Plants in Extreme Environments

0 downloads 0 Views 1MB Size Report
Authors are indebted to Ilse Kranner, Csaba Papdi and Laura Zsigmond for proofreading and correcting the chapter. Research and this publication were.
Provided for non-commercial research and educational use only. Not for reproduction, distribution or commercial use. This chapter was originally published in the book Advances in Botanical Research, Vol. 57, published by Elsevier, and the attached copy is provided by Elsevier for the author's benefit and for the benefit of the author's institution, for non-commercial research and educational use including without limitation use in instruction at your institution, sending it to specific colleagues who know you, and providing a copy to your institution’s administrator.

All other uses, reproduction and distribution, including without limitation commercial reprints, selling or licensing copies or access, or posting on open internet sites, your personal or institution’s website or repository, are prohibited. For exceptions, permission may be sought for such use through Elsevier's permissions site at: http://www.elsevier.com/locate/permissionusematerial From: László Szabados, Hajnalka Kovács, Aviah Zilberstein and Alain Bouchereau, Plants in Extreme Environments: Importance of Protective Compounds in Stress Tolerance. In Jean-Claude Kader and Michel Delseny, editors: Advances in Botanical Research, Vol. 57, Burlington: Academic Press, 2011, pp. 105-150. ISBN: 978-0-12-387692-8 © Copyright 2011 Elsevier Ltd. Academic Press.

Author's personal copy

Plants in Extreme Environments: Importance of Protective Compounds in Stress Tolerance

´ SZLO ´ CS,* ´ SZABADOS,*,1 HAJNALKA KOVA LA { AVIAH ZILBERSTEIN AND ALAIN BOUCHEREAU{

*Institute of Plant Biology, Biological Research Center, Szeged, Hungary { Faculty of Life Sciences, Tel Aviv University, Ramat Aviv, Tel Aviv, Israel { Universite´ de Rennes 1, Campus de Beaulieu, Baˆtiment 14A, Rennes Cedex, France

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . II. Global Metabolic Consequences of Osmotic Stress . . . . . . . . . . . . . . . . . . . . . . . . III. Osmoprotective Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A. Quaternary Amines: Glycine Betaine........................................ B. Sugars: Trehalose ............................................................... C. Polyalcohols: Mannitol, Pinitol, Inositol ................................... D. Amino Acids: Proline .......................................................... IV. Osmoprotective Compounds and Adaptation to Extreme Environments . . V. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

106 107 110 115 116 117 119 125 130 132 133

1

Corresponding author: E-mail: [email protected]

Advances in Botanical Research, Vol. 57 Copyright 2011, Elsevier Ltd. All rights reserved.

0065-2296/11 $35.00 DOI: 10.1016/B978-0-12-387692-8.00004-7

Author's personal copy 106

L. SZABADOS ET AL.

ABSTRACT Extreme environmental conditions such as drought, cold or high soil salinity impede plant growth and require specific adaptation capacity. In response to environmental stresses, a number of low-molecular-weight compounds can accumulate in plants: protective amino acids, sugar alcohols, sugars and betaine-type quaternary amines. The function of these compounds includes the stabilisation of cellular structures, photosynthetic complexes, specific enzymes and other macromolecules, the scavenging of reactive oxygen species or acting as metabolic signals in stress conditions. Although a correlation between the accumulation of certain osmoprotective compounds and stress tolerance certainly exists, a causal relationship between osmolyte accumulation and enhanced tolerance could not always be confirmed. Nevertheless, the importance of osmoprotective compounds for the adaptation to extreme environmental conditions is supported by numerous studies obtained with natural variants, mutants or transgenic plants with different capabilities to accumulate these metabolites. Combining genetic analysis with metabolic profiling approaches could considerably increase our understanding of plant stress responses and the importance of the protective metabolites in the adaptation to stress conditions.

I. INTRODUCTION Environmental conditions determine plant growth and development. Abiotic stresses such as drought, heat, cold and soil salinisation restrain the optimal growth of plants and require certain degree of adaptation to such extreme environments. Climate change needs more intense research on reprogramming of physiological, metabolic events during stress, regulation of growth and development under suboptimal conditions and adaptation of plants to extreme environments. Approximately 40% of the arable earth is arid, semi arid or affected by soil salinity, which reduces crop yield leading to increasing ecological, agronomical and economical impact. A correlation between increased frequency of extreme environmental events and global warming requires more efficient, environmentally compatible agricultural practices including the implementation of new crop cultivars with enhanced tolerance to environmental stresses (Ahuja et al., 2010; Boyer, 1982; Etterson and Shaw, 2001; Gregory et al., 2005; Kintisch, 2009). Water shortage, extreme temperatures or high salinity lead to a depletion of cellular water content, enhance the cellular osmotic potential generating osmotic stress in plants. Physiological responses to drought, cold and salt stress are similar and include reduced shoot growth and photosynthetic activity, accumulation of reactive oxygen species (ROS), alterations in ion transport and compartmentalisation, and changes in metabolite profiles (Lugan et al., 2009, 2010; Munns, 2002; Shulaev et al., 2008). One of the metabolic consequences of osmotic stress is the accumulation of low-molecular-weight organic compounds that do not interfere with normal metabolic reactions and

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

107

are considered to have protective functions (Ashraf and Harris, 2004; Bartels and Sunkar, 2005; Bohnert et al., 1995; Hare et al., 1998; Hasegawa et al., 2000; Yancey et al., 1982). This chapter is intended to give an overview of osmoprotective compounds with an emphasis on recent advances describing their function and significance to adaptation in improving stress tolerance.

II. GLOBAL METABOLIC CONSEQUENCES OF OSMOTIC STRESS Metabolism is of essential functional importance to support growth and development by providing building blocks and energy for plant structures and reserves, to generate signalling compounds for coordination and defence or to produce protective substances to cope with adverse environmental conditions. At a systemic level, metabolism can be viewed as the most integrated (and possibly informative) action reflecting both genotypic and environment-dependent regulations and resulting in the phenotype. Control of energy inputs and outputs is mainly driven through metabolic processes to support optimal growth and reproduction, which is a major goal for all organisms (Baena-Gonzalez and Sheen, 2008; Stitt et al., 2010). Metabolic adjustment is a challenging task for plants, which have to resist fluctuating environmental conditions frequently perceived as stress factors. As a matter of fact, plant metabolic networks are under constraint and metabolic reconfigurations are observed under abiotic stress exposition (Bohnert and Sheveleva, 1998; Guy, 1990) Major trends frequently observed in many plant species exposed to high salt, low water availability, and high or low temperatures include the accumulation of primary metabolites like amino acids, non-structural sugars and organic acids (Ahuja et al., 2010; Sanchez et al., 2008). Multiple functions have been proposed for these stress-induced metabolic adjustments like osmotic potential regulation, thermo- and osmoprotection, chaperon-like roles of macromolecules, adjustments in the carbon/nitrogen balance, scavenging of ROS or reactive nitrogen species, redox buffering, pH adjustment, carbon and nitrogen reserves deposition and signal transduction (Bartels and Sunkar, 2005; Noctor, 2006). The impact of stress on metabolic network is often associated with concurrent depression of growth and development depending on the amplitude and the duration of the physico-chemical pressure (Chaves et al., 2003). Exploring correlations between biomass production and metabolite profiles in recombinant inbred line (RIL) populations of Arabidopsis, Meyer and collaborators proposed that growth may drive metabolism and not the opposite and biomass can be described as a function of metabolic composition. Significant correlation

Author's personal copy 108

L. SZABADOS ET AL.

could be observed between biomass production and specific combination of metabolites (Meyer et al., 2007). Through the analysis of quantitative trait loci (QTL) with RIL and introgression line (IL) populations of Arabidopsis thaliana, it was demonstrated that there was a substantial and significant overlap of at least a subset of the biomass QTL, with metabolic QTL, suggesting a strong link between biomass and primary metabolism (Lisec et al., 2008). In contrast to secondary metabolism, it is proposed that perturbation of the primary metabolic network should have strong detrimental effects on plant performance. Metabolic adjustments in plants under osmotic stress are often observed through accumulation of specific metabolites. Such changes can be interpreted either as regulated adaptive or acclimation events or as biochemical consequences of growth impairment and trophic disruption. Plants that are tolerant to harsh environments (high salt, low water, high temperature) generally display low relative growth rate and often possess high solute concentrations (Alpert, 2006; Lugan et al., 2010). Although much evidence is now available about the adaptive or protective role of specific accumulation of compatible osmoprotectants, antioxidant and ROS scavenging compounds, a clear explanation of functional consequences of symptomatic global metabolic changes is still missing. Studies of metabolic adjustments in plants under stressful conditions have been encouraged in the past years, as technologies available for metabolomic approaches have been significantly improved (Fiehn et al., 2000; Ratcliffe and Shachar-Hill, 2005; Sumner et al., 2003; Weckwerth, 2003). Together with transcriptomic and proteomic approaches, metabolomic analysis gives a more comprehensive view of systems biology studies to investigate biological networks in order to understand plant responses to environmental cues and to develop tolerance improvement strategies. Metabolic profiling has been used recently for the description of plant adaptation to a wide range of biotic and abiotic stresses, creating the emerging field of environmental metabolomics (Bundy et al., 2009; Lugan et al., 2009; Schauer and Fernie, 2006; Shulaev et al., 2008). Complex changes in the metabolome during temperature stress have been characterised by studying the metabolic events associated with cold acclimation or acquired thermotolerance (Browse and Lange, 2004; Guy et al., 2008; Hannah et al., 2006; Kaplan et al., 2007; Korn et al., 2010; Shuman et al., 2006; Stitt and Hurry, 2002). Reconfiguration of central carbohydrate metabolism under stress appears to play a major role in plant response to temperature, where proposed triple function of sugars as nutrients, signalling compounds and putative ROS scavengers complicates the elucidation of the mechanisms involved (Guy et al., 2008; Rolland et al., 2006; Rosa et al., 2009; Stitt and Hurry, 2002). Moreover, the importance of DREB1/CBF

