+ All Categories
Home > Documents > TaMDAR6 acts as a negative regulator of plant cell death and participates indirectly in stomatal...

TaMDAR6 acts as a negative regulator of plant cell death and participates indirectly in stomatal...

Date post: 24-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
16
Physiologia Plantarum 2015 © 2015 Scandinavian Plant Physiology Society, ISSN 0031-9317 TaMDAR6 acts as a negative regulator of plant cell death and participates indirectly in stomatal regulation during the wheat stripe rust–fungus interaction Mohamed Awaad Abou-Attia a,b , Xiaojie Wang a,* , Mohamed Nashaat Al-Attala a , Qiang Xu a , Gangming Zhan a and Zhensheng Kang a,* a State Key Laboratory of Crop Stress Biology for Arid Areas and College of Plant Protection, Northwest A&F University, Shaanxi, People’s Republic of China b Identification of Microorganisms and Biological Control Unit, Plant Pathology Research Institute, Agricultural Research Center, Giza, Egypt Correspondence *Corresponding authors, e-mail: [email protected]; [email protected] Received 6 January 2015; revised 19 May 2015 doi:10.1111/ppl.12355 We identified a new monodehydroascorbate reductase (MDAR) gene from wheat, designated TaMDAR6, which is differentially affected by wheat– Puccinia striiformis f. sp. tritici (Pst) interactions. TaMDAR6 is a neg- ative regulator of plant cell death (PCD) triggered by the Bax gene and Pst. Transcript levels of TaMDAR6 are significantly upregulated during a com- patible wheat– Pst interaction, indicating that TaMDAR6 may contribute to plant susceptibility. In addition, H 2 O 2 production and PCD are significantly induced and initial pathogen development is significantly reduced in the TaMDAR6 knocked-down plants upon Pst infection. Thus, the suppression of TaMDAR6 enhances wheat resistance to Pst. Besides, the suppression of TaM- DAR6 during an incompatible interaction induces a change in the morphology of stomata, which leads to poor stoma recognition and as a consequence to reduced infection efficiency. The percentage of infection sites that develop substomatal vesicles decreases in the TaMDAR6 knocked-down plants during the incompatible interaction presumably due to the increase in ROS accu- mulation, which is likely to activate other resistance mechanisms that have a negative effect on substomatal vesicle formation. TaMDAR6 can therefore be considered a negative regulator of PCD and of wheat defense to Pst. Introduction Wheat (Triticum aestivum) is one of the most important cereal crops in the world, which is the primary source of vegetable protein in human food (Chandra et al. 2014, Langridge 2012). Therefore, research on wheat has Abbreviations – AA, ascorbic acid; APX, ascorbate peroxidase; Bax, mammalian Bax gene; Bgh, Blumeria graminis; BSMV, barley stripe mosaic virus; CAT, catalase; CYR, Chinese races of Pst ; DFCI, computational biology and functional genomics laboratory database; GPX, glutathione peroxidase; GSH, glutathione; HR, hypersensitive response; MDAR, monodehydroascor- bate reductase; NCBI, National Center for Biotechnology Information; ORF, open reading frame; PCD, programmed cell death; PDS, phytoene desaturase; PR proteins, pathogenesis-related proteins; Pst , Puccinia striiformis f. sp. tritici ; pyr_redox 2, pyri- dine nucleotide-disulfide oxidoreductase 2 domain; pyr_redox, pyridine nucleotide-disulfide oxidoreductase domain; PEG, polyethylene glycol; qRT-PCR, quantitative real-time PCR; ROS, reactive oxygen species; SOD, superoxide dismutase; TaCAT, catalase; TaMDAR6, chloroplastic monodehydroascorbate reductase; TaPOD, class III peroxidase; TaPR2, -1,3-glucanase; TaPR5, thaumatin-like; TOCs, tocopherols; VIGS, virus-induced gene silencing. provided a basic understanding of this food crop to improve its productivity. By contrast, plant diseases have a negative impact on wheat production, and one of the most important diseases of wheat worldwide is the wheat stripe rust, caused by Puccinia striiformis f. sp. tritici (Pst) (Zheng et al. 2013). Allison and Isenbeck (1930) were Physiol. Plant. 2015
Transcript

Physiologia Plantarum 2015 © 2015 Scandinavian Plant Physiology Society, ISSN 0031-9317

TaMDAR6 acts as a negative regulator of plant cell death andparticipates indirectly in stomatal regulation during thewheat stripe rust–fungus interactionMohamed Awaad Abou-Attiaa,b, Xiaojie Wanga,*, Mohamed Nashaat Al-Attalaa, Qiang Xua,Gangming Zhana and Zhensheng Kanga,*

aState Key Laboratory of Crop Stress Biology for Arid Areas and College of Plant Protection, Northwest A&F University, Shaanxi, People’s Republic ofChinabIdentification of Microorganisms and Biological Control Unit, Plant Pathology Research Institute, Agricultural Research Center, Giza, Egypt

Correspondence*Corresponding authors,e-mail: [email protected];[email protected]

Received 6 January 2015;revised 19 May 2015

doi:10.1111/ppl.12355

We identified a new monodehydroascorbate reductase (MDAR) genefrom wheat, designated TaMDAR6, which is differentially affected bywheat–Puccinia striiformis f. sp. tritici (Pst) interactions. TaMDAR6 is a neg-ative regulator of plant cell death (PCD) triggered by the Bax gene and Pst.Transcript levels of TaMDAR6 are significantly upregulated during a com-patible wheat–Pst interaction, indicating that TaMDAR6 may contribute toplant susceptibility. In addition, H2O2 production and PCD are significantlyinduced and initial pathogen development is significantly reduced in theTaMDAR6 knocked-down plants upon Pst infection. Thus, the suppression ofTaMDAR6 enhances wheat resistance to Pst. Besides, the suppression of TaM-DAR6 during an incompatible interaction induces a change in the morphologyof stomata, which leads to poor stoma recognition and as a consequence toreduced infection efficiency. The percentage of infection sites that developsubstomatal vesicles decreases in the TaMDAR6 knocked-down plants duringthe incompatible interaction presumably due to the increase in ROS accu-mulation, which is likely to activate other resistance mechanisms that have anegative effect on substomatal vesicle formation. TaMDAR6 can therefore beconsidered a negative regulator of PCD and of wheat defense to Pst.

Introduction

Wheat (Triticum aestivum) is one of the most importantcereal crops in the world, which is the primary sourceof vegetable protein in human food (Chandra et al.2014, Langridge 2012). Therefore, research on wheat has

Abbreviations – AA, ascorbic acid; APX, ascorbate peroxidase; Bax, mammalian Bax gene; Bgh, Blumeria graminis; BSMV,barley stripe mosaic virus; CAT, catalase; CYR, Chinese races of Pst; DFCI, computational biology and functional genomicslaboratory database; GPX, glutathione peroxidase; GSH, glutathione; HR, hypersensitive response; MDAR, monodehydroascor-bate reductase; NCBI, National Center for Biotechnology Information; ORF, open reading frame; PCD, programmed cell death;PDS, phytoene desaturase; PR proteins, pathogenesis-related proteins; Pst, Puccinia striiformis f. sp. tritici; pyr_redox 2, pyri-dine nucleotide-disulfide oxidoreductase 2 domain; pyr_redox, pyridine nucleotide-disulfide oxidoreductase domain; PEG,polyethylene glycol; qRT-PCR, quantitative real-time PCR; ROS, reactive oxygen species; SOD, superoxide dismutase; TaCAT,catalase; TaMDAR6, chloroplastic monodehydroascorbate reductase; TaPOD, class III peroxidase; TaPR2, 𝛽-1,3-glucanase;TaPR5, thaumatin-like; TOCs, tocopherols; VIGS, virus-induced gene silencing.

provided a basic understanding of this food crop toimprove its productivity. By contrast, plant diseases havea negative impact on wheat production, and one of themost important diseases of wheat worldwide is the wheatstripe rust, caused by Puccinia striiformis f. sp. tritici (Pst)(Zheng et al. 2013). Allison and Isenbeck (1930) were

Physiol. Plant. 2015

the first to establish the existence of races in Pst based ondifferential susceptibility of wheat cultivars. China is thelargest epidemic region in the world (Stubbs 1988, Asadet al. 2012); the most destructive epidemics of stripe rustoccurred in 1950, 1964 and 1990, caused yield lossesof 29.3, 13.3 and 1.8% of the national total production,respectively (Li and Zeng 2000).

Pst is a biotrophic fungus that infects living cells andtakes up water and nutrients from the living host. Infec-tion can occur from the seedling to maturity stage. Thefungus forms yellow to orange colored pustules (ure-dinia), and each uredinium contains thousands of ure-diospores, which spread from the pustules by wind.Infection requires high humidity for 4–6 h at 10–15∘C.Compared to other rust fungi, Pst prefers a lower tem-perature for development, which limits this fungus as amajor disease in many areas of the world (Stubbs 1985,Wellings and McIntosh 1990).

