Dupont et al., 2009). The escape leads to the
accumulation of host membrane components near the site of infection and subsequent host
signaling events that lead to inflammation. This likely occurs through the activation of
phospholipases that hydrolyze plasma membrane phospholipids, such as DAG, which then
activate protein kinases (Rawlings et al., 2006). Subsequently, protein kinases activate the
complex of caspase-recruitment domain (CARD), B-cell lymphoma 10 (BCL-10), and
mucosa-associated-lymphoid-tissue lymphoma-translocation gene 1(MALT1). This complex
is also referred to as the CBM complex. The CBM complex interacts with and activates
tumor-necrosis factor receptor-associated factor 6 (TRAF6). TRAF6 is an E3 ubiquitin ligase
that interacts with the Ubc13, an E2 ubiquitin-conjugating enzyme, to form lysine 63 (Lys63)linked polyubiquitin chains (Takeuchi and Akira, 2010). The presence of unconjugated
Lys63 polyubiquitin chains activates the TGF-β-activated kinase 1 (TAK1), TAK1-binding
protein 1 (TAB1), and TAK1-binding protein 2 (TAB2) complex. This leads to IκBα
phosphorylation and degradation. This allows NF-κB to translocate into the nucleus, where it
activates the expression of proinflammatory cytokine genes (Takeuchi and Akira, 2010).
S. flexneri encodes a type III effector, OspI, which modulates this host response by
deamidating Ubc13. Incubation of OspI with TRAF6, ubiquitin conjugating E2 enzymes
67 (UBC13 and UEV1A), and ubiquitin revealed that OspI causes a shift in the mobility of
Ubc13 in SDS-PAGE gels. Mass spectrometry analysis revealed that Ubc13 is deamidated
at residue Gln100 in the presence of OspI (Sanada et al., 2012). The deamidation of Ubc13
Gln100 blocks its ubiquitin-conjugating activity, thereby inhibiting the TRAF6-DAG-CBM
complex and subsequent NF-κB signaling. This is made evident by the increase in
phosphorylation of the NF-κB inhibitor, IκBα, and consequent NF-κB translocation to the
nucleus in HeLa cells infected with S. flexneri ∆ospI (Sanada et al., 2012). Consistent with
this result, NF-κB translocation to the nucleus is inhibited in the presence of OspI (Sanada
et al., 2012).
OspI forms a papain-like catalytic pocket that rearranges upon binding Ubc13.
The crystal structure of OspI reveals an α/β fold with 4 β-strands, seven α-helices,
and one 310 helix (Figure 2.8A) (Sanada et al., 2012). OspI also shares structural homology
with the papain-like superfamily, as noted above for PMT and the Cif family of deamidases.
It is most closely related to the Pseudomonas syringae cysteine protease effector AvrPphB
(Shao et al., 2002; Zhu et al., 2004).
Superimposition of the catalytic domain of AvrPphB with the OspI crystal structure
revealed that the catalytic triad in OspI comprises Cys62, His145, and Asp160 (Sanada et
al., 2012). OspI inhibition of the DAG-CBM-NF-κB pathway is dependent on the catalytic
residues as mutation of any of these residues leads to the loss of OspI activity (Sanada et
al., 2012).
Recently, the crystal structure of the OspI(Cys62Ala)/Ubc13 complex has been
solved (Figure 2.8B,C) (Fu et al., 2013). This structure revealed that an extensive binding
surface between OspI and Ubc13 that covers 11% of the exposed surface of OspI (Fu et al.,
2013). A negatively-charged region and hydrophobic pocket of OspI bind α-helix 1 and the
L1 and L2 loops of Ubc13, respectively. Analysis of point mutations in these regions
revealed that both interfaces are required for deamidation of Ubc13 (Fu et al., 2013). In
68 contrast to Cifs, significant structural rearrangements occur in the catalytic pocket of OspI
upon substrate binding (Fu et al., 2013). In uncomplexed OspI, the catalytic pocket is
blocked by Asn61 (Figure 2.8A). When OspI binds Ubc13, Asn61 rotates approximately 180
degrees. This leads to formation of a hydrogen bond with Asn54 and repositioning of Ala62.
The movement of these residues opens the pocket allowing insertion of Gln100 from Ubc13
into the catalytic site. Superposition of the catalytic pockets of AvrPphB and Ubc13-bound
OspI reveals an overlap of OspI Ala62 (Cys62 in the wild-type protein) and the catalytic
cysteine of AvrPphB (Fu et al., 2013). This suggests that the remodeled OspI catalytic
pocket (when bound to Ubc13) contains the proper orientation for catalytic activity.
Additionally, Phe95 and His96 of OspI re-orient to form a hydrophobic pocket that interacts
with Ubc13 loops L1 and L2 (Fu et al., 2013).
The OspI/Ubc13 complexed structure confirms that the molecular mechanism of
OspI is similar to that of papain-like enzymes and that suggested for Cif/NEDD8
deamidation (Figure 2.7C) (Fu et al., 2013; Storer and Menard, 1994; Zhu et al., 2004). It is
interesting to note that the catalytic cysteine is capable of forming disulfide bonds with
Cys65 (Figure 2.8C) (Sanada et al., 2012). Although, it has not been demonstrated that this
disulfide bond affects the function and positioning of the catalytic cysteine, it is possible that
it maintains the deamidase in a non-catalytic state in non-reducing conditions, as has been
demonstrated for PMT (Figure 2.3B,C) (Kitadokoro et al., 2007).
Conclusions
Recently, it has been discovered that deamidation is a common mode of action used
by several bacterial virulence factors to modify the function of host proteins. Deamidation, as
an irreversible modification of host proteins, offers many advantages to invading pathogens.
Unlike phosphorylation, host systems lack the ability to reverse the effects of deamidation.
This makes deamidases particularly potent virulence factors. Characterized deamidases
have been shown to modify specific residues on host proteins that are key to the host
69 protein’s function. Additionally, there is high specificity in the proteins that are targeted by
deamidases, as can be seen in the ability of CNF1 to deamidate some by not all Rho
GTPases. Although the deamidation modification is subtle and specific, the downstream
impacts on host cell function are extensive. This is most evident in the case of cycle
inhibiting factors, which inhibit the ubiquitin-protease system leading to both cell cycle arrest
and cytoskeletal rearrangements. In fact, it is apparent that bacterial deamidase virulence
factors have evolved to target key modulators of critical cell signaling systems in hosts.
However, our understanding of the role of deamidases in bacterial virulence is still in
its infancy. Further elucidation of the molecular mechanisms and the interactions between
deamidases and their substrates is important. To date, the use of enzymatic deamidation
as a generic “house-keeping” mechanism for modifying protein function has only been
documented in prokaryotes. Do eukaryotes also use this activity to modulate protein
activity? Or is enzymatic deamidation in eukaryotic cells an innovation of bacterial pathogen
mediated via translocated proteins? Future work would be greatly aided by new methods for
easier identification of deamidation. Current methods require knowledge of the specific
substrate of each deamidase in order for activity to be assayed. This differs from enzymes
such as cysteine proteases, for which cleavage of generic substrates can be easily
measured. Additionally, the presence of nonspecific deamidation in protein samples can
make the identification of enzymatic deamidation difficult. Another important future direction
is to define the molecular mechanisms by which deamidation specifically impacts the
function of host cell targets. At present, our knowledge is restricted to observing the
phenotypic outcomes of deamidation on particular targets during diverse steps in the
lifestyles of diverse pathogens. It would be valuable to understand how deamidation alters
function at the level of the modified protein. For example, G proteins are commonly targets
of at least some of the deamidases discussed here. But is the phenotypic outcome of these
events always merely an irreversible loss of catalytic function or could the deamidated G
70 protein now act as a dominant negative on its normal cellular partners, amplifying the impact
of the deamidation event to multiple output pathways, as noted above for PMT. Such
understanding will be important, for example, in exploration of genome editing to target
glutamine to glutamate mutations.
Regardless of the challenges of these studies, it is clear that tackling the molecular
mechanisms of deamidation, as has been done for other posttranslational modifications
used by bacterial pathogens, is essential for our understanding of bacterial pathogenesis
and, in fact, for the normal host cell biology of their targets.
71 Figures
Figure 2.1. A schematic representation of enzymatic deamidation in proteins.
Deamidases act on specific residues in the target protein. For all currently studied bacterial
virulence factors, the targets of deamidation are glutamine side chains, which are converted
to glutamic acids. However, it is also possible for the amide of the side chains of
asparagines to be converted to aspartic acid. Deamidation results in an increase in the
negative charge of the target protein, an increase of approximately 1 Dalton in the mass of
the target protein, and the release of ammonia. Each of these outputs can be measured
experimentally to characterize the activity of deamidases.
72 Figure 2.2. The catalytic domains of diverse deamidases. A) Catalytic domain of CNF1
from E. coli (PDB 1HQ0) is shown with invariant catalytic residues Cys866 and His881
shown in purple. The main chain carbonyl of Val833 (also in purple) is involved in
coordinating the imidazole ring of His881. In pink are conserved residues important for
CNF1 function, whilst loops 8 and 9 (likely important for substrate recognition) are shown in
yellow. B) BLF1 (PDB 3TU8) is a 23 kDa toxin and deamidase that forms a compact
catalytic domain structurally related to CNF1 from E. coli and other papain-like superfamily
enzymes. The residues that comprise the catalytic triad (Cys94, His106, and Thr88) are
shown in purple. α-helices are in blue, β-sheets in green, and loops in grey. All structure
figures were prepared with PyMol.
73 Figure 2.3. Crystal structure of the C-terminal region of Pasteurella multocida toxin.
A) The crystal structure of the C-terminal region of PMT has been solved (residues 5751285, PDB 2EBF). The catalytic region consists of several domains. The helical C1 domain
(purple), are the feet of the Trojan horse, and contain membrane-targeting properties. The
C2 “body” domain contains 2 α/β subdomains (shown in green and yellow) of unknown
function. The domain responsible for the deamidase activity of PMT is the C3 “head” Cterminal domain (blue). B) The crystal structure of PMT catalytic domain reveals a disulfide
bond between the catalytic cysteine, Cys1165 and Cys1159 in the binding pocket.
Reduction of this disulfide bond is required for the interaction of the catalytic triad and
catalytic activity. His1205 and Asp1220 form the other residues of the triad.
74 Figure 2.4. PMT deamidates and activates heterotrimeric Gα proteins. The activation of
diverse G protein-couple receptor signal transduction pathways by PMT leads to several
phenotypes that are disruptive to the host. PMT activates Gαi, Gαq/11, and Gα12/13, but not
Gαs. PMT also triggers the release of Gβγ from the heterotrimeric complex, which leads to
activation of the phosphoinositide-3-kinase (PI3K) pathway.
75 Figure 2.5. Model of Cif inactivation of CRL activity and ubiquitination. A) In uninfected
cells, the COP9 signalosome (CSN) deneddylates CRLs. NEDD8 (yellow) conjugation and
removal represent an important mechanism by with CRL activity is regulated. When CRLs
are neddylated with wild-type NEDD8 (green), they are maintained in an open conformation.
This allows substrates (blue) to bind the CRLs, become polyubiquitinated, and then
degraded. B) In cells infected with bacteria that express Cif, deamidation of NEDD8 (yellow)
may prevent CRL (green) from forming an open conformation. This prevents CRLs from
associating with the substrate-binding module that allows downstream ubiquitination and
degradation of substrate proteins (blue) to occur. This may lead to a reduction in the rate of
deneddylation and a decrease in unmodified CRLs (grey).
76 Figure 2.6. Crystal structures of cell cycle-inhibiting factors. A) Crystal structures of Cif
deamidases from E. coli (PDB 3EFY), B. pseudomallei (PDB 3GQM), P. luminescens (PDB
3GQJ), and Y. pseudotuberculosis (PDB 4F8C) are shown. In blue are the α-helices of the
core globular domain. In teal are the α-helices of the tail helical extension domain. The
antiparallel β-sheet is green and the loops are colored in grey, except for the occluding loop,
which is colored in black. The conserved catalytic residues histidine, cysteine, and
glutamine are in purple. B) Amino acids in the catalytic domain of CHBP are shown as sticks
in purple.
77 Figure 2.7. Crystal structure of the CifYp deamidase/NEDD8 complex. A) The CHYP (αhelices in blue, β-sheets in green, and loops in grey) and NEDD8 (red) complex consists of
an extensive interaction surface (PDB 4FBJ). The catalytic Cys117 position is modeled
based on the CHYP(Cys117Ala)/NEDD8 complex to demonstrate the proposed molecular
mechanism of deamidation. B) An up-close view of the catalytic domain of CHYP shows the
deamidation target, Gln40 of NEDD8, is positioned near the Cif catalytic residues (shown in
purple). C) The proposed molecular mechanism of deamidation involves deprotonation of
the cysteine by the imidazolium group of histidine (step 1) followed by a nucleophilic attack
by the thiol group of the catalytic cysteine on the δ-carbon of glutamine (step 2).
78 Figure 2.8. Crystal structures of unbound OspI and the OspI/Ubc13 complex. A) OspI
(α-helices in blue, β-sheets in green, and loops in grey (PDB 3B21) forms a compact
catalytic domain structurally similar to papain-like superfamily enzymes. The residues in the
catalytic triad (Cys62, His145, and Asp160) are shown in purple. The catalytic pocket is
blocked by Asn61 in free OspI. B) The OspI(Cys62Ala)/Ubc13 (red) complex includes an
extensive interaction surface involving Ubc13 α-helix 1 and loops L1 and L2 (PDB 4IP3). C)
An up-close view of the OspI(Cys62Ala)/Ubc13 complex showing Gln100 of Ubc13
positioned adjacent to the OspI catalytic center (purple, OspI residue 62 has been modeled
as a cysteine to demonstrate the nucleophilic attack (yellow dashes) by the thiol group of
Cys62 on Gln100 in Ubc13).
79 Table 2.1. Bacterial virulence factors that use deamidation to modify host proteins.
Protein
Species
Target
Effects
Key References
Cytotoxic
Necrotizing
Factors (CNFs)
E. coli, Yersinia
pseudotuberculos
is
Rho GTPases
(RhoA, Rac,
Cdc42)
Activation of Rho
GTPases
Flatau et al.,
(1997); Schmidt
et al., (1997)
Burkholderia
Lethal Factor
(BLF1)
Burkholderia
pseudomallei
VopC
PMT
Cycle inhibiting
factors (Cifs)
OspI
Vibrio
parahaemolyticus
Pasteurella
multocida
E. coli,
Burkholderia
pseudomallei,
Yersinia
pseudotuberculos
is, Photorhabdus
luminescens,
Photorhabdus
asymbiotica
Shigella flexneri
eIF4A
Rho GTPases
(Rac, Cdc42)
Heterotrimeric G
proteins (Gαi,
Gαq/11, and
Gα12/13)
NEDD8 and
ubiquitin
UBC13
80 Inactivation of
helicase activity
of translation
factor
Activation of Rho
GTPases
Activation of G
protein signaling
Cruz-Migoni et
al., (2011)
Zhang et al.,
(2012)
Orth et al., (2009)
Inactivation
ubiquitinproteasome
system
Cui et al., (2010)
Inactivation of E2
ubiquitin
conjugating
activity of UBC13
Sanada et al.,
(2012)
REFERENCES
Boh, B.K., Ng, M.Y., Leck, Y.C., Shaw, B., Long, J., Sun, G.W., Gan, Y.H., Searle, M.S.,
Layfield, R., and Hagen, T. (2011). Inhibition of cullin RING ligases by cycle inhibiting
factor: evidence for interference with Nedd8-induced conformational control. Journal
of molecular biology 413, 430-437.
Broberg, C.A., Calder, T.J., and Orth, K. (2011). Vibrio parahaemolyticus cell biology and
pathogenicity determinants. Microbes Infect 13, 992-1001.
Buetow, L., Flatau, G., Chiu, K., Boquet, P., and Ghosh, P. (2001). Structure of the Rhoactivating domain of Escherichia coli cytotoxic necrotizing factor 1. Nature structural
biology 8, 584-588.
Buetow, L., and Ghosh, P. (2003). Structural elements required for deamidation of RhoA by
cytotoxic necrotizing factor 1. Biochemistry 42, 12784-12791.
Caprioli A, F.V., Roda LG, Ruggeri FM, Zona C (1983). Partial purification and
characterization of an escherichia coli toxic factor that induces morphological cell
alterations. Infection and immunity 39, 1300-1306.
Chao, X., Muff, T.J., Park, S.Y., Zhang, S., Pollard, A.M., Ordal, G.W., Bilwes, A.M., and
Crane, B.R. (2006). A receptor-modifying deamidase in complex with a signaling
phosphatase reveals reciprocal regulation. Cell 124, 561-571.
Contamin, S., Galmiche, A., Doye, A., Flatau, G., Benmerah, A., and Boquet, P. (2000). The
p21 Rho-activating toxin cytotoxic necrotizing factor 1 is endocytosed by a clathrinindependent mechanism and enters the cytosol by an acidic-dependent membrane
translocation step. Molecular biology of the cell 11, 1775-1787.
Crow, A., Hughes, R.K., Taieb, F., Oswald, E., and Banfield, M.J. (2012). The molecular
basis of ubiquitin-like protein NEDD8 deamidation by the bacterial effector protein
Cif. Proceedings of the National Academy of Sciences of the United States of
America 109, E1830-1838.
Crow, A., Race, P.R., Jubelin, G., Varela Chavez, C., Escoubas, J.M., Oswald, E., and
Banfield, M.J. (2009). Crystal structures of Cif from bacterial pathogens
Photorhabdus luminescens and Burkholderia pseudomallei. PloS one 4, e5582.
Croxen, M.A., and Finlay, B.B. (2010). Molecular mechanisms of Escherichia coli
pathogenicity. Nature reviews Microbiology 8, 26-38.
81 Cruz-Migoni, A., Hautbergue, G.M., Artymiuk, P.J., Baker, P.J., Bokori-Brown, M., Chang,
C.T., Dickman, M.J., Essex-Lopresti, A., Harding, S.V., Mahadi, N.M., et al. (2011). A
Burkholderia pseudomallei toxin inhibits helicase activity of translation factor eIF4A.
Science 334, 821-824.
Cui, J., and Shao, F. (2011). Biochemistry and cell signaling taught by bacterial effectors.
Trends in biochemical sciences 36, 532-540.
Cui, J., Yao, Q., Li, S., Ding, X., Lu, Q., Mao, H., Liu, L., Zheng, N., Chen, S., and Shao, F.