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

109

transcription factors in the regulation of the low-temperature metabolome, associated with cold acclimation has been demonstrated in Arabidopsis (Cook et al., 2004; Maruyama et al., 2009). Changes in lipid molecular profiles during cold acclimation and freezing have also been described in Arabidopsis using ESI-MS/MS-based technologies (Wang et al., 2006; Welti et al., 2002). Few studies have used holistic metabolic fingerprinting or metabolic profiling in plants to highlight biomarkers and understand the molecular basis of tolerance to salt or drought stress (Sanchez et al., 2008). Valuable information has been obtained about global management of metabolic networks under osmotic stress (induced either by exposure to high salt concentrations or water deprivation) in crop species like grapevines (Cramer et al., 2007), tomato (Johnson et al., 2003; Semel et al., 2007), rice (Zuther et al., 2007), pear (Larher et al., 2009) or lupin (Pinheiro et al., 2004). Halophytic extremophiles and desiccation-tolerant species have also been the subject of metabolic profiling, searching for biochemical attributes of salt or dehydration tolerance (Gagneul et al., 2007; Ksouri et al., 2010; Lugan et al., 2010; Moore et al., 2009; Weber et al., 2007). Halophytes belonging to different species collected in the same inland salt marsh habitat have been compared and characterised in terms of nitrogenous compound production, thus highlighting some competitive regulations between betaines and proline accumulation (Tipirdamaz et al., 2006). The importance of antioxidant systems in desiccation tolerance has been supported by a study on Myrothamnus flabellifolia (Kranner et al., 2002). In addition to antioxidant metabolism, carbohydrate metabolism was shown to be reconfigured in resurrection plants undergoing dehydration (Moore et al., 2007; Whittaker et al., 2001). Due to the available genetic and genomic resources most of the recent molecular and integrated studies on metabolic adjustment under salt or drought stress have used Arabidopsis, a typical glycophyte species (Lugan et al., 2009; Tester and Bacic, 2005). Such analysis can combine transcriptomic, proteomic and metabolic profiling approaches, and could markedly increase our understanding of global plant stress response and adaptation to stress conditions such as drought (Mittler and Blumwald, 2010; Seki et al., 2007; Shulaev et al., 2008; Urano et al., 2010). Through multiparallel and iterative combinatorial experiments, the central role of abscisic acid in stressregulated metabolic pathways and redox control has been illustrated (Ghassemian et al., 2008; Kempa et al., 2008; Urano et al., 2010). Genetic variation between naturally occurring populations of Arabidopsis has also provided a valuable source of information to unravel metabolic traits and genetic architecture associated with the complex mechanisms of abiotic stress tolerance (Bouchabke et al., 2008; Brosche et al., 2010; Cross et al., 2006; Hannah et al., 2006; Katori et al., 2010; Wingler and

Author's personal copy 110

L. SZABADOS ET AL.

Roitsch, 2008). To decipher the metabolic attributes and the genetic bases of extreme stress tolerance, Arabidopsis-relative model systems (ARMS) have recently been developed (Amtmann, 2009; Amtmann et al., 2005; Orsini et al., 2010). Recent comparative metabolic profiling experiences between Arabidopsis and Thellungiella salsuginea (halophila), a relative halophyte, contributed to the detection of relevant variation in metabolic composition and function in response to salt or drought treatments (Gong et al., 2005; Lugan et al., 2010; Pedras and Zheng, 2010). The metabolic composition of the salt tolerant T. salsuginea seems to be more compatible with dehydration, suggesting that this halophyte has more efficient osmoprotection than Arabidopsis (Lugan et al., 2010). It can be predicted that more and more studies will include metabolomics as a comprehensive addition to the systems biology approach to decipher plant stress response (Saito and Matsuda, 2010). The predicted changes in climatic conditions will likely create combinations of various abiotic stresses and make it necessary to adapt novel breeding or genetic engineering approaches to provide new crop varieties for modern agriculture (Mittler and Blumwald, 2010; Shulaev et al., 2008). Such practices will certainly encourage the development of metabolic phenotyping procedures and their adaptation in knowledge-based breeding programmes.

III. OSMOPROTECTIVE COMPOUNDS The most common osmoprotective compounds are amino acids, sugar alcohols, sugars and betaine-type quaternary amines (Table I; Chen and Murata, 2008; Hare and Cress, 1997; Hare et al., 1998; Hasegawa et al., 2000; Kerepesi et al., 1998; Parida and Das, 2005; Shulaev et al., 2008; Szabados and Savoure, 2010; Yancey et al., 1982). Composition and concentration of the solutes in stressed plants can vary considerably, depending on species and are influenced by the environmental conditions (Evers et al., 2010; Kumar, 2009; Lugan et al., 2010; Murakeozy et al., 2003; Sanchez et al., 2008). In contrast to inorganic ions, which can be harmful at high concentrations, osmolytes are considered compatible solutes, which can contribute to cell turgour, protect cellular structures, replace inorganic salts and alleviate ion toxicity. Compatible osmolytes were thought to mediate osmotic adjustment when water supply is limited, to replace water in biochemical reactions, stabilise the internal osmotic potential and to protect macromolecular structures. Therefore, osmotic adjustment has traditionally been accepted to be the primary function of osmolytes in plants (Crowe et al., 1992; Ford, 1984; Hasegawa et al., 2000; Ingram and Bartels, 1996; Parida and Das, 2005;

Author's personal copy

TABLE I Metabolism of the Most Common Osmoprotective Compounds in Plants

Osmolyte

Type of compound

Glycine betaine Quaternary ammonium compound

Proline

Amino acid

Trehalose

Sugar

Biosynthesis pathway Choline -(CMO)Betaine aldehyde -(BADH)Glycine betaine Glutamate -(P5CS)Pyrroline-5carboxylate -(P5CR)Proline

Enzymes in biosynthesis

Degradation pathway

Enzymes in catabolism

CMO: choline monooxygenase BADH: betaine aldehyde dehydrogenase P5CS: pyrroline5-carboxylate synthase P5CR: pyrroline5-carboxylate reductase

UDP TPS: trehalose phosphate glucose þ glucosesynthase 6-phosphate TPP: trehalose phosphate -(TPS)phosphatase Trehalose-6phosphate -(TPP)Trehalose

References Burnet et al. (1995), Hanson et al. (1994), Weigel et al. (1986)

Proline -(ProDH)Pyrroline-5-carboxylate -P5CDH)Glutamate

ProDH: proline dehydrogenase P5CDH: pyrroline5-carboxylate dehydrogenase

Trehalose -(TRE)2 Glucose

TRE: trehalase

Yoshiba et al. (1995), Hu et al. (1992), Szoke et al. (1992), Delauney and Verma (1993), Verbruggen et al. (1993), Strizhov et al. (1997), Kiyosue et al. (1996), Peng et al. (1996), Deuschle et al. (2004) Blazquez et al. (1998), Frison et al. (2007), Gussin et al. (1969) Lopez et al. (2008), Muller et al. (2001), Pramanik and Imai (2005), Vogel et al. (1998, 2001)

(continues)

Author's personal copy

Table I

Osmolyte

Type of compound

Mannitol

Polyalcohol

Myo-inositol

Polyalcohol

Pinitol

Polyalcohol

Biosynthesis pathway

(continued )

Enzymes in biosynthesis

Fructose-6P -(M6PI)Mannose-6P -(M6PR)Mannitol-1P -(M1PP)D-Mannitol D-Glucose-6P -MIPSMyo-inositol-1PMyo-inositol

M6PI: mannose-6P isomerase M6PR: mannose-6P reductase M1PP: mannose-1P phosphatase

Myo-inositol -(IMT1)D-ononitol Pinitol

IMT1: inositol-Omethyltransferase ononitol epimerase

MIPS: myo-inositol1-phosphate synthase

Degradation pathway Mannitol -(MTD)Mannose D-Mannose-6P -(M6PI)Fructose-6P

Enzymes in catabolism MTD: mannitol dehydrogenase M6PI: mannose-6P isomerase

References Loescher (1987, 1992), Rumpho et al. (1983)

Majumder et al. (1997), Johnson and Sussex (1995), Yoshida et al. (1999, 2002), Abreu and Aragao (2007), Wei et al. (2010a,b) Rammesmayer et al. (1995), Sengupta et al. (2008)

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

113

Yancey et al., 1982). This model has been challenged by results obtained with transgenic plants and mutants, suggesting that osmolytes could have alternative protective functions. Concentrations of organic osmolytes were found to be lower than inorganic solutes in several halophytes, suggesting that these compounds might not be important for osmotic adjustment in these plants (Gagneul et al., 2007). For example, proline overproducing transgenic tobacco plants were shown to be more tolerant to salt and drought stress, without unequivocal data on osmotic adjustment (Blum et al., 1996; Kishor et al., 1995). Some osmoprotectants might have other protective functions, such as stabilisation of redox balance, maintenance of proper protein folding, mediating sugar or stress signals (Chen and Murata, 2008; Hare et al., 1998; Parida and Das, 2005; Rosgen, 2007; Szabados and Savoure, 2010). One important function of osmoprotective compounds is the stabilisation of proteins under conditions that can lead to protein denaturation. Equilibrium between native and unfolded forms of proteins is influenced by solvent composition. Unfavourable environmental conditions such as extreme temperatures, high salinity or dehydration can alter secondary and tertiary structure of proteins and modify the ratio of active and inactive proteins. Protein denaturation, formation of protein aggregates and accelerated protein degradation can be the cellular consequence of such stress conditions. However, the maintenance of native folding of proteins is essential for their function and can be facilitated by osmoprotective compounds (Bolen and Baskakov, 2001; Burg, 1995; Yancey et al., 1982). Protective osmolytes were shown to stabilise proteins and push the balance of the folding equilibrium towards native, actively folded proteins through raising the free energy of the unfolded proteins (Auton and Bolen, 2005; Street et al., 2006). The protein backbone but not the side chains were shown to be the preferred target for the stabilising function of osmolytes which explains their universal protecting effect (Auton and Bolen, 2004; Liu and Bolen, 1995). Protecting osmolytes improve thermodynamic stability of proteins via hydrogen bonding, which in contrast to hydrophobic interactions, does not affect other cellular functions during stress conditions (Holthauzen and Bolen, 2007; Kumar, 2009). Cyclic polyalcohols such as pinitol and myo-inositol were shown to protect plant and bacterial enzymes against thermally induced inactivation (Jaindl and Popp, 2006). By preserving the native conformation, proteins are protected from aggregation or degradation, which can therefore be considered as principal functions of osmoprotective compounds (Bolen, 2001; Kumar, 2009; Rosgen, 2007; Street et al., 2006). Different osmolytes most probably act independently and have no synergistic or competing effects in their interactions with proteins (Holthauzen and Bolen, 2007).