Reactive oxygen species (ROS) are toxic by-productsof many aerobic metabolic processes. Moreover, vari-ous stress conditions cause increases in ROS generation,leading to oxidative damage of lipids, proteins and DNA(Apel and Hirt 2004). Under biotic stress, ROS act assignaling molecules to activate pathogenesis-related pro-teins and systemic acquired resistance in cells adjacentto the infection site to prevent further pathogen spread(Draper 1997). ROS serve as signaling molecules at lowlevels, but can also induce cell death at high levels(Sharma et al. 2012). The production of ROS is normallycounter-acted by an enzymatic anti-oxidative system[catalase (CAT); ascorbate peroxidase (APX); glutathioneperoxidase (GPX); and superoxide dismutase (SOD)],and by a non-enzymatic anti-oxidative system [ascorbicacid (AA); glutathione (GSH); tocopherols (TOCs); andphenolic compounds] to protect plant cells against ROStoxicity. The monodehydroascorbate reductase (MDAR)enzyme is involved in the ascorbate–glutathione cycle,and plays an important role in directly reducing monode-hydroascorbate (oxidized ascorbate) to ascorbate usingNAD(P)H as an electron donor (Apel and Hirt 2004).

The roles of MDAR have been extensively reportedunder abiotic stress such as ozone, salt, polyethyleneglycol (PEG) (Sharma and Davis 1997, Eltayeb et al.2007) and drought (Sharma and Dubey 2005), whichshowed that MDAR genes are regulated by abioticstresses to reduce oxidative damages through maintain-ing redox status (Eltelib et al. 2011). Therefore, MDARgenes have been used as indicators of plant resistanceto several abiotic stresses (Ali et al. 2005, Sharma andDubey 2005). For example, in Brassica campestris, tran-script levels of BcMdhar were strongly upregulatedin response to oxidative stress (Yoon et al. 2004). InPhyscomitrella patens, PpMDHAR1 and PpMDHAR3

were induced during salt stress and osmotic stress (Lundeet al. 2006). In contrast, few studies have examined theroles of MDAR under biotic stress, particularly duringthe wheat–Pst interactions. In our previous studies, weidentified two members encoding MDAR from wheat(TaMDHAR4 and TaMDHAR), which demonstrated thatthe suppression of TaMDHAR4 and TaMDHAR enhanceswheat resistance to Pst (Feng et al. 2014a). Besides,TaMDHAR is regulated by miRNA PN-2013 during Pstinfection (Feng et al. 2014b). TaMDHAR is a cytoplas-mic protein. In addition, TaMDHAR4 has been consid-ered a peroxisomal protein (Feng et al. 2014a,2014b).By contrast, TaMDAR6 is a chloroplastic protein. Plantchloroplasts are the most significant generators of ROS(Ishikawa and Shigeoka 2008, Maruta et al. 2012), andare important elements in the regulation of programmedcell death (PCD), which is an essential component ofplant defense to pathogen attacks. Thus, the sub-cellularlocalization of TaMDAR6 in plant chloroplasts opens thepossibility of the involvement of TaMDAR6 in the regula-tion of PCD. Therefore, TaMDAR6 was specially investi-gated. Our findings demonstrated that TaMDAR6 acts asa negative regulator of PCD. In addition, we reported forthe first time the involvement of TaMDAR6 gene in stom-atal regulation and substomatal vesicles formation duringthe wheat–Pst interaction. Therefore, this study providesnew insights into the role of TaMDAR in the wheat–Pstinteraction.

Materials and methods

Plant material and inoculation

Compatible and incompatible interactions were estab-lished using the wheat cultivar Suwon 11 and twoChinese races of Pst (CYR23 and CYR31). Suwon 11shows high susceptibility to CYR31 and high resistanceto CYR23 (Wang et al. 2010). Plant cultivation andinoculation with Pst were performed according to Kanget al. (2003). Inoculated leaves were sampled, quicklyfrozen in liquid nitrogen and stored at −80∘C until usedfor RNA isolation.

RNA extraction and first strand cDNA synthesis

Total RNA (3 μg) was extracted using the BIOZOL TotalRNA Extraction Reagent (Bioer Technology CompanyLimited, Binjiang, China) following the manufacturer’sprotocol. Any DNA contamination was removed bytreatment with DNase1 before synthesis of cDNA.Extracted RNA was reversely transcribed into cDNAusing the i-SCRIPT Kit (Bio-Rad, Hercules, CA) accordingto the manufacturer’s instructions.

Physiol. Plant. 2015

Cloning and sequencing

The full-length sequence was obtained by screen-ing of our cDNA database of wheat–Pst interactions(Ma et al. 2009) and the wheat EST database in theNational Center for Biotechnology Information (NCBI).Homologous sequences were assembled using theCAP3 (http://pbil.univ-lyon1.fr/cap3.php/) and multiplesequence alignments were performed using the DNAMAN

6.0 software (Lynnon BioSoft, Pointe-Claire, Canada).Specific primers were designed according to the assem-bled sequence using the PRIMER 5.0 software (Table S1,Supporting Information). The full length open readingframe (ORF) was amplified from cDNA synthesis usingRNA isolated from wheat cv. Suwon 11 leaves afterinoculation with Pst CYR23 (incompatible interaction)or CYR31 (compatible interaction). Amplification wasconducted in a DNA thermocycler (Bio-Rad) usingthe following program: 95∘C for 3 min; 35 cycles of95∘C for 30 s, 55∘C for 30 s, and 72∘C for 70 s; and72∘C for 10 min. The amplified ORF was ligated intothe pGEM T-easy vector (Promega, Madison, WI) fol-lowing the manufacturer’s protocol and transformedinto competent Escherichia coli (JM109) by the CaCl2procedure (Ausubel et al. 1997). Sequencing was per-formed using an ABI PRISM 3130XL Genetic Analyzer(Applied BioSystems, Foster City, CA). The amino acidsequence and conserved domains were analyzed usingNCBI (http://www.ncbi.nlh.nih.gov/gorf/gorf.html), Inter-ProScan (http://www.ebi.ac.uk/cgi-bin/iprscan/), Com-pute pI/MW (http://web.expasy.org/compute_pi/) andProtParam (http://www.expasy.org/tools/pi_tool.html).The prediction of chromosomal location of the corre-sponding gene was performed using the InternationalWheat Genome Sequencing Consortium database(http://wheat-urgi.versailles.inra.fr/Seq-Repository/BLAST).The sequence is deposited at GenBank under accessionnumber KP201873.

Quantitative real-time PCR

TaMDAR6 transcript levels were analyzed by qRT-PCR(7500 Real-Time PCR System, Applied Biosystems) usingsynthesized cDNA, as previously explained, at differenttime points after inoculation (12, 18, 24, 48 and 120hpi) and primer pairs (Table 1) specific to the gene ofinterest. Previous time points were chosen according toour previous studies of the interactions between ChinasePst races (CYR31, CYR23) and wheat cv. Suwon 11(Wang et al. 2007). The RT-PCR (real-time PCR) Systemwas programmed as follows: 95∘C for 1 min; 40 cyclesat 95∘C for 10 s, 60∘C for 20 s, and 72∘C for 40 s; 1cycle at 95∘C for 15 s, 60∘C for 1 min and 95∘C for 15 s;and finally 60∘C for 15 s. The relative expression of the

investigated gene was normalized using the referencegene TaEF-1 (GenBank accession number: Q03033),and relative expression was estimated using the 2–ΔΔCT

method (Livak and Schmittgen 2001). Data are themeans of three independent experiments.

Sub-cellular localization of TaMDAR6in Nicotiana benthamiana

To construct the fusion vector pCaMV35S:TaMDAR-GFP,the complete ORF of TaMDAR6 was amplified using spe-cific primers that have restriction sites for SpeI and AvrII(Table 1). The digested amplicon was ligated into the 5′

end of the green fluorescent protein (GFP) coding regionof pCaMV35S:GFP. The newly constructed plasmid wasintroduced into Agrobacterium tumefaciens GV3101through electroporation (Wise et al. 2006). Transformedcells were infiltrated into Nicotiana benthamiana leaves(Van der Hoorn et al. 2000). GFP signals were detectedusing an Olympus BX-51 fluorescence microscope(Olympus Corp., Tokyo, Japan). The experiment wasperformed twice and with three replicates each time.

Agrobacterium-mediated transientgene expression

PVX:BAX, which was kindly provided by Dr D. Dou(Purdue University, West Lafayette, IN), was digestedwith ClaI and XmaI. The BAX fragment was replacedwith GFP to construct PVX:GFP or with TaMDAR6to construct the PVX:TaMDAR (Fig. S3). Agrobacteriacarrying respective plasmids were cultured in 5 ml ofLB medium, supplemented with kanamycin, rifampicinand gentamycin. Cells were harvested at an OD600of 0.6 to 1.2 and resuspended in 10 mM MgCl2. Bac-terial suspensions were adjusted to a final density of0.2–0.5 at OD600. Transient expression assays wereperformed using 4- to 5-week-old N. benthamianaplants. Suspensions were infiltrated into N. benthamianaleaves (abaxial surface) using a 2-ml disposable syringewithout a needle (Van der Hoorn et al. 2000). At leastsix leaves on different plants were used in infiltrationexperiments; each leaf was divided into four sites theninfiltrated with A. tumefaciens cell suspensions as fol-lows; first and second sites were infiltrated with A.tumefaciens carrying PVX:00 (empty vector), third andfourth sites were infiltrated with A. tumefaciens carrying(PVX:TaMDAR). Then the infiltrated leaf was incubatedat 25∘C. After 2 days, the same areas were infiltratedagain as follows; first and third sites were infiltrated withA. tumefaciens carrying (PVX:00), second and fourthsites were infiltrated with A. tumefaciens carrying(PVX:BAX). A. tumefaciens strains carrying PVX:GFP

Physiol. Plant. 2015

Table 1. Effect of TaMDAR6 on programmed cell death induced by the mouse Bax gene in Nicotiana benthamiana. aCo-infiltration was performedby infiltrating cultures in Nicotiana benthamiana leaves. bNecrosis activity at 5, 6, 7 and 8 dpi with PVX:EV+ PVX:EV (empty vector as negative control);PVX:BAX+ PVX:EV (as positive control); PVX:TaMDAR+PVX:EV and PVX:BAX +PVX:TaMDAR (co-infiltration); − = no response; + = very weak necrosis;++ = weak necrosis; +++ moderate necrosis; and ++++ = severe necrosis. cNecrotic area % = percentage of necrotic area of the total infiltratedregion.