(2010). Glutamine deamidation and dysfunction of ubiquitin/NEDD8 induced by a
bacterial effector family. Science 329, 1215-1218.
Dean, P., and Kenny, B. (2009). The effector repertoire of enteropathogenic E. coli: ganging
up on the host cell. Current opinion in microbiology 12, 101-109.
Doye, A., Mettouchi, A., Bossis, G., Clement, R., Buisson-Touati, C., Flatau, G., Gagnoux,
L., Piechaczyk, M., Boquet, P., and Lemichez, E. (2002). CNF1 exploits the ubiquitinproteasome machinery to restrict Rho GTPase activation for bacterial host cell
invasion. Cell 111, 553-564.
Duda, D.M., Borg, L.A., Scott, D.C., Hunt, H.W., Hammel, M., and Schulman, B.A. (2008).
Structural insights into NEDD8 activation of cullin-RING ligases: conformational
control of conjugation. Cell 134, 995-1006.
Dupont, N., Lacas-Gervais, S., Bertout, J., Paz, I., Freche, B., Van Nhieu, G.T., van der
Goot, F.G., Sansonetti, P.J., and Lafont, F. (2009). Shigella phagocytic vacuolar
membrane remnants participate in the cellular response to pathogen invasion and
are regulated by autophagy. Cell host & microbe 6, 137-149.
Felix, R., Fleisch, H., and Frandsen, P.L. (1992). Effect of Pasteurella multocida toxin on
bone resorption in vitro. Infection and immunity 60, 4984-4988.
Flatau, G., Landraud, L., Boquet, P., Bruzzone, M., and Munro, P. (2000). Deamidation of
RhoA glutamine 63 by the Escherichia coli CNF1 toxin requires a short sequence of
the GTPase switch 2 domain. Biochemical and biophysical research communications
267, 588-592.
Flatau, G., Lemichez, E., Gauthier, M., Chardin, P., Paris, S., Fiorentini, C., and Boquet, P.
(1997). Toxin-induced activation of the G protein p21 Rho by deamidation of
glutamine. Nature 387, 729-733.
82 Fu, P., Zhang, X., Jin, M., Xu, L., Wang, C., Xia, Z., and Zhu, Y. (2013). Complex Structure
of OspI and Ubc13: The Molecular Basis of Ubc13 Deamidation and Convergence of
Bacterial and Host E2 Recognition. PLoS pathogens 9, e1003322.
Galan, J.E., and Zhou, D. (2000). Striking a balance: modulation of the actin cytoskeleton by
Salmonella. Proceedings of the National Academy of Sciences of the United States
of America 97, 8754-8761.
Garcia, V.F. (1997). Animal bites and Pasturella infections. Pediatrics in review / American
Academy of Pediatrics 18, 127-130.
Heasman, S.J., and Ridley, A.J. (2008). Mammalian Rho GTPases: new insights into their
functions from in vivo studies. Nature reviews Molecular cell biology 9, 690-701.
Hennig, B., Orth, J., Aktories, K., and Diener, M. (2008). Anion secretion evoked by
Pasteurella multocida toxin across rat colon. European journal of pharmacology 583,
156-163.
Hoffmann, C., Pop, M., Leemhuis, J., Schirmer, J., Aktories, K., and Schmidt, G. (2004). The
Yersinia pseudotuberculosis cytotoxic necrotizing factor (CNFY) selectively activates
RhoA. The Journal of biological chemistry 279, 16026-16032.
Hsu, Y., Jubelin, G., Taieb, F., Nougayrede, J.P., Oswald, E., and Stebbins, C.E. (2008).
Structure of the cyclomodulin Cif from pathogenic Escherichia coli. Journal of
molecular biology 384, 465-477.
Jubelin, G., Chavez, C.V., Taieb, F., Banfield, M.J., Samba-Louaka, A., Nobe, R.,
Nougayrede, J.P., Zumbihl, R., Givaudan, A., Escoubas, J.M., et al. (2009). Cycle
inhibiting factors (CIFs) are a growing family of functional cyclomodulins present in
invertebrate and mammal bacterial pathogens. PloS one 4, e4855.
Jubelin, G., Taieb, F., Duda, D.M., Hsu, Y., Samba-Louaka, A., Nobe, R., Penary, M.,
Watrin, C., Nougayrede, J.P., Schulman, B.A., et al. (2010). Pathogenic bacteria
target NEDD8-conjugated cullins to hijack host-cell signaling pathways. PLoS
pathogens 6, e1001128.
Kamitani, S., Ao, S., Toshima, H., Tachibana, T., Hashimoto, M., Kitadokoro, K., FukuiMiyazaki, A., Abe, H., and Horiguchi, Y. (2011). Enzymatic actions of Pasteurella
multocida toxin detected by monoclonal antibodies recognizing the deamidated alpha
subunit of the heterotrimeric GTPase Gq. The FEBS journal 278, 2702-2712.
83 Kamitani, S., Kitadokoro, K., Miyazawa, M., Toshima, H., Fukui, A., Abe, H., Miyake, M., and
Horiguchi, Y. (2010). Characterization of the membrane-targeting C1 domain in
Pasteurella multocida toxin. The Journal of biological chemistry 285, 25467-25475.
Kaper, J.B., Nataro, J.P., and Mobley, H.L. (2004). Pathogenic Escherichia coli. Nature
reviews Microbiology 2, 123-140.
Kitadokoro, K., Kamitani, S., Miyazawa, M., Hanajima-Ozawa, M., Fukui, A., Miyake, M., and
Horiguchi, Y. (2007). Crystal structures reveal a thiol protease-like catalytic triad in
the C-terminal region of Pasteurella multocida toxin. Proceedings of the National
Academy of Sciences of the United States of America 104, 5139-5144.
Lacerda, H.M., Lax, A.J., and Rozengurt, E. (1996). Pasteurella multocida toxin, a potent
intracellularly acting mitogen, induces p125FAK and paxillin tyrosine
phosphorylation, actin stress fiber formation, and focal contact assembly in Swiss
3T3 cells. The Journal of biological chemistry 271, 439-445.
Lemichez, E., Flatau, G., Bruzzone, M., Boquet, P., and Gauthier, M. (1997). Molecular
localization of the Escherichia coli cytotoxic necrotizing factor CNF1 cell-binding and
catalytic domains. Molecular microbiology 24, 1061-1070.
Lerm, M., Selzer, J., Hoffmeyer, A., Rapp, U.R., Aktories, K., and Schmidt, G. (1999).
Deamidation of Cdc42 and Rac by Escherichia coli cytotoxic necrotizing factor 1:
activation of c-Jun N-terminal kinase in HeLa cells. Infection and immunity 67, 496503.
Lockman, H.A., Gillespie, R.A., Baker, B.D., and Shakhnovich, E. (2002). Yersinia
pseudotuberculosis produces a cytotoxic necrotizing factor. Infection and immunity
70, 2708-2714.
Marches, O., Ledger, T.N., Boury, M., Ohara, M., Tu, X., Goffaux, F., Mainil, J., Rosenshine,
I., Sugai, M., De Rycke, J., et al. (2003). Enteropathogenic and enterohaemorrhagic
Escherichia coli deliver a novel effector called Cif, which blocks cell cycle G2/M
transition. Molecular microbiology 50, 1553-1567.
Masuda, M., Betancourt, L., Matsuzawa, T., Kashimoto, T., Takao, T., Shimonishi, Y., and
Horiguchi, Y. (2000). Activation of rho through a cross-link with polyamines catalyzed
by Bordetella dermonecrotizing toxin. The EMBO journal 19, 521-530.
McDaniel, T.K., Jarvis, K.G., Donnenberg, M.S., and Kaper, J.B. (1995). A genetic locus of
enterocyte effacement conserved among diverse enterobacterial pathogens.
Proceedings of the National Academy of Sciences of the United States of America
92, 1664-1668.
84 Morikawa, H., Kim, M., Mimuro, H., Punginelli, C., Koyama, T., Nagai, S., Miyawaki, A., Iwai,
K., and Sasakawa, C. (2010). The bacterial effector Cif interferes with SCF ubiquitin
ligase function by inhibiting deneddylation of Cullin1. Biochemical and biophysical
research communications 401, 268-274.
Mushin, R., and Schoenbaum, M. (1980). A strain of Pasteurella multocida associated with
infections in rabbit colonies. Laboratory animals 14, 353-356.
Nobes, C.D., and Hall, A. (1995). Rho, rac, and cdc42 GTPases regulate the assembly of
multimolecular focal complexes associated with actin stress fibers, lamellipodia, and
filopodia. Cell 81, 53-62.
Orden, J.A., Dominguez-Bernal, G., Martinez-Pulgarin, S., Blanco, M., Blanco, J.E., Mora,
A., Blanco, J., Blanco, J., and de la Fuente, R. (2007). Necrotoxigenic Escherichia
coli from sheep and goats produce a new type of cytotoxic necrotizing factor (CNF3)
associated with the eae and ehxA genes. International microbiology : the official
journal of the Spanish Society for Microbiology 10, 47-55.
Orth, J.H., Aktories, K., and Kubatzky, K.F. (2007). Modulation of host cell gene expression
through activation of STAT transcription factors by Pasteurella multocida toxin. The
Journal of biological chemistry 282, 3050-3057.
Orth, J.H., Blocker, D., and Aktories, K. (2003). His1205 and His1223 are essential for the
activity of the mitogenic Pasteurella multocida toxin. Biochemistry 42, 4971-4977.
Orth, J.H., Fester, I., Preuss, I., Agnoletto, L., Wilson, B.A., and Aktories, K. (2008).
Activation of Galpha (i) and subsequent uncoupling of receptor-Galpha(i) signaling
by Pasteurella multocida toxin. The Journal of biological chemistry 283, 2328823294.
Orth, J.H., Fester, I., Siegert, P., Weise, M., Lanner, U., Kamitani, S., Tachibana, T., Wilson,
B.A., Schlosser, A., Horiguchi, Y., et al. (2012). Substrate specificity of Pasteurella
multocida toxin for alpha subunits of heterotrimeric G proteins. FASEB journal :
official publication of the Federation of American Societies for Experimental Biology.
Orth, J.H., Lang, S., and Aktories, K. (2004). Action of Pasteurella multocida toxin depends
on the helical domain of Galphaq. The Journal of biological chemistry 279, 3415034155.
Orth, J.H., Lang, S., Taniguchi, M., and Aktories, K. (2005). Pasteurella multocida toxininduced activation of RhoA is mediated via two families of G{alpha} proteins,
G{alpha}q and G{alpha}12/13. The Journal of biological chemistry 280, 3670136707.
85 Orth, J.H., Preuss, I., Fester, I., Schlosser, A., Wilson, B.A., and Aktories, K. (2009).
Pasteurella multocida toxin activation of heterotrimeric G proteins by deamidation.
Proceedings of the National Academy of Sciences of the United States of America
106, 7179-7184.
Oswald, E., De Rycke, J., Guillot, J.F., and Boivin, R. (1989). Cytotoxic effect of
multinucleation in HeLa cell cultures associated with the presence of Vir plasmid in
Escherichia coli strains. FEMS microbiology letters 49, 95-99.
Preuss, I., Kurig, B., Nurnberg, B., Orth, J.H., and Aktories, K. (2009). Pasteurella multocida
toxin activates Gbetagamma dimers of heterotrimeric G proteins. Cellular signalling
21, 551-558.
Pullinger, G.D., Sowdhamini, R., and Lax, A.J. (2001). Localization of functional domains of
the mitogenic toxin of Pasteurella multocida. Infection and immunity 69, 7839-7850.
Rawlings, D.J., Sommer, K., and Moreno-Garcia, M.E. (2006). The CARMA1 signalosome
links the signalling machinery of adaptive and innate immunity in lymphocytes.
Nature reviews Immunology 6, 799-812.
Ribet, D., and Cossart, P. (2010). Pathogen-mediated posttranslational modifications: A reemerging field. Cell 143, 694-702.
Ridley, A.J., and Hall, A. (1992). The small GTP-binding protein rho regulates the assembly
of focal adhesions and actin stress fibers in response to growth factors. Cell 70, 389399.
Ridley, A.J., Paterson, H.F., Johnston, C.L., Diekmann, D., and Hall, A. (1992). The small
GTP-binding protein rac regulates growth factor-induced membrane ruffling. Cell 70,
401-410.
Robinson, A.B., and Rudd, C.J. (1974). Deamidation of glutaminyl and asparaginyl residues
in peptides and proteins. Current topics in cellular regulation 8, 247-295.
Samba-Louaka, A., Nougayrede, J.P., Watrin, C., Jubelin, G., Oswald, E., and Taieb, F.
(2008). Bacterial cyclomodulin Cif blocks the host cell cycle by stabilizing the cyclindependent kinase inhibitors p21 and p27. Cellular microbiology 10, 2496-2508.
Sanada, T., Kim, M., Mimuro, H., Suzuki, M., Ogawa, M., Oyama, A., Ashida, H., Kobayashi,
T., Koyama, T., Nagai, S., et al. (2012). The Shigella flexneri effector OspI
deamidates UBC13 to dampen the inflammatory response. Nature 483, 623-626.
86 Schmidt, G., Goehring, U.M., Schirmer, J., Uttenweiler-Joseph, S., Wilm, M., Lohmann, M.,
Giese, A., Schmalzing, G., and Aktories, K. (2001). Lysine and polyamines are
substrates for transglutamination of Rho by the Bordetella dermonecrotic toxin.
Infection and immunity 69, 7663-7670.
Schmidt, G., Sehr, P., Wilm, M., Selzer, J., Mann, M., and Aktories, K. (1997). Gln 63 of Rho
is deamidated by Escherichia coli cytotoxic necrotizing factor-1. Nature 387, 725729.
Schmidt, G., Selzer, J., Lerm, M., and Aktories, K. (1998). The Rho-deamidating cytotoxic
necrotizing factor 1 from Escherichia coli possesses transglutaminase activity.
Cysteine 866 and histidine 881 are essential for enzyme activity. The Journal of
biological chemistry 273, 13669-13674.
Schroeder, G.N., and Hilbi, H. (2008). Molecular pathogenesis of Shigella spp.: controlling
host cell signaling, invasion, and death by type III secretion. Clinical microbiology
reviews 21, 134-156.
Shao, F., Merritt, P.M., Bao, Z., Innes, R.W., and Dixon, J.E. (2002). A Yersinia effector and
a Pseudomonas avirulence protein define a family of cysteine proteases functioning
in bacterial pathogenesis. Cell 109, 575-588.
Staddon, J.M., Chanter, N., Lax, A.J., Higgins, T.E., and Rozengurt, E. (1990). Pasteurella
multocida toxin, a potent mitogen, stimulates protein kinase C-dependent and independent protein phosphorylation in Swiss 3T3 cells. The Journal of biological
chemistry 265, 11841-11848.
Stoll, T., Markwirth, G., Reipschlager, S., and Schmidt, G. (2009). A new member of a
growing toxin family--Escherichia coli cytotoxic necrotizing factor 3 (CNF3). Toxicon :
official journal of the International Society on Toxinology 54, 745-753.
Storer, A.C., and Menard, R. (1994). Catalytic mechanism in papain family of cysteine
peptidases. Methods in enzymology 244, 486-500.
Striebel, F., Imkamp, F., Sutter, M., Steiner, M., Mamedov, A., and Weber-Ban, E. (2009).
Bacterial ubiquitin-like modifier Pup is deamidated and conjugated to substrates by
distinct but homologous enzymes. Nature structural & molecular biology 16, 647-651.
Sugai, M., Hatazaki, K., Mogami, A., Ohta, H., Peres, S.Y., Herault, F., Horiguchi, Y.,
Masuda, M., Ueno, Y., Komatsuzawa, H., et al. (1999). Cytotoxic necrotizing factor
type 2 produced by pathogenic Escherichia coli deamidates a gln residue in the
conserved G-3 domain of the rho family and preferentially inhibits the GTPase
activity of RhoA and rac1. Infection and immunity 67, 6550-6557.
87 Taieb, F., Nougayrede, J.P., Watrin, C., Samba-Louaka, A., and Oswald, E. (2006).
Escherichia coli cyclomodulin Cif induces G2 arrest of the host cell cycle without
activation of the DNA-damage checkpoint-signalling pathway. Cellular microbiology
8, 1910-1921.
Takasaki, J., Saito, T., Taniguchi, M., Kawasaki, T., Moritani, Y., Hayashi, K., and Kobori, M.
(2004). A novel Galphaq/11-selective inhibitor. The Journal of biological chemistry
279, 47438-47445.
Takeuchi, O., and Akira, S. (2010). Pattern recognition receptors and inflammation. Cell
140, 805-820.
Thomas, W., Pullinger, G.D., Lax, A.J., and Rozengurt, E. (2001). Escherichia coli cytotoxic
necrotizing factor and Pasteurella multocida toxin induce focal adhesion kinase
autophosphorylation and Src association. Infection and immunity 69, 5931-5935.
Toro, T.B., Toth, J.I., and Petroski, M.D. (2013). The cyclomodulin CIF alters cullin
neddylation dynamics. Journal of Biological Chemistry [Epub ahead of print].
Ward, P.N., Miles, A.J., Sumner, I.G., Thomas, L.H., and Lax, A.J. (1998). Activity of the
mitogenic Pasteurella multocida toxin requires an essential C-terminal residue.
Infection and immunity 66, 5636-5642.
Wei, N., Serino, G., and Deng, X.W. (2008). The COP9 signalosome: more than a protease.
Trends in biochemical sciences 33, 592-600.
Wiersinga, W.J., van der Poll, T., White, N.J., Day, N.P., and Peacock, S.J. (2006).
Melioidosis: insights into the pathogenicity of Burkholderia pseudomallei. Nature
reviews Microbiology 4, 272-282.
Wilson, B.A., Zhu, X., Ho, M., and Lu, L. (1997). Pasteurella multocida toxin activates the
inositol triphosphate signaling pathway in Xenopus oocytes via G(q)alpha-coupled
phospholipase C-beta1. The Journal of biological chemistry 272, 1268-1275.
Yao, Q., Cui, J., Wang, J., Li, T., Wan, X., Luo, T., Gong, Y.N., Xu, Y., Huang, N., and Shao,
F. (2012). Structural mechanism of ubiquitin and NEDD8 deamidation catalyzed by
bacterial effectors that induce macrophage-specific apoptosis. Proceedings of the
National Academy of Sciences of the United States of America 109, 20395-20400.
Yao, Q., Cui, J., Zhu, Y., Wang, G., Hu, L., Long, C., Cao, R., Liu, X., Huang, N., Chen, S.,
et al. (2009). A bacterial type III effector family uses the papain-like hydrolytic activity
88 to arrest the host cell cycle. Proceedings of the National Academy of Sciences of the
United States of America 106, 3716-3721.