Author's personal copy 114

L. SZABADOS ET AL.

Osmolytes can stabilise the structure of large multiprotein complexes as well. Glycine betaine (GB) was shown to support the oxygen-evolving activity of photosystem II (PSII) complex by preventing the dissociation of regulatory proteins from the core complex (Papageorgiou and Murata, 1995) and stabilise the efficiency of PSII photochemistry (Zhang et al., 2008). GB can ameliorate the damage of PSII and thereby stabilise photosynthesis under salt and cold stress (Ohnishi and Murata, 2006). Protection of cellular structures such as membranes and thylakoids against destabilisation during high temperatures or freezing was attributed to some of the protective compounds such as GB (Jolivet et al., 1982; Zhao et al., 1992). Some compounds can function as sinks of reducing power or carbon and nitrogen source after stress is relieved (Greenway and Munns, 1980; Hare et al., 1998). Accumulation of ROS is the result of various abiotic stresses including drought and high salinity. The protective function of osmolytes can include the suppression of radical oxygen production or scavenging of ROS. Accumulation of proline, GB and fructans in transgenic tobacco plants leads to reduced oxidative damage after freezing (Hong et al., 2000; Parvanova et al., 2004). Protection of the photosynthetic apparatus by GB during stress can reduce ROS accumulation and minimise lipid peroxidation during salt stress (Chen and Murata, 2008; Demiral and Turkan, 2004). Activation and protection of the ROS detoxification system is another key component of stress tolerance (Moradi and Ismail, 2007). Osmoprotective compounds can scavenge ROS directly, or contribute to the protection of the enzymes involved in the antioxidant system. Proline was suggested to be a singlet oxygen quencher during osmotic stress as it could reduce ROS damage such as lipid peroxidation in different plants (Hong et al., 2000; Matysik et al., 2002; Mehta and Gaur, 1999; Siripornadulsil et al., 2002; Smirnoff and Cumbes, 1989; Wang et al., 2009). GB increased catalase activity in cold-stressed tomato plants and thereby reduced ROS levels and improved chilling tolerance (Park et al., 2006). Stabilisation of the detoxifying enzymes such as ascorbate peroxidase and glutathione reductase was recently attributed to proline and can be important for the elimination of ROS during salt and drought stress (Sze´kely et al., 2008). Accumulation of ROS can be ameliorated by enhanced rate of proline biosynthesis during stress, which can help to maintain photosynthetic electron flow in the chloroplasts, stabilise redox balance and reduce photoinhibition (Hare and Cress, 1997; Szabados and Savoure, 2010). Osmolytes can control ROS-dependent damage through other, less-known pathways also. ROS was shown to promote Kþ efflux in root epidermal cells, which was significantly reduced by several osmolytes, including proline, GB, mannitol, myo-inositol and trehalose (Cuin and Shabala, 2007).

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

115

Osmolytes can have regulatory functions and modulate metabolic processes and gene activity. Proline and GB were shown to induce the expression of defence genes such as catalase and peroxidase, which can suppress cell death during stress (Banu et al., 2009). GB induced a group of low-temperature-responsive genes, which could elicit cold acclimation and freezing tolerance (Allard et al., 1998). Transcript profiling showed that exogenous GB treatment enhances the expression of numerous Arabidopsis genes, encoding transcription factors, membrane transporters and ROS detoxifying enzymes (Einset et al., 2008). Expression of catalase and an NADPH-dependent ferric reductase (FRO2) was enhanced by GB treatment, which contributed to antioxidant activity and chilling tolerance (Einset et al., 2007; Park et al., 2006). Certain regulatory functions were attributed to proline as well. Transcript profiling revealed that part of the rehydration-inducible Arabidopsis genes can be activated by proline (Oono et al., 2003). bZIP-type transcription factors were suggested to recognise the conserved PRE element in the promoter region of these genes and to activate their transcription (Satoh et al., 2004; Weltmeier et al., 2006). A. QUATERNARY AMINES: GLYCINE BETAINE

The quaternary ammonium compound GB is a methylated derivative of glycine, which accumulates at high concentrations in many halophyte plants (Chen and Murata, 2008; Hanson et al., 1983). GB is synthetised from choline in two steps. Betaine aldehyde hydrate generated from choline by the action of choline monooxygenase (CMO) produces betaine aldehyde hydrate from choline, which is converted spontaneously to betaine aldehyde and subsequently oxidised to GB by NADþ-dependent betaine aldehyde dehydrogenase (BADH; Burnet et al., 1995; Fitzgerald et al., 2009; Fujiwara et al., 2008; Hanson et al., 1983). GB accumulation can be induced by different stress conditions such as osmotic stress (Hanson and Nelsen, 1978), salinity (Hanson et al., 1991), drought (Guo et al., 2009), heat (Jolivet et al., 1982) and cold stresses (Allard et al., 1998; DeRidder and Crafts-Brandner, 2008). The significance of GB accumulation for stress tolerance has been investigated in transgenic plants engineered to produce GB by overexpressing plant-derived CMO and BADH genes (Guo et al., 1997; Nakamura et al., 1997; Shirasawa et al., 2006). Alternatively, bacterial betaine aldehyde dehydrogenase (betB) or choline oxidase (COX or codA) genes were introduced and expressed in transgenic plants, leading to increased GB accumulation (Holmstrom et al., 1994; Mohanty et al., 2002; Park et al., 2007; Sakamoto et al., 1998; Su et al., 2006) .The beneficial effects of GB accumulation

Author's personal copy 116

L. SZABADOS ET AL.

regarding salt and osmotic stress tolerance have been demonstrated in a number of engineered GB-accumulating plants, including tobacco (McNeil et al., 2001; Nuccio et al., 1998; Zhang et al., 2008), tomato (Park et al., 2004, 2007) and rice (Chen and Murata, 2002, 2008; Guo et al., 1997; Kathuri et al., 2009; Nakamura et al., 1997; Sakamoto et al., 1998; Shirasawa et al., 2006). GB accumulation in BADH-overexpressing transgenic tobacco led to enhanced salt or heat tolerance mainly by protecting photosynthetic activity through the maintenance of Rubisco activity and PSII activity (Yang et al., 2005, 2007, 2008). Targeted accumulation of GB in chloroplasts has been achieved by engineering a plastid-expressed CMO gene, leading to higher PSII activity during salt and drought stress (Zhang et al., 2008). Therefore protection of the PSII by GB-mediated osmoprotection can be a promising strategy to improve drought and salt tolerance in crops. Rice was engineered to produce GB by introducing the chloroplast-targeted choline oxidase (codA9) gene from Arthrobacter globiformis. GB synthesis enhanced the activity of PSII, led to better ROS detoxification and improved physiological and agronomic performance (Kathuria et al., 2009). Plants are usually very sensitive to environmental stress during reproduction. GB was shown to have a particularly important protective effect on reproductive organs, such as inflorescence apices and flowers during drought and cold stress (Chen and Murata, 2008; Sakamoto and Murata, 2000). Engineering of GB accumulation reduced chilling damage on tomato flowers, leading to a 10–30% increase in fruit production (Park et al., 2004). These data confirm that GB has osmoprotective qualities, which can therefore be explored to improve tolerance to salinity and probably to other abiotic stresses such as drought and cold. B. SUGARS: TREHALOSE

Trehalose is a nonreducing disaccharide of glucose, which was first described in desiccation-tolerant organisms capable of surviving dehydration, including resurrection plants (Drennan et al., 1993; Fernandez et al., 2010; Liu et al., 2008; Moore et al., 2009). Later, trehalose accumulation was detected in numerous other plants under different stress conditions such as drought, cold, high salinity (Iordachescu and Imai, 2008; Kaplan et al., 2004; Kosmas et al., 2006; Lopez et al., 2008; Pramanik and Imai, 2005) and during various plant–microbial interactions (Brodmann et al., 2002; Dominguez-Ferreras et al., 2009; Farı´as-Rodrı´guez et al., 1998; Lopez et al., 2008). Trehalose biosynthesis is a two-step pathway. First, trehalose-6-phosphate is produced from UDP glucose and glucose-6-phosphate by trehalose phosphate synthase, which is converted to trehalose by the enzyme trehalose phosphate

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

117

phosphatase (Vogel et al., 1998, 2001). Trehalose is catabolised by trehalase, which converts it to glucose (Brodmann et al., 2002; Goddijn et al., 1997). The importance of trehalose in modulating plant stress responses was demonstrated by engineering the trehalose biosynthetic pathway in transgenic plants. Stress-dependent trehalose accumulation was engineered in transgenic rice by the regulated overexpression of the fused bacterial trehalose biosynthetic genes otsA and otsB. Enhanced trehalose levels resulted in improved salt, drought and cold tolerance, and lower photo-oxidative damage, which was suggested to modulate carbohydrate metabolism (Garg et al., 2002). Several other transgenic plants were produced that accumulated trehalose at high levels, and subsequently improved tolerance to drought, freezing or high salinity (Jang et al., 2003; Romero et al., 1997; Stiller et al., 2008; Yeo et al., 2000). However, constitutive overexpression of the trehalose biosynthetic genes led to pleiotropic growth alterations, dwarfism and abnormal root structure (Garg et al., 2002; Romero et al., 1997; Schluepmann et al., 2004; Yeo et al., 2000). Using stress-induced promoters such as the drought-induced Arabidopsis AtRAB18 or potato StDS2 promoters, the environmentally controlled upregulation of trehalose biosynthesis was important to generate transgenic plants with enhanced stress tolerance, but without morphological abnormalities (Karim et al., 2007; Stiller et al., 2008). The function of trehalose in stress responses is controversial. Trehalose was shown to have the ability to stabilise membranes and protect proteins in desiccated tissues and was suggested to function as chemical chaperon (Crowe, 2007; Crowe et al., 1984). However, trehalose levels in the engineered plants usually remained well below 1 mg/g fresh weight, suggesting that trehalose does not act as a compatible solute, but has an alternative function (Garg et al., 2002). The fact that trehalose-accumulating transgenic plants exhibit abnormal morphology can also indicate that this compound is not a neutral osmolyte. Recently, trehalose-6-phosphate was proposed to function as a metabolic signal involved in the control of the SnRK1 activity, which is a central regulator of sugar and energy homeostasis (Paul et al., 2010; Zhang et al., 2009). Therefore, the signalling function of trehalose and trehalose-6P could be more important than the previously suggested chaperone or osmolyte function, although in some tissues such a protective role cannot be excluded (Fernandez et al., 2010). C. POLYALCOHOLS: MANNITOL, PINITOL, INOSITOL

Accumulation of polyalcohols, such as mannitol and pinitol, has been detected in several water-stressed plants and are considered as important compatible solutes, which serve as ROS scavengers and molecular