Symptoms Necrosis activityb Necrotic area %c

Constructs 5 dpi 6 dpi 7 dpi 8 dpi 5 dpi 6 dpi 7 dpi 8 dpi 5 dpi 6 dpi 7 dpi 8 dpi

PVX:EV & PVX:EV(co-infiltration,negative ck)

No visible No visible No visible No visible - - - - 0 0 0 0

PVX:BAX & PVX:EV(co-infiltration,positive ck)

Necrosis Necrosis Necrosis Necrosis ++++ ++++ ++++ ++++ 100 100 100 100

PVX:TaMDAR &PVX:EV(co-infiltration)

No visible No visible No visible No visible - - - - 0 0 0 0

PVX:BAX &PVX:TaMDAR(co-infiltration)a

Very smallnecroticlesions

Small necroticlesions

Large necroticlesions

Necrosis + ++ +++ ++++ 10 30±5 80±5 100

were used to confirm the efficiency of our experiments.Pictures were taken 5–8 days after the last infiltration.The experiment was repeated three times.

Virus-induced gene silencing (VIGS)

VIGS was mediated by the barley stripe mosaic virus(BSMV). BSMV vectors (𝛼, 𝛽, 𝛾 and BSMV:PDS4as) wereprovided by Dr Scofield (Purdue University). A139 bpfragment was chosen and amplified from a plasmid con-taining the TaMDAR6 gene using specific primers withrestriction sites for NotI and PacI (Table 1, Fig. S6). ThePCR product was cloned into the pGEM T-simple vector.Extracted plasmids were digested with NotI and PacI, fol-lowed by replacement of the TaPDS coding sequence inBSMV:TaPDS to construct plasmid BSMV:TaMDARas forgene silencing according to Holzberg et al. (2002). Lin-earized plasmids were transcribed into RNA using themMessage T7 in vitro transcription kit (Ambion, Austin,TX) following the manufacturer’s instructions.

The second leaves of wheat seedlings were gentlyinoculated by rubbing the leaf surfaces from base totip two times using a mixture of each transcriptionalreaction. In total, 30 plants were used for each treatment(5 plants/pot), and another 15 plants were inoculatedwith 1× FES buffer (0.1 M glycine, 0.06 M K2HPO4, 1%w/v tetrasodium pyrophosphate, 1% w/v bentonite and1% w/v celite, pH 8.5) as control. Post 9 days virus inoc-ulation, the fourth leaf was inoculated with urediosporesof Pst CYR23 or CYR31. Infection types of Pst wererecorded 15 dpi. The fourth leaves were also sampledfor RNA isolation and histological observation. Relativetranscript levels of TaMDAR6, pathogenesis-related

protein genes (TaPR2, DQ090946; TaPR5, FG618781)and reactive oxygen species-related genes (TaCAT, cata-lase, X94352; TaPOD, class III peroxidase, TC303653)were assessed using qRT-PCR as described above. Theexperiment was repeated three times.

Histological analysis

Samples were stained as described (Wang et al. 2007).According to Parlevliet (1986), aborted substomatalvesicles were considered aborted penetration attempts.Therefore, only infection sites where substomatal vesi-cles had formed were considered. Fungal developmentand host responses for each treatment (a minimum of 50infection sites for each time point) were observed underan Olympus BX-51 microscope (Olympus Corp., Tokyo,Japan) using bright field and UV light. Auto-fluorescenceof mesophyll cells in infected leaves was measuredas necrotic area using epifluorescence microscopy(excitation filter, 485 nm; dichromic mirror, 510 nmand barrier filter, 520 nm). The percentage of stomatathat remained open was determined from records of atleast 100 stomata observed on four independent leaves.Hyphal length, hyphal branches and necrotic area werecalculated using DP-BSW software. Statistical analyseswere performed using SPSS software.

Results

Cloning and sequence analyses of TaMDAR6

Specific primers were designed to amplify the cDNAfragments of the corresponding gene. The full-length

Physiol. Plant. 2015

cDNA sequence was isolated from cDNA of wheatcv. Suwon 11, which is highly susceptible to viru-lent Pst race CYR31 and highly resistant to avirulentPst race CYR23 (Wang et al. 2010). The predictedORF of TaMDAR6 spans 1458 bp and encodes a pro-tein of 485 amino acids with a predicted molecularmass of 52.04 kDa. The prediction of protein domainsusing InterProScan and NCBI databases indicated thatTaMDAR6 protein contains pyridine nucleotide-disulfideoxidoreductase domains (pyr redox 2, pyr redox andNAD(P)-binding Rossmann-like domain) (Fig. 1, Fig. S1).In addition, the corresponding gene is a homolog ofAtMDAR6, which encodes an MDAR protein localizedin chloroplasts of Arabidopsis. Hence, the correspond-ing gene was designated as TaMDAR6. The predictionof chromosomal location using the International WheatGenome Sequencing Consortium database (http://wheat-urgi.versailles.inra.fr/Seq-Repository/BLAST) indi-cated that there were three copies in the wheat genome,located on chromosomes 7AL, 7BL and 7DL, respec-tively, indicating that the TaMDAR6 gene family maybe located in the homologous group 7 chromosomes.In addition, phylogenetic analysis and cDNA multiplesequences alignment both revealed that our correspond-ing gene of this study is located on chromosome 7DL(Fig. 2, Fig. S2). Moreover, phylogenetic analysis (Fig. 2,Table S2) revealed that TaMDAR6 is most similar toTaMDAR6-a, BdMDAR chloroplastic isoforms (X1 andX2) from Brachypodium distachyon, and TaMDAR6-b,respectively, but showed the lowest similarity with otherwheat monodehydroascorbate reductase (TaMDHAR4and TaMDHAR).

TaMDAR6 is localized in chloroplastsof N. benthamiana

Agrobacteria carrying pCaMV35S:TaMDAR-GFP wereinfiltrated into N. benthamiana leaves to determinethe sub-cellular localization of TaMDAR6. Fluorescencemicroscopic analysis revealed that green fluorescence ofTaMDAR-GFP was only detected in chloroplast, totallymatched with the red autofluorescence of chlorophyll.This result indicated that TaMDAR-GFP is localized inthe chloroplast of N. benthamiana (Fig. 2B).

TaMDAR6 has a negative effect on programmedcell death (PCD) induced by the Bax gene

To improve our understanding concerning TaMDAR6function in PCD at the molecular level, TaMDAR6 tran-sient expression was performed through the infiltration ofA. tumefaciens into N. benthamiana leaves. To evaluatethe efficiency of our approach, the transient expression

of the GFP gene was investigated. N. benthamiana leaveswere infiltrated with A. tumefaciens carrying PVX:GFP.Fluorescence microscopic analysis revealed that greenfluorescence of PVX:GFP was detected in N. benthami-ana cells. These results indicated that GFP was success-fully expressed, which demonstrated that our transientexpression system was functional (Fig. S4).

N. benthamiana leaves were infiltrated withA. tumefaciens carrying PVX:00 empty vector+ PVX:00empty vector, or PVX:BAX+PVX:00 empty vector, orPVX:TaMDAR+ PVX:00 empty vector, or PVX:Bax+PVX:TaMDAR. As shown in Fig. 3 and in Table 1, nonecrotic area occurred at areas infiltrated with PVX:00+PVX:00, or PVX:TaMDAR+ PVX:00. By contrast, areasinfiltrated with A. tumefaciens carrying PVX:Bax+PVX:00 became completely necrotic within 5 dpi. Todetermine whether TaMDAR6 enhances or suppressesPCD triggered by Bax, A. tumefaciens cultures carryingPVX:Bax and PVX:TaMDAR vectors were infiltrated intoN. benthamiana leaves at the same site. The percentageof infiltrated leaf area that had become necrotic at 5, 6,7 and 8 dpi increased over time, however, this increasewas much slower than observed for the positive control(PVX:Bax+ PVX:00) (Table 1). Only extremely smallnecrotic lesions were observed 6 dpi at the infiltrationarea with PVX:Bax+PVX:TaMDAR, and these lesionsconstantly evolved to include all of the infiltrated areaafter 8 dpi. This result indicates that the area infiltratedwith PVX:Bax+ PVX:TaMDAR became completelynecrotic after 8 dpi, which was much slower than thenecrosis that occurred with PVX:Bax+PVX:00 (Fig. 3).

Transcriptional responses of TaMDAR6to Pst infection

In the incompatible interaction, transcript levels of TaM-DAR6 remained at the control level from 12 to 18 hpi, atranscript peak was observed at 24 hpi, with an approx-imately 2.2-fold increase, and subsequently a return tocontrol levels at 120 hpi (Fig. 4). In the compatible inter-action, transcript levels of TaMDAR6 were changed alittle at 12 to 18 hpi. However, t ranscript levels werestrongly upregulated at 24 and 48 hpi, by approximately3.9- and 4.7-fold, respectively. At 120 hpi, the controllevel was reached again (Fig. 4). Therefore, the relativetranscript level of TaMDAR6 in the compatible interac-tion was much higher than that in the incompatible inter-action, at least at 24 and 48 hpi (Fig. 4).

