28.01.2015 Views

Immunity in the moss Physcomitrella patens

Immunity in the moss Physcomitrella patens

Immunity in the moss Physcomitrella patens

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

UNIVERSITY OF COPENHAGEN<br />

FACULTY OF SCIENCE<br />

DEPARTMENT OF BIOLOGY<br />

PhD <strong>the</strong>sis<br />

Simon Bressendorff<br />

<strong>Immunity</strong> <strong>in</strong> <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong><br />

Academic advisor: John Mundy<br />

Submitted: 01/11/2012


Abstract............................................................................................................................... 3<br />

Resumé................................................................................................................................ 4<br />

Abbreviations...................................................................................................................... 6<br />

Introduction......................................................................................................................... 7<br />

The plant <strong>in</strong>nate immune system .................................................................................... 7<br />

MTI ............................................................................................................................. 7<br />

ROS............................................................................................................................. 9<br />

MPK signal<strong>in</strong>g pathways .......................................................................................... 10<br />

WRKY transcription factors ..................................................................................... 12<br />

ETI ............................................................................................................................ 12<br />

RAR1, SGT1 and HSP90.......................................................................................... 13<br />

HR............................................................................................................................. 13<br />

Autophagy..................................................................................................................... 14<br />

<strong>Physcomitrella</strong> <strong>patens</strong> ................................................................................................... 17<br />

Evolution................................................................................................................... 17<br />

The life cycle and morphology of <strong>Physcomitrella</strong>.................................................... 18<br />

Homologous recomb<strong>in</strong>ation...................................................................................... 19<br />

The genome of <strong>Physcomitrella</strong> ................................................................................. 20<br />

Innate immunity <strong>in</strong> <strong>Physcomitrella</strong>............................................................................... 21<br />

<strong>Physcomitrella</strong> responses to pathogens .................................................................... 21<br />

<strong>Physcomitrella</strong> responses to MAMPs....................................................................... 21<br />

Results............................................................................................................................... 23<br />

Identification of <strong>Physcomitrella</strong> homologs of defense related Arabidopsis genes....... 23<br />

CERK1...................................................................................................................... 23<br />

MEKK1..................................................................................................................... 25<br />

MKK1 and MKK2 .................................................................................................... 27<br />

MPK4........................................................................................................................ 29<br />

R-genes ..................................................................................................................... 30<br />

Creat<strong>in</strong>g targeted KOs................................................................................................... 32<br />

Sou<strong>the</strong>rn blot............................................................................................................. 35<br />

Flow cytometry ......................................................................................................... 38<br />

Pathogen <strong>in</strong>fections....................................................................................................... 39<br />

Symptoms of a Botrytis c<strong>in</strong>erea <strong>in</strong>fection ................................................................ 40<br />

Symptoms of P. irregulare <strong>in</strong>fection........................................................................ 42<br />

Evans blue sta<strong>in</strong><strong>in</strong>g ................................................................................................... 42<br />

Alternaria spore count<strong>in</strong>g assay................................................................................ 44<br />

The MAP k<strong>in</strong>ase KO ΔMPK4A ................................................................................... 45<br />

P. irregulare <strong>in</strong>fection of ΔMPK4A.......................................................................... 45<br />

Genes not differentially expressed <strong>in</strong> <strong>the</strong> ΔMPK4A-1 l<strong>in</strong>e....................................... 46<br />

Gene expression <strong>in</strong> ΔMPK4A l<strong>in</strong>es upon B. c<strong>in</strong>erea <strong>in</strong>fection ................................. 48<br />

ROS production upon chitosan treatment................................................................. 49<br />

MPK phosphorylation western blots......................................................................... 51<br />

Sporophyte <strong>in</strong>duction................................................................................................ 51<br />

MAMP growth assay ................................................................................................ 52<br />

ΔRAR-1 and R-gene KOs............................................................................................. 52<br />

Evans blue sta<strong>in</strong><strong>in</strong>g upon B. c<strong>in</strong>erea <strong>in</strong>fection of ΔPpRAR1.................................... 53<br />

1


Infections with biotrophic pathogens........................................................................ 53<br />

Yeast two-hybrid analysis of PpRAR1 and PpSGT1 ............................................... 54<br />

Gene expression <strong>in</strong> ΔRAR1 upon chitosan treatment................................................ 55<br />

MPK phosporylation <strong>in</strong> ΔRAR1................................................................................ 56<br />

The autophagy deficient mutant ΔATG5...................................................................... 56<br />

Phenotypic description of <strong>the</strong> ΔATG5 l<strong>in</strong>es.............................................................. 56<br />

ATG8 western blot.................................................................................................... 60<br />

Pathogen treatments of <strong>the</strong> ΔATG5 l<strong>in</strong>es .................................................................. 62<br />

Infection with a Sordariomycetes fungus ................................................................. 63<br />

MPK phosporylation <strong>in</strong> ΔATG5 l<strong>in</strong>es ....................................................................... 65<br />

The AtMEKK1 homologs ΔPpMEKK1A and ΔPpMEKK1B ..................................... 65<br />

Discussion......................................................................................................................... 67<br />

MPKs <strong>in</strong> <strong>Physcomitrella</strong>............................................................................................... 67<br />

ROS........................................................................................................................... 67<br />

Abiotic stress............................................................................................................. 68<br />

Sporophyte formation ............................................................................................... 68<br />

PpMPK4B................................................................................................................. 69<br />

PpRAR1 and R-prote<strong>in</strong>s ........................................................................................... 70<br />

Autophagy..................................................................................................................... 71<br />

Materials and methods ...................................................................................................... 72<br />

Identify<strong>in</strong>g Arabidopsis homologs/orthologs <strong>in</strong> <strong>Physcomitrella</strong> .............................. 72<br />

Moss growth conditions............................................................................................ 72<br />

Media ........................................................................................................................ 73<br />

Clon<strong>in</strong>g of PpMPK4B............................................................................................... 73<br />

Construct<strong>in</strong>g a USER compatible <strong>moss</strong> KO vector .................................................. 74<br />

Preparation of pMBLU ............................................................................................. 75<br />

USER clon<strong>in</strong>g ........................................................................................................... 75<br />

Protoplast isolation.................................................................................................... 76<br />

Transformation.......................................................................................................... 77<br />

Genotyp<strong>in</strong>g................................................................................................................ 78<br />

Sou<strong>the</strong>rn blot............................................................................................................. 81<br />

RNA extraction and quantitative RT-PCR ............................................................... 83<br />

Statistics .................................................................................................................... 84<br />

Flow cytometry ......................................................................................................... 84<br />

ROS accumulation .................................................................................................... 84<br />

Histochemical sta<strong>in</strong><strong>in</strong>g.............................................................................................. 85<br />

cDNA ........................................................................................................................ 85<br />

Yeast two-hybrid....................................................................................................... 85<br />

Western blots ............................................................................................................ 85<br />

PCR conditions ......................................................................................................... 85<br />

References......................................................................................................................... 87<br />

Manuscript 1 ................................................................................................................... 102<br />

Manuscript 2 ................................................................................................................... 136<br />

2


Abstract<br />

Studies <strong>in</strong> flower<strong>in</strong>g plants have provided a wealth of <strong>in</strong>formation on pathogen<br />

recognition, signal transduction and <strong>the</strong> activation of defense responses. However, very<br />

little is known about <strong>the</strong> immune system of <strong>the</strong> phylogenetically ancient <strong>moss</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>. Mosses represent some of <strong>the</strong> earliest land plants and are thus <strong>in</strong><br />

an ideal evolutionary position to provide <strong>in</strong>formation on <strong>the</strong> evolution of plant <strong>in</strong>nate<br />

immune systems. Fur<strong>the</strong>rmore, <strong>Physcomitrella</strong> has <strong>the</strong> unique ability to be genetically<br />

manipulated us<strong>in</strong>g targeted gene replacements through homologous recomb<strong>in</strong>ation.<br />

Us<strong>in</strong>g this emerg<strong>in</strong>g model system, we identify and create targeted knock out of<br />

n<strong>in</strong>e <strong>Physcomitrella</strong> homologs of defense related Arabidopsis genes. The knock-out l<strong>in</strong>es<br />

are assessed for altered immune responses to a range of different pathogens.<br />

We f<strong>in</strong>d that at least one <strong>Physcomitrella</strong> mitogen activated prote<strong>in</strong> k<strong>in</strong>ase (MPK),<br />

PpMPK4A is required for proper <strong>in</strong>nate immune responses. This is a primary example of<br />

a s<strong>in</strong>gle, non-redundant plant MPK essential for immunity without any o<strong>the</strong>r apparent<br />

phenotypes associated with <strong>the</strong> correspond<strong>in</strong>g null-mutant. We show that PpMPK4A is<br />

phosphorylated <strong>in</strong> response to microbe associated molecular patterns (MAMPs) <strong>in</strong>clud<strong>in</strong>g<br />

fungal chit<strong>in</strong> and bacterial MAMPs. The knock out of PpMPK4A renders <strong>the</strong> <strong>moss</strong> more<br />

susceptible to <strong>the</strong> pathogenic fungi Botrytis c<strong>in</strong>erea and Alternaria brassicicola and fails<br />

to accumulate several defense related transcripts and ROS production upon treatment<br />

with fungal chitosan. While related MPKs <strong>in</strong> <strong>the</strong> higher plant model Arabidopsis thaliana<br />

are activated both by pathogen <strong>in</strong>oculation and by abiotic stress, we did not detect<br />

activation of PpMPK4A or any o<strong>the</strong>r <strong>Physcomitrella</strong> MPK by several abiotic stresses.<br />

Signal transduction via PpMPK4A may <strong>the</strong>refore be specific to MAMP-triggered<br />

immunity, and <strong>the</strong> <strong>moss</strong> may use o<strong>the</strong>r signal<strong>in</strong>g components to respond to abiotic<br />

stresses.<br />

In addition, a <strong>Physcomitrella</strong> knock-out of a homolog of <strong>the</strong> autophagy related<br />

gene ATG5 provides <strong>the</strong> first analysis of autophagy <strong>in</strong> non-vascular plants. PpATG5<br />

knock-out mutants show clear signs of autophagy deficiency, such as early senescence<br />

and sensitivity to nutrient deprivation. We also show that PpATG5 is required for <strong>the</strong><br />

defense aga<strong>in</strong>st a Sordariomycetes fungus.<br />

3


Resumé<br />

Studier i blomstrende planter har givet et væld af <strong>in</strong>formationer om pathogen genkendelse,<br />

signal transduktion og aktiver<strong>in</strong>gen af forsvars respons. Men, meget lidt vides om<br />

immunsystemet i den fylogenetisk gamle mos <strong>Physcomitrella</strong> <strong>patens</strong>. Mos repræsentere<br />

nogle af de ældste landplanter og er således i en ideel evolutionær position til at give<br />

<strong>in</strong>formationer om evolutionen af det medfødte plante immunforsvar. Yderligere har<br />

Physcomitrell den unikke evne til at blive genetisk manipuleret med målrettet gen<br />

udskiftn<strong>in</strong>g ved hjælp af homolog rekomb<strong>in</strong>ation.<br />

Ved hjælp af dette nye model system identificere og sletter vi ni <strong>Physcomitrella</strong><br />

homologer af forsvars relaterede Arabidopsis gener. Disse ”knock out” l<strong>in</strong>jer bliver testet<br />

for ændrede immun respons til en række forskellige pathogener.<br />

Vi f<strong>in</strong>der at m<strong>in</strong>dst en <strong>Physcomitrella</strong> mitogen aktiveret prote<strong>in</strong> k<strong>in</strong>ase (MPK),<br />

PpMPK4A er nødvendig for korrekt immun respons. Dette er det første eksempel på en<br />

enkelt ikke-overflødig MPK der er afgørende, uden andre fænotyper knyttet til den<br />

svarende mutant. Vi viser at PpMPK4A bliver fosforyleret som respons på mikrobe<br />

aktiverede molekylære mønstre (MAMPs) som chit<strong>in</strong> fra svampe og bakterielle<br />

MAMPs. ”knock outen” af PpMPK4A gør <strong>moss</strong>en mere modtagelig overfor de<br />

pathogene svampe Botrytis c<strong>in</strong>erea og Alternaria brassicicola og forh<strong>in</strong>dre ophobn<strong>in</strong>gen<br />

af flere forsvars relaterede transkripter og ROS produktionen efter bahandl<strong>in</strong>g med chit<strong>in</strong><br />

fra svampe. Mens MPKer i den blomstrende plantemodel Arabidopsis thaliana bliver<br />

aktiveret af både pathogen <strong>in</strong>fektioner og af abiotisk stress, så vi ikke nogen aktiver<strong>in</strong>g af<br />

PpMPK4A eller nogle andre MPKer i <strong>Physcomitrella</strong> efter mange former for abiotisk<br />

stres. Signal transduktion via PpMPK4A kan derfor være specifik for MAMP-aktiveret<br />

immun system, og <strong>moss</strong>en kan benytte andre signaler<strong>in</strong>gs komponenter som respons på<br />

abiotisk stres.<br />

En <strong>Physcomitrella</strong> ”knock out” a en homolog af det autofagi relateret gen ATG5<br />

giver den første analyse af autofagi i en ikke-vaskulær plante. PpATG5 knock out<br />

planterne viser klare tegn på autofagi mangel, så som tidlig aldr<strong>in</strong>g og følsomhed overfor<br />

nær<strong>in</strong>gsstof mangel. Vi viser at PpATG5 er nødvendig for forsvaret mod en<br />

Sordariomycetes svamp.<br />

4


Preface and acknowledgements<br />

This <strong>the</strong>sis concludes my PhD work at <strong>the</strong> Department of Biology,University of<br />

Copenhagen. The project has resulted <strong>in</strong> <strong>the</strong> manuscript: “A MAP k<strong>in</strong>ase regulates <strong>in</strong>nate<br />

immunity triggered by pathogen associated molecular patterns <strong>in</strong> <strong>the</strong> <strong>moss</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>”, submitted to Pathogen (manuscript 1) and my contribution to <strong>the</strong><br />

review: “Role of autophagy <strong>in</strong> disease resistance and hypersensitive response-associated<br />

cell death” (manuscript 2).<br />

I started this PhD project as part of <strong>the</strong> newly started Center for Comparative<br />

Genomics, but after about a year <strong>the</strong> center was closed as its Director Rasmus Nielsen<br />

moved to U. Berkeley. I <strong>the</strong>refore had to refocus my project on <strong>the</strong> bryophyte plant<br />

model <strong>Physcomitrella</strong> <strong>patens</strong> which had not been previously studied <strong>in</strong> our research<br />

group. Introduc<strong>in</strong>g a new model organism meant that I had to devote much effort and<br />

time to practical and technical problems to get <strong>the</strong> many new techniques to work with<br />

little practical supervision.<br />

Over <strong>the</strong> years I was also <strong>in</strong>volved <strong>in</strong> many different side projects which are not<br />

mentioned <strong>in</strong> <strong>the</strong> report. Toge<strong>the</strong>r with Maria Christ<strong>in</strong>a Suarez I annotated MPK, MKK<br />

and MP3K genes <strong>in</strong> <strong>the</strong> newly sequenced spike <strong>moss</strong> Selag<strong>in</strong>ella Moellendorfii. Toge<strong>the</strong>r<br />

with Michael Krogh Jensen I did many experiments with <strong>the</strong> carnivorous plant Venus<br />

flytrap (Dionaea muscipula). For example I optimized protocols for do<strong>in</strong>g DNA and<br />

RNA extraction from <strong>the</strong> flytrap and I successfully estimated <strong>the</strong> very large flytrap<br />

genome size (~3GBp) based on a qPCR method. Toge<strong>the</strong>r with Klaus Petersen I<br />

fur<strong>the</strong>rmore cloned several R-genes <strong>in</strong> Arabidopsis and <strong>in</strong>troduced dom<strong>in</strong>ant negative P-<br />

loop mutations.<br />

Dur<strong>in</strong>g <strong>the</strong> PhD program I supervised several people that I thank for <strong>the</strong>ir <strong>in</strong>terest<br />

<strong>in</strong> <strong>the</strong> <strong>moss</strong> project. Some of <strong>the</strong> figures <strong>in</strong> this report are done by <strong>the</strong>se people under my<br />

supervision: i. <strong>the</strong> yeast two hybrid assay <strong>in</strong> Figure 36 was done by lab technician student<br />

Ali Fard, and ii. <strong>the</strong> western blot <strong>in</strong> Figure 34 and <strong>the</strong> qPCR <strong>in</strong> Figure 32 were done by<br />

BSc. students Søren Iversen and Mira Wilkan.<br />

I would like to thank John Mundy and Morten Petersen for support<strong>in</strong>g <strong>the</strong> project<br />

even when noth<strong>in</strong>g seemed to work. A very special thanks to everybody at IIBCE <strong>in</strong><br />

Uruguay for be<strong>in</strong>g very generous hosts, and especially Inés Ponce de León for teach<strong>in</strong>g<br />

me many <strong>moss</strong> techniques, and Alexandra Castro who helped me with <strong>the</strong> Sou<strong>the</strong>rn blot<br />

analysis. Also thanks to Anders Tolver Jensen for help with <strong>the</strong> statistics, Sarah L<strong>in</strong>e<br />

Skovbakke for help<strong>in</strong>g with <strong>the</strong> FACS analysis, Henn<strong>in</strong>g Knudsen for help<strong>in</strong>g with <strong>the</strong><br />

identification of <strong>the</strong> Sordariomycetes fungus, Christ<strong>in</strong>e Lunde for provid<strong>in</strong>g <strong>the</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong> (Gransden 2004) wild type stra<strong>in</strong>, and Andrew Cum<strong>in</strong>g for <strong>the</strong><br />

orig<strong>in</strong>al pMBL6 and pMBLU10a KO vectors.<br />

October 2012<br />

Simon Bressendorff<br />

5


Abbreviations<br />

ACC = 1-Am<strong>in</strong>ocyclopropane-1-carboxylic acid (Ethylene<br />

Precursor) MKS1 = MPK4 Substrate 1<br />

a-DOX = alfa dioxygenase<br />

MP3K = Mitogen activated prote<strong>in</strong> k<strong>in</strong>ase k<strong>in</strong>ase k<strong>in</strong>ase<br />

Amp = Ampicill<strong>in</strong><br />

MPK= Mitogen Activated Prote<strong>in</strong> K<strong>in</strong>ase<br />

Arabidopsis = Arabidopsis thaliana<br />

MTI = MAMP Triggered <strong>Immunity</strong><br />

At = Arabidopsis thaliana<br />

MYA = Million Years Ago<br />

ATG = Autophagy related<br />

Nb = Nicotiana benthamiana<br />

BAK1 = BRI1-associated receptor k<strong>in</strong>ase 1<br />

NB = Nucleotide B<strong>in</strong>d<strong>in</strong>g<br />

BCD = Basal growth medium NDR1 = Non-race specific Disease Resistance 1<br />

BCD+AT = Basal growth medium supplemented with ammonium<br />

tartrate (see Material and Methods)<br />

NOX = NADPH oxidase<br />

bla = amp-resistance gene<br />

Os = Oryza sativa<br />

bp = base pairs<br />

PAD = Phytoalex<strong>in</strong>-Deficient<br />

CC = Coiled-Coil<br />

PAL = Phenyl Allan<strong>in</strong>e Ammonialyase<br />

CDPK = calcium-dependent prote<strong>in</strong> k<strong>in</strong>ases<br />

PAMP = Pathogen Associated Molecular Pattern<br />

Ce = Caenorhabditis elegans<br />

PCD = Programmed Cell Death<br />

CEBiP = Chit<strong>in</strong> Elicitor-B<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong><br />

PEG = Polyethylene Glycol<br />

CERK = Chit<strong>in</strong> Elicitor Receptor K<strong>in</strong>ase<br />

PGN = Peptidoglycan<br />

<strong>Physcomitrella</strong> = <strong>Physcomitrella</strong> <strong>patens</strong> (Gransden 2004<br />

CHS = Chalcone Synthase<br />

stra<strong>in</strong>)<br />

CYP71A13 = cytochrome P450, family 71, subfamily A,<br />

polypeptide 13<br />

Pp = <strong>Physcomitrella</strong> <strong>patens</strong><br />

DAMP = Danger Associated Molecular Pattern<br />

PR = Pathogen Related<br />

EDS1 = Enhanced Disease Susceptibility 1<br />

PRR = Pathogen Recognition Receptor<br />

EF-Tu = Elongation Factor Tu<br />

PRX = Peroxidase<br />

ERF = Ethylene Respone Factor<br />

Pto = Pseudomonas syr<strong>in</strong>gae pv. tomato DC3000<br />

qPCR = quantitative Reverse Transcriptase Poly Cha<strong>in</strong><br />

ETI = Effector Triggered <strong>Immunity</strong><br />

Reaction<br />

flg22 = Flagell<strong>in</strong> peptide<br />

RAR1 = required for Mla12 resistance1<br />

G418 = Genetic<strong>in</strong><br />

RB = Right Boarder<br />

hpi = hours post <strong>in</strong>fection<br />

RbohB = respiratory burst oxidase homolog B<br />

hpt = hours post treatment<br />

R-gene = Resistance gene<br />

HR = Hypersensitive Response<br />

RLK = Receptor-like k<strong>in</strong>ase<br />

HSP = Heat Shock Prote<strong>in</strong><br />

ROS = Reactive Oxygen Species<br />

JA = Jasmonic Acid<br />

SA = Salicylic Acid<br />

KI = Knock In SAG101 = Senescence Associated Gene 101<br />

KO = Knock Out<br />

Sc = Saccharomyces cerevisiae<br />

LB = Left Boarder OR bacteria media, read from context<br />

SEM = Standard Error of <strong>the</strong> Mean<br />

LOX = Lipooxygenase<br />

SGT1 = suppressor of <strong>the</strong> G2 allele of skp1<br />

LRR = Leuc<strong>in</strong>e Rich Repeat<br />

SUMM = suppressor of mkk1 mkk2<br />

LysM = Lys<strong>in</strong> Motif<br />

T-DNA = Transfer DNA<br />

MAMP = Microbe Associated Molecular Pattern<br />

TIR = Toll/Interleuk<strong>in</strong>-1 Receptor<br />

TSPO = mitochondrial peripheral-type benzodiazep<strong>in</strong>e<br />

MAPK = Mitogen Activated Prote<strong>in</strong> K<strong>in</strong>ase<br />

receptor<br />

Mb = Mega Base pairs<br />

TUB6 = beta-Tubul<strong>in</strong>6<br />

MC = Meta-Caspase<br />

USER clon<strong>in</strong>g = Urasil-Specific Excision Reagent clon<strong>in</strong>g<br />

MeJA= Methyl-Jasmonate<br />

WT= wild type<br />

MEKK = MP3K<br />

Y2H = Yeast Two Hybrid<br />

MKK = Mitogen activated prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ase<br />

Δ = knockout mutant<br />

6


Introduction<br />

Ever s<strong>in</strong>ce plants colonized land <strong>the</strong>y have formed <strong>the</strong> basis of all o<strong>the</strong>r life on land as<br />

<strong>the</strong>y alone have <strong>the</strong> ability to convert solar radiation <strong>in</strong>to <strong>the</strong> chemical energy we call<br />

nutrients. As autotrophs <strong>the</strong>y provide <strong>the</strong> nutrients upon which all o<strong>the</strong>r life feeds on.<br />

Thus, viruses, bacteria, fungi and animals depend on plants to harvest <strong>the</strong> solar energy.<br />

So even humans compete with microbes to get <strong>the</strong>se nutrients. It is estimated that yearly<br />

crop losses due to microbial pathogens is between 7 and 15%, depend<strong>in</strong>g on <strong>the</strong> crop<br />

(Oerke, 2006).<br />

The plant <strong>in</strong>nate immune system<br />

Plants, like animals, have been <strong>in</strong>volved <strong>in</strong> an evolutionary arms race with <strong>the</strong> pathogens<br />

that <strong>in</strong>fect <strong>the</strong>m. This race has evolved sophisticated strategies by which pathogens attack<br />

<strong>the</strong>ir hosts, and equally sophisticated strategies by which hosts defend <strong>the</strong>mselves.<br />

Host immune system must i. recognize non-self, ii. judge if <strong>the</strong> recognized nonself<br />

posses a danger, and iii. if dangerous, <strong>in</strong>itiate an appropriate response. In order to do<br />

so, plants have evolved a multilayered <strong>in</strong>nate immune system to combat pathogen attacks<br />

(Jones and Dangl, 2006). The first layer relies on recogniz<strong>in</strong>g <strong>the</strong> pathogen at <strong>the</strong><br />

boundaries of <strong>the</strong> cell to <strong>in</strong>itiate defense responses to prevent <strong>the</strong> pathogen from enter<strong>in</strong>g<br />

<strong>the</strong> cell. This is made possible by pathogen recognition receptors (PRRs) <strong>in</strong> <strong>the</strong> plasma<br />

membrane. These receptors recognize microbe/pathogen associated molecular patterns<br />

(MAMPs or PAMPs). S<strong>in</strong>ce <strong>the</strong> patterns recognized by <strong>the</strong> receptors are characteristics of<br />

classes of microbes, whe<strong>the</strong>r <strong>the</strong>y are pathogenic or not, <strong>the</strong> term MAMP is generally<br />

more appropriate than PAMP. This first layer of defense is termed MAMP triggered<br />

immunity (MTI) or basal defense. A typical outcome of a MTI is <strong>the</strong> production of<br />

antimicrobial compounds and cell wall fortification.<br />

Some pathogens have evolved mechanisms to overcome this first layer of defense<br />

by deliver<strong>in</strong>g effector molecules <strong>in</strong>to <strong>the</strong> host cell <strong>in</strong> an attempt to disrupt <strong>the</strong> MTI. In<br />

order to avoid this immune suppression, plants have evolved a surveillance system that<br />

ei<strong>the</strong>r recognizes <strong>the</strong> delivered effector molecules directly or recognizes <strong>the</strong>m <strong>in</strong>directly<br />

by sens<strong>in</strong>g <strong>the</strong> damage <strong>the</strong>y cause, i.e. by recogniz<strong>in</strong>g altered self. Such surveillance is<br />

done by R-prote<strong>in</strong>s (resistance prote<strong>in</strong>s) that are said to guard <strong>the</strong> targets of <strong>the</strong> effector.<br />

This second layer of defense is termed effector triggered immunity (ETI). A ETI response<br />

is more drastic than <strong>the</strong> MTI response, s<strong>in</strong>ce it often triggers a rapid, localized<br />

programmed cell death (PCD) known as <strong>the</strong> hypersensitive response (HR) (DeYoung and<br />

Innes, 2006). This scorched-earth tactic is very effective <strong>in</strong> stopp<strong>in</strong>g biotrophic pathogens<br />

from spread<strong>in</strong>g.<br />

MTI<br />

Most research <strong>in</strong> plant immunity has been done <strong>in</strong> <strong>the</strong> model plant Arabidopsis or <strong>in</strong> o<strong>the</strong>r<br />

flower<strong>in</strong>g plants like tobacco, rice and tomato. In <strong>the</strong>se plants, several MAMPs and <strong>the</strong>ir<br />

correspond<strong>in</strong>g PRR have been identified. These MAMPs <strong>in</strong>clude <strong>the</strong> 22 am<strong>in</strong>o acid<br />

peptide flg22 from <strong>the</strong> flagell<strong>in</strong> prote<strong>in</strong> which is an essential build<strong>in</strong>g block of bacterial<br />

7


flagella (Felix et al., 2002). Flg22 is recognized by <strong>the</strong> leuc<strong>in</strong>e rich repeat receptor like<br />

k<strong>in</strong>ase (LRR-RLK) flagell<strong>in</strong> sens<strong>in</strong>g 2 (FLS2) (Ch<strong>in</strong>chilla et al., 2006). The extracellular<br />

doma<strong>in</strong> of FLS2 is similar to <strong>the</strong> <strong>in</strong>nate immunity receptor TOLL <strong>in</strong> Drosophila (Gómez-<br />

Gómez and Boller, 2000). The Arabidopsis Atfls2 mutant is more susceptible to <strong>the</strong><br />

bacterial pathogen Pseudomonas syr<strong>in</strong>gae pv. tomato DC3000 (Pto) (Zipfel et al., 2004)<br />

Ano<strong>the</strong>r bacterial MAMP is <strong>the</strong> 18 residue elf18 peptide from <strong>the</strong> N-term<strong>in</strong>us of<br />

elongation factor Tu (EF-Tu) that is recognized by <strong>the</strong> LRR-RLK elongation factor<br />

receptor (EFR) (Zipfel et al., 2006). Both FLS2 and EFR form complexes with ano<strong>the</strong>r<br />

transmembrane LRR-RLK receptor, <strong>the</strong> BRI1-associated receptor k<strong>in</strong>ase 1 (BAK1)<br />

which respectively perceive flg22 and elf18 and mediate immunity (Ch<strong>in</strong>chilla et al.,<br />

2007; Roux et al., 2011). Atefr mutants are more susceptible to colonization by weakly<br />

virulent mutant stra<strong>in</strong>s of Pto (Nekrasov et al., 2009).<br />

The fungal MAMP chit<strong>in</strong>, a b-(1,4)-l<strong>in</strong>ked oligosaccharide of N-acetylglucosam<strong>in</strong>e,<br />

is an essential component of <strong>the</strong> cell walls of fungi that is not found <strong>in</strong><br />

plants and vertebrates. Chit<strong>in</strong> fragments have long been known to act as elicitors of<br />

defense responses <strong>in</strong> plants (Felix et al., 1993). It is recognized by two RLKs <strong>in</strong> rice<br />

(Oryza sativa), <strong>the</strong> chit<strong>in</strong> elicitor-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> (OsCEBiP) and chit<strong>in</strong> elicitor receptor<br />

k<strong>in</strong>ase1 (OsCERK1), which cooperatively regulate chit<strong>in</strong> perception by form<strong>in</strong>g a<br />

receptor complex (Shimizu et al., 2010). In Arabidopsis <strong>the</strong> closest homolog of rice<br />

OsCEBiP (AtLYM2) does not contribute to chit<strong>in</strong> perception (Sh<strong>in</strong>ya et al., 2012), while<br />

<strong>the</strong> closest homolog of OsCERK1 is solely responsible for <strong>the</strong> detection of chit<strong>in</strong> (Miya<br />

et al., 2007). It does so by form<strong>in</strong>g homodimers <strong>in</strong> <strong>the</strong> presence of chit<strong>in</strong> (Liu et al., 2012).<br />

The Arabidopsis Atcerk1 mutant exhibits enhanced susceptibility to <strong>the</strong> necrotrophic<br />

fungus Alternaria brassicicola (Miya et al., 2007).<br />

It has recently been found <strong>in</strong> Arabidopsis that two o<strong>the</strong>r RLKs, AtLYM1 and<br />

AtLYM3 with sequence identity to rice OsCEBiP, <strong>in</strong>teract with AtCERK1 to perceive <strong>the</strong><br />

presence of <strong>the</strong> bacterial derived MAMP peptidoglycan (PGN). This hetero-oligomer<br />

perception system is similar to <strong>the</strong> OsCEBiP OsCERK1 complex that mediates chit<strong>in</strong><br />

perception and immunity to fungal <strong>in</strong>fection <strong>in</strong> rice (Willmann et al., 2011). The ability<br />

of PRRs to form complexes and engage <strong>in</strong> hetero-oligomerization <strong>in</strong> different<br />

comb<strong>in</strong>ations greatly extends <strong>the</strong> abilities of relatively few PRRs to recognize many<br />

different MAMPs.<br />

Apart from MAMPs that orig<strong>in</strong>ate from <strong>the</strong> pathogen, plant cells under attack can also<br />

produce <strong>the</strong>ir own danger signals (damaged-associated molecular patterns, or DAMPs)<br />

that are thought to amplify <strong>the</strong> same responses triggered by MAMPs (Krol et al., 2010;<br />

Segonzac and Zipfel, 2011).<br />

PRRs belongs to <strong>the</strong> very large RLK family with >600 members <strong>in</strong> Arabidopsis, >1100 <strong>in</strong><br />

rice and 157 <strong>in</strong> <strong>Physcomitrella</strong> (Shiu et al., 2004; Vij et al., 2008). The large number of<br />

genes and <strong>the</strong> expansion of <strong>the</strong> gene family through evolution <strong>in</strong>dicates that clear one to<br />

one orthologous connection cannot be established. For example, <strong>in</strong> <strong>Physcomitrella</strong> <strong>the</strong>re<br />

are no clear homologs of <strong>the</strong> RLK receptors FLS2 and ERF2, but <strong>the</strong>re is one of BAK1<br />

(Boller and Felix, 2009) and possibly also of CERK1 (Lawton and Saidasan, 2009).<br />

However, thus far no functional PRRs have been described <strong>in</strong> <strong>Physcomitrella</strong>.<br />

8


ROS<br />

With<strong>in</strong> seconds of PAMP perception by cognate PRRs, apoplastic reactive oxygen<br />

species (ROS) rapidly accumulate <strong>in</strong> a reaction known as <strong>the</strong> ROS burst. The role of ROS<br />

<strong>in</strong> disease resistance is not very well understood, but it is conserved throughout <strong>the</strong> plant<br />

k<strong>in</strong>gdom and also occurs <strong>in</strong> animal <strong>in</strong>nate immunity (Kohchi et al., 2009; Pérez-Pérez et<br />

al., 2012).<br />

Molecular l<strong>in</strong>ks between PRRs and ROS production have not been fully<br />

documented, but it is known that activation of PRRs follow<strong>in</strong>g MAMP perception<br />

stimulates a rapid <strong>in</strong>flux of Ca 2+ from <strong>the</strong> apoplast which <strong>in</strong>creases <strong>the</strong> cytoplasmic Ca 2+<br />

concentrations. This activates calcium-dependent prote<strong>in</strong> k<strong>in</strong>ases (CDPKs) that <strong>the</strong>n<br />

directly phosphorylate <strong>the</strong> NADPH oxidase RbohB (respiratory burst oxidase homolog B)<br />

(Boudsocq and Sheen). The membrane localized Rboh enzymes are a key source of ROS<br />

<strong>in</strong> plants (Torres and Dangl, 2005).<br />

In plants, ROS can contribute to defense ei<strong>the</strong>r directly as an antibiotic agent or<br />

<strong>in</strong>directly by promot<strong>in</strong>g oxidative cross-l<strong>in</strong>k<strong>in</strong>g <strong>in</strong> <strong>the</strong> cell wall (Apel and Hirt, 2004). It<br />

can also act as a secondary messenger ei<strong>the</strong>r directly <strong>in</strong> <strong>the</strong> form of H 2 O 2 , or <strong>in</strong>directly by<br />

oxidiz<strong>in</strong>g polyunsaturated fatty acids that may act as secondary messengers to trigger<br />

defense responses (Forman et al., 2010; Vellosillo et al., 2010). In mammalian <strong>in</strong>nate<br />

immunity, ROS generation is used to kill bacteria with<strong>in</strong> <strong>the</strong> phagosome (Diebold and<br />

Bokoch, 2005), but it can also function as a secondary messenger of signal transduction<br />

(Kohchi et al., 2009).<br />

The ROS burst has long been used as a tool to identify MTI signal<strong>in</strong>g components <strong>in</strong><br />

Arabidopsis, and recently <strong>the</strong> ROS burst has been described <strong>in</strong> <strong>moss</strong>es with<strong>in</strong> seconds of<br />

treatment with <strong>the</strong> fungal MAMP chitosan (Lehtonen et al., 2012). Lethonen et al. found<br />

that a class III peroxidase (PpPRX34) is essential for <strong>the</strong> ROS burst <strong>in</strong> <strong>Physcomitrella</strong>,<br />

and that <strong>the</strong> ΔPpPRX34 mutant is more susceptible to saprophytic and a necrotrophic<br />

fungi than <strong>the</strong> wild type <strong>moss</strong> (Lehtonen et al., 2009, 2012). Similarly, Arabidopsis T-<br />

DNA <strong>in</strong>sertion l<strong>in</strong>es of two apoplastic peroxidases, Atprx33 and Atprx34, exhibit reduced<br />

ROS and callose deposition <strong>in</strong> response to <strong>the</strong> bacterial MAMPs flg22 and elf26 (Daudi<br />

et al., 2012).<br />

However, ROS production is not restricted to MTI but is a general stress response that<br />

also occurs after abiotic stress and <strong>in</strong> ETI triggered HR (Zurbriggen et al., 2010; Suzuki<br />

et al., 2012). As <strong>in</strong> flower<strong>in</strong>g plants, ROS is also produced upon abiotic stress <strong>in</strong><br />

<strong>Physcomitrella</strong> (Frank et al., 2007). A <strong>Physcomitrella</strong> mutant unable to produce a<br />

mitochondrial localized prote<strong>in</strong> homologous to <strong>the</strong> mammalian mitochondrial peripheraltype<br />

benzodiazep<strong>in</strong>e receptor (ΔPpTSPO1) is unable to ma<strong>in</strong>ta<strong>in</strong> redox homeostasis and<br />

is hyper sensitive to salt stress and fungal pathogens (Frank et al., 2007; Lehtonen et al.,<br />

2012).<br />

Recently a connection between ROS and autophagy has emerged, s<strong>in</strong>ce several studies <strong>in</strong><br />

different plants have shown activation of autophagy <strong>in</strong> response to stimuli that <strong>in</strong>crease<br />

ROS generation, regardless of <strong>the</strong> orig<strong>in</strong> and location of ROS production <strong>in</strong> <strong>the</strong> cell<br />

(reviewed <strong>in</strong> Pérez-Pérez et al., 2012).<br />

9


MPK signal<strong>in</strong>g pathways<br />

Mitogen activated prote<strong>in</strong> k<strong>in</strong>ases (MPKs) are a class of enzymes present <strong>in</strong> all<br />

Eukaryotes. They transduce extracellular stimuli from <strong>the</strong> cell surface to <strong>the</strong> nucleus via a<br />

sequential phosphorylation cascade <strong>in</strong> which MPK k<strong>in</strong>ase k<strong>in</strong>ases (MP3Ks or MEKKs)<br />

phosphorylate MPK k<strong>in</strong>ases (MKKs) which phosphorylate MPKs (Ichimura et al., 2002).<br />

They are <strong>in</strong>volved <strong>in</strong> a wide range of cellular functions <strong>in</strong>clud<strong>in</strong>g stress-responses, cellcycle<br />

checkpo<strong>in</strong>t pathways, and cell differentiation and proliferation (Suarez-Rodriguez<br />

et al., 2010). The Arabidopsis genome encodes 60 MAP3Ks, 10 MAP2Ks, and 20<br />

MAPKs (Ichimura et al., 2002). This <strong>in</strong>dicates that a MPK cascade does not necessarily<br />

only consist of s<strong>in</strong>gle MAP3K, a s<strong>in</strong>gle MAP2K, and a s<strong>in</strong>gle MAPK. It ra<strong>the</strong>r <strong>in</strong>dicates<br />

that some levels of redundancy must exist.<br />

In Arabidopsis AtMPK3, 4, 6 and 11 have all been shown to be activated with<strong>in</strong><br />

m<strong>in</strong>utes of PRRs perception of <strong>the</strong> MAMPs flg22, elf18 and chit<strong>in</strong> (Wan et al., 2004;<br />

Nekrasov et al., 2009; Petutschnig et al., 2010; Roux et al., 2011; Bethke et al., 2012;<br />

Eschen-Lippold et al., 2012). AtMPK3, 4 and 6 are also activated by abiotic stresses like<br />

salt, drought, cold, UV-light and wound<strong>in</strong>g (Ichimura et al., 2000; Droillard et al., 2004;<br />

Teige et al., 2004; González Besteiro et al., 2011). The activation of AtMPK11 upon<br />

MAMP sens<strong>in</strong>g has only been shown very recently, and <strong>the</strong>re is no data on <strong>the</strong> MPK<br />

cascade lead<strong>in</strong>g to <strong>the</strong> activation of AtMPK11 (Bethke et al., 2012; Eschen-Lippold et al.,<br />

2012).<br />

However, two MPK cascades have been shown to be activated downstream of<br />

PRRs. The first to be described was <strong>the</strong> cascade downstream of AtFLS2 consist<strong>in</strong>g of <strong>the</strong><br />

two MP3Ks AtMPKKKα and AtMEKK1, <strong>the</strong> two MKKs AtMKK4 and AtMKK5 and <strong>the</strong><br />

two MPKs AtMPK3 and AtMPK6 (Asai et al., 2002; Ren et al., 2008). The o<strong>the</strong>r MPK<br />

cascade act<strong>in</strong>g downstream of PRRs consists of AtMEKK1, AtMKK1/AtMKK2 and<br />

AtMPK4 (Gao et al., 2008). All <strong>the</strong>se components have been shown to <strong>in</strong>teract, and might<br />

thus act <strong>in</strong> a scaffold<strong>in</strong>g complex (Suarez-Rodriguez et al., 2007). Usually <strong>the</strong> activity of<br />

<strong>the</strong> components <strong>in</strong> such a signal<strong>in</strong>g cascade is conferred through phosphorylation, but<br />

surpris<strong>in</strong>gly a k<strong>in</strong>ase dead version of AtMEKK1 could rescue AtMPK4 activation <strong>in</strong> a<br />

Atmekk1 mutant background. This suggests that AtMEKK1 merely acts as a scaffold<strong>in</strong>g<br />

prote<strong>in</strong>, and that ano<strong>the</strong>r MP3K may supply <strong>the</strong> k<strong>in</strong>ase activity as long as AtMEKK1 is<br />

present (Suarez-Rodriguez et al., 2007).<br />

Exactly how PRR b<strong>in</strong>d<strong>in</strong>g of MAMPs leads to <strong>the</strong> activation of a MPK cascade is<br />

not known, but <strong>the</strong> <strong>in</strong>flux of Ca 2+ and subsequent activation of CDPKs is thought to be<br />

<strong>in</strong>volved (Wurz<strong>in</strong>ger et al., 2011).<br />

AtMPK4 was orig<strong>in</strong>ally reported as a negative regulator of plant immunity because <strong>the</strong><br />

Atmpk4 mutant displayed constitutive expression of defense related genes, elevated ROS<br />

levels, high levels of salicylic acid (SA), spontaneous cell death and a dwarfed growth<br />

phenotype (Petersen et al., 2000). The Atmekk1 mutant and <strong>the</strong> Atmkk1/Atmkk2 double<br />

mutant display similar phenotypes to Atmpk4, underl<strong>in</strong><strong>in</strong>g <strong>the</strong>ir functional connections<br />

(Suarez-Rodriguez et al., 2007; Gao et al., 2008; Qiu et al., 2008b). It has recently been<br />

shown that severe phenotypes of mutants <strong>in</strong> this MPK cascade are due to <strong>the</strong> activation of<br />

10


ETI through <strong>the</strong> R-prote<strong>in</strong> AtSUMM2 (suppressor of mkk1/mkk2 2) s<strong>in</strong>ce <strong>the</strong> triple<br />

mutant Atsumm2/Atmkk1/Atmkk2 displayed normal phenotype <strong>in</strong> respect to growth,<br />

defense gene expression, ROS and hormone levels (Zhang et al., 2012). The double<br />

mutant of Atsumm2/Atmekk1 also displayed a normal phenotype, while <strong>the</strong> double mutant<br />

Atsumm2/Atmpk4 did not fully rescue <strong>the</strong> phenotype of Atmpk4 s<strong>in</strong>ce it still had elevated<br />

ROS levels, residual cell death and slightly stunted growth (Zhang et al., 2012). This<br />

reta<strong>in</strong>ed phenotype of <strong>the</strong> Atsumm2/Atmpk4 double mutant suggests that AtMPK4 could<br />

be guarded by ano<strong>the</strong>r R-prote<strong>in</strong> besides AtSUMM2.<br />

Inducible expression of <strong>the</strong> bacterial effector HopAI1 <strong>in</strong> wild-type plants gives rise to a<br />

defense phenotype similar to that seen <strong>in</strong> Atmekk1, Atmkk1/Atmkk2 and Atmpk4 mutants<br />

<strong>in</strong>clud<strong>in</strong>g elevated levels of ROS, defense gene expression and cell death (Zhang et al.,<br />

2012). This is because <strong>the</strong> bacteria Pto <strong>in</strong>jects <strong>the</strong> HopAI1 effector <strong>in</strong>to <strong>the</strong> plant cell to<br />

prevent immune responses by irreversibly <strong>in</strong>activate AtMPK3, AtMPK4 and AtMPK6,<br />

and this modified self is sensed by <strong>the</strong> guard R-prote<strong>in</strong> AtSUMM2 which <strong>the</strong>n activates a<br />

ETI response (Zhang et al., 2007, 2012). Thus, <strong>the</strong> immune responses caused by HopAI1<br />

were completely suppressed <strong>in</strong> <strong>the</strong> Atsumm2 background (Zhang et al., 2012). However,<br />

which host component <strong>the</strong> R-prote<strong>in</strong> AtSUMM2 is guard<strong>in</strong>g has not yet been identified,<br />

s<strong>in</strong>ce AtSUMM2 does not <strong>in</strong>teract directly with any of components of <strong>the</strong> AtMEKK1,<br />

AtMKK1/AtMKK2 and AtMPK4 signal<strong>in</strong>g cascade (Zhang et al., 2012). This <strong>in</strong>dicates<br />

that AtSUMM2 possibly guards a downstream target of <strong>the</strong> AtMPK4 activity.<br />

Interest<strong>in</strong>gly, <strong>the</strong> suppressor screen of <strong>the</strong> Atmkk1/Atmkk2 double mutant also<br />

identified AtSUMM1 which turned out to be identical with <strong>the</strong> MP3K AtMEKK2 (Kong<br />

et al., 2012). Mutations <strong>in</strong> AtMEKK2 were able to suppress <strong>the</strong> dwarfed phenotype and<br />

<strong>the</strong> constitutive defense responses <strong>in</strong> Atmekk1, Atmkk1/Atmkk2, and Atmpk4 mutant plants<br />

although <strong>the</strong> Atmekk2/Atmpk4 double mutant reta<strong>in</strong>ed residual cell death and slightly<br />

elevated ROS levels just like <strong>the</strong> Atsumm2/Atmpk4 mutant (Kong et al., 2012). They<br />

found that disruption of <strong>the</strong> AtMEKK1, AtMKK1/AtMKK2 and AtMPK4 signal<strong>in</strong>g<br />

cascade leads to activation of AtMEKK2, which triggers AtSUMM2 mediated immune<br />

responses. However, s<strong>in</strong>ce AtMEKK2 and AtSUMM2 did not <strong>in</strong>teract directly, <strong>the</strong> host<br />

component guarded by AtSUMM2 still rema<strong>in</strong>s to be found (Kong et al., 2012).<br />

Understand<strong>in</strong>g <strong>the</strong> function of <strong>in</strong>dividual component of <strong>the</strong> MPK signal<strong>in</strong>g cascade <strong>in</strong><br />

flower<strong>in</strong>g plants have been complicated by functional redundancy, potential promiscuity,<br />

and <strong>in</strong>appropriate auto-activation of immune responses <strong>in</strong> some mutants. Toge<strong>the</strong>r <strong>the</strong>se<br />

observations <strong>in</strong>dicate that studies of MPKs <strong>in</strong> different evolutionary models may help<br />

elucidate <strong>the</strong>ir basal and subsequently acquired functions. A recent phylogenetic study<br />

found that <strong>the</strong>re are seven MKKs and eight MPKs <strong>in</strong> <strong>Physcomitrella</strong> (Dóczi et al., 2012).<br />

However, <strong>the</strong>re are no reports on how many MP3Ks <strong>the</strong>re are <strong>in</strong> <strong>Physcomitrella</strong>, and no<br />

reports on <strong>the</strong> functions of any s<strong>in</strong>gle MPKs or MPK signal<strong>in</strong>g cascades components.<br />

S<strong>in</strong>ce <strong>Physcomitrella</strong> has fewer members of <strong>the</strong> MPK gene family, it could be imag<strong>in</strong>ed<br />

that <strong>the</strong>y exhibit less redundancy and it could thus be easier to assign a function for each<br />

MPK.<br />

11


WRKY transcription factors<br />

WRKY transcription factors are members of a large family of plant specific prote<strong>in</strong>s<br />

def<strong>in</strong>ed by <strong>the</strong> presence of one or two WRKY doma<strong>in</strong>s, a ~60 am<strong>in</strong>o acid region that<br />

conta<strong>in</strong>s <strong>the</strong> conserved signature motif WRKYGQK followed by a z<strong>in</strong>k f<strong>in</strong>ger structure.<br />

There are no WRKYs <strong>in</strong> animals or yeast, but <strong>the</strong>y have evolved rapidly with<strong>in</strong> <strong>the</strong> plant<br />

k<strong>in</strong>gdom as <strong>the</strong>re is just one WRKY <strong>in</strong> <strong>the</strong> green algae C. re<strong>in</strong>hardtii, 37 <strong>in</strong><br />

<strong>Physcomitrella</strong> and 74 <strong>in</strong> Arabidopsis (Rushton et al., 2010). WRKY transcription factors<br />

are <strong>in</strong>volved <strong>in</strong> both positive and negative gene regulation <strong>in</strong> a range of processes<br />

<strong>in</strong>clud<strong>in</strong>g biotic and abiotic stress, senescence and different developmental processes<br />

(Rushton et al., 2010).<br />

In unchallenged Arabidopsis plants, AtMPK4 forms a nuclear localized<br />

complex with AtMKS1 (MPK4 Substrate 1) and AtWRKY25 and AtWRKY33<br />

(Andreasson et al., 2005; Qiu et al., 2008a). When treated with flg22, AtMPK4<br />

phosphorylates AtMKS1 which leads to <strong>the</strong> dissociation of <strong>the</strong> complex, <strong>the</strong>reby releas<strong>in</strong>g<br />

<strong>the</strong> AtWRKY33 transcription factor to b<strong>in</strong>d to <strong>the</strong> promoters of its target genes (Qiu et al.,<br />

2008a). One target of AtWRKY33 is <strong>the</strong> gene encod<strong>in</strong>g <strong>the</strong> P450 monooxygenase<br />

AtPAD3 (Phytoalex<strong>in</strong>-Deficient 3). AtPAD3 catalyzes <strong>the</strong> f<strong>in</strong>al step <strong>in</strong> <strong>the</strong> biosyn<strong>the</strong>sis of<br />

<strong>the</strong> antimicrobial compound camalex<strong>in</strong> (Schuhegger et al., 2006). The production of<br />

camalex<strong>in</strong> is important <strong>in</strong> defense aga<strong>in</strong>st necrotrophic fungi s<strong>in</strong>ce Atpad3 and Atwrky33<br />

mutants are more susceptible to <strong>in</strong>fection with B. c<strong>in</strong>erea and A. brassicicola (Ferrari et<br />

al., 2003; Wees et al., 2003; Zheng et al., 2006). Besides AtPAD3, several o<strong>the</strong>r defense<br />

related genes have been identified as direct targets of AtWRKY33 underl<strong>in</strong><strong>in</strong>g its<br />

important role <strong>in</strong> MTI (Birkenbihl et al., 2012)<br />

Recently Mao et al. (2011) described how WRKY33 also functions downstream<br />

of AtMPK3/AtMPK6 by be<strong>in</strong>g phosphorylated <strong>in</strong> vivo by <strong>the</strong>se MPKs <strong>in</strong> response to B.<br />

c<strong>in</strong>erea <strong>in</strong>fection. AtWRKY22 and AtWRKY29 have also been reported to function<br />

downstream of <strong>the</strong> AtMEKK1, AtMKK4/AtMKK5 and AtMPK3/AtMPK6 pathway (Asai<br />

et al., 2002)<br />

To our knowledge <strong>the</strong>re are no reports on <strong>the</strong> function of any of <strong>the</strong> 37 reported WRKY<br />

transcription factors <strong>in</strong> <strong>Physcomitrella</strong>.<br />

ETI<br />

As noted above, R-prote<strong>in</strong>s can trigger ETI by direct <strong>in</strong>teraction with pathogen effectors<br />

delivered <strong>in</strong>to <strong>the</strong> cells, or by detect<strong>in</strong>g altered self by guard<strong>in</strong>g host components or<br />

pathways targeted by <strong>the</strong> effectors (Jones and Dangl, 2006; Shen and Schulze-Lefert,<br />

2007). There are some 200 Arabidopsis genes encod<strong>in</strong>g prote<strong>in</strong>s conta<strong>in</strong><strong>in</strong>g doma<strong>in</strong>s<br />

characteristic of plant resistance prote<strong>in</strong>s (Meyers et al., 2003). These characteristics are<br />

<strong>the</strong> presence of a N-term<strong>in</strong>al region consist<strong>in</strong>g of ei<strong>the</strong>r a coiled-coil structure (CC) or a<br />

TOLL/Interleuk<strong>in</strong> 1 Receptor (TIR) doma<strong>in</strong>, followed by a nucleotide b<strong>in</strong>d<strong>in</strong>g (NB)<br />

doma<strong>in</strong>, and a C-term<strong>in</strong>al with leuc<strong>in</strong> rich repeats (LRR). These two major types are<br />

referred to as ei<strong>the</strong>r CC-NB-LRR or TIR-NB-LRR (Meyers et al., 2003). While <strong>the</strong> CC<br />

doma<strong>in</strong> is a common structural doma<strong>in</strong> found <strong>in</strong> many prote<strong>in</strong>s, <strong>the</strong> TIR doma<strong>in</strong> has<br />

homology to <strong>the</strong> <strong>in</strong>nate immunity TOLL receptor <strong>in</strong> Drosophila and to <strong>the</strong> human <strong>in</strong>nate<br />

12


immunity receptor <strong>in</strong>terleuk<strong>in</strong> 1 (Shirasu, 2009). The central NB doma<strong>in</strong> is part of a<br />

larger doma<strong>in</strong>, called NB-ARC, due to its occurrence <strong>in</strong> plant R-prote<strong>in</strong>s, <strong>the</strong> apoptotic<br />

regulator human Apoptotic Protease-Activat<strong>in</strong>g Factor 1 (APAF-1), and its<br />

Caenorhabditis elegans homolog CED-4 (van der Biezen and Jones, 1998). The LRR<br />

doma<strong>in</strong> is also found <strong>in</strong> mammalian immune receptors with strik<strong>in</strong>g similarities to <strong>the</strong><br />

plant R-prote<strong>in</strong>s (Shirasu, 2009). Some of <strong>the</strong> 200 R-prote<strong>in</strong>s <strong>in</strong> Arabidopsis lack one of<br />

<strong>the</strong>se three core doma<strong>in</strong>s, and some have additional doma<strong>in</strong>s at <strong>the</strong>ir N or C-term<strong>in</strong>i<br />

(Meyers et al., 2003).<br />

Much work has been put <strong>in</strong>to decipher<strong>in</strong>g <strong>the</strong> genetic requirements of <strong>the</strong> R-<br />

prote<strong>in</strong>s. Its has been found that <strong>the</strong> function of most CC-NB-LRR R-prote<strong>in</strong>s depends on<br />

non-race specific disease resistance 1 (NDR1) (Knepper et al., 2011b). There are,<br />

however, exceptions to this rule; several CC-NB-LRR R prote<strong>in</strong>s that specify resistance<br />

to <strong>the</strong> oomycete pathogen Hyaloperonospora arabidopsidis (Hpa) function <strong>in</strong>dependently<br />

of NDR1 (Knepper et al., 2011a). TIR-NB-LRR R-prote<strong>in</strong>s require <strong>the</strong> two lipase-like<br />

prote<strong>in</strong>s enhanced disease susceptibility 1 (EDS1) and phytoalex<strong>in</strong> deficient 4 (PAD4)<br />

and an acyl hydrolase, senescence associated gene 101 (SAG101) (Feys et al., 2005).<br />

RAR1, SGT1 and HSP90<br />

Genetic screens for loss of resistance have fur<strong>the</strong>r provided strong evidence for <strong>the</strong><br />

requirement of ano<strong>the</strong>r complex to confer resistance mediated by both CC and TIR-NB-<br />

LRRs. This complex consists of Required for mlA12 Resistance 1 (RAR1), suppressor of<br />

<strong>the</strong> G2 allele of SKP1 (SGT1), and heat shock prote<strong>in</strong> 90 (HSP90) (Shirasu, 2009). Yeast<br />

two-hybrid analysis and co-immunoprecipitation experiments on plant extracts have<br />

shown that R-prote<strong>in</strong>s <strong>in</strong>teract with RAR1, SGT1 and HSP90 and that <strong>the</strong>se three<br />

components also mutually <strong>in</strong>teract (Takahashi et al., 2003; Azevedo et al., 2006). RAR1<br />

is a s<strong>in</strong>gle copy gene <strong>in</strong> plants and <strong>the</strong> Atrar1 mutant does not exhibit o<strong>the</strong>r phenotypes<br />

than loss of R-prote<strong>in</strong> mediated resistance, <strong>in</strong>dicat<strong>in</strong>g that RAR1 functions exclusively <strong>in</strong><br />

immunity <strong>in</strong> plants (Shirasu, 2009). Arabidopsis conta<strong>in</strong>s two isoforms of SGT, AtSGT1a<br />

and AtSGT1b. The double mutant Atsgt1a/Atsgt1b is embryonic lethal, <strong>in</strong>dicat<strong>in</strong>g that<br />

SGT is essential and has functions o<strong>the</strong>r than <strong>in</strong> immunity (Azevedo et al., 2006). HSP90<br />

is a molecular chaperone and is one of <strong>the</strong> most abundant cellular prote<strong>in</strong>s. HSP90<br />

homologs are found across all k<strong>in</strong>gdoms of organisms, except for Archaea (Chen et al.,<br />

2006). RAR1 and SGT1 are thought to act as co-chaperones with HSP90 to ensure R-<br />

prote<strong>in</strong> stability (Shirasu, 2009). However, not all R-prote<strong>in</strong>s depend on <strong>the</strong> presence of a<br />

functional RAR1/SGT1/HSP90 co-chaperone (Shirasu, 2009).<br />

HR<br />

R-prote<strong>in</strong> recognition of a pathogen <strong>in</strong>duces a conformational change which is thought to<br />

start <strong>the</strong> signal<strong>in</strong>g which lead to <strong>the</strong> HR (DeYoung and Innes, 2006). However, <strong>the</strong> cha<strong>in</strong><br />

of events lead<strong>in</strong>g to <strong>the</strong> HR after effector recognition via R-prote<strong>in</strong>s is not fully<br />

elucidated (Coll et al., 2011). The molecular events that lead to <strong>the</strong> HR dur<strong>in</strong>g ETI<br />

overlap partly with those associated with MTI, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> accumulation of SA, ROS,<br />

activation of MAPK cascades, changes <strong>in</strong> <strong>in</strong>tracellular calcium levels, transcriptional<br />

reprogramm<strong>in</strong>g and syn<strong>the</strong>sis of antimicrobial compounds (Mur et al., 2008). Although<br />

CC and TIR-NB-LRRs <strong>in</strong> general require different downstream components, <strong>the</strong>ir signals<br />

13


converge <strong>in</strong> <strong>the</strong> activation of SA and ROS production, which act synergistically to drive<br />

<strong>the</strong> HR (Mur et al., 2008).<br />

In mammals, a common form of PCD is apoptosis which requires <strong>the</strong> caspase<br />

proteases. However, plants do not possess caspases, ra<strong>the</strong>r <strong>the</strong>y have caspase-like<br />

prote<strong>in</strong>s called metacaspases which cleave prote<strong>in</strong>s after arg<strong>in</strong><strong>in</strong>e or lys<strong>in</strong>e residues as<br />

opposed to cyste<strong>in</strong> and aspartate for caspases (Enoksson and Salvesen, 2010). In<br />

Arabidopsis metacaspase 1 mutants (Atmc1) suppress cell death <strong>in</strong>duced by bacterial and<br />

oomycete triggered HR (Coll et al., 2011). HR mediated by both CC and TIR-NB-LRR<br />

<strong>in</strong>tracellular immune receptors is severely attenuated <strong>in</strong> <strong>the</strong> Atmc1 plants, <strong>in</strong>dicat<strong>in</strong>g<br />

convergence of <strong>the</strong> two pathways <strong>in</strong>to a s<strong>in</strong>gle cell death output. However, at least one<br />

o<strong>the</strong>r cell death pathway exists s<strong>in</strong>ce HR <strong>in</strong>duced cell death by some TIR-NB-LRRs has<br />

been shown to require autophagy genes (Hofius et al., 2009, and manuscript 2).<br />

Apparently R-genes evolved when plants made <strong>the</strong> transition from sea to land<br />

s<strong>in</strong>ce <strong>the</strong>y are not found <strong>in</strong> an extant species represent<strong>in</strong>g <strong>the</strong> last common ancestor of all<br />

land plants, <strong>the</strong> green algae Chlamydomonas (Shirasu, 2009). <strong>Physcomitrella</strong> was<br />

recently reported to encode 18 <strong>in</strong>tact R-prote<strong>in</strong>s compared to 122 <strong>in</strong>tact R-prote<strong>in</strong>s <strong>in</strong><br />

Arabidopsis (Meyers et al., 2003; Xue et al., 2012). An <strong>in</strong>terest<strong>in</strong>g observation on <strong>the</strong><br />

evolution of R-prote<strong>in</strong>s is that monocots apparently completely lack <strong>the</strong> TIR-type R-<br />

prote<strong>in</strong>s, but still have EDS1 on which <strong>the</strong> activities of most dicot TIR R-prote<strong>in</strong>s depend<br />

(Tarr and Alexander, 2009). Thus rice (Oryza sativa) encodes >400 R-prote<strong>in</strong>s, but none<br />

of <strong>the</strong>se are of <strong>the</strong> TIR type (Wang et al., 2004).<br />

Autophagy<br />

The role of autophagy <strong>in</strong> plant defense is described <strong>in</strong> detail <strong>in</strong> manuscript 2, but I will<br />

provide a brief overview of autophagy here. Several subtypes of autophagy are described,<br />

but macroautophagy (hereafter termed autophagy) is <strong>the</strong> most extensively studied (Yang<br />

and Klionsky, 2010) and will be <strong>the</strong> only form described here. Autophagy is a conserved<br />

eukaryotic mechanism, which is classically def<strong>in</strong>ed as <strong>the</strong> degradation of cytoplasmic<br />

constituents <strong>in</strong> <strong>the</strong> lytic organelle (vacuoles <strong>in</strong> yeast and plants and lysosomes <strong>in</strong><br />

mammals) (Xie and Klionsky, 2007).<br />

It was orig<strong>in</strong>ally discovered over 40 years ago <strong>in</strong> yeast (Saccharomyces cerevisiae) as a<br />

mechanism to survive dur<strong>in</strong>g starvation (Yang and Klionsky, 2010). It has s<strong>in</strong>ce been<br />

found that <strong>the</strong> core mach<strong>in</strong>ery of autophagy is conserved throughout <strong>the</strong> eukaryote<br />

k<strong>in</strong>gdoms, and that its functions are much more versatile than just survival dur<strong>in</strong>g<br />

starvation (Liu and Bassham, 2012). Thus, autophagy has been shown to be <strong>in</strong>volved<br />

several o<strong>the</strong>r pro-survival mechanisms like clearance of misfolded prote<strong>in</strong>s, aggregated<br />

prote<strong>in</strong>s and damaged, potentially deleterious organelles. These functions play an<br />

important role <strong>in</strong> human diseases like cancer and neurodegenerative diseases (Yang and<br />

Klionsky, 2010). In plants, autophagy is also <strong>in</strong>duced by <strong>the</strong> presence of ROS which is<br />

generated when <strong>the</strong> plant is subjected to biotic or abiotic stresses (Pérez-Pérez et al.,<br />

2012). The elevated ROS levels result <strong>in</strong> damaged, oxidized prote<strong>in</strong>s which are <strong>the</strong>n<br />

degraded by autophagy (Xiong et al., 2007). It is also now known that autophagy not only<br />

functions under stress conditions but also functions at a basal level under optimal<br />

