当前位置:文档之家› The role and regulation of programmed cell death in plant–pathogen interactions

The role and regulation of programmed cell death in plant–pathogen interactions

Cellular Microbiology (2004) 6(3), 201–211doi:10.1111/j.1462-5822.2004.00361.x

? 2004 Blackwell Publishing Ltd

Received 29 September, 2003; revised 10 November, 2003; accepted 12 November, 2003. *For correspondence. E-mail jgreenbe@https://www.doczj.com/doc/457717247.html,; T el. (+1) 773 834 1908; Fax (+1) 773 702 9270.Microreview

The role and regulation of programmed cell death in plant–pathogen interactions

Jean T. Greenberg* and Nan Yao

The University of Chicago, 1103 East 57th Street, EBC410, Chicago, IL 60637, USA.Summary

It is commonly known that animal pathogens often target and suppress programmed cell death (pcd)pathway components to manipulate their hosts. In contrast, plant pathogens often trigger pcd. In cases in which plant pcd accompanies disease resistance,an event called the hypersensitive response, the plant surveillance system has learned to detect pathogen-secreted molecules in order to mount a defence response. In plants without genetic disease resis-tance, these secreted molecules serve as virulence factors that act through largely unknown mecha-nisms. Recent studies suggest that plant bacterial pathogens also secrete antiapoptotic proteins to pro-mote their virulence. In contrast, a number of fungal pathogens secrete pcd-promoting molecules that are critical virulence factors. Here, we review recent progress in determining the role and regulation of plant pcd responses that accompany both resistance and susceptible interactions. We also review progress in discerning the mechanisms by which plant pcd occurs during these different interactions.Introduction

Host cell death occurs during many, but not all, interac-tions between plants and the pathogens that infect them.This cell death can be associated with disease resistance or susceptibility, depending on the lifestyle of the patho-gen. What is the role of host cell death during pathogen-esis? Do all cells in an infection zone die by the same mechanism? Which other cellular processes are in?u-enced when some host cells die during infection? As will be addressed in this review, the answers to these ques-tions depend on the host–pathogen interaction being examined. However, what is clear is that cell death regu-lation is often intimately linked to a number of signalling pathways that in?uence other defence processes. Fur-thermore, cell death can serve to amplify other defence responses as well as to promote the aggressiveness and/or dissemination of some pathogens. As such, under-standing cell death mechanisms of control and execution offers potential targets for modulating both host disease resistance and susceptibility. If cells die by different mech-anisms depending on the host–pathogen combination,this information could be used selectively to target differ-ent pathways or, at the very least, could lead to the devel-opment of diagnostic tools that would help in the dissection of cell death and defence regulatory networks.Finally, if cells even within one host–pathogen interaction die by different mechanisms, it will be important to recog-nize that the function and regulation of these cell death events could be distinct.

Recent work suggests that pathogenesis-associated cell death in plants has common regulatory and mecha-nistic features with apoptosis in animals (for example, see Liang et al ., 2003). This observation offers the possibility of understanding the basal eukaryotic cell death machin-ery and studying how its activation is linked to other cell biological and signalling events during pathogenesis. The purpose of this review is to analyse critically what is known about the role and mechanism of cell death in different host–pathogen interactions. The reader may also ?nd sev-eral older reviews useful for insight into this topic (Green-berg, 1997; Morel and Dangl, 1997; Heath, 2000).Cell death during disease resistance

Perhaps the most well-known cell death response in plants is the hypersensitive response (HR) associated with a phenomenon termed the resistance response (RR).The RR involves the co-ordinate activation of many defences that limit pathogen growth (G reenberg, 1997).For the purpose of this review, the HR is considered to be only the cell death component of the RR. An RR is trig-gered when the host has a dominant R gene that corre-sponds to a dominant avr gene in the pathogen. This

202J. T . Greenberg and N. Y ao

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

gene-for-gene interaction results from either direct or indi-rect interaction between the R gene and avr gene prod-ucts depending on the R–avr gene pair. T wo cases of direct interactions have been described (T ang et al ., 1996;Jia et al ., 2000). Additionally, one interaction was shown to occur via formation of a complex probably involving an R protein, an Avr protein and an additional host protein(s)(Leister and Katagiri, 2000). Finally, an R–Avr interaction was recently found to result from the enzymatic activity of the Avr protein (Shao et al ., 2003). Although RRs are often associated with the HR, in some cases, they can occur without or with very little cell death (for example,see Bendahmane et al ., 1999).

Despite the observation that many RRs are accompa-nied by an HR, the signal transduction requirements for the RR and the HR can be different depending on the host–pathogen combination. Sometimes, the same gene can be required for the HR in response to one avirulent pathogen and can negatively regulate or be largely dis-pensable for the HR in response to different and/or related pathogens (see AtRBOHD/AtRBOHF , NDR1 and RAR1examples in T able 1). Additionally, the major defence sig-nal salicylic acid has both cell death-enhancing and cell death-suppressing roles depending on the avirulence fac-tor or pathogen (see discussion in Vanacker et al ., 2001).These different signalling requirements do not address whether the HRs eventually feed into a common cell death mechanism, although it is widely assumed that they do.Furthermore, the events that occur during different RRs can be quantitatively different. For example, the HR can occur with very different timing in response to highly related avirulent pathogens (Ritter and Dangl, 1996).Additionally, other defence-related events, such as the induction of cell wall papillae, which are thought to impede pathogen egress into plant cells, can occur with different timing and to different degrees depending on the R gene conditioning resistance (Freialdenhoven et al ., 1994).Cell death mechanisms during the HR

The view that the HR is an active process of the host and may be a form of programmed cell death (pcd) was sup-ported by early observations that host cells must be met-abolically active and, in some cases, the HR requires active host protein synthesis for its induction by fungi (Nozue et al ., 1977 and references therein; Heath, 2000)and bacteria (Keen et al ., 1981). Subsequent studies have shown that the HR is subject to genetic control, with factors important for its positive and negative regulation being identi?ed (T able 1). The morphology of cells under-going the HR at late stages suggest that it is a form of pcd with some apoptotic features. In particular, apoptotic-like bodies with avirulent Pseudomonas syringae infec-tions were observed (Levine et al ., 1996). We know of only

a few studies using ultrastructural analysis of morpholog-ical events that occur during a time course of the HR.Bestwick et al . (1995) found early changes in mitochon-drial morphology (swelling and cristae disorganization) in avirulent P . syringae -infected lettuce, similar to what occurs in animal cells undergoing apoptosis (Wakaba-yashi and Karbowski, 2001). Later stages of the infection were accompanied by membrane dysfunction (loss of abil-ity to be plasmolysed) and progressive vacuolization of the cytoplasm. Membrane damage was proposed to be the critical event for cell death. Apoptosis-related chroma-tin condensation and endonucleolytic cleavage were not reported. However, there was a gap in the time course in which these events may have occurred. Thus, the HR in this system has a subset of apoptotic features and may also be similar to a pcd process called oncosis that involves vacuolar disruption (Jones, 2000).

Further studies looking at a range of host–pathogen interactions using detailed time courses could establish whether apoptosis-related events such as mitochondrial swelling, chromatin condensation and endonucleolytic cleavage occur before general organelle dysfunction. This would be expected if the HR occurs by an apoptotic-like mechanism. T o begin to address this question, in collab-oration with Shigeyuki Mayama at Kobe University, we have analysed the interaction between oat and the fungal pathogen Puccinia coronata as well as the interaction between Arabidopsis and avirulent P . syringae . Strikingly,we often observe that cells adjacent to the ?rst cells that die have the apoptotic features of chromatin condensation and endonucleolytic cleavage in both interactions (Fig. 1).In these cells, we found no evidence for oncosis. We suggest that cells in an infection zone may die by multiple mechanisms, a possibility that requires further investiga-tion. It is also possible that there are differences in the mechanisms of pcd used in different host–pathogen interactions.