Suppression of TaMDAR6 enhances wheatresistance to Pst

A virus-induced gene silencing (VIGS) system estab-lished in cv. Suwon 11 was used to evaluate the function

Physiol. Plant. 2015

Fig. 1. Alignment of TaMDAR6 (7DL) with monodehydroascorbate reductase proteins from Brachypodium distachyon (BdMDAR chloroplastic isoformsX1 and X2), Triticum aestivum (TaMDAR6-a (7AL), TaMDAR6-b (7BL), TaMDHAR, TaMDHAR4) and Arabidopsis thaliana (AtMDAR6). Black shadow:amino acids are conserved in all sequences, pink shadow: amino acids are conserved among at least six sequences, light blue shadow: amino acidsare conserved between at least five sequences, yellow shadow: amino acids are conserved between at least three sequences. Red line: Pyr_redox2=pyridine nucleotide-disulfide oxidoreductase 2 domains, blue line: NAD(P)-binding Rossmann-like domain, and green line: Pyr_redox=pyridinenucleotide-disulfide oxidoreductase domains. Sequence alignment was performed using DNAMAN6 software.

of TaMDAR6 gene during the interaction between wheatand Pst (Fig. S5). To evaluate the efficiency of our VIGSsystem, we tested the silencing of the wheat phytoenedesaturase gene (PDS). The second leaves of a two-leafwheat seedling was inoculated with BSMV:TaPDSas orBSMV:𝛾 (empty vector) as positive control, or FAS bufferas negative control. As shown in Fig. 5A, mild virussymptoms occurred on the new leaves at 15 dpi. Mean-while, BSMV:TaPDSas inoculated plants showed largephoto-bleached areas on the fourth leaf, while there wereno such symptoms on plants treated with FAS buffer,which showed that our VIGS system was functional.

To make the silencing specific to TaMDAR6 (7DL), a139 bp fragment was chosen and amplified using specificprimers (Table 1, Fig. S6). The fragment located at the 5′

untranslated region (UTR) and the origin of the ORF. Rig-orous efforts were made to choose the VIGS fragment toget a specific silencing construct. Nucleotide sequencesof wheat MDAR genes were downloaded from NCBI andalignment analysis was conducted. Alignment analysisindicated no homology between the TaMDAR6 gene andother wheat MDAR genes in the VIGS fragment (Fig. S6).Al-Attala et al. (2014) and Holzberg et al. (2002) sup-ported the view that a sequence with 81% identity or less

Physiol. Plant. 2015

Fig. 2. Phylogenetic and sub-cellular localization of TaMDAR6. (A) Phy-logenetic tree of TaMDAR6 and other monodehydroascorbate reduc-tase proteins: Brachypodium distachyon (BdMDAR chloroplastic iso-forms X1 and X2); Triticum aestivum (TaMDAR6, TaMDAR6-a (7AL),TaMDAR6-b (7BL), TaMDAR6-d (7DL), TaMDHAR, TaMDHAR4); Vitisvinifera (VvMDAR chloroplastic-like); Solanum lycopersicum (SlMDARchloroplastic-like); Nicotiana tabacum (NtMDAR); Cucumis sativus(CsMDAR chloroplastic-like) and Arabidopsis thaliana (AtMDAR6). TaM-DAR6 is our sequence, the sequences of TaMDAR-d (7DL), TaMDAR6-a(7AL) and TaMDAR6-b (7BL) were obtained from the InternationalWheat Genome Sequencing Consortium. The phylogenetic tree wasconstructed using DNAMAN6 software based on the maximum likelihoodmethods (JTT model); bootstrap values (>50%) of 100 replicates areindicated at each branch. (B) Sub-cellular localization of TaMDAR6 inNicotiana benthamiana. Red color is the autofluorescence of chlorophyll(B1, B3); green fluorescence shows the localization of TaMDAR-GFP (B2);yellow indicates co-localization of fusion protein and chlorophyll contain-ing chloroplasts (B3); (B3) Merger of (B1) and (B2). Scale bars=20 μm.

does not cause targeted gene silencing. Therefore, thesecond leaves of a two-leaf tested plant were inoculatedwith BSMV:TaMDARas or with BSMV:𝛾 (empty vector ascontrol). All BSMV-infected plants displayed mild virussymptoms. The fourth leaves were inoculated with aviru-lent race CYR23 for the incompatible interaction or withvirulent race CYR31 for the compatible interaction.

To clarify whether TaMDAR6 was successfullysilenced, transcript levels of all three TaMDAR6 copies(7AL, 7BL, 7DL) were examined using quantitative RT-PCR (qRT-PCR). In comparison to BSMV:𝛾-infectedplants, the expression level of TaMDAR6 (7DL) wasdecreased in the fourth leaves of BSMV:TaMDARas-infected plants at 0, 24, 48 and 120 hpi with Pst by

Fig. 3. Transient overexpression of TaMDAR6 in Nicotiana benthami-ana leaves. Transient expression of TaMDAR6 delayed programmedcell death induced by the Bax gene. A–D comparison betweenphenotypes of localized HR induced by (PVX:BAX+ PVX:EV) and by(PVX:Bax+ PVX:TaMDAR) at different time points (A) 5 dpi, (B) 6 dpi, (C)7 dpi and (D) 8 dpi. Black circles mark the positions of infiltrated areas.

approximately 80, 88, 90.5 and 80% for the incom-patible interaction and by 80, 89.7, 92.2 and 80%for the compatible interaction, respectively (Fig. S7).These results indicated that TaMDAR6 was successfullysilenced. By contrast, transcript levels of other twocopies TaMDAR6-a (7AL) and TaMDAR6-b (7BL) werenot affected (Fig. S7), indicating that the VIGS constructwas very specific to TaMDAR6 (7DL).

Disease development was scored at 15 dpi with Pstraces to evaluate the effect of gene silencing on thewheat–Pst interaction. BSMV:𝛾-infected plants (emptyvector) had no changes in their resistance or susceptibil-ity against CYR31 or CYR23 (Fig. 5B). In the incompatibleinteraction, both treatments (BSMV:TaMDARas andBSMV:𝛾) had no changes in their resistance (infec-tion type 1) according to the 0–9 scale of McNealet al. (1971), HR areas were elicited in both treat-ments with BSMV:TaMDARas and BSMV:𝛾, and nouredosori formed (Table 2, Fig. 5B). In the compatibleinteraction, although heavy sporulation was observedin both treatments (BSMV:TaMDARas and BSMV:𝛾),

Physiol. Plant. 2015

Fig. 4. Relative transcript levels of TaMDAR6 in response to Pst infection in compatible and incompatible interactions at 12, 18, 24, 48 and 120 hpi.Transcripts were quantified using the comparative threshold (2–ΔΔCT) method. Means and standard deviations were calculated using data from threeindependent biological replicates.

BSMV:TaMDARas-infected plants showed a higherdegree of resistance (infection type 8) than control plants(infection type 9) according to the 0–9 scale of McNealet al. (1971). Moreover, PCD areas were elicited aroundthe infection sites in the BSMV:TaMDARas-infectedplants, but no such cell death areas were observed inthe control (Table 3, Fig. 5B).

To demonstrate that the increases of PCD was involvedin the resistance response; the transcriptional responsesof some defense-related genes were examined, becausethose genes have been used as indicators of HR and arenecessary for plant resistance (Schaffrath et al. 1997; VanLoon and Van Strien 2002). The transcriptional responsesof defense-related genes (TaPR2, TaPR5, TaCAT andTaPOD) to TaMDAR6 suppression were determined byqRT-PCR during incompatible and compatible interac-tions at different time points 0, 24 and 48 hpi. Transcriptlevels of PR genes were significantly upregulated dur-ing compatible and incompatible interactions at all timepoints (Fig. 6). By contrast, transcript levels of catalase(TaCAT) and class III peroxidase (TaPOD) were signifi-cantly reduced in the compatible interaction (Fig. 6). Inaddition, there was no significant change in the tran-script levels of TaCAT and TaPOD (reduced a little) inthe incompatible interaction (Fig. 6). These results indi-cated that the hypersensitive and resistance responses ofwheat to Pst were enhanced. Therefore, the increasesof PCD may be involved in the resistance responsesof wheat to Pst.

Histological observation of interactions betweenPst and TaMDAR6 knocked-down plants

To improve our understanding concerning the wheat–Pstinteraction, TaMDAR6 knocked-down plants were

infected by CYR31 (virulent Pst race) or CYR23 (avir-ulent Pst race), then microscopically examined. Inthe incompatible interaction, H2O2 production at theinteraction sites gradually increased in the TaMDAR6knocked-down plants at 24 and 48 hpi. In addition,hypersensitive cell death seemed to be enhanced inthe TaMDAR6 knocked-down plants at 48 hpi (Table 2,Fig. S8). Growth of hyphae was significantly reduced atboth time points (Table 2). Numbers of hyphal branchesand of haustorial mother cells were also significantlyreduced in the TaMDAR6 knocked-down plants at 24and 48 hpi (Table 2, Fig. S8).

