24.12.2012 Views

Redox Changes during the Legume–Rhizobium Symbiosis

Redox Changes during the Legume–Rhizobium Symbiosis

Redox Changes during the Legume–Rhizobium Symbiosis

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Molecular Plant Volume 2 Number 3 Pages 370–377 May 2009 REVIEW ARTICLE<br />

<strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong><br />

<strong>Symbiosis</strong><br />

Christine Chang, Isabelle Damiani, Alain Puppo and Pierre Frendo 1<br />

UMR INRA–Université de Nice–Sophia Antipolis–CNRS ‘Interactions Biotiques et Santé Végétale’, 400 Route des Chappes, BP167, 06903 Sophia Antipolis Cedex,<br />

France<br />

ABSTRACT Reactive Oxygen Species (ROS) are continuously produced as a result of aerobic metabolism or in response to<br />

biotic and abiotic stresses. ROS are not only toxic by-products of aerobic metabolism, but are also signaling molecules<br />

involved in plant growth and environmental adaptation. Antioxidants can protect <strong>the</strong> cell from oxidative damage by scavenging<br />

<strong>the</strong> ROS. Thus, <strong>the</strong>y play an important role in optimizing cell function by regulating cellular redox state and modifying<br />

gene expression. This article aims to review recent studies highlighting <strong>the</strong> role of redox signals in establishing and<br />

maintaining symbiosis between rhizobia and legumes.<br />

Key words: Oxidative and photo-oxidative stress; cell differentiation/specialization; gene regulation; symbiosis; legume.<br />

NITROGEN FIXING SYMBIOSIS<br />

Nitrogen, a component of many bio-molecules, is essential for<br />

growth and development of all organisms. Most nitrogen<br />

exists in <strong>the</strong> atmosphere, and utilization of this source is important<br />

in avoiding nitrogen starvation. However, <strong>the</strong> ability to fix<br />

atmospheric dinitrogen (N2) via <strong>the</strong> nitrogenase enzyme complex<br />

is restricted to certain types of diazotrophic archeobacteria<br />

and eubacteria. Rhizobia, a type of gram-negative soil<br />

bacteria, are able to reduce atmospheric nitrogen through<br />

a symbiotic interaction with plants. They are located within<br />

<strong>the</strong> Rhizobiaceae phylogenetic family (a-proteobacteria) and<br />

have <strong>the</strong> unique ability to infect and establish symbiosis with<br />

leguminous plants. This symbiosis reduces <strong>the</strong> need for nitrogen<br />

fertilizers for agriculturally important plants such as soybean<br />

and alfalfa.<br />

The symbiotic interaction is initiated when <strong>the</strong> bacterium<br />

infects <strong>the</strong> root hair, and <strong>the</strong> host forms root nodules where<br />

nitrogen fixation will occur (Dénarié and Cullimore, 1993).<br />

In order for this to happen, <strong>the</strong> host must first identify rhizobia<br />

as beneficial partners (Figure 1A). Plants secrete (iso)flavonoids<br />

that are recognized by <strong>the</strong> compatible bacteria. They bind and<br />

activate bacterial NodD proteins, which are members of <strong>the</strong><br />

LysR family of transcriptional regulators, resulting in <strong>the</strong> induction<br />

of <strong>the</strong> nodulation genes (nod genes) (Peck et al., 2006).<br />

Nod genes encode proteins that syn<strong>the</strong>size and export specific<br />

lipochito-oligosaccharides called Nod factors (NFs). NFs activate<br />

<strong>the</strong> root infection process and initiate cell division in<br />

<strong>the</strong> root cortex. NFs are not <strong>the</strong> only bacterial signals necessary<br />

for symbiosis: surface lipopolysaccharides (LPS) also contribute<br />

at various stages of symbiotic development (Oldroyd and<br />

Downie, 2008).<br />

Host recognition of <strong>the</strong> NFs leads to increased calcium levels<br />

in <strong>the</strong> root hairs, followed by strong calcium spiking, and lastly<br />

alteration of <strong>the</strong> root hair cytoskeleton (Sieberer et al., 2005).<br />

Plant root hairs curl and trap rhizobial bacteria and induction<br />

of root cortical cell division establishes a meristem and nodule<br />

primordium (Figure 1B). The bacteria within this colonized<br />

curled root hair induce <strong>the</strong> formation of infection threads that<br />

are ingrowths of <strong>the</strong> root-hair cell membrane filled with bacteria<br />

(Figure 1C). In <strong>the</strong> inner plant cortex, bacteria exit <strong>the</strong> infection<br />

thread and enter nodule cells by endocytosis. Each<br />

bacterial cell is endocytosed in a membrane-bound compartment<br />

(Brewin, 2004). This organelle-like structure, including<br />

<strong>the</strong> plant-derived membrane and <strong>the</strong> bacteroid (<strong>the</strong> differentiated<br />

bacterium), is called <strong>the</strong> symbiosome. An infected plant<br />

cell may be packed with thousands of symbiosomes.<br />

This bacteroid is provided with a low oxygen environment<br />

that allows expression of enzymes of <strong>the</strong> nitrogenase complex<br />

and thus nitrogen fixation (Fischer, 1994). This microaerobic environment<br />

within <strong>the</strong> nodule is adjusted by a plant-produced<br />

1<br />

To whom correspondence should be addressed. E-mail frendo@unice.fr,<br />

fax +33-492386587, tel. +33-492386638.<br />

ª The Author 2008. Published by <strong>the</strong> Molecular Plant Shanghai Editorial<br />

Office in association with Oxford University Press on behalf of CSPP and<br />

IPPE, SIBS, CAS.<br />

doi: 10.1093/mp/ssn090, Advance Access publication 26 December 2008<br />

Received 18 September 2008; accepted 18 November 2008<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


oxygen-binding protein, <strong>the</strong> leghaemoglobin (Downie, 2005;<br />

Ott et al., 2005). Bacteroid differentiation is concomitant with<br />

<strong>the</strong> down-regulation of most metabolic processes and an increase<br />

in <strong>the</strong> gene products involved in nitrogen fixation<br />

and respiration. Respiratory activity provides nitrogenase with<br />

16 molecules of ATP and eight electrons that are estimated to<br />

be required to reduce one molecule of N2 to two molecules of<br />

NH4 + (Jones et al., 2007). Ammonium is secreted by <strong>the</strong> bacteroid<br />

and assimilated by plant cells through primarily glutamine<br />

and asparagine syn<strong>the</strong>tases. Carbon is supplied by <strong>the</strong> plant<br />

in <strong>the</strong> form of a dicarboxylic acid such as malate and this provides<br />

metabolites and energy for bacteroid differentiation and<br />

nitrogen fixation (Jones et al., 2007).<br />

MODULATION OF REACTIVE OXYGEN<br />

SPECIES DURING NITROGEN FIXING<br />

SYMBIOSIS<br />

Reactive Oxygen Species (ROS) are constantly produced by<br />

plants as a consequence of aerobic metabolism <strong>during</strong> <strong>the</strong>ir<br />

development or in response to abiotic and biotic stresses.<br />

ROS includes singlet oxygen ( 1 O2), superoxide anion (O2 .– ), hydrogen<br />

peroxide (H2O2), and hydroxyl radical (OH . ) species.<br />

These molecules are highly toxic, since <strong>the</strong>y are able to modify<br />

all <strong>the</strong> primary constituents of <strong>the</strong> cell such as lipids, DNA, carbohydrates,<br />

and proteins (Moller et al., 2007). Their toxicity<br />

leads to senescence and cell death (Overmyer et al., 2003; Rivero<br />

Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong> | 371<br />

Figure 1. The Different Steps of Nitrogen Fixing<br />

<strong>Symbiosis</strong>.<br />

(A) Recognition of <strong>the</strong> partners.<br />

(B) Root hair curling.<br />

(C) Infection thread growth.<br />

(D) M. truncatula nodule with meristematic<br />

(I), infection (II), fixing (III), and senescent<br />

(IV) zones. NFs, Nod factors.<br />

et al., 2007). ROS are also involved in <strong>the</strong> regulation of plant<br />

metabolism as secondary messengers in many pathways associated<br />

with plant development and environmental stress<br />

responses (Apel and Hirt, 2004; Carol and Dolan, 2006; Gechev<br />

et al., 2006). Such ROS production may be controlled and deliberate,<br />

enabling plants to better adapt to <strong>the</strong> environment<br />

(Pitzschke et al., 2006).<br />

Accumulation of ROS has been observed in Medicago sativa<br />

(alfalfa) <strong>during</strong> <strong>the</strong> interaction between rhizobia and legumes<br />

