Download PhD thesis MAP Kinase 4 substrates and plant innate immunity

Document related concepts

Drosophila melanogaster wikipedia , lookup

Molecular mimicry wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Innate immune system wikipedia , lookup

Immunomics wikipedia , lookup

Plant disease resistance wikipedia , lookup

Transcript
UNIVERSITY OF COPENHAGEN
FACULTY OF SCIENCE
PhD thesis
Magnus Wohlfahrt Rasmussen
MAP Kinase 4 substrates and plant innate immunity
Supervisor: Morten Petersen
Handed in: 31/12/2015
Table of Contents
PREFACE ............................................................................................................................ 5
Aknowlegements ............................................................................................................................................................... 5
ABSTRACT ......................................................................................................................... 6
SAMMENDRAG .................................................................................................................. 7
LIST OF ABBREVIATIONS ................................................................................................ 8
PLANT INNATE IMMUNITY .............................................................................................. 11
Pathogen associated molecular patterns (PAMPs) ...................................................................................................... 11
Pathogenic effector molecules and host resistance (R) proteins ................................................................................. 12
Autoimmunity ................................................................................................................................................................. 14
References........................................................................................................................................................................ 15
MAP KINASES IN ARABIDOPSIS INNATE IMMUNITY................................................... 21
Abstract ........................................................................................................................................................................... 21
Introduction .................................................................................................................................................................... 21
MAPK cascades in PAMP triggered immunity (PTI) ................................................................................................. 21
MAPK cascades in effector triggered immunity (ETI) ............................................................................................... 22
WRKY transcription factors ......................................................................................................................................... 23
MAPK in general stress signaling ................................................................................................................................. 24
References........................................................................................................................................................................ 24
MRNA DECAY IN PLANT IMMUNITY .............................................................................. 27
Abstract ........................................................................................................................................................................... 27
Introduction .................................................................................................................................................................... 27
Plant innate immunity .................................................................................................................................................... 28
1
mRNA decay and autoimmunity ................................................................................................................................... 29
Conclusion ....................................................................................................................................................................... 35
References........................................................................................................................................................................ 35
THE MRNA DECAY FACTOR PAT1 FUNCTIONS IN A PATHWAY INCLUDING MAP
KINASE 4 AND IMMUNE RECEPTOR SUMM2 ............................................................... 42
Abstract ........................................................................................................................................................................... 42
Introduction .................................................................................................................................................................... 42
Results .............................................................................................................................................................................. 43
AtPAT1 is an ScPAT1 orthologue and interacts with MPK4 in planta ....................................................................... 43
PAT1 is required for decapping of selected mRNAs ................................................................................................... 45
PAT1 is an MPK4 substrate ......................................................................................................................................... 45
pat1 mutants exhibit autoimmunity similar to mpk4 .................................................................................................... 46
PAT1 detection in P-bodies is induced by PAMPs ...................................................................................................... 47
The MPK4 suppressor summ2 also suppresses the pat1 resistance phenotype ............................................................ 47
Discussion ........................................................................................................................................................................ 48
Materials and Methods................................................................................................................................................... 51
Plant materials and growth conditions ......................................................................................................................... 51
Cloning and transgenic lines ........................................................................................................................................ 52
Mutagenesis of His6-PAT1 .......................................................................................................................................... 52
PAT1 protein purification and in vitro kinase assays................................................................................................... 52
Yeast transformation .................................................................................................................................................... 52
Flg22 kinetics ............................................................................................................................................................... 52
Semi-quantitative and qRT-PCR ................................................................................................................................. 53
RNA extraction and RACE PCR ................................................................................................................................. 53
Quantification of capped versus uncapped transcripts ................................................................................................. 53
Confocal microscopy ................................................................................................................................................... 53
Infection assays ............................................................................................................................................................ 53
Transient expression in N. benthamiana ...................................................................................................................... 53
Protein extraction and immunoprecipitation in N. benthamiana.................................................................................. 53
Arabidopsis protein extraction and immunoprecipitation for mass spectrometry analysis .......................................... 54
SDS-PAGE and immunoblotting ................................................................................................................................. 54
Antibodies .................................................................................................................................................................... 54
In-gel digestion, TiO2 chromatography and mass spectrometry .................................................................................. 54
Acknowledgements ......................................................................................................................................................... 54
2
Author contributions ...................................................................................................................................................... 55
Conflict of Interest .......................................................................................................................................................... 55
References........................................................................................................................................................................ 55
Supplemental figures ...................................................................................................................................................... 58
AOC3 IS AN MPK4 SUBSTRATE AND ITS OVEREXPRESSION INDUCE IMMUNITY
AGAINST PST DC3000..................................................................................................... 66
Introduction .................................................................................................................................................................... 66
Results .............................................................................................................................................................................. 70
AOC3 Interacts with MPK4 in planta. ......................................................................................................................... 70
Over-expression of AOC3 leads to increased resistance .............................................................................................. 72
AOC3 is an MPK4 substrate ........................................................................................................................................ 72
MPK4 associates with chloroplasts. ............................................................................................................................. 75
Discussion ........................................................................................................................................................................ 77
Materials and methods ................................................................................................................................................... 78
References........................................................................................................................................................................ 79
EIF4E-THR6 IS SPECIFICALLY PHOSPHORYLATED IN VITRO BY MPK4 AND FORM
A COMPLEX IN THE NUCLEUS IN VIVO. ....................................................................... 83
Introduction .................................................................................................................................................................... 83
Results .............................................................................................................................................................................. 87
eIF4E is an MPK4 substrate......................................................................................................................................... 87
eIF4E interacts with MPK4 in planta ........................................................................................................................... 88
eif4e mutants does not exhibit autoimmunity similar to mpk4 ..................................................................................... 88
ClYVV susceptibility is not dependent on eIF4E phosphorylation ............................................................................. 90
MPK4 and eIF4E interact in the nucleus. .................................................................................................................... 92
Discussion ........................................................................................................................................................................ 92
Materials and methods ................................................................................................................................................... 94
Plants growth conditions .............................................................................................................................................. 94
SDS-PAGE and immunoblotting ................................................................................................................................. 95
Antibodies .................................................................................................................................................................... 95
Bacterial infection assays ............................................................................................................................................. 96
Transient expression in Nicotiana benthamiana ........................................................................................................... 96
3
Protein extraction and immunoprecipitation in Nicotiana benthamiana and Arabidopsis........................................... 96
Purification of X7 polymerase ..................................................................................................................................... 97
Genotyping................................................................................................................................................................... 97
Cloning and transgenic lines ........................................................................................................................................ 97
Recombinant protein purification and in vitro kinase assays ....................................................................................... 98
Virus inoculation and detection by ELISA .................................................................................................................. 98
References........................................................................................................................................................................ 99
CONCLUDING REMARKS ............................................................................................. 104
References...................................................................................................................................................................... 105
4
Preface
This thesis concludes my PhD work at the department of Biology, University of Copenhagen. My
main research focus has been on substrates of the immune regulating MAP kinase 4 and their
involvement in immunity. The thesis consist of:
(i)
A general introduction to plant innate immunity.
(ii)
A review article “MAP Kinases in Arabidopsis innate immunity” published in “Frontiers
in Plant Science”. This article serves as an introduction, giving an overview on the
function of MAP kinases immune signaling in Arabidopsis.
(iii)
An invited review article “mRNA decay in plant immunity” under revision for
publication in “Cellular and Molecular Life Science”. This article gives an overview on
our current understanding of the involvement of mRNA decay in plant immunity and
serves as a more comprehensive introduction to (iv).
(iv)
A research article “The mRNA decay factor PAT1 functions in a pathway including
MAP kinase 4 and immune receptor SUMM2” published in ”The EMBO Journal”.
(v)
Two draft manuscripts regarding novel MPK4 substrates and their putative function in
defense.
Aknowlegements
An era has ended and I will finally leave the PMB lab! I am sure that I will be missed and trust me;
I will miss all of you. I would like to thank my main supervisor Morten Petersen and co-supervisor
John Mundy for their support and guidance during my research and the opportunity to work in your
lab. I will also take the opportunity to thank all the current and previous lab members for making
my stay in the lab such a wonderful time. I would in particular like to thank my work wife no.1
Milena for all the interesting chitchat and for being my personal scientific oracle and my work wife
no.2 Àngels (even though you left me for a real job!) for simply putting up with me and my crazy
ideas. Lastly, thank you Tine for bear with me the numerous times when I’ve told you that that I’m
almost on my way home and end up staying two hours extra in the lab.
5
Abstract
Multi-layered defense responses are activated in plants upon recognition of invading pathogens.
Transmembrane receptors recognize conserved pathogen-associated molecular patterns and activate
MAP kinase cascades, which regulate changes in gene expression to produce appropriate immune
responses. For example, Arabidopsis MPK4 regulates the expression of a subset of defense genes
via at least one WRKY transcription factor. We report here that MPK4 is found in complexes in
vivo with (i) PAT1, component of the mRNA decapping machinery, (ii) AOC3, a component in the
biosynthesis pathway of JA and (iii) eIF4E, a component in the translational initiation protein
complex. For PAT1 and eIF4E we show that MPK4 phosphorylates specific Ser and Thr residues in
vitro, and that MPK4 also phosphorylates AOC3 at an unmapped residue. Specific in vivo
phosphorylation for PAT1 is shown in response to pathogen recognition, which also induce its
localization to cytoplasmic processing bodies. All three proteins; PAT1, AOC3 and eIF4E also
interacts with MPK4 in vivo although the functional outcome of these interactions are still elusive.
The thesis comprise a general introduction to plant innate immunity followed by two review articles
“MAP kinase cascades in Arabidopsis innate immunity” published in Frontiers in Plant Science
and “mRNA decay in plant immunity” under revision for Cellular and Molecular Life Science.
Together these sections gives a comprehensive overview of Arabidopsis defense signaling. The
results are presents as one manuscript “The mRNA decay factor PAT1 functions in a pathway
including MAP kinase 4 and immune receptor SUMM2” published in The EMBO Journal and two
draft manuscripts summarizing our data on regarding AOC3 and eIF4E in respect to MPK4.
6
Sammendrag
Flere forsvarslag aktiveres i planter ved genkendelse af invaderende patogener. Transmembrane
receptorer genkender konserverede patogen-associeret molekylære mønstre og aktivere MAPkinase kaskader, som regulerer ændringer i gen-ekspression for at producere passende
immunresponser. For eksempel, Arabidopsis MPK4 regulerer ekspressionen af en undergruppe af
forsvarsgener via mindst en WRKY transkriptionsfaktor. Vi rapporterer her, at MPK4 findes i
komplekser in vivo med (i) PAT1, en komponent i ”mRNA decapping”, (ii) AOC3, en komponent i
biosyntesevejen af JA og (iii) eIF4E, en komponent i translationsinitierings komplekset. For PAT1
og eIF4E viser vi, at MPK4 phosphorylerer specifikke Ser og Thr aminosyrerestre in vitro, og at
MPK4 også phosphorylerer en ukendt aminosyreester i AOC3. Specifik in vivo phosphorylering af
PAT1 sker som reaktion på patogen genkendelse, som derudover også inducerer dens lokalisering
til cytoplasmiske ”procesing bodies”. Alle tre proteiner; PAT1, AOC3 og eIF4E interagerer med
MPK4 in vivo, selvom det funktionelle resultat af disse interaktioner stadig er ukendte.
Afhandlingen omfatter en generel introduktion til at planters medfødte immunforsvar efterfulgt af to
review-artikler " MAP kinase cascades in Arabidopsis innate immunity " offentliggjort i Frontiers
in Plant Science og "mRNA decay in plant immunity" under revision for Cellular and Molecular
Life Science. Disse afsnit giver tilsammen et samlet overblik over Arabidopsis forsvar signalering.
Resultaterne er præsenteret som et manuskript " The mRNA decay factor PAT1 functions in a
pathway including MAP kinase 4 and immune receptor SUMM2" offentliggjort i The EMBO
Journal og to udkast til manuskripter som opsummerer vores viden vedrørende AOC3 og eIF4E i
forhold til MPK4.
7
List of Abbreviations
4E-BP
eIF4E BINDING PROTEIN
ACD
ACCELERATED CELL DEATH
AGO
ARGONAUTE
AOC
ALLENE OXIDE CYCLASE
AOS
ALENE OXIDE SYNTHASE
ARF
AUXIN RESPONSE FACTOR
At
Arabidopsis thaliana
BAK1
BRI1-ASSOCIATED RECEPTOR KINASE1
BIK1
BOTRYTIS-INDUCED KINASE1
CC
Coiled-coiled
CDPK
Ca2+ dependent kinase
ClYVV
Clover Yellow Vein Virus
Col
Columbia
CSD
COPPER/ZINC SUPEROXIDE DISMUTASE
CTR
Constitutively Triple Response
DCL1
DICER-LIKE 1
DCP1
DECAPPING 1
EBF
EIN3 binding F-box protein
EDS1
ENHANCED DISEASE SUSCEPTIBILITY1
EFR
EF-TU RECEPTOR
eIF
eukaryotic INITIATION FACTOR
EIN
Ethylene Insensitive
ERF
Ethylene Response Factor
ET
Ethylene
ETI
Effector Triggered Immunity
FLS2
FLAGELLIN-SENSITIVE2
HR
Hypersensitivity response
JA
Jasmonic acid
LAZ
LAZARUS
Ler
Landsberg
LMM
Lesion Mimic Mutant
LOX
LIPOXYGENASE
8
MBP
MYELIN BASIC PROTEIN
MKS1
MAP KINASE 4 SUBSTRATE 1
MNK
MAPK-INTERACTING KINASE
MAPK
MAP Kinase
MAP2K
MAP Kinase Kinase
MAP3K
MAP Kinase Kinase Kinase
MPK4
MAP KINASE 4
NADPH
Nicotinamide adenine dinucleotide hosphate
NDR1
NONRACE-SPECIFIC DISEASE RESISTANCE1
NMD
Non-sense Meadiated Decay
NPR
NON-EXPRESSOR OF PR-GENES
OPDA
Oxo-phytodienoic acid
OPR
OPDA REDUCTASE
PAD
PHYTOALEXIN DEFICIENT
PAMP
Pathogen Associated Molecular Pattern
PABP
poly(A) binding protein
PAT
PROTEIN ASOIATED WITH TOPOISOMERASE-II
PBs
Processing Bodies
PCD
Programmed cell death
PR
Pathogenesis-Related
PRR
Pattern Recognizing Receptor
pst
Pseudomonas syringae pv. tomato
PTI
PAMP Triggered Immunity
PTC
Premature termination codon
PTGS
Post Transcriptionally Gene Silencing
RBOHD
RESPIRATORY BURST OXIDASE HOMOLOG D
R-gene
Resistance gene
R-protein
Resistance Protein
RIN4
RPM1-INTERACTING PROTEIN4
RIPK
RPM1-INDUCED PROTEIN KINASE
RISC
RNA-induced silencing complex
ROS
Reactive oxygen species
RPM1
RESISTANCE TO PSEUDOMONAS SYRINGAE
RPS2
RESISTANCE TO PSEUDOMONAS SYRINGAE2
9
SA
Salicylic acid
SAR
Systemic Acquired Resistance
SID2
ISOCHORISMATE SYNTHASE1
siRNA
Small interfering RNA
SP
Serine-Proline
SUMM2
SUPPRESSOR of MKK1 MKK2 2
TAL
Transcription Activator-Like
TF
Transcription factor
TIR
Toll and Interleukin-1 Receptor homology
TP
Threonine-Proline
VCS
VARICOSE
WT
Wild type
Xoo
Xanthomonas oryzae pathovar oryzae
XRN
Exoribonuclease
10
Plant innate immunity
Plants convert sunlight and atmospheric carbon dioxide into chemical energy in the form of
carbohydrates (Lodish et al., 2000). Thus, plants are a source of nutrient for numerous organisms. It
is logical to assume that plants defend themselves against pathogens or herbivores to avoid
depletion of their nutrients. Plants have therefore developed various defense mechanisms to avoid
hosting malevolent organisms (Katagiri and Tsuda, 2010). All biological processes, including
defense responses, are energy consuming, thus a tight regulation of defense can avoid superfluous
use of scarce resources. Defense responses are spatial and temporal regulated, in most cases by
direct or indirect recognition of pathogens. For example, plants can instigate defense by direct
recognition of bacterial flagellin or indirectly by recognizing host tissue damaged by pathogens or
herbivores (Gómez-Gómez and Boller, 2000; Krol et al., 2010). Plants have adapted both a
mechanical defense system in the form of structural fortification of the cell wall and active defense
system in form of both transmembrane and cytoplasmic immune receptors that upon activation
boosts defense responses (Boller and He, 2009; Malinovsky et al., 2014). The role of the rigid cell
wall and regulation of its components in response to pathogens are reviewed by Malinovsky et al
(2014) and will not be covered here whereas a general introduction to immune receptors and their
activation are described below.
Pathogen associated molecular patterns (PAMPs)
PAMP is a collective term used to describe conserved molecular patters found in microorganisms
that on their own is enough to induce defense responses. PAMPs are recognized by transmembrane
immune receptors located in the plasma membrane. The recognition of bacterial flagellin by the
pattern recognition receptor (PRR) FLAGELLIN-SENSING 2 (FLS2) in Arabidopsis is one of the
best characterized PAMP recognition systems in plants (Boller and Felix, 2009; Gómez-Gómez and
Boller, 2000). Another example is the perception of bacterial elongation factor-Tu by the EF-Tu
(EFR1) receptor (Kunze et al., 2004; Zipfel et al., 2006). Whereas FLS2 is widely conserved across
plant species, EFR belongs only to the Brassicaceae family (Zipfel et al., 2006). Transgenic
expression of Arabidopsis EFR in species outside the Brassicaceae family make plants sensitive to
EF-Tu induced immunity (Lacombe et al., 2010). Thus, signaling downstream of PRRs is likely
well conserved across species.
11
Both EFR and FLS2 hetero dimerizes with the co-receptor BRI1-ASSOCIATED KINASE 1
(BAK1) in a ligand induced interaction (Chinchilla et al., 2007, -; Roux et al., 2011; Schulze et al.,
2010). Phosphorylation of BOTRTYTIS-INDUCED KINASE 1 (BIK1) is induced upon activation
of both FLS2 and EFR (Lu et al., 2010; Zhang et al., 2010). In Arabidopsis BIK1 resides in a
complex with both BAK1 and FLS2 or EFR and upon PAMP perception BIK1 is rapidly
phosphorylated which trigger BIK1 to activate downstream immune responses (Lu et al., 2010;
Zhang et al., 2010).
One of the early effects of PAMP triggered immunity (PTI) is the generation of reactive oxygen
species (ROS) in an oxidative burst producing apoplastic superoxide (O2-), which can be converted
into hydrogen peroxide (H2O2) (Torres et al., 2006). The transmembrane RESPIRATORY BURST
OXIDASE HOMOLOG D (ROBHD) is a key enzyme in producing superoxide, it oxidize
intracellular NADPH and transports the electron to the apoplastic space where it reacts with
molecular oxygen producing superoxide (Torres et al., 2005, 2006). Production of ROS, like
hydrogen peroxide is directly toxic to pathogens and it induce cell wall strengthening while also
functioning as a systemic signaling molecules (Kadota et al., 2015; Miller et al., 2009). Recent
studies have shown that ROBHD sits in a complex with FLS2/EFR, BAK1 and BIK1, when PTI is
induced, BIK1 directly phosphorylates resides in RBOHD. When hindering this phosphorylation
RBOHD is not activated properly resulting in increased susceptibility against non-adapted
pathogens (Kadota et al., 2014; Li et al., 2014).
In general, PTI is relatively mild and mostly effective against non-adapted pathogens that do not
employ effector molecules evolved to circumvent PTI signaling. In addition to ROS signaling other
canonical signaling pathways include mitogen activated protein kinases (MAPKs) that for example
can adjust transcript accumulation of defense related genes. The involvement of MAPKs in PTI is
reviewed in more details in “MAP Kinases in Arabidopsis innate immunity”.
Pathogenic effector molecules and host resistance (R) proteins
Successful pathogens have evolved effector proteins or molecules that can modify the host and
support growth of the pathogen. Bacterial pathogens deliver these effectors by types of secretion
systems that can effectively inject effector proteins directly into the host cell. Mutations that disrupt
the type-three secretion system in gram-negative bacterial pathogens are often accompanied by
weakened virulence and are inadequate of overcoming PTI (Cunnac et al., 2009).
12
Effector proteins function in numerous ways. One example is Transcription Activator-Like (TAL)
effectors functioning as transcription factors, which bind the promoter of specific host genes and
induce their transcription. For effective infection in rice the pathogen Xanthomonas oryzae
pathovar oryzae (Xoo) deploy the TAL effector pthXo1 (Yang et al., 2006). PthXo1 binds the
promoter and induce the induction of OsSWEET11, which is important for Xoo infection (Chen et
al., 2010; Yang et al., 2006; Yuan et al., 2010). OsSWEET11 is presumably a sugar transporter that
induce sugar efflux in favor of the pathogen (Chen et al., 2010), however it has also been suggested
that OsSWEET11 functions as a cupper transporter that reduce the cupper concentration in the
xylem sap to facilitate infection of the cupper sensitive Xoo (Yuan et al., 2010). In Arabidopsis
infection with Pseudomonas (Pst) DC3000 induce accumulation of transcripts from 7 SWEET
genes, whereas infections with the Pst DC3000 HrcU mutant which cannot inject type threeeffector proteins into the host did not induce accumulation of three of these transcript (Chen et al.,
2010). Plants resistant to pathogens deploying specific TAL effectors have evolved defense genes
that are under control of promoters mimicking that of the TAL effector targets, which thereby are
transcribed in presence of the TAL effectors inducing a defense response (Hutin et al., 2015).
Effector proteins can also modify host proteins to quench host defense responses. RPM1
INTERACTING PROTEIN 4 (RIN4) is one of the best characterized targets of bacterial effectors.
RIN4 is anchored to the plasma membrane and resides in complex with or are in close proximity to
FLS2 and activation of FLS2 induces phosphorylation of RIN4 important for full activation of PTI
(Chung et al., 2014; Qi et al., 2011). Although the molecular function of RIN4 is not fully
understood, its role in immunity is emphasized by being a target of no less than four bacterial
effectors comprising AvrRpm1, AvrB, AvrRpt2 and HopF2 while being a guardee of the two R
proteins RPM1 and RPS2 in Arabidopsis (Axtell and Staskawicz, 2003, 2; Mackey et al., 2002,
2003; Selote and Kachroo, 2010; Wang et al., 2010; Wilton et al., 2010). These effector proteins
and conjugant R proteins function by different mechanisms. AvrRpt2 functions as a protease and
when introduced into the host it degrades RIN4, the lack of RIN4 in turns activate RPS2 causing
effector triggered immunity (ETI) (Axtell and Staskawicz, 2003; Mackey et al., 2003, 4). PTI and
ETI signaling is somewhat similar although ETI responses are more prolonged and pronounced and
can include localized cell death (Cui et al., 2014). Arabidopsis rps2 loss-of-function mutants are
more susceptible to AvrRpt2 expressing pathogens, thus it is likely that resistant plants have
evolved RPS2 in response AvrRpt2 mediated bacterial virulence to restore resistance (Afzal et al.,
2011).
13
AvrB and AvrRpm1 on the other hand mediate RIN4 phosphorylation through RPM1-INDUCED
PROTEIN KINASE (RIPK). FLS2 and RIPK mediated RIN4 phosphorylation occurs on different
residues and whereas FLS2 facilitates PTI through RIN4, RIPK mediated phosphorylation dampens
PTI. In Arabidopsis the R protein RPM1 sense RIPK mediated phosphorylation of RIN4 and
instigate ETI (Chung et al., 2011, 2014; Liu et al., 2011).
The fourth effector protein HopF2 inhibits accumulation of PAMP induced phosphorylated RIN4,
dampening PTI (Chung et al., 2014). Besides its interaction with RIN4, HopF2 also targets the
FLS2 co-receptor BAK1 and BIK1, thus HopF2 reduced accumulation of PAMP induced
phosphorylation of RIN4 is possible to occur at multiple levels (Chung et al., 2014; Wang et al.,
2010; Zhou et al., 2014). Expression of HopF2 in Arabidopsis dampens PTI and is not associated
with activation of ETI, therefore HopF2 represents a functional effector protein apparently not
detected by Arabidopsis R proteins.
Detection of AvrB, AvrRpm1 and AvrRpt2 modified RIN4 by RPM1 and RPS2 are examples of
indirect recognition of effector proteins. In this model host proteins are guarded by R proteins that
detect modification of their guardees (Dangl and Jones, 2001; Van Der Biezen and Jones, 1998).
Some R proteins directly recognize effector proteins and trigger ETI, this is true for Arabidopsis R
protein RRS1 which directly recognize the Ralstonia solanacearum effector protein AvrPopP2
(Deslandes et al., 2003).
A third class of R proteins has evolved as decoys, mimicking potential host effector targets. This
hypothesis was first suggested by Hoorn and Kamoun (2008). A direct example of this is the
Arabidopsis R protein RRS1 that contain a WRKY domain normally associated with transcription
factors. The R. solanacearum effector protein PopP2 directly acetylates a group of defense related
WRKY transcription factors inhibiting their DNA binding capabilities. PopP2 also bind the decoy
RRS1, which is paired with another R protein, RPS4 and when PopP2 bind the RRS1 decoy the
RRS1/RPS4 complex activate ETI (Le Roux et al., 2015; Sarris et al., 2015).
Autoimmunity
HR is often associated with ETI and is typically characterized by rapid, localized programmed cell
death at the site of pathogen invasion (Mur et al., 2008). Evidence linking ETI to HR include; (i)
presence of paired effectors and R proteins (ii) over-expression of R proteins and (iii) expression of
auto-active R proteins are all sufficient to in duce HR (Gao et al., 2011; Grant et al., 2000; Peart et
al., 2005; Stokes et al., 2002). Programmed cell death during HR seems to rely on light dependent
14
ROS production in the chloroplasts as inhibition or induction of chloroplast ROS production
respectively delay or increase cell death when applied with virulent pathogens (Zurbriggen et al.,
2010). Pathogen induced HR also appear to depend on autophagy as autophagy deficient mutants do
not display full HR (Hofius et al., 2009; Munch et al., 2015). It is however still unclear whether ETI
triggered HR is cause or a consequence of disease resistance. Some evidence support HR as a cause
of ETI against strictly biotrophic pathogens (Wang et al., 2011). Whereas other findings support
that HR can be uncoupled from ETI (Bendahmane et al., 1999; Bulgarelli et al., 2010; Coll et al.,
2010; Heidrich et al., 2011; Munch et al., 2015).
Autoimmune mutants display constitutive ETI phenotypes. We have previously characterized the
autoimmune mutants accelerated cell death 11 (acd11) and map kinase 4 (mpk4) (Brodersen et al.,
2002; Petersen et al., 2000). ACD11 is a putative sphingolipid transfer protein, but its precise role
during these processes is still unknown. acd11 is a so called lesion mimic mutants as it exhibits
constitutive defense responses and cell death without pathogen perception (Brodersen et al., 2002;
Palma et al., 2010). Activation of the R protein LAZARUS 5 (LAZ5) is responsible for the acd11
phenotype. LAZ5 is activated in absence ACD11, although the mechanism of this detection remains
unknown, but laz5 loss-of-function mutants completely rescue the lesion mimic phenotype (Palma
et al., 2010). The mpk4 mutant similarly exhibits autoimmunity caused at least by activation of the
R protein SUPRESSOR OF MKK1 MKK2 2 (SUMM2) (Petersen et al., 2000; Zhang et al., 2012).
The function of MPK4 is discussed in detail in “MAP Kinases in Arabidopsis innate immunity”. It
is clear that Arabidopsis loss-of-function mutants shadowed by autoimmune phenotypes are often
the result of inappropriate activation or R-proteins. However, since these mutants constitutively
exhibits immune responses they are often categorized as negative regulators of defense due to
pleiotropic effects of ETI (Rodriguez et al., 2015). It is therefore important to elucidate this issue
when investigating the true in vivo function of these proteins.
References
Afzal, A. J., Cunha, L. da, and Mackey, D. (2011). Separable Fragments and Membrane Tethering of
Arabidopsis RIN4 Regulate Its Suppression of PAMP-Triggered Immunity. Plant Cell 23,
3798–3811. doi:10.1105/tpc.111.088708.
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, 369–377.
15
Bendahmane, A., Kanyuka, K., and Baulcombe, D. C. (1999). The Rx Gene from Potato Controls
Separate Virus Resistance and Cell Death Responses. Plant Cell 11, 781–791.
doi:10.1105/tpc.11.5.781.
Boller, T., and Felix, G. (2009). A renaissance of elicitors: perception of microbe-associated
molecular patterns and danger signals by pattern-recognition receptors. Annu. Rev. Plant
Biol. 60, 379–406.
Boller, T., and He, S. Y. (2009). Innate Immunity in Plants: An Arms Race Between Pattern
Recognition Receptors in Plants and Effectors in Microbial Pathogens. Science 324, 742–
744. doi:10.1126/science.1171647.
Brodersen, P., Petersen, M., Pike, H. M., Olszak, B., Skov, S., Ødum, N., et al. (2002). Knockout of
Arabidopsis ACCELERATED-CELL-DEATH11 encoding a sphingosine transfer protein causes
activation of programmed cell death and defense. Genes Dev. 16, 490–502.
doi:10.1101/gad.218202.
Bulgarelli, D., Biselli, C., Collins, N. C., Consonni, G., Stanca, A. M., Schulze-Lefert, P., et al. (2010).
The CC-NB-LRR-Type Rdg2a Resistance Gene Confers Immunity to the Seed-Borne Barley
Leaf Stripe Pathogen in the Absence of Hypersensitive Cell Death. PLoS ONE 5, e12599.
doi:10.1371/journal.pone.0012599.
Chen, L.-Q., Hou, B.-H., Lalonde, S., Takanaga, H., Hartung, M. L., Qu, X.-Q., et al. (2010). Sugar
transporters for intercellular exchange and nutrition of pathogens. Nature 468, 527–532.
doi:10.1038/nature09606.
Chinchilla, D., Zipfel, C., Robatzek, S., Kemmerling, B., Nurnberger, T., Jones, J. D. G., et al. (2007).
A flagellin-induced complex of the receptor FLS2 and BAK1 initiates plant defence. Nature
448, 497–500. doi:10.1038/nature05999.
Chung, E.-H., da Cunha, L., Wu, A.-J., Gao, Z., Cherkis, K., Afzal, A. J., et al. (2011). Specific
Threonine Phosphorylation of a Host Target by Two Unrelated Type III Effectors Activates a
Host Innate Immune Receptor in Plants. Cell Host Microbe 9, 125–136.
doi:16/j.chom.2011.01.009.
Chung, E.-H., El-Kasmi, F., He, Y., Loehr, A., and Dangl, J. L. (2014). A Plant Phosphoswitch Platform
Repeatedly Targeted by Type III Effector Proteins Regulates the Output of Both Tiers of
Plant Immune Receptors. Cell Host Microbe 16, 484–494. doi:10.1016/j.chom.2014.09.004.
Coll, N. S., Vercammen, D., Smidler, A., Clover, C., van Breusegem, F., Dangl, J. L., et al. (2010).
Arabidopsis Type I Metacaspases Control Cell Death. Science.
doi:10.1126/science.1194980.
Cui, H., Tsuda, K., and Parker, J. E. (2014). Effector-Triggered Immunity: From Pathogen Perception
to Robust Defense. Annu. Rev. Plant Biol. doi:10.1146/annurev-arplant-050213-040012.
Cunnac, S., Lindeberg, M., and Collmer, A. (2009). Pseudomonas syringae type III secretion system
effectors: repertoires in search of functions. Curr. Opin. Microbiol. 12, 53–60.
doi:10.1016/j.mib.2008.12.003.
16
Dangl, J. L., and Jones, J. D. G. (2001). Plant pathogens and integrated defence responses to
infection. Nature 411, 826–833. doi:10.1038/35081161.
Deslandes, L., Olivier, J., Peeters, N., Feng, D. X., Khounlotham, M., Boucher, C., et al. (2003).
Physical interaction between RRS1-R, a protein conferring resistance to bacterial wilt, and
PopP2, a type III effector targeted to the plant nucleus. Proc. Natl. Acad. Sci. 100, 8024–
8029. doi:10.1073/pnas.1230660100.
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. Proc. Natl. Acad. Sci. 108, 7619–7624.
doi:10.1073/pnas.1104410108.
Gómez-Gómez, L., and Boller, T. (2000). FLS2: An LRR Receptor-like Kinase Involved in the
Perception of the Bacterial Elicitor Flagellin in Arabidopsis. Mol. Cell 5, 1003–1011.
doi:10.1016/S1097-2765(00)80265-8.
Grant, M., Brown, I., Adams, S., Knight, M., Ainslie, A., and Mansfield, J. (2000). The RPM1 plant
disease resistance gene facilitates a rapid and sustained increase in cytosolic calcium that is
necessary for the oxidative burst and hypersensitive cell death. Plant J. 23, 441–450.
doi:10.1046/j.1365-313x.2000.00804.x.
Heidrich, K., Wirthmueller, L., Tasset, C., Pouzet, C., Deslandes, L., and Parker, J. E. (2011).
Arabidopsis EDS1 Connects Pathogen Effector Recognition to Cell Compartment–Specific
Immune Responses. Science 334, 1401–1404. doi:10.1126/science.1211641.
Hofius, D., Schultz-Larsen, T., Joensen, J., Tsitsigiannis, D. I., Petersen, N. H. T., Mattsson, O., et al.
(2009). Autophagic Components Contribute to Hypersensitive Cell Death in Arabidopsis.
Cell 137, 773–783. doi:10.1016/j.cell.2009.02.036.
Hoorn, R. A. L. van der, and Kamoun, S. (2008). From Guard to Decoy: A New Model for Perception
of Plant Pathogen Effectors. Plant Cell 20, 2009–2017. doi:10.1105/tpc.108.060194.
Hutin, M., Pérez-Quintero, A. L., Lopez, C., and Szurek, B. (2015). MorTAL Kombat: the story of
defense against TAL effectors through loss-of-susceptibility. Plant Biot. Interact., 535.
doi:10.3389/fpls.2015.00535.
Kadota, Y., Shirasu, K., and Zipfel, C. (2015). Regulation of the NADPH Oxidase RBOHD During Plant
Immunity. Plant Cell Physiol. 56, 1472–1480. doi:10.1093/pcp/pcv063.
Kadota, Y., Sklenar, J., Derbyshire, P., Stransfeld, L., Asai, S., Ntoukakis, V., et al. (2014). Direct
Regulation of the NADPH Oxidase RBOHD by the PRR-Associated Kinase BIK1 during Plant
Immunity. Mol. Cell 54, 43–55. doi:10.1016/j.molcel.2014.02.021.
Katagiri, F., and Tsuda, K. (2010). Understanding the Plant Immune System. Mol. Plant. Microbe
Interact. 23, 1531–1536. doi:10.1094/MPMI-04-10-0099.
Krol, E., Mentzel, T., Chinchilla, D., Boller, T., Felix, G., Kemmerling, B., et al. (2010). Perception of
the Arabidopsis Danger Signal Peptide 1 Involves the Pattern Recognition Receptor
17
AtPEPR1 and Its Close Homologue AtPEPR2. J. Biol. Chem. 285, 13471–13479.
doi:10.1074/jbc.M109.097394.
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. Plant Cell 16,
3496–3507. doi:10.1105/tpc.104.026765.
Lacombe, S., Rougon-Cardoso, A., Sherwood, E., Peeters, N., Dahlbeck, D., van Esse, H. P., et al.
(2010). Interfamily transfer of a plant pattern-recognition receptor confers broad-spectrum
bacterial resistance. Nat. Biotechnol. 28, 365–369. doi:10.1038/nbt.1613.
Le Roux, C., Huet, G., Jauneau, A., Camborde, L., Trémousaygue, D., Kraut, A., et al. (2015). A
Receptor Pair with an Integrated Decoy Converts Pathogen Disabling of Transcription
Factors to Immunity. Cell 161, 1074–1088. doi:10.1016/j.cell.2015.04.025.
Li, L., Li, M., Yu, L., Zhou, Z., Liang, X., Liu, Z., et al. (2014). The FLS2-Associated Kinase BIK1 Directly
Phosphorylates the NADPH Oxidase RbohD to Control Plant Immunity. Cell Host Microbe
15, 329–338. doi:10.1016/j.chom.2014.02.009.
Liu, J., Elmore, J. M., Lin, Z.-J. D., and Coaker, G. (2011). A Receptor-like Cytoplasmic Kinase
Phosphorylates the Host Target RIN4, Leading to the Activation of a Plant Innate Immune
Receptor. Cell Host Microbe 9, 137–146. doi:16/j.chom.2011.01.010.
Lodish, H., Berk, A., Zipursky, S. L., Matsudaira, P., Baltimore, D., and Darnell, J. (2000). CO2
Metabolism during Photosynthesis. Available at:
http://www.ncbi.nlm.nih.gov/books/NBK21472/ [Accessed December 23, 2015].
Lu, D., Wu, S., Gao, X., Zhang, Y., Shan, L., and He, P. (2010). A receptor-like cytoplasmic kinase,
BIK1, associates with a flagellin receptor complex to initiate plant innate immunity. Proc.
Natl. Acad. Sci. 107, 496–501. doi:10.1073/pnas.0909705107.
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 RPS2-mediated resistance.
Cell 112, 379–389.
Mackey, D., Holt III, B. F., Wiig, A., and Dangl, J. L. (2002). RIN4 interacts with Pseudomonas
syringae type III effector molecules and is required for RPM1-mediated resistance in
Arabidopsis. Cell 108, 743–754.
Malinovsky, F. G., Fangel, J. U., and Willats, W. G. T. (2014). The role of the cell wall in plant
immunity. Plant Biot. Interact. 5, 178. doi:10.3389/fpls.2014.00178.
Miller, G., Schlauch, K., Tam, R., Cortes, D., Torres, M. A., Shulaev, V., et al. (2009). The Plant
NADPH Oxidase RBOHD Mediates Rapid Systemic Signaling in Response to Diverse Stimuli.
Sci Signal 2, ra45–ra45. doi:10.1126/scisignal.2000448.
Munch, D., Teh, O.-K., Malinovsky, F. G., Liu, Q., Vetukuri, R. R., Kasmi, F. E., et al. (2015). Retromer
Contributes to Immunity-Associated Cell Death in Arabidopsis. Plant Cell 27, 463–479.
doi:10.1105/tpc.114.132043.
18
Mur, L. A. J., Kenton, P., Lloyd, A. J., Ougham, H., and Prats, E. (2008). The hypersensitive response;
the centenary is upon us but how much do we know? J. Exp. Bot. 59, 501–520.
doi:10.1093/jxb/erm239.
Palma, K., Thorgrimsen, S., Malinovsky, F. G., Fiil, B. K., Nielsen, H. B., Brodersen, P., et al. (2010).
Autoimmunity in Arabidopsis acd11 Is Mediated by Epigenetic Regulation of an Immune
Receptor. PLoS Pathog 6, e1001137. doi:10.1371/journal.ppat.1001137.
Peart, J. R., Mestre, P., Lu, R., Malcuit, I., and Baulcombe, D. C. (2005). NRG1, a CC-NB-LRR Protein,
together with N, a TIR-NB-LRR Protein, Mediates Resistance against Tobacco Mosaic Virus.
Curr. Biol. 15, 968–973. doi:10.1016/j.cub.2005.04.053.
Petersen, M., Brodersen, P., Naested, H., Andreasson, E., Lindhart, U., Johansen, B., et al. (2000).