In the compatible interaction, hyphal growth wasalso significantly shorter than that observed in theBSMV:𝛾-infected plants at 24 and 48 hpi (Table 3).Numbers of hyphal branches, haustorial mother andhaustoria cells were significantly reduced in the TaM-DAR6 knocked-down plants at 48 hpi (Table 3, Fig. S8).Moreover, H2O2 production at the site of interactionwas induced in the TaMDAR6 knocked-down plantsat 48 hpi (Table 3, Fig. S8). PCD was induced in theTaMDAR6 knocked-down plants at 120 hpi. However, at120 hpi, H2O2 production areas were too large, leadingto difficulties in taking other measurements (hyphalgrowth, numbers of hyphal branches and haustorialmother cells).

In the incompatible interaction, urediospores germi-nated well in BSMV:𝛾 (empty vector as control) andBSMV:TaMDARas plants at all time points. In controlplants, Pst germ tubes reached the stomata, with suc-cessful penetration at 24 and 48 hpi. In the TaMDAR6knocked-down plants, Pst germ tubes also reachedstomata, however, with successful penetration onlyat 24 hpi. At 48 hpi, there were two observations:

Physiol. Plant. 2015

Fig. 5. Analysis of TaMDAR6 using VIGS. (A) Phenotypes of wheatleaves after inoculation with FAS buffer (Mock) or BSMV:𝛾 orBSMV:TaMDARas or BSMV:TaPDSas. Mild virus symptoms and largephoto-bleached areas appeared on the fourth leaves of infected plantsat 15 dpi. (B) Phenotypes of the fourth leaves of inoculated plants withFAS buffer or BSMV:𝛾 or BSMV:TaMDAR followed by challenge with viru-lent (CYR31) or avirulent (CYR23) races of stripe rust. BSMV:𝛾: constructcarrying only the BSMV genome (empty vector); BSMV:PDS: silencingconstruct for PDS; BSMV:TaMDAR: silencing construct for TaMDAR6. Typ-ical leaves were photographed at 15 dpi.

(1) germ tubes did not end over stomata, and (2) germtubes grew over stomata, in both cases not penetratingthe stoma (Fig. S9). Moreover, in comparison with controlplants, the percentage of stomata that remained openin BSMV:TaMDARas-infected plants decreased at 24and 48 hpi by approximately 10 and 53%, respectively(Table 2). Meanwhile, a significant change in stomatalshape (length×width) of BSMV:TaMDARas-infectedplants could be observed at 48 hpi (Table 2, Fig. 7).However, no significant difference in stomatal shape Ta

ble

2.H

isto

logi

calo

bser

vatio

nsof

inco

mpa

tible

inte

ract

ions

betw

een

the

TaM

DA

R6kn

ocke

ddo

wn

plan

tsan

dPs

t.aW

heat

leav

esw

ere

inoc

ulat

edw

ithBS

MV:𝛾

(em

pty

vect

or)o

rBS

MV:

TaM

DA

R,fo

llow

edby

inoc

ulat

ion

with

aviru

lent

Pst

race

CY

R23.

Four

leav

esw

ere

rand

omly

sele

cted

fore

very

time

poin

tper

trea

tmen

ts,a

ndm

icro

scop

ical

lyex

amin

ed.b

Sign

ifica

ntly

diff

eren

t(P<

0.05

)acc

ordi

ngto

t-te

st.c A

vera

gelin

ear

dist

ance

from

the

base

ofth

esu

bsto

mat

alve

sicl

eto

the

hyph

altip

calc

ulat

edfr

om50

infe

ctio

nsi

tes.

dA

vera

genu

mbe

rof

hyph

albr

anch

esca

lcul

ated

from

50in

fect

ion

site

s.eA

vera

genu

mbe

rof

haus

toria

lmot

her

cells

calc

ulat

edfr

om50

infe

ctio

nsi

tes.

f Ave

rage

H2O

2pr

oduc

tion

area

per

infe

ctio

nsi

te(u

nit

inμm

2)c

alcu

late

dfr

om50

infe

ctio

nsi

tes.

gA

vera

gene

crot

icar

eape

rin

fect

ion

site

(uni

tin

μm2)c

alcu

late

dfr

om50

infe

ctio

nsi

tes.

hPe

rcen

tage

ofsu

cces

sful

lyes

tabl

ishe

din

fect

ions

whe

resu

bsto

mat

alve

sicl

esha

dfo

rmed

unde

rnea

thst

omat

a.i P

erce

ntag

eof

open

stom

ata

calc

ulat

edfr

om10

0st

omat

a.j V

alue

sgi

ven

for

stom

asi

ze(le

ngth

×w

idth

,uni

tin

μm)r

epre

sent

aver

ages

of10

0st

omat

a.

Hyp

hall

engt

h

(μm

)cH

ypha

lbra

nche

sd

Hau

stor

ial

mot

her

celle

H2O

2ar

ea(μ

m2)f

Succ

essf

ul

infe

ctio

nsi

te%

h

Ope

n

stom

ata

%i

Stom

asi

ze(μ

m)j

Trea

tmen

tsa

24hp

i48

hpi

24hp

i48

hpi

24hp

i48

hpi

24hp

i48

hpi

Nec

rotic

area

(μm

2)g

48hp

i24

hpi

48hp

i24

hpi

48hp

i24

hpi

48hp

i

BSM

V:𝛾

/CY

R23

25.9

2.00

28.7

2.39

2.07

±0.

532.

00±

0.50

2.03

±0.

502.

05±

0.60

734±

196

822±

191

531±

192

7981

7373

69.4

4.27

×4.

14±

0.92

55.6

4.62

×6.

59±

1.23

BSM

V:Ta

MD

AR

/CY

R23

21.8

2.50

b22

.91±

2.00

b1.

60±

0.50

b1.

54±

0.50

b1.

71±

0.52

b0.

85±

0.20

b10

54±

201b

2398

±19

2b12

06±

190b

8025

6320

70.4

4.25

×3.

90±

0.67

123.

93±

9.93

×5.

39±

1.52

b

Physiol. Plant. 2015

Fig. 6. Expression of resistance-related genes in the TaMDAR6knocked-down wheat in response to Pst infection during (A) anincompatible interaction and (B) a compatible interaction. Tran-script levels were assayed using qRT-PCR. TaPR2, 𝛽-1,3-glucanase(DQ090946); TaPR5, Thaumatin-like protein (FG618781); TaCAT , cata-lase (X94352) and TaPOD, class III peroxidase (TC303653). BSMV:𝛾- andBSMV:TaMDARas-infected leaves were sampled 0, 24 and 48 hpi withPst. BSMV:𝛾 construct carrying only the BSMV genome (empty vector).Relative transcript levels were quantified using the comparative thresh-old (2–ΔΔCT) method. Asterisks indicate a significant difference (P <0.05)from the non-silenced plants using t-test.

was observed between BSMV:TaMDARas-infected andBSMV:𝛾-infected leaves at 24 hpi (Table 2, Fig. 7).Besides, the percentage of the infection sites where sub-stomatal vesicles had formed underneath the stomata(successful penetration) in the TaMDAR6 silenced plantswas lower than that in control plants at 48 hpi by approx-imately 56% (Table 2). In the compatible interaction,there was no significant change in stomatal behavior orin percentage of infection sites (Table 3). Ta

ble

3.H

isto

logi

calo

bser

vatio

nsof

com

patib

lein

tera

ctio

nsbe

twee

nth

eTa

MD

AR6

knoc

ked

dow

npl

ants

and

Pst.

aW

heat

leav

esw

ere

inoc

ulat

edw

ithBS

MV:𝛾

(em

pty

vect

or)

orBS

MV:

TaM

DA

R,fo

llow

edby

inoc

ulat

ion

with

viru

lent

Pst

race

CY

R31.

Four

leav

esw

ere

rand

omly

sele

cted

fore

very

time

poin

tper

trea

tmen

ts,a

ndm

icro

scop

ical

lyex

amin

ed.b

Sign

ifica

ntly

diff

eren

t(P<

0.05

)acc

ordi

ngto

t-te

st.c A

vera

gelin

ear

dist

ance

from

the

base

ofth

esu

bsto

mat

alve

sicl

eto

the

hyph

altip

calc

ulat

edfr

om50

infe

ctio

nsi

tes.

dA

vera

genu

mbe

rof

hyph

albr

anch

esca

lcul

ated

from

50in

fect

ion

site

s.eA

vera

genu

mbe

rof

haus

toria

lmot

her

cells

calc

ulat

edfr

om50

infe

ctio

nsi

tes.

f Ave

rage

num

ber

ofha

usto

riaca

lcul

ated

from

50in

fect

ion

site

s.gA

vera

geH

2O

2pr

oduc

tion

area

per

infe

ctio

nsi

te(u

nit

inμm

2)c

alcu

late

dfr

om50

infe

ctio

nsi

tes.

hA

vera

gene

crot

icar

eape

rin

fect

ion

site

(uni

tin

μm2)c

alcu

late

dfr

om50

infe

ctio

nsi

tes.

i Per

cent

age

ofsu

cces

sful

lyes

tabl

ishe

din

fect

ions

whe

resu

bsto

mat

alve

sicl

esha

dfo

rmed

unde

rnea

thst

omat

a.j P

erce

ntag

eof

open

stom

ata

calc

ulat

edfr

om10

0st

omat

a.kVa

lues

give

nfo

rst

oma

size

(leng

th×

wid

th,u

nit

inμm

)rep

rese

ntav

erag

esof

100

stom

ata.