(Santos et al., 2001). The ROS accumulation, detected 12 h<br />

post treatment with NFs, is linked to <strong>the</strong> NFs signaling transduction<br />

pathway as S. meliloti mutants that produce altered<br />

NFs or a non-nodulating Medicago truncatula mutant,<br />

dmi1-1, are impaired in <strong>the</strong> ability to elicit ROS production<br />

(Ramu et al., 2002). Recently, fast and transient ROS changes<br />

were observed in root hair cells, minutes after treatment with<br />

NFs in Phaseolus vulgaris (Cardenas et al., 2008). Inhibition of<br />

this ROS production prevents root hair curling and formation<br />

of infection threads (Peleg-Grossman et al., 2007). Thus, <strong>the</strong><br />

production of ROS may not be a plant defense response to<br />

<strong>the</strong> microbe, but ra<strong>the</strong>r a process that is needed for <strong>the</strong> development<br />

of a proper interaction. The importance of ROS<br />

production has been confirmed in a S. meliloti strain overexpressing<br />

a catalase. This bacterium, acting as an H2O2 sink, provoked<br />

a delayed nodulation and enlargement of infection<br />

threads (Jamet et al., 2007). Hydrogen peroxide production<br />

may be involved in cell wall formation and in regulation of<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


372 | Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong><br />

<strong>the</strong> infection thread rigidity that is needed for its progression<br />

through <strong>the</strong> plant. The induction of genes encoding rhizobiuminduced<br />

peroxidase such as Rip1 (Cook et al., 1995) or Srprx1<br />

(Den Herder et al., 2007) is consistent with this hypo<strong>the</strong>sis.<br />

In <strong>the</strong> semi-aquatic legume Sesbania rostrata, invasion may<br />

happen through fissures in <strong>the</strong> lateral root base by entry to <strong>the</strong><br />

cortex via intercellular cracks (Goormachtig et al., 2004). In this<br />

model, <strong>the</strong> formation of infection pockets also depends on NFs<br />

and was shown to be associated with localized cell death and<br />

<strong>the</strong> production of large amounts of H2O2 (D’Haeze et al.,<br />

2003). Moreover, pharmacological experiments showed that<br />

ROS were required for nodule initiation (D’Haeze et al., 2003).<br />

In contrast with <strong>the</strong>se results describing <strong>the</strong> production of<br />

ROS <strong>during</strong> nodule development, NFs have been shown to inhibit<br />

<strong>the</strong> ROS efflux in M. truncatula <strong>during</strong> <strong>the</strong> very early steps<br />

of <strong>the</strong> interaction, 1 h post inoculation (Shaw and Long,<br />

2003). This down-regulation of early ROS accumulation is<br />

not observed in plants deficient for genes encoding proteins<br />

involved in <strong>the</strong> symbiotic interaction such as a protein kinase<br />

(CDPK1) or a putative NFs receptor NFP (Ivashuta et al., 2005),<br />

suggesting that it is linked to <strong>the</strong> NFs signaling transduction<br />

pathway. This reduction in ROS efflux was correlated with<br />

a transient decrease of transcript accumulation of two NADPH<br />

oxidase homologs, MtRBOH2 and MtRBOH3, 1 h after NFs<br />

treatment (Lohar et al., 2007). This transient down-regulation<br />

of both MtRBOH is not observed in <strong>the</strong> mutant nfp showing<br />

that <strong>the</strong> early stage of NFs perception is required for <strong>the</strong> regulation<br />

of MtRBOH2 and MtRBOH3. These results suggest that<br />

NFs suppress <strong>the</strong> very early activity of <strong>the</strong> ROS-generating system<br />

<strong>during</strong> <strong>the</strong> interaction and, thus, may allow a compatible<br />

interaction between <strong>the</strong> plant and <strong>the</strong> bacteria. Moreover, a S.<br />

meliloti strain deficient in <strong>the</strong> production of NFs and in <strong>the</strong> infection<br />

process induces H2O2 accumulation <strong>during</strong> <strong>the</strong> first<br />

hours post inoculation as compared with <strong>the</strong> wild-type S. meliloti<br />

(Bueno et al., 2001). Fur<strong>the</strong>rmore, treatment of a M. sativa<br />

suspension culture with yeast elicitors and S. meliloti LPS is unable<br />

to induce <strong>the</strong> ROS formation observed when <strong>the</strong> cells are<br />

treated with yeast elicitors alone, showing that LPS might<br />

modulate ROS production <strong>during</strong> <strong>the</strong> interaction (Albus<br />

et al., 2001). Finally, ano<strong>the</strong>r S. meliloti mutant strain, which<br />

can not produce exopolysaccharides, is defective in invading<br />

<strong>the</strong> plant and triggers <strong>the</strong> induction of a larger number of<br />

genes belonging to <strong>the</strong> plant defense category in M. truncatula<br />

(Jones et al., 2008). Taken toge<strong>the</strong>r, <strong>the</strong>se results show that<br />

<strong>the</strong> absence of NFs or exopolysaccharides impairs <strong>the</strong> infection<br />

process, perhaps via a plant defense response that may be<br />

linked to ROS production. In conclusion, <strong>the</strong> differential<br />

ROS accumulation observed at different time points of <strong>the</strong><br />

symbiotic interaction suggests that <strong>the</strong> ROS level may be involved<br />

in different processes at multiple spatiotemporal steps<br />

<strong>during</strong> nodule formation.<br />

The production of H2O2 <strong>during</strong> symbiosis was detected in<br />

infection threads, in <strong>the</strong> thread walls of both <strong>the</strong> infection<br />

and fixing zones of M. sativa and Pisum sativum nodules (Santos<br />

et al., 2001; Rubio et al., 2004). Hydrogen peroxide is also<br />

detected in <strong>the</strong> cell walls and intercellular spaces of <strong>the</strong> cortex.<br />

This production of H2O2 may be involved in <strong>the</strong> oxidative crosslinking<br />

needed for streng<strong>the</strong>ning of <strong>the</strong> infection threads and<br />

cell wall formation. In contrast, H2O2 was not detected in <strong>the</strong><br />

nitrogen-fixing zone of <strong>the</strong> nodules (Santos et al., 2001; Rubio<br />

et al., 2004). This lack of detection may be attributed to <strong>the</strong><br />

strong antioxidant defense (see antioxidant defense in <strong>the</strong><br />

next section) or to its fast reaction with biological compounds<br />

present in <strong>the</strong> nodule. Leghaemoglobin is a candidate for oxidative<br />

damage.<br />

Leghaemoglobin is present at a very high concentration in<br />

nodules where it facilitates oxygen transport to <strong>the</strong> bacteroids.<br />