Arabidopsis MAP Kinase 4 Negatively Regulates Systemic Acquired Resistance. Cell 103,
1111–1120. doi:10.1016/S0092-8674(00)00213-0.
Qi, Y., Tsuda, K., Glazebrook, J., and Katagiri, F. (2011). Physical association of pattern-triggered
immunity (PTI) and effector-triggered immunity (ETI) immune receptors in Arabidopsis.
Mol. Plant Pathol. 12, 702–708. doi:10.1111/j.1364-3703.2010.00704.x.
Rodriguez, E., el Ghoul, H., Mundy, J., and Petersen, M. (2015). Making sense of plant
autoimmunity and “negative regulators.” FEBS J., n/a–n/a. doi:10.1111/febs.13613.
Roux, M., Schwessinger, B., Albrecht, C., Chinchilla, D., Jones, A., Holton, N., et al. (2011). The
Arabidopsis Leucine-Rich Repeat Receptor–Like Kinases BAK1/SERK3 and BKK1/SERK4 Are
Required for Innate Immunity to Hemibiotrophic and Biotrophic Pathogens. Plant Cell
Online 23, 2440–2455. doi:10.1105/tpc.111.084301.
Sarris, P. F., Duxbury, Z., Huh, S. U., Ma, Y., Segonzac, C., Sklenar, J., et al. (2015). A Plant Immune
Receptor Detects Pathogen Effectors that Target WRKY Transcription Factors. Cell 161,
1089–1100. doi:10.1016/j.cell.2015.04.024.
Schulze, B., Mentzel, T., Jehle, A. K., Mueller, K., Beeler, S., Boller, T., et al. (2010). Rapid
Heteromerization and Phosphorylation of Ligand-activated Plant Transmembrane
Receptors and Their Associated Kinase BAK1. J. Biol. Chem. 285, 9444–9451.
doi:10.1074/jbc.M109.096842.
Selote, D., and Kachroo, A. (2010). RIN4-like proteins mediate resistance protein-derived soybean
defense against Pseudomonas syringae. Plant Signal. Behav. 5, 1453–1456.
doi:10.4161/psb.5.11.13462.
Stokes, T. L., Kunkel, B. N., and Richards, E. J. (2002). Epigenetic variation in Arabidopsis disease
resistance. Genes Dev. 16, 171–182. doi:10.1101/gad.952102.
Torres, M. A., Jones, J. D. G., and Dangl, J. L. (2005). Pathogen-induced, NADPH oxidase–derived
reactive oxygen intermediates suppress spread of cell death in Arabidopsis thaliana. Nat.
Genet. 37, 1130–1134. doi:10.1038/ng1639.
Torres, M. A., Jones, J. D. G., and Dangl, J. L. (2006). Reactive Oxygen Species Signaling in Response
to Pathogens. Plant Physiol. 141, 373–378. doi:10.1104/pp.106.079467.
19
Van Der Biezen, E. A., and Jones, J. D. G. (1998). Plant disease-resistance proteins and the genefor-gene concept. Trends Biochem. Sci. 23, 454–456. doi:10.1016/S0968-0004(98)01311-5.
Wang, W., Barnaby, J. Y., Tada, Y., Li, H., Tör, M., Caldelari, D., et al. (2011). Timing of plant
immune responses by a central circadian regulator. Nature 470, 110–114.
doi:10.1038/nature09766.
Wang, Y., Li, J., Hou, S., Wang, X., Li, Y., Ren, D., et al. (2010). A Pseudomonas syringae ADPRibosyltransferase Inhibits Arabidopsis Mitogen-Activated Protein Kinase Kinases. Plant Cell
22, 2033–2044. doi:10.1105/tpc.110.075697.
Wilton, M., Subramaniam, R., Elmore, J., Felsensteiner, C., Coaker, G., and Desveaux, D. (2010).
The type III effector HopF2 Pto targets Arabidopsis RIN4 protein to promote Pseudomonas
syringae virulence. Proc. Natl. Acad. Sci. 107, 2349–2354. doi:10.1073/pnas.0904739107.
Yang, B., Sugio, A., and White, F. F. (2006). Os8N3 is a host disease-susceptibility gene for bacterial
blight of rice. Proc. Natl. Acad. Sci. 103, 10503–10508. doi:10.1073/pnas.0604088103.
Yuan, M., Chu, Z., Li, X., Xu, C., and Wang, S. (2010). The Bacterial Pathogen Xanthomonas oryzae
Overcomes Rice Defenses by Regulating Host Copper Redistribution. Plant Cell 22, 3164–
3176. doi:10.1105/tpc.110.078022.
Zhang, J., Li, W., Xiang, T., Liu, Z., Laluk, K., Ding, X., 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.
doi:10.1016/j.chom.2010.03.007.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., et al. (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.
doi:10.1016/j.chom.2012.01.015.
Zhou, J., Wu, S., Chen, X., Liu, C., Sheen, J., Shan, L., et al. (2014). The Pseudomonas syringae
effector HopF2 suppresses Arabidopsis immunity by targeting BAK1. Plant J. 77, 235–245.
doi:10.1111/tpj.12381.
Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J. D. G., Boller, T., et al. (2006). Perception of
the Bacterial PAMP EF-Tu by the Receptor EFR Restricts Agrobacterium-Mediated
Transformation. Cell 125, 749–760. doi:10.1016/j.cell.2006.03.037.
Zurbriggen, M. D., Carrillo, N., and Hajirezaei, M.-R. (2010). ROS signaling in the hypersensitive
response. Plant Signal. Behav. 5, 393–396.
20
MINI REVIEW ARTICLE
published: 24 July 2012
doi: 10.3389/fpls.2012.00169
MAP kinase cascades in Arabidopsis innate immunity
Magnus W. Rasmussen, Milena Roux, Morten Petersen and John Mundy*
Department of Biology, University of Copenhagen, Copenhagen, Denmark
Edited by:
Alex Jones, The Sainsbury Laboratory,
UK
Reviewed by:
Birgit Kersten, Johann Heinrich von
Thünen Institute, Germany
Jesus V. Jorrin Novo, University of
Cordoba, Spain
*Correspondence:
John Mundy, Department of
Biology, University of Copenhagen,
Ole Maaløes Vej 5, 2200
Copenhagen, Denmark.
e-mail: [email protected]
Plant mitogen-activated protein kinase (MAPK) cascades generally transduce extracellular
stimuli into cellular responses. These stimuli include the perception of pathogen-associated
molecular patterns (PAMPs) by host transmembrane pattern recognition receptors which
trigger MAPK-dependent innate immune responses. In the model Arabidopsis, molecular
genetic evidence implicates a number of MAPK cascade components in PAMP signaling,
and in responses to immunity-related phytohormones such as ethylene, jasmonate, and salicylate. In a few cases, cascade components have been directly linked to the transcription
of target genes or to the regulation of phytohormone synthesis. Thus MAPKs are obvious
targets for bacterial effector proteins and are likely guardees of resistance proteins, which
mediate defense signaling in response to the action of effectors, or effector-triggered immunity. This mini-review discusses recent progress in this field with a focus on the Arabidopsis
MAPKs MPK3, MPK4, MPK6, and MPK11 in their apparent pathways.
Keywords: calcium signaling, hypersensitive response, MAP kinase cascade, MAP kinase substrates, pathogen
effectors, pattern recognition receptors, reactive oxygen species, resistance proteins
INTRODUCTION
Plants have evolved an effective basal defense system to detect
and limit the growth of pathogens. Pathogens may be recognized
by the host via the perception of conserved microbial structures
termed pathogen-associated molecular patterns (PAMPs). PAMPs
are recognized via transmembrane pattern recognition receptors (PRRs) that bind specific PAMPs and initiate intracellular
immune responses (Zipfel, 2008). These PAMP-triggered immunity (PTI) responses include the generation of reactive oxygen
species (ROS), extracellular alkalinization, and protein phosphorylation with associated gene regulation that ultimately restricts
the growth of the microbial intruder (Gimenez-Ibanez and
Rathjen, 2010).
Mitogen-activated protein kinase (MAPK) signaling plays central roles in such intracellular immunity pathways. In general,
MAP kinase signaling is initiated by the stimulus-triggered activation of a MAP kinase kinase kinase (MAP3K; also called MEKK).
MAP3K activation, which may be directly or indirectly effected
by a PRR, in turn leads to the phosphorylation and activation
of downstream MAP kinase kinases (MAP2K; also called MKK or
MEK). Subsequently, the MAP2K phosphorylates the downstream
MAPK sequentially leading to changes in its subcellular localization and/or phosphorylation of downstream substrates including
transcription factors which alter patterns of gene expression (see
Figure 1). General functions of MAPK cascades in plant biology
have recently been reviewed elsewhere (Fiil et al., 2009; Rodriguez
et al., 2010; Komis et al., 2011).
MAPK CASCADES IN PTI
A few PRRs have been documented to stimulate MAPK signaling
upon perception of PAMPs. These include the flagellin receptor
FLS2 (Felix et al., 1999; Gómez-Gómez and Boller, 2000), the bacterial elongation factor EF-Tu receptor EFR (Zipfel et al., 2006),
and the chitin receptor CERK1 (Miya et al., 2007).
The Arabidopsis genome encodes 60 MAP3Ks, 10 MAP2Ks,
and 20 MAPKs (Ichimura et al., 2002). This indicates that MAPK
cascades may not simply consist of single MAP3Ks, MAP2Ks,
and MAPKs connected together. Instead, it suggests that there
is some level of redundancy, and that the spatial and temporal
activities of different components may be strictly regulated to
minimize wanton cross-talk. The three MAPKs MPK3, MPK4,
and MPK6 are the most intensively studied plant MAPKs, and all
three were implicated in defense signaling a decade ago (Petersen
et al., 2000; Asai et al., 2002). MPK11, a close homolog to MPK4,
has also recently been shown to be activated by PAMP treatment
(Bethke et al., 2012).
MPK3, MPK4, and MPK6 are all activated by PAMPs such
as flg22 (a conserved 22 amino acid flagellin peptide) and elf18
(elongation factor-Tu peptide; Felix et al., 1999; Zipfel et al.,
2006). However, these three MAPK cascades are differently regulated already at the PRR level. For example, the two receptor
kinases BAK1 and BKK1 genetically regulate PAMP signaling
through their interactions with cognate PRRs (Roux et al., 2011;
Schwessinger et al., 2011). The BAK1 mutant allele bak1-5 carries a Cys408Tyr substitution adjacent to its kinase catalytic loop.
This impairs its flg22-regulated kinase activity and inhibits phosphorylation of MPK4. However, the catalytic complex formed
between mutant BAK1 in bak1-5 and FLS2 is still able to induce
phosphorylation of MPK3/MPK6 (Roux et al., 2011; Schwessinger
et al., 2011). Interestingly, MPK3/MPK6 phosphorylation was
impaired in only the double bak1-5 bkk1 background and not
in the individual bak1-5 and bkk1 lines (Roux et al., 2011).
Asai et al. (2002) developed an elegant protoplast expression
system in an attempt to identify signaling components downstream of FLS2. With this system they were able to show a
complete MAPK cascade downstream of FLS2 consisting of the
MAP3K MEKK1, two MAP2Ks (MKK4 and MKK5), and the
MAPKs MPK3/MPK6. However, genetic evidence later showed
21
“fpls-03-00169” — 2012/7/22 — 17:06 — page 1 — #1
Rasmussen et al.
MAP kinase cascades in immunity
FIGURE 1 | (A) MAPK signaling cascades are attractive targets for
bacterial effectors. The P. syringae HopAI1 effector irreversibly inactivates
MPK4 to prevent immune responses. The R protein SUMM2 may guard
processes downstream of MPK4 independent from MKS1, and triggers
a hypersensitive response in the event of loss or inactivation of MPK4.
(B) PAMP perception by PRRs instigates a signaling cascade, often via
that MEKK1 kinase activity was dispensable for MPK3/MPK6
activation, although mekk1 plants were impaired in MPK4 activation (Rodriguez et al., 2007). Interestingly, expressing a kinase
dead version of MEKK1 in mekk1 plants completely restored the
activation of MPK4 upon treatment with flg22, suggesting that
MEKK1 may “simply” act as a scaffold protein (Rodriguez et al.,
2007). Biochemical and genetic studies further revealed that the
two MAP2Ks MKK1 and MKK2 interact with both MEKK1 and
with MPK4, and that flg22-induced MPK4 activation is impaired
in the double mkk1 mkk2 mutant. This indicates that MKK1 and
MKK2 are partially redundant in MPK4 mediated downstream
signaling (Gao et al., 2008; Qiu et al., 2008b).
MPK4 was originally reported as a negative regulator of plant
immunity because the mpk4 mutant accumulates high levels of
salicylic acid, constitutively expresses pathogenesis-related (PR)
genes, and has a severely dwarfed growth phenotype (Petersen
et al., 2000). This phenotype is very similar to that of the mekk1
single and mkk1 mkk2 double mutants, further supporting their
functional relationships (Rodriguez et al., 2007; Gao et al., 2008;
Qiu et al., 2008b).
MAPK CASCADES IN EFFECTOR-TRIGGERED IMMUNITY
In addition to PTI, plants also employ resistance (R) proteins as
cytoplasmic receptors to directly or indirectly recognize specific
pathogenic effector proteins injected into host cells as virulence
co-receptors, which causes activation of MAP3K MEKK1 and two
MAP2Ks MKK1 and MKK2. These phosphorylate and activate MPK4
which then phosphorylates its substrate MKS1, releasing MKS1 in
complex with WRKY33. MPK3/MPK6 sequentially phosphorylate
WRKY33 allowing it to promote PAD3 transcription, thus activating
plant defense.
factors. Effector-triggered immunity (ETI) and PTI share a number of responses, although ETI also includes varying levels of rapid,
localized cell death in what is called the hypersensitive response. R
protein-dependent recognition initiates immune responses in ETI.
R proteins may recognize effector proteins either directly or indirectly by monitoring changes in the effector’s host target(s). This
latter case gave rise to the guard hypothesis in which R proteins
guard host guardees that are manipulated by pathogen effectors
(Van Der Biezen and Jones, 1998).
The genetic characterization of the MEKK1/MKK1–MKK2/
MPK4 cascade as a negative regulatory pathway of defense
responses was at odds with the activation of the pathway by
PAMPs. Instead, it was possible that the severe phenotypes of
the kinase knockout mutants were caused by activation of one or
more R protein(s) guarding this kinase pathway. Indeed, in an
elegant screen for suppressors of the mkk1 mkk2 double mutant,
Zhang et al. (2012) identified the R protein SUMM2 (suppressor
of mkk1 mkk2). The T-DNA insertion line summ2-8 completely
suppressed the severe mkk1 mkk2 phenotype in respect to morphology, cell death, ROS levels and PR gene expression (Zhang
et al., 2012). The analogous knockout phenotype of the upstream
MAP3K mekk1 is also completely suppressed in the summ2-8 background. Interestingly, although the mpk4 mutant shares a similar
phenotype with the knockouts of its upstream kinase partners,
the mpk4 phenotype is not fully suppressed by the summ2-8
22
“fpls-03-00169” — 2012/7/22 — 17:06 — page 2 — #2
Rasmussen et al.
MAP kinase cascades in immunity
mutation, as double mpk4 summ2-8 mutants still retain residual cell death and low levels of ROS. This suggests that MPK4
is involved in other pathways independent of SUMM2, and that
MPK4 may be guarded by additional R proteins (Zhang et al., 2012;
Figure 1A).
The importance of MAPK signaling in immunity is emphasized by studies reporting bacterial effector proteins targeting
MAPK cascades for downregulation (Zhang et al., 2007a,b, 2012;
Cui et al., 2010). For example, the Pseudomonas syringae effector
protein HopAI1 targets and irreversibly inactivates MPK3, MPK4,
and MPK6, thereby suppressing immune responses which would
otherwise inhibit bacterial growth (Zhang et al., 2007a, 2012). In
addition, the P. syringae effector protein AvrB has been shown to
interact with and induce the phosphorylation of MPK4, although
it has not been shown if this phosphorylation occurs as a direct
effect of AvrB action or via recognition of AvrB by the plant
immune system (Cui et al., 2010).
In plants carrying functional SUMM2 alleles, immune
responses are activated by bacterial effector proteins targeting the
MPK4 pathway (Figure 1A). For example, inducible expression
of the bacterial HopAI1 effector in wild-type plants gives rise to
a defense phenotype similar to that seen in mekk1, mkk1 mkk2,
and mpk4 mutants including elevated levels of ROS, PR gene
expression, and cell death (Zhang et al., 2012). SUMM2 apparently does not interact directly with the kinase components of the
MEKK1/MKK1–MKK2/MPK4 signaling cascade, suggesting that
SUMM2 most likely guards a downstream target of MPK4 activity
(Zhang et al., 2012). At present, the best studied in vivo substrate
of MPK4 activity is MPK4 substrate 1 (MKS1) which forms a
nuclear complex with MPK4 and the WRKY33 transcription factor (Andreasson et al., 2005; Qiu et al., 2008a). Phosphorylation
of MKS1 follows MPK4 activation by flg22 perception and, once
phosphorylated, MKS1 is released from complexes with MPK4,
thereby releasing the WRKY33 transcription factor to bind to its
cognate target genes (Qiu et al., 2008a). It has therefore been proposed that MPK4 and MKS1 sequester WRKY33 in the absence of
pathogens, and free WRKY33 to induce resistance upon pathogen
perception (Figure 1B, left).
As MKS1 is the only known direct target of MPK4, Zhang
et al. (2012) tested whether MKS1 interacted with the R protein SUMM2 that seemingly guards MPK4 activity. However, no
interaction between SUMM2 and MKS1 was detected. Since mks1
mutants have a wild-type growth phenotype, and the mpk4 phenotype is strongly suppressed in the mks1 background, SUMM2
may guard a process downstream of MPK4 that is independent of
MKS1 (Petersen et al., 2010).
WRKY TRANSCRIPTION FACTORS
The plant-specific WRKY family is a large group of transcription factors which bind a conserved W-box sequence in the
promoters of numerous genes including those encoding PR
proteins. WRKY33 was found to induce the transcription of PHYTOALEXIN DEFICIENT 3 (PAD3) which encodes the cytochrome
P450 monooxygenase 71B15 required for synthesis of the antimicrobial compound camalexin (Zhou et al., 1999; Qiu et al., 2008a;
Figure 1B). The wrky33 mutant exhibits enhanced susceptibility toward necrotrophic pathogens such as Botrytis cinerea, while
WRKY33 overexpression results in increased resistance due to
enhanced PAD3 expression (Zheng et al., 2006).
MPK3 and MPK6 activities also induce the production of
camalexin. Transient overexpression of the constitutively active,
phospho-mimic mutant forms of MKK4/MKK5 (MKK4DD and
MKK5DD ), which are the upstream MAP2Ks of MPK3/MPK6,
has been reported to induce transcription of both PAD2, which
encodes γ-glutamylcysteine synthetase functioning in glutathione
biosynthesis, and PAD3. Both PAD2 and PAD3 are necessary
for camalexin production (Parisy et al., 2007; Ren et al., 2008).
Pathogen-induced camalexin accumulation is partially comprised
in mpk3 but not notably in mpk6 mutants, yet camalexin accumulation in mpk3 mpk6 double mutants is almost completely
abolished (Ren et al., 2008). While this implicates MPK3/MPK6
in camalexin synthesis, caution should be applied in evaluating results obtained from the mpk3 mpk6 double mutant as it
is arrested at the cotyledon stage and is unable to initiate true
leaves (Wang et al., 2007). Upstream of MPK3/MPK6 in camalexin
induction, MKK4 and MKK5 are activated by the MAP3Ks
MEKK1 and MAPKKKα in response to fungal pathogens (Ren
et al., 2008). Yet another MAP2K, MKK9, whose upstream
MAP3K(s) remains unidentified, is also involved in MPK3/MPK6
activation, as plants expressing phospho-mimic MKK9DD produce even more camalexin than plants expressing MKK4DD or
MKK5DD (Xu et al., 2008).
To delineate the link between MPK3/MPK6 activation and
camalexin accumulation, Mao et al. (2011) elegantly introduced
the phospho-mimic mutant NtMEK2DD , an MKK4 and/or MKK5
ortholog from Nicotiana tabacum, into an array of different wrky
mutants in a search for essential transcription factors involved
in MPK3/MPK6 mediated camalexin induction. Interestingly,
NtMEKK2DD was able to induce camalexin accumulation in all
tested mutant lines except wrky33. In addition, WRKY33 proved
to be a substrate of MKP3/MPK6 activity, and overexpression of
non-phosphorylatable forms of WRKY33 could not fully complement the inability of wrky33 mutants to express PAD3 and
accumulate camalexin (Mao et al., 2011; Figure 1B, right).
WRKY33-induced PAD3 expression therefore appears to
involve both MPK4- and MPK3/MPK6-mediated signaling
(Andreasson et al., 2005; Qiu et al., 2008a; Mao et al., 2011). Mao
et al. (2011) proposed a model in which PAD3-mediated camalexin
induction occurs differentially depending on the type of pathogen
causing the immune response. In this model, bacterial pathogens
induce an MPK4 mediated response while fungal pathogens initiate an MPK3/MPK6 mediated response. This hypothesis is
based on overexpression of the constitutively active MKK4/MKK5
ortholog NtMEKK2DD , rendering MPK3/MPK6 hyperactive and
able to induce PAD3 expression (Mao et al., 2011). In support of
this hypothesis, the mpk3 mpk6 double mutant is comprised in B.
cinerea-induced PAD3 induction (Ren et al., 2008). Nonetheless,
and as noted above, some care should be taken with experiments
based on mpk3 mpk6 double mutants given their developmental
lethality (Wang et al., 2007).
An alternative model may therefore be proposed which combines the MPK4 and MPK3/MPK6 pathways into a dual control of
PAD3 regulation in response to pathogen perception (Figure 1B).
In such a model, WRKY33 is sequestered in a nuclear complex
23
“fpls-03-00169” — 2012/7/22 — 17:06 — page 3 — #3
Rasmussen et al.
MAP kinase cascades in immunity
comprising at least MPK4 and MKS1 in unchallenged plants,
and is released following PAMP perception (Qiu et al., 2008a).
Phosphorylation is dispensable for WRKY33 to bind its cognate
W-box cis-elements, although it does promote transcriptional
activation (Mao et al., 2011). This is illustrated by the fact that
PAD3 expression is induced in mpk4 plants (Qiu et al., 2008a),
perhaps due to the basal activity of free non-phosphorylated
WRKY33 or by free WRKY33 activated by basal MPK3 and/or
MPK6 activity. In this scenario, once WRKY33 is released from
its nuclear complex with MPK4 and MKS1, it is phosphorylated
and hence activated by MPK3/MPK6, thereby inducing camalexin
levels through PAD3 expression. The elevated PAD3 expression
induced from NtMEKK2DD hyper-activated MPK3/MPK6 (Mao
et al., 2011) is not in conflict with this model, as it is likely that
hyperactive MPK3/MPK6 are able to phosphorylate residual free
WRKY33, thus bypassing other possible feedback mechanisms in
PAD3 expression.
In this model, MPK4 and MPK3/MPK6 function together
as a binary switch conferring dual level regulation. Clarification of the mode of action in which MPK4 and MPK3/MPK6
function clearly needs further elucidation and should include
experiments using catalytically inactive and/or inactivatable
MPK4 (Petersen et al., 2000; Brodersen et al., 2006). Application of fungal PAMPs to plants expressing catalytically inactive
MPK4 might indicate whether phosphorylation of free WRKY33
by endogenous MPK3/MPK6 is enough to induce expression
of PAD3.
MAPK IN GENERAL STRESS SIGNALING
The refined work of Popescu et al. (2009) identified a MAP2K–
MAPK phosphorylation network covering 570 MAPK substrates
by combinatorially pairing active MAP2Ks with MAPKs, and then
subjecting them to a protein microarray phosphorylation assay.
Interestingly, the substrates identified were enriched for transcription factors involved in stress responses. Notably, MPK6
phosphorylated 32% of the identified targets, of which 40%
overlapped with MPK3 targets (Popescu et al., 2009). This is
in agreement with earlier data, similarly obtained from a protein microarray study (Feilner et al., 2005). Equally noteworthy
is the finding that MPK3 also shared 50% of its targets with
MPK4, revealing intensive synergy in MAPK signaling (Popescu
et al., 2009).
In addition to MAPK cascades, ROS also play a pivotal role in
stress signaling (Rodriguez et al., 2010). OXI1, a serine/threonine
kinase induced by general ROS-generating stimuli, is required for
full activation of MPK3/MPK6 after treatment with H2 O2 (Rentel
et al., 2004). Although OXI1 is characterized as an upstream regulator of MPK3/MPK6 activation, MPK3/MPK6 have been shown
to phosphorylate OXI1 in vitro. This suggests that there is a
REFERENCES
Andreasson, E., Jenkins, T., Brodersen, P., Thorgrimsen, S., Petersen, N.
H., Zhu, S., Qiu, J.-L., Micheelsen,
P., Rocher, A., Petersen, M., Newman, M. A., Bjørn Nielsen, H., Hirt,
H., Somssich, I., Mattsson, O., and
Mundy, J. (2005). The MAP kinase
substrate MKS1 is a regulator of
plant defense responses. EMBO J. 24,
2579–2589.
Asai, T., Tena, G., Plotnikova, J., Willmann, M. R., Chiu, W. L., GomezGomez, L., Boller, T., Ausubel,
F. M., and Sheen, J. (2002).
MAP kinase signalling cascade in
feedback loop, but in vivo data supporting such a loop has not
been shown (Forzani et al., 2011).
In addition to MAPK cascade signaling, PAMP perception also
induces Ca2+ dependent kinases (CDPKs) by regulating Ca2+
influx channels (Ma et al., 2009; Kwaaitaal et al., 2011). Recent
findings indicate that Ca2+ ATPases regulate Ca2+ efflux and function to regulate innate immune defenses (Zhu et al., 2010). Of
particular interest is the Ca2+ ATPase ACA8 which was shown to
interact with FLS2, and which may well regulate CDPK signaling
through flg22 perception (Frei dit Frey et al., 2012).
MPK8 activity has been shown to negatively regulate the expression of OXI1 in order to maintain ROS homeostasis. Remarkably,
activation of MPK8 is not limited to the upstream MAP2K MKK3,
as the Ca2+ binding protein calmodulin (CaM) is able to bind and
activate MPK8 in an Ca2+ -dependent manner (Takahashi et al.,
2011). CaM-mediated MPK8 activation is interesting because it
bypasses the traditional, sequential activation of MAPKs and also
unequivocally links MAPK activation with the ROS burst and
ion flux during stress signaling. In addition, CaM also mediates MAPK downregulation. MAP kinase phosphatase 1 (MKP1),
which interacts with MPK3, MPK4, and MPK6 (Ulm et al., 2002),
binds CaM in a Ca2+ -dependent manner and stimulates MKP1
phosphatase activity (Lee et al., 2008). The associations between
CDPKs and MAPK cascades have recently been review elsewhere
(Wurzinger et al., 2011).
Much progress has been made in understanding how MAPK
signaling functions in plant immunity. In Arabidopsis, 3 of the 60
identified MAP3Ks are involved in defense, namely MEKK1 (Asai
et al., 2002), EDR1 (Frye et al., 2001), and MEKKα (del Pozo et al.,
2004; Ren et al., 2008). In addition, at least 6 of the 10 identified
MAP2Ks (MKK1, MKK2, MKK4, MKK5, MKK7, and MKK9) are
involved in defense signaling (Asai et al., 2002; Djamei et al., 2007;
Dóczi et al., 2007; Zhang et al., 2007b; Yoo et al., 2008). This situation requires tight regulation of the spatial and temporal kinase
activities in order to impose specificity upon downstream signaling. To shed light on this regulation, high-throughput methods
such as those used by Popescu et al. (2009) are particularly valuable
and help to outline MAPK signaling cascades. While this progress
may be lauded, further work needs to focus on identifying direct,
in vivo kinase substrates and their respective phosphorylation sites.
This may bring us closer to bridging the apparent gap between
PRRs and MAPK cascades, and to understanding how specificity
is achieved among MAPK pathways both spatially and temporally.
ACKNOWLEDGMENTS
This work was supported by grants to John Mundy from the Danish Research Council for Technology and Production (23-03-0076)
and the Strategic Research Council (09-067148) and to Milena
Roux from the Natural Science Council (11-116368).
Arabidopsis innate immunity. Nature
415, 977–983.
Bethke, G., Pecher, P., EschenLippold, L., Tsuda, K., Katagiri,
F., Glazebrook, J., Scheel, D.,
and Lee, J. (2012). Activation of
the Arabidopsis thaliana mitogenactivated protein kinase MPK11 by
the flagellin-derived elicitor peptide,
flg22. Mol. Plant Microbe Interact. 25,
471–480.
Brodersen, P., Petersen, M., Nielsen, H.
B., Zhu, S., Newman, M., Shokat, K.
M., Rietz, S., Parker, J., and Mundy,
J. (2006). Arabidopsis MAP kinase 4
regulates salicylic acid- and jasmonic
24
“fpls-03-00169” — 2012/7/22 — 17:06 — page 4 — #4
Rasmussen et al.
acid/ethylene-dependent responses
via EDS1 and PAD4. Plant J. 47,
532–546.
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.
del Pozo, O., Pedley, K. F., and Martin,
G. B. (2004). MAPKKK[alpha] is a
positive regulator of cell death associated with both plant immunity and
disease. EMBO J. 23, 3072–3082.
Djamei, A., Pitzschke, A., Nakagami, H., Rajh, I., and Hirt,
H. (2007). trojan horse strategy in
Agrobacterium transformation: abusing MAPK defense signaling. Science
318, 453–456.
Dóczi, R., Brader, G., Pettkó-Szandtner,
A., Rajh, I., Djamei, A., Pitzschke, A.,
Teige, M., and Hirt, H. (2007). The
Arabidopsis mitogen-activated protein kinase kinase MKK3 is upstream
of group C mitogen-activated protein kinases and participates in
pathogen signaling. Plant Cell 19,
3266–3279.
Feilner, T., Hultschig, C., Lee, J.,
Meyer, S., Immink, R. G. H., Koenig,
A., Possling, A., Seitz, H., Beveridge, A., Scheel, D., Cahill, D. J.,
Lehrach, H., Kreutzberger, J., and
Kersten, B. (2005). High throughput
identification of potential Arabidopsis mitogen-activated protein kinases
substrates. Mol. Cell. Proteomics 4,
1558–1568.
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. Plant J. 18, 265–276.
Fiil, B. K., Petersen, K., Petersen, M.,
and Mundy, J. (2009). Gene regulation by MAP kinase cascades. Curr.
Opin. Plant Biol. 12, 615–621.
Forzani, C., Carreri, A., de la Fuente
van Bentem, S., Lecourieux, D.,
Lecourieux, F., and Hirt, H. (2011).
The Arabidopsis protein kinase Ptointeracting 1-4 is a common target of
the oxidative signal-inducible 1 and
mitogen-activated protein kinases.
FEBS J. 278, 1126–1136.
Frei dit Frey, N., Mbengue, M.,
Kwaaitaal, M., Nitsch, L., Altenbach,
D., Haweker, H., Lozano-Duran, R.,
Njo, M. F., Beeckman, T., Huettel,
B., Borst, J. W., Panstruga, R., and
Robatzek, S. (2012). Plasma membrane calcium ATPases are important
components of receptor-mediated
signaling in plant immune responses
and development. Plant Physiol. 159,
798–809.
MAP kinase cascades in immunity
Frye, C. A., Tang, D., and Innes, R.
W. (2001). Negative regulation of
defense responses in plants by a conserved MAPKK kinase. Proc. Natl.
Acad. Sci. U.S.A. 98, 373–378.
Gao, M., Liu, J., Bi, D., Zhang, Z.,
Cheng, F., Chen, S., and Zhang,
Y. (2008). MEKK1, MKK1/MKK2
and MPK4 function together in a
mitogen-activated protein kinase cascade to regulate innate immunity in
plants. Cell Res. 18, 1190–1198.
Gimenez-Ibanez, S., and Rathjen, J.
P. (2010). The case for the defense:
plants versus Pseudomonas syringae.
Microbes Infect. 12, 428–437.
Gómez-Gómez, L., and Boller, T.
(2000). FLS2: an LRR receptor-like
kinase involved in the perception of
the bacterial elicitor flagellin in Arabidopsis. Mol. Cell 5, 1003–1011.
Ichimura, K., Shinozaki, K., Tena, G.,
Sheen, J., Henry, Y., Champion, A.,
Kreis, M., Zhang, S., Hirt, H., Wilson, C., Heberle-Bors, E., Ellis, B. E.,
Morris, P. C., Innes, R. W., Ecker, J.
R., Scheel, D., Klessig, D. F., Machida,
Y., Mundy, J., Ohashi, Y., and Walker,
J. C. (2002). Mitogen-activated protein kinase cascades in plants: a new
nomenclature. Trends Plant Sci. 7,
301–308.
Komis, G., Illés, P., Beck, M., and
Šamaj, J. (2011). Microtubules and
mitogen-activated protein kinase signalling. Curr. Opin. Plant Biol. 14,
650–657.
Kwaaitaal, M., Huisman, R., Maintz,
J., Reinstädler, A., and Panstruga, R.
(2011). Ionotropic glutamate receptor (iGluR)-like channels mediate
MAMP-induced calcium influx in
Arabidopsis thaliana. Biochem. J. 440,
355–365.
Lee, K., Song, E. H., Kim, H. S., Yoo, J.
H., Han, H. J., Jung, M. S., Lee, S. M.,
Kim, K. E., Kim, M. C., Cho, M. J.,
and Chung, W. S. (2008). Regulation
of MAPK phosphatase 1 (AtMKP1)
by calmodulin in Arabidopsis. J. Biol.
Chem. 283, 23581–23588.
Ma, W., Qi, Z., Smigel, A., Walker, R.
K., Verma, R., and Berkowitz, G. A.
(2009). Ca2+ , cAMP, and transduction of non-self perception during
plant immune responses. Proc. Natl.
Acad. Sci. U.S.A. 106, 20995–21000.
Mao, G., Meng, X., Liu, Y., Zheng, Z.,
Chen, Z., and Zhang, S. (2011). Phosphorylation of a WRKY transcription
factor by two pathogen-responsive
MAPKs drives phytoalexin biosynthesis in Arabidopsis. Plant Cell 23,
1639–1653.
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. Proc. Natl. Acad. Sci.
U.S.A. 104, 19613–19618.
Parisy, V., Poinssot, B., Owsianowski,
L., Buchala, A., Glazebrook, J., and
Mauch, F. (2007). Identification of
PAD2 as a γ-glutamylcysteine synthetase highlights the importance of
glutathione in disease resistance of
Arabidopsis. Plant J. 49, 159–172.
Petersen, K., Qiu, J.-L., Lütje, J.,
Fiil, B. K., Hansen, S., Mundy, J.,
and Petersen, M. (2010). Arabidopsis
MKS1 is involved in basal immunity
and requires an intact N-terminal
domain for proper function. PLoS
ONE 5, e14364. doi: 10.1371/journal.pone.0014364
Petersen, M., Brodersen, P., Naested,
H., Andreasson, E., Lindhart, U.,
Johansen, B., Nielsen, H. B., Lacy,
M., Austin, M. J., Parker, J. E.,
Sharma, S. B., Klessig, D. F., Martienssen, R., Mattsson, O., Jensen, A.
B., and Mundy, J. (2000). Arabidopsis MAP kinase 4 negatively regulates
systemic acquired resistance. Cell 103,
1111–1120.
Popescu, S. C., Popescu, G. V., Bachan,
S., Zhang, Z., Gerstein, M., Snyder,
M., and Dinesh-Kumar, S. P. (2009).
MAPK target networks in Arabidopsis thaliana revealed using functional
protein microarrays. Genes Dev. 23,
80–92.
Qiu, J.-L., Fiil, B. K., Petersen, K.,
Nielsen, H. B., Botanga, C. J., Thorgrimsen, S., Palma, K., SuarezRodriguez, M. C., Sandbech-Clausen,
S., Lichota, J., Brodersen, P., Grasser,
K. D., Mattsson, O., Glazebrook, J.,
Mundy, J., and Petersen, M. (2008a).
Arabidopsis MAP kinase 4 regulates
gene expression through transcription factor release in the nucleus.
EMBO J. 27, 2214–2221.
Qiu, J.-L., Zhou, L., Yun, B.-W., Nielsen,
H. B., Fiil, B. K., Petersen, K.,
MacKinlay, J., Loake, G. J., Mundy, J.,
and Morris, P. C. (2008b). Arabidopsis mitogen-activated protein kinase
kinases MKK1 and MKK2 have overlapping functions in defense signaling
mediated by MEKK1, MPK4, and
MKS1. Plant Physiol. 148, 212–222.
Ren, D., Liu, Y., Yang, K.-Y., Han, L.,
Mao, G., Glazebrook, J., and Zhang,
S. (2008). A fungal-responsive MAPK
cascade regulates phytoalexin biosynthesis in Arabidopsis. Proc. Natl. Acad.
Sci. U.S.A. 105, 5638–5643.
Rentel, M. C., Lecourieux, D., Ouaked,
F., Usher, S. L., Petersen, L., Okamoto,
H., Knight, H., Peck, S. C., Grierson, C. S., Hirt, H., and Knight,
M. R. (2004). OXI1 kinase is necessary for oxidative burst-mediated
signalling in Arabidopsis. Nature
427, 858.
Rodriguez, M. C., Adams-Phillips, L.,
Liu, Y., Wang, H., Su, S. H., Jester, P.
J., Zhang, S., Bent, A. F., and Krysan,
P. J. (2007). MEKK1 is required
for flg22-induced MPK4 activation
in Arabidopsis plants. Plant Physiol.
143, 661.
Rodriguez, M. C., Petersen, M., and
Mundy, J. (2010). Mitogen-activated
protein kinase signaling in plants.
Annu. Rev. Plant Biol. 61, 621–649.
Roux, M., Schwessinger, B., Albrecht,
C., Chinchilla, D., Jones, A., Holton,
N., Malinovsky, F. G., Tör, 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. Plant Cell
23, 2440–2455.
Schwessinger, B., Roux, M., Kadota,
Y., Ntoukakis, V., Sklenar, J.,
Jones, A., and Zipfel, C. (2011).
Phosphorylation-dependent differential regulation of plant growth,
cell death, and innate immunity by
the regulatory receptor-like kinase
BAK1. PLoS Genet. 7, e1002046. doi:
10.1371/journal.pgen.1002046
Takahashi, F., Mizoguchi, T., Yoshida,
R., Ichimura, K., and Shinozaki, K.
(2011). Calmodulin-dependent activation of MAP kinase for ROS homeostasis in Arabidopsis. Mol. Cell 41,
649–660.
Ulm, R., Ichimura, K., Mizoguchi, T.,
Peck, S. C., Zhu, T., Wang, X., Shinozaki, K., and Paszkowski, J. (2002).
Distinct regulation of salinity and
genotoxic stress responses by Arabidopsis MAP kinase phosphatase 1.
EMBO J. 21, 6483–6493.
Van Der Biezen, E. A., and Jones, J. D. G.
(1998). Plant disease-resistance proteins and the gene-for-gene concept.
Trends Biochem. Sci. 23, 454–456.
Wang, H., Ngwenyama, N., Liu, Y.,
Walker, J. C., and Zhang, S. (2007).
Stomatal development and patterning are regulated by environmentally
responsive mitogen-activated protein
kinases in Arabidopsis. Plant Cell 19,
63–73.