Evidence for endonucleolytic cleavage of DNA during the HR of cowpea was found during infections with cow-pea rust fungus (Heath, 2000). Additionally, the involve-ment of proteases in the RR and/or HR has been shown in a number of host–pathogen interactions (Heath, 2000;G rey, 2002). In animals, apoptosis often involves pro-teases called caspases (G reen and Reed, 1998).Although clear homologues of caspases have not been found in the complete genome of the model plant Arabi-dopsis , caspase-like activities in plants have been docu-mented biochemically or inferred from inhibitor studies (for example, see Elbaz et al ., 2002; Lincoln et al ., 2002; del Pozo and Lam, 2003). Lam et al . (2001) have suggested that, as in animals, mitochondria could have a role in controlling the HR. Evidence for a role of plastids in virus-induced HR is indicated by the change in the amount of cell death during infection in plants when levels of the

Cell death in plant pathogenesis 203

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

Table 1.Examples of proteins in addition to R gene products with HR-repressing (–), HR-enhancing (+) or both repressing and enhancing (–/+)effects.HR effect a Protein/functions

Comments/reference(s)

CPN1, possible phospholipid-binding protein and Ca 2+-sensitive modulator of environmental responses

Affects Arabidopsis HR with P . syringae /avrRpt2. Loss of function mutant causes activation of basal defences and spontaneous cell death (Jambunathan et al ., 2001)

–P AtCYS1, cystatin with cysteine proteinase inhibitory activity

Ectopic expression in Arabidopsis suspension cells or tobacco plants reduces cell death in response to avirulent P . syringae (Belenghi et al ., 2003)

–P

a -DOX1, oxygenates fatty acids and protects cells from oxidative stress Arabidopsis with reduced a -DOX1 levels show an increased cell death with paraquat treatment and an increased HR with P . syringae /avrRpm1. These plants also support higher growth of bacteria (Ponce de Leon et al ., 2002)–P LeETR4, ethylene receptor

T omato with antisense LeETR4 show a faster HR with X. campestris pv. vesicatoria (Ciardi et al ., 2001)

–FtsH, plastid protease presumed active during stress responses

Overexpression in tobacco enhances TMV resistance, reduced expression reduces resistance (Seo et al ., 2000)

–P HSR203J, similar to esterases/hydrolases T obacco with HSR203J antisense has an accelerated HR in response to avirulent P . syringae and P . parasitica . T ransgenic plants are also affected in the transcription of defence-related genes (T ronchet et al ., 2001)

–P b

LSD1, Zn 2+ ?nger protein. May interact with the related protein LOL1. May indirectly regulate superoxide levels

Loss of LSD1 in Arabidopsis leads to ectopic death and increased basal resistance (Jabs et al ., 1996; Dietrich et al ., 1997; Rusterucci et al ., 2001; Aviv et al ., 2002; Epple et al ., 2003)

–NAD7, mitochondria complex I subunit

Loss of NAD7 in tobacco enhances TMV resistance. Mutant has upregulated antioxidant enzymes (Dutilleul et al ., 2003)

–RDS, not yet cloned

Dominant suppressor of HR in oat–P . coronata interaction (Yu et al ., 2001)+P EDS1, lipase-like protein

Important for HR induction of Arabidopsis by P . parasitica (Parker et al ., 1996)+LOL1, Zn 2+ ?nger protein, may interact with LSD1, possible cellular redox sensor

Affects HR of Arabidopsis in response to avirulent P . syringae ; overexpression induces cell death (Epple et al ., 2003)

+

MEK, mitogen-activated protein kinase (MAPK) cascade component Gain of function allele in tobacco confers NbrbohB-dependent cell death

(Y oshioka et al ., 2003 and references therein). Other members of the MAPK cascade may also regulate the HR (Shirasu and Schulze-Lefert, 2003)

+P AtMYB30, possible transcription factor

Important for HR induction in Arabidopsis and tobacco by various pathogens (Vailleau et al ., 2002)

+

AmMYB308, possible transcription factor from antirrhinum. Negative regulator of phenolic acid metabolism when ectopically expressed in tobacco

Ectopic expression of AmMYB308 in tobacco leads to a faster HR in response to P . syringae , alterations in leaf morphology and precocious cell death (T amagnone et al ., 1998)+PTI, part of the PTO RR signal transduction cascade

Overexpression of PTI from tomato in tobacco enhances the HR induced by P . syringae /avrPto (Zhou et al ., 1995)

+

RAR1, component of the ubiquitin proteolysis system, interacts with SGT1Important for barley Mla 12-mediated HR but largely dispensable for Mlg -mediated HR during B. graminis infection (Freialdenhoven et al ., 1994). Functions with many R genes (Shirasu and Schulze-Lefert, 2003)

+

RIH, not yet cloned

Required for the HR as assayed indirectly by whole-cell auto?uorescence of oat in response to P . coronata . Resistance appears to be unaffected by the rih mutation (Yu et al ., 2001)

+E NbrohA/NbrbohB (gp91phox homologues),

components of the NADPH oxidase complex Reduced expression leads to reduced HR of tobacco infected with P . infestans (Y oshioka et al ., 2003)

+SGT1, interacts with RAR1

Important for the function of many R genes in a number of plant species (Shirasu and Schulze-Lefert, 2003)

+P SID2/EDS16, isochorismate synthase, probable salicylic acid biosynthetic enzyme

Arabidopsis mutants are hypersusceptible to many pathogens. Important for the HR in response to P . syringae /avrRpt2 (Zhang et al ., 2002)

–/+P

NDR1, possible membrane-associated protein, important for salicylic acid accumulation in response to P . syringae and UV-C, defective in H 2O 2 accumulation during the HR

HR-repressing in Arabidopsis in response to many P . syringae isolates carrying avr genes. HR-activating function in response to P . syringae /avrRpt2 (Century et al ., 1995; Shapiro and Zhang, 2001)–/+

AtRBOHD/AtRBOHF , components of the NADPH oxidase complex involved in producing

extracellular superoxide as part of the oxidative burst

The Arabidopsis double mutant has spontaneous cell death late in

development. Proteins are HR-repressing only in response to a P . parasitica isolate that induces a weak RR response; they are HR-promoting in response to avirulent P . syringae (T orres et al ., 2002)

–/+

RIN4, membrane-localized protein that can interact with R proteins RPM1 and RPS2 as well as the Avr protein AvrB. rps2– mutants expressing avrRpt2 have no RIN4 protein

Reduced expression in Arabidopsis leads to decreased HR upon infection with P . syringae /avrRpm1 and constitutive activation of cell death (Mackey et al ., 2002; Axtell and Staskawicz, 2003). Overexpression suppresses the HR induced by P . syringae /avrRpt2 (Mackey et al ., 2003)

a. P indicates pathogen-inducible gene, E indicates elicitor-inducible gene.

b. H. Lu and J. T . Greenberg, unpublished observations.

Proteins are organized in alphabetical order within each HR-in?uencing class.

204J. T . Greenberg and N. Y ao

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

plastid protease FtsH are increased or decreased (Seo et al ., 2000).