Exposure of young or mature nodules to oxidative stress results<br />

in <strong>the</strong> formation of leghaemoglobin radicals that are similar to<br />

those observed <strong>during</strong> natural senescence (Mathieu et al.,<br />

1998). Moreover, when linked to O2, this hemoprotein undergoes<br />

facile autoxidation to form ROS such as superoxide radicals<br />

or H2O2 (Puppo et al., 1981). This ROS formation is<br />

important, as plants deficient in <strong>the</strong> formation of leghemoglobin<br />

contain a significantly lower level of H2O2 (Gun<strong>the</strong>r et al.,<br />

2007). Multiple results suggest that redox balance is involved<br />

in <strong>the</strong> regulation of nodule metabolism. The diminution of nitrogen<br />

fixation under abiotic stress is correlated with a modification<br />

of <strong>the</strong> redox balance and a strong decline in <strong>the</strong><br />

antioxidant defense (Escurado et al., 1996; Gogorcena et al.,<br />

1997; Jebara et al., 2005). Moreover, this decrease in nitrogen<br />

fixation is associated with a carbon flux shortage (Matamoros<br />

et al., 1999a; Galvez et al., 2005). Sucrose synthase, which plays<br />

a major role in carbon metabolism regulation and is crucial for<br />

an adequate nitrogen fixation (Baier et al., 2007), has been<br />

shown to be regulated by redox state in nodules (Marino<br />

et al., 2008). Interestingly, modification of nodule metabolism<br />

under senescence could be partially mimicked by a direct oxidative<br />

stress on mature nodules (Mathieu et al., 1998; Marino<br />

et al., 2006).<br />

During nodule senescence, ROS have been detected around<br />

senescing symbiosomes, suggesting <strong>the</strong>ir involvement in this<br />

process (Alesandrini et al., 2003; Rubio et al., 2004). At <strong>the</strong> molecular<br />

level, a transcriptomic analysis performed with M. truncatula<br />

nodules showed that <strong>the</strong> transcriptomes of nodule and<br />

leaf senescence have a high degree of overlap, arguing for <strong>the</strong><br />

recruitment of similar pathways (Van de Velde et al., 2006).<br />

Thus, a combination of multiple factors including ROS may<br />

be involved in <strong>the</strong> control of <strong>the</strong> events leading to <strong>the</strong> rupture<br />

of <strong>the</strong> symbiotic interaction (Puppo et al., 2005).<br />

INVOLVEMENT OF ANTIOXIDANT<br />

DEFENSE IN NITROGEN FIXING<br />

SYMBIOSIS<br />

Plant cells contain an impressive antioxidant defense with enzymatic<br />

(mainly water soluble) and non-enzymatic (water and<br />

lipid soluble) ROS scavenging systems. All types of antioxidants<br />

described in plants have also been detected in nodules. This<br />

includes: (1) enzymatic systems directly involved in ROS<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


scavenging such as superoxide dismutase (SOD), catalase<br />

(KAT), and peroxidase (Prx); (2) <strong>the</strong> non-enzymatic antioxidant<br />

molecules such as ascorbate (ASC), glutathione (GSH), and<br />

a-tocopherol; (3) <strong>the</strong> ascorbate–glutathione pathway, which<br />

allows <strong>the</strong> reduction of <strong>the</strong> two antioxidant molecules by<br />

NAD(P)H; and (4) <strong>the</strong> enzymes involved in <strong>the</strong> disulfide reduction,<br />

thioredoxin (Trx) and glutaredoxin (Grx). The bacterial<br />

partner that differentiates to bacteroid to reduce atmospheric<br />

nitrogen also contains a large range of antioxidant systems including<br />

enzymatic and non-enzymatic defense. This section<br />

will focus on <strong>the</strong> involvement of <strong>the</strong> antioxidant systems in<br />

<strong>the</strong> following facets of nitrogen-fixing symbiosis: nodule formation,<br />

nitrogen fixing efficiency, and alteration of this efficiency<br />

under stress conditions.<br />

Antioxidant systems are known to be involved in <strong>the</strong> regulation<br />

of developmental processes because mutants in various<br />

antioxidant systems showed growth defect and organ malformations<br />

(for reviews, see Beveridge et al., 2007; Covarrubias<br />

et al., 2008). In this framework, <strong>the</strong> redox transduction pathway<br />

has been shown to play an important role in nodule development.<br />

Sinorhizobium meliloti genome-wide transcriptome<br />

experiments showed that sodC specifically is induced <strong>during</strong><br />

Medicago infection (Ampe et al., 2003). Bacterial mutants deficient<br />

in catalase activity are impaired in <strong>the</strong>ir symbiotic efficiency<br />

(Sigaud et al., 1999; Jamet et al., 2003). S. meliloti contains three<br />

catalase genes (katA, katB, andkatC), which are differentially<br />

expressed <strong>during</strong> <strong>the</strong> various physiological states of <strong>the</strong> bacteria.<br />

Whereas <strong>the</strong> symbiotic efficiency of single mutant strains was<br />

not affected, <strong>the</strong> double mutant katA katC nodulated poorly,<br />

and <strong>the</strong> double mutant katB katC displayed abnormal infection<br />

with a failure in bacteroid differentiation. Never<strong>the</strong>less, recent<br />

results obtained <strong>during</strong> <strong>the</strong> symbiotic interaction between Medicago<br />

sativa and an overexpressing S. meliloti RmkatB ++ mutant<br />

indicate that H2O2 is required for optimal progression of infection<br />

threads through <strong>the</strong> root hairs and plant cell layers (Jamet<br />

et al., 2007).<br />

Antioxidant molecules have also been shown to be involved<br />

in symbiosis efficiency. In M. truncatula, inhibition of GSH and<br />

homoglutathione (hGSH) syn<strong>the</strong>sis (a legume-specific GSH homolog)<br />

by pharmacological or genetic approaches inhibited<br />

root nodule formation (Frendo et al., 2005). A strong reduction<br />

in <strong>the</strong> number of nascent nodules and in <strong>the</strong> expression<br />

of <strong>the</strong> early nodulin genes, involved in nodule formation, was<br />

observed in GSH and hGSH-depleted plants. In parallel, <strong>the</strong> importance<br />

of GSH <strong>during</strong> symbiosis of S. meliloti and Rhizobium<br />

tropici has also been demonstrated (Harrison et al., 2005;<br />

Muglia et al., 2008). The S. meliloti and R. tropici strains deficient<br />

in gshB, <strong>the</strong> gene encoding <strong>the</strong> enzyme catalyzing <strong>the</strong><br />

second step of GSH syn<strong>the</strong>sis, were affected in <strong>the</strong>ir symbiotic<br />

efficiencies with a delayed nodulation coupled with a reduction<br />

in <strong>the</strong> nitrogen-fixation capacity. This phenotype was linked to<br />

abnormal nodule development with early nodule senescence.<br />

There are several indications that Trxs also play an important<br />

role in <strong>the</strong> establishment of <strong>the</strong> nitrogen-fixing symbiotic process.<br />

Recently, a detailed analysis of Trxs in M. truncatula estab-<br />

Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong> | 373<br />

lished <strong>the</strong> existence of two isoforms that do not belong to any<br />

of <strong>the</strong> previously described types (Alkhalfioui et al., 2008). Interestingly,<br />

<strong>the</strong>se novel isoforms seem to be specifically dedicated<br />

to <strong>the</strong> symbiotic interaction according to <strong>the</strong>ir pattern of expression.<br />

In Glycine max, a Trx was found to be involved in nodulation,<br />

as plants deficient in this isoform have a reduced number<br />

of nodules (Lee et al., 2005). The bacterial partner, a Bradyrhizobium<br />

japonicum mutant deficient in tlpA (a gene encoding<br />

protein having a Trx domain), was notably defective in <strong>the</strong> development<br />

of a nitrogen-fixing endosymbiosis and exhibited<br />

a strong decrease in oxidase activity compared to <strong>the</strong> wild-type<br />