Trea

tmen

tsa

Hyp

hall

engt

h(μ

m)c

Hyp

halb

ranc

hesd

Hau

stor

ial

mot

her

celle

Succ

essf

ul

infe

ctio

nsi

te%

i

Ope

n

stom

ata

%j

Stom

asi

ze(μ

m)k

24hp

i48

hpi

24hp

i48

hpi

24hp

i48

hpi

Hau

stor

iaf

48hp

i

H2O

2ar

ea(μ

m2)g

48hp

i

Nec

rotic

area

(μm

2)h

120

hpi

24hp

i48

hpi

24hp

i48

hpi

24hp

i48

hpi

BSM

V:𝛾

/CY

R31

27.7

2.50

32.8

2.10

1.82

±0.

502.

97±

0.50

1.87

±0.

502.

70±

0.55

1.83

±0.

20–

–85

9792

9572

.17±

4.25

×5.

59±

1.00

72.5

4.52

×5.

43±

1.00

BSM

V:Ta

MD

AR

/CY

R31

21.7

2.50

b25

.94±

2.00

b1.

50±

0.50

1.77

±0.

53b

1.50

±0.

491.

72±

0.55

b0.

45±

0.20

b18

5150

192

8295

9094

72.3

4.27

×5.

17±

1.03

72.5

4.67

×5.

02±

1.10

Physiol. Plant. 2015

Fig. 7. Stomatal shape on wheat leaves inoculated with BSMV:𝛾 (empty vector) or BSMV:TaMDARas during an incompatible interaction. Suppressionof TaMDAR6 induced changes in stomatal morphology. (A) Stoma on control plants at 24 hpi. (B) Stoma on the TaMDAR6 knocked-down plants at 24hpi. (C) Stoma on control plants at 48 hpi. (D) Induced changes in stomatal shape on the TaMDAR6 knocked-down plants at 48 hpi. Scale bars=20 μm.

Suppression of TaMDAR6 induces H2O2accumulation without Pst attack

To demonstrate that the suppression of TaMDAR6induces ROS accumulation, wheat leaves were infectedwith BSMV:TaMDARas or with BSMV:𝛾 without Pstinoculation. In comparison to BSMV:𝛾-infected leaves,H2O2 production was induced in the fourth leaves ofBSMV:TaMDARas-infected plants (Table S3, Fig. 8). Inaddition, there also was no significant change in stomatalshape in the fourth leaves of BSMV:TaMDARas-infectedplants (Table S3).

Discussion

TaMDAR6 functions as a negative regulator of PCD

PCD is an important element for plant immunity againstbiotic and abiotic stress as well as for plant developmentand proliferation (Gadjev et al. 2008), including vac-uolar cell death, necrosis and hypersensitive cell death(Van Doorn et al. 2011). In fact, plant hypersensitive celldeath occurs during plant infection by hemibiotrophicand biotrophic pathogens, and displays the same micro-scopic features of both necrosis and vacuolar cell death

Fig. 8. Microscopic observation of H2O2 accumulation in the TaMDAR6knocked-down wheat without pathogen infection. The fourth leaves ofthe wheat seedlings were infected with (A) BSMV:𝛾 (empty vector) or(B) BSMV:TaMDARas, and stained with 3,3-diaminobenzidine. Areas ofH2O2 production were measured by DP-BSW software and calculated fromat least 10 fields for each segment. Scale bars=20 μm.

(Van Doorn et al. 2011). In addition, ROS have been con-sidered as key factors of the induction and modulationof the PCD during plant–pathogen interaction (Desikan

Physiol. Plant. 2015

et al. 1998, Gadjev et al. 2008). The lifetime of ROSwithin the cellular environment is determined by theantioxidative system, which provides crucial protectionagainst oxidative damage (Apel and Hirt 2004). As men-tioned, MDAR proteins also play important roles in ROSaccumulation. It is worth mentioning that plant chloro-plasts are considered as the most important sources ofROS (Ishikawa and Shigeoka 2008, Maruta et al. 2012).In addition, the role of TaMDAR6 encoding chloroplas-tic TaMDAR protein in the wheat–Pst interaction andin PCD has not been reported. Therefore, it is temptingto speculate that the chloroplastic TaMDAR protein mayparticipate in PCD regulation. To study this hypothesis,we isolated the TaMDAR6 gene encoding chloroplasticTaMDAR from wheat, then studied its role in PCD trig-gered by a mammalian Bax gene and by Pst.

The mammalian Bax gene is an animal pro-apoptoticprotein, absent in plants, which induces PCD (HR-likecell death) primarily through the accumulation ofROS similar to plant hypersensitive cell death (Caiand Jones 1998, Fujita et al. 1998, Aravind et al.1999, Kawai-Yamada et al. 2001). The co-infiltrationof TaMDAR6 together with Bax resulted in a slowercell death than that induced by Bax alone. Thisfinding is consistent with most reports that haveindicated plant antioxidants as a negative regulatorof Bax-induced PCD (Kampranis et al. 2000, Moonet al. 2002 and Chen et al. 2004). In these studies,the over-expression of plant antioxidants (ascorbateperoxidase; glutathione-S-transferase/peroxidase, andphospholipid hydroperoxide glutathione peroxidase)reduces ROS accumulation and suppresses Bax expres-sion in N. benthamiana and in yeast. The authorsconclude that plant antioxidants suppress Bax-inducedPCD through decreasing ROS generation. Takentogether, we suggest that the differences observed in thenecrotic development are caused by TaMDAR6 expres-sion, which in turn functions as a negative regulator ofPCD triggered by Bax.

In this study, a VIGS approach was used to deter-mine the role of TaMDAR6 (located on chromosome7DL) in pathogen-induced cell death. Macroscopic andmicroscopic analyses showed that the average necroticarea was significantly increased in the TaMDAR6knocked-down plants upon Pst infection. El-Zahabyet al. (1995) stated that plant antioxidants suppress theformation of necrotic symptoms during barley–Bghinteractions. Therefore, we conclude that silencing ofTaMDAR6 is able to increase PCD during the wheat–Pstinteractions. Meanwhile, H2O2 accumulation was sig-nificantly increased in the TaMDAR6 knocked-downplants over time, suggesting that increase in necrotic

area may be linked to an increase in H2O2 accumu-lation, indicating that TaMDAR6 silencing increasesPCD through increased H2O2 accumulation. Thus,TaMDAR6 has a negative effect on PCD through reduc-ing H2O2 accumulation. Therefore, we concludethat TaMDAR6 is involved in PCD during wheat–Pstinteractions.

TaMDAR6 has a negative role in Pst recognition

Our previous histological studies stated that an oxida-tive burst was triggered following the recognition of elic-itor(s), released from haustoria of an avirulent race bya receptor of the host cell (Wang et al. 2007, 2010).From this study, we detected H2O2 accumulation atthe infection sites after TaMDAR6 was knocked-downin the compatible interaction, indicating that the wheatcells became able to recognize the virulent race of Pst.Thus, the interaction between pathogen elicitors andhost receptors probably activates a signal transductioncascade that involves H2O2 accumulation. ThereforeTaMDAR6 has a negative role in pathogen recognitionduring the wheat–Pst interaction. Furthermore, hyphalgrowth, the number of hyphal branches, and the num-ber of haustoria and haustorial mother cells were signif-icantly reduced in the TaMDAR6 knocked-down plantsfor both interactions. These results open the possibilitythat increased H2O2 production reduces initial growthof Pst, consistent with the findings of Feng et al. (2014a,2014b) in wheat against Pst during suppression of TaMD-HAR4 and TaMDHAR genes. Ellingboe (1972) reportedthat development of hyphae is a good indicator for estab-lishment of a compatible interaction. Therefore, we con-clude that the knocking down of TaMDAR6 enhancesplant resistance to Pst.

TaMDAR6 negatively regulates ROS accumulation

Transcript levels of TaMDAR6 were upregulated uponPst infection in compatible and incompatible interac-tions. Therefore, we suggest that TaMDAR6 is involvedin wheat–Pst interactions. Moreover, transcript levelsof TaMDAR6 were higher in a compatible interactionthan that in incompatible interaction. Burhenne andGregersen (2000) and Vanacker et al. (1998), stated thatthe inoculation of barley leaves with Blumeria graminis(Bgh) causes significant increases in the expression levelsand activities of antioxidant enzymes such as ascorbateperoxidase in compatible interactions more than incom-patible interactions, suggesting that antioxidant enzymesmay contribute to plant sensitivity against Bgh infection.Feng et al. (2014a) described that transcript levels ofTaMDHAR4 gene were downregulated at 12–18 hpi

Physiol. Plant. 2015

during incompatible interaction then upregulated at48 hpi in both interactions. The authors suggested thatsuppression of TaMDHAR4 might be important to theincompatible interaction. In contrast, transcript levelsof TaMDAR6 remained at the control level from 12to 18 hpi then upregulated at 24 and 48 hpi in bothinteractions. Therefore, it is tempting to speculate thatthere may be a required level of TaMDAR6 expression inwheat plants to show incompatible interaction againstPst, and that an increase in that level may contribute toplant sensitivity against Pst.