14


conditions as a house keep<strong>in</strong>g function to ma<strong>in</strong>ta<strong>in</strong> cell homeostasis (Liu and Bassham,<br />

2012).<br />

Morphologically, autophagy beg<strong>in</strong>s with <strong>the</strong> formation of cup-shaped double membranes,<br />

which expand to form autophagosomes engulf<strong>in</strong>g malfunction<strong>in</strong>g or un-needed<br />

macromolecules and organelles and transport <strong>the</strong>m for degradation <strong>in</strong>side <strong>the</strong> vacuole<br />

(Figure 1). In plants, <strong>the</strong> orig<strong>in</strong> of <strong>the</strong> double membrane is unknown (Liu and Bassham,<br />

2012). Upon arrival of <strong>the</strong> autophagosomes to <strong>the</strong> vacuoles, <strong>the</strong>ir outer membrane fuses<br />

with <strong>the</strong> tonoplast, creat<strong>in</strong>g s<strong>in</strong>gle membrane vesicles <strong>in</strong>side <strong>the</strong> vacuole, termed<br />

'autophagic bodies'. Autophagic bodies and <strong>the</strong>ir contents are <strong>the</strong>n degraded <strong>in</strong>side <strong>the</strong><br />

vacuole, provid<strong>in</strong>g recycled materials to build new macromolecules (Xie and Klionsky,<br />

2007).<br />

Figure 1. Schematic of <strong>the</strong> formation of an autophagic body. Cytosolic material is sequestered by an<br />

expand<strong>in</strong>g membrane, <strong>the</strong> phagophore, result<strong>in</strong>g <strong>in</strong> <strong>the</strong> formation of a double-membrane vesicle, an<br />

autophagosome. The outer membrane of <strong>the</strong> autophagosome subsequently fuses with <strong>the</strong> vacuole,<br />

expos<strong>in</strong>g <strong>the</strong> <strong>in</strong>ner s<strong>in</strong>gle membrane of <strong>the</strong> autophagosome to lysosomal hydrolases. The cargoconta<strong>in</strong><strong>in</strong>g<br />

membrane compartment is <strong>the</strong>n lysed, and <strong>the</strong> contents are degraded. Adapted from<br />

(Klionsky et al., 2008).<br />

In plants, autophagy is <strong>in</strong>volved <strong>in</strong> leaf senescence. When leaves senesce, <strong>the</strong> plant<br />

recovers and recycles valuable nutrient components that have been <strong>in</strong>corporated dur<strong>in</strong>g<br />

growth that would o<strong>the</strong>rwise be lost when <strong>the</strong> leaf dies or is shed (Figure 2). Chloroplasts<br />

conta<strong>in</strong> approximately 80% of total leaf nitrogen and represent a major source of recycled<br />

nitrogen dur<strong>in</strong>g leaf senescence (Mak<strong>in</strong>o and Osmond, 1991). Chloroplasts are thus taken<br />

up by <strong>the</strong> vacuole and degraded <strong>in</strong> an autophagy dependent manner dur<strong>in</strong>g leaf<br />

senescence, s<strong>in</strong>ce <strong>the</strong> Arabidopsis autophagy deficient mutant Atatg4a4b-1 was unable to<br />

degrade chloroplasts dur<strong>in</strong>g senescence (Wada et al., 2009). In an assay of gene<br />

expression dur<strong>in</strong>g senescence it was found that 15 Arabidopsis genes <strong>in</strong>volved <strong>in</strong><br />

autophagy are up regulated, show<strong>in</strong>g <strong>the</strong> key role of autophagy <strong>in</strong> <strong>the</strong> controlled<br />

degradation of cellular components (Breeze et al., 2011).<br />

Not only <strong>in</strong> plants, but also <strong>in</strong> yeast and animals, autophagy mutants often display<br />

early senescence under nutrient deficient conditions. Arabidopsis atg2, atg5, atg4s, atg6,<br />

atg7, atg8s, atg9, atg10, atg12s and atg18a mutants all show early senescence <strong>in</strong> carbon<br />

or nitrogen deficient conditions, and some even under nutrient-rich conditions (Liu and<br />

Bassham, 2012).<br />

15


Figure 2. Senescence leaves of Quercus dentata, <strong>the</strong> Daimyo Oak. In a controlled process that <strong>in</strong>volves<br />

autophagy, cellular content is degraded and nutrients transported via <strong>the</strong> vascular system <strong>in</strong>to <strong>the</strong><br />

stem before <strong>the</strong> leaves are shed. Senescence proceeds from leaf marg<strong>in</strong>s toward <strong>the</strong> center. Note that<br />

cells surround<strong>in</strong>g <strong>the</strong> vascular tissues senesce relatively late to facilitate nutrient mobilization from<br />

adjacent, senesc<strong>in</strong>g cells.<br />

Besides its pro-survival functions autophagy has also been found have a pro-death<br />

function <strong>in</strong> an autophagy dependent PCD pathway conserved across <strong>the</strong> Eukaryotic<br />

k<strong>in</strong>gdoms (Yang and Klionsky, 2010). Autophagic PCD has been shown to be essential<br />

to specific developmental processes <strong>in</strong> Drosophila, C. elegans and Arabidopsis (See text<br />

and Figure 2, manuscript 2).<br />

Hofius et al. (2009) showed that TIR-NB-LRR mediated cell death was autophagy<br />

dependent s<strong>in</strong>ce <strong>the</strong> TIR-NB-LRR RPS4-mediated HR response <strong>in</strong> Atatg7 and Atatg9<br />

was compromised <strong>in</strong> response to Pto DC3000 carry<strong>in</strong>g <strong>the</strong> effector avrRps4. This result<br />

contradicted previously published results show<strong>in</strong>g that autophagy was required for<br />

restrict<strong>in</strong>g <strong>the</strong> spread of cell death <strong>in</strong>duced by pathogen effector recognition (Patel and<br />

D<strong>in</strong>esh-Kumar, 2008; Yoshimoto et al., 2009). The apparent discrepancy between <strong>the</strong>se<br />

results is discussed <strong>in</strong> more detail <strong>in</strong> manuscript 2. In brief, <strong>the</strong> discrepancy may be due to<br />

<strong>the</strong> age of <strong>the</strong> plants used. In older, autophagy deficient plants, toxic compounds<br />

accumulate that would have been cleared <strong>in</strong> wild type plants. This accumulation of toxic<br />

compounds make <strong>the</strong> cells vulnerable to pathogen attack and thus <strong>the</strong> apparent <strong>in</strong>creased<br />

cell death <strong>in</strong> older autophagy deficient plants could be due to accumulative effects of not<br />

hav<strong>in</strong>g a functional autophagy house-keep<strong>in</strong>g system.<br />

Lenz et al. (2011) recently showed that Arabidopsis autophagy deficient plants are less<br />

susceptible to <strong>the</strong> biotrophic Pto DC3000 bacteria while <strong>the</strong>y are more susceptible to <strong>the</strong><br />

necrotrophic fungus A. brassicicola. They suggested that this difference could be due to<br />

slightly elevated levels of <strong>the</strong> phytohormone SA which promotes HR. Lai et al. (2011)<br />

also found that autophagy deficient Arabidopsis plants were more susceptible to <strong>the</strong><br />

necrotrophic fungus B. c<strong>in</strong>erea and that <strong>the</strong> un<strong>in</strong>fected autophagy mutant had elevated<br />

SA levels. Fur<strong>the</strong>rmore Lai et al. (2011) showed that <strong>the</strong> transcription factor WRKY33,<br />

which is important for MTI aga<strong>in</strong>st necrotrophic pathogens (Zheng et al., 2006),<br />

<strong>in</strong>teracted with ATG18a, thus po<strong>in</strong>t<strong>in</strong>g to a possible regulatory l<strong>in</strong>k between autophagy<br />

and defense aga<strong>in</strong>st necrotrophic pathogens.<br />

16


Autophagy has been demonstrated <strong>in</strong> <strong>the</strong> green algae model Chlamydomonas re<strong>in</strong>hardtii<br />

(Pérez-Pérez et al., 2010), but to our knowledge <strong>the</strong>re have been no studies published on<br />

any aspect of autophagy <strong>in</strong> <strong>Physcomitrella</strong> or any o<strong>the</strong>r nonvascular land plant. Thus <strong>the</strong><br />

autophagy deficient <strong>Physcomitrella</strong> mutant described <strong>in</strong> this report represents <strong>the</strong> first<br />

autophagic analysis <strong>in</strong> a non-vascular plant. However, s<strong>in</strong>ce <strong>the</strong> core mach<strong>in</strong>ery of<br />

autophagy is conserved throughout <strong>the</strong> Eukaryote k<strong>in</strong>gdoms, <strong>Physcomitrella</strong> should be a<br />

very good model organism <strong>in</strong> which to study this phenomenon due to its unique ability to<br />

be manipulated through reverse genetics.<br />

<strong>Physcomitrella</strong> <strong>patens</strong><br />

Evolution<br />

Bryophytes pioneered and modified <strong>the</strong> terrestrial environment 475 MYA (Wellman et al.,<br />

2003), and thus paved <strong>the</strong> way for all o<strong>the</strong>r terrestrial life forms. Bryophytes developed<br />

from an ancestor most closely related to modern green algae (Lewis and McCourt, 2004)<br />

and after mak<strong>in</strong>g <strong>the</strong> transition onto land <strong>the</strong> early land plants diverged <strong>in</strong>to l<strong>in</strong>eages<br />

adapt<strong>in</strong>g to <strong>the</strong> different habitats of <strong>the</strong> terrestrial environment. Mosses and seed plants<br />

shared <strong>the</strong>ir last common ancestor at least 420 MYA (Clarke et al., 2011), mak<strong>in</strong>g <strong>the</strong><br />

evolutionary distance between <strong>Physcomitrella</strong> and Arabidopsis similar to that of fishes<br />

and humans (Hedges, 2002). Fossil records show that <strong>the</strong> macro morphology of extant<br />

<strong>moss</strong>es appears unchanged s<strong>in</strong>ce <strong>the</strong> earliest preserved fossil records from 320 MYA<br />

(Hübers and Kerp, 2012).<br />

Mov<strong>in</strong>g from <strong>the</strong> aquatic environment to <strong>the</strong> terrestrial led to many fundamental<br />

changes. Drought, UV-light exposure and chang<strong>in</strong>g temperatures are among <strong>the</strong> major<br />

environmental differences that required adaptations.<br />

Figure 3. Land plant evolution. A schematic overview of <strong>the</strong> adaptations <strong>in</strong> <strong>the</strong> different l<strong>in</strong>eages of<br />

landplants. From (Rens<strong>in</strong>g et al., 2008)<br />

S<strong>in</strong>ce <strong>moss</strong>es are amongst <strong>the</strong> earliest land plants <strong>the</strong>y present an evolutionary l<strong>in</strong>k<br />

between green algae and flower<strong>in</strong>g plants. This makes <strong>the</strong>m ideal for study<strong>in</strong>g <strong>the</strong><br />

evolutionary changes required to conquer land. These chances <strong>in</strong>cluded adaptations to <strong>the</strong><br />

new abiotic stresses <strong>in</strong>clud<strong>in</strong>g enhanced osmoregulation and osmoprotection, desiccation<br />

and freez<strong>in</strong>g tolerance, heat resistance, syn<strong>the</strong>sis and accumulation of protective<br />

17


“sunscreens”, and enhanced DNA repair mechanisms (Rens<strong>in</strong>g et al., 2008). However, it<br />

can be assumed that plants mov<strong>in</strong>g from water to land were accompanied by pathogens<br />

many of which evolved airborne spores. This started <strong>the</strong> evolutionary arms race that<br />

resulted <strong>in</strong> advanced pathogen weapons and attack strategies and coevolved defense<br />

systems <strong>in</strong> <strong>the</strong> plants. <strong>Physcomitrella</strong> is thus <strong>in</strong> an ideal evolutionary position to provide<br />

useful <strong>in</strong>sights to <strong>the</strong> evolution of plant <strong>in</strong>nate immunity.<br />

The life cycle and morphology of <strong>Physcomitrella</strong><br />

Like o<strong>the</strong>r bryophytes, <strong>Physcomitrella</strong> exhibits a reverse alternation of generations<br />

compared to vascular plants. Thus, <strong>the</strong> gametophyte (haploid) generation comprises most<br />

of <strong>the</strong> plant while <strong>the</strong> sporophyte (diploid) generation is very small and dependent upon<br />

<strong>the</strong> haploid shoot (gametophore) (Figure 4). Follow<strong>in</strong>g meiosis, haploid spores (Figure<br />

4A) germ<strong>in</strong>ate <strong>in</strong>to filamentous networks of cells with apical growth compris<strong>in</strong>g <strong>the</strong><br />

protonema stage. The protonemal filaments consist of two cell types: chloronemata with<br />

many fully developed chloroplasts (Figure 4B) and caulonemata with faster, <strong>in</strong>vasive<br />

growth and fewer, smaller chloroplasts (Figure 4C). When <strong>moss</strong> is grown on medium<br />

overlaid with cellophane <strong>the</strong> protonema stage is prolonged and formation of<br />

gametophores is delayed compared to growth on medium without cellophane. At some<br />

po<strong>in</strong>t a branch of chloronemal cells will differentiate and form a gametophore (Figure<br />

4D). The gametophore is a leafy shoot composed of a nonvascular stem with leafs and<br />

rhizoids. On top of a gametophore both male (an<strong>the</strong>ridia) and female (archegonia) sexual<br />

organs form (Figure 4E). If water is present, motile spermatozoids can swim from <strong>the</strong><br />

an<strong>the</strong>ridia to <strong>the</strong> archegonium and fertilize <strong>the</strong> eggs. The result<strong>in</strong>g zygote develops <strong>in</strong>to<br />

<strong>the</strong> diploid sporophore (Figure 4F). The sporophyte needs to divert water and nutrients<br />

from <strong>the</strong> parental gametophyte. While <strong>the</strong> gametophyte does not have stomata, <strong>the</strong>re are<br />

fully functional stomata on <strong>the</strong> diploid sporangium (Chater et al., 2011). Meioses <strong>in</strong> <strong>the</strong><br />

spore capsule with<strong>in</strong> <strong>the</strong> sporangium produces approximately 4000 haploid spores.<br />

When grown <strong>in</strong> <strong>the</strong> lab, <strong>the</strong> life cycle of <strong>Physcomitrella</strong> can be completed with<strong>in</strong><br />

approximately 12 weeks. Cold conditions with short days of low light mimic autumnal<br />

seasonal change and <strong>in</strong>duce sporophyte formation (Hohe et al., 2002). However, for most<br />

experimental procedures <strong>the</strong>re is no need to complete <strong>the</strong> life cycle of <strong>Physcomitrella</strong> as<br />

it is easily propagated vegetatively. This is because any part of <strong>the</strong> plant will differentiate<br />

<strong>in</strong>to chloronemal cells and produce a new colony when mechanically disrupted.<br />

18


Figure 4. (A) A haploid spore germ<strong>in</strong>ates <strong>in</strong>to (B) chloronemal cells, which cont<strong>in</strong>ue to grow and<br />

differentiate <strong>in</strong>to (C) caulonemal cells. (D) Gametophores, or shoots, emerge off protonemal<br />

filaments and are ultimately anchored by rhizoids that grow by tip growth from <strong>the</strong> base of <strong>the</strong><br />

gametophore. (E) At <strong>the</strong> apex of <strong>the</strong> gametophore, both female, archegonia (arrows) and male,<br />

an<strong>the</strong>ridia (arrowheads) form. A motile flagellate sperm fertilizes <strong>the</strong> egg and <strong>the</strong> (F) sporophyte<br />

(marked with a bracket) develops at <strong>the</strong> apex of <strong>the</strong> gametophore. From (Prigge and Bezanilla, 2010).<br />

Like o<strong>the</strong>r <strong>moss</strong>es, <strong>Physcomitrella</strong> has a relatively simply morphology compared to<br />

flower<strong>in</strong>g plants. The lack of roots and vascular tissue restricts <strong>the</strong>ir size. Both <strong>the</strong> leaves,<br />

rhizoids and protonemal filaments consist of only one layer of cells without a fully<br />

developed cuticle to protect from water loss. Thus <strong>the</strong>y are poikilohydric such that <strong>the</strong><br />

water content of <strong>the</strong> plant is <strong>in</strong> equilibrium with <strong>the</strong> water content of <strong>the</strong> environment.<br />

Though most <strong>moss</strong>es can tolerate some dehydration, <strong>the</strong>y are dependent on a ra<strong>the</strong>r moist<br />

environment. S<strong>in</strong>ce <strong>the</strong> leaves of <strong>Physcomitrella</strong> consist of just one layer of cells <strong>the</strong>re is<br />

no need for stomata. However, <strong>the</strong> sporophyte which develops <strong>the</strong> multilayered<br />

sporangium has stomata (Chater et al., 2011).<br />

Homologous recomb<strong>in</strong>ation<br />

Much of <strong>the</strong> knowledge of modern genetics has been ga<strong>in</strong>ed from <strong>the</strong> discovery of gene<br />

target<strong>in</strong>g through homologous recomb<strong>in</strong>ation <strong>in</strong> <strong>the</strong> yeast Saccharomyces cerevisiae back<br />

<strong>in</strong> <strong>the</strong> late 1970s (Struhl et al., 1979). In 1991, a protocol was published for genetic<br />

transformation of <strong>Physcomitrella</strong> protoplasts (Schaefer et al., 1991). This protocol is<br />

based on <strong>the</strong> same techniques as for yeast transformation: Direct uptake of DNA <strong>in</strong> <strong>the</strong><br />

presence of polyethylene glycol (PEG) followed by heat chock. To achieve efficient gene<br />

target<strong>in</strong>g through homologous recomb<strong>in</strong>ation, <strong>the</strong> homologous sequences flank<strong>in</strong>g <strong>the</strong><br />

selection cassette have to be 500-1000 bp <strong>in</strong> <strong>Physcomitrella</strong> compared to just 30-60 bp <strong>in</strong><br />

S. cerevisiae (Schaefer and Zrÿd, 1997). However, with such longer flank<strong>in</strong>g regions <strong>the</strong><br />

19


homologous recomb<strong>in</strong>ation rate <strong>in</strong> <strong>Physcomitrella</strong> is <strong>the</strong> highest reported for any<br />

multicellular organism mak<strong>in</strong>g it a very useful model organism for do<strong>in</strong>g reverse genetic<br />

studies (Schaefer and Zrÿd, 1997). This technique has successfully enabled targeted<br />

knock out (KO) or targeted <strong>in</strong>sert (KI) or targeted mutation down to one nucleotide,<br />

mak<strong>in</strong>g <strong>Physcomitrella</strong> a model to provide useful <strong>in</strong>formation <strong>in</strong> a wide range of research<br />

fields. A schematic draw<strong>in</strong>g of <strong>the</strong> pr<strong>in</strong>ciple <strong>in</strong> targeted KO through homolog<br />

recomb<strong>in</strong>ation can be seen <strong>in</strong> Figure 16 <strong>in</strong> <strong>the</strong> result section.<br />

The genome of <strong>Physcomitrella</strong><br />

The genome of <strong>Physcomitrella</strong> <strong>patens</strong> was published <strong>in</strong> 2008 (Rens<strong>in</strong>g et al., 2008). It<br />

consists of 27 chromosomes with a total size of 480 Mb, which is almost four times <strong>the</strong><br />

size of <strong>the</strong> small genome of Arabidopsis thaliana (125 Mb, 5 chromosomes) (The<br />

Arabidopsis Genome Initiative, 2000). Accord<strong>in</strong>g to <strong>the</strong> latest gene annotation version<br />

(V1.6 at www.cos<strong>moss</strong>.org ) <strong>Physcomitrella</strong> <strong>patens</strong> has 38.357 prote<strong>in</strong> cod<strong>in</strong>g genes<br />

compared to <strong>the</strong> 27.416 of Arabidopsis thaliana (TAIR 10 at www.Arabidopsis.org ).<br />

Whole genome duplications have occurred <strong>in</strong> both <strong>the</strong> Arabidopsis and<br />

<strong>Physcomitrella</strong> l<strong>in</strong>eages. This makes it difficult to establish orthologous relationships<br />

between genes <strong>in</strong> <strong>the</strong>se two evolutionary distant species, especially for members of<br />

multigene families. It is hypo<strong>the</strong>sized that <strong>the</strong> 27 chromosomes of <strong>Physcomitrella</strong> have<br />

arisen from an orig<strong>in</strong>al chromosome number of 7 that duplicated twice to 28 and a s<strong>in</strong>gle<br />

chromosome was subsequently lost. The last whole genome duplication event has been<br />

estimated to have occurred ~45 MYA i.e. after <strong>the</strong> diversification from <strong>the</strong> last common<br />

shared ancestor of Arabidopsis (Rens<strong>in</strong>g et al., 2007). Arabidopsis has also undergone<br />

two or more rounds of whole genome duplication s<strong>in</strong>ce <strong>the</strong> diversification from <strong>the</strong> last<br />

common ancestor of <strong>moss</strong>es, <strong>the</strong> most recent duplication be<strong>in</strong>g ~112 MYA (Ku et al.,<br />

2000), as well as numerous local gene duplications and gene losses (The Arabidopsis<br />

Genome Initiative, 2000).<br />

Despite <strong>the</strong>se major genomic rearrangements and more than 400 million years of<br />

evolution, only about 130 genes are reta<strong>in</strong>ed <strong>in</strong> <strong>Physcomitrella</strong> that do not have clear<br />

homologs <strong>in</strong> flower<strong>in</strong>g plants (Rens<strong>in</strong>g et al., 2005). Thus, much of “what it takes to be a<br />

plant” is present <strong>in</strong> <strong>Physcomitrella</strong>. In fact, many basic cellular processes are conserved<br />

throughout <strong>the</strong> Eukaryote k<strong>in</strong>gdoms and thus <strong>the</strong> knowledge ga<strong>in</strong>ed from comparative<br />

and evolutionary studies of <strong>Physcomitrella</strong> can provide <strong>in</strong>sights <strong>in</strong>to related processes <strong>in</strong><br />

a wide range of o<strong>the</strong>r organisms from crop plants to humans.<br />

Summary of <strong>the</strong> strengths of <strong>Physcomitrella</strong> as a model organism<br />

• It is easily grown <strong>in</strong> simple sterile medium <strong>in</strong> Petri dishes and propagated<br />

vegetatively<br />

• The genome has been sequenced and resources are accessible at<br />

www.cos<strong>moss</strong>.org , www.phytozome.net or www.ncbi.nlm.nih.gov<br />

20


• The occurrence, at a very high frequency, of homologous recomb<strong>in</strong>ation between<br />

sequences <strong>in</strong> transform<strong>in</strong>g DNA and <strong>the</strong> correspond<strong>in</strong>g genomic sequences,<br />

allows for targeted gene <strong>in</strong>activation and for allele modification<br />

• The haploidy of <strong>the</strong> dom<strong>in</strong>ant phase of <strong>the</strong> <strong>moss</strong> life cycle allows <strong>the</strong> direct<br />

recognition of mutants hav<strong>in</strong>g a recessive phenotype<br />

• The simple morphology with monolayered leaves make it very suitable for<br />

microscopy and subcellular localization studies.<br />

Innate immunity <strong>in</strong> <strong>Physcomitrella</strong><br />

The study of <strong>Physcomitrella</strong> pathology is <strong>in</strong> its <strong>in</strong>fancy. To date, only a few studies on<br />

pathogenesis <strong>in</strong> <strong>Physcomitrella</strong> have been done, and <strong>the</strong>y all <strong>in</strong>volve broad host range<br />

necrotrophic pathogens (Andersson et al., 2005; Ponce de León et al., 2007; Lehtonen et<br />

al., 2009; Oliver et al., 2009; Akita et al., 2011). However, many of <strong>the</strong> molecular and<br />

cellular responses of <strong>Physcomitrella</strong> after pathogen <strong>in</strong>fection are similar to those<br />

observed <strong>in</strong> flower<strong>in</strong>g plants <strong>in</strong>clud<strong>in</strong>g production of ROS, altered gene expression,<br />

production of secondary metabolites, and hallmarks of PCD.<br />

<strong>Physcomitrella</strong> responses to pathogens<br />

<strong>Physcomitrella</strong> <strong>in</strong>fection with <strong>the</strong> necrotrophic fungus B. c<strong>in</strong>erea and <strong>the</strong> necrotrophic<br />

oomycete Pythium causes an <strong>in</strong>crease <strong>in</strong> ROS and SA production (Ponce de León et al.,<br />

2007; Oliver et al., 2009; Ponce De León et al., 2012). Along with ROS and SA<br />

production, o<strong>the</strong>r hallmarks of PCD <strong>in</strong>clude cytoplasmic shr<strong>in</strong>kage, accumulation of<br />

autofluorescent compounds and chloroplast breakdown (Ponce De León et al., 2012). The<br />

apparent activation of HR-like PCD after <strong>in</strong>fection with necrotrophic pathogens could be<br />

a result of <strong>the</strong> <strong>in</strong>vasion strategy of <strong>the</strong> pathogen by which it manipulates <strong>the</strong> host defense<br />

system to <strong>in</strong>duce localized host cell death to facilitate its own growth (Govr<strong>in</strong> and Lev<strong>in</strong>e,<br />

2000; Govr<strong>in</strong> et al., 2006; Frías et al., 2011).<br />

While HR seems like an <strong>in</strong>appropriate defense response aga<strong>in</strong>st necrotrophic<br />

<strong>in</strong>truders, <strong>in</strong>fection of <strong>Physcomitrella</strong> with B. c<strong>in</strong>erea and Pythium also <strong>in</strong>duces more<br />

appropriate defense responses like cell wall modifications <strong>in</strong> <strong>the</strong> form of callose<br />

deposition and accumulation of phenolic compounds (Ponce de León et al., 2007; Oliver<br />

et al., 2009). Infection with B. c<strong>in</strong>erea and Pythium also results <strong>in</strong> up-regulation of<br />

several defense related genes, <strong>in</strong>clud<strong>in</strong>g PpPAL, PpCHS, PpLOX and PpPR-1 (Ponce de<br />

León et al., 2007; Oliver et al., 2009).<br />

<strong>Physcomitrella</strong> responses to MAMPs<br />

Treatment of <strong>Physcomitrella</strong> with cell free pathogen culture filtrates and purified<br />

MAMPs <strong>in</strong>duces defense responses, <strong>in</strong>dicat<strong>in</strong>g a functional MTI system (Ponce de León<br />

et al., 2007; Lehtonen et al., 2009, 2012). Ponce de León et al. (2007) observed <strong>the</strong>se<br />

responses when treat<strong>in</strong>g <strong>Physcomitrella</strong> protonemal tissue with cell free culture filtrates<br />

of two different stra<strong>in</strong>s of Pectobacterium carotovorum ssp. carotovorum (P.c.<br />

21


carotovorum, formerly named Erw<strong>in</strong>ia carotovora ssp. carotovora), <strong>the</strong> P.c. carotovorum<br />

SCC1 stra<strong>in</strong> which produces <strong>the</strong> elicitor HrpN, and <strong>the</strong> P.c. carotovorum SCC3193 stra<strong>in</strong><br />

which is a harp<strong>in</strong> (HrpN)-negative stra<strong>in</strong>. The culture filtrates of both stra<strong>in</strong>s <strong>in</strong>duced<br />

defense responses <strong>in</strong> <strong>Physcomitrella</strong>, but filtrates of <strong>the</strong> harp<strong>in</strong>-secret<strong>in</strong>g stra<strong>in</strong> <strong>in</strong>duced<br />

much stronger responses than filtrates of <strong>the</strong> HrpN-negative stra<strong>in</strong>. These responses<br />

<strong>in</strong>cluded up regulation of defense related genes and signs of PCD <strong>in</strong>clud<strong>in</strong>g cytoplasmic<br />

shr<strong>in</strong>kage, and <strong>the</strong> accumulation of ROS and auto fluorescent compounds (Ponce de León<br />

et al., 2007).<br />

Lehtonen et al. (2009, 2012) observed that treat<strong>in</strong>g <strong>Physcomitrella</strong> with <strong>the</strong> fungal<br />

MAMP chitosan caused a rapid <strong>in</strong>crease <strong>in</strong> peroxidase activity and a ROS burst. They<br />

discovered that <strong>the</strong> peroxidase activity caused by a s<strong>in</strong>gle peroxidase, PpPRX34. The KO<br />

of PpPRX34 was unable to produce a ROS burst upon chitosan treatment and rendered<br />

plants more susceptible to two different pathogenic fungi (Lehtonen et al., 2009, 2012).<br />

These experiments <strong>in</strong>dicate that MTI is conserved between <strong>Physcomitrella</strong> and flower<strong>in</strong>g<br />

plants. However, until now noth<strong>in</strong>g is known about how MAMPs are perceived, how <strong>the</strong><br />

signal is transduced to <strong>the</strong> nucleus, and how <strong>the</strong> MAMP <strong>in</strong>duced defense responses are<br />

regulated <strong>in</strong> <strong>Physcomitrella</strong>.<br />

Whe<strong>the</strong>r <strong>Physcomitrella</strong> posses a functional ETI with R-prote<strong>in</strong>s is also not known, but<br />

<strong>the</strong> clear signs of HR-like PCD and <strong>the</strong> discovery of R-prote<strong>in</strong> homologs of functional R-<br />

prote<strong>in</strong>s <strong>in</strong> flower<strong>in</strong>g plants <strong>in</strong>dicates that <strong>the</strong> <strong>moss</strong> has a functional ETI system (Xue et<br />

al., 2012).<br />

22


Results<br />

Identification of <strong>Physcomitrella</strong> homologs of defense related<br />

Arabidopsis genes<br />

13 genes of <strong>in</strong>terest from Arabidopsis were chosen as candidates for creat<strong>in</strong>g targeted<br />

KOs of <strong>the</strong>ir closest <strong>Physcomitrella</strong> homologs. The 13 genes are <strong>the</strong> n<strong>in</strong>e listed <strong>in</strong> Table 1<br />

plus two R-genes of <strong>the</strong> CC-NB-LRR type and two of <strong>the</strong> TIR-NB-LRR type.<br />

AtCERK1 AT3G21630<br />

AtMEKK1 AT4G08500<br />

AtMKK2 AT4G29810<br />

AtMKK1 AT4G26070<br />

AtMPK4 AT4G01370<br />

AtSGT1 AT4G11260<br />

AtRAR1 AT5G51700<br />

AtACD11 AT2G34690<br />

AtATG5 AT5G17290<br />

Table 1. Arabidopsis genes of <strong>in</strong>terest and <strong>the</strong>ir TAIR IDs. Bold means a <strong>Physcomitrella</strong> ortholog<br />

could be identified by <strong>the</strong> reciprocal best hit method.<br />

For four of <strong>the</strong> 13 Arabidopsis genes, an ortholog <strong>in</strong> <strong>Physcomitrella</strong> could be identified<br />

by <strong>the</strong> reciprocal best hit method (Table 1) (Moreno-Hagelsieb and Latimer, 2008). This<br />

means that if, when perform<strong>in</strong>g a reverse BLASTp search aga<strong>in</strong>st <strong>the</strong> Arabidopsis genes<br />

with <strong>the</strong> best <strong>Physcomitrella</strong> hit as query, <strong>the</strong> best hit <strong>in</strong> Arabidopsis is <strong>the</strong> same as <strong>the</strong><br />

orig<strong>in</strong>al query, <strong>the</strong>n <strong>the</strong> two genes are considered orthologous. Analyses of <strong>the</strong> KO of<br />

such <strong>in</strong> silico orthologs may clarify if <strong>the</strong>y are <strong>in</strong>deed functional orthologs.<br />

For <strong>the</strong> rema<strong>in</strong><strong>in</strong>g n<strong>in</strong>e genes of <strong>in</strong>terest, more than one homologous gene with<br />

similar identity could be <strong>the</strong> functional orthologs. Therefore, more than one candidate<br />

was chosen for KOs. This selection was done by construct<strong>in</strong>g <strong>the</strong> phylogenetic<br />

relationship between <strong>the</strong> gene of <strong>in</strong>terest and its closest homologs <strong>in</strong> both Arabidopsis<br />

and <strong>Physcomitrella</strong>, and sometimes also by <strong>in</strong>clud<strong>in</strong>g homologs from o<strong>the</strong>r species.<br />

CERK1<br />

AtCERK1 belongs to <strong>the</strong> very large family of receptor like k<strong>in</strong>ases with >600 members <strong>in</strong><br />

Arabidopsis, >1100 <strong>in</strong> rice and 157 <strong>in</strong> <strong>Physcomitrella</strong> (Shiu et al., 2004; Vij et al., 2008).<br />

With <strong>the</strong> large number of genes and <strong>the</strong> great expansion of <strong>the</strong> gene family through<br />

evolution, it is expected that a clear one to one orthologous connection can not be<br />

established. Figure 5 shows <strong>the</strong> phylogenetic relationship between <strong>the</strong> CERK1 and<br />

CEBiP homologs with human IRAK1 as outgroup. There are no closely related homologs<br />

to <strong>the</strong> rice chit<strong>in</strong> elicitor-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> (OsCEBiP) and its three homologs <strong>in</strong><br />

Arabidopsis (AtLYM1-3). However, <strong>the</strong>re are four <strong>Physcomitrella</strong> RLKs with extensive<br />

similarity to AtCERK1. They were termed PpCERK1A-D (bolded <strong>in</strong> Figure 5). From this<br />

23


phylogenetic relationship it is not possible to establish a s<strong>in</strong>gle candidate as <strong>the</strong> functional<br />

ortholog of AtCERK1. Therefore, <strong>the</strong> phylogenetic relationship was also created us<strong>in</strong>g<br />

DNA and truncated am<strong>in</strong>o acid sequences conta<strong>in</strong><strong>in</strong>g only conserved doma<strong>in</strong>s. However,<br />

<strong>the</strong>se phylogenetic trees also did not reveal if any of <strong>the</strong> four candidates should be<br />

omitted as a putative functional ortholog (data not shown).<br />

Figure 5. Phylogenetic relationship of CERK1 and CeBiP homologs <strong>in</strong> Pp=<strong>Physcomitrella</strong> <strong>patens</strong>,<br />

Sm=Selag<strong>in</strong>ella moellendorffii, Os=Oryza sativa, At=Arabidopsis thaliana, Hs=Homo sapiens. Human<br />

IRAK1 <strong>in</strong>cluded as outgroup.<br />

The mRNAs of MAMP receptors such as FLS2, EFR and CERK1 are know to be upregulated<br />

at <strong>the</strong> level of gene expression upon perception of <strong>the</strong>ir respective MAMPs<br />

(Hruz et al., 2008). We <strong>the</strong>refore wanted to see if any of <strong>the</strong> four candidates responded by<br />

<strong>in</strong>creased expression upon chitosan treatment. As seen <strong>in</strong> Figure 6, PpCERK1A and<br />

PpCERK1B mRNA levels were <strong>in</strong>duced 8-11 fold with<strong>in</strong> one hour of chitosan treatment<br />

relative to <strong>the</strong> untreated control (0h). PpCERK1C was slightly <strong>in</strong>duced ~2 fold, while<br />

PpCERK1D did not show any <strong>in</strong>duction. Note that <strong>the</strong> graphs represent a s<strong>in</strong>gle q-PCR<br />

run and some technical variation between runs must be expected. Never<strong>the</strong>less,<br />

PpCERK1A and PpCERK1B seem to respond much more to <strong>the</strong> treatment compared to<br />

<strong>the</strong> o<strong>the</strong>r two and were <strong>the</strong>refore chosen as candidates for creat<strong>in</strong>g targeted KOs.<br />

24


Relative expression<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

CERK1A<br />

CERK1B<br />

CERK1C<br />

CERK1D<br />

0h 15m 30m 1h 2h 4h 8h<br />

Figure 6. Expression of <strong>the</strong> four CERK1 homologs <strong>in</strong> <strong>Physcomitrella</strong> upon chitosan treatment <strong>in</strong> wild<br />

type <strong>moss</strong> relative to untreated (0h). The figure is based on a s<strong>in</strong>gle technical replicate.<br />

MEKK1<br />

In Arabidopsis, <strong>the</strong>re are 10 MP3K members of <strong>the</strong> MEKK type k<strong>in</strong>ase family and two<br />

MEKK-like genes (Suarez-Rodriguez et al., 2010). There are no publications on how<br />

many MEKKs exists <strong>in</strong> <strong>Physcomitrella</strong>. Thus, to f<strong>in</strong>d a homolog of AtMEKK1 <strong>in</strong><br />

<strong>Physcomitrella</strong> we first identified all sequences conta<strong>in</strong><strong>in</strong>g <strong>the</strong> conserved signature<br />

catalytic motif of MEKK type k<strong>in</strong>ases (Figure 7).<br />

Figure 7. WebLogo of <strong>the</strong> conserved consensus motif <strong>in</strong> <strong>the</strong> catalytic doma<strong>in</strong> of MEKK type plant<br />

k<strong>in</strong>ases. (from Suarez-Rodriguez et al., 2010; I created this figure for that publication, as mentioned<br />

<strong>in</strong> its acknowledgments).<br />

Perform<strong>in</strong>g BLAST queries aga<strong>in</strong>st <strong>Physcomitrella</strong> prote<strong>in</strong> databases at<br />

www.cos<strong>moss</strong>.org, with both full length and <strong>the</strong> conserved doma<strong>in</strong> of Arabidopsis<br />

MEKKs, MEKK-like and RAF type MP3Ks, revealed 17 prote<strong>in</strong>s with <strong>the</strong> signature<br />

motif of plant MP3Ks. In order to identify <strong>the</strong> homologs with <strong>the</strong> highest similarity to<br />

AtMEKK1, phylogenetic relationships were established. Due to high variation <strong>in</strong> <strong>the</strong> N-<br />

term<strong>in</strong>al regions of <strong>the</strong> prote<strong>in</strong>s, <strong>the</strong> full length sequences did not align well. Thus,<br />

truncated versions conta<strong>in</strong><strong>in</strong>g <strong>the</strong> conserved catalytic doma<strong>in</strong> were used to perform a<br />

multiple alignment us<strong>in</strong>g ClustalW2, from which a phylogenetic tree was created us<strong>in</strong>g<br />

Treeview (Figure 8). This phylogenetic relationship reveals that <strong>Physcomitrella</strong> has some<br />

10 MP3Ks with similarity to <strong>the</strong> MEKK type, at least four with similarity to <strong>the</strong> RAF<br />

type, and three that show higher similarity to <strong>the</strong> MEKK-like. Only members of <strong>the</strong><br />

MEKK1 type constitute canonical MP3Ks that activate a MAP k<strong>in</strong>ase cascade (Suarez-<br />

Rodriguez et al., 2010). S<strong>in</strong>ce we were search<strong>in</strong>g for an AtMEKK1 ortholog, this search<br />

25


should not be considered thorough regard<strong>in</strong>g <strong>the</strong> number RAF like MP3Ks <strong>in</strong><br />

<strong>Physcomitrella</strong>: <strong>the</strong>re are probably more members of this family than shown <strong>in</strong> Figure 8.<br />

Figure 8. Phylogenetic relationship of 17 MP3K homologs <strong>in</strong> <strong>Physcomitrella</strong> and <strong>the</strong> 10 Arabidopsis<br />

MEKK type MP3Ks, two members of <strong>the</strong> RAF type MP3Ks (AtEDR1 and AtCTR1) and members of<br />

<strong>the</strong> MEKK-like family (AtMP3Ke1 and AtMP3Ke2). Human RAF1 and MEKK3 are <strong>in</strong>cluded as<br />

outgroups. Pp=<strong>Physcomitrella</strong> <strong>patens</strong>, At=Arabidopsis thaliana, Hs=Homo sapiens.<br />

From <strong>the</strong> phylogenetic tree (Figure 8) it can be seen that <strong>the</strong>re are two <strong>Physcomitrella</strong><br />

MP3K homologs with similar identity to Arabidopsis AtMEKK1, here termed<br />

PpMEKK1A and PpMEKK1B (bolded <strong>in</strong> Figure 8). Perform<strong>in</strong>g multiple alignments<br />

with full length sequences of <strong>the</strong>se two with AtMEKK1 did not favor ei<strong>the</strong>r one.<br />

Therefore, I assayed if ei<strong>the</strong>r of <strong>the</strong> two responded to chitosan at <strong>the</strong> level of gene<br />

expression.<br />

26


Relative expression<br />

6<br />

4<br />

2<br />

PpMEKK1A<br />

PpMEKK1B<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

Figure 9. Expression of <strong>the</strong> two MEKK1 homologs <strong>in</strong> <strong>Physcomitrella</strong> upon chitosan treatment <strong>in</strong> wild<br />

type <strong>moss</strong> relative to untreated (0h). The figure is based on a s<strong>in</strong>gle technical replicate.<br />

As seen <strong>in</strong> Figure 9, PpMEKK1A mRNA levels respond more strongly than does<br />

PpMEKK1B mRNA upon chitosan treatment. However, it cannot be excluded that<br />

PpMEKK1B is also differentially expressed upon chitosan perception. Thus, both<br />

homologs were chosen as candidates for creat<strong>in</strong>g KOs.<br />

MKK1 and MKK2<br />

At <strong>the</strong> time when we wanted to identify a homolog of AtMKK1 and AtMKK2 <strong>in</strong><br />

<strong>Physcomitrella</strong> <strong>the</strong>re did not exist any publications on how many MKKs <strong>Physcomitrella</strong><br />

may have. Thus, we first identified all sequences conta<strong>in</strong><strong>in</strong>g <strong>the</strong> conserved signature<br />

catalytic motif of MKK type k<strong>in</strong>ases (Figure 10).<br />

Figure 10. WebLogo of <strong>the</strong> conserved consensus motifs <strong>in</strong> <strong>the</strong> catalytic doma<strong>in</strong> of MKK type k<strong>in</strong>ases<br />

of plants (Suarez-Rodriguez et al., 2010; I created this figure for that publication as mentioned <strong>in</strong> its<br />

acknowledgments).<br />

Perform<strong>in</strong>g BLAST queries aga<strong>in</strong>st <strong>Physcomitrella</strong> prote<strong>in</strong> databases at<br />

www.cos<strong>moss</strong>.org with both full length and <strong>the</strong> conserved doma<strong>in</strong> of <strong>the</strong> 10 Arabidopsis<br />

MKKs revealed 7 MKKs <strong>in</strong> <strong>Physcomitrella</strong>. The phylogenetic relationship of <strong>the</strong>se and<br />

<strong>the</strong> 10 Arabidopsis homologs are shown <strong>in</strong> Figure 11.<br />

27


Figure 11. Phylogenetic relationship of 7 MKK homologs <strong>in</strong> <strong>Physcomitrella</strong> and <strong>the</strong> 10 Arabidopsis<br />

MKKs. Human MEK1 is <strong>in</strong>cluded as an outgroup. Pp=<strong>Physcomitrella</strong> <strong>patens</strong>, At=Arabidopsis<br />

thaliana, Hs=Homo sapiens.<br />

Recently, Dóczi et al. (2012) published a study on <strong>the</strong> phylogeny of plant MKKs and<br />

MPKs. They found <strong>the</strong> same 7 MKKs <strong>in</strong> <strong>Physcomitrella</strong>. A m<strong>in</strong>or difference was that<br />

<strong>the</strong>y used slightly different annotated models of three of <strong>the</strong> homologs: Pp1s106_83V6.1,<br />

Pp1s16_334V6.1 and Pp1s114_89V6.1. After <strong>the</strong> sequenc<strong>in</strong>g of <strong>the</strong> <strong>Physcomitrella</strong><br />

genome, <strong>the</strong> putative genes have been automatically annotated by several different<br />

annotation algorithms. Thus, at each putative gene locus several different gene models<br />

are annotated and <strong>the</strong>y might differ <strong>in</strong> <strong>the</strong>ir cod<strong>in</strong>g regions (CDS). For example, one gene<br />

model has <strong>in</strong>cluded an exon which ano<strong>the</strong>r model excluded. Until <strong>the</strong> gene models are<br />

experimentally confirmed, <strong>the</strong>y are all hypo<strong>the</strong>tical. In our study, most homologs were<br />

chosen from <strong>the</strong> latest annotation version (V1.6) at www.cos<strong>moss</strong>.org.<br />

As seen <strong>in</strong> Figure 11, <strong>the</strong>re are three PpMKKs with similar levels of identity to<br />

AtMKK1 and AtMKK2, here termed PpMKK1A-C. It can also be seen that <strong>the</strong>y all share<br />

higher similarity to AtMKK6 than <strong>the</strong>y do to AtMKK1 and AtMKK2. Arabidopsis<br />

AtMKK6 has been shown to function <strong>in</strong> a different signal<strong>in</strong>g pathway than AtMKK1 and<br />

AtMKK2, s<strong>in</strong>ce AtMKK6 activates AtMPK13 (Melikant et al., 2004). Aga<strong>in</strong>, we tested<br />

how <strong>the</strong> regulation of <strong>the</strong> putative orthologs responded to chitosan treatment (Figure 12).<br />

28


Relative expression<br />

3<br />

2<br />

1<br />

PpMKK1A<br />

PpMKK1B<br />

PpMKK1C<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

Figure 12. mRNA levels of <strong>the</strong> three MKK1 and MKK2 homologs <strong>in</strong> <strong>Physcomitrella</strong> upon chitosan<br />

treatment <strong>in</strong> wild typerelative to untreated (0h). Only three time po<strong>in</strong>ts were tested for PpMKK1A.<br />

The figure is based on a s<strong>in</strong>gle technical replicate.<br />

As seen <strong>in</strong> Figure 12, PpMKK1B and PpMKK1C responded slightly with a 2.5 fold<br />

<strong>in</strong>crease <strong>in</strong> mRNA two hours after chitosan treatment. This response was not considered<br />

robust enough to omit any of <strong>the</strong> three as putative functional orthologs. Thus, all three<br />

were chosen as candidates for creat<strong>in</strong>g KOs.<br />

MPK4<br />

At <strong>the</strong> time when we wanted to identify a homolog of AtMPK4 <strong>in</strong> <strong>Physcomitrella</strong> <strong>the</strong>re<br />

did not exist any publications on how many MPKs <strong>the</strong>re are <strong>in</strong> <strong>Physcomitrella</strong>. Thus, we<br />

first identified all sequences conta<strong>in</strong><strong>in</strong>g <strong>the</strong> conserved signature catalytic motif of MPK<br />

type k<strong>in</strong>ases (Figure 13).<br />

Figure 13. WebLogo of <strong>the</strong> conserved consensus motifs <strong>in</strong> <strong>the</strong> catalytic doma<strong>in</strong> of MPK type k<strong>in</strong>ases<br />

of plants. (Suarez-Rodriguez et al., 2010; I created this figure for that publication as mentioned <strong>in</strong> its<br />

acknowledgments).<br />

Perform<strong>in</strong>g BLAST queries aga<strong>in</strong>st <strong>Physcomitrella</strong> prote<strong>in</strong> databases at<br />

www.cos<strong>moss</strong>.org with both full length and <strong>the</strong> conserved doma<strong>in</strong> of <strong>the</strong> 20 Arabidopsis<br />

MPKs revealed 8 MPKs <strong>in</strong> <strong>Physcomitrella</strong>. The phylogenetic relationship of <strong>the</strong>se and<br />

<strong>the</strong> 20 of Arabidopsis are shown <strong>in</strong> Figure 14.<br />

29


Figure 14. Phylogenetic relationship of 8 MPK homologs <strong>in</strong> <strong>Physcomitrella</strong> and <strong>the</strong> 20 Arabidopsis<br />

MPKs. Human ERK1 is <strong>in</strong>cluded as an outgroup. Pp=<strong>Physcomitrella</strong> <strong>patens</strong>, At=Arabidopsis thaliana,<br />

Hs=Homo sapiens.<br />

Dóczi et al. (2012) later found <strong>the</strong> same 8 MPKs but used different annotated versions of<br />

Pp1s80_71V6.1 and Pp87_157V6.1. As seen from <strong>the</strong> phylogenetic tree, two MPK<br />

homologs share similar sequence identity to AtMPK4 (bolded <strong>in</strong> Figure 14), and both<br />

were chosen for fur<strong>the</strong>r studies that are described <strong>in</strong> detail <strong>in</strong> manuscript 1 and <strong>in</strong> <strong>the</strong><br />

ΔPpMPK4A section.<br />

R-genes<br />

At <strong>the</strong> time this study was undertaken <strong>the</strong>re only existed one report on <strong>the</strong> presence of R-<br />

genes <strong>in</strong> <strong>Physcomitrella</strong> (Akita and Valkonen, 2002). That study was done before <strong>the</strong><br />

genome was sequenced and was <strong>the</strong>refore probably not exhaustive. Thus, we wanted to<br />

establish how many R-genes of <strong>the</strong> CC-NB-LRR type and how many of <strong>the</strong> TIR-NB-<br />

LRR type exist <strong>in</strong> <strong>Physcomitrella</strong>. This was done by runn<strong>in</strong>g BLASTp queries with <strong>the</strong><br />

am<strong>in</strong>o acid sequence of <strong>the</strong> ~83 TIR-NB-LRR genes and 51 CC-NB-LRR genes <strong>in</strong><br />

Arabidopsis (Meyers et al., 2003) aga<strong>in</strong>st <strong>the</strong> <strong>Physcomitrella</strong> prote<strong>in</strong> sequence databases<br />

at www.cos<strong>moss</strong>.org. Many of <strong>the</strong> identified sequences conta<strong>in</strong>ed only parts of a fully<br />

<strong>in</strong>tact R-gene, for example prote<strong>in</strong>s with CC-NB and no LRR doma<strong>in</strong> or just <strong>the</strong> NB-<br />

LRR doma<strong>in</strong>s. For this study we were <strong>in</strong>terested <strong>in</strong> establish<strong>in</strong>g whe<strong>the</strong>r <strong>the</strong> R-gene<br />

homologs functioned <strong>in</strong> <strong>in</strong>nate immunity <strong>in</strong> <strong>Physcomitrella</strong> as <strong>the</strong>y do <strong>in</strong> flower<strong>in</strong>g plants.<br />

We <strong>the</strong>refore performed a conservative search and were mostly <strong>in</strong>terested <strong>in</strong> R-genes<br />

30


with all three doma<strong>in</strong>s <strong>in</strong>tact. The <strong>Physcomitrella</strong> sequences from <strong>the</strong> first BLAST search<br />

were thus used as queries <strong>in</strong> additional BLAST searches aga<strong>in</strong>st <strong>the</strong> <strong>Physcomitrella</strong><br />

genome to identify paralogs with high identity that were missed <strong>in</strong> <strong>the</strong> first search. The<br />

presence of a TIR doma<strong>in</strong> was verified us<strong>in</strong>g Pfam http://pfam.sanger.ac.uk/ (F<strong>in</strong>n et al.,<br />

2007), while presumptive coiled-coil doma<strong>in</strong>s were verified us<strong>in</strong>g Coils<br />

http://embnet.vital-it.ch/software/COILS_form.html (Lupas et al, 1991). This search<br />

identified six CC-NB-LRR R-genes, n<strong>in</strong>e unusual R-genes conta<strong>in</strong><strong>in</strong>g a Prote<strong>in</strong> K<strong>in</strong>ase<br />

doma<strong>in</strong> (Pfam: PF00069) as well as a CC doma<strong>in</strong> <strong>in</strong> <strong>the</strong> N-term<strong>in</strong>al region (CC-PK-NB-<br />

LRR) and three TIR-NB-LRRs. Figure 15 depicts <strong>the</strong> phylogenetic relationship of <strong>the</strong>se<br />

18 R genes with Arabidopsis ADR1-L2, a CC-NB-LRR, and with Arabidopsis RPS4, a<br />

TIR-NB-LRR. Also <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> phylogenetic tree is an unusual <strong>Physcomitrella</strong> R-like<br />

prote<strong>in</strong> conta<strong>in</strong><strong>in</strong>g both TIR and CC <strong>in</strong> <strong>the</strong> N-term<strong>in</strong>al region but no LRR<br />

(Pp1s32_318V6.1). The human Toll Like Receptor 8 (TLR8) is <strong>in</strong>cluded as an outgroup.<br />

Figure 15. Phylogenetic relationship of 18 R-prote<strong>in</strong> homologs and one R-prote<strong>in</strong> like prote<strong>in</strong><br />

(Pp1s32_318V6.1 R2) <strong>in</strong> <strong>Physcomitrella</strong> and two Arabidopsis R-prote<strong>in</strong>s. Human TLR8 is <strong>in</strong>cluded as<br />

outgroup. Pp=<strong>Physcomitrella</strong> <strong>patens</strong>, At=Arabidopsis thaliana, Hs=Homo sapiens.<br />

.<br />

Four different k<strong>in</strong>ds of R-genes were chosen for fur<strong>the</strong>r functional studies by creat<strong>in</strong>g<br />

KOs. These were termed R1-4 (bolded <strong>in</strong> Figure 15): R1 is CC-NB-LRR, R2 is a R-like<br />

31


prote<strong>in</strong> consist<strong>in</strong>g of CC-TIR-NB, R3 is a TIR-NB-LRR and R4 is a member of <strong>the</strong> novel<br />

prote<strong>in</strong> k<strong>in</strong>ase conta<strong>in</strong><strong>in</strong>g group CC-PK-NB-LRR.<br />

Recently Xue et al. (2012) described <strong>the</strong> phylogeny of R-genes <strong>in</strong> <strong>Physcomitrella</strong>.<br />

Besides 47 shortened NBS-encod<strong>in</strong>g genes, <strong>the</strong>y found, like us, a total of 18 <strong>in</strong>tact R-<br />

genes with N-term<strong>in</strong>al, NB, and LRR doma<strong>in</strong>s all present. Like us, three of <strong>the</strong>se are of<br />

<strong>the</strong> TIR-NB-LRR k<strong>in</strong>d. But unlike us <strong>the</strong>y found six of <strong>the</strong> novel prote<strong>in</strong> k<strong>in</strong>ase doma<strong>in</strong><br />

N-term<strong>in</strong>al conta<strong>in</strong><strong>in</strong>g k<strong>in</strong>d, and n<strong>in</strong>e of <strong>the</strong> ‘classic’ CC-NB-LRR k<strong>in</strong>d while we found 9<br />

of <strong>the</strong> novel class and just six belong<strong>in</strong>g to <strong>the</strong> CC-NB-LRR class. In addition, <strong>the</strong>y do<br />

not mention if <strong>the</strong>y, like us, found a CC doma<strong>in</strong> <strong>in</strong> addition to <strong>the</strong> prote<strong>in</strong> k<strong>in</strong>ase doma<strong>in</strong><br />

<strong>in</strong> <strong>the</strong> novel class. Unfortunately, <strong>the</strong>y did not <strong>in</strong>clude full ID numbers on <strong>the</strong> genes <strong>the</strong>y<br />

f<strong>in</strong>d. It is <strong>the</strong>refore impossible for us to <strong>in</strong>vestigate this discrepancy, and also impossible<br />

to say whe<strong>the</strong>r <strong>the</strong> four R-genes we chose to create KOs of are among <strong>the</strong> R-genes<br />

<strong>in</strong>cluded <strong>in</strong> <strong>the</strong>ir phylogenetic tree.<br />

Creat<strong>in</strong>g targeted KOs<br />

From <strong>the</strong> bio<strong>in</strong>formatics searches with <strong>the</strong>13 Arabidopsis genes of <strong>in</strong>terest <strong>in</strong> Table 1, 17<br />

homologs <strong>in</strong> <strong>Physcomitrella</strong> were chosen for creat<strong>in</strong>g targeted KOs us<strong>in</strong>g<br />

<strong>Physcomitrella</strong>’s ability to do homologous recomb<strong>in</strong>ation at a high rate. This procedure is<br />

achieved by clon<strong>in</strong>g approximately 1 kbp of <strong>the</strong> genomic flank<strong>in</strong>g regions of a target<br />

gene <strong>in</strong>to each side of a selection marker <strong>in</strong> a transformation vector. The plasmid DNA is<br />

<strong>the</strong>n transformed <strong>in</strong>to <strong>moss</strong> protoplasts us<strong>in</strong>g a PEG and heat chock based protocol<br />

(Schaefer et al., 1991). In some protoplasts, homologues recomb<strong>in</strong>ation will occur on<br />

both of <strong>the</strong> flank<strong>in</strong>g regions result<strong>in</strong>g <strong>in</strong> a swap of <strong>the</strong> target gene with <strong>the</strong> selection<br />

marker (Figure 16).<br />

Figure 16. Overview of a wild type (WT) locus of a target gene and <strong>the</strong> ideal KO locus after<br />

homologous recomb<strong>in</strong>ant <strong>in</strong>tegration of a s<strong>in</strong>gle selection marker (nptII). The pr<strong>in</strong>ciple primers used<br />

for genotyp<strong>in</strong>g are depicted P1-P8. LB – Left Boarder and RB – Right Boarder.<br />

The clon<strong>in</strong>g of <strong>the</strong> first two target genes (PpMPK4A and PpMPK4B) was done by regular<br />

restriction enzyme digestion and ligation one side at <strong>the</strong> time. The clon<strong>in</strong>g of PpMPK4B<br />

is described <strong>in</strong> Materials & Methods, and <strong>the</strong> clon<strong>in</strong>g of PpMPK4A is described <strong>in</strong><br />

manuscript 1. This was very tedious work and <strong>the</strong>refore a transformation plasmid<br />

conta<strong>in</strong><strong>in</strong>g two USER clon<strong>in</strong>g cassettes was created (Nour-Eld<strong>in</strong> et al., 2006; Jacobsen et<br />

al., 2011). This enabled four fragment clon<strong>in</strong>g, <strong>the</strong>reby clon<strong>in</strong>g of both LB and RB onto<br />

<strong>the</strong> selection marker and <strong>the</strong> backbone <strong>in</strong> one step. The rema<strong>in</strong><strong>in</strong>g KO vectors were<br />

cloned this way. The result<strong>in</strong>g 16 plasmids are listed <strong>in</strong> Table 8 with <strong>the</strong>ir name and ID#<br />

of <strong>the</strong> correspond<strong>in</strong>g target gene. One of <strong>the</strong> 16 plasmids is a knock <strong>in</strong> (KI) vector with a<br />

GFP tagged version of PpMPK4A. This KI construct was cloned <strong>in</strong>to <strong>the</strong> pMBLU-GFP<br />

32


vector which was made by PhD student Magnus Rasmussen <strong>in</strong> our group. KO contructs<br />

for PpMKK1B and PpMKK1C rema<strong>in</strong> to be cloned.<br />

Upon transformation, protoplasts are subjected to two rounds of selection. Colonies<br />

surviv<strong>in</strong>g <strong>the</strong>se selection rounds are considered stable transformants and must have a<br />

selection gene <strong>in</strong>tegrated <strong>in</strong> <strong>the</strong> genome. Although <strong>the</strong> homologous recomb<strong>in</strong>ation rate <strong>in</strong><br />

<strong>Physcomitrella</strong> is <strong>the</strong> highest reported for any multicellular organism (Schaefer and Zrÿd,<br />

1997), <strong>in</strong> practice targeted KO with <strong>the</strong> <strong>in</strong>tegration of a s<strong>in</strong>gle selection cassette at <strong>the</strong><br />

locus of <strong>the</strong> target gene only occurs at a low frequency. Random <strong>in</strong>tegration or<br />

concatemeric <strong>in</strong>sertions occur at higher frequencies than s<strong>in</strong>gle gene replacements. Thus,<br />

<strong>the</strong> surviv<strong>in</strong>g colonies are first subjected to a screen for WT KO <strong>in</strong> which a wild type<br />

PCR band fails to be amplified while a control PCR works. As an example of such a<br />

screen, part of <strong>the</strong> screen for ΔPpMEKK1A KOs is shown <strong>in</strong> Figure 17.<br />

Figure 17. Screen<strong>in</strong>g of potential ΔPpMEKK1A KOs with <strong>the</strong> WT primer set (P3-P4, Figure 16) and a<br />

control primer set (P3-P4 for PpMEKK1B). Specific primer sequences can be found <strong>in</strong> Table 9. The<br />

asterisk marks an example of a false KO s<strong>in</strong>ce <strong>the</strong> control PCR also failed to amplify a product.<br />

Next, <strong>the</strong> KOs found <strong>in</strong> <strong>the</strong>se screens were propagated on new plates. New DNA was<br />

<strong>the</strong>n extracted and subjected to fur<strong>the</strong>r genotyp<strong>in</strong>g to verify <strong>the</strong> WT KO and to check for<br />

correct <strong>in</strong>tegration at LB and RB. As an example of such <strong>the</strong> genotyp<strong>in</strong>g, PCRs of two<br />

<strong>in</strong>dependent KO l<strong>in</strong>es of <strong>the</strong> R-gene R3 are shown <strong>in</strong> Figure 18.<br />

Figure 18. Genotyp<strong>in</strong>g of two KO l<strong>in</strong>es of <strong>the</strong> R-gene R3 (ΔR3-1 = KO1 and ΔR3-2 = KO2). The<br />

numbers above each lane refer to <strong>the</strong> primers depicted <strong>in</strong> Figure 16.<br />

The production and selection of <strong>the</strong> 17 KOs were <strong>in</strong> practice done <strong>in</strong> two rounds.<br />

Included <strong>in</strong> <strong>the</strong> first round of KOs were ATG5, MPK4A, MPK4B, SGT1, RAR1, R1, R2,<br />

R3, R4. The second, ongo<strong>in</strong>g round of KO transformations <strong>in</strong>cluded CERK1A, CERK1B,<br />

MEKK1A, MEKK1B, MKK1A, MPK4B#2, and ACD11 as well as a repeat of MPK4A<br />

to produce more KO l<strong>in</strong>es. Also a targeted <strong>in</strong>sert, knock <strong>in</strong> (KI), of a GFP tag on<br />

PpMPK4A is now be<strong>in</strong>g transformed and selected.<br />

33


From <strong>the</strong> first transformation round, no KOs were found of <strong>the</strong> PpSGT1 and<br />

PpMPK4B despite at least six <strong>in</strong>dependent transformation attempts done <strong>in</strong> parallel when<br />

creat<strong>in</strong>g o<strong>the</strong>r successful KO l<strong>in</strong>es. This could <strong>in</strong>dicate that <strong>the</strong> two genes are essential. In<br />

Arabidopsis <strong>the</strong>re are two copies of SGT1, and <strong>the</strong>ir double KO is embryo lethal<br />

(Azevedo et al., 2006). S<strong>in</strong>ce <strong>Physcomitrella</strong> only has one SGT1 homolog and is haploid<br />

it is likely that this gene can not be permanently knocked out. In Arabidopsis, some MPK<br />

loss-of-function mutants also exhibit seedl<strong>in</strong>g lethality, <strong>in</strong>appropriate defense activation<br />

and dwarfism (Petersen et al., 2000; Ren et al., 2008; Zhang et al., 2012). S<strong>in</strong>ce<br />

<strong>Physcomitrella</strong> has only eight MPKs compared to <strong>the</strong> 20 <strong>in</strong> Arabidopsis, it is possible<br />

that miss<strong>in</strong>g one MPK <strong>in</strong> this more basal set of MPKs could be lethal.<br />

From <strong>the</strong> o<strong>the</strong>r target genes from <strong>the</strong> first round of transformation, two KOs of<br />

each l<strong>in</strong>e that showed correct LB and RB <strong>in</strong>tegrations were kept for fur<strong>the</strong>r experiments.<br />

However, it has previously been shown that between 40 and 85% of targeted gene<br />

replacements loci conta<strong>in</strong> multiple copies of <strong>the</strong> target<strong>in</strong>g cassette and on average 11<br />

copies are <strong>in</strong>tegrated (Kamisugi et al., 2006). Therefore, to exam<strong>in</strong>e if only one copy of<br />

<strong>the</strong> selection cassette was <strong>in</strong>serted at <strong>the</strong> target locus a PCR flank<strong>in</strong>g <strong>the</strong> whole targeted<br />

gene was conducted. The results from <strong>the</strong>se PCRs are shown <strong>in</strong> Figure 19.<br />