(Loferer et al., 1993). All <strong>the</strong>se data emphasize <strong>the</strong> role of <strong>the</strong><br />

antioxidant defense and redox state regulation in <strong>the</strong> establishment<br />

of an efficient symbiosis between legumes and rhizobia.<br />

In <strong>the</strong> functional nodule, numerous processes may contribute<br />

to a high production of ROS. The highly reducing environment<br />

needed for nitrogen fixation may lead to ROS formation,<br />

as many electron donors, including ferredoxin, uricase, and hydrogenase,<br />

are susceptible to auto-oxidation, resulting in superoxide<br />

formation (Dalton et al., 1991). Oxyleghemoglobin,<br />

which is present at a high concentration in nodules, can also<br />

produce superoxide (Puppo et al., 1991). In addition, <strong>the</strong> high<br />

level of transition metals in <strong>the</strong> nodule may lead to <strong>the</strong> generation<br />

of ROS (Becana and Klucas, 1992). In this context,<br />

a strong antioxidant defense may be crucial in order to allow<br />

efficient nodule metabolism. As mentioned previously, H2O2 is<br />

not detected in <strong>the</strong> nitrogen-fixing zone of nodules. It has<br />

been proposed that antioxidant defense participates in <strong>the</strong><br />

limitation of <strong>the</strong> entry of O 2 into <strong>the</strong> nodule interior (Dalton<br />

et al., 1998). Expression of ascorbate peroxidase (APX) is highly<br />

induced in soybean nodules in which APX represents about 1%<br />

of <strong>the</strong> protein (Dalton et al., 1987). The preferential localization<br />

of <strong>the</strong> APX is <strong>the</strong> endodermis of indeterminate nodules in<br />

alfalfa, pea, and clover and <strong>the</strong> exterior edge of <strong>the</strong> parenchyma<br />

of determinate nodules in bean and soybean. These<br />

data suggest that this enzyme may participate in <strong>the</strong> nodule<br />

oxygen barrier through <strong>the</strong> detoxication of ROS produced<br />

by high rates of respiration detected in <strong>the</strong> endodermis (indeterminate<br />

nodules) and in boundary layers (determinate nodules)<br />

(Dalton et al., 1998).<br />

In contrast to <strong>the</strong> preferential localization of APX in <strong>the</strong> peripheral<br />

layer of <strong>the</strong> nodule, <strong>the</strong> GSH and its legume-specific<br />

homolog hGSH are mainly present in <strong>the</strong> infected zone of<br />

<strong>the</strong> determinate bean nodule (Matamoros et al., 1999b). High<br />

content of <strong>the</strong>se two thiols also occurs in <strong>the</strong> meristematic<br />

zone of <strong>the</strong> nodule. Moreover, <strong>the</strong> ascorbate–glutathione<br />

pathway was more potent in effective nodules than in ineffective<br />

nodules, accentuating <strong>the</strong> important protective role that<br />

this pathway may play for nitrogen fixation in soybean and<br />

alfalfa (Dalton et al., 1993). Comparable results were also observed<br />

in <strong>the</strong> leghemoglobin-RNA interference L. japonicus<br />

lines, which are altered in <strong>the</strong>ir nitrogen-fixation activity<br />

(Ott et al., 2005; Gun<strong>the</strong>r et al., 2007).<br />

The nitrogen-fixing deficiency of S. meliloti strains mutated<br />

in genes belonging to redox homeostasis also indicates its<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


374 | Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong><br />

crucial role in nitrogen-fixation metabolism. Bacterial mutants<br />

affected in <strong>the</strong> antioxidant defense (catalase or glutathione)<br />

were not only affected in <strong>the</strong>ir capacity to infect <strong>the</strong> plant,<br />

but also in <strong>the</strong>ir nitrogen-fixation efficiency (Jamet et al.,<br />

2003; Harrison et al., 2005). Similarly, a peroxiredoxin (prxS)/bifunctional<br />

catalase–peroxidase (katG) Rhizobium etli double<br />

mutant has a significantly reduced nitrogen fixation capacity<br />

(Dombrecht et al., 2005). Additionally, S. meliloti strains mutated<br />

for a wide variety of cellular processes but not directly<br />

involved in antioxidant defense were also found to be affected<br />

in oxidative stress protection and symbiosis (Davies and<br />

Walker, 2007). Likewise, mutation of a thioredoxin-like gene<br />

involved in melanin production affected <strong>the</strong> response to paraquat-induced<br />

oxidative stress and <strong>the</strong> symbiotic nitrogen fixation<br />

in S. meliloti (Castro-Sowinski et al., 2007). Finally,<br />

a Rhizobium leguminosarum deficient in <strong>the</strong> thioredoxin-like<br />

cycY was unable to form nitrogen-fixing nodules on pea or<br />

vetch, as <strong>the</strong> mutant was defective in maturation of all c-type<br />

cytochromes (Vargas et al., 1994).<br />

The alteration of <strong>the</strong> nitrogen-fixing efficiency <strong>during</strong> senescence<br />

correlates with a general diminution of <strong>the</strong> antioxidant<br />

defense and to <strong>the</strong> detection of ROS species in <strong>the</strong><br />

nodule-senescent zone. This correlation is established <strong>during</strong><br />

both natural and stress-induced senescence. In this framework,<br />

diminution of GSH/hGSH and ASC pools was concomitant with<br />

a decrease in nitrogen-fixing efficiency <strong>during</strong> natural senescence<br />

in a large number of legumes such as soybean (Evans<br />

et al., 1999), pea (Groten et al., 2006), or common bean (Loscos<br />

et al., 2008). A similar diminution of antioxidant molecules and<br />

enzymes was observed in legumes submitted to various environmental<br />

stresses (Moran et al., 1994; Escurado et al., 1996;<br />

Gogorcena et al., 1997; Matamoros et al., 1999b; Loscos<br />

et al., 2008). The modulation of <strong>the</strong> redox balance <strong>during</strong> cell<br />

death and senescence is a general process in plants (Gechev<br />

et al., 2006; Noctor et al., 2007; Shao et al., 2008). In this framework,<br />

as mentioned previously, nodule senescence has a high<br />

degree of overlap with leaf senescence.<br />

One of <strong>the</strong> specific features of nodules is low oxygen pressure<br />

in <strong>the</strong> fixing zone to permit efficient nitrogen fixation.<br />

The regulation of respiration and <strong>the</strong> efficiency of <strong>the</strong> oxygen<br />

diffusion barrier play a crucial role in <strong>the</strong> protection of nodule<br />

metabolism. This regulation seems to be an active plantcontrolled<br />

process (Wei and Layzell, 2006). Stress-induced decrease<br />

of nitrogen-fixing activity is generally associated with<br />

a modification of <strong>the</strong> oxygen pressure in <strong>the</strong> nodule. In soybean,<br />

<strong>the</strong> drought and salt stresses caused a decrease in nitrogenase<br />

activity correlating with a diminution of nodule<br />

permeability (Del Castillo et al., 1994; Serraj et al., 1994). Similarly,<br />

soybean nodules presented modified respiration and<br />

intercellular space size <strong>during</strong> chilling, showing an altered<br />

oxygen diffusion barrier (van Heerden et al., 2008). In contrast,<br />

<strong>the</strong> sensitivity to salinity appears to be associated with an increase<br />

in nodule conductance and an increased respiration of<br />

nodules in M. truncatula (Aydi et al., 2004). Taken toge<strong>the</strong>r,<br />