Wurzinger, B., Mair, A., Pfister, B.,
and Teige, M. (2011). Cross-talk
of calcium-dependent protein kinase
and MAP kinase signaling. Plant Signal. Behav. 6, 8–12.
Xu, J., Li, Y., Wang, Y., Liu, H., Lei, L.,
Yang, H., Liu, G., and Ren, D. (2008).
Activation of MAPK kinase 9 induces
ethylene and camalexin biosynthesis
and enhances sensitivity to salt stress
in Arabidopsis. J. Biol. Chem. 283,
26996–27006.
25
“fpls-03-00169” — 2012/7/22 — 17:06 — page 5 — #5
Rasmussen et al.
Yoo, S.-D., Cho, Y.-H., Tena, G., Xiong,
Y., and Sheen, J. (2008). Dual control
of nuclear EIN3 by bifurcate MAPK
cascades in C2 H4 signalling. Nature
451, 789–795.
Zhang, J., Shao, F., Li, Y., Cui, H., Chen,
L., Li, H., Zou, Y., Long, C., Lan,
L., Chai, J., Chen, S., Tang, X., and
Zhou, J. M. (2007a). A Pseudomonas
syringae effector inactivates MAPKs
to suppress PAMP-induced immunity in plants. Cell Host Microbe 1,
175–185.
Zhang, X., Dai, Y., Xiong, Y., DeFraia, C.,
Li, J., Dong, X., and Mou, Z. (2007b).
Overexpression of Arabidopsis MAP
kinase kinase 7 leads to activation
of plant basal and systemic acquired
resistance. Plant J. 52, 1066–1079.
Zhang, Z., Wu, Y., Gao, M., Zhang, J.,
Kong, Q., Liu, Y., Ba, H., Zhou, J.,
MAP kinase cascades in immunity
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.
Zheng, Z., Qamar, S. A., Chen,
Z., and Mengiste, T. (2006). Arabidopsis WRKY33 transcription factor is required for resistance to
necrotrophic fungal pathogens. Plant
J. 48, 592–605.
Zhou, N., Tootle, T. L., and Glazebrook, J. (1999). Arabidopsis PAD3, a
gene required for camalexin biosynthesis, encodes a putative cytochrome
P450 monooxygenase. Plant Cell 11,
2419–2428.
Zhu, X., Caplan, J., Mamillapalli,
P., Czymmek, K., and DineshKumar, S. P. (2010). Function
of endoplasmic reticulum calcium
ATPase in innate immunity-mediated
programmed cell death. EMBO J. 29,
1007–1018.
Zipfel, C. (2008). Pattern-recognition
receptors in plant innate immunity.
Curr. Opin. Immunol. 20, 10–16.
Zipfel, C., Kunze, G., Chinchilla, D.,
Caniard, A., Jones, J. D. G., Boller,
T., and Felix, G. (2006). Perception
of the bacterial PAMP EF-Tu by the
receptor EFR restricts Agrobacteriummediated transformation. Cell 125,
749–760.
Conflict of Interest Statement: The
authors declare that the research was
conducted in the absence of any commercial or financial relationships that
could be construed as a potential conflict of interest.
Received: 02 May 2012; paper pending
published: 24 May 2012; accepted: 09 July
2012; published online: 24 July 2012.
Citation: Rasmussen MW, Roux M,
Petersen M and Mundy J (2012) MAP
kinase cascades in Arabidopsis innate
immunity. Front. Plant Sci. 3:169. doi:
10.3389/fpls.2012.00169
This article was submitted to Frontiers in
Plant Proteomics, a specialty of Frontiers
in Plant Science.
Copyright © 2012 Rasmussen, Roux,
Petersen and Mundy. This is an openaccess article distributed under the terms
of the Creative Commons Attribution
License, which permits use, distribution
and reproduction in other forums, provided the original authors and source
are credited and subject to any copyright notices concerning any third-party
graphics etc.
26
“fpls-03-00169” — 2012/7/22 — 17:06 — page 6 — #6
mRNA decay in plant immunity
Magnus W. Rasmussen, Àngels M. Regué, John Mundy and Morten Petersen*
Review article submitted to Cellular and Molecular Life science
*Corresponding author: [email protected]
Abstract
Strict regulation of mRNA metabolism is required in plant immunity in order to control defense
responses and regulation of immune receptors. Transcription is specific and is typically regulated
through activation and release of transcription factors, equally important is transcript degradation.
The mechanisms of mRNA decay are well studied, but the mechanisms regarding specificity is are
not well understood. Mutants defective in both nonsense mediated decay and decapping display
autoimmune phenotypes implying functions in immune regulation. In this review, we will focus on
the few published examples concerning mRNA decay and immune responses in Arabidopsis.
Introduction
mRNA metabolism is strictly regulated to control the fate of cells in response to developmental and
environmental stimuli. mRNA metabolism refers to any event in the lifecycle of mRNAs including
transcription, processing and degradation. Much is known about changes in transcript levels during
plant development (Becker and Theißen, 2003) and in response to abiotic stress (Akhtar et al., 2012;
Rasmussen et al., 2013) or biotic stresses including microbial perception (Andreasson et al., 2005; Mao
et al., 2011; Qiu et al., 2008). Changes in specific transcript levels are often attributed to changes in
their transcription rates, although there are relatively few reports of global or specific transcription
rates (Sidaway-Lee et al., 2014). While transcription of genes responding to specific stimuli clearly
affects the accumulation of mRNAs (Akhtar et al., 2012), the contribution of mRNA decay is less
studied. A primary example of the importance of mRNA decay is illustrated by Xu and Chua (Xu
and Chua, 2012, 1) who showed that phosphorylation of Decapping 1 (DCP1) by MAP kinase 6
(MPK6) upon dehydration stress promotes decapping, and that plants expressing a nonphophorylatable version of DCP1 were hypersensitive to osmotic stress.
Eukaryotic mRNAs contain two integral stability determinants, the 5’ 7-methylguanosine
triphosphate cap (m7G) and the 3’ poly(A) tail, both of which are incorporated co-transcriptionally.
These structures interact with cellular proteins: eukaryotic initiation factor 4E (eIF4E) binds directly
27
to the 5’ cap and the poly(A) binding protein (PABP) binds to the 3’ poly(A) tail. The binding of
eIF4E and PABP promotes translational initiation and, at the same time, prevents degradation by
exoribonucleases. Thus, to initiate mRNA decay these determinants must either be removed or the
mRNA must be cleaved by an endonuclease followed by an exoribonuclease attack on the newly
exposed 3’ and 5’ ends. Eukaryotic mRNAs are primarily degraded by removal of the poly(A) tail
with subsequent removal of the 5’ cap and 5’ to 3’ degradation by exoribonucleases, or via 3’ to 5’
degradation by the exosome complex. The components of the Arabidopsis exosome complex are
reviewed by Lange and Gagliardi (Lange and Gagliardi, 2010), and Chiba and Green (Chiba and Green,
2009) provide a comprehensive comparison of yeast and plant mRNA decay machinery. Transcripts
can also be degraded via nonsense mediated decay (NMD), a eukaryotic RNA surveillance
mechanism that controls transcripts with aberrant features and targets them for degradation. These
aberrant features include premature termination codons (PTC), PTC-independent long 3’
untranslated regions, 3’ UTR-positioned introns, and upstream open reading frames (Wachter and
Hartmann, 2014). In addition to controlling mRNA integrity, NMD has also been described in
mammalian cells to target different mRNA isoforms, providing an additional regulation point for
important biological processes (Smith and Baker, 2015). This review mainly focuses on mRNA
degradation in the context of immune signaling in plants. While most published work in this field
concerns microRNA (miRNA) mediated decay of transcripts, we primarily focus here on the few
examples of non-miRNA immune mediated mRNA.
Plant innate immunity
Plants have two layers of immunity (Jones and Dangl, 2006). The first layer is comprised of pattern
recognition receptors (PRRs) in the plasma membrane, which recognize pathogen-associated
molecular patterns (PAMPs). This recognition induces PAMP triggered immunity (PTI) (Greeff et
al., 2012). PAMPs and their cognate PPRs include bacterial flagellin (and its peptide derivative
flg22) recognized by the receptor FLS2 (Felix et al., 1999; Gómez-Gómez and Boller, 2000), bacterial
elongation factor EF-Tu by EFR1 (Kunze et al., 2004; Zipfel et al., 2006) and fungal chitin recognized
by CERK1 (Miya et al., 2007). Immune responses triggered by PAMP recognition include protein
phosphorylation, induction of defense genes, extracellular alkalization and production of reactive
oxygen species (Rasmussen et al., 2012). The second layer of defense employs cytoplasmic immune
receptors known as resistance (R) proteins that can detect pathogenic virulence determinants,
commonly referred to as effectors, which are injected into the host cell. R proteins recognize
28
effectors by direct interaction or via effector modification of host proteins and activate effector
triggered immunity (ETI) (Jones and Dangl, 2006). PTI and ETI signaling is somewhat similar
although ETI responses are prolonged and more pronounced and can include localized cell death
(Cui et al., 2014). Most R-proteins are nucleotide-binding domain, leucine-rich repeat (NB-NLR)
proteins that can be further separated into coiled-coil- (CC-) or Toll/interleukin-1 receptor- (TIR-)
NB-LRRs (Caplan et al., 2008). The mechanism(s) by which R-proteins confer resistance against
specific pathogens remain largely unknown, and their signaling mechanisms are probably diverse
(Bonardi and Dangl, 2012). Nonetheless, different genes are required for NB-LRR mediated resistance
as ENHANCED DISEASE SUSCEPTIBILITY1 (EDS1) and PHYTOALEXINDEFICIENT4 (PAD4)
are required for TIR-NB-LRR defenses whereas NO DISEASE RESPONSES1 (NDR1) is required
for most defense responses mediated by CC-NB-LRRs (Century et al., 1997; Parker et al., 1996). R
proteins are tightly regulated as their overexpression may lead to increased immune signaling and
autoimmune phenotypes in non-infected plants (Alcázar and Parker, 2011; Nandety et al., 2013).
Conversely, plants with loss-of-function R genes are more susceptible to specific pathogen attacks.
mRNA decay and autoimmunity
Some mutants implicated in mRNA decay exhibit autoimmune phenotypes. These include protein
associated with topoisomerase protein 1 (pat1) involved in decapping of transcripts, as well as
upframeshift1 (upf1), upf3 and suppressor with morphogenic effect on genitalia7 (smg7)
functioning in non-sense mediated decay (NMD) (Fig. 1A) (Gloggnitzer et al., 2014; Jeong et al., 2011;
Rayson et al., 2012a; Riehs-Kearnan et al., 2012; Roux et al., 2015). The autoimmune phenotypes of these
NMD deficient mutants resemble those of lesion mimic mutants such as mpk4 (Petersen et al., 2000;
Zhang et al., 2012, 2) and accelerated cell death11 (acd11) (Brodersen et al., 2002; Palma et al., 2010,
11) in which autoimmunity is caused by inappropriate activation of ETI. Similar to R protein
mediated ETI, autoimmunity in NMD deficient mutants is also repressed at elevated temperatures
or high humidity (Alcázar and Parker, 2011; Riehs-Kearnan et al., 2012; Zhu et al., 2010). In addition,
inactivation of PAD4 but not NDR1 fully suppresses the retarded growth observed in the smg7 and
upf1 NMD deficient mutants. This indicates activation of ETI by one or more TIR-NB-LRRs in
these mutants (Riehs-Kearnan et al., 2012).
29
Figure 1: Nonsense-mediated mRNA decay
A, The ribosome stalls at a premature termination codon (PTC), giving rise to a long 3’UTR sensed by UPF1
(Hogg and
Goff, 2010). UPF1 then associates with translation release factor eRF3 and subsequently recruits UPF2 and UPF3. A
phosphatidylinositol 3-kinase-related kinase phosphorylates UPF1 which enables translocation of the UPF1/2/3-mRNA
complex to p-bodies by binding SMG7 in a phospho-dependent manner (Kerényi et al., 2013).
B, NMD mutant plants display autoimmune phenotypes reassembling effector triggered immunity and resistance (Jeong et
al., 2011; Rayson et al., 2012b; Riehs-Kearnan et al., 2012). This may arise either from (i) activation of R-proteins
guarding NMD components such as UPF1-3 or SMG7 (ii) hyper-accumulation of R-gene transcripts normally downregulated by NMD (Gloggnitzer et al., 2014).
NMD is apparently regulated during immune responses because plants challenged with the virulent
bacteria Pseudomonas syringae pv. tomato (pto) DC3000 have reduced levels of UPF1 and UPF3
transcripts and accumulate canonical NMD target transcripts (Jeong et al., 2011). RNA-seq data have
shown that 39 TIR-NB-LRRs and 11 CC-NB-LRRs are upregulated in the smg7 pad4 double
mutant compared to wild type (Gloggnitzer et al., 2014). Interestingly, 19 of these TIR-NB-LRRs
30
contain putative NMD target features while these features are absent in the 11 upregulated CC-NBLRRs. The half-lives of all but one of the upregulated TIR-NB-LRR transcripts are increased in the
smg7 pad4 double mutant, whereas the half-lives of the upregulated CC-NB-LRR transcripts are
unaltered. This indicates that only TIR-NB-LRR transcripts are affected by mutations in SMG7
(Gloggnitzer et al., 2014). Although NMD seems to regulate TIR-NB-LRR transcripts post-
transcriptionally and hence regulate ETI, NMD is down regulated after PAMP treatment with flg22.
This supports a role for NMD in regulating certain PTI responses (Gloggnitzer et al., 2014). This
raises the question of whether the phenotype of NMD deficient mutants is due to an intrinsic
mechanism stabilizing NMD targeted defense transcripts such as those of TIR-NB-LRRs.
Alternatively, since a natural variant of the TIR-NB-LRR RPS6 can suppress the autoimmune smg7
phenotype (Gloggnitzer et al., 2014) this NMD deficient phenotype could also be due to activation of
RPS6 guarding putative effector targeted steps in the NMD pathway mimicked by mutations in
SMG7. Two models explaining autoimmunity in smg7 is illustrated in Figure 2B.
Thus, autoimmunity in mRNA decay mutants may arise from inappropriate activation of R proteins
caused by mutations in favored effector targets. An example of this comes from analysis of the
decapping mutant pat1 (Roux et al., 2015). Decapping or removal of the mRNA 5’ cap structure is
important for exoribonuclease mediated transcript degradation. Mutants impaired in decapping such
as dcp1, dcp2 and varicose (vcs) are all seedling lethal and accumulate high levels of capped
transcripts (Goeres et al., 2007; Xu et al., 2006, 1). Recent studies have shown that transcript
accumulation due to improper decapping or lack of sequential decay by exoribonucleases or the
exosome results in induced endogenous gene silencing through increased siRNA levels
(Martínez de Alba et al., 2015; Zhang et al., 2015). In addition to gene silencing, miRNA-guided
translational repression also seems to be elevated in plant decapping mutants, as also reported for
mammalian systems (Brodersen et al., 2008; Eulalio et al., 2007). mRNA decapping occurs in discrete
cytoplasmic foci known as processing bodies and, upon activation of PTI with flg22, these foci
increase in number (Balagopal and Parker, 2009; Parker and Sheth, 2007; Roux et al., 2015). PAT1
localizes to processing bodies where it functions as a decapping activator (Marnef and Standart,
2010). Arabidopsis PAT1 interacts and is phosphorylated by MPK4, and PAT1 phosphorylation
increases after flg22 treatment (Roux et al., 2015). MPK4 functions in transducing signals from PPRs
(Rodriguez et al., 2010) and plants lacking MPK4 activity display activation of ETI dependent on the
R protein SUMM2 which guards the MPK4 pathway (Petersen et al., 2000, 4; Zhang et al., 2012, 4).
Mutants lacking PAT1 display constitutive defense responses and, similarly to mpk4 mutants, pat1
is immunosuppressed when crossed into a summ2 background. In addition, PAT1 interacts or is in
31
complex with SUMM2. It is therefore possible that PAT1 is an effector target with a specific
function in immunity that is under surveillance by SUMM2 (Roux et al., 2015). Arabidopsis MPK6,
another stress related MAP kinase, phosphorylates DCP1 upon dehydration stress and thereby
represses decapping (Xu and Chua, 2012, 1). Taken together, these observations indicate that both
biotic and abiotic stress signaling can directly influence mRNA decapping.
An example of how mRNA decay may control important steps in a defense pathway includes the
plant hormone ethylene that regulates both developmental processes and responses to biotic and
abiotic stresses (Johnson and Ecker, 1998). Ethylene is perceived in Arabidopsis by a family of 4
receptors that function as negative regulators (Chang abnd Stadler, 2001). In the absence of ethylene,
these receptors activate Constitutively Triple Response (CTR1), a MAPK kinase kinase (MAP3K)
which also functions as a negative regulator of ethylene signaling (Kieber et al., 1993). Ethylene
Insensitive 2 (EIN2) is genetically downstream of CTR1 and is a positive regulator promoting
ethylene responses through the transcription factors EIN3 and EIN3-like proteins (Alonso et al., 1999,
2; Chao et al., 1997; Guo and Ecker, 2003). In the presence of ethylene EIN3 levels increase rapidly,
while in the absence of ethylene EIN3 is down-regulated post-translationally through a
ubiquitin/proteasome pathway mediated by the EIN3 binding F-box proteins 1 and 2 (EBF1&2
(Gagne et al., 2004; Guo and Ecker, 2003, 3; Potuschak et al., 2003; Yanagisawa et al., 2003). Both EBF1
and EBF2 transcripts have been found to be targets of the Exoribonuclease 4 (XRN4) and hence
accumulate in xrn4 mutants which results in reduced levels of EIN3 protein (Olmedo et al., 2006;
Potuschak et al., 2006, 4). Tight regulation of EIN3, likely through EBF1 and EBF2, is essential for
fine-tuning immunity as EIN3 positively regulates transcription of the PRR FLS2 by binding to its
promoter (Boutrot et al., 2010). EIN3 also regulates transcription of Isochorismate Synthase 1 (ICS1
also known as SID2), involved in the biosynthesis of the defense hormone salicylic acid. In contrast
to its effects on FLS2, EIN3 represses transcription of ICS1 and thereby inhibits salicylic acid
accumulation (Chen et al., 2009). How and why EIN3 reciprocally regulates two genes important for
mounting proper immune responses remains a bit of a mystery. Nonetheless, it seems clear that
EIN3 itself is regulated via EBF1 and EBF2 mRNA decay (Olmedo et al., 2006; Potuschak et al.,
2006). Expression of Ethylene Response Factor 1 (ERF1) is directly induced by EIN3 (Solano et al.,
1998). ERF1 links ethylene and jasmonate signaling in Arabidopsis for appropriate expression of
defense responses against pathogens (Lorenzo et al., 2003). As for EBF1 and EBF2, ERF1 transcripts
are also targeted by XRN4 and hence accumulate in xrn4 mutants (Nguyen et al., 2015). XRN4 is
homologous to yeast XRN1 which degrades RNA species with a single phosphate group at the 5’
end (Kastenmayer and Green, 2000; Souret et al., 2004). It is therefore likely that the ERF1 transcripts,
32
as well as those of EBF1 and EBF2, that accumulate in xrn4 mutants are uncapped and that
decapping in fact regulates these transcripts prior to XRN4 mediated decay. This has not yet been
tested experimentally.
In addition to the above examples, miRNAs provide an additional control over mRNA decay and a
direct link to regulation of immunity. Post transcriptional gene silencing (PTGS) is a key regulatory
mechanism in controlling immune responses in plants (Seo et al., 2013). In PTGS, double stranded
RNAs are cleaved into 20-24 nucleotide small RNA species by DICER-like enzymes. One of the
two RNA strands then associates with ARGONAUTE (AGO) proteins which are catalytic
components of the RNA-induced silencing complex (RISC) promoting mRNA cleavage (Meister,
2013). Different classes of small RNA species including micro RNAs (miRNAs) and short
interfering RNAs (siRNAs) are involved in PTGS. miRNAs are transcribed from the genome by
RNA polymerase II and, due to regions of reverse sequence complementarity, primary or primiRNAs form hairpin-like secondary structures (Ha and Kim, 2014). In plants, mature miRNAs are
formed in the nucleus by cleavage of pri-miRNAs by DICER-LIKE1 (DCL1) proteins in complex
with SERRATE and HYL1. Mature miRNAs are then exported to the cytosol where they associate
with AGO1 and promote mRNA cleavage (Ha and Kim, 2014).
miRNAs has been shown to function in plant immunity (Li et al., 2010; Navarro et al., 2006). In
response to PAMP perception, miRNA miR393 accumulates and negatively regulates auxin
signaling by targeting transcripts encoding the auxin receptors TIR1 (transport inhibitor 1), AFB2
(auxin signaling F-box protein 2) and AFB3 (Navarro et al., 2006). Arabidopsis encodes a number
auxin response factors (ARFs) that can promote or repress the transcription of genes containing
auxin responsive elements (Hagen and Guilfoyle, 2002). The auxin-regulated Aux/IAA proteins
sequester ARFs and thereby repress auxin signaling (Liscum and Reed, 2002). TIR1 is part of a
ubiquitin-ligase complex interacting with Aux/IAA proteins and marking them for degradation
(Gray et al., 2001). Therefore, as a consequence of PAMP triggered miR393 accumulation, Aux/IAA
proteins are stabilized resulting in repressed auxin signaling (Navarro et al., 2006). Overexpression of
transgenic miR393 increases resistance to pto DC3000, while overexpression of AFB1 (a paralog to
TIR1 and AFB2/3 that is resistant to miR393 mediated degradation) renders plants more susceptible
(Navarro et al., 2006). This supports a function of miRNA in regulating proper responses during PTI.
Virulent bacterial pathogens have evolved effector proteins that can suppress the miRNA pathway
and thereby repress host immunity. During infection, Pto DC3000 can employ the effector protein
AvrPtoB that represses transcription of miR393a and miR393b which both encode miR393 (Navarro
et al., 2008). In addition, the effector protein AvrPto stabilizes miR393 precursors thereby reducing
33
the levels of mature miR393 and rendering plants more susceptible to infection (Navarro et al., 2008).
Since AvrPto and AvrPtoB both interact with and inhibit PRRs including FLS2 (Dodds and Rathjen,
2010; Göhre et al., 2008; Xiang et al., 2008) it is possible that these effectors do not only target the
miRNA pathway, but also suppress PTI signaling downstream of FLS2 and other PRRs. Other
miRNAs including miR160 and miR167 are also upregulated during PTI, and both target members
of the ARF transcription factor family. This further supports the involvement of miRNA mediated
gene silencing in auxin signaling during PTI (Fahlgren et al., 2007).
Transcripts encoding R protein immune receptors have also been shown to be direct targets of
miRNAs. miR472 negatively regulates several CC-NB-LRRs by targeting conserved regions
encoding ATP binding p-loop domains (Boccara et al., 2014; Lu et al., 2006). It is likely that in the
absence of infection miR472 promotes degradation of CC-NB-LRR transcripts and that activation
of PTI down regulates or pacifies miR472 to permit accumulation of CC-NB-LRR transcripts
resulting in increased resistance (Boccara et al., 2014). Consistent with this, miR472 deficient and
over-expressor plants are respectively more resistant and susceptible to pto DC3000 (Boccara et al.,
2014). Similarly, tomato miR482 targets a group of CC-NB-LRRs and results in the production of
secondary trans-acting siRNAs that target other immune regulators. Such secondary targets include
PEN3 involved in basal immunity and a proteasomal subunit that is also likely to function in
resistance (Shivaprasad et al., 2012). In addition, and as seen for Arabidopsis miR472 and miR393,
tomato miR482 mediated silencing is impaired in infected plants, likely due to a bacterial effector.
Tomato plants infiltrated with an hrcC mutant of Pto. DC3000 which is incapable of delivering
effectors into host plants showed impaired accumulation of miR482 target transcripts compared to
plants infiltrated with virulent Pto. DC3000 (Shivaprasad et al., 2012). This effector mediated upregulation of resistance gene expression via the suppression of host RNA silencing might have
evolved as a counter defense against pathogens similar to ETI.
Plant miRNAs represent a fraction of the total pool of small RNAs which includes siRNAs and
secondary siRNAs (Meyers et al., 2008). In contrast to miRNAs, siRNAs usually perform auto
silencing, while trans-acting siRNAs target mRNAs transcribed from other loci. This amplifies the
effects of siRNA and miRNA mediated PTGS (Ghildiyal and Zamore, 2009). For example, the
Arabidopsis RPP5 locus encodes seven TIR-NB-LRR proteins including SNC1 and RPP4. Most of
these R gene paralogs are all repressed by 21-nucleotide siRNAs corresponding to their LRR
regions which are likely generated from antisense transcription from SNC1 (Yi and Richards, 2007).
34
Conclusion
Our current knowledge of immune regulated mRNA decay and its specificity remains limited.
Control of R proteins through miRNA-mediated mRNA decay implies direct roles in regulating
appropriate immune responses. The examples described above also suggest that the general decay
pathways comprising the exosome, exoribonucleases and NMD are important for immune signaling
as well as for responses and adaptations to abiotic stress. It is difficult to identify the target
transcript specificities of these pathways, in contrast to the sequence homology based identification
of miRNA-mediated decay targets. Nonetheless, since mutations in several of the key components
in mRNA decay pathways have autoimmune phenotypes, they are themselves likely targets of
pathogen effectors. This issue needs further investigation to ensure that conclusions regarding the
functions of these decay components are not based on potential pleiotropic phenotypes of their lossof-function mutants, but rather reflect their true in vivo functions.
References
Akhtar, M., Jaiswal, A., Taj, G., Jaiswal, J. P., Qureshi, M. I., and Singh, N. K. (2012). DREB1/CBF
transcription factors: their structure, function and role in abiotic stress tolerance in plants. J. Genet.
91, 385–395.
Alcázar, R., and Parker, J. E. (2011). The impact of temperature on balancing immune responsiveness and
growth in Arabidopsis. Trends Plant Sci. 16, 666–675. doi:10.1016/j.tplants.2011.09.001.
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.
doi:10.1126/science.284.5423.2148.
Andreasson, E., Jenkins, T., Brodersen, P., Thorgrimsen, S., Petersen, N. H., Zhu, S., et al. (2005). The MAP
kinase substrate MKS1 is a regulator of plant defense responses. EMBO J 24, 2579–2589.
doi:10.1038/sj.emboj.7600737.
Balagopal, V., and Parker, R. (2009). Polysomes, P bodies and stress granules: states and fates of eukaryotic
mRNAs. Curr. Opin. Cell Biol. 21, 403–408. doi:10.1016/j.ceb.2009.03.005.
Becker, A., and Theißen, G. (2003). The major clades of MADS-box genes and their role in the development
and evolution of flowering plants. Mol. Phylogenet. Evol. 29, 464–489. doi:10.1016/S10557903(03)00207-0.
Boccara, M., Sarazin, A., Thiébeauld, O., Jay, F., Voinnet, O., Navarro, L., et al. (2014). The Arabidopsis
miR472-RDR6 Silencing Pathway Modulates PAMP- and Effector-Triggered Immunity through the
Post-transcriptional Control of Disease Resistance Genes. PLoS Pathog 10, e1003883.
doi:10.1371/journal.ppat.1003883.
Bonardi, V., and Dangl, J. L. (2012). How complex are intracellular immune receptor signaling complexes?
Plant Proteomics 3, 237. doi:10.3389/fpls.2012.00237.
35
Boutrot, F., Segonzac, C., Chang, K. N., Qiao, H., Ecker, J. R., Zipfel, C., et al. (2010). Direct transcriptional
control of the Arabidopsis immune receptor FLS2 by the ethylene-dependent transcription factors
EIN3 and EIL1. Proc. Natl. Acad. Sci. 107, 14502–14507. doi:10.1073/pnas.1003347107.
Brodersen, P., Petersen, M., Pike, H. M., Olszak, B., Skov, S., Ødum, N., et al. (2002). Knockout of
Arabidopsis ACCELERATED-CELL-DEATH11 encoding a sphingosine transfer protein causes
activation of programmed cell death and defense. Genes Dev. 16, 490–502. doi:10.1101/gad.218202.
Brodersen, P., Sakvarelidze-Achard, L., Bruun-Rasmussen, M., Dunoyer, P., Yamamoto, Y. Y., Sieburth, L.,
et al. (2008). Widespread Translational Inhibition by Plant miRNAs and siRNAs. Science 320,
1185–1190. doi:10.1126/science.1159151.
Caplan, J., Padmanabhan, M., and Dinesh-Kumar, S. P. (2008). Plant NB-LRR immune receptors: from
recognition to transcriptional reprogramming. Cell Host Microbe 3, 126–135.
doi:10.1016/j.chom.2008.02.010.
Century, K. S., Shapiro, A. D., Repetti, P. P., Dahlbeck, D., Holub, E., and Staskawicz, B. J. (1997). NDR1,
a Pathogen-Induced Component Required for Arabidopsis Disease Resistance. Science 278, 1963–
1965. doi:10.1126/science.278.5345.1963.
Chang, C., and Stadler, R. (2001). Ethylene hormone receptor action in Arabidopsis. BioEssays 23, 619–627.
doi:10.1002/bies.1087.
Chao, Q., Rothenberg, M., Solano, R., Roman, G., Terzaghi, W., and Ecker†, J. R. (1997). Activation of the
Ethylene Gas Response Pathway in Arabidopsis by the Nuclear Protein ETHYLENEINSENSITIVE3 and Related Proteins. Cell 89, 1133–1144. doi:10.1016/S0092-8674(00)80300-1.
Chen, H., Xue, L., Chintamanani, S., Germain, H., Lin, H., Cui, H., et al. (2009). ETHYLENE
INSENSITIVE3 and ETHYLENE INSENSITIVE3-LIKE1 Repress SALICYLIC ACID
INDUCTION DEFICIENT2 Expression to Negatively Regulate Plant Innate Immunity in
Arabidopsis. Plant Cell 21, 2527–2540. doi:10.1105/tpc.108.065193.
Chiba, Y., and Green, P. J. (2009). mRNA Degradation Machinery in Plants. J. Plant Biol. 52, 114–124.
doi:10.1007/s12374-009-9021-2.
Cui, H., Tsuda, K., and Parker, J. E. (2014). Effector-Triggered Immunity: From Pathogen Perception to
Robust Defense. Annu. Rev. Plant Biol. doi:10.1146/annurev-arplant-050213-040012.
Dodds, P. N., and Rathjen, J. P. (2010). Plant immunity: towards an integrated view of plant–pathogen
interactions. Nat. Rev. Genet. 11, 539–548. doi:10.1038/nrg2812.
Eulalio, A., Rehwinkel, J., Stricker, M., Huntzinger, E., Yang, S.-F., Doerks, T., et al. (2007). Target-specific
requirements for enhancers of decapping in miRNA-mediated gene silencing. Genes Dev. 21, 2558–
2570. doi:10.1101/gad.443107.
Fahlgren, N., Howell, M. D., Kasschau, K. D., Chapman, E. J., Sullivan, C. M., Cumbie, J. S., et al. (2007).
High-Throughput Sequencing of Arabidopsis microRNAs: Evidence for Frequent Birth and Death of
MIRNA Genes. PLoS ONE 2, e219. doi:10.1371/journal.pone.0000219.
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. Plant J. 18, 265–276. doi:10.1046/j.1365313X.1999.00265.x.
Gagne, J. M., Smalle, J., Gingerich, D. J., Walker, J. M., Yoo, S.-D., Yanagisawa, S., et al. (2004).
Arabidopsis EIN3-binding F-box 1 and 2 form ubiquitin-protein ligases that repress ethylene action
36
and promote growth by directing EIN3 degradation. Proc. Natl. Acad. Sci. U. S. A. 101, 6803–6808.
doi:10.1073/pnas.0401698101.
Ghildiyal, M., and Zamore, P. D. (2009). Small silencing RNAs: an expanding universe. Nat. Rev. Genet. 10,
94–108. doi:10.1038/nrg2504.
Gloggnitzer, J., Akimcheva, S., Srinivasan, A., Kusenda, B., Riehs, N., Stampfl, H., et al. (2014). NonsenseMediated mRNA Decay Modulates Immune Receptor Levels to Regulate Plant Antibacterial
Defense. Cell Host Microbe 16, 376–390. doi:10.1016/j.chom.2014.08.010.
Goeres, D. C., Norman, J. M. V., Zhang, W., Fauver, N. A., Spencer, M. L., and Sieburth, L. E. (2007).
Components of the Arabidopsis mRNA Decapping Complex Are Required for Early Seedling
Development. Plant Cell 19, 1549–1564. doi:10.1105/tpc.106.047621.
Göhre, V., Spallek, T., Häweker, H., Mersmann, S., Mentzel, T., Boller, T., et al. (2008). Plant PatternRecognition Receptor FLS2 Is Directed for Degradation by the Bacterial Ubiquitin Ligase AvrPtoB.
Curr. Biol. 18, 1824–1832. doi:10.1016/j.cub.2008.10.063.
Gómez-Gómez, L., and Boller, T. (2000). FLS2: An LRR Receptor-like Kinase Involved in the Perception of
the Bacterial Elicitor Flagellin in Arabidopsis. Mol. Cell 5, 1003–1011. doi:10.1016/S10972765(00)80265-8.
Gray, W. M., Kepinski, S., Rouse, D., Leyser, O., and Estelle, M. (2001). Auxin regulates SCFTIR1dependent degradation of AUX/IAA proteins. Nature 414, 271–276. doi:10.1038/35104500.
Greeff, C., Roux, M., Mundy, J., and Petersen, M. (2012). Receptor-like kinase complexes in plant innate
immunity. Front. Plant Sci. 3. doi:10.3389/fpls.2012.00209.
Guo, H., and Ecker, J. R. (2003). Plant Responses to Ethylene Gas Are Mediated by SCFEBF1/EBF2Dependent Proteolysis of EIN3 Transcription Factor. Cell 115, 667–677. doi:10.1016/S00928674(03)00969-3.
Hagen, G., and Guilfoyle, T. (2002). Auxin-responsive gene expression: genes, promoters and regulatory
factors. Plant Mol. Biol. 49, 373–385. doi:10.1023/A:1015207114117.
Ha, M., and Kim, V. N. (2014). Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 15, 509–524.
doi:10.1038/nrm3838.
Hogg, J. R., and Goff, S. P. (2010). Upf1 senses 3′UTR length to potentiate mRNA decay. Cell 143, 379–
389. doi:10.1016/j.cell.2010.10.005.
Jeong, H.-J., Kim, Y. J., Kim, S. H., Kim, Y.-H., Lee, I.-J., Kim, Y. K., et al. (2011). Nonsense-Mediated
mRNA Decay Factors, UPF1 and UPF3, Contribute to Plant Defense. Plant Cell Physiol. 52, 2147–
2156. doi:10.1093/pcp/pcr144.
Johnson, P. R., and Ecker, J. R. (1998). THE ETHYLENE GAS SIGNAL TRANSDUCTION PATHWAY:
A Molecular Perspective. Annu. Rev. Genet. 32, 227–254. doi:10.1146/annurev.genet.32.1.227.
Jones, J. D. G., and Dangl, J. L. (2006). The plant immune system. Nature 444, 323–329.
doi:10.1038/nature05286.
Kastenmayer, J. P., and Green, P. J. (2000). Novel features of the XRN-family in Arabidopsis: Evidence that
AtXRN4, one of several orthologs of nuclear Xrn2p/Rat1p, functions in the cytoplasm. Proc. Natl.
Acad. Sci. U. S. A. 97, 13985–13990.
37
Kerényi, F., Wawer, I., Sikorski, P. J., Kufel, J., and Silhavy, D. (2013). Phosphorylation of the N- and Cterminal UPF1 domains plays a critical role in plant nonsense-mediated mRNA decay. Plant J. 76,
836–848. doi:10.1111/tpj.12346.
Kieber, J. J., Rothenberg, M., Roman, G., Feldmann, K. A., and Ecker, J. R. (1993). CTR1, a negative
regulator of the ethylene response pathway in arabidopsis, encodes a member of the Raf family of
protein kinases. Cell 72, 427–441. doi:10.1016/0092-8674(93)90119-B.
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. Plant Cell 16, 3496–
3507. doi:10.1105/tpc.104.026765.
Lange, H., and Gagliardi, D. (2010). “The Exosome and 3′–5′ RNA Degradation in Plants,” in RNA Exosome
Advances in Experimental Medicine and Biology., ed. T. H. Jensen (Springer US), 50–62. Available
at: http://link.springer.com/chapter/10.1007/978-1-4419-7841-7_5 [Accessed April 9, 2015].
Liscum, E., and Reed, J. W. (2002). Genetics of Aux/IAA and ARF action in plant growth and development.
Plant Mol. Biol. 49, 387–400. doi:10.1023/A:1015255030047.
Li, Y., Zhang, Q., Zhang, J., Wu, L., Qi, Y., and Zhou, J.-M. (2010). Identification of MicroRNAs Involved
in Pathogen-Associated Molecular Pattern-Triggered Plant Innate Immunity. Plant Physiol. 152,
2222–2231. doi:10.1104/pp.109.151803.
Lorenzo, O., Piqueras, R., Sánchez-Serrano, J. J., and Solano, R. (2003). ETHYLENE RESPONSE
FACTOR1 Integrates Signals from Ethylene and Jasmonate Pathways in Plant Defense. Plant Cell
15, 165–178. doi:10.1105/tpc.007468.
Lu, C., Kulkarni, K., Souret, F. F., MuthuValliappan, R., Tej, S. S., Poethig, R. S., et al. (2006). MicroRNAs
and other small RNAs enriched in the Arabidopsis RNA-dependent RNA polymerase-2 mutant.
Genome Res. 16, 1276–1288. doi:10.1101/gr.5530106.
Mao, G., Meng, X., Liu, Y., Zheng, Z., Chen, Z., and Zhang, S. (2011). Phosphorylation of a WRKY
Transcription Factor by Two Pathogen-Responsive MAPKs Drives Phytoalexin Biosynthesis in
Arabidopsis. Plant Cell 23, 1639–1653. doi:10.1105/tpc.111.084996.
Marnef, A., and Standart, N. (2010). Pat1 proteins: a life in translation, translation repression and mRNA
decay. Biochem. Soc. Trans. 38, 1602–1607. doi:10.1042/BST0381602.
Martínez de Alba, A. E., Moreno, A. B., Gabriel, M., Mallory, A. C., Christ, A., Bounon, R., et al. (2015). In
plants, decapping prevents RDR6-dependent production of small interfering RNAs from endogenous
mRNAs. Nucleic Acids Res. 43, 2902–2913. doi:10.1093/nar/gkv119.
Meister, G. (2013). Argonaute proteins: functional insights and emerging roles. Nat. Rev. Genet. 14, 447–
459. doi:10.1038/nrg3462.
Meyers, B. C., Axtell, M. J., Bartel, B., Bartel, D. P., Baulcombe, D., Bowman, J. L., et al. (2008). Criteria
for Annotation of Plant MicroRNAs. Plant Cell 20, 3186–3190. doi:10.1105/tpc.108.064311.
Miya, A., Albert, P., Shinya, T., Desaki, Y., Ichimura, K., Shirasu, K., et al. (2007). CERK1, a LysM
receptor kinase, is essential for chitin elicitor signaling in Arabidopsis. Proc. Natl. Acad. Sci. 104,
19613–19618. doi:10.1073/pnas.0705147104.
Nandety, R. S., Caplan, J. L., Cavanaugh, K., Perroud, B., Wroblewski, T., Michelmore, R. W., et al. (2013).
The Role of TIR-NBS and TIR-X Proteins in Plant Basal Defense Responses. Plant Physiol. 162,
1459–1472. doi:10.1104/pp.113.219162.