Signalling during the HR

A strategy for studying the signalling requirements and mechanism of cell death during the HR is to use simpli?ed experimental systems. In particular, pathogen-derived molecules, called elicitors, that induce HR-like cell death reactions have been used either by applying them directly to plant cells or by expressing the genes for these elicitors directly in plant cells. In some cases, the elicitors have been shown to be Avr products. Some researchers have also used pathogen infections of plant cell cultures. Such systems are potentially very powerful for identifying HR signalling components, for determining the relationship between cell death and other defence-related events and for studying the involvement of organelle changes in cell death.

Using such approaches, an oxidative burst (Heath,2000), ion channels (Atkinson et al ., 1996; Heath, 2000;Wendehenne et al ., 2002), NO (Delledonne et al ., 2001)and the interaction between some of these different sig-nals (Delledonne et al ., 2001) were implicated in HR con-trol. The exact role of these molecules is still the subject of intense investigations by many researchers. Using pro-toplasts of Arabidopsis , we recently found that the HR in response to speci?c Avr proteins could be https://www.doczj.com/doc/457717247.html,ing this system and careful ?ow cytometric and micro-scopic analysis, we have established that the HR involves a mitochondrial permeability transition (N. Y ao and J. T .G reenberg, unpublished observations). In animal cells,such transitions are often important components of the apoptotic mechanism because of the release of the cell death-inducing cytochrome c (G reen and Reed, 1998).Whether cytochrome c release is necessary for HR induc-tion has not yet been examined. However, Hansen (2000)found that Agrobacteria -infected maize suspension cells exhibited an apoptotic-like response that was inhibited in transgenic maize cells producing antiapoptotic proteins from animal viruses. Cell death was accompanied by cyto-chrome c release to the cytoplasm. A careful analysis of the timing of cytochrome c release relative to other apo-ptosis-related events needs to be established to deter-mine whether such release could be causal to cell https://www.doczj.com/doc/457717247.html,e of simpli?ed systems in which one or two proteins from pathogens are expressed in host cells has also revealed that the extracellular bacterial pathogen P . syrin-gae injects into plants antiapoptotic proteins in addition to the known proapoptotic Avr activities (Abramovitch et al .,2003; Espinosa et al ., 2003). Experiments by Abramovitch et al . (2003) in particular suggest that the antiapoptotic activity acts at a different place in the cell death pathway

from the proapoptotic activities. Indeed, the antiapoptotic

Fig. 1.Ultrastructural features of cells under-going the HR.

A, B and D. Primary leaves of oat cv. Shokan 1 were inoculated with the HR-inducing incom-patible race 226 of P . coronata f. sp. avenae for 24 h. The cells adjacent to the ?rst HR col-lapsed cell were observed. The corpses of the initial HR cells are fragmented, as evidenced by the chloroplast fragments (A) and membrane-bound body (representing an apoptotic-like body) containing swollen mitochondria (B) in the intercellular spaces. Cells adjacent to the fragmented initial collapsed cells had the typical apoptotic features of chromatin condensation and intact tonoplasts (A and B).

C. 21-day-old Arabidopsis (ecotype Col-0) leaves were in?ltrated with avirulent P . syringae carrying avrRpm1 at 5 ¥ 106 cfu ml -1 for 21 h. Note the initiation of chromatin condensation (arrows) in the cell adjacent to the ?rst dead cell.

D. Abundant gold label detecting endonucle-olytic cleavage DNA using an Apoptag plus ?uorescein in situ apoptosis detection kit (Inter-gen) can be seen in the condensed heterochro-matin portion. ALB, apoptotic-like body; Ch, chloroplast; CW, cell wall; Eu, euchromatin; Hc, heterochromatin; IS, intercellular space; M, mitochondrion; N, nucleus; V , vacuole. Bar in (A), (B) and (C) = 1 m m; in (D) = 200 nm.

Cell death in plant pathogenesis 205

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

activity was active in the heterologous system of yeast Saccharomyces cerevisiae (Abramovitch et al ., 2003). It is intriguing that P . syringae secretes antiapoptotic activi-ties. Doke (1983) showed that the fungal pathogen Phy-tophthora infestans also possesses both cell death-promoting and cell death-inhibitory molecules. We spec-ulate that, in some infections, the cell death pathway may be only partially or transiently activated. In support of this view, we have observed that host cells two or three cells distal to the primary fungus-infected cells undergoing an HR have some apoptotic features early after infection.However, similarly positioned cells late in the infection appeared to recover, showing thicker cell walls and fewer chloroplasts, but lacking apoptotic features (N. Y ao,unpublished observations). The role of such partially apo-ptotic and/or recovered cells remains to be elucidated.Interestingly, some viral pathogens of animals also use both proapoptotic and antiapoptotic proteins to manipu-late their hosts at different times in the infection (Munger et al ., 2003).

It is possible that cells in an infection zone undergoing an RR may not all die from the same signals. In particular,the HR has been suggested to require the correct relative levels of both NO and H 2O 2 to be induced in the host (Delledonne et al ., 2001). However, NO was found to be generated ?rst at cell surfaces in avirulent P . syringae -infected Arabidopsis at a time too late to be causal to the ?rst cell death events during the HR. The pattern of NO generation suggests that it has a role in cell–cell signalling and the spreading of cell death as an infection proceeds (Zhang et al ., 2003). Furthermore, inhibition of NO syn-thesis or action only attenuates the HR, in support of this role. Interestingly, an Arabidopsis AtbohF/AtbohD double mutant in NADPH oxidase complex components, thought to generate an oxidative burst during the HR, lacks detect-able H 2O 2 accumulation during an RR (T orres et al .,2002). This double mutant still activated some early cell death. T orres et al . (2002) suggested that the ?rst cell deaths occur independently of H 2O 2 and subsequent death requires H 2O 2 generation. Thus, initial cell deaths during the HR may be both NO and H 2O 2 independent.One caveat to this conclusion is that the studies on NO and H 2O 2 production relative to cell death relied largely on whole-tissue analyses. It is possible that highly localized subcellular generation of these signals was missed using these techniques. For example, when cell death in oat is induced by the toxin victorin, H 2O 2 accumulation is mainly found in localized mitochondrial pores (Y ao et al ., 2002a).Is the HR important for pathogen resistance?A number of attempts have been made to address the role of cell death in disease resistance. Many experiments have involved using pharmacology and genetics com-bined with an analysis of the timing of pathogen arrest relative to cell death. Interpretations of these experiments can be dif?cult as a number of genes and signal mole-cules affect more than just cell death (T able 1). What has emerged is that the importance of cell death in resistance depends on the host–pathogen interaction. This is because, as mentioned above, not all RR are the same because of differences in the strength of signalling and the downstream defences that are activated. Certainly, in some cases, pathogen arrest can occur in the absence of cell death (see below) and, conversely, cell death may not be suf?cient for pathogen arrest when other defences are compromised (Century et al ., 1995).