<strong>the</strong>se results show that nodule efficiency is strongly affected<br />

by <strong>the</strong> regulation of oxygen availability that may be linked<br />

to ROS production.<br />

Ano<strong>the</strong>r specific feature is <strong>the</strong> high carbon/nitrogen metabolism<br />

needed to feed <strong>the</strong> nitrogen-fixing bacteroids and to allow<br />

nitrogen assimilation by <strong>the</strong> plant cell. Transcriptomic<br />

approaches suggest that several metabolic pathways are coordinately<br />

up-regulated in nodules, including glycolysis, amino<br />

acid biosyn<strong>the</strong>sis, haem, and redox metabolism (Colebatch<br />

et al., 2004). Glycolysis plays a primary role in feeding <strong>the</strong> bacteroids;<br />

environmental stresses such as salt stress, drought<br />

stress, and dark stress generally down-regulate <strong>the</strong> enzymes<br />

involved in supplying carbon to <strong>the</strong> bacteroids (Gogorcena<br />

et al., 1997; Galvez et al., 2005; Lopez et al., 2008). Interestingly,<br />

mild oxidative stress using paraquat treatment mimics<br />

biological modifications observed under environmental stress,<br />

indicating a likely involvement of redox modifications in <strong>the</strong><br />

perception of environmental stress in pea nodules (Marino<br />

et al., 2006). In pea, this regulation of carbon metabolism<br />

by <strong>the</strong> cellular redox state has been recently linked to <strong>the</strong> transcriptional<br />

and post-translational regulation of sucrose synthase,<br />

<strong>the</strong> enzyme allowing sucrose assimilation in <strong>the</strong><br />

nodule (Marino et al., 2008). These data underline <strong>the</strong> involvement<br />

of redox regulation in nodule metabolism.<br />

PERSPECTIVES<br />

The data summarized in this review clearly indicate that ROS<br />

and antioxidant defense play a crucial role in symbiosis between<br />

legumes and rhizobia. The involvement of ROS and antioxidant<br />

defense in key steps of <strong>the</strong> nodule formation, such as infection<br />

thread development and nodule meristem formation, shows<br />

that redox regulation is important for <strong>the</strong> nodule development.<br />

Moreover, several bacterial strains deficient in antioxidant defense<br />

exhibit an altered nitrogen-fixation capacity without<br />

strong modifications of <strong>the</strong>ir growth efficiency in free-living<br />

conditions, suggesting that nodule environment is more stressful<br />

for <strong>the</strong> bacteria. Finally, <strong>the</strong> modifications of <strong>the</strong> plant antioxidant<br />

defense is observed in parallel with metabolic changes<br />

<strong>during</strong> nodule functioning. Thus, <strong>the</strong> redox homeostasis may<br />

play a crucial role in <strong>the</strong> control of nodule metabolism by<br />

<strong>the</strong> plant. However, <strong>the</strong> fine characterization of ROS production<br />

and modification of <strong>the</strong> redox state <strong>during</strong> symbiosis still needs<br />

to be achieved. Specific protein probes sensitive to redox potential<br />

(Meyer et al., 2007) or to ROS accumulation (Belousov et al.,<br />

2006) would be useful tools for realizing <strong>the</strong> spatiotemporal<br />

analysis of H 2O 2 production and redox changes. Fur<strong>the</strong>rmore,<br />

taking into account that plastids are involved in GSH syn<strong>the</strong>sis<br />

(Pasternak et al., 2008) and in retrograde signaling (Nott et al.,<br />

2006), it would be worth addressing <strong>the</strong> role of <strong>the</strong>se organelles<br />

in redox homeostasis. The development of two legume model<br />

systems, M. truncatula (www.medicago.org/) and L. japonicus<br />

(www.lotusjaponicus.org/) will facilitate an efficient analysis<br />

of <strong>the</strong> redox-dependent metabolism.<br />

Legume–rhizobia symbiosis is an interesting model for<br />

studying compatible plant–microorganism interactions and<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


plant development processes. Thus, information obtained<br />

with this model will be of general interest for plant biologists.<br />

From a more pragmatic point of view, legumes are unique, as<br />

<strong>the</strong>y do not need nitrogen-fertilizers (which need fossil energy<br />

for <strong>the</strong>ir production) and present high protein content. Thus,<br />

legumes have a crucial role to play in <strong>the</strong> development of more<br />

sustainable and multifunctional agro-ecosystems.<br />

FUNDING<br />

This work is funded by CNRS, INRA and <strong>the</strong> University of Nice-<br />

Sophia Antipolis. The Agence Nationale de la Recherche is supporting<br />

this work through <strong>the</strong> RHYTHMS project (contract number:<br />

BLAN07-2_182872).<br />

ACKNOWLEDGMENTS<br />

We gratefully acknowledge Julie Hopkins, Erica Stein, and Ryan<br />

Gutierrez for critical reading of <strong>the</strong> manuscript. C.C. is <strong>the</strong> recipient<br />

of a post-doctoral fellowship from <strong>the</strong> ‘Ministère de l’Enseignement<br />

Supérieur et de la Recherche’. I.D. is <strong>the</strong> recipient of<br />

a post-doctoral fellowship from <strong>the</strong> CNRS. We apologize to our colleagues<br />

whose works were not cited here. No conflict of interest<br />

declared.<br />

REFERENCES<br />

Albus, U., Baier, R., Holst, O., Pühler, A., and Niehaus, K. (2001).<br />

Suppression of an elicitor-induced oxidative burst reaction in<br />

Medicago sativa cell cultures by Sinorhizobium meliloti<br />

lipopolysaccharides. New Phytologist. 151, 597–606.<br />

Alesandrini, F., Mathis, R., Sype, G.V.d, Hérouart, D., and Puppo, A.<br />

(2003). Possible roles for a cysteine protease and hydrogen peroxide<br />

in soybean nodule development and senescence. New Phytologist.<br />

158, 131–138.<br />

Alkhalfioui, F., Renard, M., Frendo, P., Keichinger, C., Meyer, Y.,<br />

Gelhaye, E., Hirasawa, M., Knaff, D.B., Ritzenthaler, C., and<br />

Montrichard, F. (2008). A novel type of thioredoxin dedicated<br />

to symbiosis in legumes. Plant Physiol. 148, 424–435.<br />

Ampe, F., Kiss, E., Sabourdy, F., and Batut, J. (2003). Transcriptome<br />

analysis of Sinorhizobium meliloti <strong>during</strong> symbiosis. Genome<br />

Biol. 4, R15.<br />

Apel, K., and Hirt, H. (2004). Reactive oxygen species: metabolism,<br />

oxidative stress, and signal transduction. Annu. Rev. Plant Biol.<br />

55, 373–399.<br />

Aydi, S., Drevon, J.J., and Abdelly, C. (2004). Effect of salinity on<br />

root–nodule conductance to <strong>the</strong> oxygen diffusion in <strong>the</strong> Medicago<br />

truncatula–Sinorhizobium meliloti symbiosis. Plant Physiol.<br />

Biochem. 42, 833–840.<br />

Baier, M.C., Barsch, A., Kuster, H., and Hohnjec, N. (2007). Antisense<br />

repression of <strong>the</strong> Medicago truncatula nodule-enhanced sucrose<br />

synthase leads to a handicapped nitrogen fixation mirrored by<br />

specific alterations in <strong>the</strong> symbiotic transcriptome and metabolome.<br />

Plant Physiol. 145, 1600–1618.<br />

Becana, M., and Klucas, R.V. (1992). Transition metals in legume root<br />

nodules. Iron dependant free radical production increases <strong>during</strong><br />

nodule senescence. Proc. Natl Acad. Sci. U S A. 87, 7295–7299.<br />

Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong> | 375<br />

Belousov, V.V., Fradkov, A.F., Lukyanov, K.A., Staroverov, D.B.,<br />

Shakhbazov, K.S., Terskikh, A.V., and Lukyanov, S. (2006). Genetically<br />

encoded fluorescent indicator for intracellular hydrogen<br />

peroxide. Nat. Methods. 3, 281–286.<br />

Beveridge, C.A., Ma<strong>the</strong>sius, U., Rose, R.J., and Gresshoff, P.M.<br />