Figure 19. Whole locus PCR amplification on WT and two KOs from each l<strong>in</strong>e us<strong>in</strong>g <strong>the</strong> P1-P6<br />

pr<strong>in</strong>ciple primers depicted <strong>in</strong> Figure 16. Primer sequences are <strong>in</strong> Table 9.<br />

The great advantage of genotyp<strong>in</strong>g with <strong>the</strong> whole locus spann<strong>in</strong>g primers is that <strong>the</strong><br />

same primer set will amplify a product <strong>in</strong> both WT and KO but with different lengths <strong>in</strong><br />

case of s<strong>in</strong>gle <strong>in</strong>tegration of <strong>the</strong> selection cassette at <strong>the</strong> locus. Also, <strong>the</strong> PCR product<br />

from <strong>the</strong> KOs can be sent for sequenc<strong>in</strong>g to verify correct <strong>in</strong>tegration. The disadvantage<br />

is that <strong>the</strong> PCR products are very long, 4.5 – 9.5 Kbp, and can <strong>the</strong>refore be hard to<br />

amplify. In addition, if concatemeric <strong>in</strong>tegration has occurred <strong>the</strong> product will be too long<br />

to amplify.<br />

As seen <strong>in</strong> Figure 19, it was possible to amplify a whole gene product for KO1 of<br />

each l<strong>in</strong>e except for R1 for which both KO1 and KO2 could be amplified. These PCR<br />

products were all sequenced and correct <strong>in</strong>sertions were verified. The fact that correct LB<br />

and RB was confirmed <strong>in</strong> all l<strong>in</strong>es, but whole gene amplification only was possible <strong>in</strong><br />

some, could <strong>in</strong>dicate that concatemeric <strong>in</strong>tegration had occurred at <strong>the</strong> loci. If <strong>the</strong> whole<br />

selection cassette is <strong>in</strong>serted several times <strong>in</strong> a “head to tail” manner at <strong>the</strong> same locus it<br />

should be possible to amplify a PCR product with <strong>the</strong> P7-P8 primers po<strong>in</strong>t<strong>in</strong>g “outwards”<br />

<strong>in</strong> <strong>the</strong> selection cassette while this primer set should not yield any product if only one<br />

selection cassette have been <strong>in</strong>serted (Kamisugi et al., 2006). This type of PCR was<br />

conducted on all l<strong>in</strong>es and <strong>the</strong> result is shown <strong>in</strong> Figure 20.<br />

34


Figure 20. PCR amplification to check for concatemeric selection marker <strong>in</strong>tegration on WT and two<br />

KOs from each l<strong>in</strong>e us<strong>in</strong>g <strong>the</strong> P7-P8 primer set depicted <strong>in</strong> Figure 16.<br />

The presence of PCR product with <strong>the</strong> primer set P7-P8 (Figure 19) corresponds with <strong>the</strong><br />

lack of a PCR product with <strong>the</strong> primer set P1-P6 (Figure 20) except for <strong>the</strong> ΔPpMPK4A-2<br />

l<strong>in</strong>e <strong>in</strong> which nei<strong>the</strong>r primer set produced any product. This could be expla<strong>in</strong>ed by a<br />

concatemeric <strong>in</strong>tegration with only parts of <strong>the</strong> selection marker, thus lack<strong>in</strong>g a primer<br />

b<strong>in</strong>d<strong>in</strong>g site.<br />

From <strong>the</strong> new and ongo<strong>in</strong>g round of transformations, eight KOs of MEKK1A (Figure 17),<br />

four KOs of MEKK1B, and one new MPK4A KO have been confirmed by lack of WT<br />

band <strong>in</strong> two <strong>in</strong>dependent DNA extractions. But fur<strong>the</strong>r genotyp<strong>in</strong>g is still lack<strong>in</strong>g. A<br />

complete summary of <strong>the</strong> genotypic evidence for <strong>the</strong> different transformants is <strong>in</strong>cluded<br />

<strong>in</strong> Table 3.<br />

Sou<strong>the</strong>rn blot<br />

S<strong>in</strong>ce a s<strong>in</strong>gle <strong>in</strong>tegration event at a target locus does not rule out ectopic <strong>in</strong>tegration<br />

events elsewhere <strong>in</strong> <strong>the</strong> genome, Sou<strong>the</strong>rn blots were made with <strong>the</strong> selection gene nptII<br />

as probe. This assay has so far only been done on KOs from <strong>the</strong> first round.<br />

Genotype Restriction enzymes<br />

Pst1 BamH1 EcoRV<br />

ΔATG5 4107+17759 18860 19205<br />

ΔMPK4A 436+9510 11312 5065<br />

ΔRAR1 1397+3545 22k+ 7129<br />

ΔR1 545+10k+ 20k+ 20k+<br />

ΔR2 2151+4959 22k+ 7263<br />

ΔR3 2279+2429 16842 7648<br />

ΔR4 2946+5183 11300 8251<br />

Table 2. Expected band sizes (bp) for <strong>the</strong> different KO l<strong>in</strong>es when <strong>the</strong>ir gDNA is digested with <strong>the</strong><br />

<strong>in</strong>dicated restriction enzymes and hybridized to <strong>the</strong> nptII probe.<br />

Genomic DNA from WT and <strong>the</strong> different KO l<strong>in</strong>es was digested with <strong>the</strong> three<br />

restriction enzymes listed <strong>in</strong> Table 2. The restriction enzyme Pst1 cuts once <strong>in</strong>side <strong>the</strong><br />

nptII selection cassette and gDNA cut with this enzyme should thus yield two bands<br />

when annealed to nptII if only one <strong>in</strong>tegration event has occurred. BamH1 and EcoRV do<br />

not cut with<strong>in</strong> <strong>the</strong> selection cassette and should <strong>the</strong>refore only yield one band.<br />

35


Figure 21. Sou<strong>the</strong>rn blot analysis. Detection of nptII <strong>in</strong>tegration <strong>in</strong> genomic DNA of <strong>the</strong> <strong>in</strong>dicated<br />

genotypes digested with Pst1 which cuts once with<strong>in</strong> nptII and with BamH1 which does not cut <strong>in</strong><br />

nptII. Control (C) is 15 ng of nptII cut from pMBLU with <strong>the</strong> restriction enzymes Sac1 and AsiSI<br />

(1940 bp). The expected sizes <strong>in</strong> case of a s<strong>in</strong>gle <strong>in</strong>tegration event are shown <strong>in</strong> Table 2. The numbers<br />

to <strong>the</strong> left are sizes <strong>in</strong> kb of selected bands of <strong>the</strong> ladder, λ-phage cut with Pst1. A picture of <strong>the</strong><br />

correspond<strong>in</strong>g ethidium bromide sta<strong>in</strong>ed DNA gel can be found <strong>in</strong> material and methods (Figure<br />

53A).<br />

It proved difficult to extract enough clean DNA of <strong>the</strong> ΔATG5 l<strong>in</strong>es. Perhaps due to its<br />

autophagy deficiency it had a high content of some materials that could not be separated<br />

from <strong>the</strong> DNA with <strong>the</strong> phenol/chloroform DNA extraction protocol used. Thus <strong>the</strong>se<br />

bands are a bit weak. A picture of <strong>the</strong> ethidium bromide sta<strong>in</strong>ed agarose gel with <strong>the</strong><br />

DNA before it was blotted onto <strong>the</strong> membrane can be seen <strong>in</strong> Figure 53A <strong>in</strong> Materials<br />

and Method.<br />

From this first blot <strong>the</strong> very fa<strong>in</strong>t bands of ΔATG5-1 <strong>in</strong>dicate that it could be a<br />

s<strong>in</strong>gle <strong>in</strong>tegration targeted KO, while ΔATG5-2 may have more than one copy of nptII.<br />

Both of <strong>the</strong> ΔMPK4A l<strong>in</strong>es have unexpected bands and are <strong>the</strong>refore likely to have more<br />

than one copy. ΔRAR1-1 only has <strong>the</strong> two expected bands when digested with Pst1 while<br />

<strong>the</strong>re is not really any clear band when digested with BamH1. None<strong>the</strong>less, ΔRAR1-1 is<br />

most likely a s<strong>in</strong>gle <strong>in</strong>tegration targeted KO.<br />

S<strong>in</strong>ce digestions with BamH1 did not produce very clear bands, <strong>the</strong> experiment<br />

was repeated with EcoRV and <strong>the</strong> rema<strong>in</strong><strong>in</strong>g genotypes were <strong>in</strong>cluded (Figure 22 and<br />

Figure 23)<br />

36


Figure 22. Sou<strong>the</strong>rn blot analysis. Detection of nptII <strong>in</strong>tegration <strong>in</strong> genomic DNA of <strong>the</strong> <strong>in</strong>dicated<br />

genotypes digested with Pst1 which cuts once with<strong>in</strong> nptII and with EcoRV which does not cut <strong>in</strong><br />

nptII. Control (C) is, to <strong>the</strong> left 50 ng and to <strong>the</strong> right 5 ng of pMBLU cut with Sma1 (4254 bp). The<br />

expected sizes <strong>in</strong> case of a s<strong>in</strong>gle <strong>in</strong>tegration event are shown <strong>in</strong> Table 2. Numbers to <strong>the</strong> left are sizes<br />

<strong>in</strong> kbp of selected bands of <strong>the</strong> ladder, λ-phage cut with Pst1. (A) and (B) are two different exposures<br />

of <strong>the</strong> same blot. A picture of <strong>the</strong> correspond<strong>in</strong>g ethidium bromide sta<strong>in</strong>ed DNA gel can be found <strong>in</strong><br />

material and methods (Figure 53B).<br />

In <strong>the</strong> blot shown <strong>in</strong> Figure 22 <strong>the</strong>re is a band <strong>in</strong> <strong>the</strong> WT most likely due to run over from<br />

<strong>the</strong> very dark control band s<strong>in</strong>ce no bands were detected <strong>in</strong> <strong>the</strong> two WT digestions<br />

<strong>in</strong>cluded <strong>in</strong> <strong>the</strong> first blot (Figure 21). The ΔATG5-1 l<strong>in</strong>e is still fa<strong>in</strong>t, but <strong>the</strong> expected<br />

bands can just be detected whereas no o<strong>the</strong>r bands are detected. ΔATG5-1 is thus<br />

concluded to be a s<strong>in</strong>gle <strong>in</strong>tegration targeted KO. ΔATG5-2 has more than one <strong>in</strong>sertion.<br />

Aga<strong>in</strong>, both of <strong>the</strong> ΔMPK4A l<strong>in</strong>es have unexpected bands and are <strong>the</strong>refore likely to have<br />

more than one copy. In Figure 22B ΔRAR1-1, is clearly a s<strong>in</strong>gle <strong>in</strong>tegration targeted KO<br />

while ΔRAR1-2 has multiple <strong>in</strong>sertions.<br />

Figure 23. Sou<strong>the</strong>rn blot analysis. Detection of nptII <strong>in</strong>tegration <strong>in</strong> genomic DNA of <strong>the</strong> <strong>in</strong>dicated<br />

genotypes digested with Pst1 which cuts once with<strong>in</strong> nptII and with EcoRV which does not cut <strong>in</strong><br />

37


nptII. Control (C) is to <strong>the</strong> left 50 ng and to <strong>the</strong> right 5 ng of pMBLU cut with Sma1 (4254 bp). The<br />

expected sizes <strong>in</strong> case of a s<strong>in</strong>gle <strong>in</strong>tegration event are shown <strong>in</strong> Table 2. Numbers to <strong>the</strong> left are sizes<br />

<strong>in</strong> kbp of selected bands of <strong>the</strong> ladder, λ-phage cut with Pst1. (A) and (B) are two different exposures<br />

of <strong>the</strong> same blot. A picture of <strong>the</strong> correspond<strong>in</strong>g ethidium bromide sta<strong>in</strong>ed DNA gel can be found <strong>in</strong><br />

material and methods (Figure 53C)<br />

The DNA gel used to create <strong>the</strong> blots <strong>in</strong> Figure 22 and Figure 23 ran a little bit too far ,<br />

and thus DNA fragments below some 600 bp ran out. Also, <strong>in</strong> Figure 23 <strong>the</strong> DNA gel did<br />

not seem to have transferred equally onto to <strong>the</strong> membrane s<strong>in</strong>ce <strong>the</strong> bands on <strong>the</strong> left<br />

side are weaker than expected, <strong>in</strong>clud<strong>in</strong>g <strong>the</strong> control DNA which to <strong>the</strong> right is loaded<br />

with 10 times <strong>the</strong> amount of control DNA compared to <strong>the</strong> left side but which shows as<br />

equal <strong>in</strong>tensity on <strong>the</strong> film. Even so, <strong>in</strong> Figure 23B very fa<strong>in</strong>t bands at <strong>the</strong> expected sizes<br />

for ΔR1-1 and ΔR1-2 can be detected and thus <strong>the</strong>se are considered s<strong>in</strong>gle <strong>in</strong>tegration<br />

targeted KOs. For <strong>the</strong> rema<strong>in</strong><strong>in</strong>g R-genes, ΔR2-1, ΔR3-1 and ΔR4-1 are s<strong>in</strong>gle <strong>in</strong>tegration<br />

targeted KOs, while ΔR2-2, ΔR3-2 and ΔR4-2 conta<strong>in</strong> more than one selection cassette<br />

<strong>in</strong>tegrated <strong>in</strong> <strong>the</strong> genome. A summary of <strong>the</strong> Sou<strong>the</strong>rn blot evidence is <strong>in</strong>cluded <strong>in</strong> Table<br />

3.<br />

When a sou<strong>the</strong>rn blot with <strong>the</strong> selection cassette as probe shows a s<strong>in</strong>gle<br />

<strong>in</strong>tegration event, it still does not rule out <strong>the</strong> possibility of random <strong>in</strong>tegration of <strong>the</strong><br />

vector backbone or <strong>the</strong> flank<strong>in</strong>g regions. Thus, <strong>the</strong> best evidence that <strong>the</strong> observed<br />

phenotype of a transformed <strong>moss</strong> l<strong>in</strong>e is due to <strong>in</strong>tegration <strong>in</strong> <strong>the</strong> target locus and not<br />

ectopic <strong>in</strong>sertion is that more than one <strong>in</strong>dependent transformant shows <strong>the</strong> same<br />

phenotype. This is because it is very unlikely that a random <strong>in</strong>tegration event will occur<br />

at <strong>the</strong> same locus and thus cause <strong>the</strong> same molecular phenotype. Thus, at least two<br />

<strong>in</strong>dividual transformant were kept for each target locus no matter <strong>the</strong> Sou<strong>the</strong>rn blot result.<br />

Flow cytometry<br />

PEG transformation is known to cause protoplast fusion which can result <strong>in</strong> polyploidy.<br />

In a large mutant screen of PEG transformed protoplasts it was found that up to 17% of<br />

<strong>the</strong> transformants were polyploid (Schween et al., 2005). Therefore, a screen of ploidy<br />

was conducted. Protoplasts of <strong>the</strong> different genotypes were sta<strong>in</strong>ed with <strong>the</strong> DNA sta<strong>in</strong><br />

propidium iod<strong>in</strong>e and <strong>the</strong> DNA content of <strong>the</strong> protoplasts was compared to WT<br />

protoplasts <strong>in</strong> a FACS analysis us<strong>in</strong>g a flow cytometer. No difference was found between<br />

any of <strong>the</strong> genotypes exam<strong>in</strong>ed, and thus <strong>the</strong>y were all concluded to be haploid as <strong>the</strong> WT.<br />

Only <strong>the</strong> KOs from first round was subjected to this analysis (see Table 3). KOs from <strong>the</strong><br />

second round of transformation still needs to be assessed for ploidy level.<br />

PCR genotyp<strong>in</strong>g<br />

Sou<strong>the</strong>rn<br />

blot FACS<br />

WT LB RB Whole Concatemer SIKO Haploid<br />

Name P3-P4 P1-P2 P5-P6 P1-P6 P7-P8<br />

R1-1 V V V V - V V<br />

R1-2 V V V V - V V<br />

R2-1 V V V V - V V<br />

R2-2 V V V - V M V<br />

R3-1 V V V V - V V<br />

R3-2 V V V - V M V<br />

38


R4-1 V V V V - V V<br />

R4-2 V V V - V M V<br />

RAR1-1 V V V V - V V<br />

RAR1-2 V V V - V M V<br />

ATG5-1 V V V V - V V<br />

ATG5-2 V V V - V M V<br />

MPK4A-1 V V V V - M V<br />

MPK4A-2 V V V - - M V<br />

MPK4A-4 V V<br />

MEKK1A-1<br />

MEKK1A-2<br />

MEKK1A-3<br />

MEKK1A-4<br />

MEKK1B-1<br />

MEKK1B-2<br />

MEKK1B-3<br />

MEKK1B-4<br />

MEKK1B-5<br />

MEKK1B-6<br />

MEKK1B-7<br />

MEKK1B-8<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

V<br />

Table 3. Overview of KO transformants and <strong>the</strong>ir genotypic evidence. Transformant names <strong>in</strong> bold<br />

are from <strong>the</strong> first round of transformation, while <strong>the</strong> rest is from <strong>the</strong> new and recently undertaken<br />

round of transformation. Primers <strong>in</strong> Figure 16 and listed <strong>in</strong> Table 9 are yellow. “V” <strong>in</strong> <strong>the</strong> “PCR<br />

genotyp<strong>in</strong>g” columns means PCR product denom<strong>in</strong>ated <strong>in</strong> <strong>the</strong> top of <strong>the</strong> column could be amplified.<br />

“-“ means <strong>the</strong> PCR was performed but did not produce any product. “V” <strong>in</strong> <strong>the</strong> Sou<strong>the</strong>rn blot<br />

means it showed evidence of a S<strong>in</strong>gle Integration KO event (SIKO) while “M” means evidence of<br />

multiple <strong>in</strong>tegration events. “V” below “FACS” means <strong>the</strong> l<strong>in</strong>e is haploid. Noth<strong>in</strong>g <strong>in</strong> any box means<br />

that <strong>the</strong> test has not been done yet.<br />

Pathogen <strong>in</strong>fections<br />

With <strong>the</strong> KO l<strong>in</strong>es obta<strong>in</strong>ed <strong>in</strong> <strong>the</strong> first round of transformation (Table 3) I went to<br />

Instituto de Investigaciones Biológicas Clemente Estable (IIBCE) <strong>in</strong> Montevideo,<br />

Uruguay, where, under <strong>the</strong> supervision of Dr. Inés Ponce de León, I screened for altered<br />

immune responses of <strong>the</strong> mutants upon <strong>in</strong>fection with different pathogens. The pathogens<br />

tested <strong>in</strong> Uruguay are all necrotrophic and known to <strong>in</strong>fect <strong>Physcomitrella</strong> (Table 4)<br />

(Andersson et al., 2005; Ponce de León et al., 2007; Oliver et al., 2009; Ponce De León et<br />

al., 2012).<br />

Necrotrophic pathogens<br />

Botrytis c<strong>in</strong>erea<br />

Alternaria brassicicola<br />

Pectobacterium carotovorum subsp. carotovorum SCC3193<br />

Pectobacterium carotovorum subsp. carotovorum SCC1<br />

Pythium irregulare<br />

Table 4. Necrotrophic pathogens tested on <strong>Physcomitrella</strong><br />

39


The different KO l<strong>in</strong>es were all <strong>in</strong>fected with <strong>the</strong>se pathogens and screened for altered<br />

disease symptoms. After <strong>in</strong>fection, symptoms were visualized with different sta<strong>in</strong><strong>in</strong>gs and<br />

<strong>in</strong>spected by microscopy.<br />

Symptoms of a Botrytis c<strong>in</strong>erea <strong>in</strong>fection<br />

As previously reported, B. c<strong>in</strong>erea <strong>in</strong>fection causes macroscopic disease symptoms like<br />

brown<strong>in</strong>g of <strong>the</strong> stem and protonemal tissue as well as visible maceration and cell death<br />

of <strong>the</strong> <strong>in</strong>fected tissue (Ponce de León et al., 2007; Ponce De León et al., 2012) (see also<br />

Figure 26B). Cellular changes occurr<strong>in</strong>g dur<strong>in</strong>g this <strong>in</strong>fection can be exam<strong>in</strong>ed with<br />

histochemical sta<strong>in</strong><strong>in</strong>g techniques.<br />

In Figure 24A-C <strong>the</strong> <strong>in</strong>itial steps of a B. c<strong>in</strong>erea colonization are detected with<br />

Evans blue sta<strong>in</strong><strong>in</strong>g which sta<strong>in</strong>s both dead cells and fungal cell walls. Spores with<br />

elongated germ tubes attempt to penetrate <strong>the</strong> cell wall of <strong>the</strong> <strong>moss</strong> which responds by<br />

produc<strong>in</strong>g a visible brown<strong>in</strong>g at <strong>the</strong> po<strong>in</strong>t of entrance. In Figure 24A <strong>the</strong> hyphae have<br />

penetrated one cell and cont<strong>in</strong>ued to grow <strong>in</strong>side <strong>the</strong> dead cell. The brown<strong>in</strong>g of <strong>the</strong> cell<br />

wall is suspected to be caused by <strong>the</strong> accumulation of antimicrobial phenolic compounds.<br />

This suspicion is confirmed by sta<strong>in</strong><strong>in</strong>g with toluid<strong>in</strong>e blue which sta<strong>in</strong>s phenolic<br />

compounds (Figure 24D-F). Correspond<strong>in</strong>g to <strong>the</strong> brown areas of Figure 24A-C, <strong>in</strong><br />

Figure 24D-F a clear blue sta<strong>in</strong><strong>in</strong>g is visible.<br />

Two days after B. c<strong>in</strong>erea <strong>in</strong>fection larger areas of dead cells <strong>in</strong> leaves can also be<br />

visualized with Evans blue (Figure 24G). The <strong>in</strong>fected areas also respond with cell wall<br />

fortification by callose deposition (Figure 24H, I). Hyphal proliferation <strong>in</strong>side <strong>the</strong><br />

<strong>in</strong>fected cells one day after <strong>in</strong>fection is clearly visible after sta<strong>in</strong><strong>in</strong>g with solophenyl<br />

flav<strong>in</strong>e which sta<strong>in</strong>s fungal cell walls (Hoch et al., 2005) (Figure 24J). ROS production<br />

<strong>in</strong>side <strong>in</strong>fected cells is visualized by sta<strong>in</strong><strong>in</strong>g with H 2 DCFDA which becomes fluorescent<br />

<strong>in</strong> <strong>the</strong> presence of reactive oxygen (Eruslanov and Kusmartsev, 2010) (Figure 24K, L).<br />

Note that <strong>the</strong> B. c<strong>in</strong>erea spore and hyphae are also sta<strong>in</strong>ed by H 2 DCFDA and thus also<br />

produce ROS (Figure 24L).<br />

These sta<strong>in</strong><strong>in</strong>gs are very good for a qualitative description of disease symptoms and<br />

defense mechanisms of a pathogen <strong>in</strong>fection. However, it is very difficult to use to<br />

quantitatively compare genotype responses to pathogen attack.<br />

We did all of <strong>the</strong>se histochemical sta<strong>in</strong><strong>in</strong>gs and subsequent microscopy screen<strong>in</strong>gs<br />

of <strong>the</strong> different transformants. However, we did not detect obvious differences between<br />

<strong>the</strong> transformants and <strong>the</strong> wild type and <strong>the</strong>refore <strong>the</strong>se data are omitted.<br />

40


Figure 24. Symptoms of B. c<strong>in</strong>erea <strong>in</strong>fection of three week old gametophores grown on BCD+AT. (A-<br />

C) Shows germ tube elongation from spores one day post <strong>in</strong>fection, sta<strong>in</strong>ed with Evans blue. In (A) a<br />

hyphal tip has penetrated <strong>the</strong> cell wall and hyphae are grow<strong>in</strong>g <strong>in</strong>side <strong>the</strong> cell. In (B and C) <strong>the</strong><br />

hyphal tip is about to penetrate <strong>the</strong> cell wall and brown<strong>in</strong>g is visible at <strong>the</strong> po<strong>in</strong>t of penetration. (D-F)<br />

Phenolic compounds <strong>in</strong>corporated <strong>in</strong> <strong>the</strong> cell walls <strong>in</strong> contact with hyphae and especially around <strong>the</strong><br />

po<strong>in</strong>t of hyphal tip penetration. One day post <strong>in</strong>fection sta<strong>in</strong>ed with toluid<strong>in</strong>e blue. (G) A leaf tip two<br />

days post <strong>in</strong>fection show<strong>in</strong>g dead cells sta<strong>in</strong>ed with Evans blue. (H, I) Methyl blue sta<strong>in</strong><strong>in</strong>g of leaves<br />

two days post <strong>in</strong>fection. (H) Bright field image of <strong>the</strong> same cells as (I) show<strong>in</strong>g cell wall fortification<br />

by callose deposition seen as green fluorescent (marked by an arrow). (J) Hyphal growth <strong>in</strong>side<br />

<strong>in</strong>fected cells one day after <strong>in</strong>fection, sta<strong>in</strong>ed with solophenyl flav<strong>in</strong>e. (K, L) ROS production <strong>in</strong><br />

<strong>in</strong>fected cells one day post <strong>in</strong>fection, sta<strong>in</strong>ed with H 2 DCFDA. Scale bars (A-F and H-L) 20 µm and (G)<br />

200 µm.<br />

41


Symptoms of P. irregulare <strong>in</strong>fection<br />

Infection with P. irregulare results <strong>in</strong> callose deposition visualized by sta<strong>in</strong><strong>in</strong>g with<br />

methyl blue as seen <strong>in</strong> Figure 25B compared to Figure 25A which is also sta<strong>in</strong>ed with<br />

methyl blue but not <strong>in</strong>fected. Hyphal tissue grow<strong>in</strong>g on and <strong>in</strong>side cells is clearly visible<br />

after sta<strong>in</strong><strong>in</strong>g with <strong>the</strong> fluorescent dye solophenyl flav<strong>in</strong>e (Figure 25C, E). In Figure 25D,<br />

an oospore <strong>in</strong>side an <strong>in</strong>fected cell can be seen with solophenyl flav<strong>in</strong>e sta<strong>in</strong><strong>in</strong>g.<br />

As with B. c<strong>in</strong>erea <strong>in</strong>fection, symptoms upon P. irregulare <strong>in</strong>fection were<br />

assessed for <strong>the</strong> different genotypes. However, no obvious differences between <strong>the</strong>se and<br />

<strong>the</strong> wild type were found and <strong>the</strong>se data are omitted.<br />

Figure 25. Symptoms at two days upon P. irregulare <strong>in</strong>fection of three week old wild type<br />

gametophores grown on BCD+AT. (A) Un<strong>in</strong>fected leaf sta<strong>in</strong>ed with methyl blue. (B) Callose<br />

deposition and hyphal growth <strong>in</strong>side leaf two days post <strong>in</strong>fection, methyl blue sta<strong>in</strong><strong>in</strong>g. (C) Hyphal<br />

tissue sta<strong>in</strong>ed with <strong>the</strong> fluorescent dye solophenyl flav<strong>in</strong>e. (D) Oospore <strong>in</strong>side an <strong>in</strong>fected cell sta<strong>in</strong>ed<br />

with solophenyl flav<strong>in</strong>e. (E) A s<strong>in</strong>gle <strong>in</strong>fected cell with hyphal growth sta<strong>in</strong>ed with solophenyl flav<strong>in</strong>e.<br />

Scale bars represent 20 µm.<br />

Evans blue sta<strong>in</strong><strong>in</strong>g<br />

In order to compare <strong>the</strong> susceptibility of <strong>the</strong> different genotypes to <strong>the</strong> pathogens, we<br />

used Evans blue sta<strong>in</strong><strong>in</strong>g of protonemal tissue which can be quantified and thus compared<br />

42


<strong>in</strong> an objective manner. In Figure 26B it can be seen how two days of B. c<strong>in</strong>erea <strong>in</strong>fection<br />

causes a visible brown<strong>in</strong>g of <strong>the</strong> colony and maceration of <strong>the</strong> cells compared to control<br />

treated plants (Figure 26A). The result<strong>in</strong>g cell death can be visualized by sta<strong>in</strong><strong>in</strong>g with<br />

Evans blue which only sta<strong>in</strong>s a few cells <strong>in</strong> control treated plants Figure 26C-F. This<br />

sta<strong>in</strong><strong>in</strong>g can be quantified and used as a measure of cell death by desta<strong>in</strong><strong>in</strong>g <strong>the</strong> colonies<br />

<strong>in</strong> 50% methanol, 1% SDS and measur<strong>in</strong>g <strong>the</strong> OD (600nm) of this liquid (Oliver et al.,<br />

2009; Ponce De León et al., 2012). However, as seen <strong>in</strong> Figure 26F and Figure 24A-C,<br />

Evans blue sta<strong>in</strong>s both dead cells and <strong>the</strong> fungal hyphae and thus <strong>the</strong> quantification of<br />

Evans blue sta<strong>in</strong><strong>in</strong>g is more likely a mix of cell death and fungal growth. None<strong>the</strong>less,<br />

when compar<strong>in</strong>g genotypes higher OD values still mean higher susceptibility.<br />

Figure 26. Two week old <strong>moss</strong> colonies two days after spray<strong>in</strong>g with ei<strong>the</strong>r H 2 O or 2x10 5 B. c<strong>in</strong>erea<br />

spores per ml. (A) and (B) are unsta<strong>in</strong>ed while (C-F) are sta<strong>in</strong>ed with Evans blue. Bars represent <strong>in</strong><br />

(A-D) 0.5 cm and <strong>in</strong> (E, F) 100 µm.<br />

The quantitative Evans blue sta<strong>in</strong><strong>in</strong>g assay was done upon B. c<strong>in</strong>erea <strong>in</strong>fection <strong>in</strong> a<br />

prelim<strong>in</strong>ary experiment (Figure 27). This experiment showed that this method could be<br />

used to compare <strong>the</strong> susceptibility to B. c<strong>in</strong>erea between different transformants and <strong>the</strong><br />

wild type <strong>moss</strong>.<br />

43


Evans blue sta<strong>in</strong><strong>in</strong>g two days post B. c<strong>in</strong>erea <strong>in</strong>fection<br />

OD (600nM)/mg DW<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

H2O<br />

Botrytis<br />

0<br />

WT MPK4A-1 MPK4A-2 RAR1-1 RAR1-2<br />

Figure 27. Evans blue sta<strong>in</strong><strong>in</strong>g two days post B. c<strong>in</strong>erea <strong>in</strong>fection. Error bars represent SEM of four<br />

samples, each of four colonies grown on <strong>the</strong> same plate.<br />

This experiment led to <strong>the</strong> discovery that ΔMPK4A l<strong>in</strong>es are more susceptible to B.<br />

c<strong>in</strong>erea <strong>in</strong>fection compared to wild type. This discovery is described <strong>in</strong> detail <strong>in</strong><br />

manuscript 1. It also <strong>in</strong>dicated that <strong>the</strong> ΔRAR1 l<strong>in</strong>es are less susceptible to B. c<strong>in</strong>erea<br />

<strong>in</strong>fection compared to wild type. This observation will be fur<strong>the</strong>r discussed <strong>in</strong> <strong>the</strong> ΔRAR1<br />

section. What could also be observed from this prelim<strong>in</strong>ary experiment was that <strong>the</strong>re is<br />

considerable biological variance between <strong>the</strong> different plates with <strong>in</strong>fected <strong>moss</strong>, as seen<br />

from <strong>the</strong> difference between ΔMPK4A-1 and ΔMPK4A-2. This experiment was<br />

performed with just one plate with 16 colonies for each l<strong>in</strong>e. These 16 colonies were<br />

<strong>the</strong>reafter divided <strong>in</strong>to 4x4 colonies and <strong>the</strong>n sta<strong>in</strong>ed and desta<strong>in</strong>ed. Thus, each data po<strong>in</strong>t<br />

consists of four samples but from 16 colonies that grew on <strong>the</strong> same plate. In order to<br />

take <strong>in</strong>to account <strong>the</strong> variation between different plates of <strong>the</strong> same genotype, we decided<br />

to perform <strong>the</strong> assay with four plates of 16 colonies for each l<strong>in</strong>e.<br />

Alternaria spore count<strong>in</strong>g assay<br />

For some reason <strong>the</strong> quantification of cell death us<strong>in</strong>g Evans blue sta<strong>in</strong><strong>in</strong>g did not work<br />

after A. brassicicola <strong>in</strong>fection despite several attempts. This had also been <strong>the</strong>ir<br />

experience <strong>in</strong> Uruguay (Inés Ponce de León, personal communication). In order to<br />

compare A. brassicicola <strong>in</strong>fections, we <strong>the</strong>refore decided to adapt a quantitative spore<br />

count<strong>in</strong>g assay that had previously been published for A. brassicicola <strong>in</strong>fections on<br />

Arabidopsis (Wees et al., 2003).<br />

Figure 28. Alternaria brassicicola <strong>in</strong>fection. (A) Leaf of three week old wild type three days post<br />

spray<strong>in</strong>g with 2x10 5 spores/ml. (B) Wild type colony four days post drop <strong>in</strong>oculation with 5 µl of<br />

44


2x10 5 spores/ml. (C) Spores <strong>in</strong> hemocytometer. The darker spore is one of <strong>the</strong> old spores from <strong>the</strong><br />

<strong>in</strong>fection and <strong>the</strong> lighter is a newly formed spore. Bars: (A) 500 µm, (B), 5 mm and (C) 100 µm.<br />

Three days post A. brassicicola <strong>in</strong>fection <strong>the</strong> fungus starts to produce new spores (Figure<br />

28) which, unlike <strong>the</strong> B. c<strong>in</strong>erea spores, can easily be shaken off. The assay is performed<br />

by <strong>in</strong>oculat<strong>in</strong>g with a drop (5 µl) conta<strong>in</strong><strong>in</strong>g ~2500 spores on each colony. Four days post<br />

<strong>in</strong>oculation <strong>the</strong> colonies are carefully harvested, placed <strong>in</strong> a falcon tube and shaken<br />

vigorously <strong>in</strong> a 0.01% TWEEN20 solution. The detached spores <strong>in</strong> this liquid are <strong>the</strong>n<br />

counted us<strong>in</strong>g a hemocytometer. The newly formed spores are light <strong>in</strong> color and thus<br />

easily dist<strong>in</strong>guished from <strong>the</strong> older ones use <strong>in</strong> <strong>the</strong> drop <strong>in</strong>oculation which are dark brown<br />

(Figure 29C).<br />

A prelim<strong>in</strong>ary experiment us<strong>in</strong>g just one plate of 16 colonies per genotype<br />

showed that this assay could be used to compare <strong>the</strong> susceptibility to A. brassicicola of<br />

different genotypes (Figure 29).<br />

400<br />

Alternaria spore count 4 days post <strong>in</strong>fection<br />

1000 spores/colony<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

WT ATG5‐1 ATG5‐2 MPK4A‐1 MPK4A2 RAR1‐1 RAR1‐2<br />

Figure 29. Alternaria spore count 4 days post drop <strong>in</strong>oculation with ~2500 spores/colony. This<br />

experiment served as a screen with only one biological replicate per genotype. For this reason no<br />

statistics were applied. The error bars represent SEM of <strong>the</strong> six counts made per l<strong>in</strong>e <strong>in</strong> this<br />

experiment.<br />

From <strong>the</strong> prelim<strong>in</strong>ary spore count<strong>in</strong>g experiment it seemed that <strong>the</strong> ΔATG5 l<strong>in</strong>es<br />

supported more fungal growth than <strong>the</strong> wild type. This result will be fur<strong>the</strong>r discussed <strong>in</strong><br />

<strong>the</strong> ΔATG5 section. The ΔMPK4A l<strong>in</strong>es also had higher spore counts and thus are more<br />

susceptible to A. brassicicola <strong>in</strong>fection compared to <strong>the</strong> wild type.<br />

The MAP k<strong>in</strong>ase KO ΔMPK4A<br />

The <strong>in</strong>creased susceptibility of this KO mutant to <strong>the</strong> necrotrophic fungi B. c<strong>in</strong>erea and A.<br />

brassicicola seen <strong>in</strong> <strong>the</strong> prelim<strong>in</strong>ary screens (Figure 27 and Figure 29) was repeated and<br />

is described thoroughly <strong>in</strong> manuscript 1. This section describes some experiments made<br />

with <strong>the</strong> ΔMPK4A l<strong>in</strong>es and that are not <strong>in</strong>cluded <strong>in</strong> manuscript 1.<br />

P. irregulare <strong>in</strong>fection of ΔMPK4A<br />

The susceptibility of <strong>the</strong> two <strong>in</strong>dependent ΔMPK4A l<strong>in</strong>es upon P. irregulare <strong>in</strong>fection<br />

was assessed with <strong>the</strong> quantitative Evans blue sta<strong>in</strong><strong>in</strong>g assay (Figure 30). It was not<br />

45


possible to conclude anyth<strong>in</strong>g due to <strong>the</strong> high variation. This is probably due to <strong>the</strong> way<br />

<strong>the</strong> <strong>in</strong>fection is done: a small peace of PDA plate with <strong>the</strong> oomycete mycelium is<br />

harvested by press<strong>in</strong>g a filter tip <strong>in</strong>to <strong>the</strong> agar and it is <strong>the</strong>n placed on top of each colony.<br />

The <strong>in</strong>fection is done this way to <strong>in</strong>fect <strong>the</strong> colonies with an equal amount of mycelium.<br />

But it does visually gives rise to variation <strong>in</strong> <strong>the</strong> rate at which <strong>the</strong> <strong>in</strong>dividual colonies<br />

become <strong>in</strong>fected and, even with 4 times 16 colonies for each data po<strong>in</strong>t, this probably<br />

reflects <strong>the</strong> high variation <strong>in</strong> <strong>the</strong> f<strong>in</strong>al graph. This experiment was repeated with similar<br />

results.<br />

Evans blue sta<strong>in</strong><strong>in</strong>g 24 hours post P. irregulare <strong>in</strong>fection<br />

OD (600nm)/mg DW<br />

0.16<br />

0.12<br />

0.08<br />

0.04<br />

Pythium<br />

H2O<br />

0<br />

WT MPK4A-1 MPK4A-2<br />

Figure 30. Evans blue sta<strong>in</strong><strong>in</strong>g 24 hours post P. irregulare <strong>in</strong>fection. The error bars represent SEM of<br />

four <strong>in</strong>dividual biological replicates consist<strong>in</strong>g of 16 <strong>moss</strong> colonies each.<br />

Genes not differentially expressed <strong>in</strong> <strong>the</strong> ΔMPK4A-1 l<strong>in</strong>e<br />

When screen<strong>in</strong>g for genes that are differentially expressed upon chitosan treatment (Fig.<br />

4, manuscript 1) we also found genes that were not differentially expressed <strong>in</strong> <strong>the</strong><br />

ΔMPK4A-1 l<strong>in</strong>e compared to <strong>the</strong> wild type. Some of <strong>the</strong>se are shown <strong>in</strong> Figure 31.<br />

46


CYP71A13<br />

WRKY70<br />

Relative expression<br />

3<br />

2<br />

1<br />

0<br />

WT<br />

MPK4A-1 KO<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

PRX34<br />

WRKY40<br />

Relative expression<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

ERFb3<br />

WRKY53<br />

Relative expression<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

30<br />

20<br />

10<br />

0<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

80<br />

60<br />

40<br />

20<br />

0<br />

ERF5<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

40<br />

30<br />

20<br />

10<br />

0<br />

PAL4-2<br />

0h 15m 30m 1h 2h 4h 8h<br />

Figure 31. Quantitative reverse transcriptase PCR (qPCR) analysis of transcript levels <strong>in</strong> WT (blue<br />

diamonds) and ΔPpMPK4A-1 (red squares) relative to untreated WT (time 0h) follow<strong>in</strong>g treatment<br />

with 100 μg/ml chitosan. In cases where <strong>the</strong>re are no error bars, <strong>the</strong> experiment was based on a s<strong>in</strong>gle<br />

technical replicate, o<strong>the</strong>rwise error bars represent SEM of three <strong>in</strong>dependent technical replicates.<br />

Gene identifiers and primers used are <strong>in</strong> Table 10.<br />

S<strong>in</strong>ce AtWRKY25 and AtWRKY33 have been shown to be regulated downstream of<br />

AtMPK4 (Andreasson et al., 2005) we tried to identify WRKY genes that were<br />

47


differentially regulated after chitosan treatment <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 l<strong>in</strong>e compared to<br />

wild type. If a WRKY gene fails to be upregulated <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 background it<br />

implies that it is a downstream target of this signal<strong>in</strong>g cascade. However WRKY genes<br />

are part of a large family of closely related prote<strong>in</strong>s with 37 members <strong>in</strong> <strong>Physcomitrella</strong><br />

and 74 <strong>in</strong> Arabidopsis (Ülker and Somssich, 2004; Rens<strong>in</strong>g et al., 2008). Thus, a clear<br />

orthologous relationship can not be established and <strong>the</strong>refore <strong>the</strong> expression of several<br />

WRKY homologs was assessed. The closest homologs of AtWRKY25 and 33 were not<br />

differentially expressed <strong>in</strong> a prelim<strong>in</strong>ary screen of ΔPpMPK4A-1 and wild type 30 m<strong>in</strong>, 1<br />

h and 2 h after chitosan treatment (data not shown). In Figure 31, PpWRKY40 is clearly<br />

also not differentially expressed, while PpWRKY53 and PpWRKY70 could be<br />

differentially expressed. However, <strong>the</strong> result should be replicated to be conv<strong>in</strong>c<strong>in</strong>g.<br />

Ano<strong>the</strong>r downstream target of AtMPK4 <strong>in</strong> Arabidopsis is AtCYP71A13, which<br />

encode a cytochrome P450 monoxygenase required for syn<strong>the</strong>sis of <strong>the</strong> antimicrobial<br />

phytoalex<strong>in</strong> camalex<strong>in</strong> (Petersen et al., 2008). The closest AtCYP71A13 homolog <strong>in</strong><br />

<strong>Physcomitrella</strong> was not regulated by chitosan (Figure 31). But <strong>the</strong> cytochrome P450<br />

super family has 245 members <strong>in</strong> Arabidopsis (Schuler et al., 2006) and it is likely that<br />

<strong>the</strong> homolog we identified is not a functional ortholog.<br />

We looked at <strong>the</strong> expression of PpPRX34 s<strong>in</strong>ce this peroxidase was shown to be<br />

<strong>in</strong>volved <strong>in</strong> defense aga<strong>in</strong>st pathogenic fungi <strong>in</strong> <strong>Physcomitrella</strong> (Lehtonen et al., 2009).<br />

However PpPRX34 did not seem to be regulated by PpMPK4A (Figure 31).<br />

The ethylene response factor (ERF) family <strong>in</strong> Arabidopsis <strong>in</strong>cludes 122 members.<br />

The biological functions of <strong>the</strong> majority rema<strong>in</strong> unknown, although several ERF genes<br />

are upregulated <strong>in</strong> Arabidopsis upon chit<strong>in</strong> treatment (Libault et al., 2007), and <strong>the</strong><br />

Aterf5/Aterf6 double mutant showed a significant <strong>in</strong>crease <strong>in</strong> susceptibility to B. c<strong>in</strong>erea<br />

(Moffat et al., 2012). Fur<strong>the</strong>rmore, AtERF5 <strong>in</strong>teracts with AtMPK3 and AtMPK6 which<br />

are <strong>in</strong>volved <strong>in</strong> defense aga<strong>in</strong>st A. brassicicola (Son et al., 2012). PpERF2 was<br />

differentially regulated <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 l<strong>in</strong>e compared to wild type after chitosan<br />

treatment (Figure 4, manuscript 1). However, <strong>the</strong> PpERFb3 and PpERF5 homologs only<br />

seem to be differentially regulated <strong>in</strong> <strong>the</strong> very early response (at 15 m<strong>in</strong> for PpERF5 and<br />

from 15 m<strong>in</strong> to 1 h for PpERF3b) (Figure 31).<br />

PAL production is important for <strong>the</strong> defense aga<strong>in</strong>st B. c<strong>in</strong>erea <strong>in</strong> Arabidopsis<br />

(Ferrari et al., 2003). We found that PpPAL4 was differentially regulated <strong>in</strong> <strong>the</strong><br />

ΔPpMPK4A-1 l<strong>in</strong>e compared to wild type after chitosan treatment (Figure 4, manuscript<br />

1). Ano<strong>the</strong>r PAL4 homolog PpPAL4-2 may be less upregulated <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1<br />

background after chitosan treatment, but <strong>the</strong> result should be repeated to be conv<strong>in</strong>c<strong>in</strong>g<br />

(Figure 31).<br />

Gene expression <strong>in</strong> ΔMPK4A l<strong>in</strong>es upon B. c<strong>in</strong>erea <strong>in</strong>fection<br />

We wanted to test if changes <strong>in</strong> gene expression upon treatment with <strong>the</strong> fungal MAMP<br />

chitosan were also reflected <strong>in</strong> gene expression upon B. c<strong>in</strong>erea <strong>in</strong>fection. Thus, we<br />

sprayed <strong>moss</strong> colonies with a spore suspension of 2x10 6 spores/ml, <strong>the</strong> same<br />

concentration used for <strong>the</strong> successful MAPK phosphorylation shown <strong>in</strong> Figure 5,<br />

manuscript 1, but it is ten times higher than <strong>the</strong> concentration used <strong>in</strong> <strong>the</strong> quantitative<br />

Evans blue sta<strong>in</strong><strong>in</strong>g assay upon B. c<strong>in</strong>erea <strong>in</strong>fection (Figure 3B, manuscript 1). Even with<br />

this high spore concentration we did not see much <strong>in</strong>duction of <strong>the</strong> genes assessed until<br />

48


after 4 hours (Figure 32). The reason for this is that sporulation starts after approximately<br />

5 hours, and <strong>the</strong> changes <strong>in</strong> <strong>moss</strong> gene expression are probably not only controlled via a<br />

MAMP <strong>in</strong>duced signal<strong>in</strong>g cascade activated by <strong>the</strong> chit<strong>in</strong> receptor. Instead, <strong>the</strong> changes<br />

<strong>in</strong> gene regulation are probably caused by <strong>the</strong> an array of signals from <strong>the</strong> sporulat<strong>in</strong>g<br />

fungi because, upon sporulation, <strong>the</strong> fungus starts <strong>in</strong>vad<strong>in</strong>g <strong>the</strong> <strong>moss</strong> cells and it is no<br />

longer a situation of immunity triggered by MAMPs alone. HR is probably also triggered,<br />

cells are killed and DAMP molecules are released, which all are factors that could alter<br />

gene expression <strong>in</strong> <strong>the</strong> <strong>moss</strong> colony.<br />

MPK4B<br />

CHS<br />

Relative expression<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0h 15m 30m 1h 2h 4h 8h 16h<br />

Relative expression<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

WT<br />

MPK4A-1 KO<br />

MPK4A-2 KO<br />

0h 15m 30m 1h 2h 4h 8h 16h<br />

a-DOX<br />

PAL<br />

120<br />

120<br />

Relative expression<br />

80<br />

40<br />

0<br />

0h 15m 30m 1h 2h 4h 8h 16h<br />

Relative expression<br />

90<br />

60<br />

30<br />

0<br />

0h 15m 30m 1h 2h 4h 8h 16h<br />

Figure 32. Quantitative reverse transcriptase PCR (qPCR) analysis of transcript levels <strong>in</strong> WT (blue<br />

diamonds), ΔPpMPK4A-1 (magenta squares), and ΔPpMPK4A-2 (red triangles) relative to untreated<br />

WT (time 0h) follow<strong>in</strong>g <strong>in</strong>fection with 2x10 6 B. c<strong>in</strong>erea spores/ml. Error bars represent SEM of three<br />

<strong>in</strong>dependent technical replicates. Gene identifiers and primers used are <strong>in</strong> Table 10. These qPCR<br />

experiments where carried out by bachelor students Søren Iversen and Mira Wilkans under my<br />

supervision.<br />

ROS production upon chitosan treatment<br />

Production of reactive oxygen species (ROS) is a key feature of an activated immune<br />

system, and MAP k<strong>in</strong>ases have been implicated <strong>in</strong> <strong>the</strong> production of ROS (Nakagami et<br />

al., 2006). Indeed, <strong>the</strong> genes shown to be differentially regulated upon chitosan treatment<br />

described <strong>in</strong> manuscript 1 are all oxidative stress associated genes. They are ei<strong>the</strong>r<br />

directly <strong>in</strong>volved <strong>in</strong> ROS production (NOX) (Miller et al., 2009), or <strong>in</strong> ROS detoxification<br />

(a-DOX, CHS and PAL) (Ponce de León et al., 2002; Apel and Hirt, 2004), or are known<br />

to be regulated by redox level (LOX7 and ERF2) (Dietz et al., 2010; Ponce De León et al.,<br />

2012). Therefore, we wanted to <strong>in</strong>vestigate whe<strong>the</strong>r ROS accumulation upon chitosan<br />

treatment was altered <strong>in</strong> <strong>the</strong> ΔPpMPK4A l<strong>in</strong>es.<br />

49


The production of ROS was visualized by sta<strong>in</strong><strong>in</strong>g with DAB 24 hours after<br />

treat<strong>in</strong>g <strong>the</strong> wild type and <strong>the</strong> two ΔPpMPK4A mutants with chitosan (Figure 33). The<br />

wild type colonies sta<strong>in</strong>ed darker with DAB, <strong>in</strong>dicat<strong>in</strong>g that more ROS had accumulated<br />

<strong>in</strong> <strong>the</strong>se plants than <strong>in</strong> <strong>the</strong> KO l<strong>in</strong>es. Some background sta<strong>in</strong><strong>in</strong>g occurred equally <strong>in</strong> all<br />

l<strong>in</strong>es sta<strong>in</strong>ed with DAB but treated with control, and also some darken<strong>in</strong>g occurred <strong>in</strong> all<br />

l<strong>in</strong>es only treated with chitosan and not sta<strong>in</strong>ed with DAB.<br />

Figure 33. ROS accumulation. DAB sta<strong>in</strong><strong>in</strong>g of wild type and two <strong>in</strong>dependent ΔPpMPK4A (KO1 and<br />

KO2) l<strong>in</strong>es 24 hours after treatment with 100 µg/ml chitosan (Chi, <strong>in</strong> 0.01% Acetic acid, adjusted to<br />

pH 5.5 with NaOH) or control treatment (C, 0.01% Acetic acid, adjusted to pH 5.5 with NaOH). The<br />

experiment was repeated with similar results.<br />

The failure to accumulate ROS upon chitosan treatment <strong>in</strong> <strong>the</strong> ΔPpMPK4A l<strong>in</strong>es<br />

implicates this MPK <strong>in</strong> transduc<strong>in</strong>g signals from <strong>the</strong> MAMP receptor to activate defense<br />

responses. To substantiate this result, we attempted to quantify ROS accumulation <strong>in</strong> <strong>the</strong><br />

l<strong>in</strong>es us<strong>in</strong>g a fluorimetric technique with <strong>the</strong> dye Amplex Red. However, we have not<br />

succeeded with this yet.<br />

50


MPK phosphorylation western blots<br />

S<strong>in</strong>ce AtMPK4 has been implicated <strong>in</strong> regulat<strong>in</strong>g some responses to <strong>the</strong> phytohormones<br />

ethylene (ET) and jasmonic acid (JA) (Brodersen et al., 2006), we tested if treatment with<br />

precursors of <strong>the</strong>se hormones could <strong>in</strong>duce phosphorylation of MPKs <strong>in</strong> <strong>the</strong> <strong>moss</strong>.<br />

Arabidopsis MPKs are also phosphorylated <strong>in</strong> response to <strong>the</strong> ethylene precursor 1-<br />

am<strong>in</strong>ocyclopropane-1-carboxylic acid (ACC) (Yoo et al., 2008). And while JA has not<br />

been found <strong>in</strong> <strong>Physcomitrella</strong> (Stumpe et al., 2010), Ponce de león et al. (2012) recently<br />

showed that <strong>Physcomitrella</strong> is capable of respond<strong>in</strong>g to methyl jasmonate (MeJA).<br />

Therefore, we tested if this response <strong>in</strong>volved MPK phosphorylation. However, we did<br />

not detect phosphorylation of any <strong>moss</strong> MPKs <strong>in</strong> response to <strong>the</strong>se hormone treatments<br />

(Figure 34).<br />

Figure 34. Immunoblot analysis with anti-phospho-p44/42 MAPK antibodies at <strong>in</strong>dicated m<strong>in</strong>utes<br />

after spray<strong>in</strong>g with ei<strong>the</strong>r 1 mg/ml chit<strong>in</strong>, 100 µM MeJA and 1 mM 1-am<strong>in</strong>ocyclopropane-1-<br />

carboxylic acid (ACC). Load<strong>in</strong>g controls show amido black sta<strong>in</strong>ed total prote<strong>in</strong>.<br />

Sporophyte <strong>in</strong>duction<br />

We recently discovered that <strong>the</strong> ΔPpMPK4A l<strong>in</strong>es appear to be <strong>in</strong>hibited <strong>in</strong> <strong>the</strong>ir ability to<br />

form sporophytes (Figure 35). On several <strong>in</strong>dependent plates of both ΔPpMPK4A l<strong>in</strong>es,<br />

<strong>the</strong>re were almost no sporophytes while <strong>the</strong> wild type of <strong>the</strong> same age had sporophytes<br />

formed on top of virtually all gametophores. On a whole plate of both ΔPpMPK4A-1 and<br />

ΔPpMPK4A-2, only one or two sporophytes could be discerned. This is a very<br />

prelim<strong>in</strong>ary result and it is currently be<strong>in</strong>g repeated with all KO l<strong>in</strong>es <strong>in</strong>cluded.<br />

51


Figure 35. Sporophyte <strong>in</strong>duction. 10 week old gametophores of wild type and ΔPpMPK4A-1 grown<br />

for four weeks <strong>in</strong> standard conditions, <strong>the</strong>n transferred to 17°C and short day with low light<br />

conditions to <strong>in</strong>duce sporophyte formation. Scale bars (A, B) 3 mm and (C, D) 1 mm.<br />

Arabidopsis AtMPK4 has been shown to function <strong>in</strong> both somatic and meiotic cytok<strong>in</strong>esis<br />

result<strong>in</strong>g <strong>in</strong> retarded root tip growth and abnormal pollen formation <strong>in</strong> <strong>the</strong> Atmpk4 mutant<br />

(Kosetsu et al., 2010; Zeng et al., 2011). Thus, it is not unlikely that PpMPK4A is<br />

<strong>in</strong>volved <strong>in</strong> a developmental process like sporophyte formation.<br />

MAMP growth assay<br />

Grow<strong>in</strong>g Arabidopsis seedl<strong>in</strong>gs on plates supplemented with <strong>the</strong> MAMPs flg22 or elf18<br />

results <strong>in</strong> growth retardation which is a useful tool to identify mutants of genes <strong>in</strong>volved<br />

<strong>in</strong> sens<strong>in</strong>g <strong>the</strong>se two MAMPs. However, <strong>the</strong> chit<strong>in</strong> MAMP does not stunt growth <strong>in</strong><br />

Arabidopsis and this assay can <strong>the</strong>refore not be used to identify components <strong>in</strong> chit<strong>in</strong><br />

signal<strong>in</strong>g. <strong>Physcomitrella</strong> was grown on medium supplemented with 1 mg/ml chitosan<br />

but this also did not produce any visible difference from control plants (data not shown)<br />

ΔRAR-1 and R-gene KOs<br />

Most R-prote<strong>in</strong>s <strong>in</strong> flower<strong>in</strong>g plants need <strong>the</strong> RAR1-SGT1-HSP90 chaperone complex to<br />

be correctly folded and ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> a recognition competent stage <strong>in</strong> <strong>the</strong> correct cellular<br />

location. Genetic screens for loss of resistance have shown that RAR1 is required for <strong>the</strong><br />

function of multiple and dist<strong>in</strong>ct R-prote<strong>in</strong>s <strong>in</strong> both monocots and dicots (Shirasu, 2009).<br />

Activat<strong>in</strong>g <strong>the</strong> R-prote<strong>in</strong>s results <strong>in</strong> HR which will <strong>in</strong>duce localized PCD. The<br />

hypersensitive response is an effective defense aga<strong>in</strong>st biotrophic plant pathogens,<br />

52


estrict<strong>in</strong>g access to water and nutrients, but can be exploited by necrotrophic pathogens<br />

such as B. c<strong>in</strong>erea to generate dead tissue around <strong>the</strong> <strong>in</strong>fected area and thus facilitat<strong>in</strong>g its<br />

growth (Govr<strong>in</strong> and Lev<strong>in</strong>e, 2000). Specifically, B. c<strong>in</strong>erea has been shown to secrete an<br />

elicitor that <strong>in</strong>duces HR <strong>in</strong> Arabidopsis (Govr<strong>in</strong> et al., 2006) and such a secreted tox<strong>in</strong> has<br />

recently been identified as <strong>the</strong> cerato-platan<strong>in</strong> prote<strong>in</strong> BcSpl1 (Frías et al., 2011). Frías et<br />

al. (2011), showed that BcSpl1 is required for full B. c<strong>in</strong>erea virulence and that <strong>the</strong><br />

prote<strong>in</strong> <strong>in</strong>duced cell death with symptoms of HR <strong>in</strong> Arabidopsis, tobacco and tomato. It<br />

has previously been shown <strong>in</strong> tobacco (Nicotiana benthamiana) that virus-<strong>in</strong>duced gene<br />

silenc<strong>in</strong>g of NbSGT1 and NbEDS1 enhanced plant resistance to B. c<strong>in</strong>erea (El Oirdi and<br />

Bouarab, 2007). Toge<strong>the</strong>r <strong>the</strong>se data <strong>in</strong>dicate that B. c<strong>in</strong>erea is able to hijack <strong>the</strong> host<br />

immune system by secret<strong>in</strong>g an elicitor that activates <strong>the</strong> R-prote<strong>in</strong> system to <strong>in</strong>duce HR,<br />

<strong>the</strong>reby caus<strong>in</strong>g cell death which releases nutrients and facilitates necrotrophic growth.<br />

Evans blue sta<strong>in</strong><strong>in</strong>g upon B. c<strong>in</strong>erea <strong>in</strong>fection of ΔPpRAR1<br />

In <strong>the</strong> <strong>in</strong>itial screen of quantitative Evans blue sta<strong>in</strong><strong>in</strong>g upon B. c<strong>in</strong>erea <strong>in</strong>fection, <strong>the</strong> two<br />

<strong>in</strong>dependent ΔRAR1 l<strong>in</strong>es had less sta<strong>in</strong><strong>in</strong>g than <strong>the</strong> wild type and thus seemed less<br />

susceptible to B. c<strong>in</strong>erea (Figure 27). This result was very excit<strong>in</strong>g s<strong>in</strong>ce RAR1 is part of<br />

<strong>the</strong> co-chaperone RAR1/SGT1/HSP90 that stabilizes R-genes, and SGT1 has been shown<br />

to be required for full B. c<strong>in</strong>erea virulence (El Oirdi and Bouarab, 2007). Thus, <strong>the</strong> less<br />

susceptible ΔPpRAR1 l<strong>in</strong>es <strong>in</strong>dicate that <strong>Physcomitrella</strong> has a functional R-gene system<br />

capable of <strong>in</strong>duc<strong>in</strong>g programmed cell death. To our knowledge, <strong>the</strong>re have been no<br />

descriptions of functional R-genes <strong>in</strong> any non-vascular plants. Therefore we expended<br />

efforts to replicate this <strong>in</strong>itial result. However, <strong>in</strong> subsequent assays unexpla<strong>in</strong>able<br />

variation made <strong>the</strong> results <strong>in</strong>conclusive (data not shown). Therefore, <strong>the</strong> role of PpRAR1<br />

<strong>in</strong> B. c<strong>in</strong>erea <strong>in</strong>fection rema<strong>in</strong>s unknown.<br />

S<strong>in</strong>ce B. c<strong>in</strong>erea apparently manipulates <strong>the</strong> R-gene system to facilitate its own<br />

growth, we also tested whe<strong>the</strong>r <strong>the</strong> four R-gene KOs had altered susceptibility to <strong>the</strong><br />

fungus. But none of <strong>the</strong> R-gene KOs showed any significant difference <strong>in</strong> a quantitative<br />

Evans blue sta<strong>in</strong><strong>in</strong>g upon B. c<strong>in</strong>erea <strong>in</strong>fection (data not shown).<br />

Infections with biotrophic pathogens<br />

No obligate biotrophic pathogen has been described that <strong>in</strong>fects <strong>Physcomitrella</strong>. This<br />

makes it difficult to <strong>in</strong>vestigate <strong>the</strong> role of PpRAR1 and R-prote<strong>in</strong>s <strong>in</strong> <strong>Physcomitrella</strong><br />

s<strong>in</strong>ce R-prote<strong>in</strong>s are primarily <strong>in</strong>volved <strong>in</strong> defense aga<strong>in</strong>st biotrophic pathogens. We<br />

<strong>the</strong>refore hoped that ei<strong>the</strong>r <strong>the</strong> ΔRAR-1 l<strong>in</strong>es or <strong>the</strong> four R-gene KOs could render <strong>the</strong><br />

<strong>moss</strong> susceptible to a biotrophic pathogen. However, <strong>in</strong>fection with <strong>the</strong> biotrophic<br />

pathogens listed <strong>in</strong> Table 5 did not yield disease symptoms on any of <strong>the</strong> mutants from<br />

<strong>the</strong> first round of transformations (data not shown). The <strong>in</strong>fections were performed on<br />

both protonemal tissues and gametophores, both by spray<strong>in</strong>g high concentrations of<br />

bacteria and by plac<strong>in</strong>g a small colony of bacteria directly on top of a colony. Prewound<strong>in</strong>g<br />

<strong>the</strong> <strong>moss</strong> was also attempted, s<strong>in</strong>ce this has previously been shown to facilitate<br />

a pathogen <strong>in</strong>fection on <strong>moss</strong> (Andersson et al., 2005; Ponce de León et al., 2007).<br />

53


Pseudomonas syr<strong>in</strong>gae pv. Tomato (Pto) DC3000<br />

Hyaloperonospora parasitica (Hpa) isolate Noco2<br />

Hyaloperonospora parasitica (Hpa) isolate Emwa1<br />

Hyaloperonospora parasitica (Hpa) isolate Cala2<br />

Table 5. Biotrophic pathogens used to challenge <strong>Physcomitrella</strong>.<br />

Yeast two-hybrid analysis of PpRAR1 and PpSGT1<br />

The functions of <strong>the</strong> <strong>Physcomitrella</strong> PpRAR1, PpSGT1 and <strong>the</strong> R-prote<strong>in</strong>s homologs are<br />

unknown. S<strong>in</strong>ce <strong>the</strong> <strong>in</strong>teraction of RAR1 and SGT1 <strong>in</strong> flower<strong>in</strong>g plants is required for<br />

<strong>the</strong>ir function <strong>in</strong> stabiliz<strong>in</strong>g R-prote<strong>in</strong>s (Shirasu, 2009), a functionally orthologous<br />

relationship would be streng<strong>the</strong>ned if <strong>the</strong> <strong>Physcomitrella</strong> PpRAR1 and PpSGT1<br />

homologs <strong>in</strong>teract. We <strong>the</strong>refore tested if PpRAR1 and PpSGT1 <strong>in</strong>teract us<strong>in</strong>g a GAL4<br />

based, yeast two hybrid system (Matchmaker tm from Clonetech).<br />

Full length CDS of PpRAR1 and PpSGT1 were amplified from cDNA with<br />

USER extended primers (Table 9) and cloned <strong>in</strong>to both <strong>the</strong> b<strong>in</strong>d<strong>in</strong>g doma<strong>in</strong> vector<br />