A seminal paper from Baulcombe and colleagues examined Rx -mediated resistance against the PVX virus (Bendahmane et al ., 1999). They found that resistance to this virus occurs without the HR. Intriguingly, the Rx pro-tein resembles many other R proteins that are known to trigger an HR. Ectopic expression of the viral Avr elicitor and the Rx protein can induce an Rx -dependent HR. This has led to the suggestion that, in some host–pathogen interactions, a phenomenon called ‘extreme resistance’without cell death occurs because suf?cient R protein signalling can arrest the pathogen. A quantitatively greater interaction would then lead to the HR, which Baul-combe and colleagues consider as a mechanism to rein-force the front-line ‘extreme resistance’ defence reaction in some host–pathogen interactions. A prediction of their idea is that some RR’s produce defences that are extremely rapid and robust, arresting the pathogen before cell death. Work from Schulze-Lefert and colleagues has shown that resistance against the obligate pathogen pow-dery mildew (Blumeria graminis f. sp. hordei ) ?ts well with this hypothesis (Freialdenhoven et al ., 1994; Peterh?nsel et al ., 1997). In the case of Mlg -mediated resistance,rapid and robust activation of cell wall alterations restricts pathogen growth before cell death. In contrast, Mla 12-mediated resistance is very closely correlated with the HR. Similarly, increasing the HR strength against an iso-late of the obligate pathogen Peronospora parasitica ,which normally triggers a weak RR, strongly enhances disease resistance (T orres et al ., 2002). In general, it is likely that cell death may have a larger contribution to defence with obligate pathogens that require living cells in order to replicate.

T o evaluate rigorously the contribution of the HR to the RR, the ideal experiment would involve the selective inhi-bition of the HR. T o do this, one needs to know the com-ponents of the cell death machinery and the selectivity of the reagent used to inhibit the machinery. As mentioned above, in plants, there is some evidence that the HR involves the activation of caspase-like activities. If in plants these caspase-like activities are truly speci?cally involved in activating pcd, then they provide an ideal target to

206J. T . Greenberg and N. Y ao

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

disrupt in order to test the involvement of the HR in resis-tance. A challenge for future studies is to establish rigor-ously whether these plant caspase-like activities are solely involved in cell death control.

T o test the possible involvement of plant caspases in resistance, del Pozo and Lam (2003) placed the caspase-inhibitory protein p35 from baculovirus in tobacco that normally shows resistance to TMV because of the pres-ence of the N resistance gene. They infected the trans-genic plants with TMV and found that the lesions were the same size and number as in infected wild-type plants, but their appearance was ‘less dehydrated’. This suggests that the process of pcd was partially interrupted. Strikingly,in p35-expressing plants, TMV was not restricted to local lesions as in non-transgenic N plants, but rather spread systemically throughout the plant. It will be important to examine the ultrastructure of the p35-expressing TMV-infected plants to see exactly how the HR is altered. How-ever, these results argue strongly for a role for the HR in limiting TMV replication. Importantly, a catalytically inac-tive form of p35 was ineffective in altering the host–virus interaction. A similar approach was attempted using plants with altered expression of the plant-derived Bax inhibitor-1 (BI-1) protein (Huckelhoven et al ., 2003).Overexpression of BI-1 attenuates cell death caused by Magnaporthe grisiae -derived elicitors (Matsumura et al .,2003). However, BI-1 manipulation also strongly affects other defences in addition to cell death (Huckelhoven et al ., 2003), making it possible that the cell death machin-ery is not the only target of BI-1. The identi?cation of the basal cell death machinery in plants will hopefully lead to additional rigorous experiments to test the role of the HR in resistance.

Other experiments to alter the HR, mainly using genet-ics and transgenic approaches to cause its attenuation or acceleration, may also be informative for examining the role of the HR in resistance. However, the interpretation of the experiments must be done cautiously, especially when pleiotropic effects of manipulations are known. For example, the RR can be uncoupled from the HR to most avirulent P . syringae in the dnd1 and hlm1 mutants of Arabidopsis (Yu et al ., 1998; Balague et al ., 2003). How-ever, these mutants have constitutively active defences,making the interpretation of the lack of an HR dif?cult.Many studies have found that conditions that caused attenuated or increased disease resistance also caused similar changes in HR strength at least with some patho-gens (Freialdenhoven et al ., 1994; Seo et al ., 2000; Jam-bunathan et al ., 2001; T ronchet et al ., 2001; T orres et al .,2002; Vailleau et al ., 2002; Epple et al ., 2003; Y oshioka et al ., 2003). Exceptions to this pattern have also been described for some host–pathogen combinations (for example, see Century et al ., 1995; Yu et al ., 1998; Rate and Greenberg, 2001; Ponce de Leon et al ., 2002; T orres

et al ., 2002). In most of these cases, however, the alter-ations were documented to be pleiotropic, affecting other defence-related processes or causing some spontaneous cell death.

Other functions for the HR

In addition to playing a role in limiting pathogen growth directly, the HR may have additional contributions to defence. Careful work by Kauffmann and colleagues using the fungal glycoprotein elicitin from Phytophthora megasperma to elicit HR-like cell death has shown a strong association between cell death and the subsequent activation of speci?c defences in neighbouring tissue (Costet et al ., 1999). Such systemic signalling is important for protecting plants from future infections. Similarly,plants showing an HR after infection with avirulent P . syrin-gae exhibit stronger systemic resistance to subsequent infection with virulent P . syringae than if the initial response lacks an HR (Shapiro and Zhang, 2001).Other forms of disease resistance-associated cell death

Induction of systemic signalling resulting in acquired resis-tance is often accompanied by the rapid induction of local-ized cell death by pathogens that normally would not induce rapid cell death (reviewed by G reenberg, 1997).For example, plants with acquired resistance show very rapid HR-like cell death upon virulent P . syringae infection.Signalling for this rapid cell death appears not to be associated with H 2O 2 accumulation, unlike the HR (Wolfe et al ., 2000). Overexpression of the Arabidopsis protein AtMyb30 in plants can also cause otherwise slow cell death-inducing pathogens to induce an HR-like response as well as inducing the HR-associated accumulation of oxylipins (Vailleau et al ., 2002). Finally, loss of function mlo and edr1 mutants have increased resistance to Blumeria graminis f. sp. hordei and Erysiphe cichora-cearum respectively (Peterh?nsel et al ., 1997; Frye and Innes, 1998). In the case of edr1 mutants, E. cichora-cearum triggers increased host cell death around the fun-gal hyphae (Frye and Innes, 1998). In the case of mlo , the mutant plants develop some spontaneous cell death, but disease resistance upon infection is thought not to require cell death (Peterh?nsel et al ., 1997). Whether cell death described in these various cases is mechanistically and/or functionally similar is not known.

Cell death during successful pathogen infections (susceptible host–pathogen interactions)

The occurrence of cell death during susceptible host–pathogen interactions in which pathogens can replicate

Cell death in plant pathogenesis 207

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

well in their hosts is common. A recent study of oat infected with a broad variety of pathogens (fungi, viruses and bacteria) established that many virulent pathogens induce cell death with apoptotic features (Y ao et al .,2002b). Apoptotic-like events occurred with widely differ-ent timing depending on the infectious agent and occurred in the directly infected cells and/or the neigh-bouring cells. Although morphological criteria argue that cells may die by the same mechanism, a more rigorous test would be to use an inhibitor of the apoptotic cell death machinery and examine whether disease symp-toms and pathogen growth are altered. This approach has the same caveats as mentioned for the HR studies,in that the speci?city of the reagents used must be established.

T omato producing the anticaspase baculovirus protein p35 showed reduced cell death with a well-characterized mycotoxin that can promote apoptotic cell death on its own (Lincoln et al ., 2002). These plants also showed reduced symptoms with a number of pathogens. Recall that tobacco carrying the same protein had less dehy-drated cell death and enhanced spread of a virus to which the plants had genetic resistance due to the N resistance gene (del Pozo and Lam, 2003). The alteration of both susceptible and resistance responses by the same anti-caspase protein argues that a cell death pathway target is common to the HR and susceptible host responses. In support of apoptotic-like cell death being important for symptoms caused by virulent pathogens, Dickman et al .(2001) found that a number of antiapoptotic proteins from animals, when ectopically expressed in plants, can protect them from symptoms caused by fungal pathogens. How-ever, a recent retraction of some of these data (Dickman et al ., 2003) suggests that caution should be used in form-ing any ?rm conclusions based on this study. It must be emphasized that these experiments assume that there is a unique plant target for p35 and each of the other anti-apoptotic proteins from non-plant organisms, and that this target’s only function is to control cell death. A challenge for future experiments is to test these assumptions rigorously.