(2007). Common regulatory <strong>the</strong>mes in meristem development<br />

and whole-plant homeostasis. Curr. Opin. Plant Biol. 10, 44–51.<br />

Brewin, N.J. (2004). Plant cell wall remodelling in <strong>the</strong> rhizobium–<br />

legume symbiosis. Critical Reviews in Plant Sciences. 23, 293–316.<br />

Bueno, P., Soto, M.J., Rodriguez-Rosales, M.P., Sanjuan, J.,<br />

Olivares, J., and Donaire, J.P. (2001). Time-course of lipoxygenase,<br />

antioxidant enzyme activities and H2O2 accumulation <strong>during</strong><br />

early stages of Rhizobium legume symbiosis. New Phytol.<br />

152, 91–96.<br />

Cardenas, L., Martinez, A., Sanchez, F., and Quinto, C. (2008). Fast,<br />

transient and specific intracellular ROS changes in living root hair<br />

cells responding to Nod factors (NFs). Plant J. 27, in press.<br />

Carol, R.J., and Dolan, L. (2006). The role of reactive oxygen species<br />

in cell growth: lessons from root hairs. J. Exp. Bot. 57, 1829–1834.<br />

Castro-Sowinski, S., Matan, O., Bonafede, P., and Okon, Y. (2007). A<br />

thioredoxin of Sinorhizobium meliloti CE52G is required for melanin<br />

production and symbiotic nitrogen fixation. Mol. Plant Microbe<br />

Interact. 20, 986–993.<br />

Colebatch, G., Desbrosses, G., Ott, T., Krusell, L., Montanari, O.,<br />

Kloska, S., Kopka, J., and Udvardi, M.K. (2004). Global changes<br />

in transcription orchestrate metabolic differentiation <strong>during</strong> symbiotic<br />

nitrogen fixation in Lotus japonicus. PlantJ.39, 487–512.<br />

Cook, D., Dreyer, D., Bonnet, D., Howell, M., Nony, E., and<br />

Vandenbosh, K. (1995). Transient induction of a peroxidase gene<br />

in Medicago truncatula precedes infection by Rhizobium meliloti.<br />

Plant Cell. 7, 43–55.<br />

Covarrubias, L., Hernandez-Garcia, D., Schnabel, D., Salas-Vidal, E.,<br />

and Castro-Obregon, S. (2008). Function of reactive oxygen species<br />

<strong>during</strong> animal development: passive or active? Dev. Biol. 320,<br />

1–11.<br />

Dalton, D.A., Hanus, F.J., Russell, S.A., and Evans, H.J. (1987). Purification,<br />

properties, and distribution of ascorbate peroxidase in<br />

legume root nodules. Plant Physiol. 83, 789–794.<br />

Dalton, D.A., Joyner, S.L., Becana, M., Iturbe-Ormaetxe, I., and<br />

Chatfield, J.M. (1998). Antioxidant defenses in <strong>the</strong> peripheral cell<br />

layers of legume root nodules. Plant Physiol. 116, 37–43.<br />

Dalton, D.A., Langeberg, L., and Treneman, N.C. (1993). Correlations<br />

between <strong>the</strong> ascorbate–glutathione pathway and effectiveness<br />

in legume root nodules. Physiol. Plant. 84, 365–370.<br />

Dalton, D.A., Post, C.J., and Langeberg, L. (1991). Effects of ambient<br />

oxygen and of fixed nitrogen on concentrations of glutathione,<br />

ascorbate, and associated enzymes in soybean root nodules.<br />

Plant Physiol. 96, 812–818.<br />

Davies, B.W., and Walker, G.C. (2007). Identification of novel Sinorhizobium<br />

meliloti mutants compromised for oxidative stress<br />

protection and symbiosis. J. Bacteriol. 189, 2110–2113.<br />

Del Castillo, L.D., Hunt, S., and Layzell, D.B. (1994). The role of oxygen<br />

in <strong>the</strong> regulation of nitrogenase activity in droughtstressed<br />

soybean nodules. Plant Physiol. 106, 949–955.<br />

Den Herder, J., Lievens, S., Rombauts, S., Holsters, M., and<br />

Goormachtig, S. (2007). A symbiotic plant peroxidase involved<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


376 | Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong><br />

in bacterial invasion of <strong>the</strong> tropical legume Sesbania rostrata.<br />

Plant Physiol. 144, 717–727.<br />

Dénarié, J., and Cullimore, J. (1993). Lipo-oligosaccharide nodulation<br />

factors: a new class of signalling molecules mediating recognition<br />

and morphogenesis. Cell. 74, 951–954.<br />

D’Haeze, W., De Rycke, R., Mathis, R., Goormachtig, S., Pagnotta, S.,<br />

Verplancke, C., Capoen, W., and Holsters, M. (2003). Reactive oxygen<br />

species and ethylene play a positive role in lateral root base<br />

nodulation of a semiaquatic legume. Proc. Natl Acad. Sci. U S A.<br />

100, 11789–11794.<br />

Dombrecht, B., Heusdens, C., Beullens, S., Verreth, C., Mulkers, E.,<br />

Proost, P., Vanderleyden, J., and Michiels, J. (2005). Defence of<br />

Rhizobium etli bacteroids against oxidative stress involves a complexly<br />

regulated atypical 2-Cys peroxiredoxin. Mol. Microbiol.<br />

55, 1207–1221.<br />

Downie, J.A. (2005). Legume haemoglobins: symbiotic nitrogen fixation<br />

needs bloody nodules. Curr. Biol. 15, R196–R198.<br />

Escurado, P.R., Minchin, F.R., Gogoncena, Y., Iturbe-Ormaetxe, I.,<br />

Klucas, R.V., and Becana, M. (1996). Involved of activated oxygen<br />

in nitrate induced senescence of pea root nodules. Plant Physiol.<br />

110, 1187–1195.<br />

Evans, P.J., Gallesi, D., Mathieu, C., Hernandez, M.J., de Felipe, M.,<br />

Halliwell, B., and Puppo, A. (1999). Oxidative stress occurs <strong>during</strong><br />

soybean nodule senescence. Planta. 208, 73–79.<br />

Fischer, H.M. (1994). Genetic regulation of nitrogen fixation in rhizobia.<br />

Microbiol. Rev. 58, 352–386.<br />

Frendo, P., Harrison, J., Norman, C., Hernandez Jimenez, M.J., Van<br />

de Sype, G., Gilabert, A., and Puppo, A. (2005). Glutathione and<br />

homoglutathione play a critical role in <strong>the</strong> nodulation process of<br />