Author's personal copy 118

L. SZABADOS ET AL.

chaperones (Bohnert et al., 1995; Ford, 1984; Sengupta et al., 2008). Mannitol is the most common sugar alcohol and is an important photosynthetic product in a number of plant species (Loescher et al., 1992; Rumpho et al., 1983). In plants, mannitol is synthetised from fructose-6P by subsequent action of mannose-6P isomerase (phosphomannose isomerase), mannose6P reductase (M6PR) and mannose-1P phosphatase (Loescher et al., 1992; Rumpho et al., 1983). Catabolism of mannitol is controlled by mannitol dehydrogenase, which produces mannose. Phosphorylation of mannose results in mannose-6P, which is converted to fructose-6P by mannose-6P isomerase (Loescher, 1987). One of the first demonstrations of the osmoprotective function of an osmolyte was the overexpression of a bacterial mannitol 1-phosphate dehydrogenase (mtlD) in transgenic tobacco plants, which led to increased mannitol accumulation in conjunction with enhanced salt tolerance (Tarczynski et al., 1993). In agreement with these results, engineering of mannitol biosynthesis by overexpression of mtlD led to enhanced mannitol accumulation and improved salt tolerance in various transgenic plants including wheat (Abebe et al., 2003), poplar (Populus tomentosa; Hu et al., 2005) and loblolly pine (Pinus radiata; Tang et al., 2005). Overexpression of celery M6PR is an alternative way to enhance mannitol biosynthesis and was shown to be an efficient way to improve salt tolerance of Arabidopsis (Zhifang and Loescher, 2003). As an alternative use, M6PR was employed as a selectable marker for plant transformation, using mannose tolerance as selection criteria (Song et al., 2010). Myo-inositol is an essential polyalcohol in plants and all eukaryotes. Biosynthesis starts from D-glucose-6P, which is converted to myo-inositol1P by myo-inositol-1P synthase (MIPS; Johnson and Sussex, 1995; Majumder et al., 1997). Myo-inositol is produced from myo-inositol-1P by dephosphorylation and is used for the subsequent biosynthesis of all inositolcontaining compounds, including phospholipids. MIPS genes were shown to be salt-induced, leading to accumulation of myo-inositol in the halophyte ice plant, but not in the glycophyte Arabidopsis (Ishitani et al., 1996). MIPS genes can be regulated by several environmental stress factors such as drought, heat and cold stress, high light and were shown to be controlled by ABA signals (Abreu and Aragao, 2007; Wei et al., 2010a,b; Yoshida et al., 1999, 2002). Myo-inositol serves not only as osmoprotectant compound, but also functions as signal that controls metabolic responses to stress, such as sodium uptake in saline environments (Nelson et al., 1999). Phosphorylated derivatives of myo-inositol are important signalling compounds, which are involved in numerous regulatory pathway and control diverse aspects of plant development, responses to biotic and abiotic stresses.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

119

Other polyalcohols with osmoprotective features such as D-ononitol and are synthetised from myo-inositol through the extension of the metabolic pathway. Pinitol is a methylated inositol, which is synthetised from myo-inositol by inositol-o-methyltransferase (IMT1) and ononitol epimerase (OEP1) (Bohnert et al., 1995; Rammesmayer et al., 1995; Sengupta et al., 2008). Pinitol accumulation is a characteristic feature of a number of halophytic plants in saline environment and occurs in several glycophytic plants grown under osmotic stress conditions (Ford, 1984; Fougere et al., 1991; Murakeozy et al., 2003; N’Guyen and Lamant, 1988; Sengupta et al., 2008). Salt-induced pinitol hyperaccumulation was found in Porteresia coarctata, a halophytic wild relative of rice, which is missing in domesticated rice. The inositol methyl transferase 1 (PcIMT1) gene is strongly upregulated by salt in wild rice, suggesting that this gene controls the stress-induced pinitol accumulation in this halophytic plant which is an essential metabolic response for salt tolerance (Sengupta et al., 2008). Engineering of complex metabolic pathways by altered expression of single genes has limitations. Simultaneously regulated expression of several genes encoding key enzymes could have more profound effects on the metabolic pools. Biosynthesis of myo-inositol was enhanced by introduction and overexpression of the MIPS gene from P. coarctata in tobacco leading to inositol accumulation and enhanced salt tolerance (Majee et al., 2004). Even higher levels of salt tolerance were reported by the introgression and simultaneous overexpression of the MIPS coding gene from P. coarctata and the IMT1 coding gene from M. crystallinum in transgenic tobacco. By comparison, double transgenics accumulated more inositol and pinitol than plants transformed by single genes, which conferred improved growth, higher photosynthetic activity and lower oxidative damage during salt stress (Patra et al., 2010). D-pinitol

D. AMINO ACIDS: PROLINE

Proline is an essential amino acid, a common denominator of many stress responses being accumulated during diverse abiotic and biotic stresses such as high salinity (Ben Hassine et al., 2008; Voetberg and Stewart, 1984; Yoshiba et al., 1995), drought (Ben Hassine et al., 2008; Choudhary et al., 2005; Huang and Cavalieri, 1979; Rhodes et al., 1986), UV irradiation (Saradhi et al., 1995), heavy metals (Mehta and Gaur, 1999; Schat et al., 1997; Singh et al., 2010) and oxidative stress (Yang et al., 2009). Moreover, proline was reported to accumulate in plants infected by avirulent bacteria (Fabro et al., 2004) or Agrobacterium (Haudecoeur et al., 2009). In plants, proline is synthesised from glutamate in the cytosol and likely also in the chloroplast by the sequential action of delta-1-pyrroline-5-carboxylate

Author's personal copy 120

L. SZABADOS ET AL.

Synthesis

synthetase (P5CS) and P5C reductase (P5CR). P5CS produces glutamate semialdehyde, which is unstable and is immediately converted to pyrroline-5carboxylate (P5C). P5CR reduces P5C to proline, a reaction that takes place in the cytosol and according to biochemical data also in the chloroplast (Fig. 1; Delauney and Verma, 1993; Hu et al., 1992; Rayapati et al., 1989; Strizhov et al., 1997; Szabados and Savoure, 2010; Szoke et al., 1992; Verbruggen et al., 1993; Yoshiba et al., 1995). Proline synthesis is considered as an evolutionary conserved process based on the similarity that exists among the prokaryotic and eukaryotic pathways and the high homology of the involved enzymes. The human and plant P5CSs are bifunctional enzymes representing evolved fusion products of the domains responsible for the catalytic activity in the prokaryotic proB and proA genes, encoding glutamate kinase (GK) and g-glutamyl phosphate reductase (GPR), respectively (Csonka, 1989; Hu et al., 1992; Perez-Arellano et al., 2010). These genes are arranged in a single operon in bacteria and their paralogs in lower eukaryotes, such as yeast, still form two separate enzymes (Takagi, 2008). Despite sharing functional homology, the mammalian and plant P5CSs are localised to different cellular compartments. The human P5CS isoforms are functioning in the mitochondria, utilising innermitochondrial glutamate and energy sources (Perez-Arellano et al., 2010),

Chloroplast

Cytosol

Mitochondrion

Glutamate

Glutamate

Arginine

NADPH+ + H+ NADP+ ATP ADP

NADPH+ + H+

P5CS1

P5CS1 & 2

Arginase

NADP+

GSA P5C P5CR

GSA P5C

Ornithine P5CR

NADP+

Proline

Proline

NADPH+ + H+

OAT Accumulation during stress in cytosol, vacuole, and PRPs Transport to the cytosol

Transport into mitochondria

Mitochondrion

Degradation

ProDH 1 & 2

Proline

P5C FAD+

FADH2

P5CDH NAD+/NADP+

eElectron transport chain

NADH+ + H+/ NADPH+ + H+

Glutamate

Fig. 1. Scheme of proline metabolism in plants. Proline synthesis occurs in the cytosol and likely also in the chloroplast. Proline degradation is carried out in the mitochondrion. The dashed arrows indicate the proline cycle. See text for details.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

121

whereas the two plant isoforms, P5CS1 and P5CS2, are functioning in the cytosol, with a likely abiotic-stress-induced shift of P5CS1 to the chloroplasts (Sze´kely et al., 2008). In yeast, both enzymes (GK and GPR) function in the cytosol (Takagi, 2008). Another source for P5C is the mitochondrial degradation pathway of arginine, whose first step is catalysed by arginase that forms ornithine and urea. In the second step, ornithine amino transferase (OAT) produces P5C by ornithine deamination in the mitochondria (Brownfield et al., 2008; Roosens et al., 1998; Xue et al., 2009). The importance of this pathway in proline biosynthesis has recently been questioned as proline levels were not altered in the oat mutant (Funck et al., 2008). Although P5C transporters have not been identified in plant mitochondrial or chloroplast membranes, biochemical evidence support the movement of P5C from the mitochondria to the cytosol in human and plant cells which enables the reduction of mitochondrial-produced P5C to proline in the cytosol (Miller et al., 2009; Yoon et al., 2004). Proline catabolism occurs in the inner-mitochondrial membrane of all eukaryotes. Proline degradation provides electrons and glutamate for mitochondrial usage. Proline dehydrogenase (ProDH), a FAD-enzyme localised to the inner-mitochondrial membrane, catalyses the first oxidising step of proline to P5C and meanwhile delivers electrons to the mitochondrial electron transport chain (Kiyosue et al., 1996; Peng et al., 1996; Rayapati and Stewart, 1991). P5C is further oxidised to glutamate or transported back to the cytosol for proline re-synthesis by the proline cycle (Fig. 1; Deuschle et al., 2004; Miller et al., 2009). Proline cycle exists in yeast, mammals and plants and contains the cytosolic P5CR and mitochondrial ProDH (Miller et al., 2009; Phang et al., 2010; Takagi, 2008). In human cells, proline cycle acts as a suppressor of carcinogenesis. It oxidises proline and provides an excess of electrons producing ROS that are responsible for initiating programmed cell death that prevents tumour development (Phang et al., 2010). P5CDH is the second enzyme in proline degradation, catalysing the oxidation of P5C to glutamate in the mitochondria and is responsible for maintaining proline-P5C homeostasis (Deuschle et al., 2001, 2004; Forlani et al., 1997). When P5CDH activity is impaired, hyperactivity of proline-P5Cproline cycle provides excess electrons to the mitochondrial electron chain and generates ROS by using O2 as the electron acceptor (Miller et al., 2009). ROS signals can be produced during the first hour of dehydration in Arabidopsis leaves, when a transient increase in ProDH transcription occurs (Kiyosue et al., 1996; Peng et al., 1996). In some plant species, derivatisation of proline occurs following its stress-induced accumulation. In the salt tolerant salt-cider (Tamarix spp.), part of the stress-accumulated proline is

Author's personal copy 122

L. SZABADOS ET AL.

modified to N-methylproline analogues including N-methyl-L-proline, trans4-hydroxy-N-methyl-L-proline and trans-3-hydroxy-N-methyl-L-proline whose function is still unknown (Jones et al., 2006). Proline accumulation during stress has multiple protective functions. For a long time, proline was considered a neutral osmolyte that protects cellular structures and stabilises enzymes (Delauney and Verma, 1993; Kavi Kishor et al., 2005; Mishra and Dubey, 2006; Sharma and Dubey, 2005; Sharma et al., 1998). Besides osmoprotection, proline was shown to have antioxidant activity, activate detoxification pathways, contribute to cellular homeostasis by protecting the redox balance, function as protein precursor, energy source for the stress-recovery process and even as a signalling molecule (Hoque et al., 2008; Islam et al., 2009; Khedr et al., 2003; Matysik et al., 2002; Szabados and Savoure, 2010; Sze´kely et al., 2008). In plants, maintenance of PSII and PSI activity as well as electron flux through the photosynthetic electron transport chain is very important in stress conditions. Inhibition of Calvin cycle and pentose phosphate pathway can channel NADPH, ATP and glutamate for proline synthesis in the chloroplasts. Thus, proline synthesis in the chloroplast may allow an efficient oxidation of photosynthetically produced NADPH providing the required NADPþ for electron acceptor avoiding the use of O2 that leads to ROS generation (Hare and Cress, 1997; Szabados and Savoure, 2010). Numerous reports describe the importance of proline accumulation in salt and drought tolerance, although a clear relationship between proline accumulation and tolerance could not always be confirmed (Kavi Kishor et al., 2005; Lehmann et al., 2010; Szabados and Savoure, 2010; Verbruggen and Hermans, 2008). Convincing evidence for the protective function of proline was provided by studies of mutants and transgenic plants with proline deficiency or proline hyperaccumulation. Arabidopsis p5cs1 insertion mutants had only 10% proline of the wild type and were hypersensitive to salt stress, produced more ROS and lipid peroxidation products, which confirmed the importance of proline in stress tolerance, in particular, in ROS scavenging (Sze´kely et al., 2008). However, enhanced proline accumulation and improved salt tolerance was achieved by increasing the biosynthetic pathway through constitutive overexpression of the Vigna P5CS cDNA in transgenic tobacco (Kishor et al., 1995) and Clamydomonas (Siripornadulsil et al., 2002). Overexpression of feedback-insensitive P5CS in tobacco or rice could enhance proline levels even more than the wild-type enzyme, showing that post-translational regulation needs to be considered for metabolic engineering (Hong et al., 2000; Kumar et al., 2010). Alternatively the feedback-insensitive proAB gene from Bacillus subtilis was expressed in Arabidopsis, which produced more free proline and improved