(pGADT7) and <strong>the</strong> activation doma<strong>in</strong> vector (pGBKT7). Both vectors were modified to<br />

conta<strong>in</strong> a USER clon<strong>in</strong>g cassette (Elizabeth Peanuts, B.Sc. <strong>the</strong>sis) and correct <strong>in</strong>sertions<br />

of PpRAR1 and PpSGT1 were verified by sequenc<strong>in</strong>g. The vectors were transformed <strong>in</strong><br />

pairs <strong>in</strong>to yeast. The four vectors were also co-transformed with an empty vector as<br />

counterpart to check for auto-activation. F<strong>in</strong>ally, a vector pair encod<strong>in</strong>g two prote<strong>in</strong>s that<br />

have previously been shown to <strong>in</strong>teract (AtGF6 and AtLAZ1) were <strong>in</strong>cluded as a positive<br />

control. In total n<strong>in</strong>e different cotransformations were performed (Table 6).<br />

# pGADT7 pGBKT7<br />

1 RAR1 SGT1<br />

2 SGT1 RAR1<br />

3 RAR1 RAR1<br />

4 SGT1 SGT1<br />

5 RAR1 Empty<br />

6 SGT1 Empty<br />

7 Empty RAR1<br />

8 Empty SGT1<br />

9 GF6 LAZ1<br />

Table 6. Vector comb<strong>in</strong>ations co-transformed <strong>in</strong>to yeast for two-hybrid analysis<br />

Upon transformation <strong>in</strong>to yeast <strong>the</strong> presence of <strong>the</strong> expected vectors was verified by PCR<br />

(data not shown). One to four colonies from each co-transformation that had <strong>the</strong> expected<br />

vector present were restreaked onto SD-2 medium (double dropout medium lack<strong>in</strong>g Leu<br />

and Trp) (Figure 36). All colonies were able to grow on SD-2 confirm<strong>in</strong>g <strong>the</strong> presence of<br />

both <strong>the</strong> activat<strong>in</strong>g doma<strong>in</strong> and <strong>the</strong> b<strong>in</strong>d<strong>in</strong>g doma<strong>in</strong> vectors. All colonies were <strong>the</strong>n<br />

restreaked onto SD-4 (quadruple dropout medium lack<strong>in</strong>g Leu, Trp, His and Ade). Yeast<br />

colonies conta<strong>in</strong><strong>in</strong>g PpSGT1 <strong>in</strong> <strong>the</strong> activat<strong>in</strong>g doma<strong>in</strong> vector and PpRAR1 <strong>in</strong> <strong>the</strong> b<strong>in</strong>d<strong>in</strong>g<br />

doma<strong>in</strong> vector were able to grow on <strong>the</strong> SD-4 medium <strong>in</strong>dicat<strong>in</strong>g <strong>in</strong>teraction between <strong>the</strong><br />

two prote<strong>in</strong>s <strong>in</strong> <strong>the</strong> yeast (Fig. 33, 2,2 on SD-4). No colonies grew when <strong>the</strong> two prote<strong>in</strong>s<br />

were co-transformed with <strong>the</strong>mselves as <strong>in</strong>teract<strong>in</strong>g partners (#3 and #4), show<strong>in</strong>g that<br />

<strong>the</strong> two prote<strong>in</strong>s do not form homodimers <strong>in</strong> yeast. Also no auto-activation was observed<br />

54


when <strong>the</strong> vectors were transformed with an empty vector as counterpart. F<strong>in</strong>ally <strong>the</strong>re<br />

was growth <strong>in</strong> <strong>the</strong> positive control, although weak (Fig. 33, 9,4 on SD-4).<br />

Figure 36. Yeast two –hybrid analysis of PpRAR1 and PpSGT1. Co-transformed yeast colonies<br />

(Table 6 ) grown for four days on SD-2 and SD-4. This experiment was performed by lab technician<br />

student Ali Fard for his f<strong>in</strong>al <strong>the</strong>sis under my supervision.<br />

Gene expression <strong>in</strong> ΔRAR1 upon chitosan treatment<br />

When treat<strong>in</strong>g <strong>the</strong> wild type with <strong>the</strong> fungal MAMP chitosan we found that PpRAR1 and<br />

PpSGT1 expression was <strong>in</strong>duced (Figure 37). We <strong>the</strong>refore looked at marker gene<br />

expression <strong>in</strong> <strong>the</strong> ΔPpRAR1-1 l<strong>in</strong>e to see if we could f<strong>in</strong>d differentially expressed genes<br />

that could lead us to f<strong>in</strong>d a phenotype of <strong>the</strong> mutant. Despite screen<strong>in</strong>g <strong>the</strong> expression of<br />

most of <strong>the</strong> genes shown <strong>in</strong> Figure 31 and Figure 4 of manuscript 1, we did not f<strong>in</strong>d any<br />

genes that were significantly differentially expressed <strong>in</strong> ΔPpRAR1-1 compared to <strong>the</strong><br />

wild type. These data are <strong>the</strong>refore omitted and only an example, <strong>the</strong> expression of<br />

PpSGT1 <strong>in</strong> <strong>the</strong> KO and <strong>in</strong> wild type, is shown <strong>in</strong> Figure 37.<br />

RAR1<br />

SGT1<br />

Relative expression<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

WT<br />

rar1-1<br />

0h 15m 30m 1h 2h 4h 8h<br />

Relative expression<br />

3<br />

2<br />

1<br />

0<br />

WT<br />

rar1-1<br />

0h 15m 30m 1h 2h 4h 8h<br />

Figure 37. Quantitative reverse transcriptase PCR (qPCR) analysis of transcript levels <strong>in</strong> WT (blue<br />

diamonds) and ΔPpRAR1-1 (purple triangles) relative to untreated WT (time 0h) follow<strong>in</strong>g treatment<br />

with 100 μg/ml chitosan. In cases where <strong>the</strong>re are no error bars <strong>the</strong> time po<strong>in</strong>t is based on a s<strong>in</strong>gle<br />

55


technical replicate, o<strong>the</strong>rwise error bars represent SEM of three <strong>in</strong>dependent technical replicates.<br />

Gene identifiers and primers used are <strong>in</strong> Table 10.<br />

MPK phosporylation <strong>in</strong> ΔRAR1<br />

We also checked if <strong>the</strong> phosphorylation of MPKs upon MAMP treatment with chit<strong>in</strong> was<br />

altered <strong>in</strong> <strong>the</strong> ΔPpRAR1 l<strong>in</strong>es, but it was not different from <strong>the</strong> wild type (data not shown)<br />

The autophagy deficient mutant ΔATG5<br />

Autophagy is <strong>in</strong>volved <strong>in</strong> many cellular processes <strong>in</strong>clud<strong>in</strong>g nutrient recycl<strong>in</strong>g of old and<br />

damaged cellular compartments (Liu and Bassham, 2012). Many autophagy deficient<br />

mutants <strong>in</strong>clud<strong>in</strong>g Atatg5 <strong>in</strong> Arabidopsis display early senescence and cell death upon<br />

starvation (Thompson et al., 2005; Liu and Bassham, 2012).<br />

To our knowledge <strong>the</strong>re are no publications on autophagy <strong>in</strong> <strong>Physcomitrella</strong> or<br />

any o<strong>the</strong>r non-vascular plant. Thus, s<strong>in</strong>ce <strong>the</strong>re have been no prior description of an<br />

autophagy deficient mutant <strong>in</strong> <strong>Physcomitrella</strong>, <strong>the</strong> first experiments with <strong>the</strong> ΔATG5 l<strong>in</strong>es<br />

were to carefully describe any detectable phenotype when grown under different<br />

conditions.<br />

Phenotypic description of <strong>the</strong> ΔATG5 l<strong>in</strong>es<br />

Figure 38. Wild type and ΔATG5-1 grown on BCD+AT medium overlaid with cellophane for 12 days.<br />

Bar corresponds to 1 cm.<br />

The two <strong>in</strong>dependent ΔATG5 l<strong>in</strong>es display clear signs of autophagy defects. When grown<br />

on BCD+AT medium overlaid with cellophane <strong>the</strong>y start grow<strong>in</strong>g like wild type but after<br />

approximately 8 days <strong>the</strong> middle and oldest parts of <strong>the</strong> colony turn yellow and die with<strong>in</strong><br />

<strong>in</strong> a few days (Figure 38). In order to get a better description of <strong>the</strong> ΔATG5 phenotype,<br />

<strong>the</strong> two KO l<strong>in</strong>es were grown on media conta<strong>in</strong><strong>in</strong>g different nutritional supplements and<br />

also subjected to a week of low light to attempt to <strong>in</strong>duce carbon starvation (Figure 39).<br />

BCD medium is <strong>the</strong> m<strong>in</strong>imal medium established by Ashton and Cove (1977).<br />

The ΔATG5 l<strong>in</strong>es grow a bit slower than wild type on this medium and after three weeks<br />

<strong>the</strong> oldest parts of <strong>the</strong> plant start to die. An additional week of growth <strong>in</strong> low light kills<br />

<strong>the</strong> ΔATG5 l<strong>in</strong>es whereas <strong>the</strong> wild type stop proliferat<strong>in</strong>g but o<strong>the</strong>rwise appears<br />

unaffected (Figure 39).<br />

BCD+AT medium is BCD supplemented with 5 mM of ammonium tartrate as a<br />

nitrogen source. This medium makes both <strong>the</strong> wild type and <strong>the</strong> KO l<strong>in</strong>es grow faster, but<br />

still <strong>the</strong> ΔATG5 l<strong>in</strong>es grown a bit slower. After one week of low light <strong>the</strong> ΔATG5 l<strong>in</strong>es<br />

are dy<strong>in</strong>g but <strong>the</strong> leaf tips and <strong>the</strong> edge of <strong>the</strong> colony with protonemal tissue are still<br />

56


green and viable such that it can still be propagated by mov<strong>in</strong>g green parts to a new plate.<br />

The wild type was fully viable but had only grown slightly <strong>in</strong> <strong>the</strong> low light (Figure 39).<br />

Complete medium is designed to facilitate <strong>the</strong> growth of metabolic mutants<br />

(Egener et al., 2002). It conta<strong>in</strong>s a range of vitam<strong>in</strong>s and nutrient supplements <strong>in</strong>clud<strong>in</strong>g<br />

peptides but no carbon source. Complete medium clearly altered <strong>the</strong> growth of both <strong>the</strong><br />

wild type and ΔATG5 l<strong>in</strong>es to a dense growth with less gametophore formation, but <strong>the</strong><br />

KO l<strong>in</strong>es and <strong>the</strong> wild type grew at <strong>the</strong> same rate. After three weeks of growth on this<br />

medium <strong>the</strong> center of <strong>the</strong> ΔATG5 colonies started to die and an additional week of low<br />

light killed most of <strong>the</strong> colony, but not <strong>the</strong> leaf tips and <strong>the</strong> edges of <strong>the</strong> colony, whereas<br />

<strong>the</strong> wild type was completely viable but did not grow fur<strong>the</strong>r <strong>in</strong> low light (Figure 39).<br />

Complete + 5 g/L glucose medium <strong>in</strong>duced a more dense protonemal growth with<br />

less protonemal tissue differentiat<strong>in</strong>g <strong>in</strong>to gametophores <strong>in</strong> both <strong>the</strong> wild type and <strong>the</strong><br />

ΔATG5 l<strong>in</strong>es. Still, some older parts of <strong>the</strong> KO colonies started to die by three weeks of<br />

growth, but less than <strong>in</strong> <strong>the</strong> media without glucose supplement. Also, <strong>the</strong> added sugar<br />

made <strong>the</strong> KO l<strong>in</strong>es better at withstand<strong>in</strong>g a week of low light, although most of <strong>the</strong> older<br />

center parts of <strong>the</strong> colonies were still dy<strong>in</strong>g. Glucose enriched media is known to <strong>in</strong>duce<br />

<strong>the</strong> formation of caulonemal filaments (Thelander et al., 2005) and it seemed that <strong>the</strong><br />

added sugar <strong>in</strong>duced a strong radial growth of what appeared to be caulonemal filaments<br />

<strong>in</strong> <strong>the</strong> wild type which was <strong>in</strong>hibited <strong>in</strong> <strong>the</strong> KO l<strong>in</strong>es (Figure 39)<br />

Complete + 50g/L glucose medium completely stops <strong>the</strong> formation of<br />

gametophores. The colonies grow very slow and very dense, probably due to <strong>the</strong> osmotic<br />

stress caused by <strong>the</strong> high sugar content of <strong>the</strong> medium. But, when grown on this medium,<br />

it is not possible to dist<strong>in</strong>guish <strong>the</strong> ΔATG5 l<strong>in</strong>es from <strong>the</strong> wild type after three weeks of<br />

growth under standard light conditions. When grown for an additional week of low light<br />

<strong>the</strong> wild type had clearly visible radial growth of brown caulonemal filaments. But <strong>the</strong><br />

ΔATG5 l<strong>in</strong>es had less radial growth and <strong>the</strong> tissue appeared greener as if it had a higher<br />

content of chloroplasts and thus resembled chloronemal tissue more than caulonemal<br />

tissue (Figure 39 and close up <strong>in</strong> Figure 40). Fur<strong>the</strong>r microscopy analysis is required to<br />

establish if <strong>the</strong> radial grow<strong>in</strong>g filaments are <strong>in</strong>deed caulonemata <strong>in</strong> <strong>the</strong> wild type and<br />

chloronemata <strong>in</strong> <strong>the</strong> ΔATG5 l<strong>in</strong>es.<br />

In Arabidopsis it has been shown that autophagy is required for chloroplast<br />

degradation dur<strong>in</strong>g dark <strong>in</strong>duced senescence (Wada et al., 2009). There are many<br />

examples of autophagy required for cell differentiation dur<strong>in</strong>g development <strong>in</strong> both plants<br />

and animals (Liu and Bassham, 2012). As no examples have yet been published <strong>in</strong><br />

<strong>Physcomitrella</strong>, it would be <strong>in</strong>terest<strong>in</strong>g to <strong>in</strong>vestigate if autophagy is <strong>in</strong>volved <strong>in</strong> <strong>the</strong><br />

differentiation of chloronemata <strong>in</strong>to caulonemata.<br />

57


Figure 39. Wild type and two <strong>in</strong>dependent ΔATG5 l<strong>in</strong>es on <strong>the</strong> <strong>in</strong>dicated media. Left, growth after<br />

three weeks of standard light conditions (55 µE·m -2·s -1 ), and right <strong>the</strong> same colonies subjected to an<br />

additional week of low light (3 µE·m -2·s -1 ). The bar <strong>in</strong> <strong>the</strong> right bottom corner represents 8 mm. The<br />

experiment was repeated with similar results.<br />

Figure 40. Wild type and ΔATG5-1 grown for three weeks <strong>in</strong> standard light conditions and an<br />

additional week of low light on full +50 g glucose medium. Bar represent 5 mm.<br />

The ma<strong>in</strong> conclusions from this experiment are:<br />

1. There is no phenotypic difference between <strong>the</strong> two <strong>in</strong>dividual ΔATG5 l<strong>in</strong>es, although<br />

ΔATG5-2 had concatemeric <strong>in</strong>sertions of <strong>the</strong> selection cassette.<br />

2. The ΔATG5 l<strong>in</strong>es are very sensitive to both nitrogen and carbon starvation, as <strong>the</strong> KOs<br />

showed visible cell death if <strong>the</strong> medium was not supplemented with additional nutrients,<br />

especially nitrogen and carbon sources. Growth <strong>in</strong> low light/dark is known to <strong>in</strong>duce<br />

autophagic recycl<strong>in</strong>g to save energy/resources. The ΔATG5 l<strong>in</strong>es are apparently not<br />

capable of do<strong>in</strong>g so and thus quickly die <strong>in</strong> <strong>the</strong> dark if <strong>the</strong> medium is not supplemented<br />

with an exogenous carbon source <strong>in</strong> <strong>the</strong> form of sugar.<br />

58


S<strong>in</strong>ce many <strong>Physcomitrella</strong> assays are carried out on protonemal tissue ra<strong>the</strong>r than<br />

differentiated gametophores, <strong>the</strong> experiment was repeated by grow<strong>in</strong>g <strong>the</strong> <strong>moss</strong> on solid<br />

media overlaid with cellophane favor<strong>in</strong>g protonemal growth (Figure 41).<br />

Figure 41. Wild type and two <strong>in</strong>dependent ΔATG5 l<strong>in</strong>es on <strong>the</strong> <strong>in</strong>dicated media overlaid with<br />

cellophane. Left, growth after two weeks of standard light conditions (55 µE·m -2·s -1 ) and right, <strong>the</strong><br />

same colonies subjected to an additional week of low light (3 µE·m -2·s -1 ). The bar <strong>in</strong> <strong>the</strong> right bottom<br />

corner represents 8 mm. The experiment was repeated with similar results.<br />

The conclusion from this experiment is much <strong>the</strong> same: <strong>the</strong> more nutrients <strong>the</strong> medium is<br />

supplemented with, <strong>the</strong> less <strong>the</strong> difference is between growth of <strong>the</strong> ΔATG5 l<strong>in</strong>es and <strong>the</strong><br />

wild type.<br />

In an effort to <strong>in</strong>vestigate whe<strong>the</strong>r nitrogen or carbon was <strong>the</strong> most important<br />

factor for ΔATG5 survival, <strong>the</strong>y were also grown on BCD plates supplemented with<br />

different amounts of glucose. But glucose alone could not rescue <strong>the</strong> KO l<strong>in</strong>es (data not<br />

shown). It was also tested if high cont<strong>in</strong>uous light to promote carbon fixation could<br />

rescue <strong>the</strong> ΔATG5 l<strong>in</strong>es when grown on different media. However, <strong>the</strong> conclusion from<br />

this experiment was that high cont<strong>in</strong>uous light (180 µE·m -2·s -1 ) did not rescue <strong>the</strong> ΔATG5<br />

phenotype (data not shown). Thus, it seems that a sufficient supply of both nitrogen,<br />

carbon and o<strong>the</strong>r nutrients is necessary to keep <strong>the</strong> autophagy deficient ΔATG5 l<strong>in</strong>es<br />

viable. These results are <strong>in</strong> accordance with f<strong>in</strong>d<strong>in</strong>gs <strong>in</strong> Arabidopsis that AtATG5 is<br />

important not only for survival under nitrogen limit<strong>in</strong>g conditions, but also dur<strong>in</strong>g carbon<br />

starvation (Thompson et al., 2005).<br />

The ΔATG5 l<strong>in</strong>es are also hypersensitive to drought (data not shown) but surpris<strong>in</strong>gly<br />

<strong>the</strong>y appear to tolerate salt stress very well (Figure 42). When grown on BCD+AT<br />

medium supplemented with ei<strong>the</strong>r 100 or 200 mM of NaCl, it is difficult to visibly<br />

dist<strong>in</strong>guish <strong>the</strong> ΔATG5 l<strong>in</strong>es from wild type (Figure 42A). And when spray<strong>in</strong>g 10 day old<br />

59


ΔATG5 l<strong>in</strong>es with 500 mM NaCl, <strong>the</strong>y survive much longer than control colonies sprayed<br />

with H 2 O (Figure 42B). This is surpris<strong>in</strong>g s<strong>in</strong>ce it has been reported that autophagy is<br />

required for tolerance to salt and osmotic stress (Liu et al., 2009). Fur<strong>the</strong>r experiments<br />

need to clarify if this result is due to <strong>the</strong> NaCl <strong>in</strong> itself, or if it is <strong>the</strong> osmotic stress that<br />

keeps <strong>the</strong> ΔATG5 l<strong>in</strong>es from dy<strong>in</strong>g. This could be achieved by <strong>in</strong>duc<strong>in</strong>g osmotic stress<br />

with osmolytes such as mannitol or sorbitol.<br />

Figure 42. Salt tolerance of <strong>the</strong> wild type and two <strong>in</strong>dependent ΔATG5 l<strong>in</strong>es. (A) Grown for three<br />

weeks on BCD+AT medium with ei<strong>the</strong>r no NaCl, 100 mM NaCl or 200 mM NaCl supplemented. (B)<br />

Grown for 10 days on BCD+AT medium and <strong>the</strong>n sprayed with ei<strong>the</strong>r H 2 O or 500 mM NaCl and left<br />

for an additional week. Scale bars represent 5 mm. The experiment was repeated with similar results.<br />

Ano<strong>the</strong>r phenotypic feature of <strong>the</strong> ΔATG5 l<strong>in</strong>es is that <strong>the</strong>ir rhizoids are much shorter<br />

than <strong>the</strong> wild type (Figure 43).<br />

Figure 43. Gametophores of wild type and ΔATG5-1 grown on BCD+AT medium for three weeks and<br />

sta<strong>in</strong>ed with toluid<strong>in</strong>e blue to visualize <strong>the</strong> rhizoids. There were no phenotypic differences between<br />

<strong>the</strong> two PpATG5 KO l<strong>in</strong>es (data not shown). The bar is 3 mm.<br />

ATG8 western blot<br />

Tunicamyc<strong>in</strong> is widely used to study autophagy. It <strong>in</strong>duces <strong>the</strong> unfolded prote<strong>in</strong> response<br />

which activates autophagy and <strong>the</strong> formation of autophagosomes. In Arabidopsis <strong>the</strong><br />

formation of autophagosomes is studied <strong>in</strong> western blots us<strong>in</strong>g an antibody aga<strong>in</strong>st<br />

AtATG8A. Dur<strong>in</strong>g <strong>the</strong> formation of autophagosomes, ATG8 is C-term<strong>in</strong>ally cleaved and<br />

lipidated which causes it to travel faster <strong>in</strong> acrylamide gels dur<strong>in</strong>g electrophoresis,<br />

<strong>the</strong>reby caus<strong>in</strong>g a detectable band shift <strong>in</strong> a western blot with ATG8 primary antibody. In<br />

60


order to see if this tool could be used to <strong>in</strong>vestigate autophagy <strong>in</strong> <strong>Physcomitrella</strong>, we<br />

treated wild type and ΔATG5-1 with tunicamyc<strong>in</strong> and did a western blot with <strong>the</strong><br />

Arabidopsis ATG8 antibody (Figure 44A). The treatment was done by transferr<strong>in</strong>g one<br />

week old protonema tissue grown on cellophane on BCD+AT plates onto BCD+AT<br />

plates supplemented with 3 µg/ml tunicamyc<strong>in</strong>. This concentration was established <strong>in</strong> a<br />

prelim<strong>in</strong>ary experiment <strong>in</strong> which it was seen that this concentration killed <strong>the</strong> ΔATG5<br />

l<strong>in</strong>es but not <strong>the</strong> wild type, while a concentration of 15 µg/ml tunicamyc<strong>in</strong> also killed <strong>the</strong><br />

wild type (data not shown). As seen from <strong>the</strong> western blot <strong>in</strong> Figure 44, <strong>the</strong>re is a dist<strong>in</strong>ct<br />

difference <strong>in</strong> <strong>the</strong> pattern of prote<strong>in</strong>s b<strong>in</strong>d<strong>in</strong>g <strong>the</strong> antibody <strong>in</strong> <strong>the</strong> ΔATG5 l<strong>in</strong>es compared to<br />

wild type. However, <strong>the</strong>re does not seem to be <strong>in</strong>duction of any bands or band shifts <strong>in</strong><br />

<strong>the</strong> wild type dur<strong>in</strong>g <strong>the</strong> four days it was grow<strong>in</strong>g on tunicamyc<strong>in</strong> plates. Thus, ei<strong>the</strong>r<br />

tunicamyc<strong>in</strong> treatment did not <strong>in</strong>duce autophagy, or <strong>the</strong> Arabidopsis antibody cannot be<br />

used to detect lipidation of <strong>Physcomitrella</strong> PpATG8.<br />

There are n<strong>in</strong>e Arabidopsis isoforms of AtATG8 and a BLASTp search with <strong>the</strong>ir<br />

sequences detected five homologs <strong>in</strong> <strong>Physcomitrella</strong>. The polyclonal antibody used had<br />

been raised aga<strong>in</strong>st almost full length AtATG8A (AA 1-117 of 122 ) (Yoshimoto et al.,<br />

2004). The sequence identity for <strong>the</strong>se 117 am<strong>in</strong>o acids of AtATG8A versus <strong>the</strong><br />

correspond<strong>in</strong>g regions of <strong>the</strong> five <strong>Physcomitrella</strong> homologs ranges from 79-82%. With<br />

such high sequence identities, it is likely that <strong>the</strong> antibody recognizes all or most of <strong>the</strong><br />

<strong>Physcomitrella</strong> PpATG8 homologs. In Arabidopsis <strong>the</strong> antibody b<strong>in</strong>ds to all homologs<br />

except AtATG8H, but it is not known how many and which ATG8 isoforms are actually<br />

lipidated dur<strong>in</strong>g autophagosome formation. Thus, it is also not known which of <strong>the</strong> five<br />

ATG8 homologs <strong>in</strong> <strong>Physcomitrella</strong> might be bound by <strong>the</strong> antibody. However, s<strong>in</strong>ce <strong>the</strong><br />

pattern is clearly different <strong>in</strong> <strong>the</strong> ΔATG5 l<strong>in</strong>es and <strong>the</strong>re is an accumulation of prote<strong>in</strong>s<br />

b<strong>in</strong>d<strong>in</strong>g <strong>the</strong> antibody, this <strong>in</strong>dicates that <strong>the</strong> AtATG8A antibody does b<strong>in</strong>d PpATG8. In an<br />

Arabidopsis Atatg5 background AtATG8 accumulates because AtATG5 is required for<br />

AtATG8 degradation (Thompson et al., 2005).<br />

Therefore <strong>the</strong> reason that no <strong>in</strong>duction of a PpATG8 b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> is seen <strong>in</strong> <strong>the</strong><br />

tunicamyc<strong>in</strong> treated wild type plants could be that <strong>the</strong> tunicamyc<strong>in</strong> concentration was too<br />

low or that it does not diffuse freely through <strong>the</strong> cellophane. The experiment should be<br />

repeated with ano<strong>the</strong>r autophagy <strong>in</strong>duc<strong>in</strong>g treatment such as growth <strong>in</strong> <strong>the</strong> dark on<br />

BCD+AT media plates which has been shown to kill <strong>the</strong> ΔATG5 l<strong>in</strong>es but not affect <strong>the</strong><br />

wild type (Figure 39 and Figure 41) or by treatments with exogenous H 2 O 2 which has<br />

been shown to <strong>in</strong>duce autophagy (Xiong et al., 2007).<br />

61


Figure 44. (A) Immunoblot analysis with anti-AtATG8 antibodies at <strong>in</strong>dicated days post treatment.<br />

KO1=ΔATG5-1 and KO2= ΔATG5-2. (B) Phylogenetic relationships between <strong>the</strong> n<strong>in</strong>e Arabidopsis<br />

(At), <strong>the</strong> five <strong>Physcomitrella</strong> (Pp), and <strong>the</strong> Saccharomyces cerevisiae (Sc) ATG8 homologs.<br />

Pathogen treatments of <strong>the</strong> ΔATG5 l<strong>in</strong>es<br />

In Arabidopsis autophagy has been shown to play a role <strong>in</strong> defense aga<strong>in</strong>st necrotrophic<br />

fungi (Lai et al., 2011; Lenz et al., 2011). The ΔATG5 l<strong>in</strong>es were <strong>the</strong>refore also <strong>in</strong>fected<br />

with <strong>the</strong> pathogens listed <strong>in</strong> Table 4, and a visual screen showed that <strong>the</strong>y were more<br />

susceptible to all <strong>the</strong>se necrotrophic pathogens compared to wild type. The ΔATG5 l<strong>in</strong>es<br />

showed visible disease symptoms such as brown<strong>in</strong>g and cell maceration much earlier<br />

than <strong>the</strong> wild type (data not shown).<br />

An Alternaria spore count assay showed many more spores be<strong>in</strong>g produced <strong>in</strong> <strong>the</strong><br />

ΔATG5 l<strong>in</strong>es compared to wild type (Figure 29). However, it may well be that <strong>the</strong><br />

enhanced fungal growth on ΔATG5 is due to <strong>in</strong>direct effects of autophagy deficiency<br />

ra<strong>the</strong>r than a direct <strong>in</strong>volvement of autophagy <strong>in</strong> combat<strong>in</strong>g fungal <strong>in</strong>vasion. In pr<strong>in</strong>ciple,<br />

autophagy deficient cells have accumulated toxic compounds and lack energy and<br />

nutrients due to lack of <strong>the</strong> autophagic recycl<strong>in</strong>g system. It is <strong>the</strong>refore possible that a<br />

necrotrophic fungus would f<strong>in</strong>d better growth conditions <strong>in</strong> ΔATG5 than <strong>in</strong> wild type cell.<br />

Evans blue sta<strong>in</strong><strong>in</strong>g upon treatment with cell free culture filtrates of <strong>the</strong> HrpN secret<strong>in</strong>g P.<br />

carotovorum SCC1 and <strong>the</strong> HrpN negative stra<strong>in</strong> P. carotovorum SCC3193 showed that<br />

both stra<strong>in</strong>s <strong>in</strong>duced more cell death <strong>in</strong> ΔATG5 than <strong>in</strong> <strong>the</strong> wild type Figure 45.<br />

62


O.D. 600 nm/mg DW<br />

0.6<br />

0.4<br />

0.2<br />

Evans blue sta<strong>in</strong><strong>in</strong>g upon P. carotovorum culture<br />

filtrate treatment<br />

H2O<br />

P.c. 3193<br />

P.C. SCC1<br />

0<br />

WT ATG5‐1 ATG5‐2<br />

Figure 45. Evans blue sta<strong>in</strong><strong>in</strong>g of wild type and two <strong>in</strong>dependent ΔATG5 l<strong>in</strong>es 24 hours after spray<strong>in</strong>g<br />

with cell free culture filtrates of <strong>the</strong> HrpN negative stra<strong>in</strong> P. carotovorum 3193, <strong>the</strong> HrpN produc<strong>in</strong>g<br />

stra<strong>in</strong> P. carotovorum SCC1 or H 2 O. Error bars represent SEM of three <strong>in</strong>dependent biological<br />

replicates.<br />

Aga<strong>in</strong>, it is difficult to say if this <strong>in</strong>creased cell death is due to a direct or <strong>in</strong>direct effect<br />

of autophagy deficiency. An <strong>in</strong>terest<strong>in</strong>g observation from this experiment is that <strong>the</strong><br />

water treated ΔATG5 l<strong>in</strong>es do not have an elevated level of Evans blue sta<strong>in</strong><strong>in</strong>g even<br />

though <strong>the</strong>y have visibly dead cells after two weeks of growth when <strong>the</strong> experiment was<br />

performed. Apparently Evans blue only sta<strong>in</strong>s recently dead cells.<br />

Pathogen <strong>in</strong>fections were also tried on wild type and ΔATG5 l<strong>in</strong>es that had grown on full<br />

+ 50g glucose medium which rescues <strong>the</strong> phenotype of <strong>the</strong> ΔATG5 l<strong>in</strong>es (Figure 39). But<br />

<strong>the</strong> pathogen grew very differently due to <strong>the</strong> added sugar <strong>in</strong> <strong>the</strong> medium and apparently<br />

it also affected <strong>the</strong> Evans blue sta<strong>in</strong><strong>in</strong>g. Thus, <strong>the</strong>se experiments were <strong>in</strong>conclusive and<br />

<strong>the</strong> data omitted.<br />

Infection with a Sordariomycetes fungus<br />

A contam<strong>in</strong>ation of an unknown pathogen on a plate with all <strong>the</strong> KO l<strong>in</strong>es from <strong>the</strong> first<br />

round of transformation apparently only grew on <strong>the</strong> ΔATG5 l<strong>in</strong>es (Figure 46). The<br />

unknown pathogen was subsequently propagated on PDA medium and spores were<br />

isolated and re<strong>in</strong>fections performed several times by both spray<strong>in</strong>g a suspension of<br />

isolated spores and with small parts PDA with <strong>the</strong> organism grow<strong>in</strong>g on it. The same<br />

result was repeatedly found: <strong>the</strong> ΔATG5 l<strong>in</strong>es became <strong>in</strong>fected and <strong>the</strong> o<strong>the</strong>r l<strong>in</strong>es did<br />

apparently not.<br />

With <strong>the</strong> help of Henn<strong>in</strong>g Knudsen, a small part of <strong>the</strong> unknown pathogen’s DNA<br />

was sequenced and BLASTed aga<strong>in</strong>st a fungal DNA database. The best hit was to a<br />

unknown Ascomycetes, <strong>the</strong> second best hit was to Chaetosphaeria lateriphiala with 1140<br />

bp and 96% identity, and <strong>the</strong> third best hit was to Ophiocordyceps s<strong>in</strong>ensis with 93%<br />

identity for 650 bp. Both C. lateriphiala and O. s<strong>in</strong>ensis belong to <strong>the</strong> Sordariomycetes<br />

class which <strong>in</strong>cludes more than 600 genera with over 3000 species (Zhang et al., 2006).<br />

Thus, <strong>the</strong> unknown pathogen is most likely <strong>the</strong> anamorphic stage of a fungus belong<strong>in</strong>g<br />

to this class, which seems plausible s<strong>in</strong>ce <strong>the</strong> class exhibits a wide range of ecologies<br />

<strong>in</strong>clud<strong>in</strong>g pathogens and endophytes of plants, animal pathogens and mycoparasites<br />

(Zhang et al., 2006).<br />

63


Figure 46. KO l<strong>in</strong>es from <strong>the</strong> first round of transformations. (A) Grown for three weeks on BCD+AT<br />

and (B) <strong>in</strong>fected by spray<strong>in</strong>g a suspension of spores from <strong>the</strong> Sordariomycetes class fungus and grown<br />

for an additional two weeks. Note that <strong>the</strong> two pictures are not of <strong>the</strong> same plate. The experiment was<br />

repeated with similar result. Bars represent 1 cm.<br />

The growth of <strong>the</strong> Sordariomycetes fungus on <strong>the</strong> ΔATG5 l<strong>in</strong>es was not only due to <strong>the</strong>se<br />

l<strong>in</strong>es dy<strong>in</strong>g at this stage, s<strong>in</strong>ce control plants were still green and alive after six weeks<br />

(Figure 47A+B). A closer <strong>in</strong>spection under <strong>the</strong> microscope revealed that <strong>the</strong><br />

Sordariomycetes fungus actually did <strong>in</strong>fect <strong>the</strong> wild type (Figure 47C) (and all <strong>the</strong> o<strong>the</strong>r<br />

KO l<strong>in</strong>es from <strong>the</strong> first round of transformations, data not shown). But apparently <strong>the</strong><br />

<strong>in</strong>fection was suppressed so that <strong>the</strong> fungus did not produce conidia that were clearly<br />

visible on <strong>the</strong> ΔATG5 l<strong>in</strong>es (Figure 47B, D, E). It would be <strong>in</strong>terest<strong>in</strong>g to <strong>in</strong>vestigate<br />

whe<strong>the</strong>r autophagy plays an active role <strong>in</strong> combat<strong>in</strong>g this fungal pathogen, or if its<br />

susceptibility is more <strong>in</strong>directly due to <strong>the</strong> accumulation of toxic compounds and lack of<br />

nutrient recycl<strong>in</strong>g <strong>in</strong> <strong>the</strong> autophagy deficient mutants. One way to assess this could be to<br />

<strong>in</strong>vestigate if <strong>the</strong> pathogen <strong>in</strong>fection activates autophagy <strong>in</strong> <strong>the</strong> wild type, <strong>the</strong>reby<br />

<strong>in</strong>dicat<strong>in</strong>g a direct role of autophagy <strong>in</strong> combat<strong>in</strong>g this pathogen <strong>in</strong>truder.<br />

64


Figure 47. Infection with <strong>the</strong> Sordariomycetes fungus. (A) Wild type and ΔATG5-1 grown for five<br />

weeks on BCD+AT with no treatment. (B-D) Wild type and ΔATG5-1 grown for three weeks on<br />

BCD+AT and <strong>the</strong>n <strong>in</strong>fected with <strong>the</strong> Sordariomycetes fungus and grown for an additional two weeks.<br />

(E) Close up of <strong>the</strong> Sordariomycetes fungus hyphae with conidia. Scale bars: (A+B) 1 cm, (C+D) 0.5<br />

mm and (E) 100 µm.<br />

MPK phosporylation <strong>in</strong> ΔATG5 l<strong>in</strong>es<br />

We also checked if <strong>the</strong> phosphorylation of MPKs upon MAMP treatment with chit<strong>in</strong> was<br />

altered <strong>in</strong> <strong>the</strong> ΔATG5 l<strong>in</strong>es, but it was not dist<strong>in</strong>guishable from <strong>the</strong> wild type (data not<br />

shown). This f<strong>in</strong>d<strong>in</strong>g is <strong>in</strong> accordance with <strong>the</strong> f<strong>in</strong>d<strong>in</strong>gs that Arabidopsis Atatg5 shows<br />

normal MPK phosphorylation when treated with <strong>the</strong> bacterial MAMP flg22 (Lenz et al.,<br />

2011).<br />

The AtMEKK1 homologs ΔPpMEKK1A and ΔPpMEKK1B<br />

When treat<strong>in</strong>g <strong>moss</strong> colonies with <strong>the</strong> fungal MAMP chit<strong>in</strong> and subsequently perform<strong>in</strong>g<br />

a western blot with antibody aga<strong>in</strong>st <strong>the</strong> phosphorylated TxY motif of MPKs, we have<br />

found that two MPKs are phosphorylated upon chit<strong>in</strong> treatment. One of <strong>the</strong>m is<br />

PpMPK4A, as seen from a lack<strong>in</strong>g band <strong>in</strong> <strong>the</strong> ΔPpMPK4A l<strong>in</strong>es, and <strong>the</strong> o<strong>the</strong>r is most<br />

probably PpMPK4B judged from <strong>the</strong> size (see Figure 5, manuscript 1). This observation<br />

can be used to identify upstream components <strong>in</strong>volved <strong>in</strong> <strong>the</strong> signal<strong>in</strong>g cascade from <strong>the</strong><br />

perception of chit<strong>in</strong> to <strong>the</strong> activation of defense mechanisms. In Arabidopsis, AtMEKK1<br />

is required for <strong>the</strong> activation of AtMPK4 after treatment with <strong>the</strong> bacterial MAMP flg22<br />

(Suarez-Rodriguez et al., 2007). It is not known if AtMEKK1 functions upstream of<br />

AtMPK4 signal<strong>in</strong>g after chit<strong>in</strong> treatment <strong>in</strong> Arabidopsis, but s<strong>in</strong>ce treat<strong>in</strong>g Arabidopsis<br />

with chit<strong>in</strong> <strong>in</strong>duces <strong>the</strong> same AtMPK phosphorylation pattern as treat<strong>in</strong>g with flg22 or<br />

65


elf18 (Wan et al., 2004; Petutschnig et al., 2010; Roux et al., 2011; Bethke et al., 2012;<br />

Eschen-Lippold et al., 2012) it is thought that MEKK1 is also <strong>the</strong> upstream MP3K <strong>in</strong><br />

chit<strong>in</strong> <strong>in</strong>duced MPK signal<strong>in</strong>g. We identified two putative <strong>Physcomitrella</strong> orthologs of<br />

Arabidopsis AtMEKK1, PpMEKK1A and PpMEKK1B (Figure 8). In order to test if<br />

<strong>the</strong>se are <strong>the</strong> MP3K <strong>in</strong> <strong>the</strong> MAP k<strong>in</strong>ase signal<strong>in</strong>g cascade that phosphorylates PpMPK4A<br />

and probably PpMPK4B upon chit<strong>in</strong> treatment, we performed a western blot with <strong>the</strong><br />

antibody aga<strong>in</strong>st phosphorylated MPKs <strong>in</strong> <strong>the</strong> ΔPpMEKK1B and ΔPpMEKK1A l<strong>in</strong>es<br />

(Figure 48).<br />

Figure 48 Immunoblot analysis with anti-phospho-p44/42 MAPK antibodies three m<strong>in</strong>utes after<br />

spray<strong>in</strong>g with 1 mg/ml chit<strong>in</strong>. The numbers below each of <strong>the</strong> two KO l<strong>in</strong>es denote <strong>the</strong> number of <strong>the</strong><br />

transformant. Thus all four ΔMEKK1A l<strong>in</strong>es we identified are <strong>in</strong>cluded <strong>in</strong> this blot while ΔMEKK1B-<br />

3,7,8 have yet to be tested. Load<strong>in</strong>g controls show amido black sta<strong>in</strong>ed total prote<strong>in</strong>.<br />

As seen <strong>in</strong> <strong>the</strong> western blot, two bands are visible <strong>in</strong> all l<strong>in</strong>es tested. Although some of <strong>the</strong><br />

l<strong>in</strong>es (ΔPpMEKK1B-6 and ΔPpMEKK1A-3 and 4) appear to have a weaker band, all<br />

l<strong>in</strong>es still have two visible bands presumably correspond<strong>in</strong>g to phosphorylated<br />

PpMPK4A and B. That said, it may also be that <strong>the</strong> two MAP3Ks are partially redundant<br />

such that <strong>the</strong>ir s<strong>in</strong>gle knockouts will not fully abolish MPK phosphorylation <strong>in</strong> response<br />

to chit<strong>in</strong> treatment, or that nei<strong>the</strong>r PpMEKK1A nor PpMEKK1B are <strong>the</strong> upstream MP3K<br />

<strong>in</strong> <strong>the</strong> signal<strong>in</strong>g cascade lead<strong>in</strong>g to <strong>the</strong> phosphorylation of <strong>the</strong> two MPKs <strong>in</strong> <strong>the</strong> western<br />

blot (or at least not any of <strong>the</strong> two PpMEKK1 homologs alone). There are several<br />

examples of redundancy <strong>in</strong> MPK signal<strong>in</strong>g cascades (Rasmussen et al., 2012). Creat<strong>in</strong>g a<br />

double KO of <strong>the</strong> two PpMEKK1 homologs could be used <strong>in</strong> a similar assay to asses this<br />

question.<br />

66


Discussion<br />

This discussion highlights current and potential, future work on <strong>Physcomitrella</strong>.<br />

Despite <strong>the</strong> strik<strong>in</strong>g similarities between plant and animal <strong>in</strong>nate immune systems, <strong>the</strong>y<br />

are probably a consequence of convergent evolution s<strong>in</strong>ce multicellularity evolved<br />

<strong>in</strong>dependently <strong>in</strong> plants and animals (Meyerowitz, 2002). The use of many of <strong>the</strong> same<br />

build<strong>in</strong>g blocks and similarities <strong>in</strong> <strong>the</strong> overall design probably reflect <strong>in</strong>herent constra<strong>in</strong>ts<br />

on how an <strong>in</strong>nate immune system can be constructed (Ausubel, 2005). A better<br />

understand<strong>in</strong>g of <strong>the</strong> evolution of plant <strong>in</strong>nate immunity will help elucidate <strong>the</strong> orig<strong>in</strong> of<br />

its different components. <strong>Physcomitrella</strong> has proven to be a good model for this due to its<br />

evolutionary position and ability to be genetically manipulated.<br />

MPKs <strong>in</strong> <strong>Physcomitrella</strong><br />

The work presented <strong>in</strong> this <strong>the</strong>sis provides comprehensive evidence that <strong>the</strong> MAMP<br />

triggered immune system, with activation of a MPK signal<strong>in</strong>g cascade(s) upon MAMP<br />

recognition, is conserved between <strong>moss</strong>es and flower<strong>in</strong>g plants. In manuscript 1 it is<br />

described how at least one <strong>Physcomitrella</strong> MPK (PpMPK4A) is required for proper<br />

<strong>in</strong>nate immune responses. This is a primary example of a s<strong>in</strong>gle, non-redundant plant<br />

MPK essential for immunity without any o<strong>the</strong>r apparent phenotypes associated with <strong>the</strong><br />

correspond<strong>in</strong>g null-mutant. This f<strong>in</strong>d<strong>in</strong>g is important because researchers cont<strong>in</strong>ue to<br />

debate <strong>the</strong> immunity-related functions of numerous MPKs <strong>in</strong> vascular plants. For<br />

example, <strong>the</strong> contribution of s<strong>in</strong>gle MPK <strong>in</strong> immune responses of <strong>the</strong> higher plant<br />

Arabidopsis is unclear due to pleiotropic effects of specific knockout mutants <strong>in</strong>clud<strong>in</strong>g<br />

seedl<strong>in</strong>g lethality, <strong>in</strong>appropriate defense activation and dwarfism.<br />

ROS<br />

My f<strong>in</strong>d<strong>in</strong>gs (Figure 33) show that ΔPpMPK4A does not accumulate as much ROS as <strong>the</strong><br />

WT after chitosan treatment. This is <strong>in</strong>terest<strong>in</strong>g s<strong>in</strong>ce <strong>the</strong> MPK cascade AtMEKK1,<br />

AtMKK1/AtMKK2 and AtMPK4 has been reported to negatively regulate ROS<br />

production (Petersen et al., 2000; Nakagami et al., 2006; Gao et al., 2008). Although, it<br />

was recently discovered that <strong>the</strong> constitutive production of ROS <strong>in</strong> <strong>the</strong> mutants of this<br />

Arabidopsis MPK cascade is due to <strong>the</strong> activation of ETI through <strong>the</strong> AtSUMM2 R-<br />

prote<strong>in</strong>, <strong>the</strong> Atmpk4/Atsumm2 double mutant still havs constitutive ROS production<br />

(Zhang et al., 2012). Thus, caution is needed <strong>in</strong> <strong>in</strong>terpret<strong>in</strong>g <strong>the</strong> function of AtMPK4 from<br />

results obta<strong>in</strong>ed with <strong>the</strong> Atmpk4/Atsumm2 double mutant. In <strong>Physcomitrella</strong> <strong>the</strong> lack of<br />

constitutive ROS production <strong>in</strong> <strong>the</strong> ΔPpMPK4A shows that <strong>the</strong>re is no auto-activation of<br />

defense responses <strong>in</strong> this mutant (Figure 33). Thus, <strong>the</strong> results obta<strong>in</strong>ed with this KO<br />

probably reflect <strong>the</strong> direct effect of <strong>the</strong> miss<strong>in</strong>g <strong>the</strong> MPK. That ΔPpMPK4A does not<br />

accumulate as much ROS as WT after chitosan treatment po<strong>in</strong>ts to PpMPK4A as a<br />

positive regulator of defense. Thus, <strong>the</strong> signal<strong>in</strong>g cascade from <strong>the</strong> PRR is <strong>in</strong>terrupted <strong>in</strong><br />

<strong>the</strong> ΔPpMPK4A mutant and no defense responses are <strong>in</strong>itiated.<br />

67


As <strong>in</strong> Arabidopsis, we found that <strong>the</strong> same MPKs are activated by <strong>the</strong> application<br />

of exogenous H 2 O 2 as by MAMP treatment (Gao et al., 2008, Figure 5E manuscript 1).<br />

How H 2 O 2 activates MPKs is not known (Gao et al., 2008). It is possible that <strong>the</strong> ROS<br />

burst that occurs seconds after MAMP b<strong>in</strong>d<strong>in</strong>g to its cognate PRR is part of <strong>the</strong> l<strong>in</strong>k<br />

between PRR and MPK cascade activation. <strong>Physcomitrella</strong> should be a good model to<br />

study this s<strong>in</strong>ce <strong>the</strong> ROS burst has been shown to occur with<strong>in</strong> seconds after chitosan<br />

treatment (Lehtonen et al., 2012). Measurements of <strong>the</strong> ROS burst can be used to identify<br />

a <strong>Physcomitrella</strong> chit<strong>in</strong> receptor (PpCERK), s<strong>in</strong>ce a ΔPpCERK knockout should not be<br />

able to <strong>in</strong>duce ROS burst after chitosan treatment (Segonzac et al., 2011). In this<br />

connection, it seems likely that <strong>the</strong> ΔPpMPK4A mutant is fully capable of produc<strong>in</strong>g <strong>the</strong><br />

early ROS burst although it is <strong>in</strong>hibited <strong>in</strong> ROS accumulation as a result of defense<br />

responses downstream of PpMPK4A. The DAB sta<strong>in</strong><strong>in</strong>g shown <strong>in</strong> Figure 33 was<br />

performed 24 hours after chitosan treatment, and thus reflects long term ROS<br />

accumulation ra<strong>the</strong>r than <strong>the</strong> early ROS burst.<br />

Abiotic stress<br />

The lack of phosphorylated MPKs <strong>in</strong> <strong>Physcomitrella</strong> after abiotic stress is very surpris<strong>in</strong>g<br />

and should be fur<strong>the</strong>r tested. MPK activation <strong>in</strong> abiotic stresses like osmotic stress have<br />

been reported to be ubiquitous amongst eukaryotes (Kültz, 2001). A strong argument for<br />

<strong>the</strong> conservation of this function is that a MAP k<strong>in</strong>ase from pea Pisum sativum could<br />

rescue a knock out of <strong>the</strong> yeast MPK ScHOG1 <strong>in</strong>volved <strong>in</strong> <strong>the</strong> osmoregulatory pathway<br />

(Pöpp<strong>in</strong>g et al., 1996). Similarly, <strong>the</strong> mammalian MPK HsSAPK2a rescues ScHOG1-<br />

deficient yeast <strong>in</strong> hyper osmotic medium and restores <strong>the</strong> osmotolerance of mutant yeast<br />

to that of <strong>the</strong> wild-type (Han et al., 1994). However, to our knowledge <strong>the</strong>re are no<br />

reports of MPK activation under any abiotic stresses <strong>in</strong> non-flower<strong>in</strong>g plants (S<strong>in</strong>ha et al.,<br />

2011). Therefore, it could be <strong>in</strong>terest<strong>in</strong>g to test MPK activation under abiotic stress <strong>in</strong><br />

o<strong>the</strong>r <strong>moss</strong>es and <strong>in</strong> <strong>the</strong> spike <strong>moss</strong> Selag<strong>in</strong>ella moellendorffii, a nonseed vascular plant<br />

that is considered an extant evolutionary <strong>in</strong>termediate between bryophytes and flower<strong>in</strong>g<br />

plants (Banks et al., 2011). This would help elucidate if MPK <strong>in</strong>volvement <strong>in</strong> abiotic<br />

stress adaptation has been lost <strong>in</strong> <strong>Physcomitrella</strong> or has evolved later <strong>in</strong> plant evolution. If<br />

it evolved later, it would be a fasc<strong>in</strong>at<strong>in</strong>g example of convergent evolution, although not<br />

unique s<strong>in</strong>ce MPK signal<strong>in</strong>g <strong>in</strong> biotic stress also seems to have evolved <strong>in</strong>dependently <strong>in</strong><br />

<strong>the</strong> plant and animal l<strong>in</strong>eages (Ausubel, 2005).<br />

Sporophyte formation<br />

We have recently observed that <strong>the</strong> ΔPpMPK4A mutant might be <strong>in</strong>hibited <strong>in</strong> sporophyte<br />

formation (Figure 35). We are currently repeat<strong>in</strong>g this experiment and if <strong>the</strong> ΔPpMPK4A<br />

mutants keep produc<strong>in</strong>g far fewer sporophytes than <strong>the</strong> WT it will be <strong>in</strong>terest<strong>in</strong>g to study<br />

if <strong>the</strong> sporophytes that do form conta<strong>in</strong> stomata and if <strong>the</strong>y are fully functional. In<br />

Arabidopsis <strong>the</strong> MPK cascade consist<strong>in</strong>g of <strong>the</strong> MP3K YODA, AtMKK4/AtMKK5 and<br />

AtMPK3/AtMPK6 has been shown to regulate stomatal development (Wang et al., 2007).<br />

AtMPK3 and AtMPK6 share high similarity to PpMPK4A as seen <strong>in</strong> Figure 14, and<br />

AtMPK4 is strongly expressed <strong>in</strong> stomata (Petersen et al., 2000; Rodriguez et al., 2010).<br />

68


PpMPK4B<br />

Despite several transformation attempts with two different KO constructs we have not yet<br />

obta<strong>in</strong>ed a KO of PpMPK4B. Our failure to obta<strong>in</strong> a PpMPK4B KO may well be because<br />

PpMPK4B is essential. In pr<strong>in</strong>ciple, such potential loss-of-function lethality may be<br />

circumvented us<strong>in</strong>g <strong>in</strong>ducible gene silenc<strong>in</strong>g systems used <strong>in</strong> <strong>Physcomitrella</strong> (Bezanilla et<br />

al., 2005; Nakaoka et al., 2012). However, <strong>the</strong>re is a more powerful way to do this for<br />

most prote<strong>in</strong> k<strong>in</strong>ases us<strong>in</strong>g an approach that has been pioneered <strong>in</strong> plants <strong>in</strong> our lab with<br />

AtMPK4 (Brodersen et al., 2006). In this approach, conditional loss-of-function alleles<br />

are made accord<strong>in</strong>g to a chemical–genetic system for prote<strong>in</strong> k<strong>in</strong>ases (Bishop et al., 2000).<br />

In this system, a po<strong>in</strong>t mutation is <strong>in</strong>troduced that enlarges <strong>the</strong> k<strong>in</strong>ase ATP-b<strong>in</strong>d<strong>in</strong>g<br />

pocket to allow it to accommodate bulky ATP analogs such as N6-(benzyl)–ATP. Pocket<br />

enlargement has three consequences. 1 st , it does not impair <strong>the</strong> ability of <strong>the</strong> k<strong>in</strong>ase to<br />

phosphorylate its substrates with endogenous ATP. In <strong>Physcomitrella</strong> <strong>the</strong>n, a mutant<br />

pocket k<strong>in</strong>ase allele can be ‘knocked <strong>in</strong>’ to replace <strong>the</strong> wild type allele without loss-offunction.<br />

2 nd , pocket enlargement sensitizes <strong>the</strong> k<strong>in</strong>ase to <strong>in</strong>hibition by derivatives, such<br />

as C3–1’-naphtyl (NaPP1), of <strong>the</strong> Src tyros<strong>in</strong>e k<strong>in</strong>ase family <strong>in</strong>hibitor PP1. Due to <strong>the</strong><br />

simple morphology of <strong>Physcomitrella</strong> and its poikilohydric physiology <strong>the</strong>se cellpermeable<br />

molecules should easily penetrate all <strong>moss</strong> cells at any stage of development.<br />

S<strong>in</strong>ce NaPP1 is not an efficient <strong>in</strong>hibitor of wild type k<strong>in</strong>ases, treatments with it<br />

specifically <strong>in</strong>hibit <strong>the</strong> enlarged pocket k<strong>in</strong>ase. Transcript profil<strong>in</strong>g over a time course of<br />

<strong>in</strong>hibitor treatment can <strong>the</strong>refore identify target genes regulated by <strong>the</strong> specific k<strong>in</strong>ase<br />

pathway. 3rd, [γ-32P]-N6-(benzyl)–ATP can be used to specifically label <strong>the</strong> substrates<br />

of <strong>the</strong> enlarged pocket k<strong>in</strong>ase. Proteomics (2-D gel autoradiography and peptide<br />

identification by mass spectrometry) can <strong>the</strong>n be used to identify <strong>the</strong> specific substrates<br />

of <strong>the</strong> enlarged pocket k<strong>in</strong>ase (Specht and Shokat, 2002).<br />

This <strong>in</strong>ducible loss of function technique should be deployed to assess if PpMPK4B is<br />

<strong>in</strong>volved <strong>in</strong> defense, like its closest homolog PpMPK4A. This technique could also be<br />

used if any of future targeted PpMPKK or PpMP3K can not be knocked out us<strong>in</strong>g<br />

targeted gene replacement.<br />

As shown for Arabidopsis k<strong>in</strong>ases, functional fusions of green fluorescent prote<strong>in</strong> (GFP)<br />

to <strong>the</strong> k<strong>in</strong>ase can also be used to follow its spatial and temporal patterns of expression<br />

(Petersen et al., 2000; Anthony et al., 2004; Robatzek et al., 2006). The great advantage<br />

of us<strong>in</strong>g this technique <strong>in</strong> <strong>Physcomitrella</strong> is that <strong>the</strong> GFP-tag can be added to <strong>the</strong> native<br />

prote<strong>in</strong> under its own promoter at its native gene locus us<strong>in</strong>g <strong>the</strong> knock <strong>in</strong> technique.<br />

Potential pleiotropic effects of over expression or ectopic <strong>in</strong>tegration can <strong>the</strong>reby be<br />

avoided.<br />

We are currently select<strong>in</strong>g a knock <strong>in</strong> transformation of PpMPK4A:GFP.<br />

However, when we identify upstream MPKK and MP3K components <strong>in</strong> <strong>the</strong> signal<strong>in</strong>g<br />

cascade of PpMPK4B, <strong>the</strong>se should also be transformed with a GFP tag.<br />

69


PpRAR1 and R-prote<strong>in</strong>s<br />

Although we <strong>in</strong>itially saw that <strong>the</strong> ΔPpRAR1 mutants were less susceptible to B. c<strong>in</strong>erea<br />

(Figure 27), we consistently saw very high variation <strong>in</strong> <strong>the</strong> Evans blue sta<strong>in</strong><strong>in</strong>g for cell<br />

death <strong>in</strong> repeated experiments. In addition, Evans blue sta<strong>in</strong><strong>in</strong>g after B. c<strong>in</strong>erea <strong>in</strong>fection<br />

of <strong>the</strong> four R-prote<strong>in</strong> KOs was not different than WT. Thus we have not yet shown a<br />

phenotype of <strong>the</strong> ΔPpRAR1 and R-prote<strong>in</strong> mutants.<br />

None<strong>the</strong>less, <strong>the</strong>re are several approaches to establish if <strong>Physcomitrella</strong> has a functional<br />

R-prote<strong>in</strong> system. 1 st , we could use <strong>the</strong> already created ΔPpRAR1 and R-prote<strong>in</strong> mutants<br />

to screen for susceptibility to biotrophic pathogens. Wild type <strong>moss</strong> is resistant to those<br />

biotrophic pathogens that have been tested and that are recognized by R-gene systems <strong>in</strong><br />

higher plants. Thus, <strong>the</strong>re is a chance that <strong>the</strong> ΔPpRAR1 and R-prote<strong>in</strong> mutants do not<br />

recognize specific biotrophic pathogens such that <strong>the</strong>se mutants will be more susceptible<br />

to <strong>the</strong>m. We have only tested a few biotrophic pathogens and did not f<strong>in</strong>d any that were<br />

able to grow on any of <strong>the</strong> mutants. The problem with identify<strong>in</strong>g a biotrophic pathogen<br />

is that <strong>the</strong>y often are host specific and are often difficult to cultivate s<strong>in</strong>ce <strong>the</strong>y require a<br />

liv<strong>in</strong>g host if <strong>the</strong>y are obligate biotrophs).<br />

2 nd , some R-prote<strong>in</strong>s are auto-activated when over expressed <strong>in</strong> plant cells<br />

(Bendahmane et al., 2002; DeYoung and Innes, 2006). S<strong>in</strong>ce constitutive over expression<br />

of auto-activated R-prote<strong>in</strong>s could be lethal, <strong>the</strong>y could be stably transformed under a<br />

DEX <strong>in</strong>ducible promoter. However, <strong>in</strong>stead of over express<strong>in</strong>g native R-prote<strong>in</strong>, <strong>the</strong>re are<br />

several examples of so called ga<strong>in</strong> of function mutations result<strong>in</strong>g <strong>in</strong> auto-activat<strong>in</strong>g R-<br />

prote<strong>in</strong>s (DeYoung and Innes, 2006). Such a ga<strong>in</strong> of function mutation could be<br />

<strong>in</strong>troduced to one of <strong>the</strong> R-prote<strong>in</strong>s already knocked out and re-transformed <strong>in</strong>to its<br />

cognate KO <strong>moss</strong> l<strong>in</strong>e under an <strong>in</strong>ducible promoter. If <strong>the</strong> <strong>in</strong>duced expression of <strong>the</strong> autoactivated<br />

R-prote<strong>in</strong> results <strong>in</strong> HR, it would <strong>in</strong>dicate a functional ETI.<br />

A <strong>Physcomitrella</strong> auto-activated R-prote<strong>in</strong> could also be transiently expressed <strong>in</strong><br />

tobacco via Agrobacterium <strong>in</strong>filtration. If <strong>the</strong> <strong>Physcomitrella</strong> R-prote<strong>in</strong> causes HR <strong>in</strong><br />

tobacco it would be a good <strong>in</strong>dicator of conserved functions of ETI. There are several<br />

examples of such auto-activat<strong>in</strong>g R-prote<strong>in</strong>s from Arabidopsis, potato and tomato that<br />

<strong>in</strong>duce HR <strong>in</strong> tobacco when transiently expressed after Agrobacterium <strong>in</strong>filtration<br />

(Bendahmane et al., 2002; Tamel<strong>in</strong>g et al., 2006; Gabriëls et al., 2007; Swiderski et al.,<br />

2009).<br />

Ano<strong>the</strong>r approach could be to transiently over express auto-activat<strong>in</strong>g GFP tagged<br />

R-genes <strong>in</strong> <strong>moss</strong> delivered on gold particles us<strong>in</strong>g a gene gun (Šmídková et al., 2010).<br />

This could be done <strong>in</strong> WT and ΔPpRAR1 mutants to see if <strong>the</strong> R-prote<strong>in</strong> required<br />

PpRAR1 to function. We have successfully tested transient expression of 35S:GFP <strong>in</strong><br />

<strong>Physcomitrella</strong> transformed us<strong>in</strong>g a gene gun.<br />

3 rd One could transform a functional R-prote<strong>in</strong> from ano<strong>the</strong>r species <strong>in</strong>to<br />

<strong>Physcomitrella</strong>. For example, one could <strong>in</strong>troduce <strong>the</strong> tomato CC-NB-LRR SlNCR1 that<br />

does not require EDS1 or NDR1 (which are not present <strong>in</strong> <strong>Physcomitrella</strong>) (Gabriëls et<br />

al., 2007). In <strong>the</strong> <strong>Physcomitrella</strong> l<strong>in</strong>es express<strong>in</strong>g SlNCR1 one could <strong>the</strong>n transiently<br />

express an effector recognized by SlNCR1, for example Avr4 from Cladosporium fulvum.<br />

If <strong>the</strong> expression of Avr4 causes HR <strong>in</strong> <strong>Physcomitrella</strong>, it would <strong>in</strong>dicate a functional ETI.<br />

70


Autophagy<br />

Although <strong>the</strong> f<strong>in</strong>d<strong>in</strong>g that <strong>the</strong> ΔPpATG5 mutant is more susceptible to <strong>the</strong> necrotroph A.<br />

brassicicola (Figure 29) is concordant with similar f<strong>in</strong>d<strong>in</strong>gs <strong>in</strong> Arabidopsis (Lai et al.,<br />

2011; Lenz et al., 2011), <strong>the</strong> fragility of <strong>the</strong>se mutants makes it likely that <strong>the</strong> enhanced<br />

fungal growth is caused by pleiotropic effects of autophagy deficiency ra<strong>the</strong>r than <strong>the</strong><br />

direct <strong>in</strong>volvement of autophagy <strong>in</strong> <strong>the</strong> immune system.<br />

One way to study autophagy <strong>in</strong> <strong>Physcomitrella</strong> dur<strong>in</strong>g pathogen <strong>in</strong>fection could<br />

be to create a l<strong>in</strong>e express<strong>in</strong>g a PpATG8:GFP fusion prote<strong>in</strong>. AtATG8:GFP fusions have<br />

been very useful <strong>in</strong> study<strong>in</strong>g <strong>the</strong> formation of autophagosomes <strong>in</strong> Arabidopsis<br />

(Yoshimoto et al., 2004; Thompson et al., 2005; Xie and Klionsky, 2007; Slavikova et al.,<br />

2008). Such a construct could for example be used to characterize whe<strong>the</strong>r autophagy is<br />

<strong>in</strong>volved <strong>in</strong> defense aga<strong>in</strong>st <strong>the</strong> Sordariomycetes fungus by show<strong>in</strong>g if autophagosomes<br />

are formed <strong>in</strong> <strong>the</strong> wild type <strong>moss</strong>.<br />

A PpATG8:GFP express<strong>in</strong>g l<strong>in</strong>e would also be useful <strong>in</strong> determ<strong>in</strong><strong>in</strong>g if autophagy<br />

is <strong>in</strong>volved <strong>in</strong> <strong>the</strong> differentiation of chloronemata to caulonemata, as our f<strong>in</strong>d<strong>in</strong>gs <strong>in</strong><br />