Medicago truncatula. Mol. Plant Microbe Interact. 18, 254–259.<br />

Galvez, L., Gonzalez, E.M., and Arrese-Igor, C. (2005). Evidence for carbon<br />

flux shortage and strong carbon/nitrogen interactions in pea<br />

nodules at early stages of water stress. J. Exp. Bot. 56, 2551–2561.<br />

Gechev, T.S., Van Breusegem, F., Stone, J.M., Denev, I., and Laloi, C.<br />

(2006). Reactive oxygen species as signals that modulate plant stress<br />

responses and programmed cell death. Bioessays. 28, 1091–1101.<br />

Gogorcena, Y., Gordon, A.J., Escuredo, P.R., Minchin, F.R., Witty, J.F.,<br />

Moran, J.F., and Becana, M. (1997). N2 fixation, carbon metabolism<br />

and oxidative damage in nodules of dark stressed Common<br />

Bean plants. Plant Physiol. 113, 1193–1201.<br />

Goormachtig, S., Capoen, W., and Holsters, M. (2004). Rhizobium<br />

infection: lessons from <strong>the</strong> versatile nodulation behaviour of water-tolerant<br />

legumes. Trends Plant Sci. 9, 518–522.<br />

Groten, K., Dutilleul, C., van Heerden, P.D., Vanacker, H., Bernard, S.,<br />

Finkemeier, I., Dietz, K.J., and Foyer, C.H. (2006). <strong>Redox</strong> regulation<br />

of peroxiredoxin and proteinases by ascorbate and thiols <strong>during</strong><br />

pea root nodule senescence. FEBS Lett. 580, 1269–1276.<br />

Gun<strong>the</strong>r, C., Schlereth, A., Udvardi, M., and Ott, T. (2007). Metabolism<br />

of reactive oxygen species is attenuated in leghemoglobindeficient<br />

nodules of Lotus japonicus. Mol. Plant Microbe Interact.<br />

20, 1596–1603.<br />

Harrison, J., Jamet, A., Muglia, C.I., Van de Sype, G., Aguilar, O.M.,<br />

Puppo, A., and Frendo, P. (2005). Glutathione plays a fundamental<br />

role in growth and symbiotic capacity of Sinorhizobium meliloti.<br />

J. Bacteriol. 187, 168–174.<br />

Ivashuta, S., Liu, J., Liu, J., Lohar, D.P., Haridas, S., Bucciarelli, B.,<br />

VandenBosch, K.A., Vance, C.P., Harrison, M.J., and Gantt, J.S.<br />

(2005). RNA interference identifies a calcium-dependent protein<br />

kinase involved in Medicago truncatula root development. Plant<br />

Cell. 17, 2911–2921.<br />

Jamet, A., Mandon, K., Puppo, A., and Herouart, D. (2007). H2O2 is<br />

required for optimal establishment of <strong>the</strong> Medicago sativa/Sinorhizobium<br />

meliloti symbiosis. J. Bacteriol. 189, 8741–8745.<br />

Jamet, A., Sigaud, S., Van de Sype, G., Puppo, A., and Herouart, D.<br />

(2003). Expression of <strong>the</strong> bacterial catalase genes <strong>during</strong> Sinorhizobium<br />

meliloti–Medicago sativa symbiosis and <strong>the</strong>ir crucial role<br />

<strong>during</strong> <strong>the</strong> infection process. Mol. Plant Microbe Interact. 16,<br />

217–225.<br />

Jebara, S., Jebara, M., Limam, F., and Aouani, M.E. (2005). <strong>Changes</strong><br />

in ascorbate peroxidase, catalase, guaiacol peroxidase and superoxide<br />

dismutase activities in common bean (Phaseolus vulgaris)<br />

nodules under salt stress. J. Plant Physiol. 162, 929–936.<br />

Jones, K.M., Kobayashi, H., Davies, B.W., Taga, M.E., and<br />

Walker, G.C. (2007). How rhizobial symbionts invade plants:<br />

<strong>the</strong> Sinorhizobium–Medicago model. Nat. Rev. Microbiol. 5,<br />

619–633.<br />

Jones, K.M., Sharopova, N., Lohar, D.P., Zhang, J.Q.,<br />

VandenBosch, K.A., and Walker, G.C. (2008). Differential response<br />

of <strong>the</strong> plant Medicago truncatula to its symbiont Sinorhizobium<br />

meliloti or an exopolysaccharide-deficient mutant. Proc.<br />

Natl Acad. Sci. U S A. 105, 704–709.<br />

Lee, M.Y., et al. (2005). Induction of thioredoxin is required for nodule<br />

development to reduce reactive oxygen species levels in soybean<br />

roots. Plant Physiol. 139, 1881–1889.<br />

Loferer, H., Bott, M., and Hennecke, H. (1993). Bradyrhizobium<br />

japonicum TlpA, a novel membrane-anchored thioredoxin-like<br />

protein involved in <strong>the</strong> biogenesis of cytochrome aa3 and development<br />

of symbiosis. EMBO J. 12, 3373–3383.<br />

Lohar, D.P., Haridas, S., Gantt, J.S., and Vandenbosch, K.A. (2007). A<br />

transient decrease in reactive oxygen species in roots leads to<br />

root hair deformation in <strong>the</strong> legume–rhizobia symbiosis. New<br />

Phytol. 173, 39–49.<br />

Lopez, M., Herrera-Cervera, J.A., Iribarne, C., Tejera, N.A., and<br />

Lluch, C. (2008). Growth and nitrogen fixation in Lotus japonicus<br />

and Medicago truncatula under NaCl stress: nodule carbon metabolism.<br />

J. Plant Physiol. 165, 641–650.<br />

Loscos, J., Matamoros, M.A., and Becana, M. (2008). Ascorbate and<br />

homoglutathione metabolism in common bean nodules under<br />

stress conditions and <strong>during</strong> natural senescence. Plant. Physiol.<br />

146, 1282–1292.<br />

Marino, D., Gonzalez, E.M., and Arrese-Igor, C. (2006). Drought<br />

effects on carbon and nitrogen metabolism of pea nodules<br />

can be mimicked by paraquat: evidence for <strong>the</strong> occurrence of<br />

two regulation pathways under oxidative stresses. J. Exp. Bot.<br />

57, 665–673.<br />

Marino, D., Hohnjec, N., Kuster, H., Moran, J.F., Gonzalez, E.M., and<br />

Arrese-Igor, C. (2008). Evidence for transcriptional and posttranslational<br />

regulation of sucrose synthase in pea nodules by<br />

<strong>the</strong> cellular redox state. Mol. Plant Microbe Interact. 21, 622–630.<br />

Matamoros, M.A., Baird, L.M., Escuredo, P.R., Dalton, D.A.,<br />

Minchin, F.R., Iturbe-Ormaetxe, I., Rubio, M.C., Moran, J.F.,<br />

Gordon, A.J., and Becana, M. (1999a). Stress induced legume<br />

root nodule senescence. Physiological, biochemical and structural<br />

alterations. Plant Physiol. 121, 97–112.<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012


Matamoros, M.A., Moran, J.F., Iturbe-Ormaetxe, I., Rubio, M.C., and<br />

Becana, M. (1999b). Glutathione and homoglutathione syn<strong>the</strong>sis<br />

in legume root nodules. Plant Physiol. 121, 879–888.<br />

Mathieu, C., Moreau, S., Frendo, P., Puppo, A., and Davies, M.J.<br />

(1998). Direct detection of radicals in intact soybean nodules:<br />

presence of nitric oxide leghemoglobin complexes. Free Rad.<br />

Bio. Med. 24, 1242–1249.<br />

Meyer, A.J., Brach, T., Marty, L., Kreye, S., Rouhier, N., Jacquot, J.P.,<br />

and Hell, R. (2007). <strong>Redox</strong>-sensitive GFP in Arabidopsis thaliana is<br />

a quantitative biosensor for <strong>the</strong> redox potential of <strong>the</strong> cellular<br />

glutathione redox buffer. Plant J. 52, 973–986.<br />

Moller, I.M., Jensen, P.E., and Hansson, A. (2007). Oxidative modifications<br />

to cellular components in plants. Annu. Rev. Plant. Biol.<br />

58, 459–481.<br />

Moran, J.F., Becana, M., Iturbe-Ormaetxe, I., Frechilla, S.,<br />

Klucas, R.V., and Aparicio-Tejo, P. (1994). Drought induces oxidative<br />

stress in pea plants. Planta. 194, 346–352.<br />

Muglia, C., Comai, G., Spegazzini, E., Riccillo, P.M., and<br />

Aguilar, O.M. (2008). Glutathione produced by Rhizobium tropici<br />

is important to prevent early senescence in common bean<br />

nodules. FEMS Microbiol. Lett. 23, in press.<br />

Noctor, G., De Paepe, R., and Foyer, C.H. (2007). Mitochondrial redox<br />

biology and homeostasis in plants. Trends Plant Sci. 12,<br />

125–134.<br />

Nott, A., Jung, H.S., Koussevitzky, S., and Chory, J. (2006). Plastidto-nucleus<br />

retrograde signaling. Annu. Rev. Plant Biol. 57,<br />

739–759.<br />

Oldroyd, G.E., and Downie, J.A. (2008). Coordinating nodule morphogenesis<br />

with rhizobial infection in legumes. Annu. Rev. Plant<br />

Biol. 59, 519–546.<br />

Ott, T., van Dongen, J.T., Gun<strong>the</strong>r, C., Krusell, L., Desbrosses, G.,<br />