38
Navarro, L., Dunoyer, P., Jay, F., Arnold, B., Dharmasiri, N., Estelle, M., et al. (2006). A Plant miRNA
Contributes to Antibacterial Resistance by Repressing Auxin Signaling. Science 312, 436–439.
doi:10.1126/science.1126088.
Navarro, L., Jay, F., Nomura, K., He, S. Y., and Voinnet, O. (2008). Suppression of the MicroRNA Pathway
by Bacterial Effector Proteins. Science 321, 964–967. doi:10.1126/science.1159505.
Nguyen, A. H., Matsui, A., Tanaka, M., Mizunashi, K., Nakaminami, K., Hayashi, M., et al. (2015). Loss of
Arabidopsis 5’-3’ exoribonuclease AtXRN4 function enhances heat stress tolerance of plants
subjected to severe heat stress. Plant Cell Physiol., pcv096. doi:10.1093/pcp/pcv096.
Olmedo, G., Guo, H., Gregory, B. D., Nourizadeh, S. D., Aguilar-Henonin, L., Li, H., et al. (2006).
ETHYLENE-INSENSITIVE5 encodes a 5′→3′ exoribonuclease required for regulation of the EIN3targeting F-box proteins EBF1/2. Proc. Natl. Acad. Sci. 103, 13286–13293.
doi:10.1073/pnas.0605528103.
Palma, K., Thorgrimsen, S., Malinovsky, F. G., Fiil, B. K., Nielsen, H. B., Brodersen, P., et al. (2010).
Autoimmunity in Arabidopsis acd11 Is Mediated by Epigenetic Regulation of an Immune Receptor.
PLoS Pathog 6, e1001137. doi:10.1371/journal.ppat.1001137.
Parker, J. E., Holub, E. B., Frost, L. N., Falk, A., Gunn, N. D., and Daniels, M. J. (1996). Characterization of
eds1, a mutation in Arabidopsis suppressing resistance to Peronospora parasitica specified by several
different RPP genes. Plant Cell 8, 2033–2046. doi:10.1105/tpc.8.11.2033.
Parker, R., and Sheth, U. (2007). P Bodies and the Control of mRNA Translation and Degradation. Mol. Cell
25, 635–646. doi:10.1016/j.molcel.2007.02.011.
Petersen, M., Brodersen, P., Naested, H., Andreasson, E., Lindhart, U., Johansen, B., et al. (2000).
Arabidopsis MAP Kinase 4 Negatively Regulates Systemic Acquired Resistance. Cell 103, 1111–
1120. doi:10.1016/S0092-8674(00)00213-0.
Potuschak, T., Lechner, E., Parmentier, Y., Yanagisawa, S., Grava, S., Koncz, C., et al. (2003). EIN3Dependent Regulation of Plant Ethylene Hormone Signaling by Two Arabidopsis F Box Proteins:
EBF1 and EBF2. Cell 115, 679–689. doi:10.1016/S0092-8674(03)00968-1.
Potuschak, T., Vansiri, A., Binder, B. M., Lechner, E., Vierstra, R. D., and Genschik, P. (2006). The
Exoribonuclease XRN4 Is a Component of the Ethylene Response Pathway in Arabidopsis. Plant
Cell Online 18, 3047–3057. doi:10.1105/tpc.106.046508.
Qiu, J.-L., Zhou, L., Yun, B.-W., Nielsen, H. B., Fiil, B. K., Petersen, K., et al. (2008). Arabidopsis MitogenActivated Protein Kinase Kinases MKK1 and MKK2 Have Overlapping Functions in Defense
Signaling Mediated by MEKK1, MPK4, and MKS1. Plant Physiol. 148, 212–222.
doi:10.1104/pp.108.120006.
Rasmussen, M. W., Roux, M., Petersen, M., and Mundy, J. (2012). MAP kinase cascades in Arabidopsis
innate immunity. Plant Proteomics 3, 169. doi:10.3389/fpls.2012.00169.
Rasmussen, S., Barah, P., Suarez-Rodriguez, M. C., Bressendorff, S., Friis, P., Costantino, P., et al. (2013).
Transcriptome Responses to Combinations of Stresses in Arabidopsis. Plant Physiol. 161, 1783–
1794. doi:10.1104/pp.112.210773.
Rayson, S., Arciga-Reyes, L., Wootton, L., De Torres Zabala, M., Truman, W., Graham, N., et al. (2012a). A
Role for Nonsense-Mediated mRNA Decay in Plants: Pathogen Responses Are Induced in
Arabidopsis thaliana NMD Mutants. PLoS ONE 7, e31917. doi:10.1371/journal.pone.0031917.
39
Rayson, S., Arciga-Reyes, L., Wootton, L., De Torres Zabala, M., Truman, W., Graham, N., et al. (2012b). A
Role for Nonsense-Mediated mRNA Decay in Plants: Pathogen Responses Are Induced in
Arabidopsis thaliana NMD Mutants. PLoS ONE 7, e31917. doi:10.1371/journal.pone.0031917.
Riehs-Kearnan, N., Gloggnitzer, J., Dekrout, B., Jonak, C., and Riha, K. (2012). Aberrant growth and
lethality of Arabidopsis deficient in nonsense-mediated RNA decay factors is caused by
autoimmune-like response. Nucleic Acids Res., gks195. doi:10.1093/nar/gks195.
Rodriguez, M. C. S., Petersen, M., and Mundy, J. (2010). Mitogen-Activated Protein Kinase Signaling in
Plants. Annu. Rev. Plant Biol. 61, 621–649. doi:10.1146/annurev-arplant-042809-112252.
Roux, M. E., Rasmussen, M. W., Palma, K., Lolle, S., Regué, À. M., Bethke, G., et al. (2015). The mRNA
decay factor PAT1 functions in a pathway including MAP kinase 4 and immune receptor SUMM2.
EMBO J., e201488645. doi:10.15252/embj.201488645.
Seo, J.-K., Wu, J., Lii, Y., Li, Y., and Jin, H. (2013). Contribution of Small RNA Pathway Components in
Plant Immunity. Mol. Plant-Microbe Interact. MPMI 26, 617–625. doi:10.1094/MPMI-10-12-0255IA.
Shivaprasad, P. V., Chen, H.-M., Patel, K., Bond, D. M., Santos, B. A. C. M., and Baulcombe, D. C. (2012).
A MicroRNA Superfamily Regulates Nucleotide Binding Site–Leucine-Rich Repeats and Other
mRNAs. Plant Cell Online, tpc.111.095380. doi:10.1105/tpc.111.095380.
Sidaway-Lee, K., Costa, M. J., Rand, D. A., Finkenstadt, B., and Penfield, S. (2014). Direct measurement of
transcription rates reveals multiple mechanisms for configuration of the Arabidopsis ambient
temperature response. Genome Biol. 15, R45. doi:10.1186/gb-2014-15-3-r45.
Smith, J. E., and Baker, K. E. (2015). Nonsense-mediated RNA decay – a switch and dial for regulating gene
expression. BioEssays, n/a–n/a. doi:10.1002/bies.201500007.
Solano, R., Stepanova, A., Chao, Q., and Ecker, J. R. (1998). Nuclear events in ethylene signaling: a
transcriptional cascade mediated by ETHYLENE-INSENSITIVE3 and ETHYLENE-RESPONSEFACTOR1. Genes Dev. 12, 3703–3714. doi:10.1101/gad.12.23.3703.
Souret, F. F., Kastenmayer, J. P., and Green, P. J. (2004). AtXRN4 Degrades mRNA in Arabidopsis and Its
Substrates Include Selected miRNA Targets. Mol. Cell 15, 173–183.
doi:10.1016/j.molcel.2004.06.006.
Wachter, A., and Hartmann, L. (2014). NMD: Nonsense-Mediated Defense. Cell Host Microbe 16, 273–275.
doi:10.1016/j.chom.2014.08.015.
Xiang, T., Zong, N., Zou, Y., Wu, Y., Zhang, J., Xing, W., et al. (2008). Pseudomonas syringae Effector
AvrPto Blocks Innate Immunity by Targeting Receptor Kinases. Curr. Biol. 18, 74–80.
doi:10.1016/j.cub.2007.12.020.
Xu, J., and Chua, N.-H. (2012). Dehydration stress activates Arabidopsis MPK6 to signal DCP1
phosphorylation. EMBO J. 31, 1975–1984. doi:10.1038/emboj.2012.56.
Xu, J., Yang, J.-Y., Niu, Q.-W., and Chua, N.-H. (2006). Arabidopsis DCP2, DCP1, and VARICOSE Form a
Decapping Complex Required for Postembryonic Development. Plant Cell 18, 3386–3398.
doi:10.1105/tpc.106.047605.
Yanagisawa, S., Yoo, S.-D., and Sheen, J. (2003). Differential regulation of EIN3 stability by glucose and
ethylene signalling in plants. Nature 425, 521–525. doi:10.1038/nature01984.
40
Yi, H., and Richards, E. J. (2007). A Cluster of Disease Resistance Genes in Arabidopsis Is Coordinately
Regulated by Transcriptional Activation and RNA Silencing. Plant Cell 19, 2929–2939.
doi:10.1105/tpc.107.051821.
Zhang, X., Zhu, Y., Liu, X., Hong, X., Xu, Y., Zhu, P., et al. (2015). Suppression of endogenous gene
silencing by bidirectional cytoplasmic RNA decay in Arabidopsis. Science 348, 120–123.
doi:10.1126/science.aaa2618.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., et al. (2012). Disruption of PAMP-Induced MAP
Kinase Cascade by a Pseudomonas syringae Effector Activates Plant Immunity Mediated by the NBLRR Protein SUMM2. Cell Host Microbe 11, 253–263. doi:10.1016/j.chom.2012.01.015.
Zhu, Y., Qian, W., and Hua, J. (2010). Temperature Modulates Plant Defense Responses through NB-LRR
Proteins. PLoS Pathog 6, e1000844. doi:10.1371/journal.ppat.1000844.
Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J. D. G., Boller, T., et al. (2006). Perception of the
Bacterial PAMP EF-Tu by the Receptor EFR Restricts Agrobacterium-Mediated Transformation.
Cell 125, 749–760. doi:10.1016/j.cell.2006.03.03
41
Published online: January 20, 2015
Article
The mRNA decay factor PAT1 functions in a
pathway including MAP kinase 4 and immune
receptor SUMM2
Milena Edna Roux1,†, Magnus Wohlfahrt Rasmussen1,†, Kristoffer Palma2, Signe Lolle1, Àngels Mateu
Regué1, Gerit Bethke3, Jane Glazebrook3, Weiping Zhang4, Leslie Sieburth4, Martin R Larsen5,
John Mundy1 & Morten Petersen1,*
Abstract
Multi-layered defense responses are activated in plants upon
recognition of invading pathogens. Transmembrane receptors
recognize conserved pathogen-associated molecular patterns
(PAMPs) and activate MAP kinase cascades, which regulate
changes in gene expression to produce appropriate immune
responses. For example, Arabidopsis MAP kinase 4 (MPK4) regulates
the expression of a subset of defense genes via at least one WRKY
transcription factor. We report here that MPK4 is found in
complexes in vivo with PAT1, a component of the mRNA decapping
machinery. PAT1 is also phosphorylated by MPK4 and, upon flagellin PAMP treatment, PAT1 accumulates and localizes to cytoplasmic processing (P) bodies which are sites for mRNA decay.
Pat1 mutants exhibit dwarfism and de-repressed immunity dependent on the immune receptor SUMM2. Since mRNA decapping is a
critical step in mRNA turnover, linking MPK4 to mRNA decay via
PAT1 provides another mechanism by which MPK4 may rapidly
instigate immune responses.
Keywords decapping; immunity; MAP kinases; mRNA decay; phosphorylation
Subject Categories Microbiology, Virology & Host Pathogen Interaction;
Plant Biology; RNA Biology
DOI 10.15252/embj.201488645 | Received 2 April 2014 | Revised 2 December
2014 | Accepted 11 December 2014
Introduction
Plant innate immunity employs multilayered defense responses
comprised of two overlapping mechanisms. In the first layer, plant
pattern recognition receptors detect invading microorganisms by the
1
2
3
4
5
presence of conserved pathogen-associated molecular patterns
(PAMPs)(Boller & Felix, 2009). PAMP recognition is exemplified by
the binding of the bacterial flagellin-derived flg22 peptide to the
leucine-rich repeat-receptor-like kinase flagellin sensing 2 (FLS2)
(Gomez-Gomez et al, 2001; Chinchilla et al, 2006). PAMP recognition
initiates downstream signaling, including production of reactive
oxygen species, calcium influx, MAP kinase activation and global
changes in gene expression that induce PAMP-triggered immunity
(PTI) (Chisholm et al, 2006; Zipfel, 2009). Adapted pathogens have
evolved effector proteins that are delivered into host cells to compromise PTI by evading PAMP detection or suppressing defense
responses. In the second layer of immunity, plant resistance (R)
proteins have evolved to directly or indirectly recognize the activities
of pathogen effectors (Jones & Dangl, 2006). In the best-studied examples, R proteins are found to guard host proteins or complexes (guardees) manipulated by specific pathogen effectors. Pathogen detection
via R proteins leads to induction of strong defenses and to a form of
host programmed cell death known as the hypersensitive response
(HR) to sequester infections. These responses are collectively termed
effector-triggered immunity (ETI) (Jones & Dangl, 2006).
Changes in phosphorylation are important regulatory mechanisms
in cellular signaling. Activation of mitogen-activated protein (MAP)
kinases occurs within 10 min of PAMP application (Asai et al, 2002;
Boller & Felix, 2009) and relies on sequential phosphorylations
between MAPKK-kinases (MEKK), MAPK kinases (MKK) and MAP
kinases (MPK) (Pitzschke et al, 2009; Andreasson & Ellis, 2010;
Rasmussen et al, 2012). In Arabidopsis, several MPKs are activated
by PAMPs. A cascade comprising MEKK1-MKK4/5-MPK3/6 was
initially found to be involved in PTI downstream of FLS2 (Asai et al,
2002; Droillard et al, 2004; Gao et al, 2008). Similarly, flg22 treatment
activates another cascade including MEKK1, MKK1/2 and MPK4
(Petersen et al, 2000; Ichimura et al, 2006; Mészáros et al, 2006;
Suarez-Rodriguez et al, 2007; Gao et al, 2008; Qiu et al, 2008a).
Department of Biology, Faculty of Science, University of Copenhagen, Copenhagen, Denmark
New England Biolabs Ltd, Whitby, ON, Canada
Department of Plant Biology, University of Minnesota, St. Paul, MN, USA
Department of Biology, University of Utah, Salt Lake City, UT, USA
University of Southern Denmark, Institute for Biochemistry and Molecular Biology, Odense, Denmark
*Corresponding author. Tel: +45 353 22127; E-mail: [email protected]
†
These authors contributed equally to this work
42
Published online: January 20, 2015
The EMBO Journal
Interestingly, Flg22-induced activation of MPK3, MPK4 and MPK6 is
dependent on MKK1, while MPK3 and MPK6 are also activated by
MKK4 (Mészáros et al, 2006). Thus, FLS2 activates two cascades, one
with an unknown MEKK and MKK4/5-MPK3/6, the other with
MEKK1-MKK1/2-MPK4. More recently, the closest homologue of
MPK4, MPK11, was shown to also be activated by PAMPs (Bethke
et al, 2012).
mpk4 mutants were originally found to exhibit autoimmunity,
and MPK4 thus appeared to function genetically as a negative regulator of defense responses (Petersen et al, 2000; Droillard et al,
2004). However, MPK4 is activated in response to pathogens and
PAMP elicitation (Droillard et al, 2004; Teige et al, 2004; Ichimura
et al, 2006; Brader et al, 2007; Qiu et al, 2008a) which is counterintuitive for a negative regulator. Subsequently, it was shown that
activated MPK4 interacts with and phosphorylates MAP kinase
substrate 1 (MKS1), bringing about the release of the transcription
factor WRKY33 and induction of the expression of the PHYTOALEXIN DEFICIENT 3 (PAD3) gene required for biosynthesis of the
antimicrobial camalexin (Andreasson et al, 2005; Qiu et al, 2008b).
This illustrates how MPK4 functions as a positive regulator of PTI.
Since it was recently reported that MPK4 is a target for manipulation by pathogen effectors (Zhang et al, 2007, 2012), and as mpk4
mutant phenotypes are partially suppressed by mutations in the R
protein SUMM2 (suppressor or mkk1 mkk2) (Zhang et al, 2012),
one model is that MPK4 is a PTI guardee whose absence triggers
ETI.
Apart from regulation by transcription factors, mRNA translation
and degradation also regulate gene expression, especially when they
are reprogrammed to stabilize bulk mRNAs and favor mRNAs
required for an appropriate stress response (Jiao et al, 2010;
Munchel et al, 2011; Park et al, 2012; Ravet et al, 2012). mRNA is
intrinsically unstable to facilitate rapid changes in mRNA abundance
in response to stimuli. Eukaryotic mRNAs contain stability determinants including the 50 7-methylguanosine triphosphate cap (m7G)
and the 30 poly(A) tail. To achieve fine control of transcript
abundance, mRNA is attacked by a decapping complex that breaks
down RNA in a coordinated manner. mRNA decay is initiated by
deadenylation, the removal of the 30 poly(A) tail, whereafter mRNA
can either be routed to 30 –50 degradation by exosomal exonucleases,
or 50 –30 by the exoribonuclease activity of the decapping complex
(Garneau et al, 2007). The decapping complex is highly conserved
among eukaryotes (Garneau et al, 2007; Xu & Chua, 2011) and
comprises the decapping enzyme DCP1/2 which removes the 50 m7G
cap, and the exoribonuclease XRN that degrades the monophosphorylated mRNA. Deadenylation and decapping are linked in
eukaryotes by the decapping enhancer PAT1 (protein associated
with topoisomerase II) (Hatfield et al, 1996; Wang et al, 1996;
Tharun & Parker, 2001; Ozgur et al, 2010). In yeast, PAT1 promotes
the interaction between mRNA and decapping enzyme, altering
the messenger ribonucleoprotein (mRNP) organization from an
arrangement favoring translation to one promoting decay. mRNA
decay occurs in distinct cytoplasmic foci called processing bodies
(PBs) in eukaryotes (Parker & Sheth, 2007; Balagopal & Parker,
2009), where mRNPs are sequestered for degradation, or from which
they may re-enter polysomal complexes (Brengues, 2005). PBs can
also harbor translationally arrested mRNPs and RNA silencing
machinery (Balagopal & Parker, 2009; Xu & Chua, 2009; Thomas
et al, 2011).
mRNA decapping component PAT1 in immunity
Milena Edna Roux et al
In plants, the decapping machinery includes the decapping
enzyme DCP2 and its activators DCP1, DCP5, VARICOSE (VCS)
(Deyholos et al, 2003; Xu et al, 2006; Goeres et al, 2007; Brodersen
et al, 2008; Xu & Chua, 2009, 2011; Motomura et al, 2012) as well
as the exoribonuclease XRN4 (Olmedo et al, 2006; Potuschak
et al, 2006; Gregory et al, 2008; Rymarquis et al, 2011; Vogel et al,
2011). It is thought that DCP1, DHH1 and DCP5 form mRNPs for
translational repression of target mRNA, which are then subject to
decapping by recruitment of DCP2 and VCS and digestion by XRN4
(Xu & Chua, 2009, 2011). Importantly, although PAT1 functions have
not previously been studied in plants, three homologues are encoded
in the Arabidopsis genome, each of which contains a conserved
C-terminal domain. The decapping activator Sm-like (LSM) proteins,
which interact with PAT1 in eukaryotes (Salgado-Garrido et al, 1999;
Bonnerot et al, 2000; Bouveret et al, 2000; Tharun et al, 2000;
Tharun, 2009), have recently been characterized in Arabidopsis
(Perea-Resa et al, 2012; Golisz et al, 2013). It was found that
LSM1-7 proteins form a complex and lsm1 mutants accumulate
capped mRNA. Furthermore, VCS, DHH1 and PAT1 homologs were
identified in LSM1 immunoprecipitates (Golisz et al, 2013).
The decapping complex plays important roles in eukaryotic
development. In contrast, links between mRNA decapping and
stress signaling are just being uncovered (Jiao et al, 2010; Buchan
et al, 2011; Munchel et al, 2011; Park et al, 2012), and how decapping may be involved in regulating immune systems is largely
unknown. Here, we characterize the Arabidopsis homologue of the
mRNA decay regulator PAT1. We show that PAT1 functions in
decapping of mRNA. Furthermore, PAT1 is phosphorylated in
response to flg22 and localizes to discrete, punctate foci in the cytosol. PAT1 also interacts with MPK4 and with the R protein SUMM2
in planta, and the absence of PAT1 triggers SUMM2 dependent
immunity. This indicates that PAT1 is regulated by MPK4 in a pathway whose disruption leads to ETI via SUMM2.
Results
AtPAT1 is an ScPAT1 orthologue and interacts with
MPK4 in planta
We previously conducted a yeast two-hybrid screen to identify Arabidopsis proteins that interact with MPK4 (Andreasson et al, 2005). In
addition to the MPK4 substrate MKS1, we identified two clones encoding PAT1 (At1g79090), an mRNA decapping stimulator involved in
post-transcriptional gene regulation (Coller & Parker, 2005). Two PAT1
homologs are encoded in the Arabidopsis genome (AtPAT1H1
At3g22270; AtPAT1H2 At4g14990, Supplementary Fig S1A). The
steady-state expression level of PAT1 and the homologs compared to
the housekeeping gene ACTIN8 (At1g49240) was analyzed by RT-PCR
(Supplementary Fig S1B). We focused on PAT1 as the other two homologs were not identified by yeast two-hybrids. To confirm the PAT1MPK4 interaction in planta, doubly transgenic Arabidopsis lines were
generated in the Ler mpk4-1 background that expressed PAT1 with a
C-terminal Myc tag and HA-tagged MPK4 under the control of their own
promoters. Anti-Myc immunoprecipitation from either MPK4-HA or
double transgenic MPK4-HA/Pat1-Myc tissue detected a 50 kDa band
corresponding to MPK4-HA only in double transgenic lines (Fig 1A).
Thus, MPK4 and PAT1 can be found in complex in Arabidopsis.
43
Published online: January 20, 2015
Milena Edna Roux et al
The EMBO Journal
mRNA decapping component PAT1 in immunity
A
Input
Myc IP
PAT1-Myc
-
+
-
+
MPK4-HA
+
+
+
+
50
WB: anti-HA
100
PAT1-Myc
WB: anti-Myc
50
kDa
Dilution
B
0
1
2
Dilution
3
4
5
0
1
2
3
4
5
B4742
B4742+AtPAT1
pat1
HA WB
GFP WB
Rubisco
EIF4A1
GFP IP
GFP WB
Col-0 pat1-1
EXPL1
UGT87A2
Input
HA WB
D
PAT1-HA
C
LSM1-GFP
PAT1-HA+LSM1-GFP
pat1 +AtPAT1
Figure 1. MPK4 and PAT1 interact in Arabidopsis.
A MPK4-HA is detected in PAT1-Myc immunoprecipitates from double transgenic Arabidopsis plants. Immunoblots of input and anti-Myc IPs probed with anti-HA and
anti-Myc antibodies. Left panel, input; right panel, anti-Myc IP.
B Yeast pat1D mutants are unable to grow at 37°C while wild-type (B4742) strain grows at 30°C (left panel) and 37°C (right). Growth at 37°C is restored in pat1D
expressing AtPAT1. Serial dilutions of each yeast strain were plated on YPAD agar plates and grown at the indicated temperatures.
C Co-IP between PAT1-HA and LSM1-GFP. Proteins were transiently co-expressed in N. benthamiana and tissue harvested 3 days post-infiltration. Immunoblots of GFP
IPs (top panel) and inputs (20 lg each, bottom panel). Immunoblots were cut in half and probed with anti-HA antibodies and anti-GFP antibodies.
D Agarose gel electrophoresis of 50 RACE PCR to detect capped transcripts of EIF4A1, UGT87A2 and EXPL1 in 14-day-old Col-0 and pat1-1.
Source data are available online for this figure.
Yeast PAT1 engages with translating mRNPs and is involved in
translational repression and decapping activation (Marnef & Standart, 2010). Since the function of PAT1 in Arabidopsis was
unknown, we examined whether it functions similarly to yeast
PAT1. To this end, a full-length Arabidopsis PAT1 cDNA was cloned
from Col-0 (Supplementary Fig S1C and D) and transformed into
wild-type yeast (B4742) and a yeast mutant (Y15797) in which yeast
PAT1 was replaced with a G418 resistance cassette (BY4742
(YCR077c) pat1D::KanMX). In contrast to the wild-type, yeast lacking PAT1 (pat1D) display a temperature-sensitive phenotype and
are impaired at 37°C but grow normally at 30°C (Tharun et al,
2005). This phenotype is reverted to wild-type in yeast containing
Arabidopsis PAT1, as growth at 37°C was restored (Fig 1B). As an
additional control, we transformed Arabidopsis PAT1 into wild-type
44
Published online: January 20, 2015
The EMBO Journal
mRNA decapping component PAT1 in immunity
yeast (B4742/AtPAT1), and this grew similarly to wild-type at 30°C
and almost as well at the wild-type at 37°C (Fig 1B). The expression
of Arabidopsis PAT1 in yeast was confirmed by anti-PAT1 immunoblotting of yeast protein extracts (Supplementary Fig S1E). This
provides compelling evidence for the orthologous functions of these
yeast and Arabidopsis PAT1 proteins. As PAT1 is found in complex
with MPK4, these results provide a link between MPK4 and posttranscriptional regulation of mRNA stability.
We next analyzed the interaction between PAT1 and conserved
components of mRNA decapping. PAT1-LSM1-7 complexes function
in mRNA decapping and deadenylation (Bouveret et al, 2000;
Tharun, 2009; Haas et al, 2010; Ozgur et al, 2010; Totaro et al,
2011). We therefore transiently expressed in Nicotiana benthamiana
LSM1-GFP and PAT1-HA and then immunoprecipitated LSM1 with
GFP Trap beads. PAT1-HA could be detected in LSM1 immunoprecipitates but did not adhere to GFP Trap beads in the absence of
LSM1-GFP (Fig 1C). This is consistent with the detection of peptides
corresponding to PAT1 and its homologues in LSM1 immunoprecipitates (Golisz et al, 2013) and supports a role of PAT1 in mRNA
decapping. In other organisms, interactions between PAT1 and
LSM1 are robust, while those between PAT1 and other mRNA decapping proteins, including the DCP1-DCP2 complex and XRN1, are
more transient (Bouveret et al, 2000; Nissan et al, 2010; Ozgur et al,
2010). This is consistent with our difficulty in detecting DCP1 in
complex with PAT1 in Arabidopsis (Supplementary Fig S1F).
Milena Edna Roux et al
PAT1 is an MPK4 substrate
Since MPK4 and PAT1 are found in complexes in planta, we asked
whether PAT1 is an MPK substrate. PAT1 contains 5 Ser-Pro (SP)
motifs which are commonly phosphorylated by MPKs (Pearson
et al, 2001; Ubersax & Ferrell, 2007) (Supplementary Fig S1D). To
characterize PAT1 phosphorylation in vivo, we identified PAT1
phosphopeptides by mass spectrometry. Since MPK3/4/6 are activated by flagellin and by virulent strains of Pseudomonas syringae
pv. tomato (Pto) DC3000 (Asai et al, 2002; Brader et al, 2007;
Suarez-Rodriguez et al, 2007; Bethke et al, 2009, 2012; Rasmussen
et al, 2012), PAT1 was immunoprecipitated from extracts of
untreated control and flg22-treated wild-type Col-0 or PAT1-GFP
transgenic lines. Bands corresponding to PAT1-GFP (130 kDa)
were excised from the gel (Supplementary Fig S2), subjected to ingel tryptic digestion, and peptides were extracted. Phosphopeptides
were enriched by TiO2 chromatography and analyzed by liquid
chromatography (RP-HPLC) coupled to electrospray ionization
tandem mass spectrometry (LC-ESI-MS/MS). This identified several
phosphopeptides from PAT1-GFP IPs that were not detectable in
the negative control (Col-0) (Table 1; Supplementary Fig S2). The
most abundant phosphopeptide, based on the extracted ion chromatogram from the LC-MS/MS analysis, revealed phosphorylation
of Ser208 in an SP motif (Fig 2A). Another peptide was identified
with phosphorylation of Ser343. However, this site is not within
an SP motif, making it a less likely site for phosphorylation by
MPKs. Importantly, both these peptides have previously been
detected in Arabidopsis by mass spectrometry (Phosphat Database,
http://phosphat.mpimp-golm.mpg.de/). It should be noted that the
PAT1 phosphopeptides were detected both whether or not the
sample had been treated with flg22. Thus, under the conditions
used here, PAT1 was phosphorylated and remained so after exposure to flg22.
To determine whether PAT1 is a substrate of a specific MPK, we
carried out in vitro kinase assays using immunoprecipitated, PAMPactivated MPKs with purified His6-PAT1 protein. Phosphorylated
His6-PAT1 was detectable as a radioactive band around 95 kDa after
incubation with flg22-activated MPK4 (Fig 2B), while MPK6 caused
only low levels of PAT1 phosphorylation and MPK3 did not significantly alter PAT1 phosphorylation (Supplementary Fig S3). Each
MPK was also incubated with the generic MPK substrate myelin
basic protein (MBP) to verify their activation (Fig 2B; Supplementary Fig S3). This confirms that activated MPK4, and to a lesser
extent, MPK6, is responsible for the phosphorylation of PAT1. A
version of His6-PAT1 with Ser208 mutated to alanine (S208A)
was also incubated with MPK4 and MPK6 IPs, and this mutant
PAT1 is required for decapping of selected mRNAs
In order to determine whether PAT1 behaves as an activator of
mRNA decapping, we used 50 RACE to compare the levels of
capped mRNAs in Col-0 and pat1 mutants. To this end, we identified an allele, pat1-1 (Salk_040660), with a T-DNA insertion in the
last exon of PAT1 (Supplementary Fig S1C). We also generated an
anti-PAT1 antibody against a C-terminal peptide (Supplementary
Fig S1D). Immunoblotting of Col-0 protein extracts with this antibody detected a clear band around 90 kDa. In contrast, no protein
could be detected in pat1-1 mutant extracts (Supplementary Fig
S1G). This indicates that pat1-1 harbors either a truncated version
of PAT1, no PAT1 protein, or levels of the protein that are below
detection.
50 RACE was performed on transcripts known to be degraded by the
decapping complex (EXPL1; UGT87A2) (Perea-Resa et al, 2012), as well
as a housekeeping transcript EIF4A1. We found that capped EXPL1 and
UGT87A2 accumulated in pat1-1 mutants, while capped EIF4A1 mRNA
was present in equal amounts in Col-0 and pat1-1 (Fig 1D). This
indicates that PAT1 plays a role in mRNA decay via decapping.
Table 1. Phosphopeptides identified in PAT1-GFP IPs by mass spectrometry analysis.
Mascot score
Sequence
Phospho-sitesa
m/z
PAT1GFP water
PAT1GFP flg22
SSFVSYPPPGSISPDQR
1 (S208)
950.93005
71
79
SSFVSYPPPGSISPDQR
2 (S208, S200)
990.91333
46
64
SSSGNYDGMLGFGDLR
1 (S343)
929.87323
70
114
SSSGNYDGMLGFGDLR
2 (S343, S342)
969.85663
39
50
a
Phosphosites refers to the number and in brackets the location of the phosphorylation detected by MS analysis. Potential phosphorylation sites are indicated in
bold letters. m/z refers to the mass-to-charge ratio of the tabulated peptides.
45
Published online: January 20, 2015
Milena Edna Roux et al
The EMBO Journal
mRNA decapping component PAT1 in immunity
A
1230.50
y11+P
100
95
90
b6
b13
b15
SSFVSYPPPGSISPDQR
85
80
y15 y14 y13 y12 y11 y10 y9
y8
y7
y5
y4
y2
75
70
Relative Abundance
65
60
55
50
2+
(M+H)
45
40
- H3PO4
b6
35
y11+P – H3PO4
y10+P
1400
b15+P
y15+P
1708.75
1726.71
1393.42
1300
Y13+P
1598.08
b13+P – H3PO4
1480.58
y9+P
y14+P
y12+P
1132.67
901.92
y8+P
1288.25
y7+P
1036.42
y5+P
863.75
5
515.25
536.08
10
y4
y2
303.17
15
615.75
20
671.08
682.17
25
932.75
653.08
30
0
300
400
500
600
700
800
900
1000
1100
1200
1500
1600
1700
1800
1900
m/z
B
His6-PAT1
-
4
His6-PAT1 S208A
6
-
4
6
MBP
-
4
6
Autoradiogram
CBB Gel
Figure 2. PAT1 is phosphorylated in planta and in vitro.
A Tandem MS spectrum of the phosphopeptide SSFVSYPPPGSISPDQR in which the underlined serine is phosphorylated. Y- and b-ions are indicated that localize
phosphorylation to the SP site.
B MPK4 and MPK6 were immunoprecipitated from extracts of Col-0 seedlings treated with 200 nM flg22 for 10 min. For the negative control (), extracts were
incubated with agarose beads without antibodies. IPs were incubated with His6-PAT1, His6-PAT1 S208A or MBP for 60 min at 37°C before boiling and SDS-PAGE.
Autoradiogram (top panel), Coomassie-stained gel for loading control (bottom).
Source data are available online for this figure.
had significantly lower levels of phosphorylation (Fig 2B). This
supports the identification of S208 as a key phosphorylation site in
PAT1.
pat1 mutants exhibit autoimmunity similar to mpk4
Given that MPK4 is an important immune regulator in Arabidopsis
and that mpk4 mutants have de-repressed defense responses, we
examined whether PAT1 may also be involved in immunity. The
pat1-1 mutant has a distinct leaf serration phenotype and a slightly
smaller rosette than Col-0 (Fig 3). A similar rosette phenotype was
seen in another T-DNA insertion line (pat1-2, WiscDsLox_734_D04
in Supplementary Figs S1C and S4), indicating that this phenotype is
due to loss-of-function of PAT1. Importantly, however, the phenotype of pat1-1 is not as extreme as mpk4-2, which is much smaller
and has more pronounced leaf curling and reduced fertility as well
as constitutive defense gene expression (Petersen et al, 2000);
Fig 3).
In order to determine whether pat1-1 is a constitutive defense
mutant similar to mpk4, quantitative reverse-transcription PCR
(qRT-PCR) was used to measure the steady-state level of the
pathogenesis-related PR1 and PR2 genes in adult plants (Fig 4A).
Previous work showed that the enhanced defense response of mpk4
mutants is suppressed by mutations in EDS1 (Brodersen et al, 2006).
Thus, eds1-2 was crossed with pat1-1 to explore the pat1 phenotype in the absence of this regulator (Fig 3). Compared to
Col-0, pat1-1 mutants accumulated 1,000-fold more PR1 and 150-fold
more PR2 transcripts (Fig 4A). In contrast, the levels of PR mRNAs
in pat1-1 eds1-2 double mutants were similar to those in wild-type
(Fig 4A). Under the same conditions, mpk4-2 mutants accumulated
4,000-fold more PR1 and 800-fold more PR2 transcripts (Fig 4A).
We found that the elevated PR gene expression in pat1-1 correlated to an enhanced resistance to infection by syringe-infiltrated
Pto DC3000 when compared to Col-0 (Fig 4B). While Pto DC3000
growth reached 5.5 cfu/cm2 in Col-0, pat1-1 mutants supported
tenfold lower accumulation of Pto DC3000. In this experiment,
eds1-2 mutants showed enhanced bacterial growth as expected for
this mutant with compromised defense responses (Fig 4B). pat1
disease resistance is EDS1 dependent, as bacterial growth pat1-1
eds1-2 double mutants was similar to that in eds1-2 (Fig 4B). These
findings indicate that, similar to mpk4, pat1 mutants express
EDS1-dependent autoimmunity in the absence of microbes.
46
Published online: January 20, 2015
The EMBO Journal
mRNA decapping component PAT1 in immunity
Col-0
pat1-1
eds1-2
Milena Edna Roux et al
mpk4-2
pat1-1 eds1-2
Figure 3. pat1 mutants have a serrated leaf, semi-dwarf phenotype.
Plants photographed at 4 weeks of growth in short day conditions, genotypes as indicated. Col-0 and eds1-2 plants and pat1-1 and pat1-1/eds1-2, respectively, were placed in
the middle of the same pots before the pictures were taken.
PAT1 detection in P-bodies is induced by PAMPs
Processing bodies (PBs) are cytoplasmic granules involved in both
mRNA decay and translational repression pathways (Kulkarni et al,
2010). PAT1 is a conserved PB component in yeast (Rodriguez-Cousino
et al, 1995), C. elegans (Boag et al, 2008), Drosophila (Marnef et al,
2010) and mammals (Scheller et al, 2007) and PBs can be significantly induced in Arabidopsis by hypoxia and heat stress (Weber
et al, 2008). To investigate whether PAT1 is found in PBs, we
produced transgenic Arabidopsis lines that express PAT1-GFP from
its native promoter complementing the pat1 phenotype (Supplementary Fig S5A and B). We detected GFP signal in small numbers of
distinct foci by confocal microscopy in roots of young seedlings
(Fig 5A). Within 20 min of flg22 treatment, a significant increase in
the number of GFP-positive foci could be seen in the root tips
(Fig 5A). To test whether these foci correspond to PBs, we treated
the roots with cycloheximide in DMSO which is known to abrogate
PB formation in plants (Goeres et al, 2007). This revealed that
cycloheximide inhibited flg22-induced foci (Fig 5A). Importantly,
the control DMSO treatment did not reduce the number of foci
(Supplementary Fig S6A and B). As a control, we also tracked the
localization of the known decapping component VCS (Xu et al,
2006). Similar to PAT1-GFP, VCS-GFP was seen in flg22-induced
PBs and absent when treated with cycloheximide (Fig 5B). Interestingly, PAT1 protein, which is hardly detectable under steady-state
conditions, also accumulated in response to flg22 treatment, with a
peak by 60 min, and a return to normal levels after 2 h (Fig 5C).
PAT1-GFP mirrors this effect and was similarly up-regulated in
response to flg22 treatment, as detected by anti-GFP Western blotting
of seedling protein (Fig 5D). Importantly, PAT1 transcript levels were
not highly induced by flg22 treatment in Col-0 seedlings
(Supplementary Fig S6C), suggesting that PAT1 induction occurs
post-transcriptionally. These data indicate that activation of PTI
leads to up-regulation of PAT1 protein levels by an unknown posttranscriptional mechanism, and this facilitates the detection of PAT1
in PBs. In addition, PTI could induce PB formation as part of cellular
reprogramming, and PAT1 may localize to PBs to engage in this
process.
The MPK4 suppressor summ2 also suppresses the pat1
resistance phenotype
Autoimmunity caused by loss of different components of the MPK4
kinase cascade can be suppressed by mutations in the resistance
protein SUMM2 (Zhang et al, 2012). An explanation for this is that
SUMM2 keeps this PAMP responsive pathway under surveillance
and that mutations in its components mimic the effects of microbial
effectors that prevent phosphorylation within or below the cascade
(Zhang et al, 2007). To further probe the connection between MPK4
and PAT1, we generated pat1-1 summ2-8 double mutants. These
mutants retained the leaf serration phenotype of pat1-1 single
mutants (Supplementary Fig S7A). However, double mutants no
longer accumulated excessive PR1 and PR2 transcripts (Fig 6A) and
displayed summ2 levels of susceptibility to syringe-infiltrated Pto
DC3000 (Fig 6B). Importantly, the pat1 growth phenotype was not
caused by overexpression of SUMM2 transcripts, as the level of
SUMM2 in pat1-1 was similar to wild-type (Supplementary Fig S7B).