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

123

osmotolerance (Chen et al., 2007a,b). While overexpression of Arabidopsis P5CR in transgenic soybean improved drought and heat tolerance, antisense expression of P5CR leads to stress sensitivity (De Ronde et al., 2004).These reports confirm that proline accumulation in dehydrated plants is not only a consequence of stress, but also a part of the metabolic defence system against abiotic stress. 1. Regulation of proline metabolism While proline biosynthesis is augmented and proline catabolism is repressed during stress, the opposite happens during recovery, when synthesis declines rapidly and catabolism is activated. Accumulation of cellular free proline levels has been attributed to increased transcription of P5CS and silencing of ProDH, while reciprocal transcriptional regulation suppress P5CS transcrip´ braha´m et al., 2003; Kiyosue et al., tion and activates ProDH after stress (A 1996; Peng et al., 1996; Ribarits et al., 2007; Strizhov et al., 1997; Verbruggen et al., 1996; Yoshiba et al., 1995). Whereas the stress-induced drastic reduction in ProDH transcription is a general phenomenon in plants, some species lack the concomitant increase in P5CS transcription (Ginzberg et al., 1998; Miller et al., 2005). Stress-dependent P5CS activation and ProDH repression is controlled by both ABA-dependent and independent signalling pathways, ´ braha´m et al., 2003; modulated by light and brassinosteroid signals (A Sharma and Verslues, 2010; Strizhov et al., 1997; Sze´kely et al., 2008; Verslues et al., 2007). During incompatible plant–pathogen interactions P5CS activation is controlled by SA-dependent ROS signals (Fabro et al., 2004). In the ornithine pathway, OAT transcription is upregulated by salt stress in young Arabidopsis seedlings but not in mature plants, suggesting the involvement of developmental and temporal regulation in OAT expression (Roosens et al., 1998). Regulation of key enzymes involved in proline metabolism seems to vary according to environmental conditions and also during plant development (Lehmann et al., 2010; Szabados and Savoure, 2010). Calmodulin and MYB2 are likely involved in P5CS activation under salt stress, as interaction of MYB2 and the salt-induced CaM isoform (GmCaM4) was shown to upregulate P5CS1 transcription and leads to proline accumulation (Yoo et al., 2005). In addition to the MYB2, bZIP transcription factors (bZIP-TF) might also be involved in regulating transcription of the genes encoding enzymes of proline metabolism. Promoter analyses by several servers that identify promoter motifs (e.g., Agris, http://arabidopsis.med.ohio-state.edu) indicate that the promoters of all Arabidopsis genes encoding enzymes involved in proline metabolisms possess bZIP-TF recognition motifs: six sites in P5CS1 (At2g39800), 11 sites in P5CS2 (At3g55610), 14 in P5CR (At5g14800), four in ProDH1 (At3g30775), six in ProDH2 (At5g38710), two

Author's personal copy 124

L. SZABADOS ET AL.

in P5CDH (At5g62530) and five in OAT (At5g46180). As yet, only the molecular basis of induced ProDH transcription by the drought and salinity stressrecovery process has been unravelled, indicating the importance of bZIP transcription factors. bZIP-TFs from the S and C groups AtbZIP53 and AtbZIP10 (Jakoby et al., 2002) form homo- or heterodimers and promote the recruitment of the transcription complex to the ProDH promoter region. Interaction of these transcription factors with the PRE motif ACTCAT of the ProDH promoter strongly enhances its transcription (Satoh et al., 2002, 2004; Weltmeier et al., 2006). Interactions of bZIP11/ATB2 and other heterodimerforming partners from groups S and C with this PRE motif is also responsible for upregulating ProDH transcription in response to sugar depletion, an effect that also blocks the post-transcriptional inhibition of bZIP11 by sucrose (Hanson et al., 2008). The possible involvement of bZIP-TFs in the regulation of the other genes controlling proline metabolism still has to be confirmed. Upon recovery from stress, a reduction in P5CS and enhancement of ProDH transcription occur, leading to intensive proline degradation that provides energy and glutamate for the surviving cells (Kiyosue et al., 1996; Peng et al., 1996; Verbruggen and Hermans, 2008). In addition to changes in enzyme levels, mostly dictated by regulation at the transcriptional level, proline allosterically inhibits the enzymatic activity of bacterial GK/GPR and the plant P5CS, the initial enzyme in the synthesis pathway (Csonka et al., 1988; Fujita et al., 2003; Hu et al., 1992). The accumulation of proline during various stresses without exerting a feed-back regulation on P5CS activity is not well understood. Although expression of ProDH genes is induced by exogenous proline, ProDH transcription is silenced during stress, despite the high levels of accumulated free proline in the cells (Kiyosue et al., 1996; Miller et al., 2005). Hence, it is likely that a certain compartmentalisation of the stress-accumulated proline exists to avoid interaction with P5CS and cancel the silencing of ProDH transcription. Further studies are required to clearly understand the signalling pathways and define all the regulatory proteins involved. 2. Proline rich proteins and stress response Proline rich proteins (PRP) are plant-specific proteins that are mainly localised to the cell wall, playing different roles in cell wall maintenance and linkage to the plasma membrane. Their synthesis utilises a large portion of the free proline pool and normally occurs in most of the plant cells during development. These proteins are characterised by reiterated proline-rich sequence motifs. Many of the proline residues in these proteins are hydroxylated and thereafter glycosylated through their processing in the ER and Golgi network before being secreted to the outer side of the plasma membrane or linked to the membrane by a GPI-anchore or hydrophobic

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

125

transmembrane domain/s. The hydroxyproline-rich glycoproteins comprise a huge super family in Arabidopsis and other plants, which is divided into three subfamilies; arabinogalactans, extensins and PRP (Showalter et al., 2010). Most of these proteins are involved in cell wall assembly in all plant organs and along the developmental processes or are more specific to certain organs (Fowler et al., 1999). Oxidative cross-linking of tyrosine residues within the subfamily of extensin molecules strengthens the cell wall and increases resistance to invading pathogens. Whereas many hydroxyproline-rich proteins are involved in plant response to wounding and pathogen attacks, only a limited number of these PRP are involved in abiotic stress responses (Deepak et al., 2010). Most of them belong to the HyPRP family. HyPRPs, which are only found in plants, contain a proline-rich N-terminal repetitive domain suggested to protrude from the plasma membrane towards the cell wall and a hydrophobic C-terminal transmembrane domain with a certain conserved arrangement of eight cysteine residues (Jose-Estanyol et al., 2004). MsPRP2 is an HyPRP that is induced by water stress in alfalfa (Medicago sativa) cells under saline conditions (Deutch and Winicov, 1995), whereas another alfalfa HyPRP, MsACIC, reveals enhanced expression in cold-tolerant plants (Castonguay et al., 1994). The Brassica napus HyPRP, BNPRP, is also highly expressed at low temperatures, whereas low level of the BNPRP transcript is also present under normal growth conditions (Goodwin et al., 1996). As yet, the functional role of these proteins and their contribution to stress tolerance are not clear. It is likely that the Arabidopsis cell wall linker protein (CWLP) that shows 85% amino acid sequence homology to BNPRP (Goo et al., 1999), and is induced by cold treatment in Arabidopsis, is involved in the adhesion of the cell wall to the plasma membrane (Ziberstein’s group preliminary results). Since cold-induced synthesis of these proteins is linked to proline accumulation, availability of free proline might support the synthesis of PRP.

IV. OSMOPROTECTIVE COMPOUNDS AND ADAPTATION TO EXTREME ENVIRONMENTS Numerous earlier papers and reviews have discussed the importance of osmoprotective compounds in stress tolerance and the value they might have in the adaptation process to extreme environmental conditions (Hare et al., 1998; Rontein et al., 2002; Table II). While a number of reports described the coincidence of osmolyte accumulation with salt or drought tolerance, the adaptive value of osmoprotective compounds for stress tolerance was unequivocally demonstrated in a few cases only.

Author's personal copy TABLE II Plant Species or Genotypes in Which Osmolyte Accumulation is Important for Tolerance to Extreme Environmental Conditions Species

Compound

Environment

Atriplex halimus L (xero-halophyte)

Glycine betaine, sugars, proline

Arid environment, saline soil

Camphorosma annua (halophyte)

Glycine betaine, pinitol

salty–sodic soil

Cicer arietinum (chickpea genotypes with contrasting copper tolerance) Hordeum vulgare L (salt tolerant/ sensitive barley)

Proline

Copper toxicity

Proline, glycine betaine, sugars, hexose phosphates

Saline soil

Lepidium crassifolium (halophyte)

Proline, sugars

salty–sodic soil

Limonium spp. (halophyte, Plumbaginaceae spp.) Limonium gmelini (halophyte)

b-Alanine betaine, choline-Osulphate b-Alanine betaine, choline-Osulphate, pinitol Sugars, chiro-inositol, proline Proline, sugars

Hypoxic saline, sulphate-rich saline soils salty–sodic soil

Limonium latifolium (halophyte) Medicago truncatula, M. laciniata, (contrasting landraces in drought tolerance) Mesembryanthemum crystallinum (ice plant, halophyte) Oryza sativa (drought tolerant/ sensitive rice varieties) Oryza sativa (drought tolerant/ sensitive indica rice variety)

Polyols (myo-inositol, D-ononitol, D-pinitol) Proline Reducing sugars, proline, polyols

Reference Shen et al. (2002), Wang and Showalter (2004), Martinez et al. (2004) Murakeozy et al. (2003) Singh et al. (2010)

Saline soil Drought, arid environment

Chen et al. (2007b), Widodo et al. (2009) Murakeozy et al. (2003), this study Hanson et al. (1991, 1994) Murakeozy et al. (2003) Gagneul et al. (2007) Yousfi et al. (2010)

Saline soil

Nelson et al. (1999)