That pcd and other disease symptoms result from host-encoded functions is supported by a number of ?ndings.Arabidopsis mutants that form disease-like lesions spon-taneously have been documented by us and others (Greenberg et al ., 2000; Pilloff et al ., 2002). One of these mutants, acd5, shows increased disease symptoms and modestly increased growth of P . syringae when infected before spontaneous lesion formation. Interestingly, ACD5encodes a ceramide kinase (CERK) that is induced during virulent P . syringae infections (Liang et al ., 2003). Ceram-ides are known to be bioactive lipids that activate apopto-sis in animals (Hannun and Obeid, 2002). We have shown that ceramide is suf?cient to induce apoptotic-like cell

death, while its phosphorylated derivative (the product of the CERK reaction) can partially block pcd in Arabidopsis protoplasts. Finally, acd5 mutant plants accumulate increased amounts of the ACD5 CERK substrate.T ogether, these ?ndings suggest that cell death activated by ceramide is important for P . syringae virulence.

Ceramides are part of the sphingolipid family of bioac-tive lipids (Hannun and Obeid, 2002). This sphingolipid pathway may be commonly targeted by plant pathogens.Indeed, some fungal pathogens secrete cerebrosides (which are derived from sphingolipids) that induce HR-like cell death in rice (Koga et al ., 1998). Additionally, a num-ber of fungal pathogens secrete related mycotoxins (for example, AAL and fumonisin) that cause pcd and disrupt sphingolipid metabolism (Abbas et al ., 1994).Multiple signals control susceptible cell death Like the HR, disease symptoms caused by at least some pathogens may be a mixture of events, each of which is under the control of different signals. In tomato–Xanth-omonas campestris interactions, disease symptoms are in part attributable to the known host signal molecules ethylene and salicylic acid (O’Donnell et al ., 2001). How-ever, plants blocked for these pathways show a normal initial cell death response but then fail to develop subse-quent cell death and chlorophyll loss (a late pathogenesis symptom). Ethylene-insensitive, ethylene-de?cient and salicylic acid-de?cient tomato plants were described as pathogen tolerant because they did not have altered rep-lication of the pathogen even though symptoms were attenuated. A similar ?nding was made with ethylene-insensitive Arabidopsis mutants infected with bacterial pathogens (Bent et al ., 1992). We have also found that attenuation of chlorophyll catabolism in Arabidopsis dur-ing P . syringae infection reduced overall cell death without reducing pathogen growth (Mach et al ., 2001).

A few additional examples of mutants and natural vari-ants of Arabidopsis that display tolerance to bacterial pathogens have been described (reviewed by Greenberg,1997). In most cases, the molecular basis of tolerance is not known. The Arabidopsis COI gene that confers sensi-tivity to the hormone jasmonic acid is implicated in the control of lesion formation during P . syringae infection (Kloek et al ., 2001). COI1 is thought to be involved in targeting proteins for degradation through the ubiquitin pathway (Xie et al ., 1998). Because genes involved in tolerance to bacterial pathogens seem largely to involve hormone signalling, it is likely that disease symptom development is not the only process altered in the tolerant plants. We note that the initial cell death events, which usually still occur even in tolerant plants during infection,appear to be important for contributing to pathogen repli-cation (see below).

208J. T . Greenberg and N. Y ao

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

Contribution of host cell death to virulent pathogen replication

In some host–pathogen interactions, pcd has a clear role in promoting pathogen growth. This is especially true for pathogens that secrete toxins to kill host cells rapidly,presumably to gain nutrition. We have already mentioned that the AAL toxin and related molecules induce pcd.Alternaria alternata f. sp. lycopersici lacking the AAL toxin has severely reduced growth on susceptible plants (Akamatsu et al ., 1997). Likewise, the pathogenicity of Cochliobolus victoriae was correlated with the secretion of the victorin toxins (Wolpert et al ., 2002). Exogenous application of victorin to susceptible oat plants causes an apoptotic-like response, including mitochondrial alter-ations (Curtis and Wolpert, 2002). Interestingly, patho-gens that secrete AAL or victorin have a very narrow host range, only infecting speci?c hosts. For this reason, AAL and victorin are considered to be host-speci?c toxins (Wolpert et al ., 2002).

In addition to the very clear cases outlined above, evi-dence has also emerged to support pcd as an important event promoting the growth of other pathogens. T omato carrying the anticaspase p35 protein show both attenu-ated symptoms and growth of normally virulent P . syringae (Lincoln et al ., 2002). In this case, the primary disease lesions were smaller in the transgenic p35-expressing plants than in the controls, and the replication of P . syrin-gae was reduced. These transgenic tomato showed reduced symptoms with other pathogens, but their repli-cation was not measured. Importantly, basal defences in the transgenic tomato appeared to be unaltered by the presence of p35. The involvement of pcd in promoting P .syringae virulence can also be inferred by the phenotype of Arabidopsis fbr mutants (Stone et al ., 2000). These mutants were selected based on their resistance to the pcd-inducing toxin fumonisin. When infected with virulent P . syringae , the fbr mutants showed reduced disease symptoms and pathogen replication. Basal defences in fbr mutants did not appear to be altered.Other roles for susceptible cell death

For some bacterial pathogens, cell death lesions associ-ated with water-soaked tissue may be formed to facilitate both release of the bacteria on to leaf surfaces and sub-sequent transmission of the bacteria. Some bacterial pro-teins that are injected into plant cells during infection seem to play a role speci?cally in inducing water soaking (Y ang et al ., 1994) and/or causing lesion formation (Badel et al ., 2003). We have found a tight correlation between severe disease symptoms and bacterial release on to leaf surfaces using the P . syringae –Arabidopsis model system (G uttman and G reenberg, 2001). We envisage such

release to be an important step in the dissemination of bacteria, especially in cases where bacteria do not cause a systemic infection.Concluding remarks

Many questions about the role, regulation and mechanism of pcd during host–plant interactions remain unanswered.Are there multiple mechanisms of cell death execution and regulation in infection zones and in different plant species? In which host–pathogen interactions is cell death truly important for disease resistance or susceptibility? We ?nd it intriguing that ectopic expression in plants of the anticaspase p35 protein affects cell death in both the HR and susceptible host–pathogen interactions. This raises the possibility that there is a basal cell death machinery engaged during the different responses. It is possible that,although some pcd steps are common between resistant and susceptible responses, some aspects of pcd are dif-ferent. Further analysis of pcd mechanisms in different conditions will be necessary to resolve this question. For future evaluation of the role of cell death in plant patho-genesis, it will be important to try to inhibit the cell death machinery selectively and simultaneously to monitor other defence- and pathogenesis-related events. Using this approach, it should be possible to determine whether cell death can be uncoupled from other responses and, if so,what its contribution to resistance or susceptibility is.Acknowledgements

We thank members of the Greenberg laboratory for helpful dis-cussions and comments on the manuscript. We acknowledge the help and support of Shigeyuki Mayama at Kobe University in whose laboratory the oat infection study shown in Fig. 1 was performed. Our research is funded by grants to J.T .G. from the NSF and NIH.