To test whether the accumulation of 50 capped transcripts in pat1-1
(Fig 1D) was merely an effect of inappropriate activation of
SUMM2, we tested the accumulation of 50 capped UGT87A2 in an
XRN1 sensitivity assay. XRN1 degrades all uncapped RNA leaving 50
capped RNA intact (Blewett & Goldstrohm, 2012). The accumulation
of 50 capped versus uncapped UGT87A2 transcripts was at similar
levels in pat1-1 and pat1-1 summ2-8 (Fig 6C). Col-0 and summ2-8
plants had similarly reduced 50 capped versus uncapped ratios but,
importantly, these were much lower than the levels of pat1-1 and
pat1-1 summ2-8 (Fig 6C). Therefore, the accumulation of capped
transcripts in pat1-1 mutants is not an artifact of defense activation
but reflects a role for PAT1 in mRNA decay. In order to determine
whether the SUMM2-mediated resistance in pat1 mutants is specific
to pathogens of a specific class, we compared susceptibility to
pathogens with different lifestyles. Pto DC300 is a hemi-biotrophic
47
Published online: January 20, 2015
Milena Edna Roux et al
p < 0.0001
p < 0.0001
p < 0.0001
p < 0.0001
A
4000
Fold change
3500
PR1
PR2
3000
2500
2000
1500
1000
500
0
Col-0 eds1-2 mpk4-2 pat1-1 pat1-1
eds1-2
p=0.0300
B
p=0.0243
p=0.0003
p=0.0006
7
6
Log(cfu/cm2)
The EMBO Journal
mRNA decapping component PAT1 in immunity
5
4
3
2
1
0
Col-0
eds1-2
pat1-1
pat1-1
eds1-2
Figure 4. pat1 mutants display EDS1-dependent, constitutive defense
responses.
A PR gene expression is elevated in mpk4-2 and pat1-1 mutants. Fourweek-old plants were used for RNA extraction, followed by qRT-PCR.
Fold-change in PR1 (gray bars) and PR2 (black) expression is relative to
Col-0, normalized to UBQ10. Standard error of the mean is indicated by
errors bars (n = 3). Statistical significance between the mean values
was determined by ANOVA followed by Fisher’s LSD test, P-values are
only shown for mean values significantly different from Col-0.
B pat1 is more resistant to colonization by P. syringae pv. tomato DC3000.
Bacteria were syringe-infiltrated and samples taken 3 days post-infiltration.
Data are shown as log10-transformed colony-forming units/cm2 leaf tissue
(cfu/cm2). Standard error of the mean is indicated by errors bars (n = 4).
Statistical significance between the mean values was determined by
ANOVA followed by Fisher0 s LSD test.
pathogen, which grows in living tissue during early infection. In
contrast, the necrotrophic fungal pathogen Botrytis cinerea relies on
dying tissue for its propagation in plants (Glazebrook, 2005). Infection of pat1-1 summ2 double mutants with B. cinerea resulted in
similar growth of the fungus as detected in summ2-8 single mutants
(Supplementary Fig S8).
To more accurately address to what extent PAT1 phosphorylation is MPK4 dependent and to measure PAT1 phosphorylation
profiles before and after flg22 treatment, we generated PAT1-HA
transgenic lines in the mkk1/2 summ2-8 background. The
suppression of mpk4 by summ2-8 is only partial, but the autoimmunity phenotype is fully suppressed in mkk1/2 summ2 triple
mutants. We next applied quantitative mass spectrometry (iTRAQ)
to PAT1-HA IPs from Col-0 and mkk1/2-summ2 and this revealed
increased phosphorylation stoichiometry of PAT1 at Ser208 in
untreated Col-0 compared to mkk1/2 summ2 plants (Fig 6D).
Flg22 treatment leads to increased levels of PAT1 Ser208 in both
genotypes, but the levels in Col-0 were higher than what we
found for PAT1 in mkk1/2 summ2 (Fig 6D). Since flg22 treatment augments PAT1 Ser208 phosphorylation in the absence of
MKK1/2, MPK4 might be activated by other upstream kinases or
PAT1 may be phosphorylated by MPK6. Nevertheless, these data
indicate that flg22 treatment leads to increased phosphorylation of
PAT1 at Ser208 and the MKK1/2 MPK4 pathway contributes to
this phosphorylation.
Our data indicate that PAT1 is part of the pathway including
MPK4 and the SUMM2 R protein. SUMM2 and PAT1 might thus
interact in planta. To test this, we transiently expressed and immunoprecipitated PAT1-GFP in N. benthamiana also transiently
expressing MPK4-HA or SUMM2-HA. When PAT1-GFP was pulled
down using GFP Trap beads, we could easily detect MPK4-HA and
SUMM2-HA in the immunoprecipitates (Fig 7A). To verify the association of MPK4, SUMM2 and LSM1, we also pulled down PAT1-HA
and detected MPK4-GFP, SUMM2-GFP and LSM1-GFP in the immunoprecipitates but not Myc-GFP (Supplementary Fig S9). To further
confirm the association of PAT1 and SUMM2, we used bimolecular
fluorescence complementation (BiFC), which produces a fluorescent
readout upon reconstruction of YFP. The BiFC assay showed detectable YFP signal in the cytoplasm when PAT1 and SUMM2 were
co-expressed (Fig 7B). No signals were observed when PAT1 was
expressed with another R protein, At4g12010, implying that the association/reconstruction is specific (Fig 7B, bottom panel). This indicates that PAT1 is in complex with SUMM2 in planta. Thus, MPK4
and PAT1 are both physically and genetically linked to SUMM2.
Discussion
In this study, we identify the Arabidopsis decapping component
PAT1 and show it to be an interactor and substrate of MPK4 in plant
innate immunity. Furthermore, we demonstrate that Arabidopsis
PAT1 complements yeast pat1Δ mutants, indicating that the function of this conserved protein is maintained in plants. The accumulation of capped mRNAs in pat1 mutants is consistent with a role
for PAT1 in the mRNA decapping pathway.
Pat1 mutants exhibit a distinct leaf serration phenotype (Supplementary Fig S4) resembling those of miRNA-loss-of-function
mutants (Nikovics et al, 2006) such as abh1-8 (Gregory et al, 2008),
serrate (Grigg et al, 2005) and hypomorphic alleles of ago1 (Morel,
2002). This suggests that PAT1 may have a role connected to
microRNA activity. Arabidopsis decapping mutants such as dcp1,
dcp2 and vcs accumulate lower levels of certain miRNAs (Motomura
et al, 2012). Mutants with a pat1-like phenotype, such as vcs, suo6
and amp1, have revealed new components in miRNA-mediated
translational repression (Brodersen et al, 2008; Yang et al, 2012; Li
et al, 2013). In Drosophila, HPat interacts with components of the
miRNA machinery including AGO1 and GW182 (Barisic-Jäger et al,
2013). The connection between the mRNA decay activator HPat and
the miRNA effector complex may provide a link to promote the
transition of mRNA from translation to degradation. Although
PAT1 may regulate targets of miRNA-mediated translational
repression, the mechanism by which translational repression occurs
in Arabidopsis is still under investigation.
48
Published online: January 20, 2015
The EMBO Journal
mRNA decapping component PAT1 in immunity
A
Milena Edna Roux et al
B
GFP
Bright field & GFP
GFP
D
0
5 10
20
60
Col-0
VCS-GFP flg22 20 min
VCS-GFP flg22 20 min
+ cycloheximide 60 min
C
pat1-1
PAT1-GFP flg22 20 min
PAT1-GFP flg22 20 min
+ cycloheximide 60 min
VCS-GFP
PAT1-GFP
Bright field & GFP
90 180 H2O
Col-0/PAT-GFP
0
5
10
20
60
90 180
min flg22
anti-PAT1
anti-GFP
CBB
CBB
Figure 5. PAT1-GFP is present in P-bodies after PAMP treatment.
A Confocal microscopy with root elongation zones. Five-day-old Col-0/PAT1-GFP seedlings were treated with 1 lM flg22 on glass slides for 20 min (second panel)
followed by treatment with 100 lg/ml cycloheximide for 60 min (bottom panel). The scale bar corresponds to 10 lm.
B Confocal microscopy with 5-day-old VCS-GFP seedlings treated as in (A). The scale bar corresponds to 10 lm.
C PAT1 protein is induced by PAMP treatment. Immunoblot detection of equal amounts of protein from Col-0 seedlings at times in minutes as indicated following
vacuum infiltration with 1 lM flg22 or water (180 min after infiltration). Immunoblots were probed with anti-PAT1 antibodies. Negative control pat1-1 was loaded
for comparison. Coomassie brilliant blue protein loading control is indicated by CBB.
D Anti-GFP immunoblotting of proteins extracted from Col-0/PAT1-GFP seedlings treated over a time course with flg22 (top panel) and Coomassie brilliant blue (CBB)stained PVDF loading control is shown (bottom panel). No PAT1-GFP was detected in negative control Col-0.
Source data are available online for this figure.
We also find here that PAT1 is an MPK4 substrate and that pat1
mutants exhibit autoimmunity as does mpk4. The connection
between MPK4 and PAT1 is further supported by suppression of the
pat1 constitutive defense phenotype by loss-of-function of the
SUMM2 R protein (Fig 6). However, SUMM2 deficiency only
partially rescues mpk4 mutants (Zhang et al, 2012), thus it is
49
Published online: January 20, 2015
Milena Edna Roux et al
A
The EMBO Journal
mRNA decapping component PAT1 in immunity
2500
PR1
PR2
B
7
6
log (cfu/cm2)
Fold change
2000
1500
1000
500
5
4
3
2
1
0
0
pat1-1
Col-0
C
summ2-8
pat1-1
summ2-8
Col-0
mpk4-2
D
p=0.0012
p=0.0056
80%
60%
40%
20%
0%
Col-0
pat1-1
pat1-1
summ2-8
SSFVSYPPPGSISPDQR – Ser208
summ2-8
Normalized iTRAQ ratiosrelative
to Col PAT1-HA
5’ capped transcripts
p=0.0071
summ2-8
1.8
p=0.0010
100%
pat1-1
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
pat1-1
summ2-8
0
Col-0
PAT1-HA
Col-0
PAT1-HA
flg22
mkk1/2/summ2 mkk1/2/summ2
PAT1-HA
PAT1-HA
flg22
Figure 6. SUMM2 is required for the constitutive defense phenotype of pat1 mutants.
A Elevated PR gene expression in pat1 mutants is suppressed by summ2-8. Four-week-old plants were used for RNA extraction, followed by qRT-PCR. Fold-change in
PR1 (gray bars) and PR2 (black bars) expression is relative to Col-0, normalized to UBQ10. Standard error of the mean is indicated by errors bars (n = 3).
B pat1 resistance to P. syringae pv. tomato DC3000 is suppressed by summ2-8. Bacteria were syringe-infiltrated and samples taken 3 days post-infiltration. Data are
log10-transformed colony-forming units/cm2 leaf tissue (cfu/cm2). Standard error of the mean is indicated by errors bars (n = 4).
C 14-day-old seedlings grown on ms plates were used for RNA extraction. Next, 1 lg of RNA from each sample was treated with XRN1 (NEB) or mock treated before
RT-qPCR. Data were normalized to ACT2, and transcript levels were compared between XRN1 and mock-treated RNA for each genotype. Standard error of the mean is
indicated by error bars (n = 4). Statistical significance between the mean values was determined by ANOVA followed by Fisher’s LSD test.
D Phosphorylation of the PAT1 peptide SSFVSYPPPGSISPDQR, which include Ser208, in Col-0 and mkk1/1-summ2 before and after flg22 treatment. The phosphorylation
stoichiometry is illustrated relative to Col PAT1 without flg22 treatment. The ratios were obtained using quantitative iTRAQ mass spectrometry.
possible that this partial rescue represents a SUMM2 PAT1 branch in
the MPK4 pathway. Nevertheless, the pat1 constitutive defense
phenotype is suppressed by summ2 such that pat1 summ2 mutants
display a wild-type phenotype in response to biotrophic and necrotrophic pathogens (Fig 6; Supplementary Fig S8). Since PAT1 and
SUMM2 also interact in planta (Fig 7), PAT1, or a PAT1-containing
complex, is part of a pathway that includes SUMM2 as well as
MPK4. It is therefore possible that PAT1 is under SUMM2 surveillance because it is an effector target with specific functions in immunity. Since pat1 summ2 mutants are not immunosuppressed, PAT1
and its homologues in Arabidopsis may function redundantly during
PTI. An alternative explanation is that we simply have not yet tested
a pathogen whose infection strategy could reveal a role of PAT1 in
immunity.
mRNA decapping is not the only mRNA regulatory pathway characterized by constitutive defense responses. Indeed, mutants of
nonsense-mediated decay including upf3-1, upf1-5 and smg7 mutants
display autoimmune phenotypes (Jeong et al, 2011; Rayson et al,
2012a; Riehs-Kearnan et al, 2012; Shi et al, 2012). Their phenotypes
are also suppressed by mutations in immune regulators such as eds1
and pad4 (Rayson et al, 2012b; Riehs-Kearnan et al, 2012), similar
to what we find for pat1. This suggests that these components could
also be under surveillance and may be suppressed by mutations in
specific R genes. Recently, it was suggested that nonsense-mediated
decay controls turnover of R gene mRNAs and cause autoimmunity
in smg7 (Gloggnitzer et al, 2014). However, since autoimmunity in
smg7 depends on a specific allele of the R protein RPS6, it is also
possible that SMG7 is under surveillance. Most significantly, this
suggests that plants have developed complex sensors to monitor the
integrity of these pathways, which is consistent with the importance
of differential gene expression in response to pathogen perception.
We detected PAT1 phosphorylation by PAMP-activated MPK4,
and weakly also by MPK6 in vitro. We also identified PAT1 Ser208
and Ser343 as in planta phosphorylation sites by mass spectrometry. Since PAT1 phosphorylation was reduced in vitro when Ser208
was mutated to Ala, this site may be a key target of MPK4 (Fig 2B).
Interestingly, Ser208 is conserved in Physcomitrella patens (moss),
rice, the Arabidopsis PAT1 homologues, and Xenopus PATL2, but
not in human or yeast PAT1 (Supplementary Fig S10). Although
Ser208 corresponds to an SP site in Xenopus PATL1/xPAT1b, PATL1
50
Published online: January 20, 2015
The EMBO Journal
mRNA decapping component PAT1 in immunity
A
B
Merged
100
GFP IP
100
75
HA WB
50
100
75
HA WB
50
PAT1-Nc & SUMM2-Nn
150
PAT1-Nc & SUMM2-Nn
kDa
eYFP
PAT1-Nc & At4g12010-Nn
GFP WB
_
MPK4-HA
SUMM2-HA
PAT1-GFP +
Input
Milena Edna Roux et al
Figure 7. PAT1 is associated with MPK4 and SUMM2 in N. benthamiana.
A Proteins were transiently co-expressed (PAT1-GFP + MPK4-HA or SUMM2-HA) in N. benthamiana and tissue harvested 2 days post-infiltration. Anti-HA immunoblots
of inputs (20 lg each) are shown in the bottom panel. Immunoblots of GFP IPs were probed with anti-HA antibodies, then stripped and probed with anti-GFP
antibodies. Molecular weights are shown in kDa on the right.
B Confocal images of yellow fluorescent protein complementation. PAT1-Nc was transiently co-expressed with either SUMM2-Nn or At4g12010 in N. benthamiana.
Confocal microscopy on leaf disks was conducted 2 days post-infiltration. Merged pictures show overlay of brightfield, chlorophyll and eYFP. The scale bar
corresponds to 25 lm.
Source data are available online for this figure.
is known to be phosphorylated on Ser62 by an unknown kinase
(Marnef et al, 2010). Thus, Ser208 may represent a plant-specific
site or mechanism. While Ser208 is conserved in plant PAT1 orthologs, Ser342/3 is not. As Ser342/3 does not correspond to SP sites,
they may be phosphorylated by kinases other than MPKs. Furthermore, we detected PAT1 phosphorylation in planta irrespective of
PAMP treatment.
PAT1 phosphorylation by MPKs has not been shown in any
system, although several phosphosites have been identified in
human and yeast PAT1 proteins (PhosphoElm, http://phospho.
elm.eu.org/ and Phosida http://www.phosida.com/databases).
However, other decapping complex members are subject to stressinduced MPK-mediated phosphorylation. For example, human
DCP1a is phosphorylated by c-Jun N-terminal kinase in response to
stress (Rzeczkowski et al, 2011). Similarly, Ste20 phosphorylates
yeast DCP2 upon glucose deprivation (Yoon et al, 2010). In both
cases, phosphorylation seems only to be required for P-body formation and not for general decapping (Yoon et al, 2010; Rzeczkowski
et al, 2011). Arabidopsis MPK6 was shown to specifically phosphorylate DCP1 in plants during dehydration stress (Xu & Chua, 2012).
Thus, the regulation of mRNA decay machinery by MPKs during
stress responses seems to be a key mechanism in plants and other
organisms, although exactly how this affects mRNA turnover
remains elusive.
Materials and Methods
Plant materials and growth conditions
Arabidopsis thaliana ecotype Columbia (Col-0) was used as a control.
Seeds for T-DNA insertion lines were from NASC (Nottingham, UK).
51
Published online: January 20, 2015
Milena Edna Roux et al
mRNA decapping component PAT1 in immunity
The T-DNA lines for At1g79090 (PAT1), both of which have insertions in the last exon, were Salk_040660 (here named pat1-1) and
WiscDsLox437D04 (pat1-2). The T-DNA insertion in At1g12280
(SUMM2) summ2-8 (SAIL_1152A06), mkk1/2 (Zhang et al, 2012)
and eds1-2 (Parker et al, 1996) have been described. Genotyping
primers for newly described T-DNA lines are provided in
Supplementary Table S1. Arabidopsis plants were grown in 9 × 9 cm
pots at 22°C with a 8-h photoperiod, or on plates containing
Murashige–Skoog (MS) salts medium (Duchefa), 1% sucrose and 1%
agar with a 16-h photoperiod.
Cloning and transgenic lines
The genomic PAT1 (At1g79090) DNA sequence (without stop
codon), plus 2 kb upstream from the start codon, was amplified
from Col-0 genomic DNA and cloned into pENTR-D-TOPO (Invitrogen). The entry clone was recombined into pGWB513, pGWB517
and pGWB504 (Nakagawa et al, 2007) to obtain a C-terminal HA,
Myc and GFP tags, respectively. The expression clones were transformed into Agrobacterium tumefaciens strain GV3101 and transformed into Col-0 plants by floral dipping. Transformants were
selected on hygromycin (30 lg/ml) MS agar and survivors tested
for protein expression by Western blotting.
The VCS promoter region was amplified and inserted into pGEM-T
easy and next into the HindIII SalI in pBN:GFP to generate pBNpVCS:
GFP. The VCS coding region was amplified and cloned into pGEM-T
easy next inserted into SalI and KpnI in pBNpVCS:GFP and transformed into Agrobacterium LBA4404 and transformed into Col-0
plants by floral dipping. The cDNA of LSM1 (At3g14080) without
stop codon was cloned into pENTR-D-TOPO (Invitrogen) and recombined into 35S promoter-containing pGWB505 (Nakagawa et al,
2007) to obtain C-terminal GFP-tagged LSM1. Genomic DNA of
MPK4 and cDNA of SUMM2 were cloned into pENTR-D-TOPO
and recombined into pGWB514 (Nakagawa et al, 2007) to obtain
C-terminal HA-tagged constructs. The genomic DNA and promoter of
DCP1 was cloned into pENTR-D-TOPO and recombined into
pGWB513 to obtain C-terminal HA-tagged constructs under the
control of the native promoter. Clones were transformed into
A. tumefaciens strain GV3101 and used for transient expression in
N. benthamiana or floral dipping in Arabidopsis. Genomic PAT1GFP was transformed into the Col-0 background. DCP1-HA was
dipped into the Col-0 background and T3 progeny were crossed with
Col-0/PAT1-GFP T4 lines to obtain double transgenic lines. Genomic
PAT1-Myc was transformed into Ler mpk4-1/MPK4-HA transgenic
lines (described in Petersen et al, 2000) to obtain double transgenic
lines.
The entry clones of PAT1 and SUMM2 were further recombined
into pcYGW and pnYGW (Hino et al, 2011), respectively, obtaining
the N- and C-terminal split YFP tags used for BiFC. At4g12010
clones were obtained by the same procedure. Clones were transformed into A. tumefaciens strain GV3101 and used for transient
expression in N. benthamiana.
Mutagenesis of His6-PAT1
Site-directed mutagenesis of Ser208 to Ala was accomplished using
pET15b-PAT1 as a template for PCR mutagenesis. The primers used
were PAT1 S208A F&R (see Supplementary Table S1).
The EMBO Journal
PAT1 protein purification and in vitro kinase assays
For in vitro experiments, PAT1 protein was purified from E. coli.
The PAT1 cDNA was cloned into pET15b (for an N-terminal His
fusion) and transformed into E. coli BL21 (pLysS). Protein expression was induced by overnight treatment with 0.5 mM IPTG at
18°C, added to cells at OD600 = 0.6. PAT1 protein was insoluble
and thus purified from inclusion bodies using Bugbuster (Novagen). Proteins were solubilized in 6 M urea, 0.7% N-lauroylsarcosine, 100 mM Tris-HCl pH 8 and refolded overnight at 4°C in
0.88 M L-arginine, 55 mM Tris-HCl, 2 mM NaCl, 0.88 mM KCl,
protease inhibitors. Protein was then dialyzed against 20 mM TrisHCl pH 8, 100 mM NaCl using 3,500 MWCO dialysis tubing. Purified protein was concentrated using Centriprep 30K spin columns
(Millipore) and then diluted with glycerol to a final concentration
of 0.1 mg/ml.
For kinase assays, Col-0 plants were immersed in 200 nM flg22
for 10 min of to activate MAP kinases. MPK3, 4 and 6 were immunoprecipitated from 3 mg total protein extracted from flg22-treated
tissue using 2 lg of each of their specific antibodies (Sigma) and
30 ll EZview protein A agarose beads (Sigma). Four microgram of
purified myelin basic protein (MBP, Sigma) or 20 lg His6-PAT1
protein was incubated with washed MPK immunoprecipitates for
60 min at 37°C with 3 lCi c-ATP in kinase buffer (62.5 lM ATP,
100 mM Tris pH 7.5, 150 mM NaCl, 150 mM MgCl2, 10 mM EGTA,
5 mM DTT, Phosstop inhibitor (Roche)). Kinase reactions were
diluted with 4× SDS buffer and boiled for 5 min before loading on
12% SDS-PAGE gels. Following electrophoresis, gels were stained
with Coomassie Brilliant Blue and incubated with gel drying buffer,
followed by drying on a Bio-Rad gel dryer. Dried gels were exposed
to a phosphor screen overnight.
Yeast transformation
Yeast strains pat1Δ (BY4742 (YCR077c) pat1D::KanMX) and B4742
were obtained from Euroscarf. PAT1 was cloned from A. thaliana
Col-0 cDNA into pENTR-D-TOPO and recombined into yeast expression vector pVV215 (C-terminal HA tag, -URA selection; (Van
Mullem et al, 2003) by Gateway recombination. Yeast pat1Δ and
B4742 cells were transformed using lithium acetate/polyethylene
glycol according to the Clontech Yeast Protocols handbook.
Transformed yeast was selected on SD-URA agar plates and restreaked after 3–4 days onto fresh selection plates. Pat1D and wildtype-transformed yeast was grown in liquid culture overnight, and
protein was extracted from pelleted cells by vortexing with glass
beads in 1× SDS extraction buffer. Boiled proteins were subjected to
SDS-PAGE and anti-PAT1 immunoblotting. For temperature sensitivity assays, overnight cultures of wild-type, pat1D and PAT1expressing transformants were plated on YPAD and grown at 30 and
37°C for 3 days.
Flg22 kinetics
Seedlings were grown on MS agar (Col-0) or MS agar containing
30 lg/ml hygromycin (Col-0/PAT1-GFP) for 5 days before being
transferred to MS liquid medium in 24-well plates. After 10 days,
seedlings were treated by the addition of flg22 to a final concentration of 100 nM (2 seedlings per well × 3 wells per treatment time
52
Published online: January 20, 2015
The EMBO Journal
point). Seedlings were harvested at the indicated times and immediately frozen in liquid nitrogen for later RNA or protein extraction.
Semi-quantitative and qRT-PCR
Total RNA was extracted from seedlings with Tri Reagent (Sigma).
RNA samples were treated with DNase Turbo DNA-free (Ambion),
quantified with a NanoDrop spectrophotometer (Thermo Scientific)
and reverse-transcribed into cDNA with SuperScript III reverse
transcriptase (Invitrogen). For semi-quantitative reverse-transcription
PCR (RT-PCR), Col-0 seedling cDNA was used as a template for PCR
with primers specific for ACTIN8, PAT1, PAT1H1 or PAT1H2 using
Sigma Jumpstart REDTaq Readymix. PCR products (20 ll) were
separated on 2% (w/v) agarose gels and visualized with ethidium
bromide. Brilliant II SybrGreen master mix (Agilent) was used for
qPCRs. The UBQ10 (At4g05320) gene was used for normalization.
Gene expression of PR1, PR2, PAT1 and SUMM2 was measured by
qPCR analysis, normalized to UBQ10 expression and plotted relative
to Col-0 expression level. These experiments were repeated in three
independent biological replicates, each with three technical replicates, and representative data are shown. Standard error is represented by error bars on figures, and statistical significance is
indicated by letters above error bars. These are derived from oneway ANOVA with Tukey’s multiple comparison test (GraphPad
Prism).
mRNA decapping component PAT1 in immunity
Milena Edna Roux et al
Infection assays
Pseudomonas syringae pv. tomato DC3000 (Pto DC3000) strains
were grown in overnight culture in Kings B medium supplemented
with appropriate antibiotics. Cells were harvested by centrifugation
and pellets resuspended in sterile water to OD600 = 0.002. Bacteria
were infiltrated with a needleless 1-mm syringe into four leaves on
four plants per genotype and maintained in growth chambers for
3 days. Samples were taken using a cork-borer (6.5 mm) to cut leaf
disks from four leaves per plant and four plants per genotype. Leaf
disks were ground in water, serially diluted and plated on Kings B
with appropriate selection. Plates were incubated at 28°C and colonies counted 2–3 days later. These experiments were repeated in
three independent biological replicates, and representative data are
shown.
For drop inoculation of B. cinerea (strain B05.10), 10 ll of
2.5 × 105 spores/ml in Gamborg B5/2% sucrose (pH 5.6) was
placed on the adaxial surface of fully expanded leaves of 4-week-old
plants and sampled by harvesting 10–15 leaf disks per genotype
were at each time point, namely 0, 2 and 3 days. Data are the average of 3 biological replicates. Disease severity was measured as
accumulation of the B. cinerea CUTINASE A transcript by qPCR relative to A. thaliana a-SHAGGY KINASE (At5g26751) (primers are
according to Gachon & Saindrenan, 2004).
Transient expression in Nicotiana benthamiana
RNA extraction and RACE PCR
RNA extraction used TRI reagent and 10 lg RNA was used for
50 RACE according to instructions (First Choice RACE, Ambion).
PCR was carried out on 1 ll of products from reverse transcription
of capped RNA using DreamTaq polymerase (Fermentas) with 25
cycles. RACE PCR products (10 ll) were separated on 2% (w/v)
agarose gels and visualized with ethidium bromide.
Quantification of capped versus uncapped transcripts
Total RNA was extracted from seedlings with NucleoSpin RNA
columns (Machery-Nagel). To remove RNA with no 50 cap structure,
1 lg of total RNA was incubated with 1 unit XRN1 (New England
Biolabs) or no enzyme at 37°C for 1 h. Next RNA was reverse-transcribed into cDNA with SuperScript III reverse transcriptase (Invitrogen). UGT87A2 transcript accumulation was measured by qPCR
using SybrGreen master mix (Agilent) and normalized to ACT2.
Calculating 50 capped versus uncapped transcripts was done by
comparing transcript levels from XRN1 and mock-treated samples
for the individual genotypes.
Confocal microscopy
Col-0/PAT1-GFP plants were grown on MS agar containing 30 lg/ml
hygromycin for 5 days. Seedlings were placed on glass microscope
slides with water, 1% DMSO or 100 nM flg22 for 20 min. For following cycloheximide and DMSO treatment, seedlings were removed
from flg22 and placed on new glass microscope slides. Cycloheximide was used at 100 lg/ml in a 1% DMSO dilution, and the control
was done with 1% DMSO. Imaging was done using a Leica SP5
inverted microscope.
Agrobacterium tumefaciens strains were grown in LB medium
supplemented with appropriate antibiotics overnight. Cultures were
spun down and resuspended in 10 mM MgCl2 to OD600 = 0.8. A.
tumefaciens strains carrying PAT1-GFP and MPK4-HA or SUMM2HA were mixed 1:1 and syringe-infiltrated into 3-week-old N. benthamiana leaves. Samples for protein extraction were harvested
2 days post-infiltration (dpi). Agrobacterium tumefaciens strains
carrying either PAT1-HA, LSM1-GFP or PAT1-HA + LSM1-GFP were
mixed 1:1 and syringe-infiltrated into 3-week-old N. benthamiana
leaves. Samples for protein extraction were harvested 3 dpi. For
BiFC, A. tumefaciens strains carrying PAT1 fused to the N-terminal
part of YFP and SUMM2 or At4g12010 fused to the C-terminal part
were mixed 1:1 and syringe-infiltrated into N. benthamiana leaves
at an final OD600 = 0.8. Confocal microscopy on leaf disks was
conducted 2 days post-infiltration under a Leica SP5 inverted microscope.
Protein extraction and immunoprecipitation in
Nicotiana benthamiana
Leaves were ground in liquid nitrogen and extraction buffer [50 mM
Tris-HCl pH 7.5; 150 mM NaCl; 10% (v/v) glycerol; 10 mM DTT;
10 mM EDTA; 1% (w/v) PVP; protease inhibitor cocktail (Roche);
0.1% (v/v) IGEPAL CA-630 (Sigma); Phosstop (Roche) added at
2 ml/g tissue powder. Samples were clarified by 20 min centrifugation at 4°C and 13,000 rpm. Supernatants (1.5 ml) were adjusted to
2 mg/ml protein and incubated 2 h at 4°C with 20 ll GFPTrap-A
beads (Chromotek) or anti-HA antibodies (2 lg, Santa Cruz) and
30 ll EZview protein A agarose beads (Sigma). Following incubation, beads were washed four times with IP buffer, before adding
30 ll 2× SDS and heating at 95°C for 5 min.
53
Published online: January 20, 2015
Milena Edna Roux et al
The EMBO Journal
mRNA decapping component PAT1 in immunity
Arabidopsis protein extraction and immunoprecipitation for mass
spectrometry analysis
Leaves from adult plants were ground in liquid nitrogen and extraction buffer [50 mM Tris-HCl pH 7.5; 150 mM NaCl; 10% (v/v) glycerol; 5 mM DTT; 2 mM EDTA; protease inhibitor cocktail (Roche);
0,1% (v/v) IGEPAL CA-630 (Sigma) and Phosstop (Roche) added at
2 ml/g tissue powder. Samples were clarified by 20 min centrifugation at 4°C 13,000 rpm. Supernatants (45 ml) were adjusted to
3 mg/ml protein and incubated 4 h at 4°C with 50 ll GFPTrap-A
beads (Chromotek) or anti-HA antibodies (4 lg, Santa Cruz) and
100 ll EZview protein A agarose beads (Sigma). Following incubation, beads were washed four times with IP buffer before adding
2× SDS and heating to 95°C for 5 min.
SDS-PAGE and immunoblotting
SDS-PAGE gels were prepared with either 8, 10 or 12% cross-linking.
Proteins were loaded and gels run at 100–150 V for 1.5 h before electroblotting onto PVDF membrane (GE Healthcare). Membranes were
rinsed in TBS and blocked for 1 h in 5% (w/v) non-fat milk in TBSTween (0.1% (v/v)). Antibodies were diluted in blocking solution to
the following dilutions: anti-GFP (AMS Biotechnology (rabbit)
1:5,000 or Santa Cruz (mouse) 1:1,000), anti-HA 1:1,000 (Santa
Cruz). Membranes were incubated with primary antibodies for 1 h
to overnight. Membranes were washed 3 × 10 min in TBS-T (0.1%)
before 1-h incubation in secondary antibodies, anti-rabbit or antimouse-HRP or anti-rabbit or anti-mouse AP conjugate (Promega; 1:
5,000). Chemiluminescent substrate (ECL Plus, Pierce) was applied
before exposure to film (AGFA CP-BU). For AP-conjugated primary
antibodies, membranes were incubated in NBT/BCIP (Roche) until
bands were visible. For probing immunoblots with multiple antibodies, stripping was carried out using Restore Western Blot Stripping
Buffer (Pierce) for 15–30 min, following by three washes with TBSTween.
Antibodies
Polyclonal anti-PAT1 antibodies were generated by immunizing
rabbits with synthetic peptides derived from the N-terminus
[EQRIPDRTKLYPEPQ] and C-terminus [KRSMLGSQKTEPVLS] of
PAT1. Antibodies (final bleed) were affinity-purified against the
C-terminal peptide (Eurogentec). Antibody specificity was verified
by immunoblotting with plant extracts derived from Col-0 and pat1-1
tissue. Mouse anti-HA and anti-GFP antibodies were obtained from
Santa Cruz. Rabbit anti-GFP antibodies were obtained from AMS
Biotechnology. Anti-MPK3, anti-MPK-4 and anti-MPK-6 antibodies
were obtained from Sigma. Secondary antibodies were obtained
from Promega.
gel pieces. The solution was incubated in 4°C for 30 min. After
incubation, the trypsin solution was replaced with 30 ll 20 mM
TEAB pH 7.5 and the tube was incubated at 37°C overnight.
Peptides from the digestion solution after incubation were recovered
in a low binding Eppendorf tube (Sorensen Bioscience), and the gel
pieces washed with 50% acetonitrile for 15 min to extract more
peptides. The washing solution was mixed with the recovered
peptides, and the peptide solution was lyophilized. Phosphopeptides
were purified by titanium dioxide (TiO2) chromatography (Larsen
et al, 2005). Lyophilized peptides were redissolved in loading buffer
for TiO2 chromatography (80% acetonitrile, 5% TFA, 1 M glycolic
acid), and 0.3 mg TiO2 beads (GL Science, Japan) were added to the
solution and incubated for 10 min. After incubation, the beads were
pelleted by centrifugation and the supernatant was removed. The
beads were washed once with 80% acetonitrile, 1% TFA and once
with 10% acetonitrile, 0.1% TFA. Phosphopeptides were eluted
using 1% ammonium hydroxide and desalted and concentrated prior
to LC-MS/MS using a Poros Oligo R3 micro-column (Engholm-Keller
et al, 2012).
For quantitative phosphopeptide analysis, the tryptic peptide
mixtures after in-gel digestion were labeled with iTRAQ 4 plex
(according to the manufacturer’s protocol) and the samples were
mixed prior to TiO2 enrichment as described above. The nonphosphorylated peptides were concentrated and desalted using a
Poros Oligo R3 micro-column (Engholm-Keller et al, 2012). The
non-phosphorylated peptides were used to normalize the PAT1
protein level.
LC-MS/MS analysis was performed using an EASY-LC system
(Proxeon, Thermo Fischer Scientific) coupled to an LTQ-Orbitrap XL
mass spectrometer (Thermo Fisher Scientific) as described previously (Engholm-Keller et al, 2012). Peptides were separated using a
20 min gradient from 0–34% B-buffer (A-buffer: 0.1% formic acid;
B-buffer: 90% acetonitrile, 0.1% formic acid). The data-dependent
analysis was performed using one MS full scan in the area 300–
1,800 Da performed in the Orbitrap with 60,000 in resolution,
followed by the five most intense ions selected for MSMS (collision
induced dissociation) performed in the linear ion trap.
Raw data from the LTQ-Orbitrap-XL were processed using the
Proteome Discoverer (PD) program (Thermo Fisher Scientific,
Bremen, Germany). The data were searched against the Arabidopsis
database (33,596 sequences; 13,487,687 residues) using the Mascot
2.2 version database. The parameters for the database search were
as follows: Enzyme—trypsin; missed cleavages—2; MS mass accuracy—10 ppm; MSMS mass accuracy—0.8 Da; Variable modifications—Phosphorylation (S, T, Y), Oxidation (met), Propionamide
(C). The identified peptides were filtered for 1% false discovery
rate using the ‘Fixed Value PSM Validator’ in PD. All identified
phosphopeptides were manually validated. For iTRAQ-labeled phosphopeptides, the PD quantitation node was used and the data were
normalized based on the non-phosphorylated peptides.
In-gel digestion, TiO2 chromatography and mass spectrometry
Supplementary information for this article is available online:
Bands excised from SDS-PAGE were chopped into small pieces and
incubated for 1 h in 50 mM TEAB, acetonitrile (50:50) on a shaker
at room temperature. After incubation, the sample was centrifuged
shortly and the supernatant was discarded. Gel pieces were dried in
a vacuum centrifuge for 15 min and subsequently 30 ll of trypsin
(10 ng/ll) in 20 mM TEAB pH 7.5 was added to cover the dried
http://emboj.embopress.org
Acknowledgements
We thank Yuelin Zhang for providing summ2-8 and mkk1/2 summ2-8 seeds.
This work was supported by the Danish Research Council for independent
Research, Natural Sciences Grant #10-084139 (M.P.) & Postdoctoral Fellowship
54
Published online: January 20, 2015
The EMBO Journal
#11-116368 (M.E.R.) & Technology and Production Sciences #11-106302. KP
mRNA decapping component PAT1 in immunity
Milena Edna Roux et al
Brader G, Djamei A, Teige M, Palva ET, Hirt H (2007) The MAP kinase kinase
was supported by an EMBO Long-Term Fellowship and a Marie Curie
MKK2 affects disease resistance in Arabidopsis. Mol Plant Microbe Interact
International Incoming Fellowship. We thank Suksawad Vonvisuttikun for
20: 589 – 596
technical help and Dr. Tsuyoshi Nakagawa (Shimane University) for providing
Gateway binary pGWB vectors. All confocal work was done at Center for
Advanced Bioimaging. MRL was supported by the Lundbeck Foundation (Junior
Brengues M (2005) Movement of eukaryotic mRNAs between polysomes and
cytoplasmic processing bodies. Science 310: 486 – 489
Brodersen P, Petersen M, Nielsen H, Zhu S, Newman M-A, Shokat KM, Rietz S,
Group Leader Fellowship).
Parker J, Mundy J (2006) Arabidopsis MAP kinase 4 regulates salicylic acid-
Author contributions
Plant J 47: 532 – 546
and jasmonic acid/ethylene-dependent responses via EDS1 and PAD4.
MER, MWR, KP, JM and MP conceived and designed the experiments. MER,
MWR KP, SL, MRL, JG, AMR, LS, WZ and GB performed experiments. MER,
MWR, JM and MP analyzed the data. MER, MWR, JM and MP wrote the
manuscript.
Brodersen P, Sakcarelidze-Achard L, Bruun-Rasmussen M, Dunoyer P,
Yamamoto YY, Sieburth L, Voinnet O (2008) Widespread translational
inhibition by plant miRNAs and siRNAs. Science 320: 1185 – 1190
Buchan JR, Yoon J-H, Parker R (2011) Stress-specific composition, assembly
and kinetics of stress granules in Saccharomyces cerevisiae. J Cell Sci 124:
Conflict of interest
The authors declare that they have no conflicts of interest.