Drought, osmotic stress

Choudhary et al. (2005) Zuther et al. (2007), Roychoudhury et al. (2008)

Saline soil

Author's personal copy

Proline

Tolerance to Hg2þ-induced oxidative stress Drought, arid environment

Glycine betaine

Saline soil

Plumbaginaceae spp. (halophytes) Porteresia coarctata (wild rice, halophyte) Solanum tuberosum L. (drought tolerant/sensitive potato cultivars) Sorghum bicolor (sorghum: salt tolerant/sensitive variety) Thellungiella salsugiana (halophila) (halophyte)

Proline betaine Pinitol

Arid environment Saline soil

Hanson et al. (1991, 1994) Hanson et al. (1994) Sengupta et al. (2008)

Proline, inositol, galactose, galactinol Sugars, proline

Drought treatment

Evers et al. (2010)

Drought, arid environment

Proline, sugars (sucrose, fructose, glucose, raffinose, melibiose), sugar alcohols (inositol, galactinol)

Saline soil

Triticum aestivum (drought tolerant/ sensitive wheat genotypes) Triticum durum Desf. (drought tolerant/sensitive variety) Various contrasting legume species in drought tolerance

Sugars, fructan

Drought, arid environment

Premachandra et al. (1995) Lugan et al. (2010), Gong et al. (2005), Inan et al. (2004), Arbona et al. (2010), Taji et al. (2004) Kerepesi et al. (1998)

Sugars, proline

Arid environment, osmotic stress Drought, water stress

Oryza sativa (proline accumulating mutant) Phaseolus vulgaris (contrasting landraces in drought tolerance) Plumbaginaceae spp. (halophytes)

Proline

Pinitol

Such examples are listed where experimental evidence suggests the value of osmolyte accumulation for stress tolerance.

Wang et al. (2009) Tari et al. (2008)

Bajji et al. (2001) Ford (1984)

Author's personal copy 128

L. SZABADOS ET AL.

Osmoprotective compounds can accumulate to very high concentrations in extremophile plants in saline or arid environments, suggesting that these metabolites contribute considerably to the adaptation of these plants to the harsh environment. Some species produce high levels of sugars or polyols, others preferentially accumulate nitrogenous compounds. Halophytes usually accumulate one dominant compatible solute, which can be proline, GB, sorbitol, b-alanine betaine, choline-O-sulphate or sugar (Arbona et al., 2010; Hanson et al., 1991, 1994; Inan et al., 2004; Lugan et al., 2010; Tipirdamaz et al., 2006). Halophytic Limonium species were shown to accumulate high levels of quaternary ammonium compounds such as choline-O-sulphate, GB and b-alanine betaine, suggesting that these compounds are important for the salt tolerance of these species (Hanson et al., 1991). Sucrose accumulated in Juncus maritima, Phragmites communis and Scirpus maritimus, while maltose and rhamnose were abundant in Atriplex hastata and Plantago maritima, respectively. Polyols were present in Aster tripolium, Juncus maritimus, P. maritima and P. communis (Briens and Larher, 1982). Distribution of osmoprotective ammonium compounds among different species of the Plumbaginaceae family suggested that particular compounds have selective advantage in different environments. GB was dominant in species adapted to dry environments, choline-O-sulphate was advantageous in sulphatecontaining soils, b-alanine betaine is apparently more typical in species growing on hypoxic saline soils, and proline–betaine was detected in plants adapted to arid environments (Hanson et al., 1994). Composition and concentration of compatible osmolytes had species-specificity and showed seasonal fluctuation in three halophyte species in salty–sodic grasslands in Hungary. GB and pinitol was characteristic of Camphorosma annua, b-alanine betaine and pinitol accumulated in Limonium gmelini, while proline was most characteristic of Lepidium crassifolium. The highest osmolyte concentrations were measured in spring, characterised by low temperatures, hypoxic conditions and high salt concentrations in the habitats of the tested species (Murakeozy et al., 2003). Among the extremophile Chenopodiaceae species, Artiplex halimus is a known xero-halophyte plant, characterised by GB and proline accumulation in stressful environments (Shen et al., 2002; Wang and Showalter, 2004). Accumulation of GB and sugars were suggested to be responsible for osmotic adjustment during osmotic stress in this species (Martinez et al., 2004). In a more recent study, P. maritima was found to accumulate sorbitol, proline and GB, while species belonging to the Chenopodiaceae family are typical accumulators of GB in saline environments (Tipirdamaz et al., 2006). In other species, the importance of osmoprotecting compounds in stress tolerance has been studied by comparing genotypes with contrasting

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

129

drought or salt tolerance and osmolyte accumulation. The importance of osmoprotectants for salt and drought stress tolerance of cereals has long been discussed (Garcia et al., 1997). Proline and sugar levels were higher in drought and salt tolerant than sensitive rice varieties, suggesting that these protective compounds can contribute to stress tolerance of rice (Choudhary et al., 2005; Roychoudhury et al., 2008). In durum wheat, main osmolytes that correlated with drought tolerance were sugars, followed by proline and quarternary ammonium compounds (Bajji et al., 2001). A positive correlation between drought tolerance of wheat and glucose, fructose, sucrose and fructan contents was observed, suggesting that soluble carbohydrates have adaptive value for wheat (Kerepesi et al., 1998). However, accumulation of proline and GB did not correlate with salt tolerance in contrasting barley genotypes, while hexoses and TCA intermediates did, suggesting that carbohydrates contribute to salt tolerance of barley (Chen et al., 2007a,b; Widodo et al., 2009). Transcript profiling of drought tolerant and sensitive barley varieties revealed that tolerance correlated with the drought-dependent activation of genes, which are involved in the synthesis, and transport of GB, but not proline (Guo et al., 2009). GB is therefore important for drought but not for salt tolerance in barley, hexoses are significant for salt tolerance, while proline accumulation is rather a symptom of salt susceptibility. In sorghum, drastic increase of both proline and GB levels were recorded upon water deficit (Wood et al., 1996). Accumulation of solutes was higher in a drought-tolerant sorghum line than in a drought sensitive one, with the notable exception of proline, which does not seem to contribute to drought tolerance in this plant (Premachandra et al., 1995). In saline environments, pinitol accumulated in the halophytic wild rice, P. coarctata Roxb., but not in the cultivated rice. Inositol methyl transferase (IMT1) mediates pinitol synthesis in P. coarctata, while this gene is missing in the rice genome (Sengupta et al., 2008). Moreover, a salt tolerant form of L-MIPS was identified in P. coarctata, containing a 37 amino acid stretch, which conferred superior thermodynamic stability and elevated activity to the MIPS enzyme leading to enhanced myo-inositol synthesis in saline environments (Ghosh Dastidar et al., 2006). The capacity to synthesise and accumulate myo-inositol and pinitol can therefore be an important adaptive feature in P. coarctata and probably a number of other extremophile plants (Sengupta and Majumder, 2009). In legumes, drought or salt tolerance correlated with osmolyte accumulation in some but not all species. Drought-induced proline accumulation was not different in contrasting bean landraces or in Medicago truncatula and M. laciniata ecotypes characterised by contrasting drought tolerance. Therefore, proline does not seem to contribute significantly to osmotic stress tolerance in these legumes (Tari et al., 2008; Yousfi et al., 2010). In several

Author's personal copy 130

L. SZABADOS ET AL.

tropical legumes, water stress induced the accumulation of sugar alcohols and proline, while betaine contents were not changed. Tolerance to low water potential correlated with pinitol accumulation, suggesting that this sugar alcohol contributes to drought tolerance in these species (Ford, 1984). In contrasting, chickpea (Cicer arietinum) genotypes the capacity to accumulate proline correlated with copper tolerance, suggesting that in some species proline can contribute to heavy metal tolerance (Singh et al., 2010). Metabolite analysis of potato cultivars with different drought tolerance showed that enhanced galactose, inositol and proline content in drought-stressed plants correlate with higher degree of tolerance (Evers et al., 2010). The above listed examples confirm the importance of osmoprotectants for stress adaptation not only in extremophiles but in other species as well, including in crops. The elucidation of the importance of osmolyte accumulation in extremophiles became recently possible through comparative analysis of closely related halophyte and glycophyte species, especially those which are related to the model plant Arabidopsis (Orsini et al., 2010).The halophyte T. salsuginea (halophila) and L. crassifolium are close relatives of A. thaliana (a glycophyte), and can grow under extreme saline conditions. Proline contents in T. salsuginea and L. crassifolium are higher than those in Arabidopsis, even in optimal growth conditions, and proline accumulates to higher levels when salt stress is imposed (Fig. 2). Metabolite profiling revealed that in saline environments, proline is the dominant osmoprotective compound in T. salsuginea, followed by sugars, while betaines and polyols do not accumulate in this species (Arbona et al., 2010; Inan et al., 2004; Lugan et al., 2010). Sucrose, galactose and melibiose are among the carbohydrates that also accumulate at high levels in T. salsuginea (Lugan et al., 2010). Transcript analysis showed that the P5CS gene, which controls the glutamate-derived proline biosynthetic pathway, has higher constitutive expression level in T. salsuginea than in Arabidopsis, and is induced more rapidly under stress (Inan et al., 2004; Taji et al., 2004). High proline contents in T. salsuginea could also be the consequence of lower transcription of the ProDH gene, which controls proline catabolism (Kant et al., 2006). Such comparative analysis suggests that proline levels are key for the salt tolerance of T. salsuginea, which is controlled by the activities of its key biosynthetic and catabolic genes.

V. CONCLUSIONS The specific physiological responses elicited by various stresses on higher plants are relatively well described. The accumulation of osmoprotective compounds represents a specific metabolic response that is important to

Author's personal copy 131

PLANTS IN EXTREME ENVIRONMENTS

A

Arabidopsis

Thellungiella

Lepidium

Salt stress

Control B

Proline (mM/gFW)

400

300 Control

200

Salt 100

0 Arabidopsis

Thellungiella

Lepidium

Plants

Fig. 2. Comparison of salt tolerance and proline accumulation of Arabidopsis thaliana and two close relatives: Thellungiella salsuginea and Lepidium crassifolium. (A) Plants were grown in soil for 4 weeks and then watered twice a week with water or 0.5 M NaCl for 2 weeks. (B) Relative proline accumulation in the three brassicaceae species after 1 week of salt stress. Note that both halophytes contained more free proline in the absence of salt stress and accumulated more proline than Arabidopsis under saline conditions.

withstand harmful conditions. Although osmoprotective compounds belong to various categories of bioactive chemicals, they have similar cellular functions such as the stabilisation of cellular structures, protein complexes or specific enzymes or the control of redox balance and ROS production. Sugars or proline could function as metabolic signals and therefore have broader influence on physiological responses and metabolic adjustment to stress conditions.