In this study, we studied for the first time the relation-ship between TaMDAR6 gene and substomatal vesiclesformation and stomata regulation during the Pst–wheatinteraction. The knocking down of TaMDAR6 duringwheat–Pst interactions reduced disease symptoms andinduced HR, suggesting that the suppression of TaM-DAR6 increases wheat resistance against Pst. Besides,histological observations of the Pst inoculated plantsindicate that urediospores germinated successfully inboth wheat–Pst interactions. However, the percentageof the infection sites where substomatal vesicles formed(successful penetration) was influenced by the knockingdown of TaMDAR6 in the incompatible interaction.Meanwhile, H2O2 accumulation and the expressionlevel of genes encoding pathogenesis-related proteinssignificantly increased over time in the TaMDAR6knocked-down plants. Dangl and Jones (2001) statedthat H2O2 accumulation could activate many plantdefenses to eliminate the pathogen. In addition, Broersand López-Atilano (1996) during a study of the effectsof wheat resistance on the development of Pst, sug-gested that some of the formed substomatal vesiclesdisintegrated later by PR proteins that were activatedafter pathogen colonization. Therefore, we concludethat the decrease in substomatal vesicle formationmight be linked to an increase in H2O2 accumulation,which in turn may trigger other resistance mechanismsthat have a negative effect on the formed substomatalvesicles.

Furthermore, some of the germ tubes of Pst in theincompatible interaction did not reach the stomata orgrew over them without penetration, suggesting thatthe germ tubes became unable to recognize stomataon the TaMDAR6 knocked-down plants. In addition,non-recognition of stomata by the germ tubes of Pstwas associated with changes in the shape of the stom-ata and with decreases in the percentage of stomatathat remained open in the incompatible interaction.This suggests that the suppression of TaMDAR6 inducedchanges in stomatal morphology and enhanced stom-atal closure on the TaMDAR6 knocked-down plants inthe incompatible interaction. In addition, the inability of

Pst germ tubes to recognize stomata may be triggeredby changes in morphological features of the stomata.Moreover, changes of stomatal shape in the TaMDAR6knocked-down plants might be triggered by an increasein H2O2 accumulation in the incompatible interaction.Allen et al. (2000) and McAinsh et al. (1996) statedthat H2O2 has a marked effect on stomatal behavior,induces changes in guard cells and promotes stomatalclosure. In addition, TaMDAR6 localizes to the chloro-plast of N. benthamiana. Maruta et al. (2012) statedthat plant chloroplasts are a major source of H2O2.Thus, TaMDAR6 plays important role in H2O2 regula-tion. Therefore, we conclude that TaMDAR6 participatesindirectly in stomatal regulation through its effects onH2O2 accumulation during the wheat–Pst incompati-ble interaction. According to these results, the knock-ing down of TaMDAR6 enhances plant responses toavoid further penetration, reducing Pst infection effi-ciency through reducing the probability of Pst germ tubesto locate or recognize the stomata during incompatibleinteraction.

In contrast, there was no significant change in stomatalbehavior in silenced plants for either compatible inter-action or non-inoculated plants even though there wasan increase of the H2O2 accumulation in both cases.Bhattacharjee (2012) and Melillo et al. (2006), statedthat ROS may induce different types of responses dur-ing plant infection, depending on their level. Therefore,these results open the possibility that H2O2 levels in bothcases were not sufficient to induce significant changesin stomatal behavior or to reduce the percentage of theinfection sites in the compatible interaction. Therefore,we conclude that change in stomatal behavior dependson the level of ROS.

Collectively, the results of this study provide strong evi-dence that TaMDAR6 functions as a negative regulatorof PCD. In addition, the knocking down of TaMDAR6induces ROS accumulation, which triggers other resis-tance mechanisms, depending on their level, such aschanges in stomatal behavior. Thus, the knocking downof the TaMDAR6 gene enhances the resistance of wheatto Pst.

Author contributions

M. A. A. A., X. W. and Z. K. performed conception,design and interpretation of data; M. A. A. A., M. N. A.A., Q. X.; G. Z. performed the experiments; M. A. A. A.and X. W. analyzed the data; M. A. A. A., X. W. draftedthe article or revised; Z. K. contributed the final approvalof the version to be published.

Physiol. Plant. 2015

Acknowledgements – This study was supported bythe National Basic Research Program of China (No.2013CB127700), the Key Grant Project of the ChineseMinistry of Education (313048), the National NaturalScience Foundation of China (No. 31271990), the Youngstar of science and technology, Shaan’xi (2012KJXX-15)and by the 111 Project from the Ministry of Education ofChina (B07049). We thank Nagi Abou-zeid of the PlantPathology Research Institute, ARC, Egypt for his adviceand continuous guidance; S. R. Scofield of the Departmentof Agronomy, Purdue University, West Lafayette, USA forproviding BSMV vectors; and D. Dou for providing transientexpression vectors.

References

Al-Attala MN, Wang X, Abou-Attia MA, Duan X, Kang Z(2014) A novel TaMYB4 transcription factor involved inthe defence response against Puccinia striiformis f. sp.tritici and abiotic stresses. Plant Mol Biol 84: 589–603

Ali MB, Hahn EJ, Paek KY (2005) Effects of temperature onoxidative stress defense systems, lipid peroxidation andlipoxygenase activity in Phalaenopsis. Plant PhysiolBiochem 43: 213–223

Allen GJ, Chu SP, Schumacher K, Shimazaki CT, VafeadosD, Kemper A, Hawke SD, Tallman G, Tsien RY, HarperJF, Chory J, Schroeder JI (2000) Alternation ofstimulus-specific guard cell calcium oscillations andstomatal closing in Arabidopsis det3 mutant. Science289: 2338–2342

Allison C, Isenbeck K (1930) Biologische Spezialisierungvon Puccinia glumarum tritici Erikss. und Henn.Phytopathol Z 2: 87–98

Apel K, Hirt H (2004) Reactive oxygen species:metabolism, oxidative stress, and signal transduction.Annu Rev Plant Biol 55: 373–399

Aravind L, Dixit VM, Koonin EV (1999) The domains ofdeath: evolution of the apoptosis machinery. TrendsBiochem Sci 24: 47–53

Asad MA, Xia X, Wang C, He Z (2012) Molecular mappingof stripe rust resistance gene YrSN104 in Chinese wheatline Shaannong 104. Hereditas 149: 146–152

Ausubel FM, Brent R, Kingston RE, Moore DD, SeidmanJG, Smith JA, Struhl K (1997) Current Protocols inMolecular Biology. John Wiley & Sons, New York

Bhattacharjee S (2012) The language of reactive oxygenspecies signaling in plants. J Bot 2012: 1–22

Broers LM, López-Atilano RM (1996) Effect of quantitativeresistance in wheat on the development of Pucciniastriiformis during early stages of infection. Plant Disease80: 1265–1268

Burhenne K, Gregersen PL (2000) Up-regulation of theascorbate-dependent antioxidative system in barleyleaves during powdery mildew infection. Mol PlantPathol 1: 303–314

Cai J, Jones DP (1998) Superoxide in apoptosis.Mitochondrial generation triggered by cytochrome closs. J Biol Chem 273: 11401–11404

Chandra CR, Kishor SN, Mithilesh K, Rajeev K (2014)Wheat genotypes (Triticum aestivum L.) vary widely intheir responses of fertility traits to high temperature atanthesis. Int Res J Biol Sci 3: 54–60

Chen S, Vaghchhipawala Z, Li W, Asard H, Dickman MB(2004) Tomato phospholipid hydroperoxide glutathioneperoxidase inhibits cell death induced by Bax andoxidative stresses in yeast and plants. Plant Physiol 135:1630–1641

Dangl JL, Jones JD (2001) Plant pathogens and integrateddefence responses to infection. Nature 411: 826–833

Desikan R, Reynolds A, Hancock JT, Neill SJ (1998)Harpin and hydrogen peroxide both initiateprogrammed cell death but have differential effects ongene expression in Arabidopsis suspension cultures.Biochem J 330: 115–120

Draper J (1997) Salicylate, superoxide synthesis and cellsuicide in plant defence. Trends Plant Sci 2: 162–165

Ellingboe AH (1972) Genetics and physiology of primaryinfection by Erysiphe graminis f.sp. hordei.Phytopathology 62: 401–406

Eltayeb AE, Kawano N, Badawi GH, Kaminaka H,Sanekata T, Shibahara T, Inanaga S, Tanaka K (2007)Overexpression of monodehydroascorbate reductase intransgenic Nicotiana benthamiana confers enhancedtolerance to ozone, salt and polyethylene glycol stresses.Planta 225: 1255–1264

Eltelib HA, Badejo AA, Fujikawa Y, Esaka M (2011) Geneexpression of monodehydroascorbate reductase anddehydroascorbate reductase during fruit ripening and inresponse to environmental stresses in acerola (Malpighiaglabra). J Plant Physiol 168: 619–627

El-Zahaby HM, Gullner G, Király Z (1995) Effects ofpowdery mildew infection of barley on theascorbate-glutathione cycle and other antioxidants indifferent host–pathogen interactions. Phytopathology85: 1225–1230

Feng H, Liu W, Zhang Q, Wang X, Wang X, Duan X, Li F,Huang L, Kang Z (2014a) TaMDHAR4, amonodehydroascorbate reductase gene participates inthe interactions between wheat and Puccinia striiformisf. sp. tritici. Plant Physiol Biochem 76: 7–16

Feng H, Wang X, Zhang Q, Fu Y, Feng C, Wang B, HuangL, Kang Z (2014b) Monodehydroascorbate reductasegene, regulated by the wheat PN-2013 miRNA,contributes to adult wheat plant resistance to stripe rustthrough ROS metabolism. Biochim Biophys Acta 1839:1–12