Figure 38 and Figure 39 <strong>in</strong>dicates. It would also be a useful tool <strong>in</strong> try<strong>in</strong>g to understand<br />

<strong>the</strong> apparent salt stress tolerance of <strong>the</strong> ΔPpATG5 l<strong>in</strong>es (Figure 42) by show<strong>in</strong>g if salt<br />

stress <strong>in</strong>duces autophagy <strong>in</strong> <strong>the</strong> wild type.<br />

Figure 43 show that ΔPpATG5 l<strong>in</strong>es have shorter rhizoids than wild type. In Arabidopsis,<br />

autophagy has been shown to be constitutively active <strong>in</strong> root cells, and several autophagy<br />

mutants exhibitaltered root architecture and root hair formation (Yoshimoto et al., 2004;<br />

Yano et al., 2007; Slavikova et al., 2008). Although roots develop on <strong>the</strong> sporophytes <strong>in</strong><br />

vascular plants and rhizoids develop on <strong>the</strong> gametophyte of bryophytes, <strong>the</strong>re could be a<br />

connection. A recent study shows that a similar regulatory network controls <strong>the</strong><br />

development of rhizoids <strong>in</strong> <strong>moss</strong> gametophytes and root hairs on <strong>the</strong> roots of vascular<br />

plant sporophytes (Jones and Dolan, 2012).<br />

71


Materials and methods<br />

Identify<strong>in</strong>g Arabidopsis homologs/orthologs <strong>in</strong> <strong>Physcomitrella</strong><br />

Arabidopsis prote<strong>in</strong> and cDNA sequences were obta<strong>in</strong>ed from www.Arabidopsis.org<br />

from version TAIR10 conta<strong>in</strong><strong>in</strong>g 27,416 prote<strong>in</strong> cod<strong>in</strong>g genes. <strong>Physcomitrella</strong> prote<strong>in</strong>,<br />

genomic and cDNA sequences were obta<strong>in</strong>ed from www.cos<strong>moss</strong>.org, mostly from<br />

version 1.6 but some homologs with higher identity to Arabidopsis were chosen from<br />

older versions.<br />

Homologs were identified by BLASTp searches, with Arabidopsis prote<strong>in</strong><br />

sequences as queries aga<strong>in</strong>st all annotated gene models <strong>in</strong> <strong>Physcomitrella</strong><br />

www.cos<strong>moss</strong>.org us<strong>in</strong>g <strong>the</strong> BLOSUM62 matrix. Putative orthologs were identified by<br />

means of <strong>the</strong> reciprocal best hit method (Moreno-Hagelsieb and Latimer, 2008).<br />

When no such orthologous relationship could be established, a phylogenetic tree<br />

was constructed such that all <strong>the</strong> prote<strong>in</strong> sequences were aligned us<strong>in</strong>g ClustalW2 at<br />

http://www.ebi.ac.uk/Tools/msa/clustalw2/ with default sett<strong>in</strong>gs (Lark<strong>in</strong> et al., 2007). In<br />

an effort to narrow down <strong>the</strong> candidates, most phylogenetic trees were constructed us<strong>in</strong>g<br />

both full length prote<strong>in</strong> sequences, truncated versions with sequences of conserved core<br />

doma<strong>in</strong>s, as well as cDNA sequences. The phylogeny of <strong>the</strong> aligned sequences was<br />

calculated us<strong>in</strong>g <strong>the</strong> neighbor jo<strong>in</strong><strong>in</strong>g method with default sett<strong>in</strong>gs <strong>in</strong> ClustalW2. With<br />

such alignment <strong>in</strong>formation, a phylogenetic tree was <strong>the</strong>n drawn us<strong>in</strong>g TreeView V1.6.6.<br />

Scale bars <strong>in</strong> <strong>the</strong> phylogenetic trees represent 0.1 am<strong>in</strong>o acid substitutions per site.<br />

Moss growth conditions<br />

Standard conditions: 22ºC, 16 hours light and 8 hours dark with 55 µE·m -2·s -1 on<br />

BCD+AT media. L<strong>in</strong>es were rout<strong>in</strong>ely propagated by sett<strong>in</strong>g small pieces of protonema<br />

on a new plate every week to ma<strong>in</strong>ta<strong>in</strong> fresh tissue. For most assays, 16 small pieces of<br />

protonema were placed on a BCD+AT plate overlaid with cellophane and grown for 14<br />

days under standard conditions (Figure 49).<br />

Figure 49. Different <strong>Physcomitrella</strong> l<strong>in</strong>es grown on BCD+AT plates overlaid with cellophane for 14<br />

days.<br />

Sporophyte <strong>in</strong>duction: 17°C (optimal condition would be 15°C; Hohe et al., 2002) <strong>in</strong> 8<br />

hours light and 16 hours dark, and app. 20 µE·m -2·s -1 on BCD+AT media supplemented<br />

with 200 mg/l (v/w) D-Glucose.<br />

72


Media<br />

BCD: MgSO 4 . 7H 2 O 250 mg/l, KH 2 PO 4 250 mg/l, KNO 3 1010 mg/l, FeSO 4 . 7H 2 O 12.5<br />

mg/l, CaCl 2 . H 2 O 147 mg/l, trace elements*, pH 6.5 adjusted with KOH, agar 8g/l. After<br />

autoclav<strong>in</strong>g app. 30 ml poured <strong>in</strong> sterile Petri dishes (90 mm) and left to solidify. The<br />

plates were stored cold until use.<br />

*Trace elements: Made as a 1000x stock and kept at 4°C. Work<strong>in</strong>g solution <strong>in</strong> all<br />

medias are: 614 µg/l H 3 BO 3 , 389 µg MnCl 2·4H 2 O, 110 µg/l AlK(SO 4 ) 2·12H 2 O, 55 µg/l<br />

CoCl 2·6H 2 O, 55 µg/l CuSO 4·5H 2 O, 55 µg/l ZnSO 4·7H 2 O, 28 µg/l KBr, 28 µg/l KI, 28<br />

µg/l LiCl, 28 µg/l SnCl 2·2H 2 O, 25 µg/l Na 2 MoO 4·2H 2 O, 59 µg/µl NiCl 2·6H 2 O<br />

BCD+AT: As BCD but with ammonium tartrate 920 mg/l added before autoclav<strong>in</strong>g.<br />

PRMT: MgSO 4 . 7H 2 O 250 mg/l, KH 2 PO 4 250 mg/l, KNO 3 1010 mg/l, ammonium tartrate<br />

920 mg/l, FeSO 4 . 7H 2 O 12.5 mg/l, CaCl 2 . H 2 O 1470 mg/l, D-mannitol 80 mg/l, trace<br />

elements, pH 6.5, agar 6g/l was autoclaved.<br />

PRMB: MgSO 4 . 7H 2 O 250 mg/l, KH 2 PO 4 250 mg/l, KNO 3 1010 mg/l, ammonium<br />

tartrate 920 mg/l, FeSO 4 . 7H 2 O 12.5 mg/l, CaCl 2 . H 2 O 1470 mg/l, D-mannitol 60mg/l,<br />

trace elements, pH 6.5, agar 10g/l. Autoclaved and poured <strong>in</strong> sterile Petri dishes and left<br />

to solidify. The plates were stored cold until use.<br />

+G418:, 40 mg/l G418 (Sigma) was added to <strong>the</strong> liquid media when it had cooled below<br />

70°C after <strong>the</strong> autoclav<strong>in</strong>g when <strong>the</strong> media was used for selection.<br />

Full: The supplements are as described <strong>in</strong> Egener et al., (2002): MgSO 4 . 7H 2 O 250 mg/l,<br />

KH 2 PO 4 250 mg/l, KNO 3 1010 mg/l, ammonium tartrate 920 mg/l, FeSO 4 . 7H 2 O 12.5<br />

mg/l, CaCl 2 . H 2 O 1470 mg/l. Supplemented with: Myo-<strong>in</strong>ositol 4 mg/l, Chol<strong>in</strong>e chloride<br />

2.8 mg/l, Nicot<strong>in</strong>e acid 2.8 mg/l, Thiam<strong>in</strong>e-HCl 500 µg/l, Pyridox<strong>in</strong>e 250 µg/l, Biot<strong>in</strong> 10<br />

µg/l, P-am<strong>in</strong>obenzoic acid 250 µg/l, Ca-D-panto<strong>the</strong>nate 1.9 mg/l, Riboflav<strong>in</strong>e 15 µg/l,<br />

Aden<strong>in</strong>e 6.76 mg/l, Na-palmet<strong>in</strong>ic acid 3.84 mg/l and Peptone 250 mg/l<br />

+ 5g sug: As BCD+AT with 5 g/l D-Glucose<br />

+ 50g sug: As BCD+AT with 50 g/l D-Glucose<br />

Sporophyte <strong>in</strong>duction media: As BCD+AT with 200 mg/l D-Glucose added (Hohe et<br />

al., 2002)<br />

Clon<strong>in</strong>g of PpMPK4B<br />

The <strong>Physcomitrella</strong> transformation vector pMBL10a was k<strong>in</strong>dly provided by Dr. Andrew<br />

Cum<strong>in</strong>g. Flank<strong>in</strong>g regions of <strong>the</strong> Pp1s59_325V6.1 gene were PCR amplified and <strong>the</strong><br />

result<strong>in</strong>g fragments cloned <strong>in</strong>to pCR®-Blunt II-TOPO® (Invitrogen) and transformed<br />

<strong>in</strong>to competent E. coli cells. The primers used to amplify <strong>the</strong> LB and RB are listed <strong>in</strong><br />

Table 9. The LB fragment was cut from <strong>the</strong> TOPO vector us<strong>in</strong>g <strong>the</strong> restriction enzymes<br />

73


Not1 (Fermentas) and BamH1 (Fermentas), and <strong>the</strong> result<strong>in</strong>g 621 bp fragment was<br />

ligated us<strong>in</strong>g T4 fast ligase (Promega) <strong>in</strong>to pMBL10a which was also cut with Not1 and<br />

BamH1. This plasmid was transformed <strong>in</strong>to competent E. coli cells. The plasmid was<br />

extracted and correct <strong>in</strong>sertion of <strong>the</strong> LB verified. Subsequently, <strong>the</strong> RB fragment was<br />

cloned <strong>in</strong>to pMBL10a conta<strong>in</strong><strong>in</strong>g <strong>the</strong> LB. The RB fragment was cut from TOPO vector<br />

us<strong>in</strong>g <strong>the</strong> restriction enzymes Sph1 (Fermentas) and EcoRV (Fermentas) and <strong>the</strong> result<strong>in</strong>g<br />

580 bp fragment was ligated <strong>in</strong>to pMBL10a conta<strong>in</strong><strong>in</strong>g <strong>the</strong> first LB fragment also<br />

digested with Sph1 and EcoRV. This plasmid was transformed <strong>in</strong>to competent E. coli<br />

cells. The plasmid was extracted and correct <strong>in</strong>sertion of <strong>the</strong> LB and RB verified by<br />

sequenc<strong>in</strong>g with <strong>the</strong> primers listed <strong>in</strong> Table 7. The result<strong>in</strong>g vector was termed<br />

pMBL10a_MPK4B.<br />

Construct<strong>in</strong>g a USER compatible <strong>moss</strong> KO vector<br />

The <strong>moss</strong> transformation vector pMBL6 was modified for USER clon<strong>in</strong>g (Nour-Eld<strong>in</strong><br />

et al., 2010) by add<strong>in</strong>g a USER cassette on both sides of <strong>the</strong> selection gene. Thus, at <strong>the</strong><br />

left side of <strong>the</strong> nptII gene a USER cassette conta<strong>in</strong><strong>in</strong>g a Pac1 restriction site and two<br />

Nt.BbvCI (NEB) nick<strong>in</strong>g sites was <strong>in</strong>serted <strong>in</strong> pMBL6 cut with Sac1 and Kpn1.<br />

Subsequently, at <strong>the</strong> right side of <strong>the</strong> NptII selection cassette a USER cassette conta<strong>in</strong><strong>in</strong>g<br />

an AsiSi restriction site and two Nt.BbvCI nick<strong>in</strong>g sites was <strong>in</strong>serted by cutt<strong>in</strong>g <strong>the</strong> vector<br />

with Sal1 and EcoR1 restriction enzymes.<br />

Left side User cassette:<br />

5’-CGCTGAGGCTTAATTAAGGATCCTTAATTAAACCTCAGCGGTAC-3’<br />

3’-TCGAGCGACTCCGAATTAATTCCTAGGAATTAATTTGGAGTCGC-5’<br />

Right side User cassette:<br />

5’-TCGACGCTGAGGCGCGATCGCGGATCCGCGATCGCACCTCAGCG-3’<br />

3’GCGACTCCGCGCTAGCGCCTAGGCGCTAGCGTGGAGTCGCTTAA-5’<br />

The result<strong>in</strong>g vector was termed pMBLU and is depicted <strong>in</strong> Figure 50.<br />

74


Figure 50. A graphic representation of pMBLU with <strong>the</strong> two USER cassettes.<br />

Preparation of pMBLU<br />

For four fragment clon<strong>in</strong>g, 50 µg of pMBLU was digested with 10 µl Pac1 (NEB) and 10<br />

µl AsiSI (Invitrogen) <strong>in</strong> buffer 1xTango (Invitrogen) over night at 37°C. Then, 5 µl of<br />

Pac1 (NEB) and 5µl AsiSI (Invitrogen) were added along with 10 µl Nt.BbvCI (NEB) for<br />

an additional two hours of digestion at 37°C. (If only LB or RB is to be <strong>in</strong>serted <strong>in</strong> a<br />

regular two fragment clon<strong>in</strong>g, ei<strong>the</strong>r Pac1 or AsiSI are omitted). The digested plasmid<br />

was cleaned us<strong>in</strong>g a kit (High Pure PCR Product Purification Kit, Roche) accord<strong>in</strong>g to<br />

<strong>the</strong> manufacturer's specifications and diluted <strong>in</strong> H 2 O to100 ng/µl.<br />

USER clon<strong>in</strong>g<br />

1 µl digested pMBLU (100 ng) was mixed with 5 µl of LB PCR product, 5 µl of RB PCR<br />

product (both between 10-40 ng/µl), and 1 µl of USER enzyme mix (NEB). The mix was<br />

<strong>the</strong>n <strong>in</strong>cubated for 20 m<strong>in</strong> at 37°C and 20 m<strong>in</strong> at 25°C. The mix was added to competent<br />

E. coli cells and transformed by heat chock (42°C for 50 sec), and <strong>the</strong>n plated on LB<br />

plates conta<strong>in</strong><strong>in</strong>g ampicill<strong>in</strong>. Plasmid DNA was extracted from a selection of surviv<strong>in</strong>g<br />

colonies and digested with appropriate restriction enzymes to screen for correctly<br />

transformed plasmids. Correct <strong>in</strong>sertions of LB and RB were verified by sequenc<strong>in</strong>g <strong>the</strong><br />

plasmid DNA (see Table 7 for primers used).<br />

75


Sequenc<strong>in</strong>g <strong>in</strong>serts <strong>in</strong> pMBLU-GFP and pMBLU-mCherry<br />

sb410 GGTTATTGTCTCATGAGCGGA LB forward<br />

GFP CTCCTCGCCCTTGCTCACCAT LB reverse<br />

sb024 GGTATCAGAGCCATGAATAGGTC RB forward<br />

sb417 GACCATGATTACGCCAAGCTA RB reverse<br />

Sequenc<strong>in</strong>g <strong>in</strong>serts <strong>in</strong> pMBLU<br />

sb410 GGTTATTGTCTCATGAGCGGA LB forward<br />

sb027 GGCAATGGAATCCGAGGAGGT LB reverse<br />

sb024 GGTATCAGAGCCATGAATAGGTC RB forward<br />

sb059 TGTGAGCGGATAACAATTTCAC RB reverse<br />

Table 7. Primers used to sequence <strong>in</strong>sertion <strong>in</strong> pMBLU, pMBLU-GFP and pMBLU-mCherry. The<br />

table conta<strong>in</strong>s: Primer name, sequence and description<br />

Protoplast isolation<br />

This protocol is modified from (Schaefer et al., 1991). One week old protonemal <strong>moss</strong><br />

tissue grown on BCD+AT media over laid with cellophane was harvested (Figure 51) and<br />

digested <strong>in</strong> 10 ml 8% (v/w) D-mannitol with 0.5% (w/v) Driselase (Sigma), for 30 m<strong>in</strong><br />

with slow agitation on a rock<strong>in</strong>g table. The mixture was filtered through a mesh with 100<br />

μm × 100 μm pores. The protoplasts were sedimented by a 150 g, 3 m<strong>in</strong>ute centrifugation<br />

with low breaks (Standard centrifugation <strong>in</strong> this protocol), and <strong>the</strong> supernatant removed.<br />

The pellet was washed twice by resupend<strong>in</strong>g <strong>in</strong> 8% (v/w) D-mannitol and sedimented by<br />

centrifugation. The f<strong>in</strong>al pellet was resuspended <strong>in</strong> 10 ml CaPW (D-mannitol 80 g/l,<br />

CaCl 2 (2H 2 O) 7.35 g/l, distilled water), and protoplast concentration assessed us<strong>in</strong>g a<br />

hemocytometer (Figure 52).<br />

Figure 51. Harvest<strong>in</strong>g of one week old protonemal tissue grown on BCD+AT plate overlaid with<br />

cellophane.<br />

76


Figure 52. Protoplasts <strong>in</strong> a hemocytometer. One square is 250 µm X 250 µm.<br />

Transformation<br />

5x10 5 protoplasts were resuspended <strong>in</strong> 300 µl freshly made MMM solution (Filter<br />

sterilized mixture of D-mannitol 920mg, 150 µl MgCl 2 1M, 1 ml MES 1% (v/w) pH 5.6<br />

and distilled water 8.85ml) <strong>the</strong>n 30 µl (1 µg/µl) l<strong>in</strong>earized plasmid was added (See Table<br />

8 for restriction enzymes used to digest <strong>the</strong> different KO vectors). 300 µl freshly made<br />

PEGT solution (PEG 2 g (4000 MW, Sigma), 4.5 ml 8% (v/w) D-mannitol, 500 µl 1 M<br />

Ca(NO 3 ) 2 ) was added, and 5 m<strong>in</strong>utes later <strong>the</strong> tube was heat shocked <strong>in</strong> a 45ºC water<br />

bath for 5 m<strong>in</strong>utes. Samples were left at room temperature for 10 m<strong>in</strong>utes, <strong>the</strong>n 600 µl<br />

CaPW was gently mixed <strong>in</strong>to <strong>the</strong> solution followed after 5 m<strong>in</strong>utes by ano<strong>the</strong>r 1 ml, 2ml,<br />

3ml, and 4ml added each after an additional 5 m<strong>in</strong>utes. Protoplasts were <strong>the</strong>n centrifuged,<br />

<strong>the</strong> supernatant removed, and <strong>the</strong> pellet resuspended <strong>in</strong> 1 ml sterile filtered 8% (v/w) D-<br />

mannitol. 7 ml molten PRMT (45°C) was added and <strong>the</strong> mixture was spread on four<br />

PRMB plates with cellophane, 2 ml per Petri dish. The protoplasts were left to recover<br />

for one week under standard conditions. Then, <strong>the</strong>y were transferred onto PRMB with<br />

G418 (Sigma) (40 µg/l) for two weeks of selection, <strong>the</strong>n two weeks without selection,<br />

<strong>the</strong>n followed by ano<strong>the</strong>r two weeks on selection. Colonies surviv<strong>in</strong>g <strong>the</strong>se two rounds of<br />

selection were considered stable transformants.<br />

Gene Plasmid name Enzymes Fragments<br />

Plasmid<br />

size<br />

Phypa_457568 pMBLU-R1 Sac1 Sma1 4093-2246 6339<br />

Phypa_430134 pMBLU-R2 Sac1 Sma1 4020-2246 6266<br />

Phypa_452798 pMBLU-R3 Nde1 4369-2552 6921<br />

Phypa_426777 pMBLU-R4 Stu1 Snab1 4215-2467 6682<br />

Phypa_452597 pMBLU-ATG5 Sac1 4420-2335 6755<br />

Phypa_461267 pMBLU-RAR1 Sac1 Sma1 4736-2246 6982<br />

Phypa_446407 pMBL10a-MPK4A Not1 Xba1 3280-2383 5663<br />

Phypa_445237 pMBLU-MEKK1A Sac1 3731-2423 6154<br />

Phypa_429991 pMBLU-MEKK1B Sma1 4168-2376 6544<br />

Phypa_430122 pMBLU-MKK1A EcorV Xba1 4105-2258 6363<br />

Phypa_451420 pMBLU-ACD11 Sac1 Sma1 3920-2246 6166<br />

Phypa_435424 pMBLU-MPK4B2 EcoR1 6431 6431<br />

77


Phypa_432218 pMBLU-CERK1A Sac1 Sma1 4390-2246 6636<br />

Phypa_458799 pMBLU-CERK1B Sma1 6424 6424<br />

Phypa_446407 pMBLU-GFP-MPK4A Sma1 7511 7511<br />

Phypa_453593 pMBLU-SGT1 Nhe1 Sma1 4442-2311 6753<br />

Phypa_435424 pMBL10a-Mpk4B1 Not1 Nde1 3191-2331 5522<br />

Phypa_435424 pMBLU-MPK4B2 EcoR1 6431 6431<br />

Table 8. <strong>Physcomitrella</strong> transformation plasmids. The table conta<strong>in</strong>s <strong>the</strong> Cos<strong>moss</strong> ID for <strong>the</strong> targeted<br />

gene, <strong>the</strong> plasmid name, <strong>the</strong> enzymes used to l<strong>in</strong>earize it, <strong>the</strong> fragment sizes <strong>in</strong> bp obta<strong>in</strong>ed by <strong>the</strong><br />

digestion, and <strong>the</strong> uncut plasmid size. The plasmids used to create confirmed KOs are <strong>in</strong> black. The<br />

blue color plasmids have not yet been confirmed as KOs or KI, and <strong>the</strong> red plasmids have failed to<br />

produce KOs despite several attempts.<br />

Genotyp<strong>in</strong>g<br />

Initial screen<strong>in</strong>g of KO colonies was done with <strong>the</strong> Extract-N-Amp Plant PCR Kit<br />

(Sigma-Aldrich). DNA extraction and PCR amplification was accord<strong>in</strong>g to <strong>the</strong><br />

manufacturer's specifications. If WT PCR failed but a control PCR worked, <strong>the</strong> colony<br />

was propagated and new DNA extracted us<strong>in</strong>g a modified Edwards extraction protocol<br />

with an additional Phenol:Chloroform:Isoamyl Alcohol 25:24:1 purification step<br />

(Edwards et al., 1991). This DNA was <strong>the</strong>n used as template for PCRs to confirm correct<br />

left boarder and right boarder <strong>in</strong>tegration as well as whole gene amplification. Primers<br />

used for genotyp<strong>in</strong>g are <strong>in</strong> Table 9.<br />

MPK4A Phypa_446407 Pp1s149_39V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb269 TGAAAACGTCGTTGCCATTA P4 sb270 GGTCCGTATCCATCAACTCG 290 WT gDNA<br />

P1 sb94 TTCGAATCCTAAATTTGAAAACAA P2 sb103 TTTTCTCTCCTTCGTTCGTCA 1223 WT LB<br />

P5 sb200 TGATTCGTTTCCTTGCAACA P6 sb54 ATAAAAGAGGGGTGTCTGCCTGGTAGTTC 1859 WT RB<br />

P1 sb94 TTCGAATCCTAAATTTGAAAACAA P7 sb27 GGCAATGGAATCCGAGGAGGT 880 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb54 ATAAAAGAGGGGTGTCTGCCTGGTAGTTC 1419 KO RB<br />

P1 sb94 TTCGAATCCTAAATTTGAAAACAA P6 sb54 ATAAAAGAGGGGTGTCTGCCTGGTAGTTC 5646/4169 All WT/KO<br />

cDNAF sb228 GGTACAAGCCACCACTTCGT cDNAR sb270 GGTCCGTATCCATCAACTCG 270 WT cDNA<br />

LBFF sb47,2 GCGGCCGCGGGTCTTTAGATTTGCCATGT LBFR sb64 GGATCCGTGTATTCCCGCGATTCTAGGT 663 Clon<strong>in</strong>g LB<br />

RBFF sb48 GATATCAAGGCCTACTCTCAATTCTTTAG RBFR sb66 TAGTACAGTAGGCAATGGTAGAATTGTT 888 Clon<strong>in</strong>g RB<br />

RAR1 Phypa_461267 Pp1s505_15V5.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb252 AGGCAGCTCCTAGTCCTAACG P4 sb253 CTGCAAAACGAAAACAACCAT 263 WT gDNA<br />

P1 sb255 GTCTACACATGTAATGTTGAAGGATAA P2 sb245 CATACCCAGGCAGGCTTAAA 1530 WT LB<br />

P5 sb215 CTCAACAACCAACAGCTGGA P6 sb254 TCTTGATTCAACCGCTGACA 2533 WT RB<br />

P1 sb255 GTCTACACATGTAATGTTGAAGGATAA P7 sb92 GGTACCGCTGAGGTTTAAT 1555 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb254 TCTTGATTCAACCGCTGACA 1602 KO RB<br />

P1 sb255 GTCTACACATGTAATGTTGAAGGATAA P6 sb254 TCTTGATTCAACCGCTGACA 4528/4772 All WT/KO<br />

cDNAF sb252 AGGCAGCTCCTAGTCCTAACG cDNAR sb216 GGATGGCACAGCCTTCTTTA 222 WT cDNA<br />

LBFF sb195 GGCTTAAUGGTTTGAACGGAGAACGAAA LBFR sb196 GGTTTAAUGGAACAAACCGAAGCAATGT 1250 Clon<strong>in</strong>g LB<br />

RBFF sb164 GGCGCGAUCGTTGGTGCGAGTATTAATGTAAG RBFR sb165 GGTGCGAUCCATTTAGACTACGACACACTTGG 1499 Clon<strong>in</strong>g RB<br />

Y2H AFRAR1F GGCTTAAUATGACGACGACGACAGGAAGAC Y2H AFRAR1R GGTTTAAUTCACGCCGCATCCTGTGGATTG 779 Clon<strong>in</strong>g CDS<br />

ATG5 Phypa_452597 Pp1s227_54V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb223 TTTTCGTGCTCAGACATTGC P4 sb224 GTGCAGCATCAAAGAGGTCA 285 WT gDNA<br />

P1 sb221 ATTTACCACCTTCATGGTGAGG P2 sb115 TCTCGGGCTCCAACAACGGT 3326 WT LB<br />

P5 sb223 TTTTCGTGCTCAGACATTGC P6 sb230 GTCCTCTCCCTTGTGCTGAG 2627 WT RB<br />

P1 sb221 ATTTACCACCTTCATGGTGAGG P7 sb92 GGTACCGCTGAGGTTTAAT 1647 KO LB<br />

78


P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb230 GTCCTCTCCCTTGTGCTGAG 1693 KO RB<br />

P1 sb233 TCGGAGAGTGGGAAAGAAGA P6 sb230 GTCCTCTCCCTTGTGCTGAG 7448/4954 All WT/KO<br />

cDNAF sb114 TGGTCGGGAGCCGTGCCAAT cDNAR sb115 TCTCGGGCTCCAACAACGGT 124 WT cDNA<br />

LBFF sb116 GGCTTAAUTCATAACAAATTGCCAGGAGTG LBFR sb117 GGTTTAAUATGCGCACAATTGAATTAACAG 1264 Clon<strong>in</strong>g LB<br />

RBFF sb118 GGCGCGAUTCCCTTTCGTCATCTGTGTATG RBFR sb119 GGTGCGAUTACGCACTTTGTGAGATTTGCT 1262 Clon<strong>in</strong>g RB<br />

R1 Phypa_457568 Pp1s327_15V5.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb240 TCGACAAACCTCAAGGAGCTA P4 sb241 TGATAGAAGTGCGCCTCAGAT 154 WT gDNA<br />

P1 sb246 CCATTACAGGTCTTATGTCCCATT P2 sb175 AACCCTAACTTGTTCGTATTCTCG 1462 WT LB<br />

P5 sb150 TGAGAATGAGGTCGTGTAATGC P6 sb256 CTAGAAGGCGGGATAATGTCAC 2086 WT RB<br />

P1 sb246 CCATTACAGGTCTTATGTCCCATT P7 sb27 GGCAATGGAATCCGAGGAGGT 1244 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb256 CTAGAAGGCGGGATAATGTCAC 1910 KO RB<br />

P1 sb246 CCATTACAGGTCTTATGTCCCATT P6 sb256 CTAGAAGGCGGGATAATGTCAC 5747/5039 All WT/KO<br />

LBFF sb193 GGCTTAAUGGGTTCGAAGGCACTGTTTA LBFR sb194 GGTTTAAUTGCCAGTCTGGTGATCTGAG 945 Clon<strong>in</strong>g LB<br />

RBFF sb124 GGCGCGAUTTCAAAAAGTGAGCCAACAAGA RBFR sb125 GGTGCGAUGTAACCTCCATCAAATGTGTGC 1168 Clon<strong>in</strong>g RB<br />

R2 Phypa_430134 Pp1s32_318V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb126 TGATGCAGCCGATCTCAA P4 sb127 CACCCGGTCGATGATTTT 907 WT gDNA<br />

P1 sb219 TCGGGACCTCAATAACTGCTAT P2 sb236 TTCTCCATCACCTCCGAATC 1570 WT LB<br />

P5 sb226 CTGAGCCGTTTCCTCTTGTC P6 sb220 AGGCATTTGCAATGAACCTCAA 2199 WT RB<br />

P1 sb219 TCGGGACCTCAATAACTGCTAT P7 sb27 GGCAATGGAATCCGAGGAGGT 1020 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb220 AGGCATTTGCAATGAACCTCAA 1860 KO RB<br />

P1 sb219 TCGGGACCTCAATAACTGCTAT P6 sb220 AGGCATTTGCAATGAACCTCAA 8906/4624 All WT/KO<br />

LBFF sb128 GGCTTAAUTACGTCGTCATTGGAAGTTTTG LBFR sb147 GGTTTAAUATATTTCGGTCGTCCATAGATTGT 724 Clon<strong>in</strong>g LB<br />

RBFF sb130 GGCGCGAUAGTTATTGTTGTTGTGCCATCG RBFR sb131 GGTGCGAUAACCCTCCATAGCTGCAAATTA 1314 Clon<strong>in</strong>g RB<br />

R3 Phypa_452798 Pp1s231_60V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb132 CCTTGGTAGCGTCGCAAT P4 sb133 TTCACGCATGCTCCTCTG 313 WT gDNA<br />

P1 sb152 GCAACTGGATGTTCAGTCAGAG P2 sb153 GAGCTCTCGGTGGAATAGATTG 1552 WT LB<br />

P5 sb237 TGGAATCCTTACCCTGTTGC P6 sb238 ATGACCAAGTTTGCGGTTTC 1744 WT RB<br />

P1 sb152 GCAACTGGATGTTCAGTCAGAG P7 sb92 GGTACCGCTGAGGTTTAAT 1368 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb238 ATGACCAAGTTTGCGGTTTC 1568 KO RB<br />

P1 sb152 GCAACTGGATGTTCAGTCAGAG P6 sb238 ATGACCAAGTTTGCGGTTTC 6696/4536 All WT/KO<br />

LBFF sb134 GGCTTAAUAAAACATCGGTGCGAGTAGATT LBFR sb135 GGTTTAAUGAGATTGCGACATGAAAGTGAG 1256 Clon<strong>in</strong>g LB<br />

RBFF sb198 GGCGCGAUTGCTGGCTGAATTGAACTTG RBFR sb199 GGTGCGAUTCCTTCCGGACATCATCTTC 1435 Clon<strong>in</strong>g RB<br />

R4 Phypa_426777 Pp1s17_234V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb138 GGTTCAAAGAGGAAATATGA P4 sb139 TCACACCATTTTTGCATAGC 312 WT gDNA<br />

P1 sb156 ACAATCTGGGCTTATTCTGCAT P2 sb157 AACAATCATCTTGGCTTCCATC 2025 WT LB<br />

P5 sb158 CATTGACATCATTGGAAGTAGCA P6 sb225 GTCAACAAGTTGGTCACGTGTTAT 1372 WT RB<br />

P1 sb156 ACAATCTGGGCTTATTCTGCAT P7 sb92 GGTACCGCTGAGGTTTAAT 1608 KO LB<br />

P8 sb24 GGTATCAGAGCCATGAATAGGTC P6 sb225 GTCAACAAGTTGGTCACGTGTTAT 1383 KO RB<br />

P1 sb156 ACAATCTGGGCTTATTCTGCAT P6 sb225 GTCAACAAGTTGGTCACGTGTTAT 9537/4999 All WT/KO<br />

LBFF sb140 GGCTTAAUGCCAAGATTTGCTCTTCTGAGT LBFR sb141 GGTTTAAUTTTCATCAAGGTTTCACATTGC 1238 Clon<strong>in</strong>g LB<br />

RBFF sb142 GGCGCGAUGCACCTGCCAACAAGAATTTAT RBFR sb143 GGTGCGAUTATGTGTTCCGCAAGTTAGACG 1208 Clon<strong>in</strong>g RB<br />

Cloned but not transformed<br />

MPK4B(1) Phypa_435424 Pp1s59_325V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb368 AATGTCGCCACTTCTGCTCT P4 sb369 GGCTGCTCAAAATCGAACTC 134 WT gDNA<br />

cDNAF sb267 GGCGAGTACACGCAGTACAA cDNAR sb268 ATTCACAGCGGAACACACAA 117 WT cDNA<br />

LBFF sb043 CCTCACTCTTTTCCCTCTACATAACACGAG LBFR sb044 GCGGCCGCTGGCTGCTAAACAGAATTGAAA 845 Clon<strong>in</strong>g LB<br />

RBFF sb045 GTAAAGAGAGCAGTACAAGGAGTCGTG RBFR sb046 GATATCCGACAGTAAATGACCAGGATCTC 876 Clon<strong>in</strong>g RB<br />

MPK4B(2) Phypa_435424 Pp1s59_325V6.1<br />

LBFF sb364 GGCTTAAUTGGCATCAAGGATTGATAGAAGT LBFR sb365 GGTTTAAUAATCCAGATATGTTCACCACCAC 1064 Clon<strong>in</strong>g LB<br />

79


RBFF sb366 GGCGCGAUGAGAGCAGTACAAGGAGTCGTGT RBFR sb367 GGTGCGAUTCCAAAACCTCTATCCACGTAAA 1137 Clon<strong>in</strong>g RB<br />

MPK4A<br />

-GFP Phypa_446407 Pp1s149_39V6.1<br />

LBFF sb412 GGCTTAAUATCGAGTGATATTTGACGATGCTA LBFR sb413 GGTTTAAUTTACTGCACCATATCATCAGGAC 1179 Clon<strong>in</strong>g LB<br />

RBFF sb414 GGCGCGAUAGCTTCACACTTCATCTCAAGTTGTA RBFR sb415 GGTGCGAUCAGTAGGCAATGGTAGAATTGTTG 904 Clon<strong>in</strong>g RB<br />

SGT1 Phypa_453593 Pp1s244_58V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb213 TGCAACCTCTAGACCTGCTG P4 sb214 CAACGATGGCACAACTTCAG 106 WT gDNA<br />

LBFF sb166 GGCTTAAUAAAGAAGGCTAGCACTCAAAGGTA LBFR sb197 GGTTTAAUGCCAAAGCTTCCGTGTAAAG 1463 Clon<strong>in</strong>g LB<br />

RBFF sb206 GGCGCGAUTCAAGCTCATGAACCACTGC RBFR sb207 GGTGCGAUCCGACAGCCATAATCCTCAT 1017 Clon<strong>in</strong>g RB<br />

GGTTTAAUCTAAATTTCCCACTTTTTCATATCC<br />

Y2H AFSGT1F GGCTTAAUATGGCAGAGGATATTGAGAAGCG Y2H AFSGT1R ATTCCTG 1143 Clon<strong>in</strong>g CDS<br />

CERK1A Phypa_432218 Pp1s41_280V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb376 CAAGCATAGGTTGCTCGTCA P4 sb377 AGTCATAAGGGCCACCACTG 340 WT gDNA<br />

cDNAF sb336 ATCCGCCAACATACTCCTTG cDNAR sb337 AACGACTCCGAAAGCGTAGA 202 WT cDNA<br />

LBFF sb372 GGCTTAAUGGTGGTTTTGAAATTACGCATTA LBFR sb373 GGTTTAAUATCGTTCAGGGAAGTTTTATGGT 1290 Clon<strong>in</strong>g LB<br />

RBFF MW01 GGCGCGAUCATGTGCAGTAAATCTGTTTGTGA RBFR MW02 GGTGCGAUCTCATCCTCAAGTAAATGTGCAAC 1116 Clon<strong>in</strong>g RB<br />

CERK1B Phypa_458799 Pp1s364_10V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb338 CATGAGCACACCAAACCAAC P4 sb339 CAAACCAAAATCTGCCACCT 213 WT gDNA<br />

cDNAF sb338 CATGAGCACACCAAACCAAC cDNAR sb339 CAAACCAAAATCTGCCACCT 99 WT cDNA<br />

LBFF sb378 GGCTTAAUAGTCTGCATCTTTTGTTCTCCAG LBFR sb379 GGTTTAAUTCCAGGAATGAAGGAACTTAACA 806 Clon<strong>in</strong>g LB<br />

RBFF sb380 GGCGCGAUTTCGAGATTTGGAACACTTCATT RBFR sb381 GGTGCGAUTCTCCAAATCCAAACTCCTACAA 1388 Clon<strong>in</strong>g RB<br />

MEKK1A Phypa_445237 Pp1s136_5V6.3<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb348 GCAGATGAAGGAAAGCTTGG P4 sb349 GCCGGTAAGAATCTGCTCTG 234 WT gDNA<br />

cDNAF sb348 GCAGATGAAGGAAAGCTTGG cDNAR sb349 GCCGGTAAGAATCTGCTCTG 234 WT cDNA<br />

LBFF sb384 GGCTTAAUTTTTTGTTTGGAGTGCAAGATTT LBFR sb385 GGTTTAAUCCAACGTAGTATGCCTATTCTGC 996 Clon<strong>in</strong>g LB<br />

RBFF sb386 GGCGCGAUTGGGATGCAGAAGACGTATAGTT RBFR sb387 GGTGCGAUTATTCCTGAAACCTGGCTTAACA 914 Clon<strong>in</strong>g RB<br />

MEKK1B Phypa_429991 Pp1s31_252V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb370 GGATGGCTCCAGAAGTGGTA P4 sb371 CCGGAATAAGAGGACCTTCC 171 WT gDNA<br />

cDNAF sb370 GGATGGCTCCAGAAGTGGTA cDNAR sb371 CCGGAATAAGAGGACCTTCC 171 WT cDNA<br />

LBFF sb392 GGCTTAAUCAGTACATTAATAACCCCACCACA LBFR sb393 GGTTTAAUTTCAAGATCTCCAAGAACCACATA 1111 Clon<strong>in</strong>g LB<br />

RBFF sb394 GGCGCGAUTGCATTGGTGTTTATTTGGTAATC RBFR sb395 GGTGCGAUTTCTTTTTAAAACCGTCCACTCTC 1203 Clon<strong>in</strong>g RB<br />

MKK1A Phypa_430122 Pp1s32_219V6.1<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb388 GCAAGCAAGTTTTGCTAGGG P4 sb389 ATCGCTGTCGAATCCATAGG 233 WT gDNA<br />

cDNAF sb346 GCTTAGGGTTGACGCTCTTG cDNAR sb389 ATCGCTGTCGAATCCATAGG 196 WT cDNA<br />

LBFF si01 GGCTTAAUTCCTCCTACTCGATATCCTTGTTC LBFR si02 GGTTTAAUACAACAAGCCTTTACAAACACAAA 912 Clon<strong>in</strong>g LB<br />

RBFF si03 GGCGCGAUTTATGAAGGGCTTGTGTCTTGATA RBFR si04 GGTGCGAUATTGGTGGTAATTCTTGTGGAAGT 1221 Clon<strong>in</strong>g RB<br />

ACD11 Phypa_451420 Pp1s211_32V6.2<br />

Type Name Forward primer 5' - 3' Type Name Reverse primer 5' - 3'<br />

Product<br />

size<br />

Description<br />

P3 sb340 CTCATCTCGCCTCTCTTTGG P4 sb341 TCGGTGACAAGAATGTGCTC 429 WT gDNA<br />

cDNAF sb340 CTCATCTCGCCTCTCTTTGG cDNAR sb361 GGTTTAAUTTCCAAGGCAAACGAAATAGTTA 242 WT cDNA<br />

LBFF sb360 GGCTTAAUGTCTTGCAAAGTTGACACGTTTT LBFR sb361 GGTTTAAUTTCCAAGGCAAACGAAATAGTTA 1119 Clon<strong>in</strong>g LB<br />

RBFF sb362 GGCGCGAUGGATAGCGACACTGAACTCACTT RBFR sb363 GGTGCGAUTGAGTGGTATGACCCCAGATATT 817 Clon<strong>in</strong>g RB<br />

Table 9. Primers used to clone flank<strong>in</strong>g regions of target genes and to genotype KOs of <strong>the</strong>se genes.<br />

Abbreviations: P1-P8 refers to <strong>the</strong> primer type illustrated <strong>in</strong> Figure 16. cDNAF – cDNA Forward,<br />

cDNAR – cDNA Reverse, LBFF – Left Boarder Flank<strong>in</strong>g Forward, LBFR – Left Boarder Flank<strong>in</strong>g<br />

Reverse, RBFF – Right Boarder Flank<strong>in</strong>g Forward, RBFR – Right Boarder Flank<strong>in</strong>g Reverse and<br />

Y2H – Yeast two Hybrid.<br />

80


Sou<strong>the</strong>rn blot<br />

DNA was extracted from four plates of approximately one week old protonema from<br />

each genotype. 8 µg were digested over night with <strong>the</strong> restriction enzymes Pst1 (NEB)<br />

and BamH1 (Fermentas) and/or EcoRV (Fermentas). The digested DNA was run on an<br />

ethidium bromide sta<strong>in</strong>ed 1% agarose gel (Figure 53).<br />

81


Figure 53. Ethidium bromide sta<strong>in</strong>ed agarose gels with genomic DNA from <strong>the</strong> <strong>in</strong>dicated genotypes<br />

digested with <strong>the</strong> <strong>in</strong>dicated restriction enzymes used for Sou<strong>the</strong>rn blot analysis. (A) <strong>the</strong><br />

correspond<strong>in</strong>g blot is shown <strong>in</strong> Figure 21, (B) <strong>the</strong> correspond<strong>in</strong>g blot is shown <strong>in</strong> Figure 22 and (C)<br />

<strong>the</strong> correspond<strong>in</strong>g blot is shown <strong>in</strong> Figure 23.<br />

82


The gel with <strong>the</strong> digested DNA was denatured for 30 m<strong>in</strong> <strong>in</strong> denatur<strong>in</strong>g solution (NaOH<br />

0.5M, NaCl 1.5 M), <strong>the</strong>n neutralized <strong>in</strong> NaCl 1.5 M, Tris-HCL (pH 7.4) 0.5M) followed<br />

by wash<strong>in</strong>g for15 m<strong>in</strong> <strong>in</strong> 20x SCC (NaCl 3M, Na 3 -citrate 0.3 M). DNA was <strong>the</strong>n<br />

transferred to nylon membranes (Amarsham, Hybond XL) over night us<strong>in</strong>g capillary<br />

effect <strong>in</strong> 10XSCC. After transfer <strong>the</strong> lanes were marked and <strong>the</strong> membrane dried <strong>in</strong> an<br />

oven at 120°C for 30 m<strong>in</strong>.<br />

Probe label<strong>in</strong>g for Sou<strong>the</strong>rn blot was performed us<strong>in</strong>g <strong>the</strong> DIG High prime DNA<br />

Label<strong>in</strong>g and detection starter kit II (Manheim boehr<strong>in</strong>ger, cat. No. 1585614). App 500<br />

ng of <strong>the</strong> whole selection cassette 35S:ntpII:ter (1942 bp fragment <strong>in</strong> water (16 µl total),<br />

obta<strong>in</strong>ed by runn<strong>in</strong>g pMBLU digested with <strong>the</strong> restriction enzymes Pac1 and AsiSI on a<br />

gel and excis<strong>in</strong>g <strong>the</strong> band correspond<strong>in</strong>g to <strong>the</strong> selection cassette, was denatured by<br />

boil<strong>in</strong>g for 10 m<strong>in</strong> and <strong>the</strong>n cooled on ice. The sample was <strong>the</strong>n labeled over night with 4<br />

µl DIG mix. The reaction was stopped by add<strong>in</strong>g 2 µl 0.2 M EDTA (pH 8.0). To quantify<br />

<strong>the</strong> probe, a dot blot was performed with dilutions of <strong>the</strong> probe and a dilution of a DIGlabeled<br />

control DNA (data not shown). Based on this, 3 µl of <strong>the</strong> probe solution was used<br />

<strong>in</strong> 50 ml standard hybridization buffer (described <strong>in</strong> <strong>the</strong> kit manual). Development of <strong>the</strong><br />

membrane was done accord<strong>in</strong>g to <strong>the</strong> manufacturer's specifications.<br />

RNA extraction and quantitative RT-PCR<br />

RNA extraction and quantitative RT-PCR was performed as described <strong>in</strong> manuscript 1.<br />

Primers used for qPCR are listed below <strong>in</strong> Table 10.<br />

Name Phypa ID Gene ID name Forward sequence name Reverse sequence cDNA gDNA<br />

a-DOX Phypa_447346 Pp1s159_20V6.1 sb296 CCGCGAAGTTGCTATCTAGG sb297 AGAGGGTGGAGCCGTAATCT 149<br />

ATG5 Phypa_452597 Pp1s227_54V6.1 sb114 TGGTCGGGAGCCGTGCCAAT sb115 TCTCGGGCTCCAACAACGGT 124 998<br />

CERK1A Phypa_432218 Pp1s41_280V6.1 sb336 ATCCGCCAACATACTCCTTG sb337 AACGACTCCGAAAGCGTAGA 202<br />

CERK1B Phypa_458799 Pp1s364_10V6.1 sb338 CATGAGCACACCAAACCAAC sb339 CAAACCAAAATCTGCCACCT 99 213<br />

CERK1C Phypa_441050 Pp1s97_105V6.1 sb400 CATATGTGGTGCAAGCCAAC sb401 CTTGGAGTTGCCAAAAGGAG 173<br />

CERK1D Phypa_434612 Pp1s54_159V6.1 sb328 CCGAAAGACGAACAAAGAGC sb329 CTAATTGCGCCATCTTCCAT 145<br />

CHS Phypa_427901 Pp1s22_4V6.1 sb263 GGCATGGAACGAGATGTTCT sb264 CCTTGCATCTTGTCCTTGGT 99<br />

CYP71A13 Phypa_450673 Pp1s200_17V6.1 sb273 GCAGAAATATTTGCCCCGTA sb274 TAATCGCCATGTTGACTCCA 125<br />

ERF2 Phypa_422542 Pp1s2_404V6.1 sb275 GAGAGGCGTCCAAACTCTTG sb276 AGGGACTTACGGGCTTGTTT 118<br />

ERF5 Phypa_422320 Pp1s2_410V6.1 sb383 GCTCCGCTGTATCGAAAGTC sb382 TCGAAGTTGCTGACAAGGTG 204<br />

ERFB3 Phypa_431734 Pp1s38_262V6.1 sb316 CTGTCTCCCTACTCCGAACG sb317 AGTCGAACTGCGAGTCCAAT 83<br />

LOX7 Phypa_184078 Pp1s70_182V6.1 sb356 GTGGCGGTTTGATCAGGA sb357 CGTTCAGCCATCCCTCTTC 156 N/A<br />

MEKK1A Phypa_445237 Pp1s136_5V6.3 sb348 GCAGATGAAGGAAAGCTTGG sb349 GCCGGTAAGAATCTGCTCTG 234<br />

MEKK1B Phypa_429991 Pp1s31_252V6.1 sb370 GGATGGCTCCAGAAGTGGTA sb371 CCGGAATAAGAGGACCTTCC 171<br />

MKK1A Phypa_430122 Pp1s32_219V6.1 sb346 GCTTAGGGTTGACGCTCTTG sb347 ACTTTGGATCCTTCTGCAAG 195 318<br />

MKK1B Phypa_442149 Pp1s106_83V6.1 sb344 TTACGCATTGAAGGGGATTC sb345 CTGATGCAGTGTCAGCTGGT 94 210<br />

MKK1C Phypa_433829 Pp1s50_83V6.1 sb342 TTCGTTGGGACTTGCACATA sb343 CGCACACTCCAAAAGAGTCA 105 297<br />

MPK4A Phypa_446407 Pp1s149_39V6.1 sb228 GGTACAAGCCACCACTTCGT sb270 GGTCCGTATCCATCAACTCG 270 890<br />

MPK4B Phypa_435424 Pp1s59_325V6.1 sb267 GGCGAGTACACGCAGTACAA sb268 ATTCACAGCGGAACACACAA 117 N/A<br />

NOX Phypa_204103 Pp1s18_194V6.1 sb354 CACGATGTTGCAGTCGTTG sb355 TACGTGCCCTAGTGCCTGA 68 N/A<br />

PAL4 Phypa_461241 Pp1s500_4V6.1 sb284 TGGCCTACTCGGTAATGGAG sb285 GTCAACCATCCGCTTGATTT 109<br />

83


PAL4-2 Phypa_180561 Pp1s43_67V6.1 sb396 CGTTCGAGGAGGAACTCAAG sb397 TGCAGTTCTCGATCCTGTTG 108<br />

PRX34A Phypa_144797 Pp1s219_8V2.1 sb104 CAATACGCTACTCGCGACTCTGT sb105 CGTCTCTTCGACCGCCATA 73 235<br />

RAR1 Phypa_461267 Pp1s505_15V5.1 sb252 AGGCAGCTCCTAGTCCTAACG sb216 GGATGGCACAGCCTTCTTTA 222 402<br />

SGT1 Phypa_453593 Pp1s244_58V6.1 sb332 CTCGCCTGTATGGCAAGATT sb333 GGTCGAACTCCAACTGCTTC 117<br />

TUB6 Phypa_440500 Pp1s93_158V6.1 sb041 GAGTTCACGGAAGCGGAGAG sb042 ATATCTTTCAGGCTCCACCG 224<br />

Wkry 33 Phypa_431011 Pp1s35_336V6.1 sb277 GATACGGAAGGACGACGAAA sb278 CCTGGACATCATCTGTGGTG 130<br />

Wrky18 Phypa_447355 Pp1s159_105V6.1 sb322 CCAGAAATGTCGAACCACCT sb323 TTCTCAAACGCCGTCTTCTT 127<br />

Wrky25 Phypa_445119 Pp1s135_80V6.1 sb290 ATTTGTTGGCAGGAGCAATC sb291 GGAGCCTCTGCCATCAGTAG 140<br />

Wrky40 Phypa_433427 Pp1s47_251V6.1 sb318 GAAAAGCAGTGGGAGAGTCG sb319 TGTGATGGTCCGACCTTGTA 137<br />

Wrky53 Phypa_447750 Pp1s163_112V6.2 sb320 CTCATCAAAACCTCCGCAAT sb321 GCAGAAGGTGGTGAGCTTTC 152<br />

Wrky70 Phypa_447996 Pp1s166_34V6.1 sb310 GAAAGACTTGGGGCATTTGA sb311 ACAATGTCCTCCAGCAATCC 150<br />

Table 10. Primers used for qPCR. Listed are <strong>the</strong> gene name and identifiers, <strong>the</strong> primer names and<br />

sequences, and <strong>the</strong> product length when amplified on cDNA and gDNA. If no product length is listed<br />

for a cDNA, it is <strong>the</strong> same as for gDNA, and N/A means that one of <strong>the</strong> primers spans an exon/<strong>in</strong>tron<br />

junction and thus can only amplify cDNA.<br />

Statistics<br />

The statistical analysis of quantitative Evans blue sta<strong>in</strong><strong>in</strong>g and Alternaria spore count<strong>in</strong>g<br />

assays were done by first perform<strong>in</strong>g a one way analysis of variance (ANOVA) followed<br />

by a Tukey's HSD (Honestly Significant Difference) test. For <strong>the</strong> Alternaria spore<br />

count<strong>in</strong>g assay <strong>the</strong> test was performed on mean values of <strong>the</strong> count<strong>in</strong>g for each sample<br />

(Anders Tolver Jensen, personal communication). The tests were performed <strong>in</strong> “R” us<strong>in</strong>g<br />

follow<strong>in</strong>g codes:<br />

#ANOVA and Tukey HSD test<br />

rm(list=ls()) # Removes all objects<br />

your = read.csv(file.choose()) # Opens w<strong>in</strong>dos explorer, <strong>the</strong>n enter file<br />

attach(your) # attaches variables<br />

summary(aov(count ~ gt)) # Gives summery of anova test<br />

mod


control treatment (C) (0.01% Acetic acid, adjusted to pH 5.5 with NaOH). Moss colonies<br />

were <strong>the</strong>n <strong>in</strong>cubated for 2 h <strong>in</strong> 1 mg/ml DAB, HCl (pH 3.8). The colonies were r<strong>in</strong>sed<br />

twice <strong>in</strong> H 2 O and chlorophyll removed by <strong>in</strong>cubat<strong>in</strong>g overnight <strong>in</strong> 96% (v/v) ethanol.<br />

Histochemical sta<strong>in</strong><strong>in</strong>g<br />

The histochemical sta<strong>in</strong><strong>in</strong>gs and microscopic pictures <strong>in</strong> Figure 24 and Figure 25 were<br />

performed at IIBCE <strong>in</strong> Uruguay with <strong>the</strong> help of Inés Ponce de León as previously<br />

described (Oliver et al., 2009; Ponce de León et al., 2007; Ponce De León et al., 2012).<br />

Bright field microscopy and fluorescence microscopy were performed with an Olympus<br />

BX61 microscope (Sh<strong>in</strong>juku-ku, Tokyo, Japan). All images shown <strong>in</strong> Figure 24 and<br />

Figure 25 were captured with <strong>the</strong> Microsuite software package (Olympus).<br />

Cell death was visualized by sta<strong>in</strong><strong>in</strong>g for 2 h with 0.1% Evans blue (Sigma) <strong>in</strong> 0.5<br />

x PBS and wash<strong>in</strong>g four times <strong>in</strong> H 2 O to remove excess unbound dye. The <strong>in</strong>tracellular<br />

production of ROS was analyzed by <strong>in</strong>cubat<strong>in</strong>g <strong>moss</strong> tissues with 10 mM 2 ′ ,7 ′ -<br />

dichlorodihydrofluoresce<strong>in</strong> diacetate (H 2 DCFDA) for 15 m<strong>in</strong> <strong>in</strong> 0.1 M phosphate buffer<br />

(pH 7.5) <strong>in</strong> <strong>the</strong> dark. Leaves were visualized with epifluorescence. Botrytis c<strong>in</strong>erea and<br />

Pythium irregulare <strong>in</strong>oculated tissues were sta<strong>in</strong>ed with 0.1% solophenyl flav<strong>in</strong>e 7GFE<br />

500 <strong>in</strong> water for 10 m<strong>in</strong>, r<strong>in</strong>sed <strong>in</strong> water and visualized with epifluorescence. Callose<br />

detection was performed by fix<strong>in</strong>g <strong>the</strong> tissue <strong>in</strong> ethanol over night, r<strong>in</strong>s<strong>in</strong>g <strong>in</strong> water and<br />

sta<strong>in</strong><strong>in</strong>g with 0.01% methyl blue <strong>in</strong> phosphate buffer pH 7.0 for 30 m<strong>in</strong> and observed<br />

with epifluorescence. Phenolic compounds were detected by sta<strong>in</strong><strong>in</strong>g tissues with 0.05%<br />

toluid<strong>in</strong>e blue <strong>in</strong> citrate–citric acid buffer (50 mM, pH 3.5).<br />

cDNA<br />

cDNA was prepared from a total RNA extraction from wild type <strong>moss</strong> us<strong>in</strong>g <strong>the</strong><br />

Superscript III kit (Clonetech) accord<strong>in</strong>g to <strong>the</strong> manufacturer's specifications.<br />

Yeast two-hybrid<br />

The yeast two-hybrid analyses were done us<strong>in</strong>g <strong>the</strong> Matchmaker TM kit from Clonetech<br />

accord<strong>in</strong>g to <strong>the</strong> manufacturer's specifications.<br />

Western blots<br />

The MPK phosphorylation western blots were done as described <strong>in</strong> manuscript 1<br />

The ATG8 western blot was done as described <strong>in</strong> (Yoshimoto et al., 2004)<br />

PCR conditions<br />

All PCRs for USER clon<strong>in</strong>g were performed with <strong>the</strong> PfuCx polymerase, a mutant of Pfu<br />

DNA polymerase that overcomes uracil stall<strong>in</strong>g completely, allow<strong>in</strong>g <strong>the</strong> polymerase to<br />

read through uracil located <strong>in</strong> <strong>the</strong> template strand or <strong>in</strong>corporated <strong>in</strong>to <strong>the</strong> extend<strong>in</strong>g<br />

85


strand. For PCR amplification of long fragments, a mix of homemade Pfu and Phusion<br />

hot start polymerase (F<strong>in</strong>nzymes) was used.<br />

PCR for USER clon<strong>in</strong>g: Total volume of 50 μl, 39 μl H2O, 2 μl WT DNA (10<br />

ng/μl), 1 μl PFUCx, 1 μl forward primer (10mM), 1 μl reverse primer, DNTP 10 mM 1μl,<br />

5 μl 10x Pfu buffer. PCR program: 98ºC for 30 sec., 35 times (98ºC for 10 sec., 57ºC for<br />

25 sec., 72ºC for 45 sec.), 72ºC for 5 m<strong>in</strong>.<br />

86


References<br />

Akita, M., Lehtonen, M.T., Koponen, H., Martt<strong>in</strong>en, E.M., and Valkonen, J.P.T. (2011).<br />

Infection of <strong>the</strong> Sunagoke <strong>moss</strong> panels with fungal pathogens hampers susta<strong>in</strong>able<br />

green<strong>in</strong>g <strong>in</strong> urban environments. Science of The Total Environment 409, 3166–3173.<br />

Akita, M., and Valkonen, J.P.T. (2002). A Novel Gene Family <strong>in</strong> Moss (<strong>Physcomitrella</strong><br />

<strong>patens</strong>) Shows Sequence Homology and a Phylogenetic Relationship with <strong>the</strong> TIR-NBS<br />

Class of Plant Disease Resistance Genes. Journal of Molecular Evolution 55, 595–605.<br />

Andersson, R.A., Akita, M., Pirhonen, M., Gammelgård, E., and Valkonen, J.P.T. (2005).<br />

Moss-Erw<strong>in</strong>ia pathosystem reveals possible similarities <strong>in</strong> pathogenesis and pathogen<br />

defense <strong>in</strong> vascular and nonvascular plants. Journal of General Plant Pathology 71, 23–28.<br />

Andreasson, E., Jenk<strong>in</strong>s, T., Brodersen, P., Thorgrimsen, S., Petersen, N.H.T., Zhu, S.,<br />

Qiu, J.-L., Micheelsen, P., Rocher, A., Petersen, M., et al. (2005). The MAP k<strong>in</strong>ase<br />

substrate MKS1 is a regulator of plant defense responses. EMBO J 24, 2579–2589.<br />

Anthony, R.G., Henriques, R., Helfer, A., Mészáros, T., Rios, G., Tester<strong>in</strong>k, C., Munnik,<br />

T., Deák, M., Koncz, C., and Bögre, L. (2004). A prote<strong>in</strong> k<strong>in</strong>ase target of a PDK1<br />

signall<strong>in</strong>g pathway is <strong>in</strong>volved <strong>in</strong> root hair growth <strong>in</strong> Arabidopsis. EMBO J 23, 572–581.<br />

Apel, K., and Hirt, H. (2004a). REACTIVE OXYGEN SPECIES: Metabolism, Oxidative<br />

Stress, and Signal Transduction. Annual Review of Plant Biology 55, 373–399.<br />

Asai, T., Tena, G., Plotnikova, J., Willmann, M.R., Chiu, W.-L., Gomez-Gomez, L.,<br />

Boller, T., Ausubel, F.M., and Sheen, J. (2002). MAP k<strong>in</strong>ase signall<strong>in</strong>g cascade <strong>in</strong><br />

Arabidopsis <strong>in</strong>nate immunity. Nature 415, 977–983.<br />

Ashton, N.W., and Cove, D.J. (1977). The isolation and prelim<strong>in</strong>ary characterisation of<br />

auxotrophic and analogue resistant mutants of <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. Molecular<br />

and General Genetics MGG 154, 87–95.<br />

Ausubel, F.M. (2005). Are <strong>in</strong>nate immune signal<strong>in</strong>g pathways <strong>in</strong> plants and animals<br />

conserved Nature Immunology 6, 973–979.<br />

Azevedo, C., Betsuyaku, S., Peart, J., Takahashi, A., Noël, L., Sadanandom, A., Casais,<br />

C., Parker, J., and Shirasu, K. (2006). Role of SGT1 <strong>in</strong> resistance prote<strong>in</strong> accumulation <strong>in</strong><br />

plant immunity. The EMBO Journal 25, 2007–2016.<br />

Banks, J.A., Nishiyama, T., Hasebe, M., Bowman, J.L., Gribskov, M., dePamphilis, C.,<br />

Albert, V.A., Aono, N., Aoyama, T., Ambrose, B.A., et al. (2011). The Selag<strong>in</strong>ella<br />

Genome Identifies Genetic Changes Associated with <strong>the</strong> Evolution of Vascular Plants.<br />

Science 332, 960–963.<br />

87


Bendahmane, A., Farnham, G., Moffett, P., and Baulcombe, D.C. (2002). Constitutive<br />

ga<strong>in</strong>-of-function mutants <strong>in</strong> a nucleotide b<strong>in</strong>d<strong>in</strong>g site–leuc<strong>in</strong>e rich repeat prote<strong>in</strong> encoded<br />

at <strong>the</strong> Rx locus of potato. The Plant Journal 32, 195–204.<br />

Bethke, G., Pecher, P., Eschen-Lippold, L., Tsuda, K., Katagiri, F., Glazebrook, J.,<br />

Scheel, D., and Lee, J. (2012). Activation of <strong>the</strong> Arabidopsis thaliana mitogen-activated<br />

prote<strong>in</strong> k<strong>in</strong>ase MPK11 by <strong>the</strong> flagell<strong>in</strong>-derived elicitor peptide, flg22. Mol. Plant Microbe<br />

Interact. 25, 471–480.<br />

Bezanilla, M., Perroud, P.-F., Pan, A., Klueh, P., and Quatrano, R.S. (2005). An RNAi<br />

System <strong>in</strong> <strong>Physcomitrella</strong> <strong>patens</strong> with an Internal Marker for Silenc<strong>in</strong>g Allows for Rapid<br />

Identification of Loss of Function Phenotypes. Plant Biology 7, 251–257.<br />

van der Biezen, E.A., and Jones, J.D.G. (1998). The NB-ARC doma<strong>in</strong>: a novel signall<strong>in</strong>g<br />

motif shared by plant resistance gene products and regulators of cell death <strong>in</strong> animals.<br />

Current Biology 8, R226–R228.<br />

Birkenbihl, R.P., Diezel, C., and Somssich, I.E. (2012). Arabidopsis WRKY33 Is a Key<br />

Transcriptional Regulator of Hormonal and Metabolic Responses toward Botrytis c<strong>in</strong>erea<br />