Yarbrough, M.L., and Orth, K. (2009). AMPylation is a new post-translational modiFICation.
Nature chemical biology 5, 378-379.
Zhang, L., Krachler, A.M., Broberg, C.A., Li, Y., Mirzaei, H., Gilpin, C.J., and Orth, K. (2012).
Type III effector VopC mediates invasion for Vibrio species. Cell reports 1, 453-460.
Zhu, M., Shao, F., Innes, R.W., Dixon, J.E., and Xu, Z. (2004). The crystal structure of
Pseudomonas avirulence protein AvrPphB: a papain-like fold with a distinct
substrate-binding site. Proceedings of the National Academy of Sciences of the
United States of America 101, 302-307.
Zywietz, A., Gohla, A., Schmelz, M., Schultz, G., and Offermanns, S. (2001). Pleiotropic
effects of Pasteurella multocida toxin are mediated by Gq-dependent and independent mechanisms. involvement of Gq but not G11. The Journal of biological
chemistry 276, 3840-3845.
89 Chapter 3
P. syringae type III effector HopAF1 suppresses plant immunity by targeting the
methionine recycling pathway
Preface
The following chapter will be submitted for publication at the completion of some
additional experiments currently being performed by myself and our collaborator, Mark
Banfield, at the John Innes Centre in Norwich, UK. We aim to submit this paper for
publication in the next 2-3 months. Shahid Mukhtar performed the yeast two-hybrid screens
and pairwise tests in which the initial interaction between HopAF1 and MTN1 was identified.
I generated data for the rest of the figures, with the help of many resources, including the
UNC Biology Imaging Core Facility, the UNC Macromolecular Interactions Facility, and the
UNC Center for Structural Biology, along with input from members of the Dangl-Grant lab. In
addition, I wrote the manuscript under the direction of my advisor, Jeff Dangl.
I generated tools during the completion of this chapter that have contributed to other
lab projects. I generated two Gateway-compatible vectors for estradiol-inducible expression
of proteins tagged C-terminally with either citrine-HA or cerulean-HA. The former vector was
published by Akimoto-Tomiyama et al., (Akimoto-Tomiyama et al., 2012) and I was included
as 4th author on this paper for my contribution. pMDC7-cerulean-HA was also used in a
publication (Roberts et al., 2013). Finally, a plasma membrane localization marker that I
generated for confocal microscopy experiments, PLC2-YFP, contributed to work published
in Gao et al (Gao et al., 2011).
Abstract
Many plant pathogens, including Pseudomonas syringae, encode the type III
secretion system for translocating effector proteins into the host during infection. Strains of
P. syringae which are not capable of delivering the type III effectors are nonpathogenic.
Therefore, the functions of type III effectors are essential for disease. Although several type
III effectors have been demonstrated to block components of the plant defense response,
the functions of most type III effectors during disease manifestation are largely unknown.
Our lab is interested in the type III effector HopAF1, a type III effector that is present in
eleven of the nineteen sequenced strains of P. syringae and other plant pathogens.
Although the presence of HopAF1 in multiple strains of P. syringae suggests that it plays an
important role in virulence, no function has yet been associated with HopAF1. We began to
determine the function of HopAF1 with structure/function analyses. Using the BioInfoBank
Meta Server and HHPred structural prediction programs and homology searches, we
generated a tertiary structure model for HopAF1. These predictions suggest with high
confidence that HopAF1 is structurally related to bacterial deamidase proteins. Deamidation,
the irreversible substitution of an amide group with a carboxylate group, is the mechanism
by which several bacterial virulence factors manipulate the activity of a specific substrate. To
identify a potential target of the HopAF1 putative deamidation activity, we employed a
stringent yeast two-hybrid screen and identified Arabidopsis methylthioadenosine
nucleosidase (MTN1) as a putative target of HopAF1. This interaction, which we extended to
include Arabidopsis MTN2, was confirmed in planta using coimmunoprecipitations and
bimolecular fluorescence complementation assays. MTNs are enzymes in the methionine
recycling cycle, which is a cycle essential for high levels of ethylene biosynthesis. Ethylene
is a key plant hormone for plant developmental processes, such as senescence and
flowering. Ethylene is also induced during PTI. Therefore, we hypothesized that HopAF1
inhibits PTI by manipulating MTNs and levels of ethylene in plants. To this end, we used gas
91 chromatography to measure ethylene biosynthesis in plants treated with bacteriallydelivered HopAF1. We determined that HopAF1 inhibits ethylene biosynthesis in a manner
dependent on putative catalytic residues. Additionally, Yang cycle mutants mtn1 mtn2 and
mtk are more susceptible to weak pathogens. This data is consistent with the idea that
HopAF1 targets a novel component of the plant immune system, the Yang cycle. Additional
work on this project includes determining the enzymatic activity by which HopAF1 inhibits
MTN1 and MTN2.
Introduction
P. syringae is a Gram negative bacterium that causes disease in agronomically
important crops such as tomatoes, tobacco, and beans and the model plant, Arabidopsis
thaliana. P. syringae uses the type III secretion system (T3SS) to inject type III effector
(T3E) proteins into the host to cause disease in plants. P. syringae mutants lacking the type
III secretion system lack the ability to cause disease in hosts, meaning that the type III
secretion system is its main virulence determinant (Lindgren et al., 1986).
The immune response that plants use to protect themselves against pathogens
consists of multiple stages. The first line of defense begins with transmembrane pattern
recognition receptors (PRRs) perceiving conserved molecular patterns from pathogens
called pathogen-associated molecular patterns (PAMPs). Because these microbial
signatures are not limited to pathogenic microbes, they are often referred to as microbialassociated molecular patterns or MAMPs. For the purposes of this chapter, we will use the
terms PAMPs. PAMPs are often highly conserved across a wide range of microbes, fulfill a
function essential to the pathogen’s lifecycle, and yet are not present in the host (Boller and
Felix, 2009). Common bacterial PAMPs are lipopolysaccharide, peptidoglycan, elongation
factor Tu (EF-Tu), and flagellin (Boller and Felix, 2009). A 22 amino acid peptide derived
from a conserved region of the amino terminus of bacterial flagellin (flg22) is sufficient for
recognition by host PRR, FLS2 (Felix et al., 1999; Gomez-Gomez and Boller, 2000). This
92 leads to a downstream signaling cascade that includes many components such as an
increase in callose deposition, up-regulation of defense gene expression and an increase in
ethylene biosynthesis.
The second line of the plant immune system is activated by the recognition of
pathogen effectors by plant disease resistance gene products. Plants use disease
resistance (R) proteins, which often referred to as NB-LRR, or NLR, proteins due to their
domain architecture to detect effectors from a wide variety of plant pathogens. The
recognition of the presence of the effector, or more commonly the effect of an effector on a
specific host protein, that leads to effector-triggered immunity or ETI. An enhanced and
accelerated cell death response called the hypersensitivity response (HR) is an easily visible
phenotype often associated with ETI (Jones and Dangl, 2006).
Plant hormones are also key components of the plant immune response. There is
significant cross-talk between the three main stress hormones in plants, salicylic acid (SA),
jasmonic acid (JA), and ethylene that allow the plant to regulate its response to various
classes of pathogen and herbivores. As such, recently it has become clear that the
manipulation of plant hormone homeostasis is an important target of T3Es. For example, he
P. syringae T3E HopI1 suppresses SA accumulation (Jelenska et al., 2007). Additionally,
AvrPtoB stimulates ABA biosynthesis and ABA responses that antagonize SA biosynthesis
and SA-mediated defenses (de Torres-Zabala et al., 2007). AvrRpt2 alters the auxin
physiology to promote host susceptibility (Chen et al., 2007). XopD, a Xanthomonas
campestris pv. vesicatoria T3E, was recently found to target a tomato ethylene response
factor (ERF) called SIERF4. SIERF4 mutants display reduced ethylene levels and an
increase in susceptibility to Xcv (Kim et al., 2013), supporting the hypothesis that ethylene is
involved in plant defense.
Ethylene is a key component of the PTI response. Ethylene accumulation upregulates the expression of pathogenesis-related (PR) proteins and increases the
93 accumulation of hydroxyproline-rich, cell-wall strengthening proteins (Ecker and Davis,
1987; Penninckx et al., 1998). Ethylene also increases SA-induced expression of PR-1
(Lawton et al., 1994). Furthermore, ethylene directly up-regulates the expression of FLS2.
FLS2 transcription is directly controlled by binding of EIN3/EIL1/EIL2 to its promoter and in
ethylene insensitive mutants (ein2) seedlings, FLS2 expression, and FLS2 protein
accumulation are suppressed. As a result, ein2 plants are more susceptible to weakened
strains of P. syringae (Boutrot et al., 2010; Mersmann et al., 2010).
HopAF1 was identified as a T3E in P. syringae pv. tomato DC3000 (Pto DC3000)
and P. syringae pv. phaseolicola 1448A race 6 via its HrpL-dependent transcriptional
induction, a loosely defined N-terminal translocation signal sequence, and the ability to be
translocated into plant cells (Chang et al., 2005). HopAF1 is confirmed to be present in 11 of
19 strains of P. syringae (Baltrus et al., 2011). There are also HopAF1 orthologs in other
plant pathogens, such asAvrXv3 from Xanthomonas campestris pv. vesicatoria (AstuaMonge et al., 2000), Ralstonia solanacearum, Pseudomonas savastoni, and Aave_1373
from Acidovorax citrulli (Studholme et al., 2010). The presence of HopAF1 in multiple strains
of P. syringae and other plant pathogens suggests an important role in virulence.
The goal of this study was to determine the role of HopAF1 in P. syringae infections.
We found that the C-terminal domain of HopAF1 is predicted to be structural homologous to
bacterial deamidases, members of the papain-like superfamily. The C-terminus of HopAF1
also contains conserved putative catalytic residues in the predicted catalytic fold. HopAF1
inhibits PTI responses in a manner dependent on these putative catalytic residues. HopAF1
inhibits high levels of ethylene biosynthesis, also in a manner dependent on its putative
catalytic activity. HopAF1 interacts with Arabidopsis methylthioadenosine nucleosidases
(MTN) at the plant plasma membrane. These Yang cycle proteins are required for recycling
methionine to supplement de novo methionine biosynthesis. We demonstrated that the
Yang cycle is required during high levels of ethylene biosynthesis during pathogen attack. In
94 sum we demonstrated that HopAF1 is a key virulence factor of P. syringae and also
demonstrated the role of the Yang cycle and ethylene in the PTI.
Results
HopAF1 is a highly conserved type III effector with a putative C-terminal catalytic domain.
To determine if the structure of HopAF1 might suggest an enzymatic activity, we
used the BioInfoBank metaserver and HHPred structural prediction programs to generate a
structure prediction for HopAF1 from P. syringae pv. tomato DC3000 (Bujnicki et al., 2001;
Soding et al., 2005). The predicted structure of the full length HopAF1 protein did not have
homology to any known protein structure. However, when the C-terminus of HopAF1
(residues 130-284) was queried, the homology searches predicted with high confidence that
the HopAF1 C-terminus is structurally related to CheD, a single-domain chemotaxis protein
encoded by Thermotoga maritima (Figure 3.1A) (Chao et al., 2006). We validated our
HopAF1 model using Verify3D (Eisenberg et al., 1997).
We also generated structural models for all HopAF1 family members individually and
found that each, except for the Acidovorax allele which has a C-terminal truncation, contains
a predicted secondary structure with homology to CheD (Figure 3.1B). We specifically
focused on the divergent allele AvrXv3 from Xanthomonas campestris, which triggers HR in
tomato and pepper plants (Astua-Monge et al., 2000). Although HopAF1 and AvrXv3 are
only 27% identical, structural modeling of AvrXv3 also predicts that this protein has
structural homology to CheD (Figure 3.1C).
CheD is a deamidase that activates methyl-accepting chemotaxis proteins (MCPs)
(Chao et al., 2006). CheD is structurally related to the Cytotoxic Necrotizing Factor (CNF)
family of bacterial deamidases (Flatau et al., 1997; Schmidt et al., 1997). Deamidation is the
irreversible mutation of asparagine and glutamine residues to aspartic acid and glutamic
acid, respectively. Deamidation is a posttranslational modification utilized by multiple
bacterial virulence factors to modify the function of eukaryotic host target proteins
95 (Washington et al., 2013). The catalytic domains of all known bacterial deamidases contain
homology to papain-like structures (Washington et al., 2013). Therefore, we hypothesized
that the HopAF1 orthologs possess enzymatic activities associated with the papain-like
superfamily of enzymes, potentially deamidation. Bacterial deamidases require a histidine
and cysteine residue in their catalytic pocket for activity. We demonstrated that the C-termini
of both HopAF1 and AvrXv3 are predicted to form a catalytic pocket in which the invariant
catalytic residues for deamidation, cysteine and histidine, are present (Figure 3.1A,B).
Interestingly, these residues are conserved among all known HopAF1 orthologs, including
the divergent family members such as AvrXv3 (Figure 3.2). Therefore, we predicted that
these residues are likely catalytic and critical for the enzymatic function of HopAF1.
The role of the third catalytic residue of deamidases is to coordinate the imidazole
ring of histidine to ensure optimal orientation of the thiolate-imidazolium pair; therefore these
residues are required for function. The third residue of the CheD catalytic triad is T27 (Figure
3.1A). There is no threonine at this position in HopAF1 family members. However, adjacent
to the position of T27, there is a conserved aspartic acid, D154, in HopAF1 (Figure 3.1A).
We hypothesize that this residue may be involved in catalytic activity. Unfortunately,
structural modeling predicts this residue is present on a flexible loop and therefore, we are
not confident about the exact position of D154. Since the cysteine and histidine are
invariant, we did not mutate the predicted third catalytic residue of HopAF1. We simply
focused on the invariant putative catalytic residues, cysteine and histidine, in further
analyses of HopAF1 function.
HopAF1 blocks flg22-induced PTI and this requires its putative catalytic residues.
Many T3Es block PAMP-triggered immunity (PTI) and thereby contribute to bacterial
virulence (Block and Alfano, 2011). To determine if HopAF1 plays a role in virulence, we
obtained a Pto DC3000 derivative in which the HopAF1 gene has been disrupted (Jamir et
al., 2004). We did not observe a significant growth defect of Pto DC3000 in the absence of
96 HopAF1 (data not shown), therefore we decided to see if HopAF1 could enhance the growth
of a weak pathogen. To this end, we generated wild-type HopAF1 transgenic lines in the
wild-type Arabidopsis Col-0 background. We also generated the putative catalytic dead
version of the HopAF1 transgenic lines, by mutating both the C159 and H186 to alanine
(HopAF1C159AH186A). Conditional expression of HopAF1 wild-type and HopAF1C159AH186A was
evident by 6 hours after estradiol induction and was maintained for 48 hours.
The removal of 28 T3Es by sequential mutagenesis from Pto DC3000 transformed
this virulent strain into a weakly pathogenic derivative called Pto DC3000D28E (Cunnac et
al., 2011). To determine if HopAF1 inhibits PTI, we measured the growth of Pto
DC3000D28E in our transgenic lines 24 hours after induction of HopAF1 expression by
estradiol treatment. We hand inoculated Pto DC3000D28E into Col-0 and the two
independent HopAF1 transgenic lines for both wild-type HopAF1 and HopAF1C159AH186A. We
observed enhanced growth of Pto DC3000D28E on HopAF1-expressing transgenic lines
compared to Col-0, demonstrating that HopAF1 expression allowed for the enhanced growth
of a weak pathogen when overexpressed in planta (Figure 3.3A). The increase in bacterial
growth in the transgenic lines correlates with an increase in disease symptoms (Figure
3.3B). Importantly, the growth of Pto DC3000D28E on transgenic lines expressing the
putative catalytic dead HopAF1 was similar to that on Col-0, indicating that the enhanced
growth of a weak pathogen on HopAF1 transgenic lines was dependent on the putative
catalytic residues of HopAF1 (Figure 3.3A,B). The expression of HopAF1 in these lines was
confirmed by western blots (Figure 3.3C).
The pretreatment of Arabidopsis plants with flg22 induces PTI and diminishes
susceptibility to subsequent infection with virulent pathogens, such as Pto DC3000 (Zipfel et
al., 2004). To further test our hypothesis that HopAF1 inhibits PTI, we examined flg22induced disease resistance in Col-0 and the HopAF1 transgenic lines. After flg22 treatment,
we observed reduced growth of Pto DC3000 on Col-0 compared to untreated plants (Figure
97 3.4A). This effect was suppressed in transgenic Col-0 plants expressing HopAF1, but not in
transgenic Col-0 plants expressing the HopAF1 putative catalytic dead protein (Figure 3.4A).
Because Pto DC3000 is a highly virulent pathogen, HopAF1 transgenic plants are not more
susceptible to Pto DC3000 than Col-0. Expression of HopAF1 in the transgenic lines in
these experiments was also confirmed (Figure 3.4B). We conclude that HopAF1 suppresses
flg22-induced PTI in a manner dependent on its putative catalytic residues.
HR triggered by AvrXv3 in tomato is also dependent on its putative catalytic residues.
AvrXv3 is recognized by, and elicits an HR-response in, tomato cultivar Florida 216
(Fla 216) but not Florida 7060 (Fla 7060) (Astua-Monge et al., 2000). In order to determine if
the putative catalytic residues of AvrXv3 are required for recognition and HR, we generated
clones that allowed the constitutive overexpression of wild-type AvrXv3 and the AvrXv3
putative catalytic residue mutant, AvrXv3H126A, with a C-terminal myc tag. We transiently
expressed AvrXv3 wild type and mutant in both tomato cultivars via Agrobacterium mediated
T-DNA transfer, we measured their expression by western blot, and scored for HR. Wildtype AvrXv3, but not AvrXv3H126A, triggered HR on Fla 216 (Figure 3.5A). As expected, we
did not observe HR on Fla 7060 (data not shown). All clones were expressed (Figure 3.5B).
These data suggest that the recognition of AvrXv3 in tomato requires its putative catalytic
residues.
HopAF1 is targeted to the plasma membrane via acylation.
The N-termini of multiple HopAF1 orthologs contain putative sites for myristoylation
and palmitoylation (Figure 3.2). Myristoylation is a eukaryotic-specific posttranslational
modification that is used often in combination with palymitoylation to target proteins stably in
the plasma membrane lipid bilayer (Resh, 2006). Multiple P. syringae type III effectors utilize
these modes of acylation to localize to the plasma membrane once injected into the plant
cell (Dowen et al., 2009; Lewis et al., 2008; Nimchuk et al., 2000; Robert-Seilaniantz et al.,
2006; Shan et al., 2000).