228 – 239
Chinchilla D, Bauer Z, Regenass M, Boller T, Felix G (2006) The Arabidopsis
receptor kinase FLS2 binds flg22 and determines the specificity of flagellin
perception. Plant Cell 18: 465 – 476
References
Chisholm ST, Coaker G, Day B, Staskawicz BJ (2006) Host-microbe interactions:
Andreasson E, Jenkins T, Brodersen P, Thorgrimsen S, Petersen NH, Zhu S,
Coller J, Parker R (2005) General translational repression by activators of
shaping the evolution of the plant immune response. Cell 124: 803 – 814
Qiu J-L, Micheelsen P, Rocher A, Petersen M, Newman M-A, Nielsen HB,
Hirt H, Somssich IE, Mattsson O, Mundy J (2005) The MAP kinase
substrate MKS1 is a regulator of plant defense responses. EMBO J 24:
2579 – 2589
Andreasson E, Ellis B (2010) Convergence and specificity in the Arabidopsis
MAPK nexus. Trends Plant Sci 15: 106 – 113
Asai T, Tena G, Plotnikova J, Willmann MR, Chiu WL, Gómez-Gómez L, Boller
T, Ausubel FM, Sheen J (2002) MAP kinase signalling cascade in
Arabidopsis innate immunity. Nature 415: 977 – 983
Balagopal V, Parker R (2009) Polysomes, P bodies and stress granules: states
and fates of eukaryotic mRNAs. Curr Opin Cell Biol 21: 403 – 408
Barisic-Jäger E, Krez cioch I, Hosiner S, Antic S, Dorner S (2013) HPat a
decapping activator interacting with the miRNA effector complex. PLoS
One 8: e71860
Bethke G, Unthan T, Uhrig JF, Poeschl Y, Gust AA, Scheel D, Lee J (2009)
Flg22 regulates the release of an ethylene response factor substrate from
MAP kinase 6 in Arabidopsis thaliana via ethylene signaling. PNAS 106:
674
Bethke G, Pecher P, Eschen-Lippold L, Tsuda K, Katagiri F, Glazebrook J, Scheel
D, Lee J (2012) Activation of the Arabidopsis thaliana mitogen-activated
protein kinase MPK11 by the flagellin-derived elicitor peptide, flg22. Mol
Plant Microbe Interact 25: 471 – 480
Blewett NH, Goldstrohm AC (2012) A eukaryotic translation initiation factor
4E-binding protein promotes mRNA decapping and is required for PUF
repression. Mol Cell Biol 32: 4181 – 4194
Boag PR, Atalay A, Robida S, Reinke V, Blackwell TK (2008) Protection of
specific maternal messenger RNAs by the P body protein CGH-1 (Dhh1/
RCK) during Caenorhabditis elegans oogenesis. J Cell Biol 182: 543 – 557
Boller T, Felix G (2009) A renaissance of elicitors: perception of
microbe-associated molecular patterns and danger signals by
pattern-recognition receptors. Annu Rev Plant Biol 60: 379 – 406
Bonnerot C, Boeck R, Lapeyre B (2000) The two proteins Pat1p (Mrt1p) and
Spb8p interact in vivo, are required for mRNA decay, and are functionally
linked to Pab1p. Mol Cell Biol 20: 5939 – 5946
mRNA decapping. Cell 122: 875 – 886
Deyholos MK, Cavaness GF, Hall B, King E, Punwani J, van Norman J, Sieburth
L (2003) VARICOSE, a WD-domain protein, is required for leaf blade
development. Development 130: 6577 – 6588
Droillard M-J, Boudsocq M, Barbier-Brygoo H, Laurière C (2004)
Involvement of MPK4 in osmotic stress response pathways in cell
suspensions and plantlets of Arabidopsis thaliana: activation by
hypoosmolarity and negative role in hyperosmolarity tolerance. FEBS
Lett 574: 42 – 48
Engholm-Keller K, Birck P, Størling J, Pociot F, Mandrup-Poulsen T, Larsen MR
(2012) TiSH–a robust and sensitive global phosphoproteomics strategy
employing a combination of TiO2, SIMAC, and HILIC. J Proteomics 75:
5749 – 5761
Gachon C, Saindrenan P (2004) Real-time PCR monitoring of fungal
development in Arabidopsis thaliana infected by Alternaria brassicicola and
Botrytis cinerea. Plant Physiol Biochem 42: 367 – 371
Gao M, Liu J, Bi D, Zhang Z, Cheng F, Chen S, Zhang Y (2008) MEKK1,
MKK1/MKK2 and MPK4 function together in a mitogen-activated protein
kinase cascade to regulate innate immunity in plants. Cell Res 18:
1190 – 1198
Garneau NL, Wilusz J, Wilusz CJ (2007) The highways and byways of mRNA
decay. Nat Rev Mol Cell Biol 8: 113 – 126
Glazebrook J (2005) Contrasting mechanisms of defense against biotrophic
and necrotrophic pathogens. Annu Rev Phytopathol 43: 205 – 227
Gloggnitzer J, Akimcheva S, Srinivasan A, Kusenda B, Riehs N, Stampfl H,
Bautor J, Dekrout B, Jonak C, Jimenez-Gomez JM, Parker JE, Riha K (2014)
Nonsense-mediated mRNA decay modulates immune receptor levels to
regulate plant antibacterial defense. Cell Host Microbe 16: 376 – 390
Goeres D, Van Norman JM, Zhang W, Fauver NA, Spencer ML, Sieburth L
(2007) Components of the Arabidopsis mRNA decapping complex are
required for early seedling development. Plant Cell 19: 1549 – 1564
Golisz A, Sikorski PJ, Kruszka K, Kufel J (2013) Arabidopsis thaliana LSM
proteins function in mRNA splicing and degradation. Nucleic Acids Res 41:
6232 – 6249
Bouveret E, Rigaut G, Shevchenko A, Wilm M, Seraphin B (2000) A Sm-like
Gomez-Gomez L, Bauer Z, Boller T (2001) Both the extracellular leucine-rich
protein complex that participates in mRNA degradation. EMBO J 19:
repeat domain and the kinase activity of FLS2 are required for flagellin
1661 – 1671
binding and signaling in Arabidopsis. Plant Cell 13: 1155 – 1163
55
Published online: January 20, 2015
Milena Edna Roux et al
The EMBO Journal
mRNA decapping component PAT1 in immunity
Gregory BD, Lister R, OMalley R, Urich MA, Tonti-Phillipini J, Chen H, Millar
AH, Ecker JR (2008) A link between RNA metabolism and silencing
affecting Arabidopsis development. Dev Cell 14: 854 – 866
Grigg SP, Canales C, Hay A, Tsiantis M (2005) SERRATE coordinates shoot
meristem function and leaf axial patterning in Arabidopsis. Nature 437:
1022 – 1026
Haas G, Braun JE, Tritschler F, Nishihara T, Izaurralde E (2010) HPat provides
Nikovics K, Blein T, Peaucelle A, Ishida T, Morin H, Aida M, Laufs P (2006) The
balance between the MIR164A and CUC2 genes controls leaf margin
serration in Arabidopsis. Plant Cell 18: 2929 – 2945
Nissan T, Rajyaguru P, She M, Song H, Parker R (2010) Decapping activators
in Saccharomyces cerevisiae act by multiple mechanisms. Mol Cell 39:
773 – 783
Olmedo G, Guo H, Gregory BD, Nourizadeh SD, Aguilar-Henonin L, Li H, An F,
a link between deadenylation and decapping in metazoa. J Cell Biol 189:
Guzman P, Ecker JR (2006) ETHYLENE-INSENSITIVE5 encodes a 5 ‘-> 3 ‘
289 – 302
exoribonuclease required for regulation of the EIN3-targeting F-box
Hatfield L, Beelman CA, Stevens A, Parker R (1996) Mutations in trans-acting
factors affecting mRNA decapping in Saccharomyces cerevisiae. Mol Cell
Biol 16: 5830 – 5838
Hino T, Tanaka Y, Kawamukai M, Nishimura K, Mano S, Nakagawa T (2011)
Two Sec13p homologs, AtSec13A and AtSec13B, redundantly contribute to
proteins EBF1/2. PNAS 103: 13286 – 13293
Ozgur S, Chekulaeva M, Stoecklin G (2010) Human Pat1b connects
deadenylation with mRNA decapping and controls the assembly of
processing bodies. Mol Cell Biol 30: 4308 – 4323
Park S-H, Chung PJ, Juntawong P, Bailey-Serres J, Kim YS, Jung H, Bang SW,
the formation of COPII transport vesicles in Arabidopsis thaliana. Biosci
Kim Y-K, Do Choi Y, Kim J-K (2012) Posttranscriptional control of
Biotechnol Biochem 75: 1848 – 1852
photosynthetic mRNA decay under stress conditions requires 30 and 50
Ichimura K, Casais C, Peck SC, Shinozaki K, Shirasu K (2006) MEKK1 is required for
MPK4 activation and regulates tissue-specific and temperature-dependent
cell death in Arabidopsis. J Biol Chem 281: 36969 – 36976
Jeong H-J, Kim YJ, Kim SH, Kim Y-H, Lee I-J, Kim YK (2011)
Nonsense-mediated mRNA decay factors, UPF1 and UPF3, contribute to
plant defense. Plant Cell Physiol 52: 2147 – 2156
Jiao X, Xiang S, Oh C, Martin CE, Tong L, Kiledjan M (2010) Identification of a
quality-control mechanism for mRNA 50 -end capping. Nature 467:
608 – 611
Jones JDG, Dangl JL (2006) The plant immune system. Nature 444: 323 – 329
Kulkarni M, Ozgur S, Stoecklin G (2010) On track with P-bodies. Biochem Soc
Trans 38: 242 – 251
Larsen MR, Thingholm TE, Jensen ON, Roepstorff P, Jørgensen TJD (2005)
Highly selective enrichment of phosphorylated peptides from peptide
mixtures using titanium dioxide microcolumns. Mol Cell Proteomics 4:
873 – 886
Li S, Liu L, Zhuang X, Yu Y, Liu X, Cui X, Ji L, Pan Z, Cao X, Mo B, Zhang F,
untranslated regions and correlates with differential polysome association
in rice. Plant Physiol 159: 1111 – 1124
Parker JE, Holub EB, Frost LN, Falk A, Gunn ND, Daniels MJ (1996)
Characterization of eds1, a mutation in Arabidopsis suppressing resistance
to Peronospora parasitica specified by several different RPP genes. Plant
Cell 8: 2033 – 2046
Parker R, Sheth U (2007) P bodies and the control of mRNA translation and
degradation. Mol Cell 25: 635 – 646
Pearson G, Robinson F, Gibson TB, Xu BE, Karandikar M, Berman K, Cobb MH
(2001) Mitogen-activated protein (MAP) kinase pathways: regulation and
physiological functions. Endocr Rev 22: 153 – 183
Perea-Resa C, Hernández-Verdeja T, López-Cobollo R, del Mar Castellano M,
Salinas J (2012) LSM proteins provide accurate splicing and decay of
selected transcripts to ensure normal Arabidopsis development. Plant Cell
24: 4930 – 4947
Petersen M, Brodersen P, Naested H, Andreasson E, Lindhart U, Johansen B,
Nielsen HB, Lacy M, Austin MJ, Parker JE, Sharma SB, Klessig DF, Martienssen
Raikhel N, Jiang L, Chen X (2013) MicroRNAs inhibit the translation of
R, Mattsson O, Jensen AB, Mundy J (2000) Arabidopsis MAP kinase 4
target mRNAs on the endoplasmic reticulum in Arabidopsis. Cell 153:
negatively regulates systemic acquired resistance. Cell 103: 1111 – 1120
562 – 574
Marnef A, Maldonado M, Buguaut A, Balasubramanian S, Kress M, Weil D,
Standart N (2010) Distinct functions of maternal and somatic Pat1
protein paralogs. RNA 16: 2094 – 2107
Marnef A, Standart N (2010) Pat1 proteins: a life in translation, translation
repression and mRNA decay. Biochem Soc Trans 38: 1602 – 1607
Mészáros T, Helfer A, Hatzimasoura E, Magyar Z, Serazetdinova L, Rios G,
Pitzschke A, Schikora A, Hirt H (2009) MAPK cascade signalling networks in
plant defence. Curr Opin Plant Biol 12: 421 – 426
Potuschak T, Vansiri A, Binder BM, Vierstra RD, Genschik P (2006) The
exoribonuclease XRN4 is a component of the ethylene response pathway
in Arabidopsis. Plant Cell 18: 3047 – 3057
Qiu J-L, Fiil BK, Petersen K, Nielsen HB, Botanga CJ, Thorgrimsen S, Palma K,
Suarez-Rodriguez MC, Sandbech-Clausen S, Lichota J, Brodersen P, Grasser
Bardóczy V, Teige M, Koncz C, Peck S, Bögre L (2006) The Arabidopsis MAP
KD, Mattsson O, Glazebrook J, Mundy J, Petersen M (2008a) Arabidopsis
kinase kinase MKK1 participates in defence responses to the bacterial
MAP kinase 4 regulates gene expression through transcription factor
elicitor flagellin. Plant J 48: 485 – 498
release in the nucleus. EMBO J 27: 2214 – 2221
Morel JB (2002) Fertile hypomorphic ARGONAUTE (ago1) mutants impaired in
Qiu J-L, Zhou L, Yun B-W, Nielsen HB, Fiil BK, Petersen K, MacKinlay J, Loake GJ,
post-transcriptional gene silencing and virus resistance. Plant Cell 14:
Mundy J, Morris PC (2008b) Arabidopsis mitogen-activated protein kinase
629 – 639
kinases MKK1 and MKK2 have overlapping functions in defense signaling
Motomura K, Le QT, Kumakura N, Fukaya T, Takeda A, Watanabe Y (2012)
The role of decapping proteins in the miRNA accumulation in Arabidopsis
thaliana. RNA Biol 9: 644 – 652
Munchel SE, Shultzaberger RK, Takizawa N, Weis K (2011) Dynamic profiling
of mRNA turnover reveals gene-specific and system-wide regulation of
mRNA decay. Mol Biol Cell 22: 2787 – 2795
Nakagawa T, Kurose T, Hino T, Kawamukai M, Niwa Y, Toyooka K, Matsuoka
mediated by MEKK1, MPK4, and MKS1. Plant Physiol 148: 212 – 222
Rasmussen MW, Roux M, Petersen M, Mundy J (2012) MAP kinase cascades
in Arabidopsis innate immunity. Front Plant Sci 3: 169
Ravet K, Reyt G, Arnaud N, Krouk G, Djouani E, Boucherez J, Briat J, Gaymard
F (2012) Iron and ROS control of the DownSTream mRNA decay pathway
is essential for plant fitness. EMBO J 31: 175 – 186
Rayson S, Arciga-Reyes L, Wooton L, de Torres Zabala M, Truman W, Grant
K, Jinbo T, Kimura T (2007) Development of series of gateway binary
M, Davies B (2012a) A role for nonsense-mediated mRNA decay in plants:
vectors, pGWBs, for realizing efficient construction of fusion genes for
pathogen responses are induced in Arabidopsis thaliana NMD mutants.
plant transformation. J Biosci Bioeng 104: 34 – 41
PLoS One 7: e31917
56
Published online: January 20, 2015
The EMBO Journal
Rayson S, Ashworth M, de Torres Zabala M, Grant M, Davies B (2012b) The
salicylic acid dependent and independent effects of NMD in plants. Plant
Signal Behav 7: 1434 – 1437
Riehs-Kearnan N, Gloggnitzer J, Dekrout B, Jonak C, Riha K (2012) Aberrant
growth and lethality of Arabidopsis deficient in nonsense-mediated RNA
decay factors is caused by autoimmune-like response. Nucleic Acids Res
40: 5615 – 5624
Rodriguez-Cousino N, Lill R, Neupert W, Court D (1995) Identification and
mRNA decapping component PAT1 in immunity
Milena Edna Roux et al
Thomas M, Loschi M, Desbats M, Boccaccio G (2011) RNA granules: the good,
the bad and the ugly. Cell Signal 23: 324 – 334
Totaro A, Renzi F, La Fata G, Mattioli C, Raabe M, Urlaub H, Achsel T (2011)
The human Pat1b protein: a novel mRNA deadenylation factor identified
by a new immunoprecipitation technique. Nucleic Acids Res 39: 635 – 647
Ubersax JA, Ferrell JE Jr (2007) Mechanisms of specificity in protein
phosphorylation. Nat Rev Mol Cell Biol 8: 530 – 541
Van Mullem V, Wery M, De Bolle X, Vandenhaute J (2003) Construction of a
initial characterization of the cytosolic protein Ycr77p. Yeast 11:
set of Saccharomyces cerevisiae vectors designed for recombinational
581 – 585
cloning. Yeast 20: 739 – 746
Rymarquis LA, Souret FF, Green PJ (2011) Evidence that XRN4, an Arabidopsis
homolog of exoribonuclease XRN1, preferentially impacts transcripts with
certain sequences or in particular functional categories. RNA 17: 501 – 511
Rzeczkowski K, Beuerlein K, Müller H, Dittrich-Breiholz O, Schneider H,
Kettner-Buhrow D, Holtmann H, Kracht M (2011) c-Jun N-terminal kinase
phosphorylates DCP1a to control formation of P bodies. J Cell Biol 194:
581 – 596
Salgado-Garrido J, Bragado-Nilsson E, Kandels-Lewis S, Seraphin B (1999) Sm
and Sm-like proteins assemble in two related complexes of deep
evolutionary origin. EMBO J 18: 3451 – 3462
Scheller N, Resa-Infante P, la Luna de S, Galao RP, Albrecht M, Kaestner L,
Lipp P, Lengauer T, Meyerhans A, Díez J (2007) Identification of PatL1, a
human homolog to yeast P body component Pat1. Biochim Biophys Acta
1773: 1786 – 1792.
Shi C, Baldwin I, Wu J (2012) Arabidopsis plants having defects in
nonsense-mediated mRNA decay factors UPF1, UPF2, and UPF3 show
photoperiod-dependent phenotypes in development and stress responses.
J Integr Plant Biol 54: 99 – 114
Suarez-Rodriguez MC, Adams-Phillips L, Liu Y, Wang H, Su S-H, Jester PJ,
Zhang S, Bent AF, Krysan PJ (2007) MEKK1 is required for flg22-induced
MPK4 activation in Arabidopsis plants. Plant Physiol 143: 661 – 669
Teige M, Scheikl E, Eulgem T, Doczi F, Ichimura K, Shinozaki K, Dangl JL, Hirt
Vogel F, Hofius D, Paulus KE, Jungkunz I, Sonnewald U (2011) The second
face of a known player: Arabidopsis silencing suppressor AtXRN4 acts
organ-specifically. New Phytol 189: 484 – 493
Wang X, Watt P, Louis E, Borts RH, Hickson I (1996) Pat1: a topoisomerase
II-associated protein required for faithful chromosome transmission in
Saccharomyces cerevisiae. Nucleic Acids Res 24: 4791 – 4797
Weber C, Nover L, Fauth M (2008) Plant stress granules and mRNA processing
bodies are distinct from heat stress granules. Plant J 56: 517 – 530
Xu J, Yang J-Y, Niu Q-W, Chua N-H (2006) Arabidopsis DCP2, DCP1, and
VARICOSE form a decapping complex required for postembryonic
development. Plant Cell 18: 3386 – 3398
Xu J, Chua NH (2009) Arabidopsis decapping 5 is required for mRNA
decapping, P-Body formation, and translational repression during
postembryonic development. Plant Cell 21: 3270 – 3279
Xu J, Chua N-H (2011) Processing bodies and plant development. Curr Opin
Plant Biol 14: 88 – 93
Xu J, Chua NH (2012) Dehydration stress activates Arabidopsis MPK6 to signal
DCP1 phosphorylation. EMBO J 31: 1975 – 1984
Yang L, Wu G, Poethig RS (2012) Mutations in the GW-repeat protein SUO
reveal a developmental function for microRNA-mediated translational
repression in Arabidopsis. PNAS 109: 315 – 320
Yoon J-H, Choi E-J, Parker R (2010) Dcp2 phosphorylation by Ste20 modulates
H (2004) The MKK2 pathway mediates cold and salt stress signaling in
stress granule assembly and mRNA decay in Saccharomyces cerevisiae. J
Arabidopsis. Mol Cell 15: 141 – 152
Cell Biol 189: 813 – 827
Tharun S, He WH, Mayes AE, Lennertz P, Beggs JD, Parker R (2000) Yeast
Zhang J, Shao F, Li Y, Cui H, Chen L, Li H, Zou Y, Long C, Lan L, Chai J, Chai J,
Sm-like proteins function in mRNA decapping and decay. Nature 404:
Chen S, Tang X, Zhou J-M (2007) A Pseudomonas syringae effector
515 – 518
inactivates MAPKs to suppress PAMP-induced immunity in plants. Cell Host
Tharun S, Parker R (2001) Targeting an mRNA for decapping: displacement of
translation factors and association of the Lsm1p-7p complex on
deadenylated yeast mRNAs. Mol Cell 8: 1075 – 1083
Tharun S, Muhlrad D, Chowdhury A, Parker R (2005) Mutations in the
Saccharomyces cerevisiae LSM1 gene that affect mRNA decapping and
30 end protection. Genetics 170: 33 – 46
Tharun S (2009) Lsm1-7-Pat1 complex. RNA Biol 6: 228 – 232
Microbe 1: 175 – 185
Zhang Z, Wu Y, Gao M, Zhang J, Kong Q, Liu Y, Ba H, Zhou J, 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
Zipfel C (2009) Early molecular events in PAMP-triggered immunity. Curr
Opin Plant Biol 12: 414 – 420
57
Figure S1: Pat1 alleles, protein and immunoblotting
Phylogenetic tree of PAT1 orthologues (neighbor-joining tree generated from MAFFT multiple alignment in Geneious
version 5.6.6 created by Biomatters (http://www.geneious.com/). AtPAT, Arabidopsis thaliana; CePAT, C. elegans; OsPAT1,
Oryza sativa; HPat, Drosophila melanogaster; HsPAT, Homo sapiens; ScPAT, Saccharomyces cerevisiae; XlPAT, Xenopus
laevis. RT-PCR analysis of PAT1, ACTIN8, PAT1H1 and PAT1H2 gene expression in Arabidopsis Col-0 seedlings.C.
PAT1
gene structure. Promoter (grey box), exons (white boxes) and introns (black lines). Start codon is marked ATG. T-DNA
insertion sites of pat1-1 (Salk_040660) and pat1-2 (WiscDs_437D04) are marked with black triangles in exon. D. PAT1
contains N-terminal, Pro-rich, middle and Pat-C domains. Peptides for antibody generation are indicated by black
pentagons. Potential SP phosphosites are marked with black stars. Amino acid numbers are above each domain.
E. Anti-PAT1 immunoblotting of yeast protein extracts from pat1∆ transformed with AtPAT1 and untransformed pat1∆.
AtPAT1 corresponds to the 89 kDa band seen in the pat1∆ transformed with AtPAT1. A nonspecific band indicated by *.
58
F. Co-IP between PAT1-GFP and DCP1-HA. Transgenic lines expressing DCP1-HA (1,2) or PAT1-GFP and DCP1-HA (3,4)
were treated with water (-) or 200nM flg22 (+) for 15 mins. Proteins were extracted and subjected to HA and GFP IPs.
Immunoblots of inputs (20 ug each, top panel), GFP IPs (middle panel) and HA IPs (bottom panel) were probed with antiGFP and anti-HA antibodies. GFP IPs were also probed with anti-MPK4 antibodies as a control for PAT1-GFP IPs.
Molecular weights indicated on the right in kDa.
G. The anti-PAT1 antibody does not cross-react with any protein in pat1-1 mutant total protein extracts. Protein extracts (35
µg) derived from wild-type Col-0 and pat1-1 probed with 1:250 dilution of anti-PAT1 and alkaline phosphatase-conjugated
anti-rabbit antibodies.
Figure S2: Mass spectrometry analysis of PAT1-GFP.
Colloidal Coomassie-stained SDS-PAGE of GFP IPs from Col-0
or Col-0/PAT1-GFP treated with water or flg22. Molecular weight
in kDa indicated on the left.
Figure S3: MPK3 does not phosphorylate PAT1 in vitro.
MPK3 was immunoprecipitated from Col-0 seedlings treated with water (-)
or 200 nM flg22 (+) for 10 minutes. MPK3 IPs were incubated with His 6PAT1 or MBP for 30 mins at 37°C before SDS-PAGE. Autoradiogram (top
panel), Coomassie-stained gel for loading control (bottom). Molecular
weight in kDa indicated at right.
59
Figure S4: Pat1 mutants have a leaf serration phenotype.
4-week-old (left) or 8-week-old (right) plants grown under short-day conditions.
60
Figure S5: PAT1-GFP complements the pat1-1 phenotype.
A. Plants photographed at 5-weeks of growth in short-day conditions where PAT1-GFP complements the pat1-1 leaf
serration phenotype. B. The accumulation of PR1 and UGT87A2 transcripts in pat1-1 are reverted when introducing PAT1GFP under control of the native PAT1 promoter. Transcript levels were quantified by qRT-PCR using RNA extracted from
5-week-old plants grown in soil under short-day conditions. Fold change is represented relative to Col-0 and normalized to
ACT2. Standard error of the mean is indicated by error bars (n=3) and statistical significance determined by ANOVA
followed by Tukey's test is shown as letters above error bars.
61
Figure S6: effect of flg22 treatment on PAT1 transcript and PAT1-GFP protein.
Confocal microscopy with root elongation zones of five-days-old Col-0/PAT1-GFP seedlings. Roots treated in DMSO show no
significant change in localization or amounts of GFP signal. The number of GFP foci in PAT1-GFP seedlings is clearly
induced upon treatment with 1μM flg22 on glass slides for 20 min (second panel). No reduction in GFP foci is seen after
treatment with DMSO for 60 min (bottom panel). The scale bar corresponds to 10 µm. qRT-PCR analysis of PAT1
transcripts in Col-0 seedlings treated with time course of 100 nM flg22 with time in minutes post-treatment as indicated.
UBQ10 was used for normalization. Expression is displayed as fold change (where Col-0 at t=0 mins is 1). Standard error of
the mean is indicated by error bars (n=3).
62
Figure S7: pat1 summ2 mutant phenotype
A. 4-week-old plants grown under short-day
conditions.
B. qPCR analysis of SUMM2 transcripts in 4week-old Arabidopsis Col-0, pat1-1, summ2-8
and pat1-1 summ2-8 plants. UBQ10 was used for
normalization. Expression is displayed as fold
change (where Col-0 = 1). Standard error of the
mean is indicated by error bars (n=3).
Figure S8: Infection of pat1-1 summ2-8 mutants with necrotrophic fungus. B. cinerea drop inoculation infection of 4-week-old
Col-0, pat1-1, summ2-8 and pat1-1 summ2-8 plants sampled at 0, 2 and 3 days. Data are the average of 3 biological replicates.
Disease severity was measured as accumulation of the B. cinerea CUTINASE A transcript by qPCR relative to A. thaliana αSHAGGY KINASE (At5g26751).
63
Figure S9: PAT1 is associated with LSM1, MPK4 and SUMM2 in N. benthamiana. Proteins were transiently co-expressed
(PAT1-HA with LSM1-GFP, MycGFP, MPK4-GFP or SUMM2-GFP) in N. benthamiana and tissue and harvested 2 days
post-infiltration. Anti-HA immunoblot of PAT1-HA IPs are shown in the top panel. Anti-GFP immunoblot of PAT1-HA IPs
are shown in the middle panels. The bottom panel show the presence of LSM1-GFP, Myc-GFP and MPK4-GFP in the input.
Although we were unable to detect SUMM2-GFP in the input SUMM2-GFP was easily detected in the enriched PAT1-HA IP.
Molecular weights are shown in kDa at left.
64
Figure S10: Multiple alignments of key regions of PAT1 homologues from several kingdoms. Alignment used MAFFT
algorithm in Geneious version 5.6.6 created by Biomatters (http://www.geneious.com/). Pink boxes mark predicted SP sites;
yellow boxes mark phosphorylation sites identified by MS in this study.
65
AOC3 is an MPK4 substrate and its overexpression
induce immunity against pst DC3000
Introduction
In response to pathogen perception plants rapidly instigate local defense mechanisms. Two systems
for detecting pathogens are: (i) Pathogen recognition receptors on the plant cell surface that
recognize conserved pathogenic structures termed pathogen associated molecular patterns (PAMPs)
and (ii) cytoplasmic immune receptors that recognize either injected pathogen effectors or
modifications of effector host targets (Rasmussen et al., 2012). Both mechanisms induce rapid
localized responses but can also induce systemic acquired resistance mediated by phytohormones
(Robert-Seilaniantz et al., 2011). The transduction of and/or responses to defense signaling in
Arabidopsis involves at least three phytohormones that mediate appropriate responses toward to
specific classes of pathogens. These hormones are salicylic acid (SA), jasmonic acid (JA) and
ethylene (ET) or their derivatives (Bari and Jones, 2008). In response to pathogen perception, plants
produce these hormones in ratios depending on the specific class of pathogen (Robert-Seilaniantz et
al., 2011). Plants usually respond to biotrophic pathogens feeding on living plant tissue by
producing SA, whereas necrotrophic pathogen feeding on dead tissue typically induces production
of JA and ET (Glazebrook, 2005). The lifestyle of many pathogens are not easily classified as either
biotrophic or necrotrophic. Therefore, the otherwise antagonistic SA and JA/ET pathways are likely
correlated to tailor response to specific pathogens (Adie et al., 2007). Some biotrophic plant
pathogens have developed ways of suppressing SA mediated defenses by inducing the antagonistic
JA signaling pathway. For example, this is accomplished by introducing the bacterial JA analog
coronatine into the plant cell leading to suppression of defense responses against biotrophic
pathogens (Cui et al., 2005; Laurie-Berry et al., 2006). Athough coronatine in some instances mimics
JA, this is not always true. For example, plants respond to coronatine by opening their stomata
allowing pathogens to enter the apoplastic space while JA induces stomatal closure (Melotto et al.,
2006; Suhita et al., 2004).
66
The synthesis of defense related hormones including JA is upregulated in response to recognition of
specific pathogens. The first steps in the biosynthesis of JA occur in the chloroplast (Fig. 1). Here
α-linolenic acid present in the thylakoid membrane is oxygenated by a lipoxygenase (LOX) to yield
the fatty acid peroxide 13(S)‐HpOTrE. This peroxide is next dehydrated by ALLENE OXIDE
SYNTEHASE (AOS) to the unstable epoxide 12,13(S)-epoxyoctadecatrienoic acid which is further
modified by ALLENE OXIDE CYCLASE (AOC) to produce 9S,13S-12-oxo-phytodienoic acid
(OPDA) (Hamberg and Fahlstadius, 1990; Schaller et al., 2004; Vick and Zimmerman, 1979;
Zimmerman and Feng, 1978). In aqueous solutions the epoxide produced from AOS degrades
rapidly (half-life of less than 30 seconds at 0˚C) into mainly α- and γ-ketols while a minor fraction
undergoes cyclization producing a racemic mixture of OPDA (Hamberg, 1988; Ziegler et al., 1999).
This is in contrast to the AOC catalyzed conversion which produces only the optically pure
enantiomer 9S,13S-OPDA used in downstream JA biosynthesis (Hamberg and Fahlstadius, 1990).
Due to the short half-life of the AOS product, it is likely that the functions of AOS and AOC are
coupled in a complex, although no direct evidence of such an interaction has been shown.
Translocation of OPDA to the peroxisomes, where the remaining steps in the pathway occur, is not
fully understood. OPDA itself can function as a signaling compound and therefor its translocation
across the chloroplast envelope is likely to be tightly regulated (Dave and Graham, 2012).
Translocation from the cytosol to the peroxisome is actively facilitated by ABC transporters
(Gerhardt, 1983; Theodoulou et al., 2005). In the peroxisomes OPDA is reduced enzymatically by
an OPDA reductase (OPR) and subsequently undergoes three rounds of β-oxidation resulting in the
production of JA (Fig. 1) (Li et al., 2005; Schaller et al., 2000).
67
Figure 1: The JA biosynthesis pathway.
In the chloroplasts α-linolenic acid is oxygenated by LOX to generate 13(S)-HpOTrE which is further modified by AOS to
yield the unstable epoxide 12,13(S)-epoxyoctadecatrienoic acid. In the absence of AOC the epoxide rapidly hydrolyzes into
mainly α- and γ-ketols while a minor fraction undergoes cyclization into a racemic mixture of OPDA. In contrast, AOC
catalyzed cyclization of the epoxide produces only the optically pure 9S,13S-OPDA which then is transported out of the
chloroplast to the cytosol by an unknown mechanism. From the cytosol 9S,13S-OPDA is actively transported into the
peroxisomes by an ABC transporter where it is reduced by OPR and undergoes three rounds of β-oxidation resulting in the
production of JA.
68
AOC displays a greater substrate specificity than AOS and it thus appears that AOC confers
additional specificity on the JA biosynthetic pathway (Schaller et al., 2008). The Arabidopsis
genome contain four genes encoding AOC enzymes: AOC1 (At3g25770), AOC2 (At3g25780),
AOC3 (At3g25760) and AOC4 (At1g13280). These four genes share great sequence identity (Fig.
2) and are most likely functionally redundant to each other as individual T-DNA loss-of-function
mutants do not exhibit JA-related phenotypes (Stenzel et al., 2012). Promoter analysis using a GUS
reporter system has also shown partially redundant promoter activities during development: (i) In
fully developed leaves AOC1, AOC2 and AOC3 promoter activity are apparent throughout all leaf
tissue, whereas AOC4 promoter activity are specific to vascular tissue; (ii) only AOC3 and AOC4
promoter activities are detected in roots; (iii) AOC1 and AOC4 promoter activities are detected
during flower development (Stenzel et al., 2012). In Arabidopsis AOC2 is the best characterized of
the four AOC enzymes and its crystal structure confirms previous findings that AOC forms a
functional trimer (Hofmann et al., 2006). The regulation of the JA biosynthesis pathway is not fully
understood, although positive feedback regulation seems able to induce accumulation of transcripts
encoding genes involved in the biosynthetic pathway of JA (Wasternack and Hause, 2013).
Interestingly all four AOC enzymes can perform both homo and heteromerization in vivo with each
other. This might serve as a level of regulation in JA biosynthesis since some spatial and temporal
promoter activities are observed for the different AOC promoters (Stenzel et al., 2012).
Figure 2: Alignment of AOC1-4 from Arabidopsis
Peptide sequences were aligned with “CLC Main Workbench” without their chloroplast transit peptides (first 81-84 aa).
Putative MAP kinase phosphorylation sites (SP/AP) are indicated with red background coloring. Secondary structures
according the crystal structure of AOC2 are indicated above the alignment.
69
One of the early responses to PAMP perception is activation of MAP Kinase (MPK) signaling
(Rasmussen et al., 2012). MPK4 activation is linked to recognition of the bacterial PAMP flagellin
and thus functions in transducing immune signaling through phosphorylation of downstream targets
(Qiu et al., 2008; Rodriguez et al., 2010; Roux et al., 2015). MPK4 is a target for the bacterial
effector protein AvrB indicating MPK4s pivotal role in immune signaling (Cui et al., 2010). This is
confirmed by the fact that SUPPRESSOR OF MKK1 MKK2 2 (SUMM2), a cytoplasmic immune
receptor, is activated in the absence of MPK4 or MPK4 activity (Zhang et al., 2012). In
investigating the molecular function of MPK4, a yeast two-hybrid screen was initiated searching for
MPK4 interacting proteins (Andreasson et al., 2005). AOC3 was identified in this screen as a fulllength cDNA clone. Here we characterize the interaction between AOC3 and MPK4. We show that
AOC3 is a substrate of MPK4 and that overexpressing AOC3 in Arabidopsis trigger enhanced
immunity. We also provide evidence that MPK4 is associated with the chloroplast, that may
indicate that MPK4 can regulate AOC3 activity/stability inside the chloroplast in response to
pathogens.
Results
AOC3 Interacts with MPK4 in planta.
In a previously conducted yeast two-hybrid screen we identified AOC3 as an interactor of MPK4.
To confirm the AOC3-MPK4 interaction in planta, we transiently expressed in Nicotiana
benthamiana AOC3-GFP along with MPK4-HA and then immunoprecipitated AOC3-GFP. MPK4HA was detected in AOC3-GFP immunoprecipitates from plants transiently expressing both
MPK4-HA and AOC3-GFP, but not in control plants only expressing MPK4-HA (Fig. 3). Thus,
MPK4 and AOC3 can be found in complex in planta.
We also generated double transgenic Arabidopsis lines in the mpk4-1 background expressing
MPK4 with a C-terminal HA tag under control of its native promoter and AOC3 tagged Cterminally with either Myc or GFP under control of a 35S promoter. MPK4-HA were detected in
AOC3-Myc and AOC3-GFP immunoprecipitates only from double transgenic lines but not from
control lines expressing only MPK4-HA (Fig. 4).
70
Figure 3: MPK4 is in complex with AOC3 in planta.
MPK4-HA is detected in AOC3-GFP immunoprecipitates from N. benthamiana transiently expressing AOC3-GFP and
MPK-HA. Immunoblots of input and anti-GFP immunoprecipitates are probed with anti-HA and anti-GFP antibodies
Figure 4: MPK4 is in complex with AOC3 in Arabidopsis.
Immunoprecipitates from four-week-old plants expressing MPK4-HA and AOC3-GFP or AOC3-Myc. (A) MPK4-HA is
detected in AOC3-GFP immunoprecipitates. Immunoblots of input and anti-GFP immunoprecipitates are probed with antiHA and anti-GFP antibodies. (B) MPK4-HA is detected in AOC3-Myc immunoprecipitates. Immunoblots of input and antiMyc immunoprecipitates are probed with anti-HA and anti-Myc antibodies.
MPK4 is activated by treatment with the bacterial flagellin-derived flg22 peptide (Rodriguez et al.,
2010). We questioned whether activation of Arabidopsis immune responses including MPK
activation would affect the interaction between AOC3 and MPK4. We therefore pre-treated doubly
transgenic plants expressing AOC3-GFP and MPK4-HA for 10 minutes with 100nM flg22 before
immunoprecipitation of AOC3-GFP. MPK4-HA was detected in immunoprecipitations from both
flg22 and mock treated plants (Fig. 5). A less intense band was observed in the
immunoprecipitation from flg22 treated plants, but the experiment has not been repeated and
therefore it is not possible to conclude if the lower intensity arises from stochastic events or if the
affinity of the interactions decreases upon flg22 perception.
71
Figure 5: The AOC3-MPK4 interaction is independent of flg22.
Three-week-old plants grown in liquid media were incubated with or without 200 nM flg22 for 15 min. MPK4-HA were
detected in anti-GFP immunoprecipitates from both mock and flg22 treated double transgenic Arabidopsis plants.
Immunoblots of input and anti-GFP immunoprecipitates are probed with anti-HA and anti-GFP antibodies.
Over-expression of AOC3 leads to increased resistance
While mpk4-1 plants expressing MPK4-HA under its native promoter are complemented and thus
indistinguishable from wild type overexpression of AOC3-Myc or AOC3-GFP in these plants
caused dwarfism (Fig. 6A). Mpk4 mutants also exhibit dwarfism and we therefore asked if
overexpression of tagged AOC3 also caused accumulation of PR1 protein. Plants overexpressing
AOC3-Myc accumulated PR1 protein to a similar extent as the mpk4-1 mutant (Fig. 6B). AOC3GFP overexpressing plants also accumulated PR1 protein but to a lesser extent and mpk4
complemented lines had no detectable accumulation of PR1 protein (Fig. 6B). Since mpk4 plants
also exhibit increased resistance towards virulent pathogens, we next assayed the susceptibility of
AOC overexpressing plants. Both AOC3-GFP and AOC3-Myc overexpressing plant exhibited
increased resistance against pst DC3000 (Fig. 6C). Thus, overexpression of AOC3 interestingly
leads to accumulation of PR1 protein and induced resistance against pst DC3000.