Vigeolas, H., Bock, V., Czechowski, T., Geigenberger, P., and<br />

Udvardi, M.K. (2005). Symbiotic leghemoglobins are crucial for<br />

nitrogen fixation in legume root nodules but not for general<br />

plant growth and development. Curr. Biol. 15, 531–535.<br />

Overmyer, K., Brosche, M., and Kangasjarvi, J. (2003). Reactive oxygen<br />

species and hormonal control of cell death. Trends Plant Sci.<br />

8, 335–342.<br />

Pasternak, M., Lim, B., Wirtz, M., Hell, R., Cobbett, C.S., and<br />

Meyer, A.J. (2008). Restricting glutathione biosyn<strong>the</strong>sis to <strong>the</strong> cytosol<br />

is sufficient for normal plant development. Plant J. 53,<br />

999–1012.<br />

Peck, M.C., Fisher, R.F., and Long, S.R. (2006). Diverse flavonoids<br />

stimulate NodD1 binding to nod gene promoters in Sinorhizobium<br />

meliloti. J. Bacteriol. 188, 5417–5427.<br />

Peleg-Grossman, S., Volpin, H., and Levine, A. (2007). Root hair curling<br />

and Rhizobium infection in Medicago truncatula are mediated<br />

by phosphatidylinositide-regulated endocytosis and<br />

reactive oxygen species. J. Exp. Bot. 58, 1637–1649.<br />

Pitzschke, A., Forzani, C., and Hirt, H. (2006). Reactive oxygen species<br />

signaling in plants. Antioxid. <strong>Redox</strong> Signal. 8, 1757–1764.<br />

Puppo, A., Groten, K., Bastian, F., Carzaniga, R., Soussi, M.,<br />

Lucas, M.M., de Felipe, M.R., Harrison, J., Vanacker, H., and<br />

Foyer, C.H. (2005). Legume nodule senescence: roles for redox<br />

and hormone signalling in <strong>the</strong> orchestration of <strong>the</strong> natural aging<br />

process. New Phytol. 165, 683–701.<br />

Chang et al. d <strong>Redox</strong> <strong>Changes</strong> <strong>during</strong> <strong>the</strong> <strong>Legume–Rhizobium</strong> <strong>Symbiosis</strong> | 377<br />

Puppo, A., Herrada, G., and Rigaud, J. (1991). Lipid peroxidation in<br />

peribacteroid membranes from French-bean nodules. Plant<br />

Physiol. 96, 826–830.<br />

Puppo, A., Rigaud, J., and Job, D. (1981). Role of superoxide anion<br />

in leghemoglobin autoxidation. Plant Science Let. 22, 353–360.<br />

Ramu, S.K., Peng, H.M., and Cook, D.R. (2002). Nod factor induction<br />

of reactive oxygen species production is correlated with expression<br />

of <strong>the</strong> early nodulin gene rip1 in Medicago truncatula. Mol.<br />

Plant Microbe Interact. 15, 522–528.<br />

Rivero, R.M., Kojima, M., Gepstein, A., Sakakibara, H., Mittler, R.,<br />

Gepstein, S., and Blumwald, E. (2007). Delayed leaf senescence<br />

induces extreme drought tolerance in a flowering plant. Proc.<br />

Natl Acad. Sci. U S A. 104, 19631–19636.<br />

Rubio, M.C., James, E.K., Clemente, M.R., Bucciarelli, B.,<br />

Fedorova, M., Vance, C.P., and Becana, M. (2004). Localization<br />

of superoxide dismutases and hydrogen peroxide in legume root<br />

nodules. Mol. Plant Microbe Interact. 17, 1294–1305.<br />

Santos, R., Herouart, D., Sigaud, S., Touati, D., and Puppo, A. (2001).<br />

Oxidative burst in alfalfa–Sinorhizobium meliloti symbiotic interaction.<br />

Mol. Plant Microbe Interact. 14, 86–89.<br />

Serraj, R., Roy, G., and Drevon, J.J. (1994). Salt stress induces a decrease<br />

in <strong>the</strong> oxygen uptake of soybean nodules and in <strong>the</strong>ir permeability<br />

to oxygen diffusion. Physiologia Plantarum. 91, 161–168.<br />

Shao, H.B., Chu, L.Y., Shao, M.A., Jaleel, C.A., and Mi, H.M. (2008).<br />

Higher plant antioxidants and redox signaling under environmental<br />

stresses. C.R. Biol. 331, 433–441.<br />

Shaw, S.L., and Long, S.R. (2003). Nod factor inhibition of reactive<br />

oxygen efflux in a host legume. Plant Physiol. 132, 2196–2204.<br />

Sieberer, B.J., Timmers, A.C., and Emons, A.M. (2005). Nod factors<br />

alter <strong>the</strong> microtubule cytoskeleton in Medicago truncatula root<br />

hairs to allow root hair reorientation. Mol. Plant Microbe Interact.<br />

18, 1195–1204.<br />

Sigaud, S., Becquet, V., Frendo, P., Puppo, A., and Hérouart, D.<br />

(1999). Differential regulation of two divergent Sinorhizobium<br />

meliloti genes for HPII-like catalases <strong>during</strong> free-living growth<br />

and protective role of both catalases <strong>during</strong> symbiosis. J. Bacteriol.<br />

181, 2634–2639.<br />

Van de Velde, W., Guerra, J.C., De Keyser, A., De Rycke, R.,<br />

Rombauts, S., Maunoury, N., Mergaert, P., Kondorosi, E.,<br />

Holsters, M., and Goormachtig, S. (2006). Aging in legume symbiosis.<br />

A molecular view on nodule senescence in Medicago truncatula.<br />

Plant Physiol. 141, 711–720.<br />

van Heerden, P.D., Kiddle, G., Pellny, T.K., Mokwala, P.W., Jordaan, A.,<br />

Strauss, A.J., de Beer, M., Schluter, U., Kunert, K.J., and Foyer, C.H.<br />

(2008). Roles for <strong>the</strong> regulation of respiration and <strong>the</strong> oxygen<br />

diffusion barrier in soybean in <strong>the</strong> protection of symbiotic nitrogen<br />

fixation from chilling-induced inhibition and shoots<br />

from premature senescence. Plant Physiol. 148, 316–327.<br />

Vargas, C., Wu, G., Davies, A.E., and Downie, J.A. (1994). Identification<br />

of a gene encoding a thioredoxin-like product necessary<br />

for cytochrome c biosyn<strong>the</strong>sis and symbiotic nitrogen fixation in<br />

Rhizobium leguminosarum. J. Bacteriol. 176, 4117–4123.<br />

Wei, H., and Layzell, D.B. (2006). Adenylate-coupled ion movement:<br />

a mechanism for <strong>the</strong> control of nodule permeability to O2 diffusion.<br />

Plant Physiol. 141, 280–287.<br />

Downloaded from<br />

http://mplant.oxfordjournals.org/ by guest on December 23, 2012

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!