98 We predicted that HopAF1 G2 and C4 are myristoylated and palmitoylated,
respectively, to target HopAF1 to the plasma membrane. To test this hypothesis, we
generated an acylation minus HopAF1 variant, HopAF1G2AC4S. For the HopAF1 orthologs
that lack the putative acylation sequence, such as AvrXv3, we predicted that they will
localize to the cytoplasm (Figure 3.2). We used confocal microscopy to detect the
localization of transiently expressed estradiol-inducible HopAF1 and AvrXv3 in N.
benthamiana. Both proteins were tagged on the C-terminus with a cerulean-HA tag. To
avoid aberrant protein localization, we determined the lowest amount of Agrobacterium and
estradiol necessary to detect HopAF1 via western blot (Figure 3.6A). We also confirmed that
we were assaying the localization of HopAF1 and not free, soluble cerulean (Figure 3.6B).
We used a plasma membrane marker, PLC2-YFP, as a plasma membrane localization
control (Otterhag et al., 2001). HopAF1 is localized at the plasma membrane and colocalizes with PLC2 (Figure 3.7A). However, the acylation minus variant was localized in the
plant cell cytoplasm, where it co-localized with free YFP, a marker for nucleocytoplasmiclocalized proteins (Figure 3.8A,B,C). Similar to the acylation minus HopAF1 variant, AvrXv3
is found in the cytoplasm (Figure 3.7A). Both of the putative catalytic dead variants,
HopAF1H186A and AvrXv3H126A, remained localized to the plasma membrane, suggesting that
the putative catalytic activity of HopAF1 orthologs is not required for localization (Figure
3.7A).
To confirm the confocal data, we performed cell fractionation assays with N.
benthamiana tissues transiently expressing the same constructs described above. The
control proteins, anti-APX and anti-H+-ATPase were used as controls for the soluble and
membrane fractions, respectively, and confirm that the separation of the fractions was
complete. Wild-type HopAF1 is detected in the microsomal fraction, which contains the
plasma membrane, whereas the acylation minus variant of HopAF1 is solely in the soluble
fraction (Figure 3.7B). AvrXv3 is also in the soluble fraction. Consistent with the confocal
99 results, HopAF1H186A and AvrXv3H126A remained localized to the microsomal and soluble
fractions, respectively (Figure 3.7B). We conclude that HopAF1 is targeted to the plant
plasma membrane via the acylation signal sequence and mutation of the putative catalytic
residues does not affect this localization.
HopAF1 associates with Arabidopsis MTN proteins at the plasma membrane.
We identified Arabidopsis methylthioadenosine nucleosidase (AtMTN1; At4g38800)
as a protein that interacts with HopAF1 in a yeast two-hybrid screen (Mukhtar et al., 2011).
This interaction was confirmed with pairwise yeast two-hybrid assays (data not shown).
Methylthioadenosine nucleosidases function in the methionine recycling pathway, also
known as the Yang cycle (Figure 3.9A) (Miyazaki and Yang, 1987). In Arabidopsis, two
proteins, MTN1 and MTN2 (67% identical), are responsible for methylthioadenosine
nucleosidase activity (Park et al., 2009). The Yang cycle generates high levels of methionine
to supplement de novo synthesis of methionine. This process is essential when levels of
methionine are limiting (Burstenbinder et al., 2007). This may be the case, for example,
during PTI. PTI results in the induction of high levels of ethylene, and methionine is the
precursor for ethylene biosynthesis. Therefore, if de novo biosynthesis of methionine does
not produce sufficient levels of methionine, the Yang cycle may be required for the PTItriggered induction of ethylene biosynthesis. We hypothesized that disrupting the function of
the Arabidopsis MTNs would block ethylene biosynthesis induced by PTI. Additionally, we
predicted that the virulence function of HopAF1 was to inhibit MTN function and block
ethylene-dependent PTI. This hypothesis is consistent with the finding that the type III
secretion system deficient mutant, Pto DC3000 hrcC, triggers the accumulation of high
levels of ethylene; whereas Pto DC3000 does not (Figure 3.9B). This suggests that the one
or more members of the suite of T3Es from Pto DC3000 can suppress ethylene
accumulation.
100 To extend the yeast two-hybrid interaction, we performed co-immunoprecipitation
studies using Agrobacterium-mediated transient expression assays in N. benthamiana
leaves. We co-expressed myc-tagged HopAF1 with the putative targets, MTN1-HA or
MTN2-HA. The cell lysates were incubated with anti-myc conjugated beads. Bound MTN
proteins were then detected with anti-HA westerns. We detected both MTN1-HA and MTN2HA interacting with HopAF1-myc (Figure 3.10A). Furthermore, HopAF1H186A-4x-myc coimmunoprecipated with MTN1-HA.The negative control demonstrates that there is very little
non-specific binding of MTN1-HA to the anti-myc beads (Figure 3.10A). We conclude that
HopAF1 also interacts with Arabidopsis proteins MTN1 and MTN2 when transiently
overexpressed in N. benthamiana.
Because HopAF1 is targeted to the plant plasma membrane, we predicted the
plasma membrane to be the site of interaction with MTNs. This is consistent with literature
showing that MTN1 can be found in the plant plasma membrane fraction and may interact
with CBL3 in the plasma membrane (Marmagne et al., 2007; Oh et al., 2008). Therefore,
MTN1 and MTN2 may be transiently tethered to the plant plasma membrane via proteinprotein interactions.
We used confocal microscopy to determine if HopAF1 and MTN1 could interact at
the plasma membrane. We performed bimolecular fluorescence complementation (BiFC)
experiments using a Gateway BiFC vector set with ubiquitin promoters providing low but
detectable levels of expression to minimize aberrant interactions due to over-expression
(Grefen et al., 2010). We generated constructs with MTN1 and MTN2 fused at their Ctermini to the C-terminal half of YFP. HopAF1 variants were fused at the C-terminus to the
N-terminal half of YFP. As a negative control, we assayed whether HopAF1 associated with
known plasma membrane protein PLC2 (Otterhag et al., 2001). Reconstitution of YFP,
resulting in a fluorescent signal emitting light at 514nm, occurred at the plasma membrane
when wild-type HopAF1-nYFP was co-expressed in N. benthamiana cells with either MTN1-
101 cYFP or MTN2-cYFP (Figure 3.10B). However, there was no reconstitution of YFP when
HopAF1-nYFP was co-expressed with PLC2-cYFP (Figure 3.10B). Reconstitution at the
plasma membrane was not dependent on the putative catalytic activity of HopAF1 (Figure
3.10B). Consistent with the presence of Arabidopsis MTNs in the cytoplasmic fraction
(Figure 3.5B, above), the acylation minus HopAF1G2AC4S-nYFP variant co-expressed with
MTN1-cYFP or MTN2-cYFP reconstituted the YFP signal, but in the cytoplasm (Figure
3.10B). We conclude that wild-type HopAF1 and Arabidopsis MTNs interact specifically at
the plasma membrane, consistent with the demonstrated localization of HopAF1, and the
observation that MTN1 and MTN2 may transiently localize to the plasma membrane via
protein-protein interactions (Marmagne et al., 2007; Oh et al., 2008).
HopAF1 inhibits PAMP-induced ethylene production in a manner dependent on its
putative catalytic activity and MTN proteins.
Recognition of flg22 triggers activation of a MAPK cascade that leads to an increase
in ethylene biosynthesis (Felix et al., 1991a; Felix et al., 1991b; Liu and Zhang, 2004). In
order to determine the effect of HopAF1 on PAMP-induced ethylene biosynthesis, we
established a system with which we could both induce ethylene biosynthesis with PAMPs
and measure the effect of HopAF1 on subsequent ethylene biosynthesis.
To induce PAMP-mediated ethylene production, we used a P. fluorescens strain
(Pfo-1) lacking known T3Es but engineered to carry a functional T3SS, as a delivery vehicle
for PAMPs, and, if desired, specific T3Es (Thomas et al., 2009). Pfo-1 triggers an increase
in ethylene production (Figure 3.11A). We transformed HopAF1 variants expressed from the
native promoter into Pfo-1. All HopAF1 alleles are expressed equally in Pfo-1 and are
capable of translocation via the type III secretion system, as assayed via their ability to
deliver truncated ∆79’avrRpt2 to trigger an RPS2-dependent HR (Figure 3.11B,C) (Mudgett
and Staskawicz, 1999).
102 HopAF1 inhibited the PAMP-induced ethylene biosynthesis compared to Pfo-1 and
empty vector controls (Figure 3.12A). Inhibition of ethylene production required the HopAF1
putative catalytic residue H186, supporting our hypothesis that HopAF1 inhibits MTN
function with catalytic activity (Figure 3.12B). However, the inhibition of PAMP-induced
ethylene production was not dependent on HopAF1 localization in the plasma membrane
(Figure 3.12B). This is consistent with our finding (Figure 3.10B), and literature (Marmagne
et al., 2007; Oh et al., 2008) showing that MTN1 and MTN2 localized to both the plasma
membrane and the cytoplasm.
HopAF1 phenocopies Yang cycle intermediate mutants for PAMP-induced increase in
ethylene production.
We predicted that HopAF1 suppression of PAMP-induced ethylene production would
be phenocopied in mutants of Yang cycle intermediates. As noted above, there are two
MTN paralogs in Arabidopsis, and it is likely that they function redundantly in the Yang cycle
(Figure 3.9A). MTN activity is required for maximal ethylene biosynthesis in tomatoes
(Kushad et al., 1985). The enzyme downstream of MTN in the Yang cycle, MTK, is also
required for maximal levels of ethylene biosynthesis (Burstenbinder et al., 2007). We
obtained Arabidopsis mtn1 and mtn2 mutants and generated anew the mtn1 mtn2 double
mutant (Burstenbinder et al., 2010). We demonstrated that Pfo-1 induced high levels of
ethylene in each mtn1 and mtn2 single mutant, consistent with these proteins functioning
redundantly for Yang cycle-mediated PAMP-induced ethylene biosynthesis (Figure 3.13A).
However, in both the mtk single mutant and the mtn1 mtn2 double mutant, Pfo-1 was unable
to induce high levels of ethylene production, demonstrating that the Yang cycle is necessary
for PAMP-induced ethylene production in this assay (Figure 3.13B). These results are
consistent with our suggestion that HopAF1 can inhibit PAMP-induced ethylene biosynthesis
through modulating or inhibiting the functions of Arabidopsis MTN1 and MTN2.
103 HopAF1 inhibits MTN1 activity in vitro in manner dependent on its putative catalytic
residues.
We used recombinant proteins purified from E. coli to determine if HopAF1 can
inhibit MTN1 function directly. A protocol for purifying Arabidopsis MTN proteins from E. coli
was previously established (Park et al., 2009; Park et al., 2006). To purify HopAF1, we
generated N-terminal 6xHis fusion constructs for wild-type and putative catalytic dead
HopAF1. Because the coding sequence of HopAF1 likely contains subdomains, we also
generated 6xHisMBP-tagged truncations of HopAF1 to identify the minimal region of
HopAF1 required for activity (Figure 3.14A). We used circular dichroism to confirm that
MTN1, 6xHis-HopAF1, and 6xHis-HopAF1H186A are folded (Figure 3.14B).
Using an established assay to measure the activity of MTN1 in vitro (Dunn et al.,
1994), we demonstrated that HopAF1 inhibited the ability of MTN1 to cleave its substrate
(Figure 3.15A,B). However, HopAF1H186A was not able to inhibit MTN1 activity (Figure
3.15B). Furthermore, the proposed catalytic domain of HopAF1 (residues 130-284) is
sufficient to inhibit MTN1 function (Figure 3.15B). The ∆147HopAF1 truncated protein was
not able to inhibit MTN1 function, suggesting that an essential part of the HopAF1 structure
exists between residues 130 and 147 (Figure 3.15B). These in vitro data support our in
planta findings that HopAF1 inhibits PAMP-induced ethylene biosynthesis by blocking MTN
function.
MTN1 and MTN2 play a role in innate immunity.
To determine the role of MTNs in PTI, we assayed the susceptibility of mtn1 mtn2
and mtk to infection with Pto DC3000D28E. As a control, we used the ein2 (Ethylene
Insensitive 2) null mutant (Alonso et al., 1999). Ethylene insensitive mutants allow enhanced
growth of weak pathogens and are defective in flg22-mediated priming of the immune
system (Boutrot et al., 2010; Mersmann et al., 2010). We show that Yang cycle mutants,
similar to ein2, are more susceptible to Pto DC3000D28E (Figure 3.16). Additionally, the
104 enhanced growth of Pto DC3000D28E in mtn1 mtn2 was similar to levels of the same strain
observed in transgenic plants expressing wild-type HopAF1 (Figure 3.16).
Discussion
Many plant bacterial pathogens use type III effectors (T3Es) as their main virulence
determinants. The type III secretion system is essential to the ability of these plant
pathogens to cause disease on their hosts. As such, it is critical to understand which host
cellular systems T3Es are targeting and how they are disrupting those systems. Additionally,
we learn how these potent virulence factors, which are common among plant and animal
pathogens, function.
HopAF1 is a reasonably well conserved T3E. The presence of HopAF1 in multiple
strains of P. syringae and other plant pathogens suggests an important role in virulence and
we sought to define its specific molecular function. Here we report that HopAF1 contributes
to P. syringae virulence and we propose a model (Figure 3.17), in which HopAF1 interacts
with Arabidopsis Yang cycle proteins, MTN1 and MTN2, at the plant plasma membrane.
HopAF1 inhibits PAMP-induced ethylene biosynthesis and the mtn1 mtn2 phenocopies the
effect of HopAF1 on PAMP-induced ethylene biosynthesis. This is consistent with HopAF1
inhibition of MTN1 activity in vitro. We have also demonstrated that MTN1 and MTN2, and,
therefore the Yang cycle, are key components of the plant immune response.
The role of ethylene in the plant defense response is well-documented, yet
complicated because of the multiple, and sometimes conflicting, roles that ethylene plays.
The contradictory results from pathology assays might result from differences in bacterial
strains used and infection methods. Ethylene plays a role in defense against necrotrophs by
antagonizing salicylic acid responses which are involved in defense against biotrophs
(Robert-Seilaniantz et al., 2007). Ethylene is also required for symptom development during
infection with virulent and avirulent pathogens (Bent et al., 1992).
105 However, more recently, it has been demonstrated in multiple studies that ethylene
plays a significant role in defense against P. syringae. Ethylene directly controls FLS2
expression by inducing the activation of the EIN3/EIL family of transcription facts which
directly bind the FLS2 promoter (Boutrot et al., 2010). This leads to an increase in FLS2
protein accumulation (Boutrot et al., 2010; Mersmann et al., 2010). Mutants deficient in
ethylene signaling pathways, such as ein2, have reduced FLS2 transcript levels with or
without flg22 treatment (Boutrot et al., 2010). This likely contributes to the susceptibility of
ein2 to weakened plant pathogens such as P. syringae pv. tomato DC3000
∆AvrPto/∆AvrPtoB (Mersmann et al., 2010). Furthermore, XopD desumoylates and
represses an ethylene-responsive transcription factor and suppresses ethylene
accumulation (Kim et al., 2013). Ethylene is required for the immune response against
Xanthomonas campestris pv. vesicatoria (Kim et al., 2013).
The role of the Yang cycle in PTI responses is thus far undocumented. This is likely
due to the lack of the link between Yang cycle proteins and pathogen-induced ethylene
accumulation in the literature. The Yang cycle kinase, MTK, was demonstrated to be
required for accumulation of high level of ethylene by crossing mtk with an ethylene
overproducing mutant, eto (Burstenbinder et al., 2007). This suggested that the Yang cycle
is required for high levels of ethylene biosynthesis, such as that which occurs during PTI.
However, subsequent papers have stated that the MTN proteins are not required for
ethylene biosynthesis (Burstenbinder et al., 2010). Unfortunately, they did not investigate
ethylene accumulation in the double mtn1 mtn2 mutant; nor did they attempt to induce high
levels of ethylene accumulation (Burstenbinder et al., 2010). Therefore, the burden of proof
was on us to demonstrate the requirement of MTN proteins for high levels of ethylene
accumulation in Arabidopsis. This highlights the importance of the results in figure 13. We
conclude that the Yang cycle is not required for steady-state levels of ethylene biosynthesis.
106 However, our results consistently show that Yang cycle proteins, MTN1 and MTN2 or MTK
are required for the high levels of PAMP-induced ethylene biosynthesis.
However, there are still some questions that remain unanswered. For example, are
there other type III effectors from P. syringae pv. tomato DC3000 that inhibit ethylene
accumulation? In figure 9, we demonstrate that the type III secretion system deficient
mutant, Pto DC3000 hrcC, triggers the accumulation of high levels of ethylene; whereas Pto
DC3000 does not. Assuming there is an equal presence of PAMPs associated with these
pathogens, the likely conclusion is that the fully virulent Pto DC3000 is suppressing ethylene
accumulation. Incorporation of the Pto DC3000∆HopAF1 mutant into this assay will help us
discern whether or not other effectors are targeting ethylene accumulation. For example, if
Pto DC3000∆HopAF1 accumulates ethylene similar to Pto DC3000 hrcC, HopAF1 is likely
to be the only Pto DC3000 T3E that can inhibit ethylene biosynthesis. Conversely, if Pto
DC3000∆HopAF1 still inhibits ethylene accumulation similar to Pto DC3000, there are still
T3Es other than HopAF1 that can inhibit ethylene biosynthesis.
Although we have predicted, based on structural modeling, that HopAF1 is a
member of the family of bacterial toxins/effectors that are deamidases, we have not yet
demonstrated that HopAF1 uses this activity to affect the function of Arabidopsis MTNs or
any other proteins. We have, however, demonstrated that the ability of HopAF1 to block PTI,
inhibit MTN1 function in vitro, and inhibit PAMP-induced ethylene biosynthesis is dependent
on putative catalytic residues. This suggests that HopAF1 has enzymatic activity. We will
attempt to determine the enzymatic function of HopAF1 using mass spectrometry to identify
deamidation on MTN1 (see Chapter 4). Mass spec has been an invaluable tool in identifying
posttranslational modifications on host proteins due to pathogen virulence proteins, such as
type III effectors. However, identifying a 1 Dalton increase in the size of a target protein,
such as would occur during deamidation, with mass spectrometry is not trivial. For this
reason, we are collaborating with Mark Banfield, a biochemist at the John Innes Centre with
107 extensive experience with deamidases (Crow et al., 2012; Crow et al., 2009; Jubelin et al.,
2009).