References

Abbas, H.K., Tanaka, T., Duke, S.O., Porter, J.K., Wray, E.M.,Hodges, L., et al. (1994) Fumonisin- and AAL-toxin-induced disruption of sphingolipid metabolism with accumulation of free sphingoid bases. Plant Physiol 106: 1085–1093.

Abramovitch, R., Kim, Y., Chen, S., Dickman, M., and Martin,G. (2003) Pseudomonas type III effector AvrPtoB induces plant disease susceptibility by inhibition of host pro-grammed cell death. EMBO J 22: 60–69.

Akamatsu, H., Itoh, Y., Kodama, M., Otani, H., and Kohmoto,K. (1997) AAL-toxin-de?cient mutants of Alternaria alter-nata tomato pathotype by restriction enzyme-mediated integration. Phytopathology 87: 967–972.

Atkinson, M.M., Mildland, S.L., Sims, J.J., and Keen, N.T.(1996) Syringolide triggers Ca 2+ in?ux, K + ef?ux, and extra-cellular alkalization in soybean cells carrying the disease-resistance gene rpG4. Plant Physiol 112: 297–302.

Aviv, D.H., Rusterucci, C., Holt, B.F., III, Dietrich, R.A.,

Cell death in plant pathogenesis 209

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

Parker, J.E., and Dangl, J.L. (2002) Runaway cell death,but not basal disease resistance. lsd1 is SA- and NIM1/NPR1-dependent. Plant J 29: 381–391.

Axtell, M.J., and Staskawicz, B.J. (2003) Initiation of RPS2-speci?ed disease resistance in Arabidopsis is coupled to the AvrRpt2-directed elimination of RIN4. Cell 112: 369–377.

Badel, J.L., Nomura, K., Bandyopadhyay, S., Shimizu, R.,Collmer, A., and He, S.Y. (2003) Pseudomonas syringae pv. tomato DC3000 HopPtoM (CEL ORF3) is important for lesion formation but not growth in tomato and is secreted and translocated by the Hrp type III secretion system in a chaperone-dependent manner. Mol Microbiol 49: 1239–1251.

Balague, C., Lin, B., Alcon, C., Flottes, G., Malmstrom, S.,Kohler, C., et al. (2003) HLM1, an essential signaling com-ponent in the hypersensitive response, is a member of the cyclic nucleotide-gated channel ion channel family. Plant Cell 15: 365–379.

Belenghi, B., Acconcia, F., Trovato, M., Perazzolli, M.,Bocedi, A., Polticelli, F., et al. (2003) AtCYS1, a cystatin from Arabidopsis thaliana , suppresses hypersensitive cell death. Eur J Biochem 270: 2593–2604.

Bendahmane, A., Kanyuka, K., and Baulcombe, D.C. (1999)The Rx gene from potato controls separate virus resistance and cell death responses. Plant Cell 11: 781–791.

Bent, A.F., Innes, R.W., Ecker, J.R., and Staskawicz, B.J.(1992) Disease development in ethylene-insensitive Arabidopsis thaliana infected with virulent and avirulent Pseudomonas and Xanthomonas pathogens. Mol Plant–Microbe Interact 5: 372–378.

Bestwick, C.S., Bennet, M.H., and Mans?eld, J.W. (1995)Hrp mutant of Pseudomonas syringae pv. phaseolicola induces alterations but not membrane damage leading to the hypersensitive reaction in lettuce. Plant Physiol 108:503–516.

Century, K.S., Holub, E.B., and Staskawicz, B.J. (1995)NDR1, a locus of Arabidopsis thaliana that is required for disease resistance to both a bacterial and a fungal patho-gen. Proc Natl Acad Sci USA 92: 6597–6601.

Ciardi, J.A., Tieman, D.M., Jones, J.B., and Klee, H.J. (2001)Reduced expression of the tomato ethylene receptor gene LeETR4 enhances the hypersensitive response to Xanth-omonas campestris pv. vesicatoria . Mol Plant–Microbe Interact 14: 487–495.

Costet, L., Cordelier, S., Dorey, S., Baillieul, F., Fritig, B., and Kauffmann, S. (1999) Relationship between localized acquired resistance (LAR) and the hypersensitive response (HR): HR is necessary for LAR to occur and salicylic acid is not suf?cient to trigger LAR. Mol Plant–Microbe Interact 12: 655–662.

Curtis, M.J., and Wolpert, T.J. (2002) The oat mitochondrial permeability transition and its implication in victorin binding and induced cell death. Plant J 29: 295–312.

Delledonne, M., Zeier, J., Marocco, A., and Lamb, C. (2001)Signal interactions between nitric oxide and reactive oxy-gen intermediates in the plant hypersensitive disease resis-tance response. Proc Natl Acad Sci USA 98: 13454–13459.

Dickman, M.B., Park, Y.K., Oltersdorf, T., Li, W., Clemente,T., and French, R. (2001) Abrogation of disease develop-ment in plants expressing animal antiapoptotic genes. Proc Natl Acad Sci USA 98: 6957–6962.

Dickman, M.B., Park, Y.K., Oltersdorf, T., Li, W., Clemente,

T., and French, R. (2003) Correction for: Abrogation of disease development in plants expressing animal antiapo-ptotic genes. Proc Natl Acad Sci USA 100: 11816.

Dietrich, R.A., Richberg, M.H., Schmidt, R., Dean, C., and Dangl, J.L. (1997) A novel zinc ?nger protein is encoded by the Arabidopsis LSD1 gene and functions as a negative regulator of plant cell death. Cell 88: 685–694.

Doke, N. (1983) G eneration of superoxide anion by potato tuber protoplasts during the hypersensitive response to hyphal wall components of Phytophthora infestans and speci?c inhibition of the reaction by suppressors of hyper-sensitivity. Physiol Plant Pathol 23: 359–367.

Dutilleul, C., Garmier, M., Noctor, G., Chantal, M., Chetrit, P.,Foyer, C.H., et al. (2003) Leaf mitochondria modulate whole cell redox homeostasis, set antioxidant capacity, and determine stress resistance through altered signaling and diurnal regulation. Plant Cell 15: 1212–1226.

Elbaz, M., Avni, A., and Weil, M. (2002) Constitutive caspase-like machinery executes programmed cell death in plant cells. Cell Death Differ 9: 726–733.

Epple, P., Mack, A.A., Morris, V.R.F., and Dangl, J.L. (2003)Antagonistic control of oxidative stress-induced cell death in Arabidopsis by two related, plant-speci?c zinc ?nger proteins. Proc Natl Acad Sci USA 100: 6831–6836.

Espinosa, A., Guo, M., Tam, V.C., Fu, Z.Q., and Alfano, J.R.(2003) The Pseudomonas syringae type III-secreted pro-tein HopPtoD2 possesses protein tyrosine phosphatase activity and suppresses programmed cell death in plants.Mol Microbiol 49: 377–387.

Freialdenhoven, A., Scherag, B., Hollricher, K., Collinge, D.B.,Thordal-Christensen, H., and Schulze-Lefert, P. (1994) Nar-1and Nar-2, two loci required for Mla 12-speci?ed race-speci?c resistance to powdery mildew in barley. Plant Cell 6: 983–994.

Frye, C.A., and Innes, R.W. (1998) An Arabidopsis mutant with enhanced resistance to powdery mildew. Plant Cell 10: 947–956.

Green, D.R., and Reed, J.C. (1998) Mitochondria and apop-tosis. Science 281: 1309–1312.