AOC3 is an MPK4 substrate
Since MPK4 and AOC3 are found in complexes in planta, it is likely that AOC3 also represent an
MPK4 substrate. Therefore, we carried out in vitro kinase assays using immunoprecipitated flg22activated MPK4 or MPK3 or MPK6 to determine whether AOC3 is also an MPK4 specific
substrate. These MPKs were then incubated with recombinant His6-AOC3 protein lacking the
chloroplast transit peptide. Phosphorylated His6-AOC3 was detectable as a radioactive band around
20 kDa after incubation with flg22-activated MPK4 (Fig. 7), while MPK6 caused only low levels of
AOC3 phosphorylation and MPK3 did not significantly alter AOC3 phosphorylation. Each MPK
72
Figure 6: Overexpression of AOC3 induce resistance.
(A) Plants photographed at 4 weeks of growth in short day conditions, genotypes as indicated. (B) PR1 protein accumulation
is elevated in mpk4-1 complemented lines overexpressing AOC3-myc and AOC3-GFP. Four-week-old plants were used for
protein extraction and Bradford determined protein concentrations were adjusted to equal levels before SDS page (Samples
are run in duplicates). Immunoblots were probed with anti-PR1. (C) Tagged AOC3 lines are more resistant to colonization
by Pst DC3000 (OD600=0.001). Bacteria were syringe-infiltrated and samples taken 4 days post-infiltration. Data are shown
as log10-transformed colony-forming units/cm2 leaf tissue. Standard error of the mean is indicated by errors bars (n = 7).
Statistical significance between the mean values was determined by ANOVA followed by Tukey's test. Multiplicity adjusted P
values are shown for significantly different mean values.
73
Figure 6: AOC3 is phosphorylated in vitro by MPK4
MPK-3, -4 and -6 were immunoprecipitated from extracts of Col-0 seedlings treated with 200 nM flg22 for 10 min. For the
negative control (-), extracts were incubated with agarose beads without antibodies. Immunoprecipitates were incubated with
His6-AOC3 or MBP for 60 min at 30°C before SDS-PAGE. Autoradiogram (top panel), Coomassie-stained gel for loading
control (bottom).
was also incubated with the generic MPK substrate myelin basic protein (MBP) to verify their
activation (Fig. 7). Thus, AOC3 is an in vitro substrate of MPK4, if this is also true in vivo is the
subject of further research.
AOC3 contains 1 Ser-Pro (SP) and 1 Thr-Pro (TP) motif which are commonly phosphorylated by
MPKs (Pearson et al., 2001; Ubersax and Ferrell Jr, 2007). The SP motif is conserved in all 4
Arabidopsis AOC enzymes (Fig. 2) and is central for their enzyme activity (Hofmann et al., 2006;
Schaller et al., 2008). In AOC2 the SP motif (Ser-96,Pro-97) and two Asn residues (Asn-90,Asn118) are important for stabilizing the reactive oxygen in the epoxide group in catalyzing the
cyclization of 13-HPOT into OPDA (Hofmann et al., 2006; Schaller et al., 2008). Online database
searches (PhosPhAt and P3DB) showed that AOC2 from Arabidopsis, as well as AOC4 from
Glycine max (Soybean) and Medicago truncatula (barrel clover), are phosphorylated at this specific
SP motif (Durek et al., 2010; Yao et al., 2014). It is reasonable to believe that this phosphorylation
might regulate the activity of AOC enzymes as addition of the negatively charged phosphate group
to the SP motif could affect the stabilization of the epoxide group in the 13-HPOT substrate. To test
if MPK4 can phosphorylate the SP site we incubated a mutant version of His6-AOC3 with Ser-100
mutated to alanine (S100A) with MPK4 immunoprecipitates and measured the incorporation of
radioactive labeled AT32P (Fig. 8). There was no clear difference in the phosphorylation of AOC3
and AOC3-S100A meaning that (i) MPK4 does not phosphorylate AOC3-Ser100 in vitro or (ii) the
74
phosphorylation of an additional residue(s) mask the reduction in phosphorylation of the AOC3S100A mutant.
To investigate the MPK4 mediated phosphorylation of AOC3 we have cloned, expressed and
purified the additional phospho site mutants AOC3-T233A and AOC-S100A/T233A, but we have
not yet tested their specificity as MPK4 substrates.
Figure 6: AOC3A100A is phosphorylated by MPK4.
MPK4 was immunoprecipitated from extracts of Col-0 seedlings treated with 200 nM flg22 for 10 min. For the negative
control (-), extracts were incubated with agarose beads without antibodies. immunoprecipitates were incubated with His6AOC3 or His6-AOC3S100A for 60 min at 30°C before SDS-PAGE. Autoradiogram (top panel), Coomassie-stained gel for
loading control (bottom).
MPK4 associates with chloroplasts.
AOC3 is transcribed in the nucleus, translated in the cytosol and an N-terminal transit peptide
guides the protein to the chloroplast where it functions in JA biosynthesis (Schaller et al., 2008).
MPK4 has previously been shown to be present in the cytoplasm and nucleus, but no association to
plastids has so far been shown (Andreasson et al., 2005; Petersen et al., 2000, 4). A large scale
phosphoprotein analysis discovered that numerous chloroplast proteins are phosphorylated at
putative MAP kinase phosphorylation motifs (Reiland et al., 2009). However, no MAP kinase has
been shown to associate with chloroplasts. We therefore questioned where the interaction between
MPK4 and AOC3 occurs and investigated whether MPK4 associates with chloroplasts.
To answer these questions, we isolated intact chloroplasts from transgenic plants expressing MPK4HA and wild type Ler plants and probed for MPK4-HA by immunoblotting (Fig. 9). A clear,
distinct band for MPK4-HA was observed in the chloroplast fraction from MPK4-HA expressing
plants, but was absent in the control plants. We further probed for the presence of MPK6, another
immunity regulating MAPK (Rasmussen et al., 2012). MPK6 was detected only in the total protein
75
fraction and not in the chloroplast fraction (Fig. 9). To exclude the possibility that the chloroplast
fractions were contaminated with cytoplasmic protein we probed for the cytoplasmic protein
DECAPPING 1 (DCP1) (Fig. 9) (Xu et al., 2006, 2). We detected DCP1 only in the total protein
sample and not in the chloroplast fractions. This indicates that these fractions where not
contaminated with cytoplasmic proteins. We also probed for the presence of the chloroplast protein
COPPER/ZINC SUPEROXIDE DISMUTASE 2 (CDS2) to verify presence of chloroplastic
proteins (Fig. 9) (Pilon et al., 2011). CSD2 was observed in both total protein samples and the
chloroplast fractions. Whereas CSD2 is present in the chloroplast, CSD1 is a cytoplasmic protein
(Pilon et al., 2011). The CSD2 antibody cross-reacts with CSD1 and a lower band is observed in the
total protein samples but is absent in the chloroplast fractions. This band correlates to the size of
CSD1 (15 kDa). These results confirm that the chloroplast fractions are free of cytoplasmic
proteins.
Figure 9: MPK4 is detected in isolated chloroplasts.
Intact chloroplasts were isolated from leaf tissue of four-week-old plants expressing MPK4-HA under control of its native
promoter or from Ler wild type plants. Total protein samples and chloroplast fractions were then subjected to SDS-PAGE
and immunoblotting.
76
MPK4 does not have a chloroplast transit peptide. However, proteomic data suggest that up to 30%
of the protein species in the chloroplast lack encoded transit peptides and therefore are translocated
to the chloroplasts by unknown mechanisms (Armbruster et al., 2009). It is therefore possible that
MPK4 can translocate to the chloroplasts even in absence of a transit peptide. Alternatively MPK4
could also simply associate with the outer envelope of the chloroplast without actually being
translocated into the chloroplast.
Discussion
In this study, we identify AOC3 as an interactor and substrate of the immunity regulating MPK4 in
Arabidopsis. AOC3 is located in the chloroplasts where it catalyzes the production of the JA
precursor ODPA (Hamberg and Fahlstadius, 1990). Although the enzymatic steps in the
biosynthesis of JA are well understood, information regarding regulation is scarce. A possible point
of regulation might be controlling the availability of enzymes necessary in the pathway. LOX2
catalyzes the first step in the JA biosynthesis by converting α-linolenic acid into 13(S)‐HpOTrE
(Stenzel et al., 2003). LOX2 interacts with the eukaryotic translational initiating factor 4E (eIF4E)
in the cytosol (Freire et al., 2000). It is possible that this interaction sequesters LOX2 in the cytosol
thwarting the initial step in JA biosynthesis. Alternatively, it is also possible that LOX2 actually
down-regulates translation through its interaction with eIF4E. Confirming or rejecting one or both
of these hypotheses requires further investigation.
Regulation of the pathway is also likely to occur through phosphorylation. Numerous sites of
chloroplast proteins are mapped through phospho-proteome profiling including several motifs
thought to be phosphorylated by MAP kinases (Reiland et al., 2009). Online database searches
(PhosPhAt and P3DB) show that AOC2 from Arabidopsis as well as AOC4 from Glycine max
(Soybean) and Medicago truncatula (barrel clover) are phosphorylated at a conserved SP motif
(AOC3 Ser-100 Pro-101) (Durek et al., 2010; Yao et al., 2014). This SP motif is important for the
function of the AOC enzymes as it functions in stabilizing the reactive oxygen of the epoxide group
in its substrate when 13-HPOT cyclizes into OPDA (Hofmann et al., 2006; Schaller et al., 2008).
This however does not seem to occur though the activity of MPK4 since an SP to AP mutant was
still phosphorylated by MPK4 in vitro (Fig. 8). The database searches further showed that AOC4
from Medicago truncatula is phosphorylated on an SP site between α-helix 3 and α-helix 4. In
Arabidopsis AOC1-3 has a TP site between α-helix 3 and α-helix 4 while AOC4 has an EP site
putatively mimicking a phosphorylated motif (Fig. 2). It is likely that this motif can be
phosphorylated by MPK4, although this and its possible effects have to be tested experimentally.
77
To gain insight into where the interaction between AOC3 and MPK4 occurs, we investigated if
MPK4 associates with chloroplasts. MPK4 was clearly observed in immunoblots from isolated
chloroplast (Fig. 9). This, however, does not provide information of whether MPK4 is translocated
into the chloroplast or if it merely resides attached to the outer membrane. Therefore, to further
explore the location of the molecular interaction between AOC3 and MPK4 studies such as
bimolecular fluorescence could be applied for a more precise confirmation of where in the cell this
interaction occurs. Nonetheless, the presence of phosphorylated MAP kinase motifs in chloroplast
proteins sustains the idea of MPK4 functioning inside the chloroplast as well as in the cytosol and
nucleus (Andreasson et al., 2005; Reiland et al., 2009). If MPK4 resides inside the chloroplast, it
raises the question whether it is activated by perception of pathogens similar to its cytoplasmic
counterpart? And, if this is the case, can MPK4 regulate JA biosynthesis through phosphorylation
of AOC3 (and AOC1 and -2) upon pathogen perception?
MPK4 was initially considered to regulate JA induced gene transcription as the mpk4 mutant fails to
accumulate JA induced PDF1.2 transcripts (Petersen et al., 2000, 4). However, crossing the mpk4
mutant into an eds1 mutant background which blocks ETI restored the exogenous JA induced
accumulation of PFD1.2 transcripts (Brodersen et al., 2006). This implies that the blockage of JA
induced genes in the mpk4 mutant is a pleiotropic effect of activating ETI and not necessarily linked
to MPK4 lack of function. Further investigation of the phosphorylation of AOC3 by MPK4 may
shed light on a putative regulatory point in JA biosynthesis and how this converges with defense
signaling.
Materials and methods
See material and methods for “eIF4E-Thr6 is specifically phosphorylated in vitro by MPK4 and
form a complex in the nucleus in vivo.”
78
References
Adie, B. A. T., Pérez-Pérez, J., Pérez-Pérez, M. M., Godoy, M., Sánchez-Serrano, J.-J., Schmelz, E.
A., et al. (2007). ABA Is an Essential Signal for Plant Resistance to Pathogens Affecting JA
Biosynthesis and the Activation of Defenses in Arabidopsis. Plant Cell 19, 1665–1681.
doi:10.1105/tpc.106.048041.
Andreasson, E., Jenkins, T., Brodersen, P., Thorgrimsen, S., Petersen, N. H., Zhu, S., et al. (2005).
The MAP kinase substrate MKS1 is a regulator of plant defense responses. EMBO J 24,
2579–2589. doi:10.1038/sj.emboj.7600737.
Armbruster, U., Hertle, A., Makarenko, E., Zühlke, J., Pribil, M., Dietzmann, A., et al. (2009).
Chloroplast Proteins without Cleavable Transit Peptides: Rare Exceptions or a Major
Constituent of the Chloroplast Proteome? Mol. Plant 2, 1325–1335. doi:10.1093/mp/ssp082.
Bari, R., and Jones, J. D. G. (2008). Role of plant hormones in plant defence responses. Plant Mol.
Biol. 69, 473–488. doi:10.1007/s11103-008-9435-0.
Brodersen, P., Petersen, M., Nielsen, H. B., Zhu, S., Newman, M.-A., Shokat, K. M., et al. (2006).
Arabidopsis MAP kinase 4 regulates salicylic acid- and jasmonic acid/ethylene-dependent
responses via EDS1 and PAD4. Plant J. 47, 532–546. doi:10.1111/j.1365313X.2006.02806.x.
Cui, H., Wang, Y., Xue, L., Chu, J., Yan, C., Fu, J., et al. (2010). Pseudomonas syringae Effector
Protein AvrB Perturbs Arabidopsis Hormone Signaling by Activating MAP Kinase 4. Cell
Host Microbe 7, 164–175. doi:10.1016/j.chom.2010.01.009.
Cui, J., Bahrami, A. K., Pringle, E. G., Hernandez-Guzman, G., Bender, C. L., Pierce, N. E., et al.
(2005). Pseudomonas syringae manipulates systemic plant defenses against pathogens and
herbivores. Proc. Natl. Acad. Sci. U. S. A. 102, 1791–1796. doi:10.1073/pnas.0409450102.
Dave, A., and Graham, I. A. (2012). Oxylipin signaling: a distinct role for the jasmonic acid
precursor cis-(+)-12-oxo-phytodienoic acid (cis-OPDA). Plant Physiol. 3, 42.
doi:10.3389/fpls.2012.00042.
Durek, P., Schmidt, R., Heazlewood, J. L., Jones, A., MacLean, D., Nagel, A., et al. (2010).
PhosPhAt: the Arabidopsis thaliana phosphorylation site database. An update. Nucleic Acids
Res. 38, D828–D834. doi:10.1093/nar/gkp810.
Freire, M. A., Tourneur, C., Granier, F., Camonis, J., Amrani, A. E., Browning, K. S., et al. (2000).
Plant lipoxygenase 2 is a translation initiation factor-4E-binding protein. Plant Mol. Biol.
44, 129–140. doi:10.1023/A:1006494628892.
Gerhardt, B. (1983). Localization of β-oxidation enzymes in peroxisomes isolated from nonfatty
plant tissues. Planta 159, 238–246. doi:10.1007/BF00397531.
79
Glazebrook, J. (2005). Contrasting Mechanisms of Defense Against Biotrophic and Necrotrophic
Pathogens. Annu. Rev. Phytopathol. 43, 205–227.
doi:10.1146/annurev.phyto.43.040204.135923.
Hamberg, M. (1988). Biosynthesis of 12-oxo-10,15(Z)-phytodienoic acid: Identification of an
allene oxide cyclase. Biochem. Biophys. Res. Commun. 156, 543–550. doi:10.1016/S0006291X(88)80876-3.
Hamberg, M., and Fahlstadius, P. (1990). Allene oxide cyclase: A new enzyme in plant lipid
metabolism. Arch. Biochem. Biophys. 276, 518–526. doi:10.1016/0003-9861(90)90753-L.
Hofmann, E., Zerbe, P., and Schaller, F. (2006). The Crystal Structure of Arabidopsis thaliana
Allene Oxide Cyclase: Insights into the Oxylipin Cyclization Reaction. Plant Cell 18, 3201–
3217. doi:10.1105/tpc.106.043984.
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. Mol. Plant. Microbe
Interact. 19, 789–800. doi:10.1094/MPMI-19-0789.
Li, C., Schilmiller, A. L., Liu, G., Lee, G. I., Jayanty, S., Sageman, C., et al. (2005). Role of βOxidation in Jasmonate Biosynthesis and Systemic Wound Signaling in Tomato. Plant Cell
17, 971–986. doi:10.1105/tpc.104.029108.
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.
doi:10.1016/j.cell.2006.06.054.
Pearson, G., Robinson, F., Beers Gibson, T., Xu, B., Karandikar, M., Berman, K., et al. (2001).
Mitogen-Activated Protein (MAP) Kinase Pathways: Regulation and Physiological
Functions. Endocr. Rev. 22, 153–183. doi:10.1210/edrv.22.2.0428.
Petersen, M., Brodersen, P., Naested, H., Andreasson, E., Lindhart, U., Johansen, B., et al. (2000).
Arabidopsis MAP Kinase 4 Negatively Regulates Systemic Acquired Resistance. Cell 103,
1111–1120. doi:10.1016/S0092-8674(00)00213-0.
Pilon, M., Ravet, K., and Tapken, W. (2011). The biogenesis and physiological function of
chloroplast superoxide dismutases. Biochim. Biophys. Acta BBA - Bioenerg. 1807, 989–998.
doi:10.1016/j.bbabio.2010.11.002.
Qiu, J.-L., Zhou, L., Yun, B.-W., Nielsen, H. B., Fiil, B. K., Petersen, K., et al. (2008). Arabidopsis
Mitogen-Activated Protein Kinase Kinases MKK1 and MKK2 Have Overlapping Functions
in Defense Signaling Mediated by MEKK1, MPK4, and MKS1. Plant Physiol. 148, 212–
222. doi:10.1104/pp.108.120006.
Rasmussen, M. W., Roux, M., Petersen, M., and Mundy, J. (2012). MAP kinase cascades in
Arabidopsis innate immunity. Plant Proteomics 3, 169. doi:10.3389/fpls.2012.00169.
Reiland, S., Messerli, G., Baerenfaller, K., Gerrits, B., Endler, A., Grossmann, J., et al. (2009).
Large-Scale Arabidopsis Phosphoproteome Profiling Reveals Novel Chloroplast Kinase
Substrates and Phosphorylation Networks. Plant Physiol. 150, 889–903.
doi:10.1104/pp.109.138677.
80
Robert-Seilaniantz, A., Grant, M., and Jones, J. D. G. (2011). Hormone Crosstalk in Plant Disease
and Defense: More Than Just JASMONATE-SALICYLATE Antagonism. Annu. Rev.
Phytopathol. 49, 317–343. doi:10.1146/annurev-phyto-073009-114447.
Rodriguez, M. C. S., Petersen, M., and Mundy, J. (2010). Mitogen-Activated Protein Kinase
Signaling in Plants. Annu. Rev. Plant Biol. 61, 621–649. doi:10.1146/annurev-arplant042809-112252.
Roux, M. E., Rasmussen, M. W., Palma, K., Lolle, S., Regué, À. M., Bethke, G., et al. (2015). The
mRNA decay factor PAT1 functions in a pathway including MAP kinase 4 and immune
receptor SUMM2. EMBO J., e201488645. doi:10.15252/embj.201488645.
Schaller, F., Biesgen, C., Müssig, C., Altmann, T., and Weiler, E. W. (2000). 12-Oxophytodienoate
reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis. Planta 210, 979–
984. doi:10.1007/s004250050706.
Schaller, F., Schaller, A., and Stintzi, A. (2004). Biosynthesis and metabolism of jasmonates. J.
Plant Growth Regul. 23, 179–199. doi:10.1007/BF02637260.
Schaller, F., Zerbe, P., Reinbothe, S., Reinbothe, C., Hofmann, E., and Pollmann, S. (2008). The
allene oxide cyclase family of Arabidopsis&nbsp; thaliana &#x2013; localization and
cyclization. FEBS J. 275, 2428–2441. doi:10.1111/j.1742-4658.2008.06388.x.
Stenzel, I., Hause, B., Miersch, O., Kurz, T., Maucher, H., Weichert, H., et al. (2003). Jasmonate
biosynthesis and the allene oxide cyclase family of Arabidopsis thaliana. Plant Mol. Biol.
51, 895–911. doi:10.1023/A:1023049319723.
Stenzel, I., Otto, M., Delker, C., Kirmse, N., Schmidt, D., Miersch, O., et al. (2012). ALLENE
OXIDE CYCLASE (AOC) gene family members of Arabidopsis thaliana: tissue- and organspecific promoter activities and in vivo heteromerization*. J. Exp. Bot. 63, 6125–6138.
doi:10.1093/jxb/ers261.
Suhita, D., Raghavendra, A. S., Kwak, J. M., and Vavasseur, A. (2004). Cytoplasmic Alkalization
Precedes Reactive Oxygen Species Production during Methyl Jasmonate- and Abscisic
Acid-Induced Stomatal Closure. Plant Physiol. 134, 1536–1545.
doi:10.1104/pp.103.032250.
Theodoulou, F. L., Job, K., Slocombe, S. P., Footitt, S., Holdsworth, M., Baker, A., et al. (2005).
Jasmonic Acid Levels Are Reduced in COMATOSE ATP-Binding Cassette Transporter
Mutants. Implications for Transport of Jasmonate Precursors into Peroxisomes. Plant
Physiol. 137, 835–840. doi:10.1104/pp.105.059352.
Ubersax, J. A., and Ferrell Jr, J. E. (2007). Mechanisms of specificity in protein phosphorylation.
Nat. Rev. Mol. Cell Biol. 8, 530–541. doi:10.1038/nrm2203.
Vick, B. A., and Zimmerman, D. C. (1979). Substrate Specificity for the Synthesis of Cyclic Fatty
Acids by a Flaxseed Extract. Plant Physiol. 63, 490–494. doi:10.1104/pp.63.3.490.
Wasternack, C., and Hause, B. (2013). Jasmonates: biosynthesis, perception, signal transduction
and action in plant stress response, growth and development. An update to the 2007 review
in Annals of Botany. Ann. Bot. 111, 1021–1058. doi:10.1093/aob/mct067.
81
Xu, J., Yang, J.-Y., Niu, Q.-W., and Chua, N.-H. (2006). Arabidopsis DCP2, DCP1, and
VARICOSE Form a Decapping Complex Required for Postembryonic Development. Plant
Cell 18, 3386–3398. doi:10.1105/tpc.106.047605.
Yao, Q., Ge, H., Wu, S., Zhang, N., Chen, W., Xu, C., et al. (2014). P3DB 3.0: From plant
phosphorylation sites to protein networks. Nucleic Acids Res. 42, D1206–D1213.
doi:10.1093/nar/gkt1135.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., et al. (2012). Disruption of PAMPInduced MAP Kinase Cascade by a Pseudomonas syringae Effector Activates Plant
Immunity Mediated by the NB-LRR Protein SUMM2. Cell Host Microbe 11, 253–263.
doi:10.1016/j.chom.2012.01.015.
Ziegler, J., Wasternack, C., and Hamberg, M. (1999). On the specificity of allene oxide cyclase.
Lipids 34, 1005–1015. doi:10.1007/s11745-999-0451-z.
Zimmerman, D. C., and Feng, P. (1978). Characterization of a prostaglandin-like metabolite of
linolenic acid produced by a flaxseed extract. Lipids 13, 313–316.
doi:10.1007/BF02533720.
82
eIF4E-Thr6 is specifically phosphorylated in vitro by
MPK4 and form a complex in the nucleus in vivo.
Introduction
Translation of mRNAs into proteins is primarily regulated at the point of initiation (Jackson et al.,
2010). Canonical translational initiation in most eukaryotes involves scanning of the 5’ untranslated
region (UTR) for the start codon by a ribosomal preinitiation complex, which is comprised of the
40S ribosomal subunit end several eukaryotic initiation factors (eIFs) (Jackson et al., 2010;
Topisirovic et al., 2011). A preliminary step in the formation of the preinitiation complex is the
binding of eIF4F complex to the 5’cap of the mRNA (Merrick, 2015). The eIF4F complex comprise
the proteins eIF4E which is the actual 5’cap binding protein, the RNA helicase eIF4A and the
scaffold protein eIF4G (Merrick, 2015). At the center of this complex is eIF4G a multidomain
protein, which functions as a scaffold protein for both eIF4E and eIF4A (Merrick, 2015). When
eIF4G binds eIF4E it induce allosteric changes and enhance the affinity of eIF4E for the 5’cap of
the mRNA (Gross et al., 2003; Volpon et al., 2006). Intriguingly the binding of the 5’cap of the
mRNA also enhances the interaction between eI4E and eIF4G, thus stabilizing the complex (Volpon
et al., 2006). eIF4G also enhance the activity of the RNA helicase eIF4A and functions as a scaffold
protein that induce allosteric conformational changes to both eIF4A and eIF4E (Gross et al., 2003;
Schütz et al., 2008; Volpon et al., 2006).
eIF4G also interacts with eIF3 and poly-A binding protein (PABP). After the eIF4F complex has
bound the 5’cap of the mRNA eIF3 recruits the 40S ribosomal subunit and several other eIFs
resulting in the formation of the 48S ribosomal initiation complex that scan the 5’UTR for a
transitional start site (Jackson et al., 2010; Topisirovic et al., 2011). The binding of PABP by eIF4G
is thought to circularize the mRNA which is believed to enhance translation due to recycling of the
ribosome after ended translation (Topisirovic et al., 2011).
Initiation of translation is somehow conserved in all eukaryotes although differences occur. For
example, in contrast to the three-subunit eIF4F complex identified in mammals and plants, only a
two-subunit complex comprising eIF4G and eIF4E have been identified in yeast. However, eIF4A
helicase activity in yeast is believed to function in association with eIF4E and eIF4G (Mayberry et
al., 2011; Merrick, 2015). Plants also stand out as in both yeast and mammalian cell lines, but no
plants have eIF4E binding proteins (4E-BP) that regulate the availability of eIF4E by sequestration
thereby inhibiting the formation of the eIF4F complex (Lukhele et al., 2013; Peter et al., 2015). In
83
mammals and yeast phosphorylation of 4E-BPs functions to regulate translational initiation and
phosphorylated 4E-BPs lose the affinity for eIF4E, promoting the formation of the eIF4F complex
and thereby translation (Merrick, 2015). How plants regulate translational initiation in absence of
4E-BPs is unclear. In mammals, a MAPK-interacting kinase (MNK) phosphorylates eIF4E at a
conserved residue (Ser-209) in the C-terminal region (Fig. 1). This phosphorylation upregulates
translation of a subset of genes involved in inflammation responses and tumor progression (Furic et
al., 2010). Using phosphoproteomic several other phosphorylated residues in human eIF4E has been
identified, but no function has been assigned to these phosphorylated residues (Fig.1)
(PhosphoSitePlus Database: Eto, 2010; Franz-Wachtel et al., 2012; Hornbeck et al., 2015; Li et al.,
2009; Rychlik et al., 1987; Sharma et al., 2014; Van Hoof et al., 2009; Zhou et al., 2013). The Ser209 of human eIF4E does not align to any Ser or Thr residues in the Arabidopsis eIF4E (Fig. 1), so
it is plausible that this specific phosphorylation is explicit for humans/mamels. Nonetheless one
phosphosite has been identified in Arabidopsis eIF4E (Fig. 1), but no function has been assigned to
it (PhosPhAt Database: Durek et al., 2010).
At_eIF4E
Hs_eIF4E
*
1 MAVE-----DTPKSVVTEEAKPNSIENPIDRYHEEGDDAEEGEIAGGEGDGNVDESSKSG
1 MATVEPETTPTPNPPTTEEEKTESNQ----------------EVANPEH-----------
At_eIF4E
Hs_eIF4E
56 VPESHPLEHSWTFWFDNPAVKSKQTSWGSSLRPVFTFSTVEEFWSLYNNMKHPSKLAHGA
34 -YIKHPLQNRWALWFFKN---DKSKTWQANLRLISKFDTVEDFWALYNHIQLSSNLMPGC
At_eIF4E
Hs_eIF4E
116 DFYCFKHIIEPKWEDPICANGGKWTMTFPKEK----SDKSWLYTLLALIGEQFD-HGDEI
90 DYSLFKDGIEPMWEDEKNKRGGRWLITLNKQQRRSDLDRFWLETLLCLIGESFDDYSDDV
At_eIF4E
Hs_eIF4E
171 CGAVVNIRGKQERISIWTKNASNEAAQVSIGKQWKEFLDYNN--SIGFIIHED-AKKLDR
150 CGAVVNVRAKGDKIAIWTTECENREAVTHIGRVYKERLGLPPKIVIGYQSHADTATKSGS
At_eIF4E
Hs_eIF4E
228 NAKNAYTA
210 TTKNRFVV
Figure 1: Alignment of human and Arabidopsis peptide sequence
Sequences were aligned with T-Coffee and colored by BoxShade3.2. Known phospho-sites (PhosPhAt and PhosphoSitePlus)
are indicated by red background coloring and putative MPK phospo-sites are marked above the alignment with (*).
84
The Arabidopsis genome contain four different eIF4E genes namely eIF4E (AT4G18040), eIF4E-B
(AT1G29550), eIF4E-C (AT1G29590) and eIF(ISO)4E (AT5G35620). eIF4EB and eIF4EC are
expressed at very low levels and are therefore not thought to function in general translation, but
expression of eIF4EB is upregulated in flowering tissue implying a possible role in reproduction
(Patrick et al., 2014; Rodriguez et al., 1998). eIF4E are expressed throughout all green tissue
whereas eIF(iso)4E are expressed mainly in flowers (Rodriguez et al., 1998). Albeit differential
expression patterns of eIF4E and eIF(iso)4E they are able to complement each other’s function in
Arabidopsis as individual loss-of-function mutants reassemble a wild type phenotype whereas the
double eif4e eif(iso)4e mutants are embryo lethal (Callot and Gallois, 2014). In Arabidopsis eIF4E
and eIF(iso)4E interacts with eIF4G and eIF(iso)4G respectively forming eIF4F and eIF(iso)4F
complexes (Patrick et al., 2012). It is therefore more likely that the complementation occurs through
the eIF4F and eIF(iso)4F complexes. In vitro assays shows that eIF4(iso)4E has an up to 10 fold
lower affinity for a 5’cap analog than eIF4E (Kropiwnicka et al., 2015). This raises the question of
how these proteins compete with each other in vivo for in translational initiation or perhaps it is
more appropriate to question whether they actually compete? Albeit they are able to complement
each other meaning that they are both able to initiate general translation, it is possible that they both
have favored affinity for a subset of differential mRNA species.
Some viruses depend on the host translational machinery for translation of their genome, and a few
of these have been shown to rely specifically on either eIF4E or eIF(iso)4E. Thus, mutations in
these proteins are often associated with loss of susceptibility (Sanfaçon, 2015). Genomic potyvirus
RNA contains a 5′ end covalently linked virus-encoded protein (VPg), which is required for virus
infectivity (Riechmann et al., 1992). The VPg linked to the viral genome interacts with the host
translational machinery on which it is dependent on translation of the viral genome (Wang and
Krishnaswamy, 2012). Therefore, mutations or loss-of-function of indispensable host factors
required by the virus can thwart viral infection (Diaz-Pendon et al., 2004). So far all mutations in
genes conferring recessive resistance in plants are found in eIFs and the vast majority are found in
genes coding for eIF4E or eIF(iso)4E (Wang and Krishnaswamy, 2012). Pytoviruses display
selective involvement of either eIF4E or eIF(iso)4E during infection for example Clover yellow
vein virus (ClYVV) dependents on the interaction between the viral VPg and the host eIF4E and not
eIF(iso)4E for proliferation and (Sato et al., 2005). Whereas both Turnip mosaic virus (TuMV) and
Lettuce mosaic virus (LMV) on the other hand rely on interactions between their viral VPg and host
eIF(iso)4E and not eIF4E (Duprat et al., 2002). In some cases the VPg compete with host 5’capped
mRNAs for their interaction with eIF4E or eI4(iso)4E, but because the specific motifs required for
85
interaction differs for each pytovirus. Thus in some cases eIF4E or eIF(iso)4E interacts with both
host 5’capped mRNAs and viral VPg proteins. And therefore the is no consensus regarding whether
or not domains for binding 5’capped host mRNA and viral VPg proteins overlap (Charron et al.,
2008; Truniger and Aranda, 2009).
Immune signaling is often involve Mitogen Activated Protein kinases (MPKs). In Arabidopsis
MPK3, MPK4 and MPK6 are described as canonical defense regulating kinases (Rasmussen et al.,
2012). All three kinases are activated by perception of Pathogen Associated Molecular Patterns
(PAMPs) by plasma membrane spanning Pathogen Recognition Receptors (PPRs). Activated MPKs
can then transduce immune signaling through for example phosphorylation of defense related
transcription factors (Rasmussen et al., 2012). In plants, virulent pathogens can inject effector protein
directly into the host cell. These effectors can then modify the host to suppress immune responses.
Both MPK3, MPK4 and MPK6 are pathogenic effector targets, implying that they are important
immune regulators (Deslandes and Rivas, 2012). Successful plants have evolved resistance (R)
proteins that can directly sense pathogenic effector proteins or indirectly recognize their effect.
SUPRESSOR OF MKK1 MKK2 1 (SUMM2) surveillances the MPK4 pathway and is activated in
absence of MPK4 activity (Zhang et al., 2012). Activation of SUMM2 activates effectortriggered immunity (ETI) which is somehow similar to PAMP-triggered immunity (PTI), but
defense response are both amplified and prolonged and can include localized programmed cell
death to effectively thwart pathogen growth (Cui et al., 2015; Zhang et al., 2012).
A screen for MPK substrates using functional protein microarrays published by Popescu et al.
(2009) identified 152 possible MPK4 substrates. However, by focusing on substrates
phosphorylated only by MPK4 we could reduce this list to 14 potential MPK4 specific substrates.
We scrutinized these 14 proteins and further eliminated three proteins that lacked the canonical
Ser/The-Pro MPK phosphorylation site (Pearson et al., 2001). We cloned the remaining list of
eleven proteins and co-expressed them in N. benthamiana with MPK4 to identify any possible
interactions by co-immunoprecipitation. We identified eIF4E as an MPK4 interacting protein and
further investigated its relation to MPK4 both in vitro and in Arabidopsis.
86
Results
eIF4E is an MPK4 substrate
In the MPK substrates published by Popescu et al. (2009) we focused on eIF4E as a putative
substrate of MPK4 kinase activity. Not all substrates reported by Popescu et al. (2009) contained the
canonical Ser/Thr-Pro MPK phosphorylation motif (Pearson et al., 2001). We therefore first
performed an in vitro kinase experiment to confirm eIF4E as an MPK4 substrate. We
immunoprecipitated PAMP activated MPKs and incubated them with recombinant His6-eIF4E
protein purified from e.coli. Phosphorylated His6-eIF4E was observed as a radioactive band after
incubation with flg22 activated MPK4, while nor MPK3 or MPK6 resulted in His6-eIF4E
phosphorylation (Fig. 2). Each MPK was also incubated with the generic MPK substrate myelin
basic protein (MBP) to verify their activation (Fig. 2). The eIF4E peptide sequence contains one
putative MPK phosphorylation site (Thr6-Pro7). A version of eIF4E with Thr6 mutated to an
alanine (eIF4E-T6A) was also incubated with immunoprecipitated MPK-3, -4 and -6, this mutant
had significantly lower levels of phosphorylation after incubation with MPK4 (Fig. 2). Thus,
eIF4E-Thr6 is phosphorylated in vitro by MPK4 and not by MPK3 or MPK6. This phosphorylated
TP site aligns with a TP site in the human eIF4E, which is also phosphorylated (Fig. 1). The effect
of phosphorylating this TP site in human eIF4E is unknown, but given its homology to the TP site
in Arabidopsis eIF4E it might have a conserved regulatory function.
Figure 2: MPK4 specifically phosphorylates eIF4E in vitro.
MPK-3, -4 and -6 were immunoprecipitated from extracts of Col-0 seedlings treated with 200 nM flg22 for 10 min. For the
negative control (-), extracts were incubated with agarose beads without antibodies. Immunoprecipitates were incubated with
His6-eIF4E, His6-eIF4E-T6A or MBP for 60 min at 30°C before SDS-PAGE. Autoradiogram (top panel), Coomassie-stained
gel for loading control (bottom).
87
eIF4E interacts with MPK4 in planta
To examine the interaction we generated transgenic Arabidopsis plants in an eif4e T-DNA loss-offunction line (SALK_145583) expressing eIF4E-Myc under control of a 35S promoter. In anti-Myc
immunoprecipitations from transgenic eIF4E-Myc tissue, we detected a 40-kDa band corresponding
to MPK4 (Fig. 3). We questioned whether a non-phophorylatable form of eIF4E would affect the
interaction. We therefore generated transgenic plants expressing eIF4E-T6A-Myc under control of a
35S promoter in the eif4e T-DNA loss-of-function mutant. Anti-Myc immunoprecipitation from
transgenic eIF4E-T6A-Myc tissue detected a more intense band than from wild type eIF4E-Myc
(Fig. 3). Thus, the interaction between MPK4 and the non-phosphorylatable eIF4E-T6A is stronger
than the interaction with eIF4E. This may indicate the interaction with MPK4 is dependent on
whether or not eIF4E is phosphorylated.
Figure 3: MPK4 is in complex with eIF4E eIF4E-T6A in Arabidopsis.
Immunoprecipitates from five-week-old plants expressing eIF4E-Myc or the phospho-dead mutant eIF4E-T6A-Myc. MPK4
is detected in anti-Myc immunoprecipitates. Immunoblots of input and anti-Myc immunoprecipitates are probed with antiMPK4 and anti-Myc antibodies.
eif4e mutants does not exhibit autoimmunity similar to mpk4
eif4e plants exhibit normal growth and are indistinguishable from Col-0 plants (Fig. 4A).
Nevertheless, given that MPK4 is an important regulator of defense, we tested whether eIF4E may
also be involved in immunity. To this end, we first examined for altered bacterial resistance by
syringe infiltrating pst DC3000, into eif4e leaf tissue. However, there was no noticeable difference
in bacterial growth after four days when compared to Col-0 (Fig. 4B). Given that eIF4E and
eIF(iso)4E can complement each other, we hypothesized that if phosphorylation of eIF4E by MPK4
is important for defense then the non-phophorylatable eIF4E-T6A might dominantly inhibit defense
88
regulatory functions normally mediated by eIF4E. We therefore also tested for altered resistance in
the eIF4E-T6A complementing lines by syringe infiltrating pst DC3000, but no noticeable
difference was observed in bacterial growth after four-days post infiltration between these plants,
Col-0 and eIF4E-Myc complementing lines (Fig. 4B)
This implies that phosphorylation of eIF4E-Thr6 is not critical for normal defense responses against
pst DC3000, nonetheless it is possible that eIF4E might be involved in defense responses against
other types of pathogens.
Figure 4: eif4e does not induce autoimmunity
(A) Photographs of five-week old plants, genotypes are as indicated.