Multiple P. syringae T3Es target PTI signaling pathways using diverse biochemical
mechanisms. In this study, we have described a new component of the plant innate immune
system, the Yang cycle. Furthermore, we will continue to determine whether a unique
biochemical mechanism, namely deamidation, is being utilized by the P. syringae T3E
HopAF1 to target Arabidopsis Yang cycle proteins, MTN1 and MTN2.
108 Materials and Methods
Plant Lines and Growth Conditions
We used wild-type Arabidopsis thaliana ecotype Columbia (Col-0) and all mutant
Arabidopsis lines we used are in the Col-0 background. For generation of transgenic
estradiol-inducible HopAF1-cerulean-HA, HopAF1C159AH186A-cerulean-HA lines, the HopAF1
coding sequences were cloned into the pMDC7-cerulean-HA Gateway-compatible vector
(Akimoto-Tomiyama et al., 2012) and transformed into Agrobacterium GV3101. Arabidopsis
transgenics were generated using Agrobacterium-mediated floral dip transformation (Clough
and Bent, 1998). Hygromycin selection of transgenic plants was performed by growing the
seedlings on plates with ½ MS media with 15 μM hygromycin in the dark for 3 days and then
in light conditions for 10 days. mtn1-1 (Burstenbinder et al., 2010), mtn1-2 (Burstenbinder et
al., 2010), mtn2-1 (Burstenbinder et al., 2010), mtn2-2 (Burstenbinder et al., 2010), mtn1-1
mtn2-1 (Burstenbinder et al., 2010), mtk (Sauter et al., 2004), and ein2-5 (Alonso et al.,
1999) have previously been described elsewhere. Except where otherwise noted, plants
were grow under controlled short day conditions (9 hours light, 21°C, 15 hours dark, 18°C).
Tomato cultivars Florida 216 (Fla 216) and Florida 7060 (Fla 7060) are described elsewhere
(Astua-Monge et al., 2000) and were grown in the greenhouse. Nicotiana benthamiana
plants were also grown in the greenhouse.
Immunoblot analysis of plant proteins
Unless otherwise noted, total proteins were extracted by grinding either fresh tissue or
tissue flash-frozen in liquid nitrogen in a buffer containing 50 mM Tris-HCl pH 8.0, 1% SDS,
1 mM EDTA pH 8.0, 14 mM β-mercaptoethanol, and protease inhibitor cocktail. Samples
were incubated on ice for 20 minutes and then the cell debris was pelleted by centrifuging
the samples for 15 minutes at 20,800 g at 4°C. The supernatant was separated and
quantified using the Bradford reagent. Protein was separated on 12% SDS-PAGE gels and
109 transferred to a nitrocellulose membrane. Western blots were performed using standard
methods. Anti-HA (Santa Cruz Biotechnology) antibody was used at a 1:5000 dilution, antiAPX (Agrisera) was used at a 1:5000 dilution, and anti-H+-ATPase (Agrisera) was used at a
1:1000 dilution. Anti-myc polyclonal antisera was used at a 1:5 dilution. Anti-MTN1 and antiMTN2 were used at a 1:5000 dilution (Burstenbinder et al., 2010).
Homology Modeling
We initially detected the homology with the C-terminal domain of HopAF1 and CheD by
querying the BioInfoBank Institut’s Metaserver and the HHPred structural prediction
programs (Bujnicki et al., 2001; Soding et al., 2005). The 3-D model for the HopAF1 Cterminus structure based on the Thermotoga maritima CheD (PDB 2F9Z) structure was
generated with MODELLER (Sali and Blundell, 1993). The validity of the HopAF1 structural
model was assessed using Verify3D (Eisenberg et al., 1997). Graphical images and
structural alignments were produced with Pymol.
P. syringae pv. tomato DC3000D28E Growth Curve Assay
Leaves of five-week-old plants were hand inoculated with 1x105 cfu/mL of P. syringae pv.
tomato DC3000D28E (Cunnac et al., 2011). Leaf bacterial populations were determined the
day of inoculation and three days post-inoculation. Each data point consisted of six
replicates. Plants expressing either estradiol-inducible HopAF1-cerulean-HA or estradiolinducible HopAF1C159AH186A-cerulean-HA were sprayed with 20 μM estradiol 24 hours before
hand inoculations.
Flg22-protection Growth Curve Assay
Leaves of five-week-old plants were sprayed with 20 μM estradiol 12 hours before flg22
infiltration. Plants were infiltrated with 1 μM flg22 or water 24 hours before infiltrating with
110 1x105 cfu/mL P. syringae pv. tomato DC3000. Leaf bacterial populations were determined at
the indicated times. Each data point consisted of six replicates.
Transient expression of AvrXv3 in tomato
The coding sequences of AvrXv3 and AvrXv3H126A were cloned in the binary Gateway
vector, pGWB17, which carries a C-terminal 4x-myc tag. Triparental mating was used to
transfer this construct to Agrobacterium strain C58C1. For transient expression of AvrXv3 in
tomato plants, the strains expressing AvrXv3 variants were grown overnight in LB media at
28°C. The bacteria were centrifuged at 6800 g for 5 minutes and resuspended in 10 mM
MgCl2 at an OD of 0.1. Fully expanded leaves were infiltrated with the bacterial strains and
maintained in the greenhouse for 48 hours (immunoblot samples) or 72 hours (HR).
Immunoblot analysis of protein samples was completed as previously described.
Cell fractionations
200 mg of tissue was flash-frozen in liquid nitrogen and ground in sucrose buffer (20 mM
Tris-HCl pH 8.0, 0.33M sucrose, 1 mM EDTA pH 8.0, 5 mM DTT, plant protease inhibitors
cocktail). The homogenate was centrifuged for 10 minutes at 2000 g at 4°C. The
supernatant was collected as the total fraction. A portion of the total fraction was spun again
for 30 minutes at 20,000 rpm at 4°C. The supernatant from this spin was collected as the
soluble fraction. The pellet, which became the microsomal fraction, was resuspended in
equal volumes of sucrose buffer.
Agrobacterium-mediated transient transformations in N. benthamiana
2 mL of overnight cultures were centrifuged at 10,600 g for 1 minute at room temperature.
Cultures were resuspended in induction media containing 10 mM MgCl2 and 10 mM MES.
Cultures were re-centrifuged and then resuspended in infiltration media plus 100 μM
111 acetosyringone and agitated at room temperature for 1 hour. Cultures were diluted to the
appropriate concentrations for each experiment and hand-infiltrated into fully expanded
Nicotiana benthamiana leaves.
Coimmunoprecipitation
For co-immunoprecipitation experiments we generated 35S::MTN1-HA, 35S::MTN2-HA,
35S::HopAF1-myc, 35S::HopAF1H186A-myc using the Gateway constructs pGWB14 (Cterminal 3x-HA tag) and pGWB17 (C-terminal 4x-myc tag). Constructs were infiltrated at
concentrations of 0.2 OD each with 0.1 OD P19 (Voinnet et al., 2003). Samples were
collected and flash-frozen in liquid nitrogen. Protein was ground in lysis buffer (20 mM TrisHCl pH 8.0, 1 mM EDTA pH 8.0, 150 mM NaCl, protease inhibitor cocktail), centrifuged for
15 minutes at 6000 g in 4°C. The supernatant was diluted to a concentration of 2 mg/mL.
50 μL of anti-myc beads was added to the protein sample and incubated at 4°C for 4 hours
with gentle agitation. The bound protein complexes were washed in extraction buffer (20
mM Tris-HCl pH 8.0, 1 mM EDTA pH 8.0, 300 mM NaCl) and eluted in pre-warmed 6X SDSPAGE loading buffer. The input and bound proteins were analyzed by SDS-PAGE and
immunoblot analysis.
Subcellular localization
HopAF1, HopAF1H186A, HopAF1G2AC4S, AvrXv3, and AvrXv3H126A coding sequences were
cloned into the Gateway-compatible pMDC7-cer-HA vector (Akimoto-Tomiyama et al., 2012)
and transferred into Agrobacterium strain C58C1 by triparental mating. Agrobacteriummediated transformations into Nicotiana benthamiana leaves were performed as described
above. Agrobacterium strains expressing HopAF1, HopAF1H186A, HopAF1G2AC4S, AvrXv3,
and AvrXv3H126A were inoculated at an OD of 0.01 in combination with the PLC2 plasma
membrane marker at 0.1 OD. Three days post-infiltration, leaves were sprayed with 5 μM
112 estradiol for 3 hours. The subcellular localization of the proteins was observed using a Zeiss
LSM 710 confocal scanning laser microscope. The plasma membrane marker, PLC2-YFP,
was generated by cloning the PLC2 ORF from Col-0 cDNA into the pGWB41 Gateway
vector.
Bimolecular fluorescence complementation
HopAF1 and MTN1 coding sequences were cloned into bimolecular fluorescence (BiFC)
constructs that contain an ubiquitin promoter (Gehl et al., 2009). HopAF1-nYFP,
HopAF1H186A-nYFP, HopAF1G2AC4S-nYFP were transiently co-expressed with MTN1-cYFP,
MTN2-cYFP, and PLC2-cYFP in N. benthamiana leaves using Agrobacterium-mediated
transient transformation. Strains expressing either HopAF1, MTN1, MTN2, or PLC2 were
inoculated at 0.4 OD with 0.1 OD of P19 (Voinnet et al., 2003). 3 days after infiltration,
confocal microscopy was used to image the reconstituted YFP signal.
HopAF1 purification
N-terminal 6xHis affinity tag constructs for HopAF1 and HopAF1H186A were generated by
ligation independent cloning (Aslanidis and de Jong, 1990). 6xHisMBP-∆130HopAF1 and
6xHisMBP-∆147HopAF1 were generated by cloning a TEV protease cleavage site onto the
N-terminus of the truncated HopAF1 ORFs and cloning them into the Gateway-compatible
pDEST-HisMBP vector.
All HopAF1 bacterial expression constructs were transformed into ArticExpress (DE3) RIL
(Agilent) cells. To initiate protein production, overnight cultures were used to inoculate 5 L of
LB media containing 100 μg/mL ampicillin at a 1:200 dilution. Cells were grown at 28°C until
the OD reached 0.6-0.8 and then the temperature was reduced to 14°C. Protein production
was induced with 1 mM IPTG and cells continued to grow overnight at 14°C.
113 Purification of 6xHis-HopAF1 and 6xHis-HopAF1H186A involved affinity purification on a
HisTrap column followed additional purification with the HiPrep 16/60 Sephacryl S-200 High
Resolution (GE Healthcare) size exclusion column. The HopAF1 fractions that were not in
the void volume were collected for in vitro assays.
Purification of 6xHisMBP-∆130HopAF1 and 6xHisMBP-∆147HopAF1 also began with a His
affinity purification. This was followed by dialysis into TEV cleavage buffer (20 mM Tris pH
8.0, 50 mM NaCl, 1 mM DTT, and 0.5 mM EDTA) and TEV cleavage at room temperature
for 6 hours. Reverse affinity chromatography was used to collect the HopAF1 truncations
into relatively pure fractions. These fractions were dialyzed into anion exchange buffer (20
mM Tris pH 8.0, 50 mM NaCl overnight at 4°C and further purified with the anion exchange
HiLoad 16/10 Q sepharose column (GE Healthcare). Size exclusion chromatography using
a HiPrep 16/60 Sephacryl S-200 High Resolution column (GE Healthcare) resulted in
monodispersed protein.
MTN1 purification
MTN1 (At4g38800) was purified as previously described (Park et al., 2009). In brief, a
PreScission Protease cleavage site was cloned onto the N-terminus of MTN1, which was
then cloned into the Gateway vector pDEST15, which carries an N-terminal glutathione Stransferase (GST) tag. This clone was transformed into Novagen® Rosetta E. coli
BL21(DE3) cells. A 5 mL overnight culture was used to inoculate 1L of LB media containing
100 μg/mL ampicillin. Cells were grown at 37°C to an OD of 0.6-0.8. The temperature was
reduced to 22°C. Protein production was induced with 1 mM IPTG and cells continued to
grow at 22°C overnight. MTN1 was purified with a GST affinity purification step followed by
114 PreScission Protease cleavage and reverse chromatography to isolate “free” MTN1 in the
flow through fractions.
MTN activity assays
MTN activity assays were performed as previously described (Dunn et al., 1994). Briefly,
reactions were carried out in 50 mM imidazole pH 8.0, 0.28 units of grade III buttermilk
xanthine oxidase, 1 mM 2-(4-iodophenyl)-3-(4-nitrophenyl)-5-phenyltetrazolium chloride
(INT) and 0.12 μg recombinant MTN1. The reaction was monitored at 470 nm on a Tecan
GENios microplate reader (Tecan, Ltd., Switzerland) for 15 minutes with 1 minute
increments. Changes in absorbance were converted to the amount of adenine released
using the molar absorption coefficient (15,400 M-1 cm-1 at pH 8.0). To generate the
Michaelis-Menten plot, the experiment was completed with a range of MTA concentrations
(0 μM MTA – 150 μM MTA). Specific activity assays were completed with 50 μM MTA. All
enzyme reactions had a final volume at 800 μL and were performed in triplicate at 25°C. All
reagents for the kinetic assay were purchased from Sigma-Aldrich Chemicals (St. Louis,
MO).
Ethylene measurements
Three-week-old seedlings were vacuum-infiltrated with Pfo-1 strains measured at 0D 0.04 in
10 mM MgCl2. After three hours, the fresh weight of the seedlings was measured. Seedlings
were sealed into gas chromatography vials and ethylene was allowed to accumulate for 24
hours. Ethylene accumulation was measured using a Clarus 500 gas chromatograph (Perkin
Elmer, USA). A pure ethylene gas control was used in each experiment. Each data point
represents 3 to 4 replicates.
115 Induction of native promoter HopAF1 in minimal media
Genomic clones of wild-type HopAF1, HopAF1H186A, and HopAF1G2AC4S, including the native
promoter with a hrp-box, were cloned into a Gateway-compatible vector, pJC531, carrying a
C-terminal HA-tag. This clone was transformed into P. fluorescens, a strain engineered to
express the T3SS (Thomas et al., 2009). Strains were spread onto KB plates overnight and
resuspended in minimal media (7.8 mM ammonium sulfate, 50 mM potassium phosphate,
1.7 mM sodium chloride, 1.7 mM NaCl and 10 mM mannitol) at an OD of 0.02. After growth
in 28°C for 12 hours, 2 mL of each cultures was centrifuged at room temperature for 5
minutes at 8600 g. The pellet was resuspended in 100uL of 6X SDS loading buffer, boiled
for 10 minutes, and 10 μL was loaded onto SDS-PAGE gel.
Circular dichroism
6xHis-HopAF1, 6xHis-HopAF1H186A, and MTN1 were purified as described above and
dialyzed into a buffer containing 20 mM sodium phosphate pH 7.4, 150 mM NaF. Samples
of approximately 0.4 mg/mL (HopAF1 variants) or 0.5 mg/mL (MTN1) were placed in a 1.0
mm cuvette. Molar ellipticity measurements were collected at 25°C from 185 nm to 260 nm
on a Chirascan Plus (Applied Photophysics, Ltd, Surrey, United Kingdom).
Generating the mtn1 mtn2 double mutant
In order to generate the mtn1 mtn2 double mutant, which is sterile (Burstenbinder et al.,
2010), we obtained and genotyped the previously described GABI Kat lines mtn1-2 and
mtn2-2 (Burstenbinder et al., 2010). We crossed mtn1-2 and mtn2-2, selfed the resulting
T1s and screened the T2s for lines which were homozygous for mtn1 and segregating mtn2.
We maintained several stocks of these lines, which we sowed out and PCR genotyped to
identify double homozygous plants at the time of experimentation.
116 Translocation Assay from Pfo-1 and trypan blue staining
Genomic HopAF1 constructs were cloned into the pJC532, a Gateway-compatible vector
carrying a C-terminal tag of the C-terminal 177 residues of the type III effector protein
AvrRpt2, from which the N-terminal 79 amino acids of AvrRpt2 were deleted (‘∆79AvrRpt2)
(Chang et al., 2005; Mudgett and Staskawicz, 1999). HopAF1-∆79AvrRpt2, HopAF1H186A∆79AvrRpt2, HopAF1G2AC4S-∆79AvrRpt2 were transferred into Pfo-1 using triparental
mating. Sixteen leaves of Arabidopsis Col-0 were hand-infiltrated per strain and the HR was
scored 26 hours post inoculation. The HR was compared to leaves infiltrated with Pfo-1
alone or Pfo-1 carrying an empty vector.
117 Figures
Figure 3.1. The predicted structure of HopAF1 exhibits structural homology to CheD,
a bacterial deamidase. A) Homology model of HopAF1 reference allele (blue) from P.
syringae pv. tomato DC3000 (Pto DC3000) residues 130-284 aligned with CheD (PDB
2F9Z) (silver). The side chains of the putative catalytic residues of HopAF1, C159, H186,
and D154, are highlighted in green. The side chains of the CheD catalytic residues are in
silver. B) Independent homology models for the additional HopAF1 family members: P.
syringae pv. maculicola M4 (teal), P. syringae pv. aceris (grey), P. syringae pv. lachrymans
str. M302278 (green), and Acidovorax citrulli Aave_1373 (pink). HopAF1 from Pto DC3000
is included in blue as a reference. C) Homology model of AvrXv3 from Xanthomonas
vesicatoria (magenta) aligned with CheD (silver). Putative catalytic residues, C105 and
H126, of AvrXv3 are highlighted in yellow and align with the catalytic residues of CheD
(silver).
.
118 Figure 3.2. Alignment of HopAF1 sequences from P. syringae pathovars and other
plant pathogens. Protein sequences from HopAF1 homologs from P. syringae pv. tomato
DC3000 (accession no. NP_791393), P. syringae pv. aceris strain M302273 (accession no.
WP_003402453), P.syringae pv. actinadae str. M02091 (accession no. WP_003381937), P.
syringae pv. lachrymans str. M302278 (Lac106) (accession no. WP_0057667510), P.
syringae pv.lachrymans str. M301315 (Lac107) (WP_005746129), P. syringae pv.
maculicola M4 str. ES4326 (accession no. WP_007248647), P. syringae pv. mori str.
301020 (accession no. WP_005752338), P. syringae pv. pisi (accession no.