Author's personal copy 132

L. SZABADOS ET AL.

Despite of the enormous amount of information accumulated in the past decades, the exact function of low-molecular-weight protective compounds in the adaptation to extreme environmental conditions is still not completely understood. Combining genomic, proteomic and metabolic profiling approaches could increase our understanding of plant stress responses on a global scale and of the metabolic bases of adaptation to drought, salinity or extreme temperatures. Engineering of model and crop plans via genetic transformation is a promising tool to study the significance of osmoprotective compounds in stress responses and to improve the performance of crop plants under suboptimal conditions. Enhanced accumulation of a metabolite can be achieved via activation of the biosynthetic pathway or inhibition of the catabolic pathway. Alternatively, novel pathways can be established in plants, by introducing genes from other species. Various examples have been published for engineering the synthesis and accumulation of osmoprotective compounds. Although osmoprotectant levels in such transgenic plants are often low, and increase in stress tolerance is small, manipulation of the metabolism of osmoprotective compounds have already produced promising results (Chen and Murata, 2008; Rontein et al., 2002; Szabados and Savoure, 2010; Verbruggen and Hermans, 2008). Adaptation to extreme environmental conditions is a complex trait, controlled by numerous genes. Engineering several traits by simultaneous manipulation of two or more genes is a real alternative to handle complex traits such as stress tolerance. Significant enhancement of drought tolerance have been achieved by pyramiding of transgenes in maize, by combining codA from E. coli and TsVP (V-H(þ)-PPase) from Thellungiella. Expression of the two transgenes led to higher GB accumulation and enhanced H(þ)-PPase activity compared with the parental lines, resulting in lower cell damage and higher yields under drought conditions (Wei et al., 2010a,b). Simultaneous engineering of osmolyte accumulation and various other traits by transgenic technology therefore is a feasible strategy to improve abiotic stress tolerance in crops.

ACKNOWLEDGEMENTS Authors are indebted to Ilse Kranner, Csaba Papdi and Laura Zsigmond for proofreading and correcting the chapter. Research and this publication were supported by OTKA grant K-68226, Cross-Border Cooperation Programme HURO/0801/167 and COST Action FA0901.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

133

REFERENCES Abebe, T. et al. (2003). Tolerance of mannitol-accumulating transgenic wheat to water stress and salinity. Plant Physiology 131(4), 1748–1755. ´ braha´m, E. et al. (2003). Light-dependent induction of proline biosynthesis by A abscisic acid and salt stress is inhibited by brassinosteroid in Arabidopsis. Plant Molecular Biology 51(3), 363–372. Abreu, E. F. and Aragao, F. J. (2007). Isolation and characterization of a myoinositol-1-phosphate synthase gene from yellow passion fruit (Passiflora edulis f. flavicarpa) expressed during seed development and environmental stress. Annals of Botany 99(2), 285–292. Ahuja, I., de Vos, Ric C. H., Bones, Atle M. and Hall, Robert D. (2010). Plant molecular stress responses face climate change. Trends in Plant Science 15, 664–674. Allard, F., Houde, M., Kro¨l, M., Ivanov, A., Huner, N. P. A. and SArhan, F. (1998). Betaine improves freezing tolerance in wheat. Plant and Cell Physiology 39, 1194–1202. Alpert, P. (2006). Constraints of tolerance: Why are desiccation-tolerant organisms so small or rare? The Journal of Experimental Biology 209(Pt. 9), 1575–1584. Amtmann, A. (2009). Learning from evolution: Thellungiella generates new knowledge on essential and critical components of abiotic stress tolerance in plants. Molecular Plant 2(1), 3–12. Amtmann, A., Bohnert, H. J. and Bressan, R. A. (2005). Abiotic stress and plant genome evolution. Search for new models. Plant Physiology 138(1), 127–130. Arbona, V., Argamasilla, R. and Gomez-Cadenas, A. (2010). Common and divergent physiological, hormonal and metabolic responses of Arabidopsis thaliana and Thellungiella halophila to water and salt stress. Journal of Plant Physiology 167, 1342–1350. Ashraf, M. and Harris, P. J. C. (2004). Potential biochemical indicators of salinity tolerance in plants. Plant Science 166, 3–16. Auton, M. and Bolen, D. W. (2004). Additive transfer free energies of the peptide backbone unit that are independent of the model compound and the choice of concentration scale. Biochemistry 43(5), 1329–1342. Auton, M. and Bolen, D. W. (2005). Predicting the energetics of osmolyte-induced protein folding/unfolding. Proceedings of the National Academy of Sciences of the United States of America 102(42), 15065–15068. Baena-Gonzalez, E. and Sheen, J. (2008). Convergent energy and stress signaling. Trends in Plant Science 13(9), 474–482. Bajji, M., Lutts, S. and Kinet, J. (2001). Water deficit effects on solute contribution to osmotic adjustment as a function of leaf ageing in three durum wheat (Triticum durum Desf.) cultivars performing differently in arid conditions. Plant Science 160(4), 669–681. Banu, N. A. et al. (2009). Proline and glycinebetaine induce antioxidant defense gene expression and suppress cell death in cultured tobacco cells under salt stress. Journal of Plant Physiology 166(2), 146–156. Bartels, D. and Sunkar, R. (2005). Drought and salt tolerance in plants. Critical Reviews in Plant Sciences 24, 23–58. Ben Hassine, A. et al. (2008). An inland and a coastal population of the Mediterranean xero-halophyte species Atriplex halimus L. differ in their ability to accumulate proline and glycinebetaine in response to salinity and water stress. Journal of Experimental Botany 59(6), 1315–1326.

Author's personal copy 134

L. SZABADOS ET AL.

Blazquez, M. A. et al. (1998). Isolation and molecular characterization of the Arabidopsis TPS1 gene, encoding trehalose-6-phosphate synthase. The Plant Journal 13(5), 685–689. Blum, A., Munns, R., Passioura, J. B., Turner, N. C., Sharp, R. E., Boyer, J. S., Nguyen, H. T., Hsiao, T. C., Verma, D. P. S. Hong, Z. et al. (1996). Genetically engineered plants resistant to soil drying and salt stress: How to interpret osmotic relations? Plant Physiology 110(4), 1051–1053. Bohnert, H. J. and Sheveleva, E. (1998). Plant stress adaptations–making metabolism move. Current Opinion in Plant Biology 1(3), 267–274. Bohnert, H. J., Nelson, D. E. and Jensen, R. G. (1995). Adaptations to environmental stresses. The Plant Cell 7(7), 1099–1111. Bolen, D. W. (2001). Protein stabilization by naturally occurring osmolytes. Methods in Molecular Biology 168, 17–36. Bolen, D. W. and Baskakov, I. V. (2001). The osmophobic effect: Natural selection of a thermodynamic force in protein folding. Journal of Molecular Biology 310 (5), 955–963. Bouchabke, O., Chang, F., Simon, M., Voisin, R., Pelletier, G. Durand-Tardif, M. et al. (2008). Natural variation in Arabidopsis thaliana as a tool for highlighting differential drought responses. PLoS ONE 3(2), e1705. Boyer, J. S. (1982). Plant productivity and environment. Science 218(4571), 443–448. Briens, M. and Larher, F. (1982). Osmoregulation in halophytic higher plants: A comparative study of soluble carbohydrates, polyols, betaines and free proline. Plant, Cell & Environment 5, 287–292. Brodmann, A., Schuller, A., Ludwig-Mu¨ller, J., Aeschbacher, R. A., Wiemken, A., Boller, T. Wingler, A. et al. (2002). Induction of trehalase in Arabidopsis plants infected with the trehalose-producing pathogen Plasmodiophora brassicae. Molecular Plant Microbe Interaction 15(7), 693–700. Brosche, M., Merilo, E., Mayer, F., Pechter, P., Puzo˜rjova, I., Brader, G., Kangasja¨rvi, J. Kollist, H. et al. (2010). Natural variation in ozone sensitivity among Arabidopsis thaliana accessions and its relation to stomatal conductance. Plant, Cell & Environment 33(6), 914–925. Brownfield, D. L., Todd, C. D. and Deyholos, M. K. (2008). Analysis of Arabidopsis arginase gene transcription patterns indicates specific biological functions for recently diverged paralogs. Plant Molecular Biology 67(4), 429–440. Browse, J. and Lange, B. M. (2004). Counting the cost of a cold-blooded life: Metabolomics of cold acclimation. Proceedings of the National Academy of Sciences of the United States of America 101(42), 14996–14997. Bundy, J. G., Davey, M. P. and Viant, M. R. (2009). Environmental metabolomics: A critical review and future perspectives. Metabolomics 5, 3–21. Burg, M. B. (1995). Molecular basis of osmotic regulation. The American Journal of Physiology 268(6 Pt 2), F983–F996. Burnet, M., Lafontaine, P. J. and Hanson, A. D. (1995). Assay, purification, and partial characterization of choline monooxygenase from spinach. Plant Physiology 108(2), 581–588. Castonguay, Y. et al. (1994). A cold-induced gene from Medicago sativa encodes a bimodular protein similar to developmentally regulated proteins. Plant Molecular Biology 24(5), 799–804. Chaves, M. M., Pereira, J. S. and Maroco, J. (2003). Understanding plant response to drought—From genes to the whole plant. Functional Plant Biology 30, 239–264. Chen, T. H. and Murata, N. (2002). Enhancement of tolerance of abiotic stress by metabolic engineering of betaines and other compatible solutes. Current Opinion in Plant Biology 5(3), 250–257.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