(B)Four-week old plants we infiltrated with pst DC3000 (OD600=0.001). Bacteria were syringe-infiltrated and samples taken
4 days post-infiltration. Data are shown as log10-transformed colony-forming units/cm2 leaf tissue. Standard error of the
mean is indicated by errors bars (n = 4).
89
ClYVV susceptibility is not dependent on eIF4E phosphorylation
Potyviruses including ClYVV and TuMV seize the host translational machinery to proliferate
themselves. ClYVV and TuMV depends on interactions with eIF4E and eIF(iso)4E respectively to
initiate translation of their genomic RNA (Duprat et al., 2002; Sato et al., 2005). Thus, mutations in
these genes can result in recessive resistance against viruses. We questioned whether
phosphorylation of eIF4E by MPK4 would affect eIF4E activity and thus have impact on virus
resistance against ClYVV. Given that mpk4 exhibits autoimmunity and suppression by summ2 is
only partial, we tested for ClYVV susceptibility in mkk1/2-summ2 plants. MKK1/2 are the upstream
MPKKs of MPK4. The mkk1/2 double mutant reassembles the phenotype of mpk4, however this
phenotype is fully suppressed by summ2, while MPK4 is not activated by PAMP perception. Col-0
was susceptible to both ClYVV and TuMV, whereas respectively eif4e and eif(iso)4e were resistant
(Fig. 5A+B).
Figure 5: ClYVV and TuMV susceptibility iss not dependent on MPK4 activation.
Five-week-old plants were inoculated with (A) ClYVV or (B) TuMV. Virus accumulation was assayed 24 dpi by ELISA.
Error bars indicate SEM (n=6). The red line indicates threshold for susceptibility equal to three times the mean value for
uninfected controls (n=3)
90
Both summ2 and mkk1/2-summ2 were susceptible to ClYVV and TuMV, thus lack of MPK4
activity do not impair eIF4E and eIF(iso)4E mediated ClYVV or TuMV proliferation.
It is possible that other kinases can phosphorylate eIF4E-Thr6 in vivo, we therefore tested ClYVV
susceptibility in the complementing eIF4E-T6A-Myc line. Plants expressing either eIF4E-Myc or
eIF4E-T6A-Myc similar to Col-0 were susceptible to ClYVV (Fig. 6). We therefore conclude that
absence of MPK4 meditated eIF4E phosphorylation does not affect eIF4E mediated ClYVV
proliferation. Nevertheless, phosphorylation of eIF4E by MPK4 may regulate its activity in assays
not tested here and might still be important for proper regulation.
Figure 6: ClYVV susceptibility is not dependent phosphorylation eIF4E Thr6.
Five-week-old plants were inoculated with (A) ClYVV. Virus accumulation was assayed after 24 days by ELISA. Error bars
indicate SEM (n=6). The red line indicates threshold for susceptibility equal to three times the mean value for uninfected
controls (n=3)
91
MPK4 and eIF4E interact in the nucleus.
Apart from its function in translational initiation eIF4E also function in export of mRNAs out of the
nucleus, in both yeast and humans up to 68% of eIF4E is present in the nucleus (Iborra et al., 2001;
Strudwick and Borden, 2002). eIF4E interacts with a secondary structure in the 3’UTR of specific
mRNA and promotes their export out of the nucleus (Culjkovic et al., 2006). We were interested in
where the MPK4-eIF4E interaction occurs in the cell and therefore employed a bimolecular
fluorescence complementation assay (BiFC). We cloned MPK4 with a C-terminal fusion of nYFP
and eIF4E with a C-terminal fusion of cYFP, both under control of a 35S promoter and coexpressed them transiently in Nicotiana benthamiana. When expressed together we detected distinct
signal in the nucleus indicating reconstitution of YFP while no signal was observed when eIF4EcYFP were co-expressed MPK6-nYFP (Fig. 7). When eIF4E-T6A-nYFP was co-expressed with
MPK4-cYFP we observed an even more intense nuclear fluorescent signal compared to the eIF4EnYFP (Fig. 7), which is in agreement with the CoIP data (Fig. 3). Thus, MPK4 and eIF4E interacts
in the nucleus and the non-phophorylatable eIF4E-T6A gives rise to a stronger interaction.
Discussion
In this study, we characterize the translational initiation factor eIF4E as an interactor (Fig. 3+7) and
substrate (Fig. 2) of the immune regulating kinase MPK4. MPK4 specifically phosphorylates
eIF4E-Thr6, which align well with a phosphorylation site in human eIF4E (Fig. 1). Although no
function has been assigned to this phosphorylation, it is possible that its function can be regulatory
or serve as a (de)stabilizing factor. Given that MPK4 is an important immune regulator, we tested a
possible role of eIF4E functioning in immunity. However we did not find evidence of eIF4E
mediated resistance against pst DC3000 (Fig. 4B). Similar we did not find MPK4 activity (Fig. 5A)
or phosphorylation of eIF4E-Thr6 (Fig. 6) important in conferring resistance against the potyvirys
ClYVV that normally rely on eIF4E in Arabidopsis for its own proliferation. However, since we
only tested a very limited set of pathogens we cannot exclude the possibility that phosphorylation of
eIF4E can affect immunity in Arabidopsis.
92
Figure 7: eIF4E interacts with MPK4 in the nucleus
Confocal images of yellow fluorescent protein complementation. eIF4E-cYFP and eIF4E-T6A-cYFP were was transiently coexpressed with either MPK4-nYFP or MPK6-nYFP in N. benthamiana. Confocal microscopy on leaf disks was conducted 2
days post-infiltration. Merged pictures show overlay of brightfield, chlorophyll and eYFP. Blue arrows indicate nuclei and
the scale bar corresponds to 10 µm.
93
The interaction between MPK4 and eIF4E occurs in the nucleus (Fig. 7). Up to 68% of eIF4E is
nuclei in both yeast and humans, in the nucleus, eIF4E serves different role than cytoplasmic
translational initiation, instead eIF4E function in specific mRNA transport out of the nucleus
(Culjkovic et al., 2006; Iborra et al., 2001; Strudwick and Borden, 2002). In humans overexpression
of eIF4E leads to enhanced translation of a subset of mRNAs with long highly structured 5’UTRs.
Interestingly these mRNAs are not upregulated at the transcriptional level, but rather actively
transported out of the nucleus to the cytoplasm for translation (Culjkovic et al., 2007). In humans
eIF4E mediated mRNA transport out of the nucleus depends on specific secondary structures in the
3’UTR distinct from bulk mRNA (Culjkovic et al., 2006). Thus, both structured 5’ and 3’UTR
influence eIF4E mediated translation (Osborne and Borden, 2015). We show that in Arabidopsis,
MPK4 interacts with eIF4E in the nucleus. It is therefore appealing to hypothesize that MKP4 can
somehow regulate eIF4E mediated mRNA transport through phosphorylation. However, to this end
we have no data showing eIF4E mediated transport in response to for example immunity. Further
examination of the relationship between MPK4 and eIF4E are needed to clarify this hypothesis.
Recently eIF4E has also been implicated in non-sense mediated decay (NMD) of mRNA in
mammalian cells in response to insulin (Park et al., 2015). eIF4E seemingly interacts with the NMD
factor Upstream Frameshift protein 1 (UPF1) as well as the general decapping factor Decapping 1
(DCP1) to promote NMD of eIF4E associated mRNAs (Park et al., 2015). We have recently shown
that MPK4 interacts with and phosphorylates the mRNA decay factor Protein Associated with
Topoisomerase II (PAT1) (Roux et al., 2015). PAT1 interacts with the cytoplasmic immunereceptor SUMM2 which in absence of PAT instigates ETI (Roux et al., 2015). Similar, ETI is also
observed in NMD deficient mutants including upf1, suggesting a role of NMD in immunity (Jeong
et al., 2011; Rayson et al., 2012). Any possible link between MPK4 mediated phosphorylation of
eIF4E in Arabidopsis and mRNA decay are however still unclear and will be a subject for further
investigation.
Materials and methods
Plants growth conditions
Arabidopsis thaliana Col-0 accession was used as wild-type in all experiments, except if otherwise
stated. Salk lines were obtained from the Nottingham Arabidopsis Seed Centre (NASC). Lines
obtained from NASC include SALK_101850 (aoc3), SALK_145583 (eif4e). mkks1/2 summ2 was
a kind gift from Yuelin Zhang. Arabidopsis plants were grown in 9 × 9 cm pots at 22°C with a 8-h
94
photoperiod, or on plates containing Murashige–Skoog (MS) salts medium (Duchefa), 1% sucrose
and 1% agar with a 16-h photoperiod.
SDS-PAGE and immunoblotting
SDS-tris/glycine or bis/tris gels were prepared with either 8, 10 or 12% acrylamide. Proteins were
loaded and gels run at 100–150 V until the dye front reached the end of the gel. Gels were next
electroblotted onto PVDF membranes (GE Healthcare) by wet transfer for 1h at 70 V. Membranes
were rinsed in TBS and blocked for 30min in 5% (w/v) non-fat milk in TBS-Tween (0.1% (v/v)).
Membranes were incubated with primary antibodies for 1 h to overnight. Membranes were washed
3 × 10 min in TBS-Tween (0.1%) before 1-h incubation in secondary antibodies. Membrnaes were
incubated in chemiluminescent substrate (100 mM Tris/HCl pH 8.8, 1.25 mM luminol, 5.3 mM
hydrogenperoxide and 2 mM 4IPBA)(Haan and Behrmann, 2007) before exposure to a Sony a7S
DSLR camera(Khoury et al., 2010). For AP-conjugated secondary antibodies membranes were
incubated in NBT/BCIP (Roche) until bands were visible. For probing immunoblots with multiple
antibodies, stripping was carried out using Restore Western Blot Stripping Buffer (Pierce) for 15–
30 min, following three washes with TBS-Tween (0.1% (v/v)).
Antibodies
Primary antibodies were diluted in 1% (w/v) non-fat milk in TBS-Tween (0.1% (v/v)) to the
following dilutions:
Anti
Host
Dilution
Suplier
Myc
Mouse
1:1000
Santa Cruz
GFP
Mouse
1:1000
Santa Cruz
HA
Mouse
1:1000
Santa Cruz
MPK3
Rabbit
1:2000
Sigma
MPK4
Rabbit
1:2000
Sigma
MPK6
Rabbit
1:2000
Sigma
CSD2
Rabbit
1:4000
Agrisera
PR1
Rabbit
1:4000
Agrisera
DCP1
Rabbit
1:1000
Gift from
eIF4E
Rabbit
1:5000
Gift from Jean-Luc Gallois
95
Secondary antiboides were diluted in 1% (w/v) non-fat milk in TBS-Tween (0.1% (v/v)) to the
following dilutions:
anti-mouse-HRP 1:10.000 (Promega)
anti-Rabbit-HRP 1:10.000 (Promega)
ProteinA-HRP 1:1000 (GE Healthcare)
Bacterial infection assays
Pseudomonas syringae pv. tomato DC3000 (Pst DC3000) strains were grown in overnight culture in
NYG medium supplemented with appropriate antibiotics. Cells were harvested by centrifugation
and pellets resuspended in sterile water to indicated concentrations. Bacteria were infiltrated with a
needleless 1-mm syringe into adult leafs and maintained in growth chambers for 3-4 days. Samples
were taken using a cork-borer (6.5 mm) to cut leaf disks from infected leafs. Leaf disks were
ground in water, serially diluted and plated on NYG plates with appropriate selection. Plates were
incubated at 28°C and colonies counted 2–3 days later.
Transient expression in Nicotiana benthamiana
Agrobacterium tumefaciens strains were grown in LB medium supplemented with appropriate
antibiotics overnight. Cultures were spun down and resuspended in 10 mM MgCl2 to OD600 = 0.8.
For co-infiltrations bacterial cultures were mixed 1:1. Bacterial cultures were syringe-infiltrated into
3-week-old N. benthamiana leaves. Samples for protein extraction were harvested 2 days postinfiltration.
Protein extraction and immunoprecipitation in Nicotiana benthamiana and Arabidopsis
Plant material were ground in liquid nitrogen and extraction buffer [10 mM phosphate pH 7.2, 150
mM NaCl, 5 mM DTT, 0.1 % Nonidet P-40, supplemented with EDTA-free protease inhibitor cocktail
form Roche or 110 mM potassium acetate, 100 mM NaCl, 2 mM MgCl, 0.5%(v/v) TritonX-100
supplemented with EDTA-free protease inhibitor cocktail form Roche]
Samples were clarified by 10 min centrifugation at 4°C and 13,000 rpm. Supernatants were adjusted
to 2 mg/ml protein and incubated 1-2 h at 4°C with GFPTrap-A beads (Chromotek), anti-Myc
magnetic beads (Miltenyi Biotech) or EZview™ Red Protein A Affinity Gel supplemented with
specific antibodies. Following incubation, beads were washed four times with extraction buffer,
before adding 6XSDS loading buffer [300 mM Tris HCL pH 6.8, 300 mM DTT, 6% SDS, 6 mM
EDTA, 0.03%(w/v) bromphenol blue, 60%(v/v) glycerol] and heating at 70°C for 10 min.
96
Purification of X7 polymerase
E.coli containing a plasmid encoding His tagged PfuX7 under control of a IPTG inducible promotor
was grown in liquid LB media (Nørholm, 2010). At OD600=0.4 IPTG was added to a final
concentration of 0.5mM. Cells where harvested by centrifugation and re-suspended in lysis buffer []
and lysed by 3-4 freeze/thaw cycles. Soluble protein were purified nickel spin columns () according
to manufactures instruction. The enzyme was desalted by dialysis o.n. against storage buffer [50%
glycerol, 100 mM Tris/HCl pH 8.0, 0.2 mM EDTA, 2 mM DTT, 0.2% NP40, 0.2% Tween 20.] and
stored at -20C.
Genotyping
Leaf tissue was used for Edwards DNA extraction (Edwards et al., 1991). PfuX7 polymerase were
used for genotyping PCR reactions in with a final concentration of 0.4 mM dATP, 0.4 mM dTTP,
0.4 mM dCTP, 0.4 mM dGTP, 20 mM Tris/HCl pH 8.8, 10 mM KCl, 6 mM (NH4)2SO4, 2 mM
MgSO4, 0.1 mg/ml BSA and 0.1% Triton X-100 (Nørholm, 2010). Cycling conditions included a
60 second 96°C template denaturation step followed by 30 cycles of 15 second 56-65°C annealing
step and a 15 second long primer extension step at 72°C. PCR products were electrophoresed on a
1% agarose gel in 0.5X TBE [45 mM Tris-borate/1 mM EDTA] with etBr and PCR products were
visualized by UV light. Primers designed using the iSect tool
(http://signal.salk.edu/isectprimers.html) were used to genotype lines for homozygosity.
Cloning and transgenic lines
Phusion Hot Start II High-Fidelity polymerase (Thermo) or PfuX7 (Nørholm, 2010) were used to
amplify PCR products for all subsequent cloning reactions.
MPK6
MPK6 was cloned into pETNR D-TOPO (Invitrogen) from genomic DNA without stop codon with
the following primers fwd “5’caccATGGACGGTGGTTCAGGTCA” reverse
“5’TTGCTGATATTCTGGATTGAAAGC”
AOC3
AOC3 without signal peptide was cloned into pETNR pOPINf for expression in E.coli from cDNA
DNA without start codon with the following primers fwd
“5’AAGTTCTGTTTCAGGGCCCGAGACCAAGTAAGATCCAAGAACTAA” reverse
“5’ATGGTCTAGAAAGCTTTAATTAGTAAAGTTACTTATAACTCCA”. For cloning AOC3S100A the following start primer where used
“5’AAGTTCTGTTTCAGGGCCCGCCAAGTAAGATCCAAGAACTAAACGTGTATGAGCTC
97
AACGAAGGAGATAGAAACgcTCCAGC”. For cloning AOC3-T235A the following revers
primer were used
“5’ATGGTCTAGAAAGCTTTAATTAGTAAAGTTACTTATAACTCCACTAGGCTCCATCGC
CTTAGCTTCCGGTGCCGGTTTCACATCCTTCGACGGCGCAACCGCCG”.
eIF4E
eIF4E was cloned into pETNR D-TOPO from genomic DNA without stop codon with the following
primers fwd “5’caccATGGCGGTAGAAGACACTCC”reverse
“5’AGCGGTGTAAGCGTTCTTTG “. For cloning eIF4E-T6A the following fwd primer was used
“5’caccATGGCGGTAGAAGACgCTCCCAAATC”. eIF4E was cloned into pOPINf for expression
in E.coli from cDNA without start codon with the following primers fwd
“5’AAGTTCTGTTTCAGGGCCCGGCGGTAGAAGACACTCCCAAA” reverse
“5’ATGGTCTAGAAAGCTTTAAGCGGTGTAAGCGTTCTTTGC”. For cloning eIF4E-T6A the
following start primer where used
“5’AAGTTCTGTTTCAGGGCCCGGCGGTAGAAGACgCTCCCAAATC”.
MPK6, AOC3 and eIF4E were recombined into binary gateway destination vectors for expression
in plants using LR Clonase II (Invitrogen). For expression of C-terminally tagged GFP, HA, Myc,
nYFP and cYFP construct pENTR clones were recombined into pGWB605, pGWB615, pGWB617,
pcYCW and pnYGW respectively (HINO et al., 2011; Nakagawa et al., 2007).
Clones were transformed into A. tumefaciens strain GV3101 and used for transient expression in N.
benthamiana or floral dipping in Arabidopsis.
Recombinant protein purification and in vitro kinase assays
For in vitro experiments, PAT1 protein was purified from E. coli. The AOC3 and eIF4E CDS was
cloned into pET15b (for an N-terminal His fusion) and transformed into E. coli BL21 (pLysS).
Protein expression was induced by overnight treatment with 0.5 mM IPTG at 37°C, added to cells
at OD600 = 0.6. Proteins were purified under native conditions Ni-NTA spin columns (Qiagen). For
kinase assay see Roux et al. (2015).
Virus inoculation and detection by ELISA
TuMV and ClYVV were propagated on turnip (Brassica rapa) and Nicotiana benthamiana before
virus inoculating of five-week-old Arabidopsis plants. Virus accumulation was assayed after 24
days by ELISA using respective viral antisera (DSMZ). Six plants per genotype were used in each
98
experiment. The susceptibility threshold is indicated for each experiment as a red line and refers to
an absorbance value at 405 nm in ELISA equal to three times the mean value for uninfected
controls.
References
Callot, C., and Gallois, J.-L. (2014). Pyramiding resistances based on translation initiation factors in
Arabidopsis is impaired by male gametophyte lethality. Plant Signal. Behav. 9.
doi:10.4161/psb.27940.
Charron, C., Nicolaï, M., Gallois, J.-L., Robaglia, C., Moury, B., Palloix, A., et al. (2008). Natural
variation and functional analyses provide evidence for co-evolution between plant eIF4E
and potyviral VPg. Plant J. 54, 56–68. doi:10.1111/j.1365-313X.2008.03407.x.
Cui, H., Tsuda, K., and Parker, J. E. (2015). Effector-Triggered Immunity: From Pathogen
Perception to Robust Defense. Annu. Rev. Plant Biol. 66, 487–511. doi:10.1146/annurevarplant-050213-040012.
Culjkovic, B., Topisirovic, I., and Borden, K. L. B. (2007). Controlling Gene Expression through
RNA Regulons: The Role of the Eukaryotic Translation Initiation Factor eIF4E. Cell Cycle
6, 65–69. doi:10.4161/cc.6.1.3688.
Culjkovic, B., Topisirovic, I., Skrabanek, L., Ruiz-Gutierrez, M., and Borden, K. L. B. (2006).
eIF4E is a central node of an RNA regulon that governs cellular proliferation. J. Cell Biol.
175, 415–426. doi:10.1083/jcb.200607020.
Deslandes, L., and Rivas, S. (2012). Catch me if you can: bacterial effectors and plant targets.
Trends Plant Sci. 17, 644–655. doi:10.1016/j.tplants.2012.06.011.
Diaz-Pendon, J. A., Truniger, V., Nieto, C., Garcia-Mas, J., Bendahmane, A., and Aranda, M. A.
(2004). Advances in understanding recessive resistance to plant viruses. Mol. Plant Pathol.
5, 223–233. doi:10.1111/j.1364-3703.2004.00223.x.
Duprat, A., Caranta, C., Revers, F., Menand, B., Browning, K. S., and Robaglia, C. (2002). The
Arabidopsis eukaryotic initiation factor (iso)4E is dispensable for plant growth but required
for susceptibility to potyviruses. Plant J. 32, 927–934. doi:10.1046/j.1365313X.2002.01481.x.
Durek, P., Schmidt, R., Heazlewood, J. L., Jones, A., MacLean, D., Nagel, A., et al. (2010).
PhosPhAt: the Arabidopsis thaliana phosphorylation site database. An update. Nucleic
Acids Res. 38, D828–D834. doi:10.1093/nar/gkp810.
Edwards, K., Johnstone, C., and Thompson, C. (1991). A simple and rapid method for the
preparation of plant genomic DNA for PCR analysis. Nucleic Acids Res. 19, 1349.
99
Eto, I. (2010). Upstream molecular signaling pathways of p27(Kip1) expression: effects of 4hydroxytamoxifen, dexamethasone, and retinoic acids. Cancer Cell Int. 10, 3.
doi:10.1186/1475-2867-10-3.
Franz-Wachtel, M., Eisler, S. A., Krug, K., Wahl, S., Carpy, A., Nordheim, A., et al. (2012). Global
detection of protein kinase D-dependent phosphorylation events in nocodazole-treated
human cells. Mol. Cell. Proteomics MCP 11, 160–170. doi:10.1074/mcp.M111.016014.
Furic, L., Rong, L., Larsson, O., Koumakpayi, I. H., Yoshida, K., Brueschke, A., et al. (2010).
eIF4E phosphorylation promotes tumorigenesis and is associated with prostate cancer
progression. Proc. Natl. Acad. Sci. 107, 14134–14139. doi:10.1073/pnas.1005320107.
Gross, J. D., Moerke, N. J., von der Haar, T., Lugovskoy, A. A., Sachs, A. B., McCarthy, J. E. G.,
et al. (2003). Ribosome Loading onto the mRNA Cap Is Driven by Conformational
Coupling between eIF4G and eIF4E. Cell 115, 739–750. doi:10.1016/S00928674(03)00975-9.
Haan, C., and Behrmann, I. (2007). A cost effective non-commercial ECL-solution for Western blot
detections yielding strong signals and low background. J. Immunol. Methods 318, 11–19.
doi:10.1016/j.jim.2006.07.027.
HINO, T., TANAKA, Y., KAWAMUKAI, M., NISHIMURA, K., MANO, S., and NAKAGAWA,
T. (2011). Two Sec13p Homologs, AtSec13A and AtSec13B, Redundantly Contribute to the
Formation of COPII Transport Vesicles in Arabidopsis thaliana. Biosci. Biotechnol.
Biochem. 75, 1848–1852. doi:10.1271/bbb.110331.
Hornbeck, P. V., Zhang, B., Murray, B., Kornhauser, J. M., Latham, V., and Skrzypek, E. (2015).
PhosphoSitePlus, 2014: mutations, PTMs and recalibrations. Nucleic Acids Res. 43, D512–
D520. doi:10.1093/nar/gku1267.
Iborra, F. J., Jackson, D. A., and Cook, P. R. (2001). Coupled Transcription and Translation Within
Nuclei of Mammalian Cells. Science 293, 1139–1142. doi:10.1126/science.1061216.
Jackson, R. J., Hellen, C. U. T., and Pestova, T. V. (2010). The mechanism of eukaryotic translation
initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11, 113–127.
doi:10.1038/nrm2838.
Jeong, H.-J., Kim, Y. J., Kim, S. H., Kim, Y.-H., Lee, I.-J., Kim, Y. K., et al. (2011). NonsenseMediated mRNA Decay Factors, UPF1 and UPF3, Contribute to Plant Defense. Plant Cell
Physiol. 52, 2147–2156. doi:10.1093/pcp/pcr144.
Khoury, M. K., Parker, I., and Aswad, D. W. (2010). Acquisition of chemiluminescent signals from
immunoblots with a digital SLR camera. Anal. Biochem. 397, 129–131.
doi:10.1016/j.ab.2009.09.041.
Kropiwnicka, A., Kuchta, K., Lukaszewicz, M., Kowalska, J., Jemielity, J., Ginalski, K., et al.
(2015). Five eIF4E isoforms from Arabidopsis thaliana are characterized by distinct features
of cap analogs binding. Biochem. Biophys. Res. Commun. 456, 47–52.
doi:10.1016/j.bbrc.2014.11.032.
100
Li, H., Ren, Z., Kang, X., Zhang, L., Li, X., Wang, Y., et al. (2009). Identification of tyrosinephosphorylated proteins associated with metastasis and functional analysis of FER in human
hepatocellular carcinoma cells. BMC Cancer 9, 366. doi:10.1186/1471-2407-9-366.
Lukhele, S., Bah, A., Lin, H., Sonenberg, N., and Forman-Kay, J. D. (2013). Interaction of the
Eukaryotic Initiation Factor 4E with 4E-BP2 at a Dynamic Bipartite Interface. Structure 21,
2186–2196. doi:10.1016/j.str.2013.08.030.
Mayberry, L. K., Allen, M. L., Nitka, K. R., Campbell, L., Murphy, P. A., and Browning, K. S.
(2011). Plant Cap-binding Complexes Eukaryotic Initiation Factors eIF4F and eIFISO4F
MOLECULAR SPECIFICITY OF SUBUNIT BINDING. J. Biol. Chem. 286, 42566–
42574. doi:10.1074/jbc.M111.280099.
Merrick, W. C. (2015). eIF4F: A Retrospective. J. Biol. Chem. 290, 24091–24099.
doi:10.1074/jbc.R115.675280.
Nakagawa, T., Suzuki, T., Murata, S., Nakamura, S., Hino, T., Maeo, K., et al. (2007). Improved
Gateway binary vectors: high-performance vectors for creation of fusion constructs in
transgenic analysis of plants. Biosci. Biotechnol. Biochem. 71, 2095–2100.
Nørholm, M. H. (2010). A mutant Pfu DNA polymerase designed for advanced uracil-excision
DNA engineering. BMC Biotechnol. 10, 21. doi:10.1186/1472-6750-10-21.
Osborne, M. J., and Borden, K. L. B. (2015). The eukaryotic translation initiation factor eIF4E in
the nucleus: taking the road less traveled. Immunol. Rev. 263, 210–223.
doi:10.1111/imr.12240.
Park, J., Ahn, S., Jayabalan, A. K., Ohn, T., Koh, H. C., and Hwang, J. Insulin signaling augments
eIF4E-dependent nonsense-mediated mRNA decay in mammalian cells. Biochim. Biophys.
Acta BBA - Gene Regul. Mech. doi:10.1016/j.bbagrm.2015.12.006.
Patrick, R. M., Browning, K. S., Patrick, R. M., and Browning, K. S. (2012). The eIF4F and
eIFiso4F Complexes of Plants: An Evolutionary Perspective, The eIF4F and eIFiso4F
Complexes of Plants: An Evolutionary Perspective. Int. J. Genomics Int. J. Genomics 2012,
2012, e287814. doi:10.1155/2012/287814, 10.1155/2012/287814.
Patrick, R. M., Mayberry, L. K., Choy, G., Woodard, L. E., Liu, J. S., White, A., et al. (2014). Two
Arabidopsis Loci Encode Novel Eukaryotic Initiation Factor 4E Isoforms That Are
Functionally Distinct from the Conserved Plant Eukaryotic Initiation Factor 4E. Plant
Physiol. 164, 1820–1830. doi:10.1104/pp.113.227785.
Pearson, G., Robinson, F., Beers Gibson, T., Xu, B., Karandikar, M., Berman, K., et al. (2001).
Mitogen-Activated Protein (MAP) Kinase Pathways: Regulation and Physiological
Functions. Endocr. Rev. 22, 153–183. doi:10.1210/edrv.22.2.0428.
Peter, D., Igreja, C., Weber, R., Wohlbold, L., Weiler, C., Ebertsch, L., et al. (2015). Molecular
Architecture of 4E-BP Translational Inhibitors Bound to eIF4E. Mol. Cell 57, 1074–1087.
doi:10.1016/j.molcel.2015.01.017.
Popescu, S. C., Popescu, G. V., Bachan, S., Zhang, Z., Gerstein, M., Snyder, M., et al. (2009).
MAPK Target Networks in Arabidopsis Thaliana Revealed Using Functional Protein
Microarrays. Genes Dev. 23, 80–92. doi:10.1101/gad.1740009.
101
Rasmussen, M. W., Roux, M., Petersen, M., and Mundy, J. (2012). MAP kinase cascades in
Arabidopsis innate immunity. Plant Proteomics 3, 169. doi:10.3389/fpls.2012.00169.
Rayson, S., Arciga-Reyes, L., Wootton, L., De Torres Zabala, M., Truman, W., Graham, N., et al.
(2012). A Role for Nonsense-Mediated mRNA Decay in Plants: Pathogen Responses Are
Induced in Arabidopsis thaliana NMD Mutants. PLoS ONE 7, e31917.
doi:10.1371/journal.pone.0031917.
Riechmann, J. L., Lain, S., and Garcia, J. A. (1992). Highlights and prospects of potyvirus
molecular biology. J. Gen. Virol. 73, 1–16. doi:10.1099/0022-1317-73-1-1.
Rodriguez, C. M., Freire, M. A., Camilleri, C., and Robaglia, C. (1998). The Arabidopsis thaliana
cDNAs coding for eIF4E and eIF(iso)4E are not functionally equivalent for yeast
complementation and are differentially expressed during plant development. Plant J. 13,
465–473. doi:10.1046/j.1365-313X.1998.00047.x.
Roux, M. E., Rasmussen, M. W., Palma, K., Lolle, S., Regué, À. M., Bethke, G., et al. (2015). The
mRNA decay factor PAT1 functions in a pathway including MAP kinase 4 and immune
receptor SUMM2. EMBO J., e201488645. doi:10.15252/embj.201488645.
Rychlik, W., Russ, M. A., and Rhoads, R. E. (1987). Phosphorylation site of eukaryotic initiation
factor 4E. J. Biol. Chem. 262, 10434–10437.
Sanfaçon, H. (2015). Plant Translation Factors and Virus Resistance. Viruses 7, 3392–3419.
doi:10.3390/v7072778.
Sato, M., Nakahara, K., Yoshii, M., Ishikawa, M., and Uyeda, I. (2005). Selective involvement of
members of the eukaryotic initiation factor 4E family in the infection of Arabidopsis
thaliana by potyviruses. FEBS Lett. 579, 1167–1171. doi:10.1016/j.febslet.2004.12.086.
Schütz, P., Bumann, M., Oberholzer, A. E., Bieniossek, C., Trachsel, H., Altmann, M., et al. (2008).
Crystal structure of the yeast eIF4A-eIF4G complex: An RNA-helicase controlled by
protein–protein interactions. Proc. Natl. Acad. Sci. 105, 9564–9569.
doi:10.1073/pnas.0800418105.
Sharma, K., D’Souza, R. C. J., Tyanova, S., Schaab, C., Wiśniewski, J. R., Cox, J., et al. (2014).
Ultradeep human phosphoproteome reveals a distinct regulatory nature of Tyr and Ser/Thrbased signaling. Cell Rep. 8, 1583–1594. doi:10.1016/j.celrep.2014.07.036.
Strudwick, S., and Borden, K. L. B. (2002). The emerging roles of translation factor eIF4E in the
nucleus. Differentiation 70, 10–22. doi:10.1046/j.1432-0436.2002.700102.x.
Topisirovic, I., Svitkin, Y. V., Sonenberg, N., and Shatkin, A. J. (2011). Cap and cap-binding
proteins in the control of gene expression. Wiley Interdiscip. Rev. RNA 2, 277–298.
doi:10.1002/wrna.52.
Truniger, V., and Aranda, M. A. (2009). “Chapter 4 - Recessive Resistance to Plant Viruses,” in
Advances in Virus Research Natural and Engineered Resistance to Plant Viruses, Part I., ed.
G. L. and J. P. Carr (Academic Press), 119–231. Available at:
http://www.sciencedirect.com/science/article/pii/S0065352709075046 [Accessed December
17, 2015].
102
Van Hoof, D., Muñoz, J., Braam, S. R., Pinkse, M. W. H., Linding, R., Heck, A. J. R., et al. (2009).
Phosphorylation dynamics during early differentiation of human embryonic stem cells. Cell
Stem Cell 5, 214–226. doi:10.1016/j.stem.2009.05.021.
Volpon, L., Osborne, M. J., Topisirovic, I., Siddiqui, N., and Borden, K. L. (2006). Cap-free
structure of eIF4E suggests a basis for conformational regulation by its ligands. EMBO J.
25, 5138–5149. doi:10.1038/sj.emboj.7601380.
Wang, A., and Krishnaswamy, S. (2012). Eukaryotic translation initiation factor 4E-mediated
recessive resistance to plant viruses and its utility in crop improvement. Mol. Plant Pathol.
13, 795–803. doi:10.1111/j.1364-3703.2012.00791.x.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., et al. (2012). Disruption of PAMPInduced MAP Kinase Cascade by a Pseudomonas syringae Effector Activates Plant
Immunity Mediated by the NB-LRR Protein SUMM2. Cell Host Microbe 11, 253–263.
doi:10.1016/j.chom.2012.01.015.
Zhou, H., Di Palma, S., Preisinger, C., Peng, M., Polat, A. N., Heck, A. J. R., et al. (2013). Toward
a comprehensive characterization of a human cancer cell phosphoproteome. J. Proteome
Res. 12, 260–271. doi:10.1021/pr300630k.
103
Concluding remarks
MPK4 was initially characterized as a negative regulator of defense, due to MPK4 the loss-offunction mutant constitutively induction of defense responses (Petersen et al., 2000). MPK4 is
activated by perception of the flagellin-derived flg22 peptide by the pathogen recognition receptor
FLS2 (Rodriguez et al., 2010). FLS2 and its co-receptor BAK1 sequentially signals the activation of
the three tiered MAPK signaling cascade comprising MEKK1, MKK1/2 and MPK4 (Chinchilla et
al., 2007; Gao et al., 2008; Qiu et al., 2008a). Inactive MPK4 is bound to MKS1 and the WRKY33
transcription factor, upon activation, MPK4 phosphorylates MKS1, which release WRKY33 from
the complex to induce transcription of the defense gene PAD3 (Andreasson et al., 2005; Qiu et al.,
2008a, 2008b). The function of MPK4 as a negative regulator of defense is in contrast to MPK4
positively inducing transcription of PAD3. The MPK4 loss-of-function defense phenotype is
suppressed my loss-of-function of EDS1 (Brodersen et al., 2006), which is essential for fully
functional effector-triggered immunity (Falk et al., 1999). This indicated that the mpk4 defense
phenotype was due to inappropriate activation of ETI and not due to loss of negative suppression of
defense. Zhang et al. (2012) identified the R-protein SUMM2, which surveillance the MPK4
pathway and induce ETI upon disruption of this pathway. This endorse the hypothesis of the mpk4
defense phenotype arising from inappropriate activation of ETI and enables the analysis of the true
in vivo function of MPK4 without pleiotropic effects in the mpk4 loss-of-function mutant.
In the previous sections, we have described and characterized three novel MPK4 substrates (i)
PAT1 a protein functioning in decapping of mRNAs which was previously undescribed in plants.
(ii) AOC3 functioning in JA biosynthesis and (iii) eIF4E with a dual function in both translational
initiation and export of specific mRNAs out of the nucleus. Both PAT1 and AOC3 seems to be
regulated through immune signaling whereas we were not able to position eIF4E in immunity.
PAT1 is particular interesting given that its loss-of-function defense phenotype is dependent on
SUMM2. PAT1 and SUMM2 interacts in planta and it is possible that SUMM2 guards a specific
branch in MPK4 mediated immunity comprising PAT1. Equally interesting is the link between
MPK4 and the JA biosynthesis component AOC3. The possible regulatory role that MPK4 have on
AOC3 is still not clear, but the in vivo interaction between these two proteins prompted us to test if
MPK4 is present in chloroplast. To this end, we have shown that MPK4, as the first MAPK so far
shown, associates with the chloroplasts. For eIF4E, we have not found a putative role in MPK4
mediated defense, however we localized the interaction to take place in the nucleus. Given the dual
function of eIF4E in both translation and mRNA export, it is still not clear what its connection to
MPK4 is, but it is possible that MPK4 regulate either translation or mRNA export or both in
104
response to pathogens. With the findings presented here in this thesis, MPK4 is evidently a
multifunctional MAPK functioning spatial and temporal in numerous cellular processes.
References
Andreasson, E., Jenkins, T., Brodersen, P., Thorgrimsen, S., Petersen, N. H., Zhu, S., et al. (2005).
The MAP kinase substrate MKS1 is a regulator of plant defense responses. EMBO J 24,
2579–2589. doi:10.1038/sj.emboj.7600737.
Brodersen, P., Petersen, M., Nielsen, H. B., Zhu, S., Newman, M.-A., Shokat, K. M., et al. (2006).
Arabidopsis MAP kinase 4 regulates salicylic acid- and jasmonic acid/ethylene-dependent
responses via EDS1 and PAD4. Plant J. 47, 532–546. doi:10.1111/j.1365313X.2006.02806.x.
Chinchilla, D., Zipfel, C., Robatzek, S., Kemmerling, B., Nurnberger, T., Jones, J. D. G., et al.
(2007). A flagellin-induced complex of the receptor FLS2 and BAK1 initiates plant defence.
Nature 448, 497–500. doi:10.1038/nature05999.
Falk, A., Feys, B. J., Frost, L. N., Jones, J. D. G., Daniels, M. J., and Parker, J. E. (1999). EDS1, an
essential component of R gene-mediated disease resistance in Arabidopsis has homology to
eukaryotic lipases. Proc. Natl. Acad. Sci. U. S. A. 96, 3292–3297.
Gao, M., Liu, J., Bi, D., Zhang, Z., Cheng, F., Chen, S., et al. (2008). MEKK1, MKK1/MKK2 and
MPK4 function together in a mitogen-activated protein kinase cascade to regulate innate
immunity in plants. Cell Res 18, 1190–1198.
Petersen, M., Brodersen, P., Naested, H., Andreasson, E., Lindhart, U., Johansen, B., et al. (2000).
Arabidopsis MAP Kinase 4 Negatively Regulates Systemic Acquired Resistance. Cell 103,
1111–1120. doi:10.1016/S0092-8674(00)00213-0.
Qiu, J. L., Fiil, B. K., Petersen, K., Nielsen, H. B., Botanga, C. J., Thorgrimsen, S., et al. (2008a).
Arabidopsis MAP kinase 4 regulates gene expression through transcription factor release in
the nucleus. EMBO J. 27, 2214–2221.
Qiu, J.-L., Zhou, L., Yun, B.-W., Nielsen, H. B., Fiil, B. K., Petersen, K., et al. (2008b).
Arabidopsis Mitogen-Activated Protein Kinase Kinases MKK1 and MKK2 Have
Overlapping Functions in Defense Signaling Mediated by MEKK1, MPK4, and MKS1.
Plant Physiol. 148, 212–222. doi:10.1104/pp.108.120006.
Rodriguez, M. C. S., Petersen, M., and Mundy, J. (2010). Mitogen-Activated Protein Kinase
Signaling in Plants. Annu. Rev. Plant Biol. 61, 621–649. doi:10.1146/annurev-arplant042809-112252.
Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., et al. (2012). Disruption of PAMPInduced MAP Kinase Cascade by a Pseudomonas syringae Effector Activates Plant
Immunity Mediated by the NB-LRR Protein SUMM2. Cell Host Microbe 11, 253–263.
doi:10.1016/j.chom.2012.01.015.
105