WP_003346247), P. syringae pv. phaseolicola 1448A race 6 (accession no. YP_273699),
P.syringae pv. syringae B728A (accession no. YP_236881), P. syringae pv. tomato T1
(accession no. EEB59276), Pseudomonas savastoni (accession no. YP_006961546),
Ralstonia solanacearum str. Po82 (accession no. YP_006028103), Xanthomonas
campestris pv. muscacearum (accession no. WP_010369035), Xanthomonas vesicatoria
AvrXv3 (accession no. WP_008577605), and Acidovorax citrulli AAC00-1 Aave_1373
(accession no. YP_969738) were aligned using VectorNTI AlignX. Amino acids with 80%
identity are shaded in grey. Putative catalytic residues are highlighted in yellow. Sites for
myristoylation and palmitoylation in denoted in red and orange, respectively. Residues
highlighted in green highlight the beginning of the putative catalytic domain.
119 Figure 3.3. HopAF1 transgenic lines are susceptible to weak pathogens. A) Bacterial
growth of P.syringae pv. tomato DC3000D28E (Cunnac et al., 2011) in Col-0 and two
independent transgenic lines expressing either estradiol-inducible HopAF1-cerulean-HA or
estradiol-inducible HopAF1C159AH186A-cerulean-HA was measured 0 and 3 days after hand
inoculation with 1x105 cfu (colony-forming units)/mL bacteria. Transgene expression was
induced with 20 μM estradiol 24 hours prior to bacterial inoculation. Error bars represent
standard error (SEM). An analysis of variance (ANOVA) was performed among the day 3
samples, followed by Tukey’s post-hoc analysis (p < 0.05). Groups that differ significantly
are indicated with different letters. B) Disease symptoms in two independent lines of
HopAF1-cerulean-HA and HopAF1C159AH186A-cerulean-HA were photographed 3 days after
hand-inoculation with 1 x 105 cfu/mL of P. syringae pv. tomato DC3000D28E in HopAF1
transgenic lines. C) Immunoblot analysis with anti-HA shows the expression of HopAF1cerulean-HA and HopAF1C159AH186A-cerulean-HA in transgenic plants 3 days after spray
application of 20 μM estradiol.
120 Figure 3.4. Arabidopsis transgenic lines expressing wild-type HopAF1 inhibit flg22mediated protection against P. syringae pv. tomato DC3000. A) Col-0, estradiolinducible HopAF1-cerulean-HA, and estradiol-inducible HopAF1C159AH186A-cerulean-HA were
sprayed with 20 μM estradiol for 12 hours and then infiltrated with 1 μM flg22 or water for 24
hours prior to infiltration with 1x105 cfu/mL P. syringae pv. tomato DC3000. Bacterial growth
was determined 0 and 3 days after bacterial inoculation. Error bars represent SEM. An
ANOVA was performed among the day 3 samples, followed by Tukey’s post-hoc analysis (p
< 0.05). Groups that differ significantly are indicated with different letters. B) Immunoblot
analysis with anti-HA shows the expression of HopAF1-cerulean-HA and HopAF1C159AH186Acerulean-HA in transgenic plants 3 days after bacterial infiltration, which corresponds to 108
hours (or 4.5 days) after estradiol induction.
121 Figure 3.5. HR triggered by AvrXv3 is dependent on putative catalytic residues. A)
AvrXv3-4x myc and AvrXv3H126A-4x myc were transiently expressed in tomato cultivar Fla
216. Cores containing both infiltrated and uninfiltrated tissue where removed from the
tomato leaves 3 dpi and trypan blue staining was used to highlight the cell death response
associated with HR. B) Anti-myc immunoblot analysis demonstrates the expression of
AvrXv3-4x-myc and AvrXv3H126A-4x-myc in tomato cultivar Fla 216 dpi. Ponceau staining of
the membrane indicates equal loading.
122 Figure 3.6. Establishing a system to study localization of transiently expressed
HopAF1 in N. benthamiana. A) Levels of HopAF1-cerulean-HA transiently expressed in N.
benthamiana were titrated by varying the concentration of estradiol (μM) and OD of
Agrobacterium. Immunoblots with anti-HA were used to detect the expression of HopAF1cerulean-HA. N. benthamiana leaves were sprayed with 5 μM estradiol 3 days postinfiltration with Agrobacterium and leaf samples were collected 3 hours after estradiol
application for immunoblot analysis. B) Immunoblot analysis with anti-HA shows the
expression of full-length HopAF1-cerulean-HA, HopAF1H186A-cerulean-HA, and
HopAF1G2AC4S-cerulean-HA and the absence of free cerulean (black arrow) at approximately
27 kDa. N. benthamiana leaves were sprayed with 5 μM estradiol 3 days post-infiltration
with Agrobacterium. Protein samples were collected 3 hours after estradiol application.
Ponceau staining indicates equal loading.
123 Figure 3.7. HopAF1 is targeted to the plasma membrane via acylation. A,B) Estradiolinducible HopAF1-cerulean-HA, HopAF1G2AC4S-cerulean-HA, HopAF1H186A-cerulean-HA,
AvrXv3-cerulean-HA, and AvrXv3H126A-cerulean HA were expressed transiently in N.
benthamiana using Agrobacterium-mediated transient transformations. Transgene
expression was induced with 5 μM estradiol 3 days after bacterial infiltration. A) The
localization of transiently expressed proteins was determined with scanning confocal laser
microscopy 3 hours after estradiol induction in N. benthamiana epidermal cells. PLC2-YFP
is a marker for the plasma membrane localization. Scale bars represent 20 μm. White
arrowheads indicate the plasma membrane (PM), cytoplasmic streaming (CS), and nuclei
(N). B) Cellular fractionations into total (T), soluble (S), and microsomal (M) fractions were
performed. Equal cell equivalents were loaded onto SDS-PAGE gels followed by anti-HA
immunoblots to detect the fractions with the expressed proteins of interest. Anti-APX and
anti-H+-ATPase were included as controls for the soluble (S) and microsomal (M) fractions,
respectively.
124 Figure 3.8. The acylation-minus variant of HopAF1 co-localizes with free YFP, a
soluble protein. A-C) Estradiol-inducible HopAF1-cerulean-HA and HopAF1G2AC4Scerulean-HA were expressed transiently in N. benthamiana using Agrobacterium-mediated
transient transformations. Transgene expression was induced with 5 μM estradiol 3 days
after bacterial infiltration. A) The localization of transiently expressed proteins was
determined with scanning confocal laser microscopy 3 hours after estradiol induction in N.
benthamiana epidermal cells. Free YFP is a soluble protein and, therefore, can be used as a
marker of the plant cytoplasm. Scale bars represent 20 μm. B) Merged image of estradiolinducible HopAF1-cerulean-HA and YFP. C) Merged image of estradiol-inducible
HopAF1G2AC4S-cerulean-HA and YFP. White arrowheads indicate the plasma membrane
(PM), cytoplasmic streaming (CS), and nuclei (N).
125 Figure 3.9. The Yang cycle contributes to ethylene biosynthesis, which is suppressed
by Pto DC3000. A) In this figure adapted from Rzewuski et al., (Rzewuski et al., 2007) the
relationship between the Yang cycle (also known as the methionine recycling cycle) and the
phytohormone ethylene in Arabidopsis is highlighted. Arabidopsis MTN1 and MTN2 are two
Yang cycle proteins. Both can catabolize MTA by cleaving the adenine to create the product
MTR. MTR is then phosphorylated by the kinase MTK. MTNs also functions in a secondary
pathway to recycle methionine, in which S-adenosylhomocysteine is its substrate.
Methionine is a precursor for the ethylene biosynthesis pathway. It is important to note that
recycling of methionine supplements the continual de novo synthesis of methionine. B)
Arabidopsis seedlings were vacuum-infiltrated with multiple concentrations of P. syringae pv.
tomato DC3000 and the type III secretion system deficient mutant, P. syringae pv. tomato
DC3000 hrcC. 3 hours post infiltration, the fresh weight of 4 seedlings was measured and
these seedlings were sealed in a vial for 24 hours before ethylene accumulation was
measured. Error bars represent standard error and an ANOVA was performed, followed by
Tukey’s post-hoc analysis (p < 0.05). Groups that differ significantly are indicated with
different letters.
126 Figure 3.10. HopAF1 associates with Arabidopsis MTN proteins at the plasma
membrane. A) Protein extracts from N. benthamiana leaves transiently expressing HopAF14x-myc, HopAF1 H186A-4x-myc, MTN1-HA, and MTN2-HA were collected 3 days after
infiltration and subjected to immunoprecipitation using anti-myc-coupled magnetic beads.
400 μg of input and a 50x concentrated elution fraction were analyzed by anti-HA and antimyc immunoblots. B) HopAF1-nYFP, HopAF1H186A-nYFP, HopAF1G2AC4S-nYFP were
transiently co-expressed with MTN1-cYFP, MTN2-cYFP, and PLC2-cYFP in N. benthamiana
leaves using Agrobacterium-mediated transient transformation. 3 days after infiltration,
confocal microscopy was used to image the reconstituted YFP signal. Scale bar represents
50 μm.
127 Figure 3.11. Establishing P. fluorescens as a tool to determine the effect of HopAF1
on PAMP-induced ethylene accumulation. A) This dose response curve demonstrates
the increase in ethylene biosynthesis elicited by increasing concentrations of Pfo-1. B) Pfo-1
expressing native promoter HopAF1 variants was inoculated into Col-0 leaves and observed
after 24 hours. The ability to elicit an RPS2-dependent HR indicates translocation of the
fusion proteins. The number of leaves in which HR was evident compared to the number of
leaves inoculated is displayed next to representative leaves for each Pfo-1 inoculation. C)
Expression of native promoter HopAF1 variants in Pfo-1 was induced with minimal media for
12 hours and detected with anti-HA immunoblot.
128 Figure 3.12. HopAF1 inhibits PAMP-induced ethylene biosynthesis. A) Arabidopsis
seedlings were vacuum-infiltrated with either MgCl2, Pfo-1, Pfo-1 carrying the empty vector
(EV), Pfo-1 expressing HopAF1. 3 hours post infiltration, the fresh weight of 4 seedlings was
measured and these seedlings were sealed in a vial for 24 hours before ethylene
accumulation was measured. B) In addition to the aforementioned treatments, Arabidopsis
seedlings were vacuumed-infiltrated with Pfo-1 expressing HopAF1H186A and Pfo-1
expressing HopAF1G2AC4S. 3 hours after infiltration, the fresh weight of 4 seedlings was
measured and these seedlings were sealed in a vial for 24 hours before ethylene
accumulation was measured. Error bars represent standard error and an ANOVA was
performed, followed by Tukey’s post-hoc analysis (p < 0.05). Groups that differ significantly
are indicated with different letters.
129 Figure 3.13. Pfo-1 is unable to induce high levels of ethylene biosynthesis in Yang
cycle mutants, mtn1 mtn2 and mtk. A) Seedlings of Col-0 and Arabidopsis mutants mtn11, mtn1-2, mtn2-1, and mtn2-2, were vacuum-infiltrated with Pfo-1 carrying the empty vector
(EV) or Pfo-1 expressing wild-type, native promoter HopAF1. 3 hours post infiltration, the
fresh weight of 4 seedlings was measured and these seedlings were sealed in a vial for 24
hours before ethylene accumulation was measured. B) Seedlings of Col-0 and Arabidopsis
mutants mtn1-2 mtn2-2 and mtk were vacuum-infiltrated with either MgCl2, Pfo-1 carrying
the empty vector (EV) or Pfo-1 expressing wild-type, native promoter HopAF1. 3 hours post
infiltration, the fresh weight of 4 seedlings was measured and these seedlings were sealed
in a vial for 24 hours before ethylene accumulation was measured. Error bars represent
standard error and an ANOVA was performed, followed by Tukey’s post-hoc analysis (p <
0.05). Groups that differ significantly are indicated with different letters.
130 Figure 3.14. HopAF1 domain architecture and circular dichroism spectra of HopAF1,
HopAF1H186A, and MTN1. A) Schematic of HopAF1 predicted domains and the truncations
generated for in vitro analyses. B) Circular dichroism spectra collected at 25°C at the far-UV
CD spectrum (185 nm to 260 nm) demonstrates that wild-type MTN1, 6xHis-HopAF1 and
6xHis-HopAF1H186A have secondary structure. The H186A mutation in HopAF1 does not
perturb the secondary structure. The buffer control (20 mM sodium phosphate pH 7.4, 150
mM NaF) was set as the baseline and subtracted from the protein-associated spectra
above.
131 Figure 3.15. HopAF1 inhibits MTN1 activity in vitro in a manner dependent on catalytic
activity. A) Michaelis-Menten plot of MTN1 activity measured in vitro using a xanthine
oxidase-coupled spectrometric assay. Recombinant MTN1 was incubated for 18 hours in
buffer at 37°C alone or with equimolar amounts of recombinant wild-type 6xHis-HopAF1 or
GST. B) MTN1 enzyme specific activity measured in vitro using xanthine oxidase-coupled
spectrometric assay. MTN1 was incubated at 37°C for 18 hours with equimolar amounts of
6xHis-HopAF1, 6xHis-HopAF1 H186A, ∆147-HopAF1, ∆130-HopAF1, and GST. Error bars
represent standard error and an ANOVA was performed, followed by Tukey’s post-hoc
analysis (p < 0.05). Groups that differ significantly are indicated with different letters.
132 Figure 3.16. The Yang cycle is required for PTI in Arabidopsis. Bacterial growth of
P.syringae pv. tomato DC3000D28E (Cunnac et al., 2011) in Col-0, mtn1 mtn2, mtk, and
ein2 was measured 0 and 3 days after hand inoculation with 1x105 cfu (colony-forming
units)/mL bacteria. Error bars represent standard error (SEM). An analysis of variance
(ANOVA) was performed among the day 3 samples, followed by Tukey’s post-hoc analysis
(p < 0.05). Groups that differ significantly are indicated with different letters.
133 Figure 3.17. Proposed model of HopAF1 suppression of plant innate immunity. Plant
pathogens such as P. syringae and X. campestris inject multiple type III effector (T3E)
proteins into plant cells, including HopAF1 and AvrXv3, respectively. Recognition of PAMPs,
such as flg22, a peptide derived from bacterial flagellin, by pattern-recognition receptors
(PRRs) such as FLS2 triggers a signal cascade that results in a multiple of outputs, which
we collectively call PTI for PAMP-triggered immunity. Additionally, recognition of T3Es by
plant Resistance proteins, often called NLRs, triggers ETI. The hypersensitive response
(HR) is often associated with ETI. Through perception of flg22 and activation of a MAPK
cascade, PTI leads to an increase in ethylene biosynthesis. Our data suggests that HopAF1
inhibits PTI, including flg22-induced PTI. Our model predicts that HopAF1 is targeted to the
plant plasma membrane via acylation, where it inhibits the function of Yang cycle proteins,
MTN1 and MTN2, which are required for the induced level of ethylene that occurs during
PTI. As a result, the induced amount of ethylene that occurs during PTI may be an important
output, making the Yang cycle an important component of the plant defense pathway and
HopAF1 a critical T3E that prevents high ethylene accumulation during the plant defense
response by targeting the Yang cycle. Since we have confirmed that HopAF1 requires the
putative catalytic residues, H186 and C159, for these functions, future work will include
determining the molecular mechanism by which HopAF1 inhibits MTN activity and how this
potential modification translates into a loss-of-function phenotype for Arabidopsis MTNs.
134 REFERENCES
Akimoto-Tomiyama, C., Furutani, A., Tsuge, S., Washington, E.J., Nishizawa, Y., Minami,
E., and Ochiai, H. (2012). XopR, a type III effector secreted by Xanthomonas oryzae
pv. oryzae, suppresses microbe-associated molecular pattern-triggered immunity in
Arabidopsis thaliana. Molecular plant-microbe interactions : MPMI 25, 505-514.
Alonso, J.M., Hirayama, T., Roman, G., Nourizadeh, S., and Ecker, J.R. (1999). EIN2, a
bifunctional transducer of ethylene and stress responses in Arabidopsis. Science
284, 2148-2152.
Aslanidis, C., and de Jong, P.J. (1990). Ligation-independent cloning of PCR products (LICPCR). Nucleic acids research 18, 6069-6074.
Astua-Monge, G., Minsavage, G.V., Stall, R.E., Davis, M.J., Bonas, U., and Jones, J.B.
(2000). Resistance of tomato and pepper to T3 strains of Xanthomonas campestris
pv. vesicatoria is specified by a plant-inducible avirulence gene. Molecular plantmicrobe interactions : MPMI 13, 911-921.
Baltrus, D.A., Nishimura, M.T., Romanchuk, A., Chang, J.H., Mukhtar, M.S., Cherkis, K.,
Roach, J., Grant, S.R., Jones, C.D., and Dangl, J.L. (2011). Dynamic evolution of
pathogenicity revealed by sequencing and comparative genomics of 19
Pseudomonas syringae isolates. PLoS pathogens 7, e1002132.
Bent, A.F., Innes, R.W., Ecker, J.R., and Staskawicz, B.J. (1992). Disease development in
ethylene-insensitive Arabidopsis thaliana infected with virulent and avirulent
Pseudomonas and Xanthomonas pathogens. Molecular plant-microbe interactions :
MPMI 5, 372-378.
Block, A., and Alfano, J.R. (2011). Plant targets for Pseudomonas syringae type III effectors:
virulence targets or guarded decoys? Current opinion in microbiology 14, 39-46.
Boller, T., and Felix, G. (2009). A renaissance of elicitors: perception of microbe-associated
molecular patterns and danger signals by pattern-recognition receptors. Annual
review of plant biology 60, 379-406.
Boutrot, F., Segonzac, C., Chang, K.N., Qiao, H., Ecker, J.R., Zipfel, C., and Rathjen, J.P.
(2010). Direct transcriptional control of the Arabidopsis immune receptor FLS2 by the
ethylene-dependent transcription factors EIN3 and EIL1. Proceedings of the National
Academy of Sciences of the United States of America 107, 14502-14507.
Bujnicki, J.M., Elofsson, A., Fischer, D., and Rychlewski, L. (2001). Structure prediction
meta server. Bioinformatics 17, 750-751.
135 Burstenbinder, K., Rzewuski, G., Wirtz, M., Hell, R., and Sauter, M. (2007). The role of
methionine recycling for ethylene synthesis in Arabidopsis. The Plant journal : for cell
and molecular biology 49, 238-249.
Burstenbinder, K., Waduwara, I., Schoor, S., Moffatt, B.A., Wirtz, M., Minocha, S.C.,
Oppermann, Y., Bouchereau, A., Hell, R., and Sauter, M. (2010). Inhibition of 5'methylthioadenosine metabolism in the Yang cycle alters polyamine levels, and
impairs seedling growth and reproduction in Arabidopsis. The Plant journal : for cell
and molecular biology 62, 977-988.