Fujita N, Nagahashi A, Nagashima K, Rokudai S, Tsuruo T(1998) Acceleration of apoptotic cell death after thecleavage of Bcl-XL protein by caspase-3-like proteases.Oncogene 17: 1295–1304

Physiol. Plant. 2015

Gadjev I, Stone JM, Gechev TS (2008) Programmed celldeath in plants: new insights into redox regulation andthe role of hydrogen peroxide. In: Kwang WJ (ed)International Review of Cell and Molecular Biology.Academic Press, Waltham, pp 87–144

Holzberg S, Brosio P, Gross C, Pogue GP (2002) Barleystripe mosaic virus-induced gene silencing in a monocotplant. Plant J 30: 315–327

Ishikawa T, Shigeoka S (2008) Recent advances inascorbate biosynthesis and the physiologicalsignificance of ascorbate peroxidase inphotosynthesizing organisms. Biosci BiotechnolBiochem 72: 1143–1154

Kampranis SC, Damianova R, Atallah M, Toby G, Kondi G,Tsichlis PN, Makris AM (2000) A novel plant glutathioneS-transferase/peroxidase suppresses Bax lethality inyeast. J Biol Chem 275: 29207–29216

Kang ZS, Wang Y, Huang LL, Wei GR, Zhao J (2003)Histology and ultrastructure of incompatiblecombination between Puccinia striiformis and wheatcultivars with resistance of low reaction type. Sci AgricSin 36: 1026–1031

Kawai-Yamada M, Jin L, Yoshinaga K, Hirata A, UchimiyaH (2001) Mammalian Bax-induced plant cell death canbe down regulated by overexpression of Arabidopsis BaxInhibitor-1 (AtBI-1). Proc Natl Acad Sci USA 98:12295–12300

Langridge P (2012) Genomics: decoding our daily bread.Nature 491: 678–680

Li Z, Zeng S (2000) Wheat Rusts in China. ChinaAgriculture Press, Beijing

Livak KJ, Schmittgen TD (2001) Analysis of relative geneexpression data using real-time quantitative PCR and the2–ΔΔCT method. Methods 25: 402–408

Lunde C, Baumann U, Shirley N, Drew D, Fincher G(2006) Gene structure and expression pattern analysis ofthree monodehydroascorbate reductase (Mdhar) genesin Physcomitrella patens: implications for the evolutionof the MDHAR family in plants. Plant Mol Biol 60:259–275

Ma J, Huang X, Wang X, Chen X, Qu Z, Huang L, Kang Z(2009) Identification of expressed genes duringcompatible interaction between stripe rust (Pucciniastriiformis) and wheat using a cDNA library. BMCGenom 10: 586

Maruta T, Noshi M, Tanouchi A, Tamoi M, Yabuta Y,Yoshimura K, Ishikawa T, Shigeoka S (2012)H2O2-triggered retrograde signaling from chloroplasts tonucleus plays specific role in response to stress. J BiolChem 287: 11717–11729

McAinsh MR, Clayton H, Mansfield TA, Hetherington AM(1996) Changes in stomatal behavior and guard cellcytosolic free calcium in response to oxidative stress.Plant Physiol 111: 1031–1042

McNeal FH, Konzak CS, Smith EP, Tate WS, Russell TS(1971) A Uniform System for Recording and ProcessingCereal Research Data. USDA, Agricultural ResearchService, Washington, DC, pp 34–121

Melillo MT, Leonetti P, Bongiovanni M,Castagnone-Sereno P, Bleve-Zacheo T (2006)Modulation of reactive oxygen species activities andH2O2 accumulation during compatible andincompatible tomato–root-knot nematode interactions.New Phytol 170: 501–512

Moon H, Baek D, Lee B, Prasad DT, Lee SY, Cho MJ, LimCO, Choi MS, Bahk J, Kim MO, Hong JC, Yun DJ (2002)Soybean ascorbate peroxidase suppresses Bax-inducedapoptosis in yeast by inhibiting oxygen radicalgeneration. Biochem Biophys Res Commun 290:457–462

Parlevliet JE (1986) Pleiotropic association of infectionfrequency and latent period of two barley cultivarspartially resistant to barley leaf rust. Euphytica 35:267–272

Schaffrath U, Freydl E, Dudler R (1997) Evidence fordifferent signaling pathways activated by inducers ofacquired resistance in wheat. Mol Plant–MicrobeInteract 10: 779–783

Sharma YK, Davis KR (1997) The effects of ozone onantioxidant responses in plants. Free Radic Biol Med 23:480–488

Sharma P, Dubey RS (2005) Drought induces oxidativestress and enhances the activities of antioxidantenzymes in growing rice seedlings. Plant Growth Regul46: 209–221

Sharma P, Jha AB, Dubey RS, Pessarakli M (2012) Reactiveoxygen species, oxidative damage, and antioxidativedefense mechanism in plants under stressful conditions.J Bot 2012: 1–26

Stubbs RW (1985) Stripe rust. In: Roelfs AP, Bushnell WR(eds) The Cereal Rusts, 2nd Edn. Academic Press, NewYork, pp 61–101

Stubbs RW (1988) Pathogenicity analysis of yellow (stripe)rust of wheat and its significance in a global context. In:Simmonds NW, Rajaram S (eds) Breeding Strategies forResistance to the Rusts of Wheat. CIMMYT, Mexico, pp23–38

Van der Hoorn RA, Laurent F, Roth R, De Wit PJ (2000)Agroinfiltration is a versatile tool that facilitatescomparative analyses of Avr9/Cf-9-induced andAvr4/Cf-4-induced necrosis. Mol Plant–Microbe Interact13: 439–446

Van Doorn WG, Beers EP, Dangl JL, Franklin-Tong VE,Gallois P, Hara-Nishimura I, Jones AM, Kawai-YamadaM, Lam E, Mundy J, Mur LA, Petersen M, Smertenko A,Taliansky M, Van Breusegem F, Wolpert T, Woltering E,Zhivovtosky B, Bozhkov PV. (2011) Morphologicalclassification of plant cell deaths. Cell Death Differ 18:1241–1246

Physiol. Plant. 2015

Van Loon LC, Van Strien EA (2002) The families ofpathogenesis-related proteins, their activities, andcomparative analysis of PR-1 type proteins. Physiol MolPlant Pathol 55: 85–97

Vanacker H, Carver TL, Foyer CH (1998) Pathogen-inducedchanges in the antioxidant status of the apoplast inbarley leaves. Plant Physiol 117: 1103–1114

Wang CF, Huang LL, Buchenauer H, Han QM, Zhang HC,Kang ZS (2007) Histochemical studies on theaccumulation of reactive oxygen species (O2

− andH2O2) in the incompatible and compatible interactionof wheat–Puccinia striiformis f. sp. tritici. Physiol MolPlant Pathol 71: 230–239

Wang CF, Huang L, Zhang HC, Han QM, Buchenauer H,Kang ZS (2010) Cytochemical localization of reactiveoxygen species (O2

− and H2O2) and peroxidase in theincompatible and compatible interaction ofwheat–Puccinia striformis f.sp. tritici. Physiol Mol PlantPathol 74: 221–229

Wellings CR, McIntosh RA (1990) Puccinia striiformis f.sp.tritici in Australasia: pathogenic changes during the first10 years. Plant Pathol 39: 316–325

Wise AA, Liu Z, Binns AN (2006) Three methods for theintroduction of foreign DNA into Agrobacterium.Methods Mol Biol 343: 43–53

Yoon HS, Lee H, Lee IA, Kim KY, Jo J (2004) Molecularcloning of the monodehydroascorbate reductase genefrom Brassica campestris and analysis of its mRNA levelin response to oxidative stress. Biochim Biophys Acta1658: 181–186

Zheng W, Huang L, Huang J, Wang X, Chen X, Zhao J,Guo J, Zhuang H, Qiu C, Liu J, Liu H, Huang X, Pei G,Zhan G, Tang C, Cheng Y, Liu M, Zhang J, Zhao Z,Zhang S, Han Q, Han D, Zhang H, Zhao J, Gao X,Wang J, Ni P, Dong W, Yang L, Yang H, Xu J-R, ZhangG, Kang Z. (2013) High genome heterozygosity andendemic genetic recombination in the wheat stripe rustfungus. Nat Commun 4: 1–10

Supporting Information

Additional Supporting Information may be found in theonline version of this article:

Table S1. Primers used in this study.

Table S2. Homology matrix of TaMDAR6 and othermonodehydroascorbate reductase proteins from differentplants.

Table S3. Histological observations of the TaMDAR6knocked-down plants without pathogen infection.

Fig. S1. Prediction of conserved domains using the NCBIdatabase.

Fig. S2. cDNA multiple sequence alignment of all TaM-DAR6 copies.

Fig. S3. Maps of the binary PVX-based expression vec-tors.

Fig. S4. Transient overexpression GFP in Nicotiana ben-thamiana.

Fig. S5. BSMV-mediated VIGS system.

Fig. S6. Alignment of the nucleotide sequences of TaM-DAR6 with other MDAR genes from wheat.

Fig. S7. Relative transcript levels of all three TaMDAR6copies in silenced and non-silenced plants.

Fig. S8. Wheat–Pst interactions in response to suppres-sion of TaMDAR6 gene.

Fig. S9. Micrographs of germlings of Pst on the TaM-DAR6 knocked-down wheat leaves in the incompatibleinteraction.

Edited by H. Saitoh

Physiol. Plant. 2015


Recommended