Infection. Plant Physiol. 159, 266–285.<br />

Bishop, A.C., Ubersax, J.A., Petsch, D.T., Ma<strong>the</strong>os, D.P., Gray, N.S., Blethrow, J.,<br />

Shimizu, E., Tsien, J.Z., Schultz, P.G., Rose, M.D., et al. (2000). A chemical switch for<br />

<strong>in</strong>hibitor-sensitive alleles of any prote<strong>in</strong> k<strong>in</strong>ase. Nature 407, 395–401.<br />

Boller, T., and Felix, G. (2009). A Renaissance of Elicitors: Perception of Microbe-<br />

Associated Molecular Patterns and Danger Signals by Pattern-Recognition Receptors.<br />

Annual Review of Plant Biology 60, 379–406.<br />

Boudsocq, M., and Sheen, J. CDPKs <strong>in</strong> immune and stress signal<strong>in</strong>g. Trends <strong>in</strong> Plant<br />

Science.<br />

Breeze, E., Harrison, E., McHattie, S., Hughes, L., Hickman, R., Hill, C., Kiddle, S., Kim,<br />

Y., Penfold, C.A., Jenk<strong>in</strong>s, D., et al. (2011). High-Resolution Temporal Profil<strong>in</strong>g of<br />

Transcripts dur<strong>in</strong>g Arabidopsis Leaf Senescence Reveals a Dist<strong>in</strong>ct Chronology of<br />

Processes and Regulation. Plant Cell 23, 873–894.<br />

Brodersen, P., Petersen, M., Bjørn Nielsen, H., Zhu, S., Newman, M.-A., Shokat, K.M.,<br />

Rietz, S., Parker, J., and Mundy, J. (2006). Arabidopsis MAP k<strong>in</strong>ase 4 regulates salicylic<br />

acid- and jasmonic acid/ethylene-dependent responses via EDS1 and PAD4. The Plant<br />

Journal 47, 532–546.<br />

Chater, C., Kamisugi, Y., Movahedi, M., Flem<strong>in</strong>g, A., Cum<strong>in</strong>g, A.C., Gray, J.E., and<br />

Beerl<strong>in</strong>g, D.J. (2011). Regulatory Mechanism Controll<strong>in</strong>g Stomatal Behavior Conserved<br />

across 400 Million Years of Land Plant Evolution. Current Biology 21, 1025–1029.<br />

Chen, B., Zhong, D., and Monteiro, A. (2006). Comparative genomics and evolution of<br />

<strong>the</strong> HSP90 family of genes across all k<strong>in</strong>gdoms of organisms. BMC Genomics 7, 156.<br />

88


Ch<strong>in</strong>chilla, D., Bauer, Z., Regenass, M., Boller, T., and Felix, G. (2006). The Arabidopsis<br />

Receptor K<strong>in</strong>ase FLS2 B<strong>in</strong>ds flg22 and Determ<strong>in</strong>es <strong>the</strong> Specificity of Flagell<strong>in</strong><br />

Perception. Plant Cell 18, 465–476.<br />

Ch<strong>in</strong>chilla, D., Zipfel, C., Robatzek, S., Kemmerl<strong>in</strong>g, B., N|[uuml]|rnberger, T., Jones,<br />

J.D.G., Felix, G., and Boller, T. (2007). A flagell<strong>in</strong>-<strong>in</strong>duced complex of <strong>the</strong> receptor<br />

FLS2 and BAK1 <strong>in</strong>itiates plant defence. Nature 448, 497–500.<br />

Clarke, J.T., Warnock, R.C.M., and Donoghue, P.C.J. (2011). Establish<strong>in</strong>g a time-scale<br />

for plant evolution. New Phytologist 192, 266–301.<br />

Coll, N.S., Epple, P., and Dangl, J.L. (2011). Programmed cell death <strong>in</strong> <strong>the</strong> plant immune<br />

system. Cell Death & Differentiation 18, 1247–1256.<br />

Daudi, A., Cheng, Z., O’Brien, J.A., Mammarella, N., Khan, S., Ausubel, F.M., and<br />

Bolwell, G.P. (2012). The Apoplastic Oxidative Burst Peroxidase <strong>in</strong> Arabidopsis Is a<br />

Major Component of Pattern-Triggered <strong>Immunity</strong>[W][OA]. Plant Cell 24, 275–287.<br />

DeYoung, B.J., and Innes, R.W. (2006). Plant NBS-LRR prote<strong>in</strong>s <strong>in</strong> pathogen sens<strong>in</strong>g<br />

and host defense. Nature Immunology 7, 1243–1249.<br />

Diebold, B.A., and Bokoch, G.M. (2005). Rho GTPases and <strong>the</strong> control of <strong>the</strong> oxidative<br />

burst <strong>in</strong> polymorphonuclear leukocytes. Curr. Top. Microbiol. Immunol. 291, 91–111.<br />

Dietz, K.-J., Vogel, M., and Viehhauser, A. (2010). AP2/EREBP transcription factors are<br />

part of gene regulatory networks and <strong>in</strong>tegrate metabolic, hormonal and environmental<br />

signals <strong>in</strong> stress acclimation and retrograde signall<strong>in</strong>g. Protoplasma 245, 3–14.<br />

Dóczi, R., Ökrész, L., Romero, A.E., Paccanaro, A., and Bögre, L. (2012). Explor<strong>in</strong>g <strong>the</strong><br />

evolutionary path of plant MAPK networks. Trends <strong>in</strong> Plant Science 17, 503–562.<br />

Droillard, M.-J., Boudsocq, M., Barbier-Brygoo, H., and Laurière, C. (2004).<br />

Involvement of MPK4 <strong>in</strong> osmotic stress response pathways <strong>in</strong> cell suspensions and<br />

plantlets of Arabidopsis thaliana: activation by hypoosmolarity and negative role <strong>in</strong><br />

hyperosmolarity tolerance. FEBS Letters 574, 42–48.<br />

Edwards, K., Johnstone, C., and Thompson, C. (1991). A simple and rapid method for <strong>the</strong><br />

preparation of plant genomic DNA for PCR analysis. Nucleic Acids Res 19, 1349.<br />

Egener, T., Granado, J., Guitton, M.-C., Hohe, A., Holtorf, H., Lucht, J.M., Rens<strong>in</strong>g, S.A.,<br />

Schl<strong>in</strong>k, K., Schulte, J., Schween, G., et al. (2002). High frequency of phenotypic<br />

deviations <strong>in</strong> <strong>Physcomitrella</strong> <strong>patens</strong> plants transformed with a gene-disruption library.<br />

BMC Plant Biology 2, 6.<br />

El Oirdi, M., and Bouarab, K. (2007). Plant signall<strong>in</strong>g components EDS1 and SGT1<br />

enhance disease caused by <strong>the</strong> necrotrophic pathogen Botrytis c<strong>in</strong>erea. New Phytologist<br />

175, 131–139.<br />

89


Enoksson, M., and Salvesen, G.S. (2010). Metacaspases are not caspases – always doubt.<br />

Cell Death & Differentiation 17, 1221–1221.<br />

Eruslanov, E., and Kusmartsev, S. (2010). Identification of ROS Us<strong>in</strong>g Oxidized DCFDA<br />

and Flow-Cytometry. In Advanced Protocols <strong>in</strong> Oxidative Stress II, D. Armstrong, ed.<br />

(Humana Press), pp. 57–72.<br />

Eschen-Lippold, L., Bethke, G., Palm-Forster, M., Pecher, P., Bauer, N., Glazebrook, J.,<br />

Scheel, D., and Lee, J. (2012). MPK11—a fourth elicitor-responsive mitogen-activated<br />

prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong> Arabidopsis thaliana. Plant Signal<strong>in</strong>g & Behavior 7, 1203–1205.<br />

Felix, G., Duran, J.D., Volko, S., and Boller, T. (2002). Plants have a sensitive perception<br />

system for <strong>the</strong> most conserved doma<strong>in</strong> of bacterial flagell<strong>in</strong>. The Plant Journal 18, 265–<br />

276.<br />

Felix, G., Regenass, M., and Boller, T. (1993). Specific perception of subnanomolar<br />

concentrations of chit<strong>in</strong> fragments by tomato cells: <strong>in</strong>duction of extracellular<br />

alkal<strong>in</strong>ization, changes <strong>in</strong> prote<strong>in</strong> phosphorylation, and establishment of a refractory state.<br />

The Plant Journal 4, 307–316.<br />

Ferrari, S., Plotnikova, J.M., De Lorenzo, G., and Ausubel, F.M. (2003). Arabidopsis<br />

local resistance to Botrytis c<strong>in</strong>erea <strong>in</strong>volves salicylic acid and camalex<strong>in</strong> and requires<br />

EDS4 and PAD2, but not SID2, EDS5 or PAD4. The Plant Journal 35, 193–205.<br />

Feys, B.J., Wiermer, M., Bhat, R.A., Moisan, L.J., Med<strong>in</strong>a-Escobar, N., Neu, C., Cabral,<br />

A., and Parker, J.E. (2005). Arabidopsis SENESCENCE-ASSOCIATED GENE101<br />

Stabilizes and Signals with<strong>in</strong> an ENHANCED DISEASE SUSCEPTIBILITY1 Complex<br />

<strong>in</strong> Plant Innate <strong>Immunity</strong>. Plant Cell 17, 2601–2613.<br />

F<strong>in</strong>n, R.D., Tate, J., Mistry, J., Coggill, P.C., Sammut, S.J., Hotz, H.-R., Ceric, G.,<br />

Forslund, K., Eddy, S.R., Sonnhammer, E.L.L., et al. (2007). The Pfam prote<strong>in</strong> families<br />

database. Nucleic Acids Research 36, D281–D288.<br />

Forman, H.J., Maior<strong>in</strong>o, M., and Urs<strong>in</strong>i, F. (2010). Signal<strong>in</strong>g Functions of Reactive<br />

Oxygen Species. Biochemistry 49, 835–842.<br />

Frank, W., Baar, K.-M., Qudeimat, E., Woriedh, M., Alawady, A., Ratnadewi, D.,<br />

Gremillon, L., Grimm, B., and Reski, R. (2007). A mitochondrial prote<strong>in</strong> homologous to<br />

<strong>the</strong> mammalian peripheral-type benzodiazep<strong>in</strong>e receptor is essential for stress adaptation<br />

<strong>in</strong> plants. The Plant Journal 51, 1004–1018.<br />

Frías, M., González, C., and Brito, N. (2011). BcSpl1, a cerato-platan<strong>in</strong> family prote<strong>in</strong>,<br />

contributes to Botrytis c<strong>in</strong>erea virulence and elicits <strong>the</strong> hypersensitive response <strong>in</strong> <strong>the</strong><br />

host. New Phytologist 192, 483–495.<br />

Gabriëls, S.H.E.J., Vossen, J.H., Ekengren, S.K., Ooijen, G. van, Abd-El-Haliem, A.M.,<br />

Berg, G.C.M. van den, Ra<strong>in</strong>ey, D.Y., Mart<strong>in</strong>, G.B., Takken, F.L.W., Wit, P.J.G.M. de, et<br />

90


al. (2007). An NB-LRR prote<strong>in</strong> required for HR signall<strong>in</strong>g mediated by both extra- and<br />

<strong>in</strong>tracellular resistance prote<strong>in</strong>s. The Plant Journal 50, 14–28.<br />

Gao, M., Liu, J., Bi, D., Zhang, Z., Cheng, F., Chen, S., and Zhang, Y. (2008). MEKK1,<br />

MKK1/MKK2 and MPK4 function toge<strong>the</strong>r <strong>in</strong> a mitogen-activated prote<strong>in</strong> k<strong>in</strong>ase<br />

cascade to regulate <strong>in</strong>nate immunity <strong>in</strong> plants. Cell Research 18, 1190–1198.<br />

Gómez-Gómez, L., and Boller, T. (2000). FLS2: An LRR Receptor–like K<strong>in</strong>ase Involved<br />

<strong>in</strong> <strong>the</strong> Perception of <strong>the</strong> Bacterial Elicitor Flagell<strong>in</strong> <strong>in</strong> Arabidopsis. Molecular Cell 5,<br />

1003–1011.<br />

González Besteiro, M.A., Bartels, S., Albert, A., and Ulm, R. (2011). Arabidopsis MAP<br />

k<strong>in</strong>ase phosphatase 1 and its target MAP k<strong>in</strong>ases 3 and 6 antagonistically determ<strong>in</strong>e UV-<br />

B stress tolerance, <strong>in</strong>dependent of <strong>the</strong> UVR8 photoreceptor pathway. The Plant Journal<br />

68, 727–737.<br />

Govr<strong>in</strong>, E.M., and Lev<strong>in</strong>e, A. (2000). The hypersensitive response facilitates plant<br />

<strong>in</strong>fection by <strong>the</strong> necrotrophic pathogen Botrytis c<strong>in</strong>erea. Current Biology 10, 751–757.<br />

Govr<strong>in</strong>, E.M., Rachmilevitch, S., Tiwari, B.S., Solomon, M., and Lev<strong>in</strong>e, A. (2006). An<br />

Elicitor from Botrytis c<strong>in</strong>erea Induces <strong>the</strong> Hypersensitive Response <strong>in</strong> Arabidopsis<br />

thaliana and O<strong>the</strong>r Plants and Promotes <strong>the</strong> Gray Mold Disease. Phytopathology<br />

96, 299–307.<br />

Han, J., Lee, J.D., Bibbs, L., and Ulevitch, R.J. (1994). A MAP k<strong>in</strong>ase targeted by<br />

endotox<strong>in</strong> and hyperosmolarity <strong>in</strong> mammalian cells. Science 265, 808–811.<br />

Hedges, S.B. (2002). The orig<strong>in</strong> and evolution of model organisms. Nature Reviews<br />

Genetics 3, 838–849.<br />

Hoch, H.C., Galvani, C.D., Szarowski, D.H., and Turner, J.N. (2005). Two new<br />

fluorescent dyes applicable for visualization of fungal cell walls. Mycologia 97, 580–588.<br />

Hofius, D., Schultz-Larsen, T., Joensen, J., Tsitsigiannis, D.I., Petersen, N.H.T., Mattsson,<br />

O., Jørgensen, L.B., Jones, J.D.G., Mundy, J., and Petersen, M. (2009). Autophagic<br />

Components Contribute to Hypersensitive Cell Death <strong>in</strong> Arabidopsis. Cell 137, 773–783.<br />

Hohe, A., Rens<strong>in</strong>g, S.A., Mildner, M., Lang, D., and Reski, R. (2002). Day Length and<br />

Temperature Strongly Influence Sexual Reproduction and Expression of a Novel MADS-<br />

Box Gene <strong>in</strong> <strong>the</strong> Moss <strong>Physcomitrella</strong> <strong>patens</strong>. Plant Biology 4, 595–602.<br />

Hruz, T., Laule, O., Szabo, G., Wessendorp, F., Bleuler, S., Oertle, L., Widmayer, P.,<br />

Gruissem, W., and Zimmermann, P. (2008). Genevestigator V3: A Reference Expression<br />

Database for <strong>the</strong> Meta-Analysis of Transcriptomes. Advances <strong>in</strong> Bio<strong>in</strong>formatics 2008, 1–<br />

5.<br />

Hübers, M., and Kerp, H. (2012). Oldest known <strong>moss</strong>es discovered <strong>in</strong> Mississippian (late<br />

Visean) strata of Germany. Geology 40, 755–758.<br />

91


Ichimura, K., Mizoguchi, T., Yoshida, R., Yuasa, T., and Sh<strong>in</strong>ozaki, K. (2000). Various<br />

abiotic stresses rapidly activate Arabidopsis MAP k<strong>in</strong>ases ATMPK4 and ATMPK6. The<br />

Plant Journal 24, 655–665.<br />

Ichimura, K., Sh<strong>in</strong>ozaki, K., Tena, G., Sheen, J., Henry, Y., Champion, A., Kreis, M.,<br />

Zhang, S., Hirt, H., Wilson, C., et al. (2002). Mitogen-activated prote<strong>in</strong> k<strong>in</strong>ase cascades<br />

<strong>in</strong> plants: a new nomenclature. Trends <strong>in</strong> Plant Science 7, 301–308.<br />

Jacobsen, J., Rosgaard, L., Sakuragi, Y., and Frigaard, N.-U. (2011). One-step plasmid<br />

construction for generation of knock-out mutants <strong>in</strong> cyanobacteria: studies of glycogen<br />

metabolism <strong>in</strong> Synechococcus sp. PCC 7002. Photosyn<strong>the</strong>sis Research 107, 215–221.<br />

Jones, J.D.G., and Dangl, J.L. (2006). The plant immune system. Nature 444, 323–329.<br />

Jones, V.A.S., and Dolan, L. (2012). The evolution of root hairs and rhizoids. Ann Bot.<br />

Kamisugi, Y., Schl<strong>in</strong>k, K., Rens<strong>in</strong>g, S.A., Schween, G., Stackelberg, M. von, Cum<strong>in</strong>g,<br />

A.C., Reski, R., and Cove, D.J. (2006). The mechanism of gene target<strong>in</strong>g <strong>in</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>: homologous recomb<strong>in</strong>ation, concatenation and multiple<br />

<strong>in</strong>tegration. Nucl. Acids Res. 34, 6205–6214.<br />

Klionsky, D.J., Abeliovich, H., Agost<strong>in</strong>is, P., Agrawal, D.K., Aliev, G., Askew, D.S.,<br />

Baba, M., Baehrecke, E.H., Bahr, B.A., Ballabio, A., et al. (2008). Guidel<strong>in</strong>es for <strong>the</strong> use<br />

and <strong>in</strong>terpretation of assays for monitor<strong>in</strong>g autophagy <strong>in</strong> higher eukaryotes. Autophagy 4,<br />

151–175.<br />

Knepper, C., Savory, E.A., and Day, B. (2011a). Arabidopsis NDR1 Is an Integr<strong>in</strong>-Like<br />

Prote<strong>in</strong> with a Role <strong>in</strong> Fluid Loss and Plasma Membrane-Cell Wall Adhesion1[W][OA].<br />

Plant Physiol 156, 286–300.<br />

Knepper, C., Savory, E.A., and Day, B. (2011b). The role of NDR1 <strong>in</strong> pathogen<br />

perception and plant defense signal<strong>in</strong>g. Plant Signal<strong>in</strong>g & Behavior 6, 1114–1116.<br />

Kohchi, C., Inagawa, H., Nishizawa, T., and Soma, G.-I. (2009). ROS and Innate<br />

<strong>Immunity</strong>. Anticancer Res 29, 817–821.<br />

Kong, Q., Qu, N., Gao, M., Zhang, Z., D<strong>in</strong>g, X., Yang, F., Li, Y., Dong, O.X., Chen, S.,<br />

Li, X., et al. (2012). The MEKK1-MKK1/MKK2-MPK4 K<strong>in</strong>ase Cascade Negatively<br />

Regulates <strong>Immunity</strong> Mediated by a Mitogen-Activated Prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ase K<strong>in</strong>ase <strong>in</strong><br />

Arabidopsis. Plant Cell.<br />

Kosetsu, K., Matsunaga, S., Nakagami, H., Colcombet, J., Sasabe, M., Soyano, T.,<br />

Takahashi, Y., Hirt, H., and Machida, Y. (2010). The MAP K<strong>in</strong>ase MPK4 Is Required for<br />

Cytok<strong>in</strong>esis <strong>in</strong> Arabidopsis thaliana. Plant Cell 22, 3778–3790.<br />

Krol, E., Mentzel, T., Ch<strong>in</strong>chilla, D., Boller, T., Felix, G., Kemmerl<strong>in</strong>g, B., Postel, S.,<br />

Arents, M., Jeworutzki, E., Al-Rasheid, K.A.S., et al. (2010). Perception of <strong>the</strong><br />

92


Arabidopsis Danger Signal Peptide 1 Involves <strong>the</strong> Pattern Recognition Receptor<br />

AtPEPR1 and Its Close Homologue AtPEPR2. J Biol Chem 285, 13471–13479.<br />

Ku, H.-M., Vision, T., Liu, J., and Tanksley, S.D. (2000). Compar<strong>in</strong>g sequenced<br />

segments of <strong>the</strong> tomato and Arabidopsis genomes: Large-scale duplication followed by<br />

selective gene loss creates a network of synteny. PNAS 97, 9121–9126.<br />

Kültz, D. (2001). Evolution of Osmosensory MAP K<strong>in</strong>ase Signal<strong>in</strong>g Pathways. Amer.<br />

Zool. 41, 743–757.<br />

Lai, Z., Wang, F., Zheng, Z., Fan, B., and Chen, Z. (2011). A critical role of autophagy <strong>in</strong><br />

plant resistance to necrotrophic fungal pathogens. The Plant Journal 66, 953–968.<br />

Lark<strong>in</strong>, M.A., Blackshields, G., Brown, N.P., Chenna, R., McGettigan, P.A., McWilliam,<br />

H., Valent<strong>in</strong>, F., Wallace, I.M., Wilm, A., Lopez, R., et al. (2007). Clustal W and Clustal<br />

X Version 2.0. Bio<strong>in</strong>formatics 23, 2947–2948.<br />

Lawton, M., and Saidasan, H. (2009). Pathogenesis <strong>in</strong> Mosses. In Annual Plant Reviews<br />

Volume 36: The Moss <strong>Physcomitrella</strong> Patens, C.D. Knight, P.-F. Perroud, and D.J. Cove,<br />

eds. (Wiley-Blackwell), pp. 298–338.<br />

Lehtonen, M.T., Akita, M., Frank, W., Reski, R., and Valkonen, J.P.T. (2012).<br />

Involvement of a Class III Peroxidase and <strong>the</strong> Mitochondrial Prote<strong>in</strong> TSPO <strong>in</strong> Oxidative<br />

Burst Upon Treatment of Moss Plants with a Fungal Elicitor. Molecular Plant-Microbe<br />

Interactions 25, 363–371.<br />

Lehtonen, M.T., Akita, M., Kalkk<strong>in</strong>en, N., Ahola-Iivar<strong>in</strong>en, E., Rönnholm, G., Somervuo,<br />

P., Thelander, M., and Valkonen, J.P.T. (2009). Quickly-released peroxidase of <strong>moss</strong> <strong>in</strong><br />

defense aga<strong>in</strong>st fungal <strong>in</strong>vaders. New Phytologist 183, 432–443.<br />

Lenz, H.D., Haller, E., Melzer, E., Kober, K., Wurster, K., Stahl, M., Bassham, D.C.,<br />

Vierstra, R.D., Parker, J.E., Bautor, J., et al. (2011). Autophagy differentially controls<br />

plant basal immunity to biotrophic and necrotrophic pathogens. The Plant Journal 66,<br />

818–830.<br />

Lewis, L.A., and McCourt, R.M. (2004). Green algae and <strong>the</strong> orig<strong>in</strong> of land plants. Am. J.<br />

Bot. 91, 1535–1556.<br />

Libault, M., Wan, J., Czechowski, T., Udvardi, M., and Stacey, G. (2007). Identification<br />

of 118 Arabidopsis Transcription Factor and 30 Ubiquit<strong>in</strong>-Ligase Genes<br />

Respond<strong>in</strong>g to Chit<strong>in</strong>, a Plant-Defense Elicitor. Molecular Plant-Microbe Interactions 20,<br />

900–911.<br />

Liu, T., Liu, Z., Song, C., Hu, Y., Han, Z., She, J., Fan, F., Wang, J., J<strong>in</strong>, C., Chang, J., et<br />

al. (2012). Chit<strong>in</strong>-Induced Dimerization Activates a Plant Immune Receptor. Science 336,<br />

1160–1164.<br />

93


Liu, Y., and Bassham, D.C. (2012). Autophagy: Pathways for Self-Eat<strong>in</strong>g <strong>in</strong> Plant Cells.<br />

Annual Review of Plant Biology 63, 215–237.<br />

Liu, Y., Xiong, Y., and Bassham, D.C. (2009). Autophagy is required for tolerance of<br />

drought and salt stress <strong>in</strong> plants. Autophagy 5, 954–964.<br />

Lupas et al, V.D., M., and Stock, J. (1991). Predict<strong>in</strong>g coiled coils from prote<strong>in</strong><br />

sequences. Science 252, 1162–1164.<br />

Mak<strong>in</strong>o, A., and Osmond, B. (1991). Effects of Nitrogen Nutrition on Nitrogen<br />

Partition<strong>in</strong>g between Chloroplasts and Mitochondria <strong>in</strong> Pea and Wheat. Plant Physiol. 96,<br />

355–362.<br />

Mao, G., Meng, X., Liu, Y., Zheng, Z., Chen, Z., and Zhang, S. (2011). Phosphorylation<br />

of a WRKY Transcription Factor by Two Pathogen-Responsive MAPKs Drives<br />

Phytoalex<strong>in</strong> Biosyn<strong>the</strong>sis <strong>in</strong> Arabidopsis. Plant Cell 23, 1639–1653.<br />

Melikant, B., Giuliani, C., Halbmayer-Watz<strong>in</strong>a, S., Limmongkon, A., Heberle-Bors, E.,<br />

and Wilson, C. (2004). The Arabidopsis thaliana MEK AtMKK6 activates <strong>the</strong> MAP<br />

k<strong>in</strong>ase AtMPK13. FEBS Letters 576, 5–8.<br />

Meyerowitz, E.M. (2002). Plants Compared to Animals: The Broadest Comparative<br />

Study of Development. Science 295, 1482–1485.<br />

Meyers, B.C., Kozik, A., Griego, A., Kuang, H., and Michelmore, R.W. (2003). Genome-<br />

Wide Analysis of NBS-LRR–Encod<strong>in</strong>g Genes <strong>in</strong> Arabidopsis. Plant Cell 15, 809–834.<br />

Miller, G., Schlauch, K., Tam, R., Cortes, D., Torres, M.A., Shulaev, V., Dangl, J.L., and<br />

Mittler, R. (2009). The Plant NADPH Oxidase RBOHD Mediates Rapid Systemic<br />

Signal<strong>in</strong>g <strong>in</strong> Response to Diverse Stimuli. Sci. Signal. 2, ra45.<br />

Miya, A., Albert, P., Sh<strong>in</strong>ya, T., Desaki, Y., Ichimura, K., Shirasu, K., Narusaka, Y.,<br />

Kawakami, N., Kaku, H., and Shibuya, N. (2007). CERK1, a LysM Receptor K<strong>in</strong>ase, Is<br />

Essential for Chit<strong>in</strong> Elicitor Signal<strong>in</strong>g <strong>in</strong> Arabidopsis. PNAS 104, 19613–19618.<br />

Moffat, C.S., Ingle, R.A., Wathugala, D.L., Saunders, N.J., Knight, H., and Knight, M.R.<br />

(2012). ERF5 and ERF6 Play Redundant Roles as Positive Regulators of JA/Et-Mediated<br />

Defense aga<strong>in</strong>st Botrytis c<strong>in</strong>erea <strong>in</strong> Arabidopsis. PLoS ONE 7, e35995.<br />

Moreno-Hagelsieb, G., and Latimer, K. (2008). Choos<strong>in</strong>g BLAST options for better<br />

detection of orthologs as reciprocal best hits. Bio<strong>in</strong>formatics 24, 319–324.<br />

Mur, L.A.J., Kenton, P., Lloyd, A.J., Ougham, H., and Prats, E. (2008). The<br />

hypersensitive response; <strong>the</strong> centenary is upon us but how much do we know J. Exp. Bot.<br />

59, 501–520.<br />

94


Nakagami, H., Soukupová, H., Schikora, A., Zárský, V., and Hirt, H. (2006). A Mitogenactivated<br />

Prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ase K<strong>in</strong>ase Mediates Reactive Oxygen Species Homeostasis<br />

<strong>in</strong> Arabidopsis. J. Biol. Chem. 281, 38697–38704.<br />

Nakaoka, Y., Miki, T., Fujioka, R., Uehara, R., Tomioka, A., Obuse, C., Kubo, M.,<br />

Hiwatashi, Y., and Goshima, G. (2012). An Inducible RNA Interference System <strong>in</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong> Reveals a Dom<strong>in</strong>ant Role of Augm<strong>in</strong> <strong>in</strong> Phragmoplast Microtubule<br />

Generation. Plant Cell 24, 1478–1493.<br />

Nekrasov, V., Li, J., Batoux, M., Roux, M., Chu, Z.-H., Lacombe, S., Rougon, A., Bittel,<br />

P., Kiss-Papp, M., Ch<strong>in</strong>chilla, D., et al. (2009). Control of <strong>the</strong> pattern-recognition<br />

receptor EFR by an ER prote<strong>in</strong> complex <strong>in</strong> plant immunity. The EMBO Journal 28,<br />

3428–3438.<br />

Nour-Eld<strong>in</strong>, H.H., Hansen, B.G., Nørholm, M.H.H., Jensen, J.K., and Halkier, B.A.<br />

(2006). Advanc<strong>in</strong>g uracil-excision based clon<strong>in</strong>g towards an ideal technique for clon<strong>in</strong>g<br />

PCR fragments. Nucl. Acids Res. 34, e122–e122.<br />

Oerke, E.-C. (2006). Crop losses to pests. The Journal of Agricultural Science 144, 31–43.<br />

Oliver, J., Castro, A., Gaggero, C., Cascón, T., Schmelz, E., Castresana, C., and Ponce de<br />

León, I. (2009). Pythium <strong>in</strong>fection activates conserved plant defense responses <strong>in</strong> <strong>moss</strong>es.<br />

Planta 230, 569–579.<br />

Patel, S., and D<strong>in</strong>esh-Kumar, S.P. (2008). Arabidopsis ATG6 is required to limit <strong>the</strong><br />

pathogen-associated cell death response. Autophagy 4, 20–27.<br />

Pérez-Pérez, M.E., Florencio, F.J., and Crespo, J.L. (2010). Inhibition of Target of<br />

Rapamyc<strong>in</strong> Signal<strong>in</strong>g and Stress Activate Autophagy <strong>in</strong> Chlamydomonas re<strong>in</strong>hardtii.<br />

Plant Physiol. 152, 1874–1888.<br />

Pérez-Pérez, M.E., Lemaire, S.D., and Crespo, J.L. (2012). Reactive Oxygen Species and<br />

Autophagy <strong>in</strong> Plants and Algae. Plant Physiol. 160, 156–164.<br />

Petersen, K., Fiil, B.K., Mundy, J., and Petersen, M. (2008). Downstream targets of<br />

WRKY33. Plant Signal Behav 3, 1033–1034.<br />

Petersen, M., Brodersen, P., Naested, H., Andreasson, E., L<strong>in</strong>dhart, U., Johansen, B.,<br />

Nielsen, H.B., Lacy, M., Aust<strong>in</strong>, M.J., Parker, J.E., et al. (2000). Arabidopsis MAP<br />

K<strong>in</strong>ase 4 Negatively Regulates Systemic Acquired Resistance. Cell 103, 1111–1120.<br />

Petutschnig, E.K., Jones, A.M.E., Serazetd<strong>in</strong>ova, L., Lipka, U., and Lipka, V. (2010). The<br />

Lys<strong>in</strong> Motif Receptor-like K<strong>in</strong>ase (LysM-RLK) CERK1 Is a Major Chit<strong>in</strong>-b<strong>in</strong>d<strong>in</strong>g<br />

Prote<strong>in</strong> <strong>in</strong> Arabidopsis thaliana and Subject to Chit<strong>in</strong>-<strong>in</strong>duced Phosphorylation♦. J Biol<br />

Chem 285, 28902–28911.<br />

95


Ponce de León, I.P., Sanz, A., Hamberg, M., and Castresana, C. (2002). Involvement of<br />

<strong>the</strong> Arabidopsisα-DOX1 fatty acid dioxygenase <strong>in</strong> protection aga<strong>in</strong>st oxidative stress and<br />

cell death. The Plant Journal 29, 61–72.<br />

Ponce de León, I., Oliver, J., Castro, A., Gaggero, C., Bentancor, M., and Vidal, S.<br />

(2007). Erw<strong>in</strong>ia carotovora elicitors and Botrytis c<strong>in</strong>erea activate defense responses <strong>in</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>. BMC Plant Biology 7, 52.<br />

Ponce De León, I., Schmelz, E.A., Gaggero, C., Castro, A., Álvarez, A., and Montesano,<br />

M. (2012). <strong>Physcomitrella</strong> <strong>patens</strong> activates re<strong>in</strong>forcement of <strong>the</strong> cell wall, programmed<br />

cell death and accumulation of evolutionary conserved defence signals, such as salicylic<br />

acid and 12‐oxo‐phytodienoic acid, but not jasmonic acid, upon Botrytis c<strong>in</strong>erea<br />

<strong>in</strong>fection. Molecular Plant Pathology 13, 960–974.<br />

Prigge, M.J., and Bezanilla, M. (2010). Evolutionary crossroads <strong>in</strong> developmental<br />

biology: <strong>Physcomitrella</strong> <strong>patens</strong>. Development 137, 3535–3543.<br />

Pöpp<strong>in</strong>g, B., Gibbons, T., and Watson, M.D. (1996). The Pisum sativum MAP k<strong>in</strong>ase<br />

homologue (PsMAPK) rescues <strong>the</strong> Saccharomyces cerevisiae hog1 deletion mutant under<br />

conditions of high osmotic stress. Plant Mol. Biol. 31, 355–363.<br />

Qiu, J.-L., Fiil, B.K., Petersen, K., Nielsen, H.B., Botanga, C.J., Thorgrimsen, S., Palma,<br />

K., Suarez-Rodriguez, M.C., Sandbech-Clausen, S., Lichota, J., et al. (2008a).<br />

Arabidopsis MAP k<strong>in</strong>ase 4 regulates gene expression through transcription factor release<br />

<strong>in</strong> <strong>the</strong> nucleus. The EMBO Journal 27, 2214–2221.<br />

Qiu, J.-L., Zhou, L., Yun, B.-W., Nielsen, H.B., Fiil, B.K., Petersen, K., MacK<strong>in</strong>lay, J.,<br />

Loake, G.J., Mundy, J., and Morris, P.C. (2008b). Arabidopsis Mitogen-Activated<br />

Prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ases MKK1 and MKK2 Have Overlapp<strong>in</strong>g Functions <strong>in</strong> Defense<br />

Signal<strong>in</strong>g Mediated by MEKK1, MPK4, and MKS1. Plant Physiol. 148, 212–222.<br />

Rasmussen, M.W., Roux, M., Petersen, M., and Mundy, J. (2012). MAP K<strong>in</strong>ase Cascades<br />

<strong>in</strong> Arabidopsis Innate <strong>Immunity</strong>. Front Plant Sci 3, 169.<br />

Ren, D., Liu, Y., Yang, K.-Y., Han, L., Mao, G., Glazebrook, J., and Zhang, S. (2008). A<br />

fungal-responsive MAPK cascade regulates phytoalex<strong>in</strong> biosyn<strong>the</strong>sis <strong>in</strong> Arabidopsis.<br />

Rens<strong>in</strong>g, S., Fritzowsky, D., Lang, D., and Reski, R. (2005). Prote<strong>in</strong> encod<strong>in</strong>g genes <strong>in</strong> an<br />

ancient plant: analysis of codon usage, reta<strong>in</strong>ed genes and splice sites <strong>in</strong> a <strong>moss</strong>,<br />

<strong>Physcomitrella</strong> <strong>patens</strong>. BMC Genomics 6, 43.<br />

Rens<strong>in</strong>g, S., Ick, J., Fawcett, J., Lang, D., Zimmer, A., Peer, Y.V. de, and Reski, R.<br />

(2007). An ancient genome duplication contributed to <strong>the</strong> abundance of metabolic genes<br />

<strong>in</strong> <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. BMC Evolutionary Biology 7, 130.<br />

Rens<strong>in</strong>g, S.A., Lang, D., Zimmer, A.D., Terry, A., Salamov, A., Shapiro, H., Nishiyama,<br />

T., Perroud, P.-F., L<strong>in</strong>dquist, E.A., Kamisugi, Y., et al. (2008). The <strong>Physcomitrella</strong><br />

96


Genome Reveals Evolutionary Insights <strong>in</strong>to <strong>the</strong> Conquest of Land by Plants. Science 319,<br />

64–69.<br />

Robatzek, S., Ch<strong>in</strong>chilla, D., and Boller, T. (2006). Ligand-<strong>in</strong>duced endocytosis of <strong>the</strong><br />

pattern recognition receptor FLS2 <strong>in</strong> Arabidopsis. Genes Dev. 20, 537–542.<br />

Roux, M., Schwess<strong>in</strong>ger, B., Albrecht, C., Ch<strong>in</strong>chilla, D., Jones, A., Holton, N.,<br />

Mal<strong>in</strong>ovsky, F.G., Tör, M., De Vries, S., and Zipfel, C. (2011). The Arabidopsis Leuc<strong>in</strong>e-<br />

Rich Repeat Receptor–Like K<strong>in</strong>ases BAK1/SERK3 and BKK1/SERK4 Are Required for<br />

Innate <strong>Immunity</strong> to Hemibiotrophic and Biotrophic Pathogens. Plant Cell 23, 2440–2455.<br />

Rushton, P.J., Somssich, I.E., R<strong>in</strong>gler, P., and Shen, Q.J. (2010). WRKY transcription<br />

factors. Trends <strong>in</strong> Plant Science 15, 247–258.<br />

Schaefer, D., Zryd, J.-P., Knight, C.D., and Cove, D.J. (1991). Stable transformation of<br />

<strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. Molecular and General Genetics MGG 226, 418–424.<br />

Schaefer, D.G., and Zrÿd, J.-P. (1997). Efficient gene target<strong>in</strong>g <strong>in</strong> <strong>the</strong> <strong>moss</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>. The Plant Journal 11, 1195–1206.<br />

Schuhegger, R., Nafisi, M., Mansourova, M., Petersen, B.L., Olsen, C.E., Svatoš, A.,<br />

Halkier, B.A., and Glawischnig, E. (2006). CYP71B15 (PAD3) Catalyzes <strong>the</strong> F<strong>in</strong>al Step<br />

<strong>in</strong> Camalex<strong>in</strong> Biosyn<strong>the</strong>sis. Plant Physiol. 141, 1248–1254.<br />

Schuler, M.A., Duan, H., Bilg<strong>in</strong>, M., and Ali, S. (2006). Arabidopsis cytochrome P450s<br />

through <strong>the</strong> look<strong>in</strong>g glass: a w<strong>in</strong>dow on plant biochemistry. Phytochemistry Reviews 5,<br />

205–237.<br />

Schween, G., Egener, T., Fritzowsky, D., Granado, J., Guitton, M.-C., Hartmann, N.,<br />

Hohe, A., Holtorf, H., Lang, D., Lucht, J.M., et al. (2005). Large-Scale Analysis of 73<br />

329 <strong>Physcomitrella</strong> Plants Transformed with Different Gene Disruption Libraries:<br />

Production Parameters and Mutant Phenotypes. Plant Biology 7, 228–237.<br />

Segonzac, C., Feike, D., Gimenez-Ibanez, S., Hann, D.R., Zipfel, C., and Rathjen, J.P.<br />

(2011). Hierarchy and Roles of Pathogen-Associated Molecular Pattern-Induced<br />

Responses <strong>in</strong> Nicotiana benthamiana. Plant Physiol. 156, 687–699.<br />

Segonzac, C., and Zipfel, C. (2011). Activation of plant pattern-recognition receptors by<br />

bacteria. Current Op<strong>in</strong>ion <strong>in</strong> Microbiology 14, 54–61.<br />

Shen, Q.-H., and Schulze-Lefert, P. (2007). Rumble <strong>in</strong> <strong>the</strong> nuclear jungle:<br />

compartmentalization, traffick<strong>in</strong>g, and nuclear action of plant immune receptors. The<br />

EMBO Journal 26, 4293–4301.<br />

Shimizu, T., Nakano, T., Takamizawa, D., Desaki, Y., Ishii-M<strong>in</strong>ami, N., Nishizawa, Y.,<br />

M<strong>in</strong>ami, E., Okada, K., Yamane, H., Kaku, H., et al. (2010). Two LysM receptor<br />

molecules, CEBiP and OsCERK1, cooperatively regulate chit<strong>in</strong> elicitor signal<strong>in</strong>g <strong>in</strong> rice.<br />

Plant J 64, 204–214.<br />

97


Sh<strong>in</strong>ya, T., Motoyama, N., Ikeda, A., Wada, M., Kamiya, K., Hayafune, M., Kaku, H.,<br />

and Shibuya, N. (2012). Functional Characterization of CEBiP and CERK1 Homologs <strong>in</strong><br />

Arabidopsis and Rice Reveals <strong>the</strong> Presence of Different Chit<strong>in</strong> Receptor Systems <strong>in</strong><br />

Plants. Plant Cell Physiol 53, 1696–1706.<br />

Shirasu, K. (2009). The HSP90-SGT1 Chaperone Complex for NLR Immune Sensors.<br />

Annual Review of Plant Biology 60, 139–164.<br />

Shiu, S.-H., Karlowski, W.M., Pan, R., Tzeng, Y.-H., Mayer, K.F.X., and Li, W.-H.<br />

(2004). Comparative Analysis of <strong>the</strong> Receptor-Like K<strong>in</strong>ase Family <strong>in</strong> Arabidopsis and<br />

Rice. Plant Cell 16, 1220–1234.<br />

S<strong>in</strong>ha, A.K., Jaggi, M., Raghuram, B., and Tuteja, N. (2011). Mitogen-activated prote<strong>in</strong><br />

k<strong>in</strong>ase signal<strong>in</strong>g <strong>in</strong> plants under abiotic stress. Plant Signal Behav 6, 196–203.<br />

Slavikova, S., Ufaz, S., Av<strong>in</strong>-Wittenberg, T., Levanony, H., and Galili, G. (2008). An<br />

autophagy-associated Atg8 prote<strong>in</strong> is <strong>in</strong>volved <strong>in</strong> <strong>the</strong> responses of Arabidopsis seedl<strong>in</strong>gs<br />

to hormonal controls and abiotic stresses. J. Exp. Bot. 59, 4029–4043.<br />

Šmídková, M., Holá, M., and Angelis, K.J. (2010). Efficient biolistic transformation of<br />

<strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. Biologia Plantarum 54, 777–780.<br />

Son, G.H., Wan, J., Kim, H.J., Nguyen, X.C., Chung, W.S., Hong, J.C., and Stacey, G.<br />

(2012). Ethylene-responsive element-b<strong>in</strong>d<strong>in</strong>g factor 5, ERF5, is <strong>in</strong>volved <strong>in</strong> chit<strong>in</strong><strong>in</strong>duced<br />

<strong>in</strong>nate immunity response. Mol. Plant Microbe Interact. 25, 48–60.<br />

Specht, K.M., and Shokat, K.M. (2002). The emerg<strong>in</strong>g power of chemical genetics.<br />

Current Op<strong>in</strong>ion <strong>in</strong> Cell Biology 14, 155–159.<br />

Struhl, K., St<strong>in</strong>chcomb, D.T., Scherer, S., and Davis, R.W. (1979). High-frequency<br />

transformation of yeast: autonomous replication of hybrid DNA molecules. PNAS 76,<br />

1035–1039.<br />

Stumpe, M., Göbel, C., Falt<strong>in</strong>, B., Beike, A.K., Hause, B., Himmelsbach, K., Bode, J.,<br />

Kramell, R., Wasternack, C., Frank, W., et al. (2010). The <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong><br />

conta<strong>in</strong>s cyclopentenones but no jasmonates: mutations <strong>in</strong> allene oxide cyclase lead to<br />

reduced fertility and altered sporophyte morphology. New Phytologist 188, 740–749.<br />

Suarez-Rodriguez, C.M., Petersen, M., and Mundy, J. (2010). Mitogen-Activated Prote<strong>in</strong><br />

K<strong>in</strong>ase Signal<strong>in</strong>g <strong>in</strong> Plants. Annual Review of Plant Biology 61, 621–649.<br />

Suarez-Rodriguez, M.C., Adams-Phillips, L., Liu, Y., Wang, H., Su, S.-H., Jester, P.J.,<br />

Zhang, S., Bent, A.F., and Krysan, P.J. (2007). MEKK1 Is Required for Flg22-Induced<br />

MPK4 Activation <strong>in</strong> Arabidopsis Plants. Plant Physiol. 143, 661–669.<br />

Suzuki, N., Koussevitzky, S., Mittler, R., and Miller, G. (2012). ROS and redox<br />

signall<strong>in</strong>g <strong>in</strong> <strong>the</strong> response of plants to abiotic stress. Plant, Cell & Environment 35, 259–<br />

270.<br />

98


Swiderski, M.R., Birker, D., and Jones, J.D.G. (2009). The TIR Doma<strong>in</strong> of TIR-NB-LRR<br />

Resistance Prote<strong>in</strong>s Is a Signal<strong>in</strong>g Doma<strong>in</strong> Involved <strong>in</strong> Cell Death Induction. Molecular<br />

Plant-Microbe Interactions 22, 157–165.<br />

Takahashi, A., Casais, C., Ichimura, K., and Shirasu, K. (2003). HSP90 <strong>in</strong>teracts with<br />

RAR1 and SGT1 and is essential for RPS2-mediated disease resistance <strong>in</strong> Arabidopsis.<br />

PNAS 100, 11777–11782.<br />

Tamel<strong>in</strong>g, W.I.L., Vossen, J.H., Albrecht, M., Lengauer, T., Berden, J.A., Har<strong>in</strong>g, M.A.,<br />

Cornelissen, B.J.C., and Takken, F.L.W. (2006). Mutations <strong>in</strong> <strong>the</strong> NB-ARC Doma<strong>in</strong> of I-<br />

2 That Impair ATP Hydrolysis Cause Autoactivation. Plant Physiol. 140, 1233–1245.<br />

Tarr, D.E., and Alexander, H. (2009). TIR-NBS-LRR genes are rare <strong>in</strong> monocots:<br />

evidence from diverse monocot orders. BMC Research Notes 2, 197.<br />

Teige, M., Scheikl, E., Eulgem, T., Dóczi, R., Ichimura, K., Sh<strong>in</strong>ozaki, K., Dangl, J.L.,<br />

and Hirt, H. (2004). The MKK2 pathway mediates cold and salt stress signal<strong>in</strong>g <strong>in</strong><br />

Arabidopsis. Mol. Cell 15, 141–152.<br />

The Arabidopsis Genome Initiative (2000). Analysis of <strong>the</strong> genome sequence of <strong>the</strong><br />

flower<strong>in</strong>g plant Arabidopsis thaliana. Nature 408, 796–815.<br />

Thelander, M., Olsson, T., and Ronne, H. (2005). Effect of <strong>the</strong> energy supply on<br />

filamentous growth and development <strong>in</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. J. Exp. Bot. 56, 653–662.<br />

Thompson, A.R., Doell<strong>in</strong>g, J.H., Suttangkakul, A., and Vierstra, R.D. (2005). Autophagic<br />

Nutrient Recycl<strong>in</strong>g <strong>in</strong> Arabidopsis Directed by <strong>the</strong> ATG8 and ATG12 Conjugation<br />

Pathways. Plant Physiol. 138, 2097–2110.<br />

Thordal-Christensen, H., Zhang, Z., Wei, Y., and Coll<strong>in</strong>ge, D.B. (1997). Subcellular<br />

localization of H2O2 <strong>in</strong> plants. H2O2 accumulation <strong>in</strong> papillae and hypersensitive<br />

response dur<strong>in</strong>g <strong>the</strong> barley—powdery mildew <strong>in</strong>teraction. The Plant Journal 11, 1187–<br />

1194.<br />

Torres, M.A., and Dangl, J.L. (2005). Functions of <strong>the</strong> respiratory burst oxidase <strong>in</strong> biotic<br />

<strong>in</strong>teractions, abiotic stress and development. Current Op<strong>in</strong>ion <strong>in</strong> Plant Biology 8, 397–<br />

403.<br />

Vellosillo, T., Vicente, J., Kulasekaran, S., Hamberg, M., and Castresana, C. (2010).<br />

Emerg<strong>in</strong>g Complexity <strong>in</strong> Reactive Oxygen Species Production and Signal<strong>in</strong>g dur<strong>in</strong>g <strong>the</strong><br />

Response of Plants to Pathogens. Plant Physiol. 154, 444–448.<br />

Vij, S., Giri, J., Dansana, P.K., Kapoor, S., and Tyagi, A.K. (2008). The Receptor-Like<br />

Cytoplasmic K<strong>in</strong>ase (OsRLCK) Gene Family <strong>in</strong> Rice: Organization, Phylogenetic<br />

Relationship, and Expression dur<strong>in</strong>g Development and Stress. Mol. Plant 1, 732–750.<br />

99


Wada, S., Ishida, H., Izumi, M., Yoshimoto, K., Ohsumi, Y., Mae, T., and Mak<strong>in</strong>o, A.<br />

(2009). Autophagy Plays a Role <strong>in</strong> Chloroplast Degradation dur<strong>in</strong>g Senescence <strong>in</strong><br />

Individually Darkened Leaves. Plant Physiol. 149, 885–893.<br />

Wan, J., Zhang, S., and Stacey, G. (2004). Activation of a mitogen-activated prote<strong>in</strong><br />

k<strong>in</strong>ase pathway <strong>in</strong> Arabidopsis by chit<strong>in</strong>. Molecular Plant Pathology 5, 125–135.<br />

Wang, H., Ngwenyama, N., Liu, Y., Walker, J.C., and Zhang, S. (2007). Stomatal<br />

Development and Pattern<strong>in</strong>g Are Regulated by Environmentally Responsive Mitogen-<br />

Activated Prote<strong>in</strong> K<strong>in</strong>ases <strong>in</strong> Arabidopsis. PLANT CELL 19, 63–73.<br />

Wang, Y., Chen, J.-Q., Araki, H., J<strong>in</strong>g, Z., Jiang, K., Shen, J., Tian, D., and Zhou, T.<br />

(2004). Genome-wide identification of NBS genes <strong>in</strong> japonica rice reveals significant<br />

expansion of divergent non-TIR NBS-LRR genes. Molecular Genetics and Genomics 271,<br />

402–415.<br />

Wees, V., C.m, S., Chang, H.-S., Zhu, T., and Glazebrook, J. (2003). Characterization of<br />

<strong>the</strong> Early Response of Arabidopsis to Alternaria Brassicicola Infection Us<strong>in</strong>g Expression<br />

Profil<strong>in</strong>g. Plant Physiol. 132, 606–617.<br />

Wellman, C.H., Osterloff, P.L., and Mohiudd<strong>in</strong>, U. (2003). Fragments of <strong>the</strong> earliest land<br />

plants. Nature 425, 282–285.<br />

Willmann, R., Lajunen, H.M., Erbs, G., Newman, M.-A., Kolb, D., Tsuda, K., Katagiri,<br />

F., Fliegmann, J., Bono, J.-J., Cullimore, J.V., et al. (2011). Arabidopsis lys<strong>in</strong>-motif<br />

prote<strong>in</strong>s LYM1 LYM3 CERK1 mediate bacterial peptidoglycan sens<strong>in</strong>g and immunity to<br />

bacterial <strong>in</strong>fection. PNAS 108, 19824–19829.<br />

Wurz<strong>in</strong>ger, B., Mair, A., Pfister, B., and Teige, M. (2011). Cross-talk of calciumdependent<br />

prote<strong>in</strong> k<strong>in</strong>ase and MAP k<strong>in</strong>ase signal<strong>in</strong>g. Plant Signal Behav 6, 8–12.<br />

Xie, Z., and Klionsky, D.J. (2007). Autophagosome formation: core mach<strong>in</strong>ery and<br />

adaptations. Nature Cell Biology 9, 1102–1109.<br />

Xiong, Y., Contento, A.L., Nguyen, P.Q., and Bassham, D.C. (2007). Degradation of<br />

Oxidized Prote<strong>in</strong>s by Autophagy dur<strong>in</strong>g Oxidative Stress <strong>in</strong> Arabidopsis. Plant Physiol<br />

143, 291–299.<br />

Xue, J.-Y., Wang, Y., Wu, P., Wang, Q., Yang, L.-T., Pan, X.-H., Wang, B., and Chen,<br />

J.-Q. (2012). A Primary Survey on Bryophyte Species Reveals Two Novel Classes of<br />

Nucleotide-B<strong>in</strong>d<strong>in</strong>g Site (NBS) Genes. PLoS ONE 7, e36700.<br />

Yang, Z., and Klionsky, D.J. (2010). Eaten alive: a history of macroautophagy. Nature<br />

Cell Biology 12, 814–822.<br />

Yano, K., Suzuki, T., and Moriyasu, Y. (2007). Constitutive Autophagy <strong>in</strong> Plant Root<br />

Cells. Autophagy 3, 360–362.<br />

100


Ülker, B., and Somssich, I.E. (2004). WRKY transcription factors: from DNA b<strong>in</strong>d<strong>in</strong>g<br />

towards biological function. Current Op<strong>in</strong>ion <strong>in</strong> Plant Biology 7, 491–498.<br />

Yoo, S.-D., Cho, Y.-H., Tena, G., Xiong, Y., and Sheen, J. (2008). Dual control of<br />

nuclear EIN3 by bifurcate MAPK cascades <strong>in</strong> C2H4 signall<strong>in</strong>g. Nature 451, 789–795.<br />

Yoshimoto, K., Hanaoka, H., Sato, S., Kato, T., Tabata, S., Noda, T., and Ohsumi, Y.<br />

(2004). Process<strong>in</strong>g of ATG8s, Ubiquit<strong>in</strong>-Like Prote<strong>in</strong>s, and Their Deconjugation by<br />

ATG4s Are Essential for Plant Autophagy. Plant Cell 16, 2967–2983.<br />

Yoshimoto, K., Jikumaru, Y., Kamiya, Y., Kusano, M., Consonni, C., Panstruga, R.,<br />

Ohsumi, Y., and Shirasu, K. (2009). Autophagy Negatively Regulates Cell Death by<br />

Controll<strong>in</strong>g NPR1-Dependent Salicylic Acid Signal<strong>in</strong>g dur<strong>in</strong>g Senescence and <strong>the</strong> Innate<br />

Immune Response <strong>in</strong> Arabidopsis. Plant Cell 21, 2914–2927.<br />

Zeng, Q., Chen, J.-G., and Ellis, B.E. (2011). AtMPK4 is required for male-specific<br />

meiotic cytok<strong>in</strong>esis <strong>in</strong> Arabidopsis. The Plant Journal 67, 895–906.<br />

Zhang, J., Shao, F., Li, Y., Cui, H., Chen, L., Li, H., Zou, Y., Long, C., Lan, L., Chai, J.,<br />

et al. (2007). A Pseudomonas syr<strong>in</strong>gae Effector Inactivates MAPKs to Suppress PAMP-<br />

Induced <strong>Immunity</strong> <strong>in</strong> Plants. Cell Host & Microbe 1, 175–185.<br />

Zhang, N., Castlebury, L.A., Miller, A.N., Huhndorf, S.M., Schoch, C.L., Seifert, K.A.,<br />

Rossman, A.Y., Rogers, J.D., Kohlmeyer, J., Volkmann-Kohlmeyer, B., et al. (2006). An<br />

overview of <strong>the</strong> systematics of <strong>the</strong> Sordariomycetes based on a four-gene phylogeny.<br />

Mycologia 98, 1076–1087.<br />

Zhang, Z., Wu, Y., Gao, M., Zhang, J., Kong, Q., Liu, Y., Ba, H., Zhou, J., and Zhang, Y.<br />

(2012). Disruption of PAMP-Induced MAP K<strong>in</strong>ase Cascade by a Pseudomonas syr<strong>in</strong>gae<br />

Effector Activates Plant <strong>Immunity</strong> Mediated by <strong>the</strong> NB-LRR Prote<strong>in</strong> SUMM2. Cell Host<br />

& Microbe 11, 253–263.<br />

Zheng, Z., Qamar, S.A., Chen, Z., and Mengiste, T. (2006). Arabidopsis WRKY33<br />

transcription factor is required for resistance to necrotrophic fungal pathogens. The Plant<br />

Journal 48, 592–605.<br />

Zipfel, C., Kunze, G., Ch<strong>in</strong>chilla, D., Caniard, A., Jones, J.D.G., Boller, T., and Felix, G.<br />

(2006). Perception of <strong>the</strong> Bacterial PAMP EF-Tu by <strong>the</strong> Receptor EFR Restricts<br />

Agrobacterium-Mediated Transformation. Cell 125, 749–760.<br />

Zipfel, C., Robatzek, S., Navarro, L., Oakeley, E.J., Jones, J.D.G., Felix, G., and Boller,<br />

T. (2004). Bacterial disease resistance <strong>in</strong> Arabidopsis through flagell<strong>in</strong> perception. Nature<br />

428, 764–767.<br />

Zurbriggen, M.D., Carrillo, N., and Hajirezaei, M.-R. (2010). ROS signal<strong>in</strong>g <strong>in</strong> <strong>the</strong><br />

hypersensitive response. Plant Signal Behav 5, 393–396.<br />

101


Manuscript 1<br />

Submitted to PLOS Pathogen<br />

102


A MAP k<strong>in</strong>ase regulates <strong>in</strong>nate immunity triggered by pathogen<br />

associated molecular patterns <strong>in</strong> <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong><br />

Simon Bressendorff 1 , Inés Ponce de León 2 , Raquel Azevedo 1 and Morten Petersen 1*<br />

1 Department of Molecular Biology, University of Copenhagen, Copenhagen, Denmark<br />

2 Departamento de Biología Molecular, Instituto de Investigaciones Biológicas Clemente<br />

Estable, Montevideo, Uruguay<br />

* E-mail: shutko@bio.ku.dk<br />

Abstract<br />

In eukaryotes, <strong>the</strong> activation of pattern recognition receptors by pathogen-associated<br />

molecular patterns (PAMPs), such as fungal chit<strong>in</strong> or bacterial flagell<strong>in</strong>, elicit PAMPtriggered<br />

immunity via MAP k<strong>in</strong>ase cascades. Given <strong>the</strong> agronomic importance of<br />

disease resistance, much is known about such immune pathways <strong>in</strong> higher plants and<br />

crops. In contrast, very little is known about how early land plants evolved immune<br />

responses. Here we provide a primary example of a MAP k<strong>in</strong>ase that is required for<br />

pathogen resistance <strong>in</strong> <strong>the</strong> phylogenetically ancient <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. We<br />

show that P. <strong>patens</strong> MAP k<strong>in</strong>ase 4A is phosphorylated <strong>in</strong> response to fungal chit<strong>in</strong> or a<br />

complement of bacterial PAMPs, but not <strong>in</strong> response to a PAMP derived from bacterial<br />

flagell<strong>in</strong> (flg22 peptide). We also show that knock-out of MAP k<strong>in</strong>ase 4A renders <strong>the</strong><br />

<strong>moss</strong> more susceptible to <strong>the</strong> pathogenic fungi Botrytis c<strong>in</strong>erea and Alternaria<br />

brassisicola, and that defense-related <strong>moss</strong> transcripts fail to accumulate <strong>in</strong> <strong>the</strong> knock-out<br />

103


mutant upon treatment with fungal chitosan. While related MAP k<strong>in</strong>ases <strong>in</strong> <strong>the</strong> higher<br />

plant model Arabidopsis thaliana are activated both by pathogen <strong>in</strong>oculation and by<br />

abiotic stress, we did not detect activation of MAP k<strong>in</strong>ase 4A or any o<strong>the</strong>r P. <strong>patens</strong> MAP<br />

k<strong>in</strong>ase by several abiotic stresses. Signal transduction via MAP k<strong>in</strong>ase 4A may <strong>the</strong>refore<br />

be specific to PAMP-triggered immunity, and <strong>the</strong> <strong>moss</strong> may use o<strong>the</strong>r signal<strong>in</strong>g<br />

components to respond to abiotic stresses.<br />

Introduction<br />

In eukaryotes, <strong>in</strong>nate immune systems play predom<strong>in</strong>ant roles <strong>in</strong> pathogen detection and<br />

elim<strong>in</strong>ation. Plants and animals have evolved surface-localized surveillance systems<br />

consist<strong>in</strong>g of pattern-recognition receptors (PRRs) that recognize evolutionarily<br />

conserved pathogen-associated molecular patterns (PAMPs) [1]. Examples of PAMP<br />

recognition <strong>in</strong> plants <strong>in</strong>clude <strong>the</strong> perception of bacterial flagell<strong>in</strong> by <strong>the</strong> receptor-like<br />

k<strong>in</strong>ase (RLK) FLS2, perception of bacterial elongation factor Tu by <strong>the</strong> closely related<br />

EFR, and fungal chit<strong>in</strong> by <strong>the</strong> LysM RLK CERK1 [2,3,4].<br />

In higher eukaryotes, MAP k<strong>in</strong>ase cascades transmit and amplify signals from<br />

PRRs to <strong>the</strong> nucleus. MAP k<strong>in</strong>ase cascades consist of a MAP k<strong>in</strong>ase k<strong>in</strong>ase k<strong>in</strong>ase<br />

(MPKKK), a MAP k<strong>in</strong>ase k<strong>in</strong>ase (MPKK) and a MAP k<strong>in</strong>ase (MPK) and signals from<br />

upstream components [5]. In A. thaliana, four MPKs (MPK3/4/6 and 11) are <strong>in</strong>volved <strong>in</strong><br />

signall<strong>in</strong>g upon PRR activation [6,7]. In addition to <strong>the</strong>ir roles <strong>in</strong> immunity, <strong>the</strong>se MPKs<br />

have also been implicated <strong>in</strong> responses to abiotic stresses and <strong>in</strong> developmental processes<br />

[5]. However, <strong>the</strong> contribution and importance of <strong>the</strong> s<strong>in</strong>gle MPKs <strong>in</strong> A. thaliana<br />

104


immunity is unclear, and some of <strong>the</strong>ir loss-of-function mutants exhibit seedl<strong>in</strong>g lethality,<br />

<strong>in</strong>appropriate defense activation and dwarfism which complicates <strong>the</strong>ir analysis [8,9,10].<br />

The <strong>moss</strong> P. <strong>patens</strong> is a nonvascular bryophyte and represents an early branch of land<br />

plants. Although P. <strong>patens</strong> shares basic processes with flower<strong>in</strong>g plants, <strong>the</strong> two l<strong>in</strong>eages<br />

shared a common ancestor at least 450 million years ago [11]. Recent studies have shown<br />

that P. <strong>patens</strong> may be a good model for study<strong>in</strong>g plant pathogen <strong>in</strong>teractions. It is<br />

susceptible to a range of pathogens <strong>in</strong>clud<strong>in</strong>g fungi, oomycetes and bacteria which<br />

activate several defense mechanisms <strong>in</strong>clud<strong>in</strong>g production of apoplastic reactive oxygen<br />

species (ROS), cell wall fortifications and expression of defense related genes [12,13,14].<br />

However, little is known about how P. <strong>patens</strong> perceives pathogens and activates<br />

responses, and <strong>the</strong> functions of PRRs and MPKs <strong>in</strong> <strong>moss</strong> signal<strong>in</strong>g cascades has not been<br />

studied.<br />

The A. thaliana genome encodes 20 MPKs whereas a recent study on <strong>the</strong><br />

phylogeny of MPKs showed that <strong>the</strong>re are some 8 MPKs <strong>in</strong> P. <strong>patens</strong> [15]. This limited<br />

set of <strong>moss</strong> MPKs may represent a “basal” set of MAP k<strong>in</strong>ase pathways, and targeted<br />

deletion of s<strong>in</strong>gle <strong>moss</strong> MPKs via homologous recomb<strong>in</strong>ation may help to elucidate <strong>the</strong><br />

complexity of MPK activation <strong>in</strong> higher plants. Study<strong>in</strong>g <strong>the</strong> repertoire of PAMPs that<br />

can activate MPKs and <strong>in</strong>duce <strong>in</strong>nate immune responses <strong>in</strong> organisms that diverged <strong>in</strong> <strong>the</strong><br />

l<strong>in</strong>eage to vascular plants will also help us understand how plant surveillance systems<br />

adapted to different types of microbes.<br />

A. thaliana MPK4 (AtMPK4) has been proposed to function <strong>in</strong> a cascade(s) that<br />

<strong>in</strong>cludes <strong>the</strong> MAP k<strong>in</strong>ase k<strong>in</strong>ases MKK1 and MKK2 and <strong>the</strong> MPKKK MEKK1[16,17].<br />