Chang, J.H., Urbach, J.M., Law, T.F., Arnold, L.W., Hu, A., Gombar, S., Grant, S.R.,
Ausubel, F.M., and Dangl, J.L. (2005). A high-throughput, near-saturating screen for
type III effector genes from Pseudomonas syringae. Proceedings of the National
Academy of Sciences of the United States of America 102, 2549-2554.
Chao, X., Muff, T.J., Park, S.Y., Zhang, S., Pollard, A.M., Ordal, G.W., Bilwes, A.M., and
Crane, B.R. (2006). A receptor-modifying deamidase in complex with a signaling
phosphatase reveals reciprocal regulation. Cell 124, 561-571.
Chen, Z., Agnew, J.L., Cohen, J.D., He, P., Shan, L., Sheen, J., and Kunkel, B.N. (2007).
Pseudomonas syringae type III effector AvrRpt2 alters Arabidopsis thaliana auxin
physiology. Proceedings of the National Academy of Sciences of the United States of
America 104, 20131-20136.
Clough, S.J., and Bent, A.F. (1998). Floral dip: a simplified method for Agrobacteriummediated transformation of Arabidopsis thaliana. The Plant journal : for cell and
molecular biology 16, 735-743.
Crow, A., Hughes, R.K., Taieb, F., Oswald, E., and Banfield, M.J. (2012). The molecular
basis of ubiquitin-like protein NEDD8 deamidation by the bacterial effector protein
Cif. Proceedings of the National Academy of Sciences of the United States of
America 109, E1830-1838.
Crow, A., Race, P.R., Jubelin, G., Varela Chavez, C., Escoubas, J.M., Oswald, E., and
Banfield, M.J. (2009). Crystal structures of Cif from bacterial pathogens
Photorhabdus luminescens and Burkholderia pseudomallei. PloS one 4, e5582.
Cunnac, S., Chakravarthy, S., Kvitko, B.H., Russell, A.B., Martin, G.B., and Collmer, A.
(2011). Genetic disassembly and combinatorial reassembly identify a minimal
functional repertoire of type III effectors in Pseudomonas syringae. Proceedings of
the National Academy of Sciences of the United States of America 108, 2975-2980.
136 de Torres-Zabala, M., Truman, W., Bennett, M.H., Lafforgue, G., Mansfield, J.W., Rodriguez
Egea, P., Bogre, L., and Grant, M. (2007). Pseudomonas syringae pv. tomato hijacks
the Arabidopsis abscisic acid signalling pathway to cause disease. The EMBO
journal 26, 1434-1443.
Dowen, R.H., Engel, J.L., Shao, F., Ecker, J.R., and Dixon, J.E. (2009). A family of bacterial
cysteine protease type III effectors utilizes acylation-dependent and -independent
strategies to localize to plasma membranes. The Journal of biological chemistry 284,
15867-15879.
Dunn, S.M., Bryant, J.A., and Kerr, M.W. (1994). A simple spectrophotometric assay for
plant 5′-deoxy-5′-methylthioadenosine nucleosidase using xanthine oxidase as a
coupling enzyme. Phytochemical Analysis 5, 286.
Ecker, J.R., and Davis, R.W. (1987). Plant defense genes are regulated by ethylene.
Proceedings of the National Academy of Sciences of the United States of America
84, 5202-5206.
Eisenberg, D., Luthy, R., and Bowie, J.U. (1997). VERIFY3D: assessment of protein models
with three-dimensional profiles. Methods in enzymology 277, 396-404.
Felix, G., Duran, J.D., Volko, S., and Boller, T. (1999). Plants have a sensitive perception
system for the most conserved domain of bacterial flagellin. The Plant journal : for
cell and molecular biology 18, 265-276.
Felix, G., Grosskopf, D.G., Regenass, M., Basse, C.W., and Boller, T. (1991a). Elicitorinduced ethylene biosynthesis in tomato cells: characterization and use as a
bioassay for elicitor action. Plant physiology 97, 19-25.
Felix, G., Grosskopf, D.G., Regenass, M., and Boller, T. (1991b). Rapid changes of protein
phosphorylation are involved in transduction of the elicitor signal in plant cells.
Proceedings of the National Academy of Sciences of the United States of America
88, 8831-8834.
Flatau, G., Lemichez, E., Gauthier, M., Chardin, P., Paris, S., Fiorentini, C., and Boquet, P.
(1997). Toxin-induced activation of the G protein p21 Rho by deamidation of
glutamine. Nature 387, 729-733.
Gao, Z., Chung, E.H., Eitas, T.K., and Dangl, J.L. (2011). Plant intracellular innate immune
receptor Resistance to Pseudomonas syringae pv. maculicola 1 (RPM1) is activated
at, and functions on, the plasma membrane. Proceedings of the National Academy of
Sciences of the United States of America 108, 7619-7624.
137 Gehl, C., Waadt, R., Kudla, J., Mendel, R.R., and Hansch, R. (2009). New GATEWAY
vectors for high throughput analyses of protein-protein interactions by bimolecular
fluorescence complementation. Molecular plant 2, 1051-1058.
Gomez-Gomez, L., and Boller, T. (2000). FLS2: an LRR receptor-like kinase involved in the
perception of the bacterial elicitor flagellin in Arabidopsis. Molecular cell 5, 10031011.
Grefen, C., Donald, N., Hashimoto, K., Kudla, J., Schumacher, K., and Blatt, M.R. (2010). A
ubiquitin-10 promoter-based vector set for fluorescent protein tagging facilitates
temporal stability and native protein distribution in transient and stable expression
studies. The Plant journal : for cell and molecular biology 64, 355-365.
Jamir, Y., Guo, M., Oh, H.S., Petnicki-Ocwieja, T., Chen, S., Tang, X., Dickman, M.B.,
Collmer, A., and Alfano, J.R. (2004). Identification of Pseudomonas syringae type III
effectors that can suppress programmed cell death in plants and yeast. The Plant
journal : for cell and molecular biology 37, 554-565.
Jelenska, J., Yao, N., Vinatzer, B.A., Wright, C.M., Brodsky, J.L., and Greenberg, J.T.
(2007). A J domain virulence effector of Pseudomonas syringae remodels host
chloroplasts and suppresses defenses. Current biology : CB 17, 499-508.
Jones, J.D., and Dangl, J.L. (2006). The plant immune system. Nature 444, 323-329.
Jubelin, G., Chavez, C.V., Taieb, F., Banfield, M.J., Samba-Louaka, A., Nobe, R.,
Nougayrede, J.P., Zumbihl, R., Givaudan, A., Escoubas, J.M., et al. (2009). Cycle
inhibiting factors (CIFs) are a growing family of functional cyclomodulins present in
invertebrate and mammal bacterial pathogens. PloS one 4, e4855.
Keogh, R.C., Deverall, B.J., and McLeod, S. (1980). Transactions of the British Mycological
Society 74, 329.
Kim, J.G., Stork, W., and Mudgett, M.B. (2013). Xanthomonas type III effector XopD
desumoylates tomato transcription factor SlERF4 to suppress ethylene responses
and promote pathogen growth. Cell host & microbe 13, 143-154.
Kushad, M.M., Richardson, D.G., and Ferro, A.J. (1985). 5'-Methylthioadenosine
Nucleosidase and 5-Methylthioribose Kinase Activities and Ethylene Production
during Tomato Fruit Development and Ripening. Plant physiology 79, 525-529.
Lawton, K.A., Potter, S.L., Uknes, S., and Ryals, J. (1994). Acquired Resistance Signal
Transduction in Arabidopsis Is Ethylene Independent. The Plant cell 6, 581-588.
138 Lewis, J.D., Abada, W., Ma, W., Guttman, D.S., and Desveaux, D. (2008). The HopZ family
of Pseudomonas syringae type III effectors require myristoylation for virulence and
avirulence functions in Arabidopsis thaliana. Journal of bacteriology 190, 2880-2891.
Lindgren, P.B., Peet, R.C., and Panopoulos, N.J. (1986). Gene cluster of Pseudomonas
syringae pv. "phaseolicola" controls pathogenicity of bean plants and hypersensitivity
of nonhost plants. Journal of bacteriology 168, 512-522.
Liu, Y., and Zhang, S. (2004). Phosphorylation of 1-aminocyclopropane-1-carboxylic acid
synthase by MPK6, a stress-responsive mitogen-activated protein kinase, induces
ethylene biosynthesis in Arabidopsis. The Plant cell 16, 3386-3399.
Marmagne, A., Ferro, M., Meinnel, T., Bruley, C., Kuhn, L., Garin, J., Barbier-Brygoo, H.,
and Ephritikhine, G. (2007). A high content in lipid-modified peripheral proteins and
integral receptor kinases features in the arabidopsis plasma membrane proteome.
Molecular & cellular proteomics : MCP 6, 1980-1996.
Mersmann, S., Bourdais, G., Rietz, S., and Robatzek, S. (2010). Ethylene signaling
regulates accumulation of the FLS2 receptor and is required for the oxidative burst
contributing to plant immunity. Plant physiology 154, 391-400.
Miyazaki, J.H., and Yang, S.F. (1987). Metabolism of 5-methylthioribose to methionine.
Plant physiology 84, 277-281.
Mudgett, M.B., and Staskawicz, B.J. (1999). Characterization of the Pseudomonas syringae
pv. tomato AvrRpt2 protein: demonstration of secretion and processing during
bacterial pathogenesis. Molecular microbiology 32, 927-941.
Mukhtar, M.S., Carvunis, A.R., Dreze, M., Epple, P., Steinbrenner, J., Moore, J., Tasan, M.,
Galli, M., Hao, T., Nishimura, M.T., et al. (2011). Independently evolved virulence
effectors converge onto hubs in a plant immune system network. Science 333, 596601.
Nimchuk, Z., Marois, E., Kjemtrup, S., Leister, R.T., Katagiri, F., and Dangl, J.L. (2000).
Eukaryotic fatty acylation drives plasma membrane targeting and enhances function
of several type III effector proteins from Pseudomonas syringae. Cell 101, 353-363.
Oh, S.I., Park, J., Yoon, S., Kim, Y., Park, S., Ryu, M., Nam, M.J., Ok, S.H., Kim, J.K., Shin,
J.S., et al. (2008). The Arabidopsis calcium sensor calcineurin B-like 3 inhibits the 5'methylthioadenosine nucleosidase in a calcium-dependent manner. Plant physiology
148, 1883-1896.
139 Otterhag, L., Sommarin, M., and Pical, C. (2001). N-terminal EF-hand-like domain is
required for phosphoinositide-specific phospholipase C activity in Arabidopsis
thaliana. FEBS letters 497, 165-170.
Park, E.Y., Choi, W.S., Oh, S.I., Kim, K.N., Shin, J.S., and Song, H.K. (2009). Biochemical
and structural characterization of 5'-methylthioadenosine nucleosidases from
Arabidopsis thaliana. Biochemical and biophysical research communications 381,
619-624.
Park, E.Y., Oh, S.I., Nam, M.J., Shin, J.S., Kim, K.N., and Song, H.K. (2006). Crystal
structure of 5'-methylthioadenosine nucleosidase from Arabidopsis thaliana at 1.5-A
resolution. Proteins 65, 519-523.
Penninckx, I.A., Thomma, B.P., Buchala, A., Metraux, J.P., and Broekaert, W.F. (1998).
Concomitant activation of jasmonate and ethylene response pathways is required for
induction of a plant defensin gene in Arabidopsis. The Plant cell 10, 2103-2113.
Resh, M.D. (2006). Trafficking and signaling by fatty-acylated and prenylated proteins.
Nature chemical biology 2, 584-590.
Robert-Seilaniantz, A., Navarro, L., Bari, R., and Jones, J.D. (2007). Pathological hormone
imbalances. Current opinion in plant biology 10, 372-379.
Robert-Seilaniantz, A., Shan, L., Zhou, J.M., and Tang, X. (2006). The Pseudomonas
syringae pv. tomato DC3000 type III effector HopF2 has a putative myristoylation site
required for its avirulence and virulence functions. Molecular plant-microbe
interactions : MPMI 19, 130-138.
Roberts, M., Tang, S., Stallmann, A., Dangl, J.L., and Bonardi, V. (2013). Genetic
requirements for signaling from an autoactive plant NB-LRR intracellular innate
immune receptor. PLoS genetics 9, e1003465.
Rzewuski, G., Cornell, K.A., Rooney, L., Burstenbinder, K., Wirtz, M., Hell, R., and Sauter,
M. (2007). OsMTN encodes a 5'-methylthioadenosine nucleosidase that is upregulated during submergence-induced ethylene synthesis in rice (Oryza sativa L.).
Journal of experimental botany 58, 1505-1514.
Sali, A., and Blundell, T.L. (1993). Comparative protein modelling by satisfaction of spatial
restraints. Journal of molecular biology 234, 779-815.
Sauter, M., Cornell, K.A., Beszteri, S., and Rzewuski, G. (2004). Functional analysis of
methylthioribose kinase genes in plants. Plant physiology 136, 4061-4071.
140 Schmidt, G., Sehr, P., Wilm, M., Selzer, J., Mann, M., and Aktories, K. (1997). Gln 63 of Rho
is deamidated by Escherichia coli cytotoxic necrotizing factor-1. Nature 387, 725729.
Shan, L., Thara, V.K., Martin, G.B., Zhou, J.M., and Tang, X. (2000). The pseudomonas
AvrPto protein is differentially recognized by tomato and tobacco and is localized to
the plant plasma membrane. The Plant cell 12, 2323-2338.
Soding, J., Biegert, A., and Lupas, A.N. (2005). The HHpred interactive server for protein
homology detection and structure prediction. Nucleic acids research 33, W244-248.
Studholme, D.J., Kemen, E., MacLean, D., Schornack, S., Aritua, V., Thwaites, R., Grant,
M., Smith, J., and Jones, J.D. (2010). Genome-wide sequencing data reveals
virulence factors implicated in banana Xanthomonas wilt. FEMS microbiology letters
310, 182-192.
Thomas, W.J., Thireault, C.A., Kimbrel, J.A., and Chang, J.H. (2009). Recombineering and
stable integration of the Pseudomonas syringae pv. syringae 61 hrp/hrc cluster into
the genome of the soil bacterium Pseudomonas fluorescens Pf0-1. The Plant journal
: for cell and molecular biology 60, 919-928.
Voinnet, O., Rivas, S., Mestre, P., and Baulcombe, D. (2003). An enhanced transient
expression system in plants based on suppression of gene silencing by the p19
protein of tomato bushy stunt virus. The Plant journal : for cell and molecular biology
33, 949-956.
Washington, E.J., Banfield, M.J., and Dangl, J.L. (2013). What a Difference a Dalton Makes:
Bacterial Virulence Factors Modulate Eukaryotic Host Cell Signaling Systems via
Deamidation. Microbiology and molecular biology reviews : MMBR 77, 527-539.
Zipfel, C., Robatzek, S., Navarro, L., Oakeley, E.J., Jones, J.D., Felix, G., and Boller, T.
(2004). Bacterial disease resistance in Arabidopsis through flagellin perception.
Nature 428, 764-767.
141 Chapter 4
Conclusions and Future Directions
Many plant bacterial pathogens use type III effectors (T3Es) as their main virulence
determinants (Deslandes and Rivas, 2012). The type III secretion system is essential to the
ability of these plant pathogens to cause disease on their hosts. As such, it is critical to
understand which host cellular systems T3Es are targeting and how they are disrupting
those systems. Much progress has been made in this field during the course of my graduate
career. Because T3Es can be used as probes of the plant immune system, the knowledge
of which plant cellular systems T3Es disrupt can be used to decipher ways to make the
plants more resistant to pathogen attack. Additionally, we learn how these potent virulence
factors, which are common in plant and animal pathogens, function. Although recently
several functions and targets for T3Es have been identified, the sheer number of plant
pathogen T3Es ensures that there is still a lot of work to be done.
In Chapter 3, I presented research that describes the function of a conserved type III
effector, HopAF1, and I noted how it contributes to disease in Arabidopsis. HopAF1 inhibits
the function of Yang cycle proteins, MTN1 and MTN2. This prevents the induction of
ethylene biosynthesis that occurs during the plant immune response after the perception of
PAMPs (termed PTI). The result is that HopAF1 inhibits a component of PTI, leading to
enhanced growth of weakened strains of P. syringae.
The research described in Chapter 3 is the beginning of our efforts to fully elucidate
the function of HopAF1 and the specific effect that it has on Arabidopsis MTN proteins.
Although we have predicted, based on structural modeling, that HopAF1 is a member of the
family of bacterial toxins/effectors that are deamidases (Washington et al., 2013), we have
not yet demonstrated that HopAF1 activity leads to deamidation of Arabidopsis MTNs or any
other proteins. We have, however, demonstrated that the ability of HopAF1 to block PTI,
inhibit MTN1 function in vitro, and inhibit PAMP-induced ethylene biosynthesis is dependent
on putative catalytic residues. Therefore, we demonstrated that HopAF1 has enzymatic
activity.
In this chapter, I will discuss the future directions to significantly extend this work,
which include determining the specific biochemical effect of HopAF1 on MTN1, solving the
HopAF1 and MTN1 co-crystal structure, and using the Yang cycle to increase plant
resistance to pathogen attack.
Using a shotgun approach and mass spectrometry to determine if HopAF1
modulates MTN function via deamidation.
Bacterial virulence factor deamidases typically alter a single conserved residue to
modulate target protein function (Washington et al., 2013). Thus, we sought to identify
conserved Arabidopsis MTN asparagine and glutamine residues that conferred loss-offunction phenotypes when mutated to the ‘deamidated’ residues, aspartic acid and glutamic
acid, respectively. We discovered only three fully conserved residues across the
evolutionary distance between plants and E. coli that could be targets of putative
deamidases: MTN1N113, MTN1N169, and MTN1N194 (Figure 4.1). Since the crystal structures of
Arabidopsis MTN1 and MTN2 are solved, we determined where the putative deamidation
target residues are in relation to the MTN ligand binding site and catalytic residues.
MTN1N113 is involved in ligand binding via coordination of a water molecule in the ligand
binding pocket. MTN1N194 and MTN1N169 are both surface-exposed and at the end of helices
α4 and α3, respectively (Figure 4.1). These conserved residues are located in approximately
the same place in the MTN2 crystal structure.