G reenberg, J.T. (1997) Programmed cell death in plant–pathogen interactions. Annu R ev Plant Physiol Plant Mol Biol 48: 525–545.

G reenberg, J.T., Silverman, F.P., and Liang, H. (2000)Uncoupling salicylic acid-dependent cell death and defense-related responses from disease resistance in the Arabidopsis mutant ACD5. Genetics 156: 341–350.

Grey, W.M. (2002) Plant defence: a new weapon in the arse-nal. Curr Biol 14: R352–R354.

G uttman, D.S., and G reenberg, J.T. (2001) Functional analysis of type III effectors AvrRpt2 and AvrRpm1 of Pseudomonas syringae with the use of a single copy genomic integration system. Mol Plant–Microbe Interact 14: 145–155.

Hannun, Y.A., and Obeid, L.M. (2002) The ceramide-centric universe of lipid-mediating cell regulation: stress encoun-ters of the lipid kind. J Biol Chem 277: 25847–25850.Hansen, G. (2000) Evidence for Agrobacterium -induced apo-ptosis in maize cells. Mol Plant–Microbe Interact 13: 649–657.

Heath, M. (2000) Hypersensitive response-related cell death.Plant Mol Biol 44: 321–334.

Huckelhoven, R., Dechert, C., and Kogen, K.-H. (2003) Over-expression of barley BAX inhibitor 1 induces breakdown of mlo -mediated penetration resistance to Blumeria graminis .Proc Natl Acad Sci USA 100: 5555–5560.

210J. T . Greenberg and N. Y ao

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

Jabs, T., Dietrich, R.A., and Dangl, J.L. (1996) Initiation of runaway cell death in an Arabidopsis mutant by extracellu-lar superoxide. Science 273: 1853–1856.

Jambunathan, N., Siani, J.M., and McNellis, T.W. (2001) A humidity-sensitive Arabidopsis copine mutant exhibits pre-cocious cell death and increased disease resistance. Plant Cell 13: 2225–2240.

Jia, Y., McAdams, S.A., Bryan, G.T., Hershey, H.P., and Valent, B. (2000) Direct interaction of resistance gene and avirulence gene products confers rice blast resistance.EMBO J 19: 4004–4014.

Jones, A. (2000) Does the plant mitochondrion integrate cel-lular stress and regulate programmed cell death? Trends Plant Sci 5: 225–230.

Keen, N.T., Ersek, T., Long, M., Bruegger, B., and Holliday,M. (1981) Inhibition of the hypersensitive reaction of soy-bean leaves to incompatible Pseudomonas spp. by blasti-cidin S, streptomycin or elevated temperature. Physiol Plant Pathol 18: 325–337.

Kloek, A.P., Verbsky, M.L., Sharma, S.B., Schoelz, J.E.,Vogel, J., Kessig, D.F., et al. (2001) Resistance to Pseudomonas syringae conferred by an Arabidopsis thaliana coronatine-insensitive (coi1) mutation occurs through two distinct mechanisms. Plant J 26: 509–522.

Koga, J., Y amauchi, T., Shimura, M., Ogawa, N., Oshima, K.,Umemura, K., et al. (1998) Cerebrosides A and C, sphin-golipid elicitors of hypersensitive cell death and phytoalexin accumulation in rice plants. J Biol Chem 273: 31985–31991.

Lam, E., Kato, N., and Lawton, M. (2001) Programmed cell death, mitochondria and the plant hypersensitive response.Nature 411: 848–853.

Leister, R.T., and Katagiri, F. (2000) A resistance gene prod-uct of the nucleotide binding site–leucine rich repeats class can form a complex with bacterial avirulence proteins in vivo . Plant J 22: 345–354.

Levine, A., Pennell, R.I., Alvarez, M.E., Palmer, R., and Lamb, C. (1996) Calcium-mediated apoptosis in a plant hypersensitive disease resistance response. Curr Biol 6:427–437.

Liang, H., Y ao, N., Song, J.T., Luo, S., Lu, H., and Greenberg,J.T. (2003) Ceramides modulate programmed cell death in plants. Genes Dev 17: 2636–2641.

Lincoln, J.E., Richael, C., Overduin, B., Smith, K., Bostock,R., and Gilchrist, D.G. (2002) Expression of the antiapop-totic baculovirus p35 gene in tomato blocks programmed cell death and provides broad-spectrum resistance to dis-ease. Proc Natl Acad Sci USA 99: 15217–15221.

Mach, J.M., Castillo, A.R., Hoogstraten, R., and Greenberg,J.T. (2001) The Arabidopsis -accelerated cell death gene ACD2 encodes red chlorophyll catabolite reductase and suppresses the spread of disease symptoms. Proc Natl Acad Sci USA 98: 771–776.

Mackey, D., Holt, B.F., Wiig, A., and Dangl, J.L. (2002) RIN4interacts with Pseudomonas syringae type III effector mol-ecules and is required for RPM1-mediated resistance in Arabidopsis . Cell 108: 743–754.

Mackey, D., Belkhadir, Y., Alonso, J.M., Ecker, J.R., and Dangl, J.L. (2003) Arabidopsis RIN4 is a target of the type III virulence effector AvrRpt2 and modulates RPS2-mediated resistance. Cell 112: 379–389.

Matsumura, H., Nirasawa, S., Kiba, A., Urasaki, N., Saitoh,H., Ito, M., et al. (2003) Overexpression of Bax inhibitor

suppresses the fungal elicitor-induced cell death in rice (Oryza sativa L.) cells. Plant J 33: 425–434.

Morel, J.-B., and Dangl, J.L. (1997) The hypersensitive response and the induction of cell death in plants. Cell Death Differ 4: 671–683.

Munger, J., Zhou, G., and Roizman, B. (2003) Cell death on demand: herpes simplex viruses and apoptosis. In SGM Symposium 62: Microbial Subversion of Host Cells .O’Connor, C.D., and Smith, D.G .E. (eds). Cambridge:Cambridge University Press, University of Edinburgh, pp.229–245.

Nozue, M., Tomiyama, K., and Doke, N. (1977) Effect of blasticidin S on development of potential of potato tuber cells to react hypersensitively to infection by Phytophthora infestans . Physiol Plant Pathol 10: 181–189.

O’Donnell, P.J., Jones, J.B., Antoine, F.R., Ciardi, J., and Klee, H.J. (2001) Ethylene-dependent salicylic acid regu-lates an expanded cell death response to a plant pathogen.Plant J 25: 315–323.

Parker, J.E., Holub, E.B., Frost, L.N., Falk, A., G unn, N.D.,and Daniels, M.J. (1996) Characterization of eds1, a muta-tion in Arabidopsis suppressing resistance to Peronospora parasitica speci?ed by several different RPP genes. Plant Cell 8: 2033–2046.

Peterh?nsel, C., Freialdenhoven, A., Kurth, J., Kolsch, R.,and Schulze-Lefert, P. (1997) Interaction analyses of genes required for resistance responses to powdery mil-dew in barley reveal distinct pathways leading to leaf cell death. Plant Cell 9: 1397–1409.

Pilloff, R.K., Devadas, S.K., Enyedi, A., and Raina, R. (2002)The Arabidopsis gain-of-function mutant dll1 spontane-ously develops lesions mimicking cell death associated with disease. Plant J 30: 61–70.