105


Different abiotic stresses can activate AtMPK4, and PAMPs activate this k<strong>in</strong>ase and<br />

release <strong>the</strong> WRKY33 transcription factor from AtMPK4 complexes [17,18,19]. In<br />

addition, A. thaliana mpk4 loss-of-function mutants exhibit extreme dwarfism and<br />

ectopically express defense responses, suggest<strong>in</strong>g that AtMPK4 functions as a negative<br />

regulator of defense [9]. Whereas a recent report implicated AtMPK4 as a negative<br />

regulator of <strong>the</strong> SUMM1 MPKKK [20], ano<strong>the</strong>r report partially expla<strong>in</strong>ed how AtMPK4<br />

could function as a positive regulator of immunity [10]. These differ<strong>in</strong>g observations<br />

<strong>in</strong>dicate that studies of MPKs <strong>in</strong> different evolutionary models may elucidate <strong>the</strong>ir basal<br />

and subsequently acquired functions.<br />

To <strong>in</strong>itiate studies on <strong>the</strong> <strong>in</strong>nate immune responses of P. <strong>patens</strong>, and to address<br />

questions about MPKs <strong>in</strong> basal land plants, we identified <strong>moss</strong> homologs of A. thaliana<br />

MPK4. Interest<strong>in</strong>gly, <strong>the</strong> targeted knock out mutant of one P. <strong>patens</strong> MPK4 homolog<br />

(ΔPpMPK4A) had no obvious phenotypic differences compared to wild type plants and<br />

did not exhibit constitutive defense activation. However, we found that <strong>the</strong> PpMPK4A<br />

k<strong>in</strong>ase is phosphorylated, and presumably activated, follow<strong>in</strong>g exposure to bacterial and<br />

fungal PAMPs. Activation of PpMPK4A contributes to immunity s<strong>in</strong>ce ΔPpMPK4A<br />

mutant is more susceptible to <strong>the</strong> necrotrophic fungi. Fur<strong>the</strong>rmore, we could not detect<br />

activation of PpMPK4A or o<strong>the</strong>r MPKs <strong>in</strong> P. <strong>patens</strong> <strong>in</strong> response to various abiotic<br />

stresses. These results <strong>in</strong>dicate that PpMPK4A specifically functions <strong>in</strong> pathways below<br />

PAMP reception, and that MPK functions <strong>in</strong> abiotic stress responses may have appeared<br />

at later stages of plant evolution, or <strong>the</strong>se traits were lost <strong>in</strong> <strong>the</strong> <strong>moss</strong> l<strong>in</strong>eage.<br />

Results<br />

106


Two MPKs <strong>in</strong> P. <strong>patens</strong> are homologs of AtMPK4<br />

Dóczi et al. (2012) recently published <strong>the</strong> phylogenetic relationship of MPKs from ten<br />

plant genomes <strong>in</strong>clud<strong>in</strong>g P. <strong>patens</strong> [15]. In P. <strong>patens</strong> <strong>the</strong>re are eight putative MPKs, two<br />

with homology to group B MPKs <strong>in</strong> A. thaliana which <strong>in</strong>clude AtMPK4/5/11/12 and13<br />

[21]. Dóczi et al. (2012) could not establish a clear orthologous relationships between<br />

MPKs from A. thaliana and basal land plants <strong>in</strong>clud<strong>in</strong>g P. <strong>patens</strong>. None<strong>the</strong>less, <strong>the</strong> two<br />

P. <strong>patens</strong> MPKs <strong>in</strong> group B had <strong>the</strong> highest sequence similarity to AtMPK4, and we<br />

<strong>the</strong>refore named <strong>the</strong>m PpMPK4A (Phypa 446407) and PpMPK4B (Phypa 435424). A<br />

pairwise sequence alignment us<strong>in</strong>g <strong>the</strong> BLOSSUM30 matrix showed 83.6% similarity<br />

between AtMPK4 and PpMPK4A and 86.1% similarity between AtMPK4 and PpMPK4B,<br />

while <strong>the</strong> similarity between <strong>the</strong> two P. <strong>patens</strong> homologs is 84.8% (Figure 1S).<br />

PpMPK4A accumulates <strong>in</strong> <strong>the</strong> presence of chitosan<br />

Treatments that affect <strong>the</strong> biochemical activity of regulatory prote<strong>in</strong>s, <strong>in</strong>clud<strong>in</strong>g MAP<br />

k<strong>in</strong>ases, can also alter <strong>the</strong> expression of <strong>the</strong>ir correspond<strong>in</strong>g mRNAs [22]. P. <strong>patens</strong><br />

responds to fungal-derived elicitors or PAMPs like chitosan [23], but lacks obvious<br />

homologs of <strong>the</strong> well-studied PAMP receptors FLS2 and EFR that b<strong>in</strong>d <strong>the</strong> bacterial<br />

PAMPs flagell<strong>in</strong> and EF-Tu [24]. We <strong>the</strong>refore exam<strong>in</strong>ed <strong>the</strong> expression of PpMPK4A<br />

and PpMPK4B upon treatment of P. <strong>patens</strong> with chitosan. To this end, 14 day-old wildtype<br />

<strong>moss</strong> colonies where sprayed with 100 µg/ml chitosan and mRNA was isolated at<br />

different time po<strong>in</strong>ts. Interest<strong>in</strong>gly, PpMPK4A mRNA was highly <strong>in</strong>duced by chitosan<br />

treatment already after 15 m<strong>in</strong>utes, peak<strong>in</strong>g at an eight-fold <strong>in</strong>duction after two hours and<br />

107


eturn<strong>in</strong>g to normal level by eight hours (Figure 1). In contrast, PpMPK4B mRNA was<br />

only marg<strong>in</strong>ally affected by this treatment.<br />

Generation of targeted PpMPK4B knock-out mutants<br />

The correlation between chitosan treatment and <strong>the</strong> expression of PpMPK4A mRNA<br />

<strong>in</strong>dicated that this MPK could play a role <strong>in</strong> P. <strong>patens</strong> <strong>in</strong>nate immune responses. For lossof-function<br />

analysis of PpMPK4A, <strong>the</strong> gene was replaced with a selection marker cassette<br />

by homologous recomb<strong>in</strong>ation [25]. After two rounds of selection, surviv<strong>in</strong>g colonies<br />

were exam<strong>in</strong>ed by PCR and two l<strong>in</strong>es <strong>in</strong> which PpMPK4A specific primers failed to<br />

amplify DNA were selected (Figure 2A and Table S1). First, we exam<strong>in</strong>ed for correct 3’<br />

and 5’ <strong>in</strong>tegration as well as full-length amplification (Figure 2B). Full length<br />

amplification with an external primer set was only possible <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 l<strong>in</strong>e,<br />

<strong>in</strong>dicat<strong>in</strong>g that <strong>the</strong> disruption <strong>in</strong> ΔPpMPK4A-2 was caused by concatemeric <strong>in</strong>tegration at<br />

<strong>the</strong> locus s<strong>in</strong>ce <strong>the</strong> 3’ primer and 5’ primer sets showed correct <strong>in</strong>tegration (Figure 2B).<br />

The full length PCR product from l<strong>in</strong>e ΔPpMPK4A-1 was sequenced to verify correct<br />

<strong>in</strong>tegration. Gene disruption was fur<strong>the</strong>r verified s<strong>in</strong>ce RT-PCR amplification of <strong>the</strong><br />

correspond<strong>in</strong>g PpMPK4A mRNAs was only possible <strong>in</strong> wild type (WT) and not <strong>in</strong> <strong>the</strong><br />

two <strong>in</strong>dependent KO l<strong>in</strong>es (Figure 2C).<br />

As noted above, loss of function mutants of A. thaliana MPK4 exhibit derepression<br />

of <strong>in</strong>nate immune responses and stunted growth [9]. Interest<strong>in</strong>gly, <strong>the</strong><br />

ΔPpMPK4A KO l<strong>in</strong>es did not exhibit such characteristics and appeared similar to <strong>the</strong> WT<br />

under different growth regimes. This <strong>in</strong>dicates that AtMPK4 and/or <strong>the</strong> pathways <strong>in</strong><br />

which it acts may have evolved additional functions compared to that of PpMPK4A.<br />

108


We also repeatedly attempted to make PpMPK4B KO l<strong>in</strong>es with two different KO<br />

vectors, but could not obta<strong>in</strong> any viable PpMPK4B KO l<strong>in</strong>es. A. thaliana AtMPK4 and<br />

AtMPK6 have recently been shown to be <strong>in</strong>volved <strong>in</strong> important developmental features of<br />

cytoskeletal regulation [26,27]. It is <strong>the</strong>refore possible that PpMPK4B performs essential<br />

functions <strong>in</strong> P. <strong>patens</strong> whose fewer MPK genes may provide less functional redundancy<br />

than <strong>the</strong> larger MPK gene families of higher plants.<br />

ΔPpMPK4A is highly susceptible to necrotrophic pathogens<br />

The four A. thaliana MAP k<strong>in</strong>ases AtMPK3/4/6 and 11 have been shown to function <strong>in</strong><br />

PTI signal<strong>in</strong>g, but some controversy exists about <strong>the</strong>ir <strong>in</strong>dividual functions <strong>in</strong> <strong>in</strong>nate<br />

immunity [7,8,10,17,28]. In contrast, noth<strong>in</strong>g is known about <strong>the</strong> functions of MPKs <strong>in</strong> P.<br />

<strong>patens</strong> <strong>in</strong>nate immunity. S<strong>in</strong>ce PpMPK4A mRNA responds to fungal chitosan, and<br />

necrotrophic fungi have previously been shown to <strong>in</strong>duce defense responses <strong>in</strong> <strong>moss</strong><br />

[29,30], we exam<strong>in</strong>ed <strong>the</strong> susceptibility of <strong>the</strong> two ΔPpMPK4A l<strong>in</strong>es to <strong>the</strong> necrotrophic<br />

fungus Alternaria brassicicola compared to wild type us<strong>in</strong>g a previously described spore<br />

count assay [31]. 14-day-old colonies of wild type and <strong>the</strong> ΔPpMPK4A l<strong>in</strong>es were drop<strong>in</strong>oculated<br />

with a fungal spore suspension and colonies were carefully harvested after<br />

four days for spore count<strong>in</strong>g. As seen <strong>in</strong> Figure 3A, <strong>the</strong> fungus produced significantly<br />

more spores <strong>in</strong> <strong>the</strong> two <strong>in</strong>dependent KO l<strong>in</strong>es of ΔPpMPK4A compared to wild type.<br />

We <strong>the</strong>n extended this analysis and <strong>in</strong>oculated <strong>the</strong> ΔPpMPK4A mutants and wild<br />

type colonies with Botrytis c<strong>in</strong>erea. A hallmark of B. c<strong>in</strong>erea <strong>in</strong>fection is host cell death<br />

detectable by Evans blue sta<strong>in</strong><strong>in</strong>g [12], and both ΔPpMPK4A mutant l<strong>in</strong>es exhibited<br />

massive cell death compared to wild type (Figure 3B). Taken toge<strong>the</strong>r, <strong>the</strong>se results<br />

109


<strong>in</strong>dicate that PpMPK4A is required for defense responses aimed at combat<strong>in</strong>g <strong>the</strong><br />

necrotrophic fungal pathogens Alternaria brassicicola and Botrytis c<strong>in</strong>erea.<br />

Defense-related gene transcripts fail to accumulate <strong>in</strong> ΔPpMPK4A<br />

To analyze whe<strong>the</strong>r <strong>the</strong> expression of defense related genes was altered <strong>in</strong> ΔPpMPK4A,<br />

we characterized <strong>the</strong> expression of several genes upon chitosan treatment. The genes<br />

chosen are known to be <strong>in</strong>volved <strong>in</strong> defense aga<strong>in</strong>st pathogens and/or are <strong>in</strong>duced by<br />

chitosan <strong>in</strong> both A. thaliana and P. <strong>patens</strong>: PAL4, CHS, ERF2, a-DOX, LOX7 and NOX<br />

[14,32]. In wild type plants <strong>the</strong>se genes were <strong>in</strong>duced after chitosan treatment compared<br />

to controls (Figure 4) and treatment without chitosan did not <strong>in</strong>duce any of <strong>the</strong> genes<br />

(CHS is shown as an example, FigureS2A). Notably, all of <strong>the</strong>se genes were upregulated<br />

to a lesser extent <strong>in</strong> <strong>the</strong> ΔPpMPK4B-1 l<strong>in</strong>e compared to wild type upon chitosan<br />

treatment (Figure 4). In wild type, PAL, CHS, ERF2 and LOX7 transcript levels after<br />

chitosan treatment were already elevated after 15 m<strong>in</strong> and peaked 2h after treatment,<br />

while NOX peaked after 4h. a-DOX started accumulat<strong>in</strong>g after 1h and peaked at 4h after<br />

treatment. In ΔPpMPK4A-1, accumulation of <strong>the</strong> transcripts occurred later and to a lesser<br />

extent than <strong>in</strong> wild type (Figure 4). This failure to accumulate a group of defense related<br />

genes <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 l<strong>in</strong>e po<strong>in</strong>ts to a role of PpMPK4A <strong>in</strong> signal transduction from<br />

chit<strong>in</strong> perception at <strong>the</strong> cell wall to <strong>the</strong> activation of defense genes.<br />

S<strong>in</strong>ce <strong>the</strong> expression of <strong>the</strong> defense-related genes is not fully <strong>in</strong>hibited <strong>in</strong><br />

ΔPpMPK4A, it is possible that PpMPK4B and PpMPK4A are partially redundant. As<br />

<strong>the</strong>re was no significant difference <strong>in</strong> <strong>the</strong> expression pattern of PpMPK4B <strong>in</strong> wild type<br />

110


versus ΔPpMPK4A-1 (Figure S2B), <strong>the</strong> expression of PpMPK4B was not altered <strong>in</strong><br />

ΔPpMPK4A to compensate for <strong>the</strong> absence of PpMPK4A.<br />

PAMP-<strong>in</strong>duced phosphorylation of PpMPK4A<br />

A. thaliana MPK3/4/6/11 are all thought to function below PRRs and to be<br />

phosphorylated upon PAMP treatment [2,7]. S<strong>in</strong>ce our data suggested that PpMPK4A<br />

functions as a signal<strong>in</strong>g component below PRRs <strong>in</strong> P. <strong>patens</strong>, we exam<strong>in</strong>ed if PpMPK4A<br />

and o<strong>the</strong>r MAP k<strong>in</strong>ases were phosphorylated and <strong>the</strong>reby activated upon different PAMP<br />

treatments.<br />

To this end we first performed immunoblot analysis of prote<strong>in</strong>s extracted from<br />

wild-type and <strong>the</strong> two PpMPK4A KO l<strong>in</strong>es after spray<strong>in</strong>g with 1 mg/ml chit<strong>in</strong> with antiphospho-p44/42<br />

antibodies. In wild-type controls, two MPKs became rapidly<br />

phosphorylated upon chit<strong>in</strong> treatment. Significantly, phosphorylation of only one MPK<br />

form was detected <strong>in</strong> <strong>the</strong> PpMPK4A KO l<strong>in</strong>es (Figure 5A). Given <strong>the</strong> predicted molecular<br />

weights of PpMPK4A (42.84 kD) and PpMPK4B (43.49 kD), <strong>the</strong> upper band present <strong>in</strong><br />

<strong>the</strong> wild type and KO l<strong>in</strong>es is probably phosphorylated PpMPK4B, while <strong>the</strong> lower band<br />

only seen <strong>in</strong> wild type is phosphorylated PpMPK4A. A similar phosphorylation pattern<br />

was seen upon treatment with 100 µg/ml chitosan, a deacetylated<br />

β-1,4-l<strong>in</strong>ked<br />

glucosam<strong>in</strong>e derivative of chit<strong>in</strong> (Figure S3A).<br />

To assess if activation of <strong>the</strong> MPKs by chit<strong>in</strong> and chitosan was similar to that<br />

upon pathogen <strong>in</strong>fection, <strong>moss</strong> colonies were sprayed with spores of B. c<strong>in</strong>erea. As seen<br />

<strong>in</strong> Figure 5A and S3A, several bands were detected <strong>in</strong> <strong>the</strong> immunoblot upon <strong>in</strong>fection<br />

with B. c<strong>in</strong>erea spores, <strong>in</strong>dicat<strong>in</strong>g phosphorylation of several MPKs <strong>in</strong>clud<strong>in</strong>g PpMPK4A<br />

111


(Figure 5B). The phosphorylation signals detected after <strong>in</strong>fection with fungal spores were<br />

weaker than <strong>the</strong> signals detected after chit<strong>in</strong> treatment. The film was <strong>the</strong>refore exposed<br />

for a longer time, result<strong>in</strong>g <strong>in</strong> some background signals seen as weak bands <strong>in</strong> <strong>the</strong><br />

untreated samples of both WT and ΔPpMPK4A (0 m<strong>in</strong> samples, Figure5B) Interest<strong>in</strong>gly,<br />

<strong>the</strong> band correspond<strong>in</strong>g to PpMPK4A was aga<strong>in</strong> absent <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 mutant.<br />

Significantly, <strong>the</strong> strong phosphorylation of wild type PpMPK4A at 16 m<strong>in</strong>utes and 60<br />

m<strong>in</strong>utes after <strong>in</strong>fection, comb<strong>in</strong>ed with <strong>the</strong> <strong>in</strong>creased susceptibility of ΔPpMPK4A to B.<br />

c<strong>in</strong>erea <strong>in</strong>fection, po<strong>in</strong>ts to <strong>the</strong> importance of PpMPK4A activation <strong>in</strong> signal transduction<br />

from a receptor to activate defense responses.<br />

A. thaliana MPK3/4/6/11 are all phosphorylated when <strong>the</strong> FLS2 receptor is<br />

activated by <strong>the</strong> flg22 PAMP [2,7], a small, conserved peptide of bacterial flagell<strong>in</strong> [33].<br />

Ano<strong>the</strong>r PAMP that triggers MPK phosphorylation is elf18 [34], a conserved peptide of<br />

bacterial elongation factor EF-Tu [4]. To assess <strong>the</strong> specificity of <strong>the</strong> chit<strong>in</strong>-<strong>in</strong>duced<br />

activation of PpMPK4A & B, we also treated wild-type <strong>moss</strong> colonies with flg22 and<br />

with elf18. No MPK phosphorylation was detected <strong>in</strong> response to flg22 or elf18, <strong>in</strong><br />

keep<strong>in</strong>g with <strong>the</strong> apparent absence of close homologues of <strong>the</strong> respective cognate<br />

immune receptors FLS2 and EFR <strong>in</strong> <strong>the</strong> P. <strong>patens</strong> genome (Figure 5C).<br />

We extended our analysis by attempt<strong>in</strong>g to detect whe<strong>the</strong>r P. <strong>patens</strong> MPKs are<br />

phosphorylated upon treatment with o<strong>the</strong>r bacterial PAMPs. To this end we treated <strong>the</strong><br />

<strong>moss</strong> with cell-free culture filtrates from two stra<strong>in</strong>s of Pectobacterium carotovorum ssp.<br />

carotovorum (P.c. carotovorum, formerly named Erw<strong>in</strong>ia carotovora ssp. carotovora). .<br />

The P.c. carotovorum SCC1 stra<strong>in</strong> produces <strong>the</strong> PAMP HrpN, while P.c. carotovorum<br />

SCC3193 is a harp<strong>in</strong> (HrpN)-negative stra<strong>in</strong>. While culture filtrates of both stra<strong>in</strong>s <strong>in</strong>duce<br />

112


defense responses <strong>in</strong> P. <strong>patens</strong>, filtrates of <strong>the</strong> harp<strong>in</strong>-secret<strong>in</strong>g stra<strong>in</strong> <strong>in</strong>duce much<br />

stronger responses than filtrates of <strong>the</strong> HrpN-negative stra<strong>in</strong> [30]. In keep<strong>in</strong>g with <strong>the</strong>se<br />

f<strong>in</strong>d<strong>in</strong>gs, culture filtrates from both stra<strong>in</strong>s <strong>in</strong>duced <strong>the</strong> activation of PpMPK4A and o<strong>the</strong>r<br />

MPK(s) (Figure 5D), <strong>in</strong>dicat<strong>in</strong>g that different MPKs are phosphorylated whe<strong>the</strong>r or not<br />

harp<strong>in</strong> has been secreted <strong>in</strong>to <strong>the</strong> culture. However, <strong>in</strong> <strong>the</strong> ΔPpMPK4A-1 mutant, <strong>the</strong><br />

lower band correspond<strong>in</strong>g to PpMPK4A was aga<strong>in</strong> absent (Figure 5D). This <strong>in</strong>dicates<br />

that <strong>the</strong> <strong>moss</strong> possesses PRRs that also recognize bacterial PAMPs, and that PpMPK4A<br />

most likely functions to transduce signals from a number of <strong>the</strong>m.<br />

We next assessed whe<strong>the</strong>r PpMPK4A could act downstream of ROS production<br />

<strong>in</strong> PAMP signal<strong>in</strong>g. We found that PpMPK4A&B were phosphorylated <strong>in</strong> wild type<br />

treated with exogenous hydrogen peroxide [35] and, aga<strong>in</strong>, <strong>the</strong> lower band was absent <strong>in</strong><br />

ΔPpMPK4A (Figure 5E). This <strong>in</strong>dicates that PpMPK4A functions downstream of ROS <strong>in</strong><br />

PAMP signal<strong>in</strong>g.<br />

A. thaliana MPK3/4/6 are all phosphorylated by various abiotic stresses <strong>in</strong>clud<strong>in</strong>g<br />

osmotic and cold stress, UV light and wound<strong>in</strong>g [19,36,37,38], and P. <strong>patens</strong> have<br />

previously been shown to respond at <strong>the</strong> level of gene expression upon exposure to UVlight,<br />

salt stress and cold [39,40,41]. Surpris<strong>in</strong>gly, we did not detect phosphorylation of<br />

any MPKs after stress treatments <strong>in</strong>clud<strong>in</strong>g salt (Figure S3B), cold, UV light, or<br />

wound<strong>in</strong>g (Figure S3C).<br />

Discussion<br />

Map k<strong>in</strong>ases are key transducers of extracellular stimuli <strong>in</strong> eukaryotes [5]. In vascular<br />

plants, numerous MAP k<strong>in</strong>ases have been shown to regulate signal<strong>in</strong>g cascades<br />

113


downstream of immune receptors [42]. However, <strong>the</strong> specific requirement for <strong>in</strong>dividual<br />

MAP k<strong>in</strong>ases <strong>in</strong> <strong>the</strong>se processes rema<strong>in</strong>s poorly understood. The results presented here<br />

show that immunity <strong>in</strong> <strong>the</strong> <strong>moss</strong> P. <strong>patens</strong> <strong>in</strong>volves PAMP recognition that activates<br />

defense responses via MAP k<strong>in</strong>ases <strong>in</strong>clud<strong>in</strong>g PpMPK4A. In addition, PpMPK4A<br />

represents a primary example <strong>in</strong> early land plants of a MAP k<strong>in</strong>ase that does not respond<br />

to various stress treatments. Instead, our results <strong>in</strong>dicate that this MAP k<strong>in</strong>ase primarily<br />

functions <strong>in</strong> <strong>in</strong>nate immunity and thus differs from its closest homologues <strong>in</strong> higher<br />

plants such as A. thaliana. That we were unable to detect phosphorylation of MAP<br />

k<strong>in</strong>ases upon exposure to different abiotic stresses is surpris<strong>in</strong>g. Although we cannot<br />

exclude <strong>the</strong> possibility that our stress conditions may be suboptimal, rapid responses to<br />

similar treatments have been documented <strong>in</strong> vascular plants [19,36,37,38]. An<br />

explanation for this apparent difference may be that <strong>the</strong> limited set of MAP k<strong>in</strong>ases <strong>in</strong><br />

<strong>moss</strong> does not function <strong>in</strong> abiotic stress responses. If so, o<strong>the</strong>r types of k<strong>in</strong>ases or<br />

regulators may be <strong>in</strong>volved <strong>in</strong> abiotic stress signal transduction <strong>in</strong> <strong>moss</strong>, and MPK<br />

functions <strong>in</strong> abiotic stress signal<strong>in</strong>g evolved subsequently <strong>in</strong> higher plant l<strong>in</strong>eages.<br />

The apparent phosphorylation of PpMPK4B <strong>in</strong>dicates that it also play roles <strong>in</strong><br />

PAMP signal<strong>in</strong>g. However, our <strong>in</strong>ability to produce a ΔPpMPK4B KO l<strong>in</strong>es makes it<br />

difficult to clarify such a function at this stage. However, <strong>the</strong> strong immunity<br />

phenotypes of ΔPpMPK4A <strong>in</strong>dicates that <strong>the</strong>se two MPKs have different roles <strong>in</strong> PAMP<br />

signal<strong>in</strong>g. Future studies us<strong>in</strong>g k<strong>in</strong>ase chemical-genetic approaches <strong>in</strong> P. <strong>patens</strong> may help<br />

to clarify PpMPK4B functions [43].<br />

PpMPK4A is phylogenetically related to AtMPK4 but, <strong>in</strong> apparent contrast to <strong>the</strong><br />

Atmpk4 mutant, <strong>the</strong> ΔPpMPK4A mutant appears phenotypically normal under all growth<br />

114


conditions tested. While this suggests that PpMPK4A and AtMPK4 have different<br />

functions, both k<strong>in</strong>ases are responsive to PAMPs and many studies have shown that<br />

AtMPK4 positively regulates PAMP triggered immunity [17]. One explanation for <strong>the</strong>ir<br />

apparently different functions is that AtMPK4 and its orthologues <strong>in</strong> higher plants have<br />

evolved additional functions. A second explanation relates to <strong>the</strong> evolution of host plant<br />

systems for monitor<strong>in</strong>g pathogen <strong>in</strong>fections. Successful pathogens deliver <strong>in</strong>to host cells<br />

various effectors which modify host prote<strong>in</strong>s and manipulate <strong>the</strong> host cell mach<strong>in</strong>ery to<br />

suppress immune responses [44]. Plants have developed mechanisms to detect <strong>the</strong>se<br />

microbial effectors via cytoplasmic immune receptors (termed resistance R-prote<strong>in</strong>s) that<br />

typically trigger a rapid, localized programmed cell death (PCD) reaction known as <strong>the</strong><br />

hypersensitive response (HR) [45]. R prote<strong>in</strong>s may thus guard effector targets or guardees<br />

[46]. Zhang et al. (2012) recently reported that <strong>the</strong> absence of AtMPK4 <strong>in</strong> <strong>the</strong> A. thaliana<br />

mpk4 mutant triggers immunity via a resistance prote<strong>in</strong> [10]. S<strong>in</strong>ce AtMPK4 is a key<br />

signal<strong>in</strong>g component <strong>in</strong> PAMP surveillance, AtMPK4 (or a complex or pathway<br />

conta<strong>in</strong><strong>in</strong>g it) may be targeted by pathogen effectors to block its activation by PAMPs<br />

and <strong>the</strong>reby prevent defense gene activation. It is <strong>the</strong>refore possible that, <strong>in</strong> <strong>the</strong><br />

evolutionarily older <strong>moss</strong>, PpMPK4A is not guarded. This may have enabled us to assess<br />

<strong>the</strong> phenotype and function of <strong>the</strong> ΔPpMPK4A mutant more readily than has been <strong>the</strong><br />

case for <strong>the</strong> Atmpk4 mutant. In addition, it would be <strong>in</strong>terest<strong>in</strong>g to exam<strong>in</strong>e whe<strong>the</strong>r our<br />

failure to isolate PpMPK4B mutants is because this k<strong>in</strong>ase performs essential functions or<br />

is <strong>in</strong> fact guarded. Although some 18 R genes are found <strong>in</strong> P. <strong>patens</strong> [47] it still needs to<br />

be demonstrated that <strong>the</strong>y function like <strong>the</strong> well studied R genes <strong>in</strong> vascular plants.<br />

115


Material and methods<br />

Growth conditions<br />

<strong>Physcomitrella</strong> <strong>patens</strong> (Gransden 2004 stra<strong>in</strong>) was grown on BCDAT media (250 mg/l<br />

MgSO 4·7H 2 O, 250 mg/l KH 2 PO 4 , 1010 mg/l KNO 3 , 920 mg/l Ammonium tartrate, 12.5<br />

mg/l FeSO 4·7H 2 O, 147 mg/l CaCl 2·2H 2 O, trace elements (614 μg/l H 3 BO 3 , 389 μg/l<br />

MnCl 2·4H 2 0, 110 μg/l AlK(S0 4 ) 2·12H 2 O, 55 μg/l CoCl 2·6H 2 0, 55 μg/l CuSO 4·5H 2 0, 55<br />

μg/l ZnS0 4·7H 2 0, 28 μg/l KBr, 28 μg/l KI, 28 μg/l LiCl, 28 μg/l SnCl 2·2H 2 0, 25 μg/l<br />

Na 2 MoO 4·2H 2 O, 59 μg/μl NiCl 2·6H 2 0)) [48], pH 6.5 adjusted with KOH, solidified with<br />

agar 8 g/l and overlaid with cellophane discs (AA Packag<strong>in</strong>g Ltd.). Colonies were grown<br />

at 22°C for two weeks <strong>in</strong> 55 µE m -2 s -1 <strong>in</strong> a long day regime of 16h light/8h dark. Botrytis<br />

c<strong>in</strong>erea and Alternaria brassicicola were cultivated on 24 g/l potato dextrose agar (PDA)<br />

(Difco, Detroit, MI, USA) at 22°C for app. two weeks until sporulation was dense.<br />

Spores were collected <strong>in</strong> H 2 O, filtered through Miracloth and concentration was<br />

determ<strong>in</strong>ed us<strong>in</strong>g a hemocytometer. Pectobacterium carotovorum subsp. carotovorum<br />

SCC3193 and Pectobacterium carotovorum subsp. carotovorum SCC1 were grown over<br />

night <strong>in</strong> LB media at 28°C. Cell densities were adjusted to equal O.D. with LB media.<br />

The cells were removed by centrifugation and <strong>the</strong> supernatant were sterile filtered us<strong>in</strong>g a<br />

0.20 µm filter (Thermo Scientific).<br />

Generation of knock out mutants<br />

The P. <strong>patens</strong> transformation vector pMBL10a was k<strong>in</strong>dly provided by David Cove.<br />

For PpMPK4A transformation, vector flank<strong>in</strong>g regions of gene Phypa 446407<br />

(Pp1s14939V6.1) was amplified us<strong>in</strong>g primer LBF+LBR and RBF+RBR (Table S1). The<br />

116


esult<strong>in</strong>g PCR fragments were cloned <strong>in</strong>to pCR®-Blunt II-TOPO® (Invitrogen) and<br />

transformed <strong>in</strong>to E. coli. The LB fragment was cut from <strong>the</strong> TOPO vector us<strong>in</strong>g<br />

restriction enzymes Not1 and BamH1, and <strong>the</strong> result<strong>in</strong>g 651 bp fragment ligated us<strong>in</strong>g T4<br />

fast ligase (Promega) <strong>in</strong>to pMBL10a l<strong>in</strong>earized with Not1 and BamH1. Subsequently, <strong>the</strong><br />

RB fragment was cloned <strong>in</strong>to pMBL10a conta<strong>in</strong><strong>in</strong>g <strong>the</strong> LB. The RB fragment was cut<br />

from Topo vector with EcoRV, and <strong>the</strong> result<strong>in</strong>g 737 bp fragment ligated <strong>in</strong>to pMBL10a<br />

conta<strong>in</strong><strong>in</strong>g LB l<strong>in</strong>earized with EcoRV and desphosphorylated us<strong>in</strong>g shrimp alkal<strong>in</strong>e<br />

phosphatase (Affymetrix). The transformation cassette was excised from <strong>the</strong> result<strong>in</strong>g<br />

vector, with Not1 and Xba1. P. <strong>patens</strong> protoplasts were transformed with 30 μg of vector<br />

DNA us<strong>in</strong>g PEG as previously described [25]. Stable transformants were identified by<br />

two rounds of selection: <strong>in</strong>itial selection on media supplemented with 40 μg/ml G418<br />

(Sigma) for two weeks, <strong>the</strong>n a non-selective release step of two weeks, and f<strong>in</strong>ally two<br />

weeks of G418 selection.<br />

Plant treatments for MAPK phosphorylation assay<br />

Unless o<strong>the</strong>rwise stated, all treatments were applied by spray<strong>in</strong>g 3 ml of <strong>the</strong> solution used<br />

for <strong>the</strong> treatment onto a Petri dish (9 cm) with 16 colonies of 14 day- old <strong>moss</strong> grown on<br />

BCDAT media overlaid with cellophane. Chit<strong>in</strong> oligosaccharide (Yaizu Suisankagaku<br />

Industry) was diluted to 1 mg/ml <strong>in</strong> H 2 O. The chitosan (Sigma Aldrich, 448869) used <strong>in</strong><br />

this study conta<strong>in</strong>s 75-85% deacetylated, low molecular weight chit<strong>in</strong> (50,000 - 190,000<br />

daltons). It was made water-soluble by dissolv<strong>in</strong>g 100 µg/ml <strong>in</strong> 0.010% acetic acid and<br />

<strong>the</strong>n adjust<strong>in</strong>g <strong>the</strong> pH to 5.5 with NaOH. 0.010% acetic acid adjusted to pH 5.5 with<br />

117


NaOH was used as a control. For MPK phosphorylation Western blot with B. c<strong>in</strong>erea,<br />

2x10 6 spores per ml were sprayed onto <strong>the</strong> colonies. Cold stress was applied by mov<strong>in</strong>g<br />

<strong>the</strong> cellophane disc with 14 day-old colonies onto prechilled BCDAT plated on ice. The<br />

surface temperature was measured dur<strong>in</strong>g <strong>the</strong> experiment at 1-2°C. UV light stress was<br />

applied <strong>in</strong> two ways. For UV(1), colonies were subjected to UV light from a Geldoc 2000<br />

conta<strong>in</strong><strong>in</strong>g six UV-B tubes (G8T5E, Ushio) and flash-frozen at <strong>the</strong> <strong>in</strong>dicated times. For<br />

UV(2), colonies were subjected to 120 mJ cm -2 UV-B (Stratal<strong>in</strong>ker 1800, Stratagene) and<br />

flash-frozen after 15 m<strong>in</strong>. Wound<strong>in</strong>g was conducted by squeez<strong>in</strong>g two colonies <strong>in</strong> an<br />

Eppendorf tube us<strong>in</strong>g a pestle which was <strong>the</strong>n flash-frozen at <strong>the</strong> <strong>in</strong>dicated time.<br />

Alternaria spore count<strong>in</strong>g assay<br />

The Alternaria spore count<strong>in</strong>g assay was adapted from Wees et al. [31]. A. brassicicola<br />

spores were harvested <strong>in</strong> H 2 O from 14-day-old plates and counted <strong>in</strong> a hemocytometer.<br />

Moss colonies were <strong>in</strong>fected by plac<strong>in</strong>g 5 µl of 2x10 5 A. brassicicola spores/ml on top of<br />

each colony. Four plates of each l<strong>in</strong>e, each with 16 colonies, were drop <strong>in</strong>oculated. Four<br />

days later <strong>the</strong> 16 colonies from each plate were carefully transferred to a tube conta<strong>in</strong><strong>in</strong>g<br />

8 ml of 0.01% (v/v) Tween 20 (Sigma-Aldrich) and shaken vigorously. The numbers of<br />

spores <strong>in</strong> <strong>the</strong> liquid were counted <strong>in</strong> a hemocytometer. Each data po<strong>in</strong>t is an average of<br />

four pools of 16 colonies per genotype. An analysis of variance followed by a Tukey's<br />

test was used to assess <strong>the</strong> statistical difference between <strong>the</strong> WT and <strong>the</strong> two KOs. The<br />

experiment was repeated twice with similar results.<br />

Evans blue sta<strong>in</strong><strong>in</strong>g for detection of cell death<br />

118


Measurement of cell death with Evans blue sta<strong>in</strong><strong>in</strong>g was done similarly to <strong>the</strong> assay<br />

described by Oliver et al. [32]. Four plates of each l<strong>in</strong>e, each with 16 colonies, were<br />

sprayed with 3 ml of 2x10 5 B. c<strong>in</strong>erea spores per ml and one plate sprayed with water.<br />

Two days later, <strong>the</strong> colonies were <strong>in</strong>cubated <strong>in</strong> 0.1% Evans blue (Bie & Berntsen) <strong>in</strong> 0.5x<br />

PBS for 2 hours followed by wash<strong>in</strong>g four times <strong>in</strong> H 2 O. They were <strong>the</strong>n desta<strong>in</strong>ed <strong>in</strong><br />

50% (v/v) methanol, 1% SDS at 60°C for 30 m<strong>in</strong>. and <strong>the</strong> O.D. at 600 nm measured.<br />

Colonies were dried over night at 70°C and weighed. Each data po<strong>in</strong>t is presented as O.D.<br />

(600nm)/mg dry weight as an average of four samples of 16 colonies each. An analysis of<br />

variance followed by a Tukey's test was applied to assess <strong>the</strong> significance of <strong>the</strong><br />

difference. The experiment was repeated twice with similar results.<br />

RNA extraction and quantitative RT-PCR<br />

Two 14 day-old colonies were sampled at each time po<strong>in</strong>t after spray<strong>in</strong>g with 100 µg/ml<br />

chitosan or control and flash frozen <strong>in</strong> liquid nitrogen. RNA was extracted with<br />

NucleoSp<strong>in</strong> RNA plant (Macherey-Nagel) accord<strong>in</strong>g to <strong>the</strong> manufacturer’s protocol.<br />

RNA concentrations were measured us<strong>in</strong>g a Nano Drop 1000 (Thermo Scientific) and<br />

adjusted to 5 ng/μl. Quantitative PCR was done us<strong>in</strong>g Brilliant II SYBR green one step<br />

kit (Agilent Technologies) with 10 pmol of each primer and 12.5 ng total RNA <strong>in</strong> 10 µl.<br />

Reactions were run on a CFX 96 Thermocycler (BioRad) three times, and means of<br />

normalized expression were calculated accord<strong>in</strong>g to Pfaffl [49]. Primers used for qPCR<br />

are <strong>in</strong> supplementary Table S2.<br />

MAP K<strong>in</strong>ase Assay<br />

119


Three 14-day-old colonies were flash-frozen at <strong>the</strong> given time po<strong>in</strong>t after treatment.<br />

Prote<strong>in</strong> was extracted by gr<strong>in</strong>d<strong>in</strong>g <strong>in</strong> Lacus buffer (50 mM Tris-HCl pH 7.5, 10 mM<br />

MgCl2, 15 mM EGTA, 100 mM NaCl, 2 mM DTT, 30 mM β-glycero-phosphate, 0.1%<br />

NP-40) plus Phosphatase <strong>in</strong>hibitor cocktail (PhosSTOP, Roche) and Protease <strong>in</strong>hibitor<br />

cocktail (Complete, Roche). Samples were cleared by centrifugation, boiled for 5 m<strong>in</strong> <strong>in</strong><br />

load<strong>in</strong>g buffer and subjected to SDS-PAGE and electroblott<strong>in</strong>g. Immunoblots were<br />

blocked by <strong>in</strong>cubat<strong>in</strong>g for 1h <strong>in</strong> TBS-Tween (0.1%, (v/v)) and 5% (w/v) milk powder.<br />

Phosphorylated MAP k<strong>in</strong>ases where detected by <strong>in</strong>cubation overnight with primary<br />

antibody anti-p42/p44-erk (1/2000, CST) <strong>in</strong> 5% (w/v) BSA, followed by <strong>in</strong>cubation for<br />

1h with anti-rabbit-HRP secondary antibody (1/10000, Sigma-Aldrich) <strong>in</strong> TBS-Tween<br />

(0.1%, (v/v)) and 5% (w/v) milk powder. Antibodies were visualized with Pierce ECL<br />

Plus Western Blott<strong>in</strong>g Substrate (Thermo Scientific) and exposed to X-ray film (Ultra<br />

UV-G, AGFA).<br />

Acknowledgement<br />

Thank you to: Dr. Andrew Cum<strong>in</strong>g for provid<strong>in</strong>g <strong>the</strong> pMBL10a transformation vector,<br />

Christ<strong>in</strong>e Lunde for provid<strong>in</strong>g <strong>the</strong> wild type <strong>Physcomitrella</strong> <strong>patens</strong> (Gransden 2004 stra<strong>in</strong>)<br />

and E.Tapio Palva for <strong>the</strong> P.c. carotovorum stra<strong>in</strong>s.<br />

120


Refferences<br />

1. Zipfel C (2009) Early molecular events <strong>in</strong> PAMP-triggered immunity. Current<br />

Op<strong>in</strong>ion <strong>in</strong> Plant Biology 12: 414–420. doi:10.1016/j.pbi.2009.06.003.<br />

2. Gómez-Gómez L, Boller T (2000) FLS2: an LRR receptor-like k<strong>in</strong>ase <strong>in</strong>volved <strong>in</strong><br />

<strong>the</strong> perception of <strong>the</strong> bacterial elicitor flagell<strong>in</strong> <strong>in</strong> Arabidopsis. Mol Cell 5: 1003–1011.<br />

3. Miya A, Albert P, Sh<strong>in</strong>ya T, Desaki Y, Ichimura K, et al. (2007) CERK1, a LysM<br />

Receptor K<strong>in</strong>ase, Is Essential for Chit<strong>in</strong> Elicitor Signal<strong>in</strong>g <strong>in</strong> Arabidopsis. PNAS 104:<br />

19613–19618. doi:10.1073/pnas.0705147104.<br />

4. Zipfel C, Kunze G, Ch<strong>in</strong>chilla D, Caniard A, Jones JDG, et al. (2006) Perception<br />

of <strong>the</strong> Bacterial PAMP EF-Tu by <strong>the</strong> Receptor EFR Restricts Agrobacterium-Mediated<br />

Transformation. Cell 125: 749–760. doi:10.1016/j.cell.2006.03.037.<br />

5. Suarez-Rodriguez CM, Petersen M, Mundy J (2010) Mitogen-Activated Prote<strong>in</strong><br />

K<strong>in</strong>ase Signal<strong>in</strong>g <strong>in</strong> Plants. Annual Review of Plant Biology 61: 621–649.<br />

doi:10.1146/annurev-arplant-042809-112252.<br />

6. Fiil BK, Petersen K, Petersen M, Mundy J (2009) Gene regulation by MAP k<strong>in</strong>ase<br />

cascades. Current Op<strong>in</strong>ion <strong>in</strong> Plant Biology 12: 615–621. doi:10.1016/j.pbi.2009.07.017.<br />

7. Bethke G, Pecher P, Eschen-Lippold L, Tsuda K, Katagiri F, et al. (2012)<br />

Activation of <strong>the</strong> Arabidopsis thaliana mitogen-activated prote<strong>in</strong> k<strong>in</strong>ase MPK11 by <strong>the</strong><br />

flagell<strong>in</strong>-derived elicitor peptide, flg22. Mol Plant Microbe Interact 25: 471–480.<br />

doi:10.1094/MPMI-11-11-0281.<br />

8. Ren D, Liu Y, Yang K-Y, Han L, Mao G, et al. (2008) A Fungal-Responsive<br />

MAPK Cascade Regulates Phytoalex<strong>in</strong> Biosyn<strong>the</strong>sis <strong>in</strong> Arabidopsis. PNAS 105: 5638–<br />

5643. doi:10.1073/pnas.0711301105.<br />

9. Petersen M, Brodersen P, Naested H, Andreasson E, L<strong>in</strong>dhart U, et al. (2000)<br />

Arabidopsis MAP K<strong>in</strong>ase 4 Negatively Regulates Systemic Acquired Resistance. Cell<br />

103: 1111–1120. doi:10.1016/S0092-8674(00)00213-0.<br />

10. Zhang Z, Wu Y, Gao M, Zhang J, Kong Q, et al. (2012) Disruption of PAMP-<br />

Induced MAP K<strong>in</strong>ase Cascade by a Pseudomonas syr<strong>in</strong>gae Effector Activates Plant<br />

<strong>Immunity</strong> Mediated by <strong>the</strong> NB-LRR Prote<strong>in</strong> SUMM2. Cell Host & Microbe 11: 253–263.<br />

doi:10.1016/j.chom.2012.01.015.<br />

11. Zimmer A, Lang D, Richardt S, Frank W, Reski R, et al. (2007) Dat<strong>in</strong>g <strong>the</strong> early<br />

evolution of plants: detection and molecular clock analyses of orthologs. Molecular<br />

Genetics and Genomics 278: 393–402. doi:10.1007/s00438-007-0257-6.<br />

12. Ponce De León I, Schmelz EA, Gaggero C, Castro A, Álvarez A, et al. (2012)<br />

<strong>Physcomitrella</strong> <strong>patens</strong> activates re<strong>in</strong>forcement of <strong>the</strong> cell wall, programmed cell death<br />

and accumulation of evolutionary conserved defence signals, such as salicylic acid and<br />

121


12‐oxo‐phytodienoic acid, but not jasmonic acid, upon Botrytis c<strong>in</strong>erea <strong>in</strong>fection.<br />

Molecular Plant Pathology 13: 960–974. doi:10.1111/j.1364-3703.2012.00806.x.<br />

13. Andersson RA, Akita M, Pirhonen M, Gammelgård E, Valkonen JPT (2005)<br />

Moss-Erw<strong>in</strong>ia pathosystem reveals possible similarities <strong>in</strong> pathogenesis and pathogen<br />

defense <strong>in</strong> vascular and nonvascular plants. Journal of General Plant Pathology 71: 23–28.<br />

doi:10.1007/s10327-004-0154-3.<br />

14. Lehtonen MT, Akita M, Frank W, Reski R, Valkonen JPT (2012) Involvement of<br />

a Class III Peroxidase and <strong>the</strong> Mitochondrial Prote<strong>in</strong> TSPO <strong>in</strong> Oxidative Burst Upon<br />

Treatment of Moss Plants with a Fungal Elicitor. Molecular Plant-Microbe Interactions<br />

25: 363–371. doi:10.1094/MPMI-10-11-0265.<br />

15. Dóczi R, Ökrész L, Romero AE, Paccanaro A, Bögre L (n.d.) Explor<strong>in</strong>g <strong>the</strong><br />

evolutionary path of plant MAPK networks. Trends <strong>in</strong> Plant Science 17: 503–562.<br />

doi:10.1016/j.tplants.2012.05.009.<br />

16. Ichimura K, Mizoguchi T, Irie K, Morris P, Giraudat J, et al. (1998) Isolation of<br />

ATMEKK1 (a MAP K<strong>in</strong>ase K<strong>in</strong>ase K<strong>in</strong>ase)-Interact<strong>in</strong>g Prote<strong>in</strong>s and Analysis of a MAP<br />

K<strong>in</strong>ase Cascade <strong>in</strong>Arabidopsis. Biochemical and Biophysical Research Communications<br />

253: 532–543. doi:10.1006/bbrc.1998.9796.<br />

17. Qiu J-L, Fiil BK, Petersen K, Nielsen HB, Botanga CJ, et al. (2008) Arabidopsis<br />

MAP k<strong>in</strong>ase 4 regulates gene expression through transcription factor release <strong>in</strong> <strong>the</strong><br />

nucleus. The EMBO Journal 27: 2214–2221. doi:10.1038/emboj.2008.147.<br />

18. Suarez-Rodriguez MC, Adams-Phillips L, Liu Y, Wang H, Su S-H, et al. (2007)<br />

MEKK1 Is Required for Flg22-Induced MPK4 Activation <strong>in</strong> Arabidopsis Plants. Plant<br />

Physiol 143: 661–669. doi:10.1104/pp.106.091389.<br />

19. Teige M, Scheikl E, Eulgem T, Dóczi R, Ichimura K, et al. (2004) The MKK2<br />

pathway mediates cold and salt stress signal<strong>in</strong>g <strong>in</strong> Arabidopsis. Mol Cell 15: 141–152.<br />

doi:10.1016/j.molcel.2004.06.023.<br />

20. Kong Q, Qu N, Gao M, Zhang Z, D<strong>in</strong>g X, et al. (2012) The MEKK1-<br />

MKK1/MKK2-MPK4 K<strong>in</strong>ase Cascade Negatively Regulates <strong>Immunity</strong> Mediated by a<br />

Mitogen-Activated Prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ase K<strong>in</strong>ase <strong>in</strong> Arabidopsis. Plant Cell.<br />

doi:10.1105/tpc.112.097253.<br />

21. Ichimura K, Sh<strong>in</strong>ozaki K, Tena G, Sheen J, Henry Y, et al. (2002) Mitogenactivated<br />

prote<strong>in</strong> k<strong>in</strong>ase cascades <strong>in</strong> plants: a new nomenclature. Trends <strong>in</strong> Plant Science<br />

7: 301–308. doi:10.1016/S1360-1385(02)02302-6.<br />

22. Hruz T, Laule O, Szabo G, Wessendorp F, Bleuler S, et al. (2008) Genevestigator<br />

V3: A Reference Expression Database for <strong>the</strong> Meta-Analysis of Transcriptomes.<br />

Advances <strong>in</strong> Bio<strong>in</strong>formatics 2008: 1–5. doi:10.1155/2008/420747.<br />

122


23. Lehtonen MT, Akita M, Kalkk<strong>in</strong>en N, Ahola-Iivar<strong>in</strong>en E, Rönnholm G, et al.<br />

(2009) Quickly-released peroxidase of <strong>moss</strong> <strong>in</strong> defense aga<strong>in</strong>st fungal <strong>in</strong>vaders. New<br />

Phytologist 183: 432–443. doi:10.1111/j.1469-8137.2009.02864.x.<br />

24. Albert M, K. Jehle A, Lipschis M, Mueller K, Zeng Y, et al. (2010) Regulation of<br />

cell behaviour by plant receptor k<strong>in</strong>ases: Pattern recognition receptors as prototypical<br />

models. European Journal of Cell Biology 89: 200–207. doi:10.1016/j.ejcb.2009.11.015.<br />

25. Schaefer D, Zryd J-P, Knight CD, Cove DJ (1991) Stable transformation of <strong>the</strong><br />

<strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. Molecular and General Genetics MGG 226: 418–424.<br />

doi:10.1007/BF00260654.<br />

26. Kosetsu K, Matsunaga S, Nakagami H, Colcombet J, Sasabe M, et al. (2010) The<br />

MAP K<strong>in</strong>ase MPK4 Is Required for Cytok<strong>in</strong>esis <strong>in</strong> Arabidopsis thaliana. Plant Cell 22:<br />

3778–3790. doi:10.1105/tpc.110.077164.<br />

27. Müller J, Beck M, Mettbach U, Komis G, Hause G, et al. (2010) Arabidopsis<br />

MPK6 is <strong>in</strong>volved <strong>in</strong> cell division plane control dur<strong>in</strong>g early root development, and<br />

localizes to <strong>the</strong> pre-prophase band, phragmoplast, trans-Golgi network and plasma<br />

membrane. The Plant Journal 61: 234–248. doi:10.1111/j.1365-313X.2009.04046.x.<br />

28. Mao G, Meng X, Liu Y, Zheng Z, Chen Z, et al. (2011) Phosphorylation of a<br />

WRKY Transcription Factor by Two Pathogen-Responsive MAPKs Drives Phytoalex<strong>in</strong><br />

Biosyn<strong>the</strong>sis <strong>in</strong> Arabidopsis. Plant Cell 23: 1639–1653. doi:10.1105/tpc.111.084996.<br />

29. Akita M, Lehtonen MT, Koponen H, Martt<strong>in</strong>en EM, Valkonen JPT (2011)<br />

Infection of <strong>the</strong> Sunagoke <strong>moss</strong> panels with fungal pathogens hampers susta<strong>in</strong>able<br />

green<strong>in</strong>g <strong>in</strong> urban environments. Science of The Total Environment 409: 3166–3173.<br />

doi:10.1016/j.scitotenv.2011.05.009.<br />

30. Ponce de León I, Oliver J, Castro A, Gaggero C, Bentancor M, et al. (2007)<br />

Erw<strong>in</strong>ia carotovora elicitors and Botrytis c<strong>in</strong>erea activate defense responses <strong>in</strong><br />

<strong>Physcomitrella</strong> <strong>patens</strong>. BMC Plant Biology 7: 52. doi:10.1186/1471-2229-7-52.<br />

31. Wees V, C.m S, Chang H-S, Zhu T, Glazebrook J (2003) Characterization of <strong>the</strong><br />

Early Response of Arabidopsis to Alternaria Brassicicola Infection Us<strong>in</strong>g Expression<br />

Profil<strong>in</strong>g. Plant Physiol 132: 606–617. doi:10.1104/pp.103.022186.<br />

32. Oliver J, Castro A, Gaggero C, Cascón T, Schmelz E, et al. (2009) Pythium<br />

<strong>in</strong>fection activates conserved plant defense responses <strong>in</strong> <strong>moss</strong>es. Planta 230: 569–579.<br />

doi:10.1007/s00425-009-0969-4.<br />

33. Felix G, Duran JD, Volko S, Boller T (2002) Plants have a sensitive perception<br />

system for <strong>the</strong> most conserved doma<strong>in</strong> of bacterial flagell<strong>in</strong>. The Plant Journal 18: 265–<br />

276. doi:10.1046/j.1365-313X.1999.00265.x.<br />

34. Roux M, Schwess<strong>in</strong>ger B, Albrecht C, Ch<strong>in</strong>chilla D, Jones A, et al. (2011) The<br />

Arabidopsis Leuc<strong>in</strong>e-Rich Repeat Receptor–Like K<strong>in</strong>ases BAK1/SERK3 and<br />

123


BKK1/SERK4 Are Required for Innate <strong>Immunity</strong> to Hemibiotrophic and Biotrophic<br />

Pathogens. Plant Cell 23: 2440–2455. doi:10.1105/tpc.111.084301.<br />

35. Nakagami H, Soukupová H, Schikora A, Zárský V, Hirt H (2006) A Mitogenactivated<br />

Prote<strong>in</strong> K<strong>in</strong>ase K<strong>in</strong>ase K<strong>in</strong>ase Mediates Reactive Oxygen Species Homeostasis<br />

<strong>in</strong> Arabidopsis. J Biol Chem 281: 38697–38704. doi:10.1074/jbc.M605293200.<br />

36. Droillard M-J, Boudsocq M, Barbier-Brygoo H, Laurière C (2004) Involvement<br />

of MPK4 <strong>in</strong> osmotic stress response pathways <strong>in</strong> cell suspensions and plantlets of<br />

Arabidopsis thaliana: activation by hypoosmolarity and negative role <strong>in</strong> hyperosmolarity<br />

tolerance. FEBS Letters 574: 42–48. doi:10.1016/j.febslet.2004.08.001.<br />

37. Ichimura K, Mizoguchi T, Yoshida R, Yuasa T, Sh<strong>in</strong>ozaki K (2000) Various<br />

abiotic stresses rapidly activate Arabidopsis MAP k<strong>in</strong>ases ATMPK4 and ATMPK6. The<br />

Plant Journal 24: 655–665. doi:10.1046/j.1365-313x.2000.00913.x.<br />

38. González Besteiro MA, Bartels S, Albert A, Ulm R (2011) Arabidopsis MAP<br />

k<strong>in</strong>ase phosphatase 1 and its target MAP k<strong>in</strong>ases 3 and 6 antagonistically determ<strong>in</strong>e UV-<br />

B stress tolerance, <strong>in</strong>dependent of <strong>the</strong> UVR8 photoreceptor pathway. The Plant Journal<br />

68: 727–737. doi:10.1111/j.1365-313X.2011.04725.x.<br />

39. Lunde C, Drew DP, Jacobs AK, Tester M (2007) Exclusion of Na+ via Sodium<br />

ATPase (PpENA1) Ensures Normal Growth of <strong>Physcomitrella</strong> <strong>patens</strong> under Moderate<br />

Salt Stress. Plant Physiol 144: 1786–1796. doi:10.1104/pp.106.094946.<br />

40. Richardt S, Timmerhaus G, Lang D, Qudeimat E, Corrêa LGG, et al. (2009)<br />

Microarray analysis of <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong> reveals evolutionarily conserved<br />

transcriptional regulation of salt stress and abscisic acid signall<strong>in</strong>g. Plant Molecular<br />

Biology 72: 27–45. doi:10.1007/s11103-009-9550-6.<br />

41. Wolf L, Rizz<strong>in</strong>i L, Stracke R, Ulm R, Rens<strong>in</strong>g SA (2010) The Molecular and<br />

Physiological Responses of <strong>Physcomitrella</strong> Patens to Ultraviolet-B Radiation. Plant<br />

Physiol 153: 1123–1134. doi:10.1104/pp.110.154658.<br />

42. Rasmussen MW, Roux M, Petersen M, Mundy J (2012) MAP K<strong>in</strong>ase Cascades <strong>in</strong><br />

Arabidopsis Innate <strong>Immunity</strong>. Front Plant Sci 3: 169. doi:10.3389/fpls.2012.00169.<br />

43. Brodersen P, Petersen M, Bjørn Nielsen H, Zhu S, Newman M-A, et al. (2006)<br />

Arabidopsis MAP k<strong>in</strong>ase 4 regulates salicylic acid- and jasmonic acid/ethylenedependent<br />

responses via EDS1 and PAD4. The Plant Journal 47: 532–546.<br />

doi:10.1111/j.1365-313X.2006.02806.x.<br />

44. Bent AF, Mackey D (2007) Elicitors, Effectors, and R Genes: The New Paradigm<br />

and a Lifetime Supply of Questions. Annual Review of Phytopathology 45: 399–436.<br />

doi:10.1146/annurev.phyto.45.062806.094427.<br />

45. DeYoung BJ, Innes RW (2006) Plant NBS-LRR prote<strong>in</strong>s <strong>in</strong> pathogen sens<strong>in</strong>g and<br />

host defense. Nature Immunology 7: 1243–1249. doi:10.1038/ni1410.<br />

124


46. Jones JDG, Dangl JL (2006) The plant immune system. Nature 444: 323–329.<br />

doi:10.1038/nature05286.<br />

47. Xue J-Y, Wang Y, Wu P, Wang Q, Yang L-T, et al. (2012) A Primary Survey on<br />

Bryophyte Species Reveals Two Novel Classes of Nucleotide-B<strong>in</strong>d<strong>in</strong>g Site (NBS) Genes.<br />

PLoS ONE 7: e36700. doi:10.1371/journal.pone.0036700.<br />

48. Ashton NW, Cove DJ (1977) The isolation and prelim<strong>in</strong>ary characterisation of<br />

auxotrophic and analogue resistant mutants of <strong>the</strong> <strong>moss</strong> <strong>Physcomitrella</strong> <strong>patens</strong>. Molecular<br />

and General Genetics MGG 154: 87–95. doi:10.1007/BF00265581.<br />

49. Pfaffl MW (2001) A new ma<strong>the</strong>matical model for relative quantification <strong>in</strong> realtime<br />

RT–PCR. Nucleic Acids Res 29: e45.<br />

125


Figures<br />

Figure 1. PpMPK4A is responsive to fungal chit<strong>in</strong>. Quantitative reverse transcriptase<br />

PCR (qPCR) analysis of PpMPK4B (blue diamonds) and PpMPK4A (red squares)<br />

expression relative to untreated (time 0h) follow<strong>in</strong>g treatment with 100 μg/ml chitosan.<br />

Expression level relative to wild type (WT) untreated was calculated as previously<br />

described by Pfaffl [49], with β-TUBULIN (Phypa_440500) as an <strong>in</strong>ternal control for<br />

normalization. Error bars represent S.E.M. of three <strong>in</strong>dependent technical replicates. The<br />

experiment was repeated with similar results.<br />

126


Figure 2. Identification of ΔPpMPK4A l<strong>in</strong>es by PCR. See table S1 for primer<br />

sequences. (A) Overview of <strong>the</strong> PpMPK4A wild type and KO locus. Primers used to<br />

identify KO l<strong>in</strong>es are <strong>in</strong>dicated as p1-p8. (B) Verification of DNA <strong>in</strong>sertion <strong>in</strong> <strong>the</strong><br />

genome by PCR amplification with <strong>the</strong> <strong>in</strong>dicated primer comb<strong>in</strong>ations on DNA from WT<br />

and two <strong>in</strong>dependent ΔPpMPK4A KO l<strong>in</strong>es. (C) Evaluation of PpMPK4A expression<br />

assessed by RT-PCR <strong>in</strong> WT and two <strong>in</strong>dependent ΔPpMPK4A KO l<strong>in</strong>es.<br />

127


Figure 3. ΔPpMPK4A plants are more susceptible to necrotrophic pathogens. (A) A.<br />

brassicicola spore counts four days post-drop-<strong>in</strong>oculation with ~2500<br />

spores per P. <strong>patens</strong> colony. Each data po<strong>in</strong>t is an average of four pools of 16 colonies<br />

per genotype. Analysis of variance followed by Tukey's test showed a statistical<br />

difference between <strong>the</strong> WT and <strong>the</strong> two <strong>in</strong>dependent ΔPpMPK4A (KO1 and KO2) l<strong>in</strong>es,<br />

<strong>in</strong>dicated by two asterisks (p


Figure 4. PpMPK4A is required for PAMP <strong>in</strong>duced gene expression. Quantitative<br />

reverse transcriptase PCR (qPCR) analysis of transcript levels <strong>in</strong> WT (blue diamonds)<br />

and ΔPpMPK4A-1 (red squares) relative to untreated WT (time 0h) follow<strong>in</strong>g treatment<br />

with 100 μg/ml chitosan. Expression level relative to untreated WT was calculated as<br />

previously [49], with β-TUBULIN (Phypa_440500) as an <strong>in</strong>ternal control for<br />

normalization. Error bars represent S.E.M. of three <strong>in</strong>dependent technical replicates. (A)<br />

PAL4 (Phypa_461241), (B) CHS (Phypa_427901), (C) ERF2 (Phypa_422542), (D) α-<br />

DOX (Phypa_447346), (E) LOX7 (Phypa_184078) and (F) NOX (Phypa_204103).<br />

129


Figure 5. PpMPK4A is phosphorylated upon PAMP treatment. Immunoblot analysis<br />

with anti-phospho-p44/42 MAPK antibodies at <strong>in</strong>dicated m<strong>in</strong>utes after different<br />

treatments. Load<strong>in</strong>g controls show amido black sta<strong>in</strong>ed total prote<strong>in</strong>.<br />

(A) WT and two <strong>in</strong>dependent ΔPpMPK4A (KO1 and KO2) l<strong>in</strong>es sprayed with 1 mg/ml<br />

chit<strong>in</strong>. The arrow po<strong>in</strong>ts to PpMPK4A. (B) WT sprayed with ei<strong>the</strong>r 100 nM flg22, 100<br />

nM elf18 or 100 μg/ml chitosan. (C) WT and ΔPpMPK4A-1 (KO1) sprayed with 2x10 6 B.<br />

c<strong>in</strong>erea spores/ml. (D) WT and ΔPpMPK4A-1 (KO1) sprayed with ei<strong>the</strong>r LB medium, 1<br />

mg/ml chit<strong>in</strong>, cell free culture filtrate of <strong>the</strong> HrpN negative stra<strong>in</strong> P.c. carotovorum<br />

130


SCC3193, or cell free culture filtrate of <strong>the</strong> HrpN positive stra<strong>in</strong> P.c. carotovorum SCC1.<br />

PpMPK4A is marked with an arrow. (E) WT and ΔPpMPK4A-1 (KO1) sprayed with 10<br />

mM H 2 O 2 .<br />

Supplementary <strong>in</strong>ventory<br />

Figure S1. Provides an alignment of AtMPK4, PpMPK4A and PpMPK4B (Figure 1).<br />