We generated individual deamidation mimic MTN1 variants (MTN1N113D, MTN1N169D,
and MTN1N194D) and tested their function using an assay to measure MTN activity in vitro
143 (Dunn et al., 1994). For negative controls, we also cloned and purified MTN1D225N and
MTN2D212N, catalytic residue loss-of-function alleles (Bretes et al., 2011). We determined
that both MTN1N113D and MTN1N194D lost activity in vitro, while MTN1N169D retained wild-type
levels of activity (Figure 4.2). Circular dichroism spectra confirmed that all of the MTN1
deamidation mimic variants retain elements of secondary structure, similar to wild-type
MTN1 (Figure 4.3).
We also determined whether the loss-of-function phenotypes we noted above are
dependent on the introduction of a negative charge in the deamidation mimics. The
MTN1N194A and MTN1N194V variants retained wild-type levels of activity (Figure 4.2). This
suggests that a charge change at MTN1N194, as would occur during deamidation, causes a
loss-of-function phenotype but that conservative replacement at MTN1N194 does not alter
activity. Conversely, MTN1N113A and MTN1N113V resulted in loss-of-function (Figure 4.2),
suggesting that MTN1N113 is sensitive to multiple mutations. This is consistent with the idea
that MTN1N113 is involved in ligand binding. We generated the same set of variants for
Arabidopsis MTN2 and determined that they functioned similar to MTN1 (Figure 4.2).
Based on the results using our deamidation mimics of MTN1 and MTN2, we
hypothesize that HopAF1 targets MTN1N113, MTN1N194, or both. Mass spectrometry can be a
critical tool for identifying specific residues that are modified via posttranslational
modifications (Witze et al., 2007). As such, mass spectrometry has been used to identify a
single residue on each target protein that is deamidated by known bacterial deamidase
toxins (Washington et al., 2013). Our initial attempts to identify MTN residues deamidated by
HopAF1 via mass spectrometry resulted in a general identification of nonenzymatically
deamidated residues (not shown). Nonenzymatic deamidation is a natural process that
occurs as proteins age, leading to their eventually misfolding and degradation. We
concluded that mutated residues identified in our original mass spectrometry experiments
were a result of nonenzymatic deamidation because multiple MTN residues were
144 deamidated after no treatment, treatment with HopAF1, and treatment with catalytic dead
HopAF1.
As an alternative, we will generate peptides containing the residues described above
and incubate them with HopAF1. Although, deamidation of peptides derived from a target
protein has not been demonstrated using mass spectrometry, Lerm et al. demonstrated that
CNF1 can deamidate a peptide derived from RhoA residues 59-78 using the glutamate
dehydrogenase assay (Lerm et al., 1999). It was also demonstrated that CNF1 could
deamidate a peptide derived from RhoA residues 59-69. Deamidation was evident due to
increased elution times on a C18 hydrophobic column (Flatau et al., 2000). After incubating
HopAF1 with the peptides derived from MTN1, we will use a hydrophobic column to purify
the peptides and analyze them using mass spectrometry. We will be able to determine
whether the putative target residues have been converted to aspartic acids. We will also
attempt to detect deamidation of peptides generated from recombinant wild-type MTN1
protein. These peptides may allow us to detect the modification of residues not considered
in our shotgun approach.
Determine whether a functional deamidation site variant of MTN1 is impervious to
HopAF1 activity in planta.
We demonstrated above that deamidation mimics MTN1N113D and MTN1N194D lose
enzymatic activity in vitro. We also showed that MTN1N194A and MTN1N194V retain activity. We
are currently performing in planta complementation experiments using ubiquitin promoter
driven constructs of wild-type MTN1, MTN1N194D, MTN N194V, MTN1N113D, MTN1N113V, and
MTN1D225N transformed into the mtn1 mtn2 double mutant. This double mutant is viable, but
fails to set seed (Burstenbinder et al., 2010). It has previously been demonstrated that a
ubiquitin promoter-driven MTN1 can complement the mtn1 mtn2 infertility phenotype
(Waduwara-Jayabahu et al., 2012). We will first address whether each of our constructs can
complement the loss of seed set phenotype. Based on the results of our in vitro assays
145 described above, we hypothesize that MTN1 wild-type and MTN1N194V will complement the
loss of seed set phenotype because these are functional variants, but that the deamidation
mimic MTN1N194D will not. Similarly, we expect the MTN1N113D and MTN1D225N
complementation lines to remain infertile because MTN1N113D and MTN1D225N are loss-offunction variants involved in ligand binding and catalysis, respectively.
We will then repeat our experiment to measure PAMP-induced ethylene biosynthesis
in transgenic seedlings expressing each MTN1 variant. If our hypothesis is true, we expect
that MTN1 wild-type and MTN1N194V will support PAMP-dependent ethylene formation, but
that MTN1N194D will not. We also expect that HopAF1 delivery will suppress Pfo-1-induced
ethylene production in plants expressing wild-type MTN1. The key test will be whether or not
MTN1N194V, which we hypothesize will still be functional based on our MTN activity
determination (see Figure 4.2 above), will be unaffected by HopAF1. If so, we will conclude
that MTN1N194 is the sole site of action for HopAF1.
In Chapter 3, we demonstrate that the mtn1 mtn2 double mutant is more susceptible
to weakened Pto DC3000D28E compared to Col-0 wild-type. We will determine the
susceptibility of mtn1 mtn2 plants to Pto DC3000D28E expressing wild-type HopAF1 under
a native promoter with a C-terminal HA tag. If Pto DC3000D28E(HopAF1) grows to equal
levels in Col-0 as Pto DC3000D28E does in mtn1 mtn2 and there is no further increase in
growth of Pto DC3000D28E(HopAF1) on mtn1 mtn2 plants, this will indicate the MTN1 and
MTN2 are necessary for the HopAF1-mediated increase in the growth ofDC3000D28E that
we described in Chapter 3.
Furthermore, we will determine the ability of Pto DC3000D28E and Pto
DC3000D28E(HopAF1) to grow in the mtn1 mtn2 complementation lines in which we
express either wild-type MTN1, MTN1N194V, MTN1N194D, MTN1N113D, MTN1N113V or MTN1D225N
under the ubiquitin promoter. Complementing mtn1 mtn2 with non-functional variants of
MTN1, such as MTN1N194D and MTN1N113D will likely result in plants in which both Pto
146 DC3000D28E and Pto DC3000D28E(HopAF1) grow to an equally high titer. However, since
MTN1N194V is functional and no longer a target of the putative deamidation activity of
HopAF1, we expect MTN1N194V to be blind to HopAF1 and therefore to not be susceptible to
either Pto DC3000D28E or Pto DC3000D28E(HopAF1). This result will further confirm that
the ability of HopAF1 to inhibit PTI is dependent upon modification of Arabidopsis MTN1N194.
Generate a HopAF1/MTN1 co-crystal to determine the molecular mechanism
behind HopAF1 inhibition of MTN activity.
The generation of co-crystals of bacterial deamidases and their targets addresses
the critical question of how deamidases can modulate target protein activity (Crow et al.,
2012; Fu et al., 2013; Yao et al., 2012). This will also provide information on how target
proteins function, which is highly relevant for several known deamidase targets such as
RhoA, ubiquitin, NEDD8, and Gα proteins (Cui et al., 2010; Flatau et al., 1997; Orth et al.,
2009; Schmidt et al., 1997). It is equally important to understand how plant
methylthioadenosine nucleosidases function, as these proteins are involved in recycling
methionine, the precursor for protein translation, polyamine biosynthesis, and ethylene
biosynthesis.
Using published methods (Park et al., 2006), we are able to purify MTN1 from E. coli.
We generated an expression construct using the Gateway-compatible pDEST15 vector
(Gateway, Invitrogen). This vector includes an IPTG-inducible promoter with an N-terminal
GST tag. We cloned a PreScission Protease (GE Healthcare) cleavage site between the
GST tag and MTN1 to allow for specific and efficient cleavage at 4°C. We used affinity
chromatography to separate GST-tagged MTN1 from the crude lysate. Then we cleaved
used PreScission Protease to cleave the protein and used reverse affinity to
chromatography to capture ‘free’ MTN1 in the flow-through. This protein was approximately
90% pure (Figure 4.4). To increase the purity of MTN1 for the purposes of crystallization, we
147 will run the partially purified protein over a size exclusion column. This will also determine if
the protein is folded and monodispersed.
After attempting many methods, we have made significant progress in optimizing the
purification of HopAF1. We used a disorder prediction program, RONN, which revealed that
the N-terminus of HopAF1 contains regions of disorder (Yang et al., 2005). These regions of
disorder may likely make it difficult to purify large quantities of full-length HopAF1 and may
also inhibit crystal formation in the future. Again using Gateway technology, we cloned a
6xHis-MBP (maltose-binding protein) on the N-terminus of HopAF1 truncations and we
identified the minimal region of HopAF1 required for activity (see Figure 3.14 and Figure
3.15 in Chapter 3). 6xHisMBP-∆55HopAF1 was insoluble, while 6xHis-∆147HopAF1 was
not able to inhibit MTN activity in vitro. However, purification of 6xHis-∆130HopAF1, the
residues of HopAF1 that correspond to the putative catalytic domain, was successful (Figure
4.5) and ∆130HopAF1 can inhibit MTN1 activity (see Figure 3.15 in Chapter 3). Therefore,
we decided to focus on purifying ∆130HopAF1 for crystallography studies.
Our current protocol, similar to the purification scheme of MTN1, uses affinity
chromatography was used to capture 6xHisMBP-∆130HopAF1 from crude extract (figure
5A). Then after cleavage with TEV protease, reverse affinity chromatography was used to
capture ∆130HopAF1 in the flow-through (Figure 4.5B). Finally, ∆130HopAF1 was run over
a size exclusion chromatography column, resulting in a monodispersed peak (Figure 4.5C).
Interestingly, the ∆30HopAF1 monodispersed peak elution time suggests that the protein is
forming oligomers. We will perform size exclusion chromatography multi-angle light
scattering (SEC-MALS) experiments in the UNC Macromolecular Interactions Facility to
determine the nature of these oligomers.
To start the crystallization trials, we will first confirm that HopAF1 and MTN1 form a
stable complex in solution using size exclusion chromatography. Since the crystal structure
of MTN1 from Arabidopsis has been solved (Park et al., 2006), we will start with the
148 conditions used for MTN1 and begin the crystallization trials with HopAF1/MTN1 in the UNC
Protein Crystallography core facility.
149 Figures
Figure 4.1. Alignment and structural localization of MTN1 conserved putative
deamidase target residues. A) Alignment of methylthioadenosine nucleosidase amino acid
sequences from plants and E. coli. B) Ribbon representation of the MTN1 crystal structure
(PDB 2H8G). Monomers are represented in sand and silver. The ligand analog (MTT) is
shown in green. The ligand-binding and catalytic residues from MTN1 are represented as
blue sticks. The putative conserved targets of deamidation are labeled and colored as red
stick representations. Images were generated using Pymol.
150 Figure 4.2. MTN1 and MTN2 variants display loss-of-function phenotypes when
residues N113 and N194 are ‘deamidated’. A) MTN1 enzyme specific activity measured in
vitro using xanthine oxidase-coupled spectrometric assay. MTN1 catalytic dead mutant is
MTN1D225N. B) MTN2 enzyme specific activity measured in vitro using xanthine-oxidase
coupled spectrometric assay. MTN2 catalytic dead mutant is MTN2D212N. Letters represent
treatments with significant difference according to the post-hoc ANOVA Tukey’s test (p <
0.05).
151 Figure 4.3. Circular dichroism spectra of MTN1 ‘deamidated’ variants does not vary
from wild-type MTN1. Circular dichroism spectra were recorded for MTN1 variants in 20
mM sodium phosphate buffer pH 7.4, 150 mM NaF from 185 nm to 260 nm. The differential
spectrum was obtained by subtracting the buffer spectrum before conversion to molar
ellipticity.
152 Figure 4.4. Purification of wild-type MTN1. A) The construct used to purify MTN1 consists
of an N-terminal glutathione S-transferase (GST) tag, followed by a PreScission Protease
cleavage site (PP), and the coding sequence of Arabidopsis MTN1 (At4g38800). B) SDSPAGE analysis and Coomassie stain of affinity purification using the GST tag. Fractions 1317 represent the eluted fractions that contain full-length GST-PP-MTN1. C) SDS-PAGE
analysis and Coomassie stain of MTN1 after cleavage with PreScission Protease to remove
the GST tag. The cleaved protein was run over a fresh GST affinity column to capture the
GST and PreScission Protease, allowing MTN1 to be collected in the flow through fractions.
Black arrows denote the proteins of interest from each step of the purification.
153 Figure 4.5. Purification of ∆130 HopAF1. A) The construct used to purify ∆130-HopAF1
consists of an N-terminal 6xHis-Maltose Binding Protein (MBP) tag, followed by a TEV
Protease cleavage site (TEV), and the coding sequence of P. syringae pv. tomato DC3000
HopAF1. B) SDS-PAGE analysis and Coomassie stain of affinity purification using the 6xHis
tag. The eluted fractions contain full-length 6xHisMBP-TEV-∆130HopAF1. C) SDS-PAGE
analysis on a 15% gel and Coomassie stain of ∆130HopAF1 after cleavage with TEV
Protease to remove the 6xHisMBP tag. The cleaved protein was run over a fresh His affinity
column to capture the 6xHisMBP and TEV Protease, allowing ∆130HopAF1 to be collected
in the flow through fractions. D) ∆130HopAF1 was run over a size exclusion column
resulting in a monodispersed elution of protein. Black arrows denote the proteins of interest
from each step of the purification.
154 REFERENCES
Bretes, E., Guranowski, A., and Nuc, K. (2011). 5'-methylthioadenosine nucleosidase from
yellow lupine (Lupinus luteus): molecular characterization and mutational analysis.
Protein and peptide letters 18, 817-824.
Burstenbinder, K., Waduwara, I., Schoor, S., Moffatt, B.A., Wirtz, M., Minocha, S.C.,
Oppermann, Y., Bouchereau, A., Hell, R., and Sauter, M. (2010). Inhibition of 5'methylthioadenosine metabolism in the Yang cycle alters polyamine levels, and
impairs seedling growth and reproduction in Arabidopsis. The Plant journal : for cell
and molecular biology 62, 977-988.
Crow, A., Hughes, R.K., Taieb, F., Oswald, E., and Banfield, M.J. (2012). The molecular
basis of ubiquitin-like protein NEDD8 deamidation by the bacterial effector protein
Cif. Proceedings of the National Academy of Sciences of the United States of
America 109, E1830-1838.
Cui, J., Yao, Q., Li, S., Ding, X., Lu, Q., Mao, H., Liu, L., Zheng, N., Chen, S., and Shao, F.
(2010). Glutamine deamidation and dysfunction of ubiquitin/NEDD8 induced by a
bacterial effector family. Science 329, 1215-1218.
Deslandes, L., and Rivas, S. (2012). Catch me if you can: bacterial effectors and plant
targets. Trends in plant science 17, 644-655.
Dunn, S.M., Bryant, J.A., and Kerr, M.W. (1994). A simple spectrophotometric assay for
plant 5′-deoxy-5′-methylthioadenosine nucleosidase using xanthine oxidase as a
coupling enzyme. Phytochemical Analysis 5, 286.
Flatau, G., Landraud, L., Boquet, P., Bruzzone, M., and Munro, P. (2000). Deamidation of
RhoA glutamine 63 by the Escherichia coli CNF1 toxin requires a short sequence of
the GTPase switch 2 domain. Biochemical and biophysical research communications
267, 588-592.
Flatau, G., Lemichez, E., Gauthier, M., Chardin, P., Paris, S., Fiorentini, C., and Boquet, P.
(1997). Toxin-induced activation of the G protein p21 Rho by deamidation of
glutamine. Nature 387, 729-733.
Fu, P., Zhang, X., Jin, M., Xu, L., Wang, C., Xia, Z., and Zhu, Y. (2013). Complex Structure
of OspI and Ubc13: The Molecular Basis of Ubc13 Deamidation and Convergence of
Bacterial and Host E2 Recognition. PLoS pathogens 9, e1003322.
Lerm, M., Selzer, J., Hoffmeyer, A., Rapp, U.R., Aktories, K., and Schmidt, G. (1999).
Deamidation of Cdc42 and Rac by Escherichia coli cytotoxic necrotizing factor 1:
155 activation of c-Jun N-terminal kinase in HeLa cells. Infection and immunity 67, 496503.
Orth, J.H., Preuss, I., Fester, I., Schlosser, A., Wilson, B.A., and Aktories, K. (2009).
Pasteurella multocida toxin activation of heterotrimeric G proteins by deamidation.
Proceedings of the National Academy of Sciences of the United States of America
106, 7179-7184.
Park, E.Y., Oh, S.I., Nam, M.J., Shin, J.S., Kim, K.N., and Song, H.K. (2006). Crystal
structure of 5'-methylthioadenosine nucleosidase from Arabidopsis thaliana at 1.5-A
resolution. Proteins 65, 519-523.
Schmidt, G., Sehr, P., Wilm, M., Selzer, J., Mann, M., and Aktories, K. (1997). Gln 63 of Rho
is deamidated by Escherichia coli cytotoxic necrotizing factor-1. Nature 387, 725729.
Waduwara-Jayabahu, I., Oppermann, Y., Wirtz, M., Hull, Z.T., Schoor, S., Plotnikov, A.N.,
Hell, R., Sauter, M., and Moffatt, B.A. (2012). Recycling of methylthioadenosine is
essential for normal vascular development and reproduction in Arabidopsis. Plant
physiology 158, 1728-1744.
Washington, E.J., Banfield, M.J., and Dangl, J.L. (2013). What a Difference a Dalton Makes:
Bacterial Virulence Factors Modulate Eukaryotic Host Cell Signaling Systems via
Deamidation. Microbiology and molecular biology reviews : MMBR 77, 527-539.
Witze, E.S., Old, W.M., Resing, K.A., and Ahn, N.G. (2007). Mapping protein posttranslational modifications with mass spectrometry. Nature methods 4, 798-806.
Yang, Z.R., Thomson, R., McNeil, P., and Esnouf, R.M. (2005). RONN: the bio-basis
function neural network technique applied to the detection of natively disordered
regions in proteins. Bioinformatics 21, 3369-3376.
Yao, Q., Cui, J., Wang, J., Li, T., Wan, X., Luo, T., Gong, Y.N., Xu, Y., Huang, N., and Shao,
F. (2012). Structural mechanism of ubiquitin and NEDD8 deamidation catalyzed by
bacterial effectors that induce macrophage-specific apoptosis. Proceedings of the
National Academy of Sciences of the United States of America 109, 20395-20400.
156