Ponce de Leon, I., Sanz, A., Hamberg, M., and Castresana,C. (2002) Involvement of the Arabidopsis a -DOX1 fatty acid dioxygenase in protection against oxidative stress and cell death. Plant J 29: 61–72.

del Pozo, O., and Lam, E. (2003) Expression of the baculov-irus p35 protein in tobacco affects cell death progression and compromises N gene-mediated disease resistance response to tobacco mosaic virus . Mol Plant–Microbe Interact 16: 485–494.

Rate, D.N., and G reenberg, J.T. (2001) The Arabidopsis aberrant growth and death2 mutant shows resistance to Pseudomonas syringae and reveals a role for NPR1 in suppressing hypersensitive cell death. Plant J 27: 203–211.

Ritter, C., and Dangl, J. (1996) Interference between two speci?c pathogen recognition events mediated by distinct plant disease resistance genes. Plant Cell 8:251–257.

Rusterucci, C., Aviv, D.H., Holt, B.F., III, Dangl, J.L., and Parker, J.E. (2001) The disease resistance signaling com-ponents EDS1 and PAD4 are essential regulators of the cell death pathway controlled by LSD1 in Arabidopsis .Plant Cell 13: 2211–2224.

Seo, S., Okamoto, M., Iwai, T., Iwano, M., Fukui, K., Isogai,A., et al. (2000) Reduced levels of chloroplast FtsH pro-tein in tobacco mosaic virus-infected tobacco leaves accelerate the hypersensitive reaction. Plant Cell 12:917–932.

Shao, F., Golstein, C., Ade, J., Stoutemyer, M., Dixon, J.E.,and Innes, R.W. (2003) Cleavage of Arabidopsis PBS1 by a bacterial type III effector. Science 301: 1230–1233.

Cell death in plant pathogenesis 211

? 2004 Blackwell Publishing Ltd, Cellular Microbiology , 6, 201–211

Shapiro, A.D., and Zhang, C. (2001) The role of NDR1 in avirulence gene-directed signaling and control of pro-grammed cell death in Arabidopsis . Plant Physiol 127:1089–1101.

Shirasu, K., and Schulze-Lefert, P. (2003) Complex forma-tion, promiscuity and multi-functionality: protein interac-tions in disease resistance pathways. Trends Plant Sci 6:252–258.

Stone, J.M., Heard, J.E., Asai, T., and Ausubel, F.M. (2000)Simulation of fungal-mediated cell death by fumonisin B1and selection of fumonisin B1-resistant (fbr ) Arabidopsis mutants. Plant Cell 12: 1811–1822.

Tamagnone, L., Merida, A., Stacey, N., Plaskitt, K., Parr, A.,Chang, C.F., et al. (1998) Inhibition of phenolic acid metab-olism results in precocious cell death and altered cell mor-phology in leaves of transgenic tobacco plants. Plant Cell 10: 1801–1816.

Tang, X., Frederick, R.D., Zhou, J., Halterman, D.A., Jia, Y.,and Martin, G.B. (1996) Initiation of plant disease resis-tance by physical interaction of AvrPto and Pto kinase.Science 274: 2060–2063.

Torres, M.A., Dangl, J.L., and Jones, J.D.G. (2002) Arabidop-sis gp91phox homologues AtrbohD and AtrbohF are required for accumulation of reactive oxygen intermediates in the plant defense response. Proc Natl Acad Sci USA 99: 517–522.

Tronchet, M., Ranty, B., Marco, Y., and Roby, D. (2001)HSR203 antisense suppression in tobacco accelerates development of hypersensitive cell death. Plant J 27: 115–127.

Vailleau, F., Daniel, X., Tronchet, M., Montillet, J.-L., Trianta-phylides, C., and Roby, D. (2002) A R2R3-MYB gene,AtMYB30, acts as a positive regulator of the hypersensitive cell death program in plants in response to pathogen attack. Proc Natl Acad Sci USA 99: 10179–10184.

Vanacker, H., Lu, H., Rate, D.N., and Greenberg, J.T. (2001)A role for salicylic acid and NPR1 in regulating cell growth in Arabidopsis. Plant J 28: 209–216.

Wakabayashi, Y., and Karbowski, M. (2001) Structural changes of mitochondria related to apoptosis. Biol Signals Recept 10: 26–56.

Wendehenne, D., Lamotte, O., Frachisse, J.-M., Barbier-Brygoo, H., and Pugin, A. (2002) Nitrate ef?ux is an essen-tial component of the cryptogein signaling pathway leading to defense responses and hypersensitive cell death in tobacco. Plant Cell 14: 1937–1951.

Wolfe, J., Hutcheon, W.J., Higgins, V.J., and Cameron,R.K. (2000) A functional gene-for-gene interaction is required for the production of an oxidative burst in response to infection with avirulent Pseudomonas syrin-gae pv. tomato . Arabidopsis thaliana . Physiol Mol Plant Pathol 56: 253–361.

Wolpert, T.J., Dunkle, L.D., and Ciuffetti, L.M. (2002) Host-selective toxins and avirulence determinants: what’s in a name? Annu Rev Phytopathol 40: 251–285.

Xie, S.-X., Feys, B.F., James, S., Nieto-Rostro, M., and Turner, J.G. (1998) COI1: an Arabidopsis gene required for jasmonate-regulated defenses and fertility. Science 280:1091–1094.

Y ang, Y., De Feyter, R., and G abriel, D.W. (1994) Host-speci?c symptoms and increased release of Xanthomonas citri and X. campestris pv. malvacearum from leaves are determined by the 102-bp tandem repeats of pthA and avrb6, respectively. Mol Plant–Microbe Interact 7: 345–355.

Y ao, N., Tada, Y., Sakamoto, M., Nakayashiki, H., Park, P.,Tosa, Y., et al. (2002a) Mitochondrial oxidative burst involved in apoptotic response in oats. Plant J 30: 567–579.

Y ao, N., Imai, S., Tada, Y., Nakayashiki, H., Tosa, Y., Park,P., et al. (2002b) Apoptotic cell death is a common response to pathogen attack in oats. Mol Plant–Microbe Interact 15: 1000–1007.

Y oshioka, H., Numata, N., Nakajima, K., Katou, S., Kawakita,K., Rowland, O., et al. (2003) Nicotiana benthamiana gp91phox homologs NbrbohA and NbrbohB participate in H 2O 2 accumulation and resistance to Phytophthora infestans. Plant Cell 15: 706–718.

Yu, G.-X., Braun, E., and Wise, R.P. (2001) Rds and Rih mediate hypersensitive cell death independent of gene-for-gene resistance to the oat crown rust pathogen Puccinia coronata f. sp . avenae . Mol Plant–Microbe Interact 14: 1376–1383.

Yu, I.C., Parker, J., and Bent, A.F. (1998) G ene-for-gene disease resistance without the hypersensitive response in Arabidopsis dnd1 mutant. Proc Natl Acad Sci USA 95:7819–7824.

Zhang, C., Czymmek, K.J., and Shapiro, A.D. (2002) Two pathways act in an additive rather than obligatorily syner-gistic fashion to induce systemic acquired resistance and PR gene expression in MBC . Plant Biol 2: 9.

Zhang, C., Czymmek, L.J., and Shapiro, A.D. (2003) Nitric oxide does not trigger early programmed cell death events but may contribute to cell-to-cell signaling governing pro-gression of the Arabidopsis hypersensitive response. Mol Plant–Microbe Interact 16: 962–972.

Zhou, J., Loh, Y.-T., Bressan, R.A., and Martin, G.B. (1995)The tomato gene Pti encodes a serine/threonine kinase that is phosphorylated by Pto and is involved in the hyper-sensitive response. Cell 83: 925–935.

相关主题
文本预览
相关文档 最新文档