Figure S2. Provides a control that CHS gene expression is not <strong>in</strong>duced by <strong>the</strong> control<br />

treatment and that <strong>the</strong> expression of PpMPK4B is not affected by <strong>the</strong> knock out of<br />

PpMPK4A (Figure 4).<br />

Figure S3. Provides immunoblot analysis with anti-phospho-p44/42 MAPK with<br />

additional treatments not shown <strong>in</strong> Figure 5.<br />

Table S1. Primers used for clon<strong>in</strong>g and genotyp<strong>in</strong>g PCR of ΔPpMPK4A KO l<strong>in</strong>es<br />

(Figure 2)<br />

Table S2. Primers used for quantitative RT-PCR shown <strong>in</strong> Figure 4.<br />

131


Figure S1. Alignment of AtMPK4, PpMPK4A and PpMPK4B. Related to Figure 1.<br />

Sequence alignment of <strong>the</strong> Arabidopsis MPK4 with its two homologs <strong>in</strong> <strong>Physcomitrella</strong>,<br />

PpMPK4A and PpMPK4B, constructed us<strong>in</strong>g CLUSTALW (available at<br />

http://www.ebi.ac.uk/Tools/clustalw). Color codes for am<strong>in</strong>o acids: red = small<br />

hydrophobic (<strong>in</strong>clud<strong>in</strong>g aromatics), blue = acidic, magenta = basic, green = hydroxyl<br />

+ am<strong>in</strong>e + basic, o<strong>the</strong>rs grey. Consensus symbols denot<strong>in</strong>g degree of conservation<br />

observed <strong>in</strong> each column: "*", identical <strong>in</strong> all sequences <strong>in</strong> <strong>the</strong> alignment; ":", conserved<br />

substitutions, ".", semi-conserved substitutions.<br />

132


Figure S2. Expression of CHS upon control treatment and expression of PpMPK4B<br />

<strong>in</strong> ΔPpMPK4A. Related to Figure 4.<br />

(A) Expression of CHS <strong>in</strong> <strong>the</strong> wild type upon spray<strong>in</strong>g with 100µg/ml chitosan (blue<br />

diamonds) or with control treatment (0.01% Acetic acid, adjusted to pH 5.5 with NaOH)<br />

(red triangle). (B) Expression of PpMPK4B <strong>in</strong> WT (blue diamonds) and ΔPpMPK4A (red<br />

squares) upon spray<strong>in</strong>g with 100 µg/ml chitosan.<br />

133


Figure S3. Immunoblot analysis with anti-phospho-p44/42 MAPK. Related to<br />

Figure 5.<br />

(A) WT and ΔPpMPK4A-1 (KO1) upon spray<strong>in</strong>g with 100ug/ml chitosan or control<br />

treatment (C) (0.01% Acetic acid, adjusted with NaOH to pH 5.5). Ponceau sta<strong>in</strong><strong>in</strong>g used as<br />

total prote<strong>in</strong> load<strong>in</strong>g control. (B) WT and ΔPpMPK4A-2 (KO2) upon spray<strong>in</strong>g with 500<br />

mM NaCl or 100 µg/ml chitosan. Load<strong>in</strong>g controls show amido black sta<strong>in</strong>ed total<br />

prote<strong>in</strong>. (C) Wild type subjected to 100 µg/ml chitosan, UV light, cold or wound<strong>in</strong>g (see<br />

material and methods for details). Load<strong>in</strong>g controls show amido black sta<strong>in</strong>ed total<br />

prote<strong>in</strong>.<br />

134


Table S1. Clon<strong>in</strong>g and genotyp<strong>in</strong>g primers<br />

Name Forward primer Name Reverse primer<br />

Product<br />

size<br />

Description<br />

LBF GCGGCCGCGGGTCTTTAGATTTGCCATGT LBR GGATCCGTGTATTCCCGCGATTCTAGGT 663 Flank<strong>in</strong>g LB<br />

RBF GATATCAAGGCCTACTCTCAATTCTTTAG RBR TAGTACAGTAGGCAATGGTAGAATTGTT 888 Flank<strong>in</strong>g RB<br />

P3 TGAAAACGTCGTTGCCATTA P4 GGTCCGTATCCATCAACTCG 290 WT gDNA<br />

P1 TTCGAATCCTAAATTTGAAAACAA P2 TTTTCTCTCCTTCGTTCGTCA 1223 WT LB<br />

P5 TGATTCGTTTCCTTGCAACA P6 ATAAAAGAGGGGTGTCTGCCTGGTAGTTC 1859 WT RB<br />

P1 TTCGAATCCTAAATTTGAAAACAA P7 GGCAATGGAATCCGAGGAGGT 880 KO LB<br />

P8 GGTATCAGAGCCATGAATAGGTC P6 ATAAAAGAGGGGTGTCTGCCTGGTAGTTC 1419 KO RB<br />

P3 GGTACAAGCCACCACTTCGT P4 GGTCCGTATCCATCAACTCG 270 WT cDNA<br />

Primers used for clon<strong>in</strong>g and genotyp<strong>in</strong>g PCR of ΔPpMPK4A KO l<strong>in</strong>es, shown <strong>in</strong> Figure<br />

2. WT- wild type, LB – left boarder, RB – right boarder.<br />

Table S2. qPCR primers<br />

Name Gene accession Forward primer Reverse primer<br />

β-tubul<strong>in</strong>1 (TUB) Phypa_440500 GAGTTCACGGAAGCGGAGAG ATATCTTTCAGGCTCCACCG<br />

MPK4A Phypa_446407 GGTACAAGCCACCACTTCGT GGTCCGTATCCATCAACTCG<br />

MPK4B Phypa_435424 GGCGAGTACACGCAGTACAA ATTCACAGCGGAACACACAA<br />

Phenylalan<strong>in</strong>e Ammonia-Lyase<br />

family (PAL4) Phypa_461241 TGGCCTACTCGGTAATGGAG GTCAACCATCCGCTTGATTT<br />

Chalcone Synthases (CHS) Phypa_427901 GGCATGGAACGAGATGTTCT CCTTGCATCTTGTCCTTGGT<br />

Ethylene Response Factor 2<br />

(ERF2) Phypa_422542 GAGAGGCGTCCAAACTCTTG AGGGACTTACGGGCTTGTTT<br />

alfa dioxygenase (α-DOX) Phypa_447346 CCGCGAAGTTGCTATCTAGG AGAGGGTGGAGCCGTAATCT<br />

Lipooxygenase LOX7 Phypa_184078 GTGGCGGTTTGATCAGGA CGTTCAGCCATCCCTCTTC<br />

NADPH-oxidase (NOX) Phypa_204103 CACGATGTTGCAGTCGTTG TACGTGCCCTAGTGCCTGA<br />

Primers used for quantitative RT-PCR shown <strong>in</strong> Figure 4.<br />

135


Manuscript 2<br />

136


Review<br />

Cell Death and Differentiation (2011) 18, 1257–1262<br />

& 2011 Macmillan Publishers Limited All rights reserved 1350-9047/11<br />

www.nature.com/cdd<br />

Role of autophagy <strong>in</strong> disease resistance and<br />

hypersensitive response-associated cell death<br />

D Hofius 1,3 , D Munch 1 , S Bressendorff 1 , J Mundy 1,2 and M Petersen* ,1<br />

Ancient autophagy pathways are emerg<strong>in</strong>g as key defense modules <strong>in</strong> host eukaryotic cells aga<strong>in</strong>st microbial pathogens. Apart<br />

from actively elim<strong>in</strong>at<strong>in</strong>g <strong>in</strong>tracellular <strong>in</strong>truders, autophagy is also responsible for cell survival, for example by reduc<strong>in</strong>g <strong>the</strong><br />

deleterious effects of endoplasmic reticulum stress. At <strong>the</strong> same time, autophagy can contribute to cellular suicide. The<br />

concurrent engagement of autophagy <strong>in</strong> <strong>the</strong>se processes dur<strong>in</strong>g <strong>in</strong>fection may sometimes mask its contribution to differ<strong>in</strong>g<br />

pro-survival and pro-death decisions. The importance of autophagy <strong>in</strong> <strong>in</strong>nate immunity <strong>in</strong> mammals is well documented, but how<br />

autophagy contributes to plant <strong>in</strong>nate immunity and cell death is not that clear. A few research reports have appeared recently to<br />

shed light on <strong>the</strong> roles of autophagy <strong>in</strong> plant–pathogen <strong>in</strong>teractions and <strong>in</strong> disease-associated host cell death. We present a first<br />

attempt to reconcile <strong>the</strong> results of this research.<br />

Cell Death and Differentiation (2011) 18, 1257–1262; doi:10.1038/cdd.2011.43; published onl<strong>in</strong>e 29 April 2011<br />

Autophagy mediates <strong>the</strong> degradation of bulk prote<strong>in</strong>s and is<br />

also <strong>in</strong>volved <strong>in</strong> <strong>the</strong> clearance of damaged organelles,<br />

<strong>in</strong>soluble prote<strong>in</strong> aggregates and lipids. 1–3 Autophagic digestion<br />

and recycl<strong>in</strong>g can occur as a survival mechanism to<br />

ma<strong>in</strong>ta<strong>in</strong> cellular homeostasis and to respond to environmental<br />

stresses, such as nutrient depletion or pathogen<br />

attack, but may also function as a mediator and/or mechanism<br />

of programmed cell death. 4–8 Several subtypes of autophagy<br />

are described, but macroautophagy (hereafter termed autophagy)<br />

is <strong>the</strong> most extensively studied 9 and will be <strong>the</strong> only form<br />

described here. The process is characterized by <strong>the</strong> formation<br />

of large, double-membrane vesicles called autophagosomes.<br />

These structures arise from expand<strong>in</strong>g s<strong>in</strong>gle membranes<br />

(termed phagophores), which enclose cytoplasmic material<br />

and organelles for degradation. Completed autophagosomes<br />

fuse with <strong>the</strong> vacuole/lysosome to release <strong>the</strong> <strong>in</strong>ner s<strong>in</strong>glemembrane<br />

vesicle, called <strong>the</strong> autophagic body, <strong>in</strong>to <strong>the</strong> lumen<br />

for hydrolytic degradation and recycl<strong>in</strong>g. 2,10<br />

The mechanism of autophagy is conserved <strong>in</strong> yeast, plants<br />

and metazoans, and <strong>in</strong>volves <strong>the</strong> action of canonical<br />

autophagy related genes (ATG) that syn<strong>the</strong>size and coord<strong>in</strong>ate<br />

membrane rearrangements to allow cellular catabolism.<br />

1,2 The core sets of ATG genes seem to be present <strong>in</strong> all<br />

eukaryotes and to be essential for <strong>the</strong> autophagy pathway<br />

(Figure 1). For <strong>in</strong>stance, <strong>in</strong>duction of autophagy requires <strong>the</strong><br />

negative regulator target of rapamyc<strong>in</strong> (TOR) k<strong>in</strong>ase and <strong>the</strong><br />

ATG1 k<strong>in</strong>ase complex, which control <strong>the</strong> activity of<br />

<strong>the</strong> phosphatidyl<strong>in</strong>ositol 3-k<strong>in</strong>ase complex conta<strong>in</strong><strong>in</strong>g, for<br />

example, ATG6/Becl<strong>in</strong>1. 11 Initiation and completion of<br />

autophagosome formation <strong>in</strong>volves two ubiquit<strong>in</strong>-like conjugation<br />

systems to produce ATG12-ATG5 and ATG8-phosphatidylethanolam<strong>in</strong>e<br />

(ATG8-PE) conjugates. ATG8-PE<br />

conjugation <strong>in</strong>volves <strong>the</strong> cyste<strong>in</strong>e prote<strong>in</strong>ase ATG4 and <strong>the</strong><br />

E1-like prote<strong>in</strong> ATG7, and lipidated ATG8 is l<strong>in</strong>ked to and<br />

translocated with autophagosomes to <strong>the</strong> vacuole. 12 Therefore,<br />

conversion from soluble to lipid bound ATG8, as well as<br />

subcellular localization of green fluorescent prote<strong>in</strong> (GFP)-<br />

fused prote<strong>in</strong>, have been used to monitor temporal dynamics<br />

and spatial regulation of autophagy. 13 F<strong>in</strong>ally, recycl<strong>in</strong>g and<br />

retrieval of autophagy prote<strong>in</strong>s require <strong>the</strong> ATG9 complex,<br />

conta<strong>in</strong><strong>in</strong>g ATG2, ATG9 and ATG18. 2,10<br />

A number of excellent reviews provide more details about<br />

<strong>the</strong> molecular mechanisms of autophagy and <strong>the</strong> <strong>in</strong>dividual<br />

components required for autophagic complexes and processes<br />

7,14–17 (see also Figure 1). In this review, we focus on<br />

<strong>the</strong> role of autophagy <strong>in</strong> programmed cell death and <strong>in</strong>nate<br />

immune responses, with special emphasis on <strong>the</strong> plant<br />

hypersensitive response associated with disease resistance.<br />

Autophagy <strong>in</strong> Plants<br />

Much has been learned about <strong>the</strong> requirement for specific<br />

ATG genes <strong>in</strong> <strong>the</strong> model plant Arabidopsis. Loss-of-function<br />

mutations <strong>in</strong> ATG genes such as ATG7 and ATG5 implicate<br />

1 Department of Biology, Copenhagen University, Ole Maaloes Vej 5, Copenhagen 2200, Denmark; 2 K<strong>in</strong>g Saud University, College of Science, Riyadh 11451, Saudi Arabia<br />

*Correspond<strong>in</strong>g author: M Petersen, Department of Biology, Copenhagen University, Ole Maaloees Vej 5, Copenhagen 2200, Denmark. Tel: þ 45 3532 2137;<br />

E-mail: shutko@bio.ku.dk<br />

3 Present address: Uppsala BioCenter, Department of Plant Biology and Forest Genetics, The Swedish University of Agricultural Sciences (SLU), Box, 750 07 Uppsala,<br />

Sweden.<br />

Keywords: autophagy; ATG genes; <strong>in</strong>nate immunity; plants<br />

Abbreviations: ATG, autophagy related genes; ATG8-PE, ATG8-phosphatidylethanolam<strong>in</strong>e; DAMPs, danger-associated molecular patterns; EDS1, enhanced<br />

disease susceptibility1; GFP, green fluorescent prote<strong>in</strong>; HR, hypersensitive response; MAMP, microbial associated molecular patterns; npr1, non expressor of PR<br />

genes; PR, pathogenesis-related; R prote<strong>in</strong>s, resistance prote<strong>in</strong>s; SA, salicylic acid; TLRs, toll-like receptors; TMV, tobacco mosaic virus; TOR k<strong>in</strong>ase, target of<br />

rapamyc<strong>in</strong>; UDP, un<strong>in</strong>fected dy<strong>in</strong>g tissue<br />

Received 01.2.11; revised 09.3.11; accepted 21.3.11; Edited by J Dangl; published onl<strong>in</strong>e 29.4.11


1258<br />

Autophagy <strong>in</strong> disease resistance and HR-associated cell death<br />

D Hofius et al<br />

Figure 1 The autophagy pathway <strong>in</strong> plants. Upon <strong>in</strong>duction by environmental and developmental stimuli, macroautophagy starts by nucleation and expansion of a preautophagosomal<br />

membrane, <strong>the</strong> phagophore, which engulfs cytoplasmic material dest<strong>in</strong>ed for degradation. Autophagosomes are <strong>the</strong>n transported to and dock with <strong>the</strong><br />

vacuole, which leads to release of <strong>the</strong> <strong>in</strong>ner s<strong>in</strong>gle-membrane vesicle, <strong>the</strong> autophagic body, <strong>in</strong>to <strong>the</strong> vacuolar lumen and hydrolytic breakdown of <strong>the</strong> enclosed cargo.<br />

Autophagy <strong>in</strong>duction and vesicle nucleation require <strong>the</strong> action of <strong>the</strong> TOR k<strong>in</strong>ase and ATG1-complex, which activates <strong>the</strong> class III PI3K complex conta<strong>in</strong><strong>in</strong>g ATG6/Becl<strong>in</strong>1 and<br />

possibly o<strong>the</strong>r yet unknown <strong>in</strong>teractors. Additionally, as known from o<strong>the</strong>r eukaryotes, <strong>the</strong> ATG9 complex is required for recycl<strong>in</strong>g and retrieval of autophagy prote<strong>in</strong>s, but this<br />

function awaits verification <strong>in</strong> plants. Membrane elongation and autophagosome formation require <strong>the</strong> action of two-ubiquit<strong>in</strong>-like conjugation systems, which modify two<br />

ubiquit<strong>in</strong>-like molecules, ATG5 and ATG8, to mediate association with <strong>the</strong> phagophore and its subsequent fold<strong>in</strong>g and expansion. *ATG prote<strong>in</strong>s not identified <strong>in</strong> plants yet<br />

autophagy as a central player <strong>in</strong> cellular homeostasis. 18,19<br />

Process<strong>in</strong>g and delivery of ATG8 to <strong>the</strong> vacuole under<br />

nitrogen-starved condition requires <strong>the</strong> cyste<strong>in</strong>e protease<br />

ATG4 and <strong>the</strong> ATG12-ATG5 conjugate, 20,21 and atg5, atg7,<br />

atg10, as well as atg12a/b double mutants are hypersensitive<br />

to both nitrogen and carbon starvation. 21–23 Thus, both<br />

autophagic-related conjugation pathways seem to be required<br />

for autophagy <strong>in</strong> plants and, as <strong>in</strong> yeast and o<strong>the</strong>r models, <strong>the</strong><br />

process is required to recycle nutrients dur<strong>in</strong>g starvation.<br />

Several reports have documented <strong>the</strong> roles of autophagy <strong>in</strong><br />

plant development and under stress conditions. Dur<strong>in</strong>g<br />

senescence of Arabidopsis leaves kept <strong>in</strong> darkness (a form<br />

of carbon starvation for photosyn<strong>the</strong>tic autotrophs), autophagy<br />

seems to be responsible for degradation of <strong>the</strong><br />

chloroplasts, 24 and root development also becomes impaired<br />

<strong>in</strong> different atg mutants dur<strong>in</strong>g nitrogen starvation. 18,20<br />

Perhaps not surpris<strong>in</strong>gly, autophagy functions <strong>in</strong> <strong>the</strong> removal<br />

of oxidized prote<strong>in</strong>s dur<strong>in</strong>g oxidative stress <strong>in</strong> Arabidopsis, 25<br />

and downregulation of ATG18a us<strong>in</strong>g <strong>in</strong>terference RNA<br />

(RNAi) renders plants more sensitive to salt and drought<br />

stress. 26 Collectively, <strong>the</strong>se reports demonstrate that autophagy<br />

affects plants <strong>in</strong> many aspects of <strong>the</strong>ir life cycle.<br />

In contrast to autophagy mechanisms <strong>in</strong> yeast and<br />

mammals, <strong>in</strong>formation about <strong>the</strong> signal<strong>in</strong>g pathways trigger<strong>in</strong>g<br />

<strong>the</strong> <strong>in</strong>duction of plant autophagy <strong>in</strong> response to developmental,<br />

nutritional and environmental cues is largely lack<strong>in</strong>g.<br />

Only recently, direct genetic evidence has been provided that<br />

<strong>the</strong> TOR k<strong>in</strong>ase is a negative regulator of autophagy <strong>in</strong> higher<br />

plants. 27 Although knockout of <strong>the</strong> s<strong>in</strong>gle TOR gene <strong>in</strong><br />

Arabidopsis proved to be embryo-lethal, 28,29 knockdown by<br />

RNAi resulted <strong>in</strong> constitutive autophagy under non-stressed<br />

conditions <strong>in</strong> an ATG18-dependent fashion. 27 In addition,<br />

Tap46, <strong>the</strong> regulatory subunit of prote<strong>in</strong> phosphatase 2A, was<br />

recently identified as a downstream effector of <strong>the</strong> TOR<br />

signal<strong>in</strong>g pathway. Depletion of Tap46 reproduced <strong>the</strong><br />

signature phenotypes of TOR <strong>in</strong>activation, <strong>in</strong>clud<strong>in</strong>g autophagy<br />

<strong>in</strong>duction. 30<br />

Autophagy <strong>in</strong> <strong>Immunity</strong><br />

As autophagy has <strong>the</strong> ability to elim<strong>in</strong>ate unwanted cellular<br />

structures, it is not surpris<strong>in</strong>g that this complex and<br />

evolutionary ancient pathway also evolved to combat unwanted<br />

<strong>in</strong>tracellular microbes. That autophagy could contribute<br />

to cellular clearance of microbes was evident already <strong>in</strong><br />

<strong>the</strong> 1980s, 31 but it was first a decade later that <strong>the</strong> molecular<br />

tools to study autophagy <strong>in</strong> immune responses became<br />

available. More recently, autophagy has been shown <strong>in</strong> a<br />

number of cases to contribute to defenses aga<strong>in</strong>st microbial<br />

<strong>in</strong>vasion. For example, autophagy defends mammalian cells<br />

aga<strong>in</strong>st <strong>in</strong>vad<strong>in</strong>g Streptococcus. 32 In contrast to wild type<br />

cells, Streptococcus survives and multiplies <strong>in</strong> ATG5 deficient<br />

cells, suggest<strong>in</strong>g that <strong>the</strong> autophagic mach<strong>in</strong>ery is engaged to<br />

actively kill <strong>the</strong> bacteria. These data were supported by<br />

micrographs of Streptococci trapped <strong>in</strong>side autophagosomal<br />

structures and <strong>the</strong>se are absent <strong>in</strong> ATG5 deficient cells. 32<br />

Likewise, <strong>in</strong>duction of autophagy suppressed <strong>in</strong>tracellular<br />

survival of Mycobacterium tuberculosis <strong>in</strong> macrophages. In<br />

this case, <strong>the</strong> bacteria are also trapped <strong>in</strong>side autophagosomal-like<br />

structures positive for Becl<strong>in</strong>1. 33 S<strong>in</strong>ce <strong>the</strong>se reports,<br />

a number of excellent reviews have discussed o<strong>the</strong>r studies<br />

document<strong>in</strong>g autophagy as an <strong>in</strong>nate defense mechanism for<br />

controll<strong>in</strong>g <strong>in</strong>tracellular pathogens <strong>in</strong> mammals. 34–36<br />

In <strong>the</strong> evolutionary arms race between pathogens and <strong>the</strong>ir<br />

hosts, some pathogens have also developed mechanisms to<br />

avoid or even exploit this defense mechanism to survive and<br />

establish <strong>in</strong>fection. Shigella bacteria can escape autophagy<br />

Cell Death and Differentiation


Autophagy <strong>in</strong> disease resistance and HR-associated cell death<br />

D Hofius et al<br />

1259<br />

by secret<strong>in</strong>g effectors by means of <strong>the</strong> type III secretion<br />

system. However, mutant bacteria lack<strong>in</strong>g specific effectors<br />

become trapped by autophagy dur<strong>in</strong>g multiplication with<strong>in</strong><br />

host cells and fail to establish <strong>in</strong>fection. 37 Interest<strong>in</strong>gly,<br />

autophagosomal-like structures <strong>in</strong> human cells provide<br />

membranous supports for poliovirus RNA replication and, <strong>in</strong><br />

cells <strong>in</strong> which autophagy is <strong>in</strong>hibited via drugs or RNAi aga<strong>in</strong>st<br />

a collection of ATG genes, poliovirus yield is dim<strong>in</strong>ished. 38<br />

More recently, <strong>in</strong>creased autophagic activity has been<br />

l<strong>in</strong>ked directly to pathogen surveillance systems <strong>in</strong> different<br />

organisms. For example, mammalian Toll-like receptors<br />

(TLRs) detect microbial-associated molecular patterns<br />

(MAMP) and <strong>in</strong>duce defense responses upon ligand detection.<br />

39 Accord<strong>in</strong>gly, stimulation of TLR7 <strong>in</strong> macrophages leads<br />

to <strong>in</strong>creased autophagic activity and elim<strong>in</strong>ation of<br />

M. tuberculosis <strong>in</strong> an ATG5-dependent manner. 40 Ano<strong>the</strong>r<br />

example <strong>in</strong>cludes <strong>the</strong> pathogen receptor CD46 that b<strong>in</strong>ds<br />

Streptococci and triggers autophagy. 41 Thus, <strong>the</strong>se two<br />

examples provide evidence that pathogen recognition <strong>in</strong><br />

different animal systems is directly l<strong>in</strong>ked to higher levels of<br />

autophagic activity.<br />

Autophagy and Cell Death<br />

Apart from be<strong>in</strong>g required to tolerate nutrient deprivation and<br />

o<strong>the</strong>r stresses, autophagy also represents a cell death<br />

pathway conserved across <strong>the</strong> eukaryotic k<strong>in</strong>gdom. In 2004,<br />

Yu et al. 42 reported <strong>the</strong> requirement for ATG7 and Becl<strong>in</strong>1 <strong>in</strong><br />

certa<strong>in</strong> types of cell death <strong>in</strong> mammalian cell cultures and<br />

provided a primary example of autophagic cell death. S<strong>in</strong>ce<br />

<strong>the</strong>n, numerous reports have argued for or aga<strong>in</strong>st autophagy<br />

as a cell death mechanism, but evidence <strong>in</strong> favor of<br />

autophagic cell death has recently emerged <strong>in</strong> various genetic<br />

models. For example, <strong>the</strong> conidium of <strong>the</strong> rice blast fungus<br />

Magnapor<strong>the</strong> grisea undergoes autophagic cell death to<br />

establish an <strong>in</strong>fection and, accord<strong>in</strong>gly, M. oryzae atg8<br />

null-mutants are unable to <strong>in</strong>fect plants. 43 In Drosophila,<br />

physiological cell death of <strong>the</strong> salivary gland requires <strong>the</strong> action<br />

of ATG genes 44 and autophagy is essential for midgut cell<br />

death as well. 45 In C. elegans, necrotic breakdown of neurons<br />

is autophagy-dependent and necrotic cell death is accompanied<br />

by elevated autophagic activity. 46 A recent report<br />

documented that cell death <strong>in</strong> <strong>the</strong> formation of tracheary<br />

elements <strong>in</strong> Arabidopsis is <strong>in</strong>hibited <strong>in</strong> atg5 null-mutants and<br />

stimulated by <strong>in</strong>creased levels of autophagy. 47 Such examples<br />

<strong>in</strong>dicate that autophagy effectuates cell death pathways critical<br />

for many aspects of eukaryotic development (Figure 2).<br />

Becl<strong>in</strong>1 provides a primary example of a molecular<br />

connection between autophagy and apoptosis, an important<br />

pathway to cellular destruction <strong>in</strong> metazoans. Becl<strong>in</strong>1 is a<br />

haplo-<strong>in</strong>sufficient tumor suppressor <strong>in</strong> mice 48 <strong>in</strong>volved <strong>in</strong> <strong>the</strong><br />

<strong>in</strong>itial formation of autophagosomes. Becl<strong>in</strong>1 also forms<br />

complexes with <strong>the</strong> anti-apoptotic prote<strong>in</strong> Bcl-2 <strong>in</strong> mammalian<br />

cells 49 and loss of Becl<strong>in</strong>1 <strong>in</strong> C. elegans triggers apoptotic cell<br />

death. 50 In Arabidopsis, Becl<strong>in</strong>1 is essential for pollen<br />

germ<strong>in</strong>ation, a feature not associated with o<strong>the</strong>r ATG-null<br />

mutants, 51 but knockdown of Becl<strong>in</strong>1 through antisense<br />

or viral-<strong>in</strong>duced gene silenc<strong>in</strong>g leads to premature chlorosis and<br />

cell death <strong>in</strong> both Arabidopsis and Nicotiana benthamiana. 52,53<br />

It is <strong>the</strong>refore tempt<strong>in</strong>g to speculate that, like Becl<strong>in</strong>1 <strong>in</strong><br />

Figure 2 Examples of cell death that depend on autophagy components <strong>in</strong><br />

different eukaryotic models. In <strong>the</strong> rice blast fungus Magnapor<strong>the</strong> oryzae,<br />

autophagic cell death is required for degradation of conidia and thus fungal<br />

pathogenicity, which can be <strong>in</strong>hibited by knockout of ATG8. 43 In Drosophila<br />

melanogaster, knockouts of ATG1, ATG2 and ATG18 demonstrated <strong>the</strong><br />

requirement of autophagy for mid gut cell death. 45 In Caenorhabiditis elegans,<br />

autophagosome formation is required for necrotic cell death of neurons, which has<br />

been shown to be <strong>in</strong>hibited by knockout of ATG1 or RNAi knockdown of ATG6,<br />

ATG8 and ATG18. 46 F<strong>in</strong>ally, <strong>in</strong> Arabidopsis thaliana, development of tracheary<br />

elements depends on autophagic cell death, which is <strong>in</strong>hibited <strong>in</strong> atg5 knockout<br />

mutants 47<br />

metazoans, plant Becl<strong>in</strong>1 could also represent a molecular<br />

l<strong>in</strong>k between autophagy and ano<strong>the</strong>r cell death route. ATG5<br />

represents an additional molecular l<strong>in</strong>k between autophagy<br />

and apoptosis, because calpa<strong>in</strong>-mediated cleavage of ATG5<br />

promotes apoptosis through mitochondrial cytochrome C<br />

release and caspase activation. 54 Recently, a conjugate<br />

between ATG12 and ATG3 was shown to affect mitochondrial<br />

homeostasis and sensitize cells to apoptosis <strong>in</strong> a context<br />

completely separated from <strong>the</strong> established roles of <strong>the</strong> ATG<br />

prote<strong>in</strong>s <strong>in</strong> <strong>the</strong> autophagic pathway. 55 Collectively, <strong>the</strong>se<br />

f<strong>in</strong>d<strong>in</strong>gs imply that specific ATG genes can act as molecular<br />

switches and have autophagy-<strong>in</strong>dependent functions <strong>in</strong><br />

homeostatic processes and cell death.<br />

Autophagy <strong>in</strong> Plant <strong>Immunity</strong> and Hypersensitive Cell<br />

Death<br />

Plants rely on a multilayered <strong>in</strong>nate immune system to prevent<br />

pathogen <strong>in</strong>vasion and proliferation. Pathogen recognition<br />

can occur on <strong>the</strong> cell surface by pattern recognition receptors<br />

that detect MAMPs and <strong>in</strong>duce immune responses such as<br />

cell wall thicken<strong>in</strong>g and production of antimicrobial prote<strong>in</strong>s. 56<br />

However, diverse pathogens deliver a variety of virulence<br />

determ<strong>in</strong>ants, commonly referred to as effectors, <strong>in</strong>to plant<br />

cells to evade or suppress MAMP-triggered immunity and to<br />

manipulate <strong>the</strong> host mach<strong>in</strong>ery for <strong>the</strong>ir own benefit. 57 In turn,<br />

plants have evolved ano<strong>the</strong>r layer of defense to recognize<br />

effector modifications of host target prote<strong>in</strong>s via host<br />

surveillance prote<strong>in</strong>s (resistance (R) prote<strong>in</strong>s). These<br />

R-mediated defenses often <strong>in</strong>clude a localized programmed<br />

cell death reaction known as <strong>the</strong> hypersensitive response<br />

(HR) to limit pathogen spread. 58<br />

Several examples of <strong>the</strong> <strong>in</strong>volvement of autophagy <strong>in</strong> plant<br />

immunity and hypersensitive-related cell death have emerged<br />

Cell Death and Differentiation


1260<br />

Autophagy <strong>in</strong> disease resistance and HR-associated cell death<br />

D Hofius et al<br />

<strong>in</strong> <strong>the</strong> past few years. However, <strong>the</strong>re rema<strong>in</strong>s some doubt<br />

and apparent contradictions concern<strong>in</strong>g <strong>the</strong> function(s) of<br />

autophagy as a pro-survival or pro-death pathway. In 2005,<br />

Liu et al., 52 l<strong>in</strong>ked <strong>the</strong> activation of autophagy to <strong>in</strong>fection <strong>in</strong><br />

plants and nicely demonstrated that autophagy contributes to<br />

resistance. In addition, <strong>the</strong>y also presented evidence that<br />

autophagy was required to restrict <strong>the</strong> spread of plant<br />

hypersensitive cell death, thus function<strong>in</strong>g as a pro-survival<br />

pathway. Activation of <strong>the</strong> N-resistance gene <strong>in</strong> N. benthamiana<br />

by <strong>the</strong> p50 helicase prote<strong>in</strong> of tobacco mosaic virus (TMV)<br />

dur<strong>in</strong>g <strong>in</strong>fection or upon transient expression triggered<br />

hypersensitive cell death, and a few days after local <strong>in</strong>fection,<br />

dead cell patches became visible <strong>in</strong> non-<strong>in</strong>fected tissues on<br />

plants silenced for Becl<strong>in</strong>1. 52 A similar approach <strong>in</strong> Arabidopsis<br />

supported <strong>the</strong> observations <strong>in</strong> N. benthamiana, because<br />

activation of hypersensitive cell death via <strong>the</strong> R gene RPM1<br />

upon <strong>in</strong>fection with bacteria also led to macroscopic cell death<br />

beyond <strong>the</strong> <strong>in</strong>fection site <strong>in</strong> plants silenced for Becl<strong>in</strong>1, start<strong>in</strong>g<br />

roughly 5 days post-<strong>in</strong>fection. 53 In this context, it should be<br />

noted that cell death triggered by RPM1 is executed rapidly<br />

and ends after roughly 6–8 h. 59 This raises <strong>the</strong> question of<br />

whe<strong>the</strong>r cell death emerg<strong>in</strong>g several days later is directly<br />

connected to uncontrolled HR cell death. Never<strong>the</strong>less, <strong>the</strong><br />

data led to <strong>the</strong> hypo<strong>the</strong>sis that autophagy prevents unrestricted<br />

HR cell death by yet unknown mechanisms and thus<br />

functions as a pro-survival pathway <strong>in</strong> plant–pathogen<br />

<strong>in</strong>teractions. 60,61<br />

More recently, a pro-death function of autophagy dur<strong>in</strong>g<br />

hypersensitive cell death was reported. 62 Here, autophagic<br />

activity <strong>in</strong> <strong>the</strong> <strong>in</strong>fected tissue accompanied <strong>the</strong> onset of cell<br />

death execution triggered by some, but not all types of R<br />

prote<strong>in</strong>s. In addition, <strong>in</strong> cases <strong>in</strong> which autophagy was<br />

<strong>in</strong>duced, cell death was suppressed <strong>in</strong> local <strong>in</strong>fected tissues<br />

<strong>in</strong> different atg mutants. Strongest suppression was found for<br />

cell death conditioned by <strong>the</strong> R prote<strong>in</strong>s RPS4 and RPP1<br />

that signal through <strong>the</strong> signal<strong>in</strong>g component Enhanced<br />

Disease Susceptibility1 (EDS1). 62 HR cell death triggered<br />

by RPM1 was also significantly suppressed <strong>in</strong> atg<br />

mutants, but <strong>in</strong> this case suppression was most prom<strong>in</strong>ent <strong>in</strong><br />

comb<strong>in</strong>ation with ca<strong>the</strong>ps<strong>in</strong> <strong>in</strong>hibitors. 62 Ca<strong>the</strong>ps<strong>in</strong>s contribute<br />

to HR triggered by Phytophthora <strong>in</strong>festans <strong>in</strong><br />

N. benthamiana HR, 63 suggest<strong>in</strong>g that both autophagy and<br />

ca<strong>the</strong>ps<strong>in</strong>s are engaged <strong>in</strong> RPM1-triggered hypersensitive<br />

cell death. 62 In this context, Hatsugai et al. 64 recently reported<br />

that <strong>the</strong> 20S proteasome subunit PBA1 has caspase-3-like<br />

activity and contributes to cell death triggered by RPM1. This<br />

fur<strong>the</strong>r supported <strong>the</strong> view that RPM1 recruits autophagic<br />

mechanisms for suicidal cell death, <strong>in</strong> parallel with various<br />

o<strong>the</strong>r execution pathways. 65<br />

How can <strong>the</strong>se apparent discrepancies about <strong>the</strong> role of<br />

autophagy <strong>in</strong> HR cell death be expla<strong>in</strong>ed First, it is important<br />

to emphasize differences <strong>in</strong> <strong>the</strong> experimental systems. Liu<br />

et al. 52 primarily <strong>in</strong>vestigated plants visually and concluded<br />

that collapse of tissues next to <strong>in</strong>fection sites a few days after<br />

local <strong>in</strong>fections is caused by uncontrolled HR. Patel and<br />

D<strong>in</strong>esh-Kumar 53 performed similar assays with similar<br />

observations <strong>in</strong> Arabidopsis. In contrast, Hofius et al. 62<br />

exam<strong>in</strong>ed cell death <strong>in</strong> <strong>the</strong> actual <strong>in</strong>fection site, us<strong>in</strong>g a widely<br />

accepted electrolyte leakage assay 59 dur<strong>in</strong>g <strong>the</strong> first hours<br />

after bacterial <strong>in</strong>fection, as well as trypan blue sta<strong>in</strong><strong>in</strong>g of<br />

Figure 3 Clarification of studies on autophagic cell death <strong>in</strong> plant immunity.<br />

HR, hypersensitive response; UDP, un<strong>in</strong>fected, dy<strong>in</strong>g tissue<br />

leaves <strong>in</strong>fected with avirulent Hyaloperonospora arabidopsidis<br />

and bacteria. So, <strong>the</strong> apparently different conclusions come<br />

from studies of non-<strong>in</strong>fected tissue, days after <strong>in</strong>fection versus<br />

<strong>in</strong>fected tissue dur<strong>in</strong>g HR execution (summarized <strong>in</strong> Figure 3).<br />

In addition, <strong>the</strong> l<strong>in</strong>k between direct activation of autophagy <strong>in</strong> R<br />

gene-triggered immunity is supported by recent discoveries <strong>in</strong><br />

mammals. Here, an R prote<strong>in</strong> homolog, <strong>the</strong> cytosolic NOD1<br />

receptor that provides a surveillance system for <strong>the</strong> detection<br />

of <strong>in</strong>tracellular pathogens, recruits <strong>the</strong> autophagic prote<strong>in</strong><br />

ATG16L to <strong>the</strong> plasma membrane at <strong>the</strong> site of bacterial<br />

entry. 66<br />

In any event, <strong>the</strong>se reports also seem to underscore <strong>the</strong><br />

importance of <strong>the</strong> autophagic mach<strong>in</strong>ery <strong>in</strong> limit<strong>in</strong>g pathogen<br />

<strong>in</strong>fection <strong>in</strong> plants. Liu et al. 52 and co-workers found <strong>in</strong>creased<br />

titers of avirulent TMV <strong>in</strong> <strong>in</strong>fected tissues of Becl<strong>in</strong>1 silenced<br />

N. benthamiana plants, and Patel and D<strong>in</strong>esh-Kumar 53<br />

observed <strong>in</strong>creased susceptibility towards virulent stra<strong>in</strong>s of<br />

P. syr<strong>in</strong>gae <strong>in</strong> Becl<strong>in</strong>1 antisense Arabidopsis plants. Similarly,<br />

Hofius et al. 62 observed <strong>in</strong>creased growth of both virulent<br />

P. syr<strong>in</strong>gae and Hyaloperonospora arabidopsidis <strong>in</strong> different<br />

Arabidopsis atg mutants. Toge<strong>the</strong>r, all <strong>the</strong>se f<strong>in</strong>d<strong>in</strong>gs demonstrate<br />

that autophagy can function <strong>in</strong> plant <strong>in</strong>nate immunity.<br />

Ano<strong>the</strong>r important contribution comes from Yoshimoto et al. 67<br />

Primarily scor<strong>in</strong>g macroscopic lesions or visible death <strong>in</strong> leaves,<br />

<strong>the</strong> authors found no difference <strong>in</strong> RPM1-triggered cell death<br />

beyond <strong>the</strong> <strong>in</strong>itial <strong>in</strong>fection site <strong>in</strong> younger atg mutants. However,<br />

<strong>in</strong> older atg mutants such as atg5, <strong>the</strong>y observed lesions <strong>in</strong> non<strong>in</strong>fected<br />

tissues 6–9 days after <strong>in</strong>fection. Interest<strong>in</strong>gly, <strong>the</strong>se<br />

effects were suppressed by removal of <strong>the</strong> phytohormone<br />

salicylic acid (SA) and by mutations <strong>in</strong> non expressor of pr genes<br />

(npr1). 67 The authors thus proposed that autophagy negatively<br />

regulates cell death by controll<strong>in</strong>g NPR1-dependent SA signal<strong>in</strong>g,<br />

although it is unclear why <strong>the</strong>re is a difference between<br />

young and old atg mutants.<br />

Autophagy-deficient mutants lack <strong>the</strong> autophagic mach<strong>in</strong>ery<br />

to remove accumulat<strong>in</strong>g cellular ‘garbage’, and <strong>in</strong> contrast<br />

to younger or newly emerged leaves, older atg mutant leaves<br />

conta<strong>in</strong> higher levels of metabolites, disrupted organelles and<br />

oxidized prote<strong>in</strong>s. 24,25 Such accumulated cellular debris may<br />

well disrupt homeostasis, lead<strong>in</strong>g to pleiotropic effects<br />

Cell Death and Differentiation


Autophagy <strong>in</strong> disease resistance and HR-associated cell death<br />

D Hofius et al<br />

1261<br />

<strong>in</strong>clud<strong>in</strong>g accumulation of danger-associated molecular<br />

patterns (DAMPs) 68 trigger<strong>in</strong>g SA accumulation, and subsequent<br />

production of secreted pathogenesis-related (PR)<br />

prote<strong>in</strong>s accompanied by ER stress. Autophagy is required<br />

to dampen <strong>the</strong> deleterious effects caused by ER stress and it<br />

is well described <strong>in</strong> o<strong>the</strong>r models that atg mutants die upon<br />

<strong>in</strong>creased ER stress. 69 Thus, accumulated ER stress would<br />

be expected to <strong>in</strong>crease <strong>the</strong> susceptibility of older atg mutants<br />

to additional stresses. Interest<strong>in</strong>gly, npr1 has reduced ER<br />

stress and expression of PR genes, and mutants like bip2 die<br />

upon SA-analog treatment. 70 This would expla<strong>in</strong> why npr1<br />

(and SA-deficient mutants) rescue older atg mutants; even <strong>in</strong><br />

un<strong>in</strong>fected tissues, DAMP signals and ER stress-related<br />

effects are reduced due to reduced expression of defense<br />

genes <strong>in</strong> npr1/atg5 double mutants. Aga<strong>in</strong>, Figure 3 attempts<br />

to summarize some of <strong>the</strong> most important observation done<br />

by <strong>the</strong> different groups.<br />

Conclud<strong>in</strong>g Remarks<br />

Many questions rema<strong>in</strong> to be addressed on <strong>the</strong> roles of<br />

autophagy <strong>in</strong> plants. Because autophagy is required for<br />

cellular homeostasis, more specific autophagic functions <strong>in</strong><br />

plant–microbe <strong>in</strong>teractions and immunity are hard to unravel.<br />

For example, are plant atg mutants, <strong>in</strong> contrast to those of<br />

o<strong>the</strong>r organisms, able to cope with prolonged ER stress and, if<br />

not, what is <strong>the</strong> outcome Similarly, it may be problematic to<br />

use older atg mutants plants, because of pleiotropic effects<br />

caused by lifelong accumulation of <strong>the</strong> k<strong>in</strong>d of cellular<br />

‘garbage’ normally removed by autophagy. In addition, we<br />

need to be circumspect <strong>in</strong> <strong>the</strong> selection of autophagic mutants<br />

amenable for specific studies. For example, <strong>the</strong> strong<br />

chlorotic phenotype of Becl<strong>in</strong>1 antisense plants suggests that<br />

Becl<strong>in</strong>1 may also be <strong>in</strong>volved <strong>in</strong> o<strong>the</strong>r cell death programs, as<br />

is now apparent <strong>in</strong> metazoans. Moreover, <strong>the</strong> available<br />

collection of atg mutants needs to be fur<strong>the</strong>r explored to<br />

analyze <strong>the</strong> role of autophagy <strong>in</strong> MAMP- and effectortriggered<br />

immune responses of various host-pathogen<br />

systems. Similarly, autophagy components and mechanisms<br />

might be specifically targeted by pathogen effector prote<strong>in</strong>s to<br />

ei<strong>the</strong>r suppress defense responses or to promote pathogenicity,<br />

for example, of necrotrophic pathogens. F<strong>in</strong>ally, diverse<br />

microbial life styles and <strong>the</strong> chang<strong>in</strong>g ‘rules of engagement’ <strong>in</strong><br />

<strong>the</strong> evolutionary arms race <strong>in</strong>dicate that autophagy may be coopted<br />

for various purposes by hosts and microbes alike.<br />

Therefore, it is possible that host-derived, perimicrobial<br />

membranes 71 are actually autophagic <strong>in</strong> orig<strong>in</strong>. If so, this<br />

could represent an armistice to enable symbiosis for <strong>the</strong><br />

benefit of both microbes and plants.<br />

Conflict of <strong>in</strong>terest<br />

The authors declare no conflict of <strong>in</strong>terest.<br />

1. Lev<strong>in</strong>e B, Klionsky DJ. Development by self-digestion: molecular mechanisms and<br />

biological functions of autophagy. Dev Cell 2004; 6: 463–477.<br />

2. Mizushima N. Autophagy: process and function. Genes Dev 2007; 21: 2861–2873.<br />

3. Rodriguez-Navarro JA, Cuervo AM. Autophagy and lipids: tighten<strong>in</strong>g <strong>the</strong> knot. Sem<strong>in</strong><br />

Immunopathol 2010; 32: 343–353.<br />

4. Kourtis N, Tavernarakis N. Autophagy and cell death <strong>in</strong> model organisms. Cell Death Differ<br />

2009; 16: 21–30.<br />

5. Scarlatti F, Granata R, Meijer AJ, Codogno P. Does autophagy have a license to kill<br />

mammalian cells Cell Death Differ 2009; 16: 12–20.<br />

6. Kuma A, Mizushima N. Physiological role of autophagy as an <strong>in</strong>tracellular recycl<strong>in</strong>g system:<br />

with an emphasis on nutrient metabolism. Sem<strong>in</strong> Cell Dev Biol 2010; 21: 683–690.<br />

7. Lev<strong>in</strong>e B, Mizushima N, Virg<strong>in</strong> HW. Autophagy <strong>in</strong> immunity and <strong>in</strong>flammation. Nature 2011;<br />

469: 323–335.<br />

8. Mizushima N, Lev<strong>in</strong>e B. Autophagy <strong>in</strong> mammalian development and differentiation.<br />

Nat Cell Biol 2010; 12: 823–830.<br />

9. Yang Z, Klionsky DJ. Eaten alive: a history of macroautophagy. Nat Cell Biol 2010; 12:<br />

814–822.<br />

10. Klionsky DJ. Autophagy: from phenomenology to molecular understand<strong>in</strong>g <strong>in</strong> less than a<br />

decade. Nat Rev Mol Cell Biol 2007; 8: 931–937.<br />

11. Patt<strong>in</strong>gre S, Espert L, Biard-Piechaczyk M, Codogno P. Regulation of macroautophagy by<br />

mTOR and Becl<strong>in</strong> 1 complexes. Biochimie 2008; 90: 313–323.<br />

12. Geng J, Klionsky DJ. The Atg8 and Atg12 ubiquit<strong>in</strong>-like conjugation systems <strong>in</strong><br />

macroautophagy. ‘Prote<strong>in</strong> modifications: beyond <strong>the</strong> usual suspects’ review series. EMBO<br />

Rep 2008; 9: 859–864.<br />

13. Klionsky DJ, Abeliovich H, Agost<strong>in</strong>is P, Agrawal DK, Aliev G, Askew DS et al. Guidel<strong>in</strong>es for<br />

<strong>the</strong> use and <strong>in</strong>terpretation of assays for monitor<strong>in</strong>g autophagy <strong>in</strong> higher eukaryotes.<br />

Autophagy 2008; 4: 151–175.<br />

14. Yang Z, Klionsky DJ. Mammalian autophagy: core molecular mach<strong>in</strong>ery and signal<strong>in</strong>g<br />

regulation. Curr Op<strong>in</strong> Cell Biol 2010; 22: 124–131.<br />

15. He C, Klionsky DJ. Regulation mechanisms and signal<strong>in</strong>g pathways of autophagy. Annu<br />

Rev Genet 2009; 43: 67–93.<br />

16. Klionsky DJ, Codogno P, Cuervo AM, Deretic V, Elazar Z, Fueyo-Margareto J et al. A<br />

comprehensive glossary of autophagy-related molecules and processes. Autophagy 2010;<br />

6: PMID: 20484971.<br />

17. He C, Lev<strong>in</strong>e B. The Becl<strong>in</strong> 1 <strong>in</strong>teractome. Curr Op<strong>in</strong> Cell Biol 2010; 22: 140–149.<br />

18. Doell<strong>in</strong>g JH, Walker JM, Friedman EM, Thompson AR, Vierstra RD. The APG8/12-<br />

activat<strong>in</strong>g enzyme APG7 is required for proper nutrient recycl<strong>in</strong>g and senescence <strong>in</strong><br />

Arabidopsis thaliana. J Biol Chem 2002; 277: 33105–33114.<br />

19. Hanaoka H, Noda T, Shirano Y, Kato T, Hayashi H, Shibata D et al. Leaf senescence and<br />

starvation-<strong>in</strong>duced chlorosis are accelerated by <strong>the</strong> disruption of an Arabidopsis autophagy<br />

gene. Plant Physiol 2002; 129: 1181–1193.<br />

20. Yoshimoto K, Hanaoka H, Sato S, Kato T, Tabata S, Noda T et al. Process<strong>in</strong>g of ATG8s,<br />

ubiquit<strong>in</strong>-like prote<strong>in</strong>s, and <strong>the</strong>ir deconjugation by ATG4s are essential for plant autophagy.<br />

Plant Cell 2004; 16: 2967–2983.<br />

21. Chung T, Phillips AR, Vierstra RD. ATG8 lipidation and ATG8-mediated autophagy <strong>in</strong><br />

Arabidopsis require ATG12 expressed from <strong>the</strong> differentially controlled ATG12A and<br />

ATG12B loci. Plant J 2010; 62: 483–493.<br />

22. Thompson AR, Doell<strong>in</strong>g JH, Suttangkakul A, Vierstra RD. Autophagic nutrient recycl<strong>in</strong>g <strong>in</strong><br />

Arabidopsis directed by <strong>the</strong> ATG8 and ATG12 conjugation pathways. Plant Physiol 2005;<br />

138: 2097–2110.<br />

23. Phillips AR, Suttangkakul A, Vierstra RD. The ATG12-conjugat<strong>in</strong>g enzyme ATG10 is<br />

essential for autophagic vesicle formation <strong>in</strong> Arabidopsis thaliana. Genetics 2008; 178:<br />

1339–1353.<br />

24. Wada S, Ishida H, Izumi M, Yoshimoto K, Ohsumi Y, Mae T et al. Autophagy plays a role <strong>in</strong><br />

chloroplast degradation dur<strong>in</strong>g senescence <strong>in</strong> <strong>in</strong>dividually darkened leaves. Plant Physiol<br />

2009; 149: 885–893.<br />

25. Xiong Y, Contento AL, Nguyen PQ, Bassham DC. Degradation of oxidized<br />

prote<strong>in</strong>s by autophagy dur<strong>in</strong>g oxidative stress <strong>in</strong> Arabidopsis. Plant Physiol 2007; 143:<br />

291–299.<br />

26. Liu Y, Xiong Y, Bassham DC. Autophagy is required for tolerance of drought and salt stress<br />

<strong>in</strong> plants. Autophagy 2009; 5: 954–963.<br />

27. Liu Y, Bassham DC. TOR is a negative regulator of autophagy <strong>in</strong> Arabidopsis thaliana.<br />

PLoS ONE 2010; 5: e11883.<br />

28. Menand B, Desnos T, Nussaume L, Berger F, Bouchez D, Meyer C et al. Expression and<br />

disruption of <strong>the</strong> Arabidopsis TOR (target of rapamyc<strong>in</strong>) gene. Proc Natl Acad Sci USA<br />

2002; 99: 6422–6427.<br />

29. Deprost D, Yao L, Sormani R, Moreau M, Leterreux G, Nicolai M et al. The Arabidopsis<br />

TOR k<strong>in</strong>ase l<strong>in</strong>ks plant growth, yield, stress resistance and mRNA translation. EMBO Rep<br />

2007; 8: 864–870.<br />

30. Ahn CS, Han JA, Lee HS, Lee S, Pai HS. The PP2A regulatory subunit Tap46, a<br />

component of <strong>the</strong> TOR signal<strong>in</strong>g pathway, modulates growth and metabolism <strong>in</strong> plants.<br />

Plant Cell 2011; 23: 185–209.<br />

31. Rikihisa Y. Glycogen autophagosomes <strong>in</strong> polymorphonuclear leukocytes <strong>in</strong>duced by<br />

rickettsiae. Anat Rec 1984; 208: 319–327.<br />

32. Nakagawa I, Amano A, Mizushima N, Yamamoto A, Yamaguchi H, Kamimoto T et al.<br />

Autophagy defends cells aga<strong>in</strong>st <strong>in</strong>vad<strong>in</strong>g group A Streptococcus. Science 2004; 306:<br />

1037–1040.<br />

33. Gutierrez MG, Master SS, S<strong>in</strong>gh SB, Taylor GA, Colombo MI, Deretic V. Autophagy is a<br />

defense mechanism <strong>in</strong>hibit<strong>in</strong>g BCG and Mycobacterium tuberculosis survival <strong>in</strong> <strong>in</strong>fected<br />

macrophages. Cell 2004; 119: 753–766.<br />

34. Sumpter Jr R, Lev<strong>in</strong>e B. Selective autophagy and viruses. Autophagy 2011; 7: PMID:<br />

21150267.<br />

35. Sumpter Jr R, Lev<strong>in</strong>e B. Autophagy and <strong>in</strong>nate immunity: trigger<strong>in</strong>g, target<strong>in</strong>g and tun<strong>in</strong>g.<br />

Sem<strong>in</strong> Cell Dev Biol 2010; 21: 699–711.<br />

Cell Death and Differentiation


1262<br />

Autophagy <strong>in</strong> disease resistance and HR-associated cell death<br />

D Hofius et al<br />

36. Deretic V, Lev<strong>in</strong>e B. Autophagy, immunity, and microbial adaptations. Cell Host Microbe<br />

2009; 5: 527–549.<br />

37. Ogawa M, Yoshimori T, Suzuki T, Sagara H, Mizushima N, Sasakawa C. Escape of<br />

<strong>in</strong>tracellular Shigella from autophagy. Science 2005; 307: 727–731.<br />

38. Jackson WT, Gidd<strong>in</strong>gs Jr TH, Taylor MP, Mul<strong>in</strong>yawe S, Rab<strong>in</strong>ovitch M, Kopito RR et al.<br />

Subversion of cellular autophagosomal mach<strong>in</strong>ery by RNA viruses. PLoS Biol 2005;<br />

3: e156.<br />

39. Medzhitov R. Recognition of microorganisms and activation of <strong>the</strong> immune response.<br />

Nature 2007; 449: 819–826.<br />

40. Zhao Z, Fux B, Goodw<strong>in</strong> M, Dunay IR, Strong D, Miller BC et al. Autophagosome<strong>in</strong>dependent<br />

essential function for <strong>the</strong> autophagy prote<strong>in</strong> Atg5 <strong>in</strong> cellular immunity to<br />

<strong>in</strong>tracellular pathogens. Cell Host Microbe 2008; 4: 458–469.<br />

41. Joubert PE, Meiffren G, Gregoire IP, Pont<strong>in</strong>i G, Richetta C, Flacher M et al. Autophagy<br />

<strong>in</strong>duction by <strong>the</strong> pathogen receptor CD46. Cell Host Microbe 2009; 6: 354–366.<br />

42. Yu L, Alva A, Su H, Dutt P, Freundt E, Welsh S et al. Regulation of an ATG7-becl<strong>in</strong> 1<br />

program of autophagic cell death by caspase-8. Science 2004; 304: 1500–1502.<br />

43. Veneault-Fourrey C, Barooah M, Egan M, Wakley G, Talbot NJ. Autophagic fungal cell<br />

death is necessary for <strong>in</strong>fection by <strong>the</strong> rice blast fungus. Science 2006; 312: 580–583.<br />

44. Berry DL, Baehrecke EH. Growth arrest and autophagy are required for salivary gland cell<br />

degradation <strong>in</strong> Drosophila. Cell 2007; 131: 1137–1148.<br />

45. Denton D, Shravage B, Sim<strong>in</strong> R, Mills K, Berry DL, Baehrecke EH et al. Autophagy, not<br />

apoptosis, is essential for midgut cell death <strong>in</strong> Drosophila. Curr Biol 2009; 19: 1741–1746.<br />

46. Samara C, Syntichaki P, Tavernarakis N. Autophagy is required for necrotic cell death <strong>in</strong><br />

Caenorhabditis elegans. Cell Death Differ 2008; 15: 105–112.<br />

47. Kwon SI, Cho HJ, Jung JH, Yoshimoto K, Shirasu K, Park OK. The Rab GTPase RabG3b<br />

functions <strong>in</strong> autophagy and contributes to tracheary element differentiation <strong>in</strong> Arabidopsis.<br />

Plant J 2010; 64: 151–164.<br />

48. Yue Z, J<strong>in</strong> S, Yang C, Lev<strong>in</strong>e AJ, He<strong>in</strong>tz N. Becl<strong>in</strong> 1, an autophagy gene essential for early<br />

embryonic development, is a haplo<strong>in</strong>sufficient tumor suppressor. Proc Natl Acad Sci USA<br />

2003; 100: 15077–15082.<br />

49. Patt<strong>in</strong>gre S, Tassa A, Qu X, Garuti R, Liang XH, Mizushima N et al. Bcl-2 antiapoptotic<br />

prote<strong>in</strong>s <strong>in</strong>hibit Becl<strong>in</strong> 1-dependent autophagy. Cell 2005; 122: 927–939.<br />

50. Takacs-Vellai K, Vellai T, Puoti A, Passannante M, Wicky C, Streit A et al. Inactivation of<br />

<strong>the</strong> autophagy gene bec-1 triggers apoptotic cell death <strong>in</strong> C. elegans. Curr Biol 2005; 15:<br />

1513–1517.<br />

51. Fujiki Y, Yoshimoto K, Ohsumi Y. An Arabidopsis homolog of yeast ATG6/VPS30 is<br />

essential for pollen germ<strong>in</strong>ation. Plant Physiol 2007; 143: 1132–1139.<br />

52. Liu Y, Schiff M, Czymmek K, Talloczy Z, Lev<strong>in</strong>e B, D<strong>in</strong>esh-Kumar SP. Autophagy regulates<br />

programmed cell death dur<strong>in</strong>g <strong>the</strong> plant <strong>in</strong>nate immune response. Cell 2005; 121: 567–577.<br />

53. Patel S, D<strong>in</strong>esh-Kumar SP. Arabidopsis ATG6 is required to limit <strong>the</strong> pathogen-associated<br />

cell death response. Autophagy 2008; 4: 20–27.<br />

54. Yousefi S, Perozzo R, Schmid I, Ziemiecki A, Schaffner T, Scapozza L et al. Calpa<strong>in</strong>mediated<br />

cleavage of Atg5 switches autophagy to apoptosis. Nat Cell Biol 2006; 8: 1124–1132.<br />

55. Radoshevich L, Murrow L, Chen N, Fernandez E, Roy S, Fung C et al. ATG12 conjugation<br />

to ATG3 regulates mitochondrial homeostasis and cell death. Cell 2010; 142: 590–600.<br />

56. Dodds PN, Rathjen JP. Plant immunity: towards an <strong>in</strong>tegrated view of plant-pathogen<br />

<strong>in</strong>teractions. Nat Rev Genet 2010; 11: 539–548.<br />

57. Jones JD, Dangl JL. The plant immune system. Nature 2006; 444: 323–329.<br />

58. Nimchuk Z, Eulgem T, Holt III BF, Dangl JL. Recognition and response <strong>in</strong> <strong>the</strong> plant immune<br />

system. Annu Rev Genet 2003; 37: 579–609.<br />

59. Mackey D, Belkhadir Y, Alonso JM, Ecker JR, Dangl JL. Arabidopsis RIN4 is a target of <strong>the</strong><br />

type III virulence effector AvrRpt2 and modulates RPS2-mediated resistance. Cell 2003;<br />

112: 379–389.<br />

60. Seay M, Patel S, D<strong>in</strong>esh-Kumar SP. Autophagy and plant <strong>in</strong>nate immunity. Cell Microbiol<br />

2006; 8: 899–906.<br />

61. Yoshimoto K, Takano Y, Sakai Y. Autophagy <strong>in</strong> plants and phytopathogens. FEBS Lett<br />

2010; 584: 1350–1358.<br />

62. Hofius D, Schultz-Larsen T, Joensen J, Tsitsigiannis DI, Petersen NH, Mattsson O et al.<br />

Autophagic components contribute to hypersensitive cell death <strong>in</strong> Arabidopsis. Cell 2009;<br />

137: 773–783.<br />

63. Gilroy EM, He<strong>in</strong> I, van der Hoorn R, Boev<strong>in</strong>k PC, Venter E, McLellan H et al. Involvement<br />

of ca<strong>the</strong>ps<strong>in</strong> B <strong>in</strong> <strong>the</strong> plant disease resistance hypersensitive response. Plant J 2007; 52:<br />

1–13.<br />

64. Hatsugai N, Iwasaki S, Tamura K, Kondo M, Fuji K, Ogasawara K et al. A novel membrane<br />

fusion-mediated plant immunity aga<strong>in</strong>st bacterial pathogens. Genes Dev 2009; 23:<br />

2496–2506.<br />

65. Pajerowska-Mukhtar K, Dong X. A kiss of death–proteasome-mediated membrane fusion<br />

and programmed cell death <strong>in</strong> plant defense aga<strong>in</strong>st bacterial <strong>in</strong>fection. Genes Dev 2009;<br />

23: 2449–2454.<br />

66. Travassos LH, Carneiro LA, Ramjeet M, Hussey S, Kim YG, Magalhaes JG et al. Nod1 and<br />

Nod2 direct autophagy by recruit<strong>in</strong>g ATG16L1 to <strong>the</strong> plasma membrane at <strong>the</strong> site of<br />

bacterial entry. Nat Immunol 2010; 11: 55–62.<br />

67. Yoshimoto K, Jikumaru Y, Kamiya Y, Kusano M, Consonni C, Panstruga R et al.<br />

Autophagy negatively regulates cell death by controll<strong>in</strong>g NPR1-dependent salicylic acid<br />

signal<strong>in</strong>g dur<strong>in</strong>g senescence and <strong>the</strong> <strong>in</strong>nate immune response <strong>in</strong> Arabidopsis. Plant Cell<br />

2009; 21: 2914–2927.<br />

68. Krol E, Mentzel T, Ch<strong>in</strong>chilla D, Boller T, Felix G, Kemmerl<strong>in</strong>g B et al. Perception of <strong>the</strong><br />

Arabidopsis danger signal peptide 1 <strong>in</strong>volves <strong>the</strong> pattern recognition receptor AtPEPR1<br />

and its close homologue AtPEPR2. J Biol Chem 2010; 285: 13471–13479.<br />

69. Bernales S, McDonald KL, Walter P. Autophagy counterbalances endoplasmic reticulum<br />

expansion dur<strong>in</strong>g <strong>the</strong> unfolded prote<strong>in</strong> response. PLoS Biol 2006; 4: e423.<br />

70. Wang D, Weaver ND, Kesarwani M, Dong X. Induction of prote<strong>in</strong> secretory pathway is<br />

required for systemic acquired resistance. Science 2005; 308: 1036–1040.<br />

71. Ivanov S, Fedorova E, Bissel<strong>in</strong>g T. Intracellular plant microbe associations: secretory<br />

pathways and <strong>the</strong> formation of perimicrobial compartments. Curr Op<strong>in</strong> Plant Biol 2010; 13:<br />

372–377.<br />

Cell Death and Differentiation

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!