135

Chen, T. H. and Murata, N. (2008). Glycinebetaine: An effective protectant against abiotic stress in plants. Trends in Plant Science 13(9), 499–505. Chen, M. et al. (2007a). Expression of Bacillus subtilis proBA genes and reduction of feedback inhibition of proline synthesis increases proline production and confers osmotolerance in transgenic Arabidopsis. Journal of Biochemistry and Molecular Biology 40(3), 396–403. Chen, Z. et al. (2007b). Compatible solute accumulation and stress-mitigating effects in barley genotypes contrasting in their salt tolerance. Journal of Experimental Botany 58(15–16), 4245–4255. Choudhary, N. L., Sairam, R. K. and Tyagi, A. (2005). Expression of delta1-pyrroline5-carboxylate synthetase gene during drought in rice (Oryza sativa L.). Indian Journal of Biochemistry & Biophysics 42(6), 366–370. Cook, D., Sarah, F., Oliver, F. Thomashow, M. F. et al. (2004). A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 101(42), 15243–15248. Cramer, G. R., Ergu¨l, A., Grimplet, J., Tillett, R. L., Tattersall, E. A., Bohlman, M. C., Vincent, D., Sonderegger, J., Evans, J., Osborne, C., Quilici, D. Schlauch, K. A. et al. (2007). Water and salinity stress in grapevines: Early and late changes in transcript and metabolite profiles. Functional and Integrative Genomics 7(2), 111–134. Cross, J. M., Krasensky, J., Dal Santo, S., Kopka, J. Jonak, C. et al. (2006). Variation of enzyme activities and metabolite levels in 24 Arabidopsis accessions growing in carbon-limited conditions. Plant Physiology 142(4), 1574–1588. Crowe, J. H. (2007). Trehalose as a ‘‘chemical chaperone’’: Fact and fantasy. Advances in Experimental Medicine and Biology 594, 143–158. Crowe, J. H., Crowe, L. M. and Chapman, D. (1984). Preservation of membranes in anhydrobiotic organisms: The role of trehalose. Science 223(4637), 701–703. Crowe, J. H., Hoekstra, F. A. and Crowe, L. M. (1992). Anhydrobiosis. Annual Review of Physiology 54, 579–599. Csonka, L. N. (1989). Physiological and genetic responses of bacteria to osmotic stress. Microbiological Reviews 53, 121–147. Csonka, L. N. et al. (1988). Nucleotide sequence of a mutation in the proB gene of Escherichia coli that confers proline overproduction and enhanced tolerance to osmotic stress. Gene 64(2), 199–205. Cuin, T. A. and Shabala, S. (2007). Compatible solutes reduce ROS-induced potassium efflux in Arabidopsis roots. Plant, Cell & Environment 30(7), 875–885. De Ronde, J. A. et al. (2004). Photosynthetic response of transgenic soybean plants, containing an Arabidopsis P5CR gene, during heat and drought stress. Journal of Plant Physiology 161(11), 1211–1224. Deepak, S., Shailasree, S., Kini, R. K., Muck, A., Mithfer, A. and Shetty, S. H. (2010). Hydroxyproline-rich glycoproteins and plant defence. Journal of Phytopathology 158, 585–593. Delauney, A. J. and Verma, D. P. S. (1993). Proline biosynthesis and osmoregulation in plants. The Plant Journal 4, 215–223. Demiral, T. and Turkan, I. (2004). Does exogenous glycinebetaine affect antioxidative system of rice seedlings under NaCl treatment? Journal of Plant Physiology 161(10), 1089–1100. DeRidder, B. P. and Crafts-Brandner, S. J. (2008). Chilling stress response of postemergent cotton seedlings. Physiologia Plantarum 134(3), 430–439. Deuschle, K. et al. (2001). A nuclear gene encoding mitochondrial Delta-pyrroline-5carboxylate dehydrogenase and its potential role in protection from proline toxicity. The Plant Journal 27(4), 345–356.

Author's personal copy 136

L. SZABADOS ET AL.

Deuschle, K. et al. (2004). The role of [Delta]1-pyrroline-5-carboxylate dehydrogenase in proline degradation. The Plant Cell 16(12), 3413–3425. Deutch, C. E. and Winicov, I. (1995). Post-transcriptional regulation of a salt-inducible alfalfa gene encoding a putative chimeric proline-rich cell wall protein. Plant Molecular Biology 27(2), 411–418. Dominguez-Ferreras, A., Soto, M. J., Pe´rez-Arnedo, R., Olivares, J. Sanjua´n, J. et al. (2009). Importance of trehalose biosynthesis for Sinorhizobium meliloti Osmotolerance and nodulation of Alfalfa roots. Journal of Bacteriology 191(24), 7490–7499. Drennan, P. M., Smith, M. T., Goldsworth, D. and Van Staden, J. (1993). The occurence of trehalose in the leaves of the desiccation-tolerant angiosperm Myrothamnus flabellifolius welw. Journal of Plant Physiology 142, 493–496. Einset, J., Nielsen, E., Conoly, E. L., Bones, A., Sparstad, T., Winge, P. and Zhu, J.-K. (2007). Membrane-trafficking RabA4c involved in the effect of glycine betaine on recovery from chilling stress in Arabidopsis. Physiologia Plantarum 130, 511–518. Einset, J., Winge, P., Bones, A. M. Connolly, E. L. et al. (2008). The FRO2 ferric reductase is required for glycine betaine’s effect on chilling tolerance in Arabidopsis roots. Plant Physiology 134(2), 334–341. Etterson, J. R. and Shaw, R. G. (2001). Constraint to adaptive evolution in response to global warming. Science 294(5540), 151–154. Evers, D., Lefe`vre, I., Legay, S., Lamoureux, D., Hausman, J.-F., Rosales, R. O., Gutierrez, M., Tincopa, L. R., Hoffmann, L., Bonierbale, M. Schafleitner, R. et al. (2010). Identification of drought-responsive compounds in potato through a combined transcriptomic and targeted metabolite approach. Journal of Experimental Botany 61(9), 2327–2343. Fabro, G. et al. (2004). Proline accumulation and AtP5CS2 gene activation are induced by plant-pathogen incompatible interactions in Arabidopsis. Molecular Plant Microbe Interaction 17(4), 343–350. Farı´as-Rodrı´guez, R., Mellor, R. B., Arias, C. and Pen˜a-Cabriales, J. J. (1998). The accumulation of trehalose in nodules of several cultivars of common bean (Phaseolus vulgaris) and its correlation with resistance to drought stress. Physiologia Plantarum 102, 353–359. Fernandez, O., Be´thencourt, L., Quero, A., Sangwan, R. S. Cle´ment, C. et al. (2010). Trehalose and plant stress responses: Friend or foe? Trends Plant Science 15 (7), 409–417. Fiehn, O., Kopka, J., Do¨rmann, P., Altmann, T., Trethewey, R. N. Willmitzer, L. et al. (2000). Metabolite profiling for plant functional genomics. Nature Biotechnology 18(11), 1157–1161. Fitzgerald, T. L., Waters, D. L. and Henry, R. J. (2009). Betaine aldehyde dehydrogenase in plants. Plant Biology 11(2), 119–130. Ford, C. W. (1984). Accumulation of low molecular weight solutes in water-stressed tropical legumes. Phytochemistry 23, 1007–1015. Forlani, G., Scainelli, D. and Nielsen, E. (1997). [delta]1-pyrroline-5-carboxylate dehydrogenase from cultured cells of potato (purification and properties). Plant Physiology 113(4), 1413–1418. Fougere, F., Le Rudulier, D. and Streeter, J. G. (1991). Effects of salt stress on amino acid, organic acid, and carbohydrate composition of roots, bacteroids, and cytosol of Alfalfa (Medicago sativa L.). Plant Physiology 96(4), 1228–1236. Fowler, T. J., Bernhardt, C. and Tierney, M. L. (1999). Characterization and expression of four proline-rich cell wall protein genes in Arabidopsis encoding two distinct subsets of multiple domain proteins. Plant Physiology 121(4), 1081–1092.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

137

Frison, M. et al. (2007). The Arabidopsis thaliana trehalase is a plasma membranebound enzyme with extracellular activity. FEBS Letters 581(21), 4010–4016. Fujita, T. et al. (2003). Identification of regions of the tomato gamma-glutamyl kinase that are involved in allosteric regulation by proline. The Journal of Biological Chemistry 278(16), 14203–14210. Fujiwara, T., Hori, K., Ozaki, K., Yokota, Y., Mitsuya, S., Ichiyanagi, T., Hattori, T. Takabe, T. et al. (2008). Enzymatic characterization of peroxisomal and cytosolic betaine aldehyde dehydrogenases in barley. Plant Physiology 134 (1), 22–30. Funck, D., Stadelhofer, B. and Koch, W. (2008). Ornithine-delta-aminotransferase is essential for arginine catabolism but not for proline biosynthesis. BMC Plant Biology 8, 40. Gagneul, D., Aı¨nouche, A., Duhaze´, C., Lugan, R., Larher, F. R. Bouchereau, A. et al. (2007). A reassessment of the function of the so-called compatible solutes in the halophytic plumbaginaceae Limonium latifolium. Plant Physiology 144(3), 1598–1611. Garcia, A. B. et al. (1997). Effects of osmoprotectants upon NaCl stress in rice. Plant Physiology 115(1), 159–169. Garg, A. K., Kim, J.-K., Owens, T. G., Ranwala, A. P., Choi, Y. D., Kochian, L. V. Wu, R. J. et al. (2002). Trehalose accumulation in rice plants confers high tolerance levels to different abiotic stresses. Proceedings of the National Academy of Sciences of the United States of America 99(25), 15898–15903. Ghassemian, M., Lutes, J., Chang, H. S., Lange, I., Chen, W., Zhu, T., Wang, X. Lange, B. M. et al. (2008). Abscisic acid-induced modulation of metabolic and redox control pathways in Arabidopsis thaliana. Phytochemistry 69(17), 2899–2911. Ghosh Dastidar, K. et al. (2006). An insight into the molecular basis of salt tolerance of L-myo-inositol 1-P synthase (PcINO1) from Porteresia coarctata (Roxb.) Tateoka, a halophytic wild rice. Plant Physiology 140(4), 1279–1296. Ginzberg, I. et al. (1998). Isolation and characterization of two different cDNAs of delta1-pyrroline-5-carboxylate synthase in alfalfa, transcriptionally induced upon salt stress. Plant Molecular Biology 38(5), 755–764. Goddijn, O. J., Verwoerd, T. C., Voogd, E., Krutwagen, R. W., de Graaf, P. T., van Dun, K., Poels, J., Ponstein, A. S., Damm, B. Pen, J. et al. (1997). Inhibition of trehalase activity enhances trehalose accumulation in transgenic plants. Plant Physiology 113(1), 181–190. Gong, Q., Li, P., Ma, S., Indu Rupassara, S. Bohnert, H. J. et al. (2005). Salinity stress adaptation competence in the extremophile Thellungiella halophila in comparison with its relative Arabidopsis thaliana. The Plant Journal 44(5), 826–839. Goo, J. H. et al. (1999). Selection of Arabidopsis genes encoding secreted and plasma membrane proteins. Plant Molecular Biology 41(3), 415–423. Goodwin, W., Pallas, J. A. and Jenkins, G. I. (1996). Transcripts of a gene encoding a putative cell wall-plasma membrane linker protein are specifically coldinduced in Brassica napus. Plant Molecular Biology 31(4), 771–781. Greenway, H. and Munns, R. (1980). Mechanisms of salt tolerance in nonhalophytes. Annual Review of Plant Physiology 31, 149–190. Gregory, P. J., Ingram, J. S. and Brklacich, M. (2005). Climate change and food security. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences 360(1463), 2139–2148. Guo, Y., Zhang, L., Xiao, G., Cao, S., Gu, D., Tian, W. Chen, S. et al. (1997). Expression of betaine aldehyde dehydrogenase gene and salinity tolerance in rice transgenic plants. Science in China. Series C: Life Sciences 40(5), 496–501.

Author's personal copy 138

L. SZABADOS ET AL.

Guo, P., Baum, M., Grando, S., Salvatore, C., Guihua, B., Li, R., Maria, V. K., Varshney, R. K., Andreas, G. Valkoun, J. et al. (2009). Differentially expressed genes between drought-tolerant and drought-sensitive barley genotypes in response to drought stress during the reproductive stage. Journal of Experimental Botany 60(12), 3531–3544. Gussin, A. E. et al. (1969). Trehalase: A new pollen enzyme. Plant Physiology 44(8), 1163–1168. Guy, C. L. (1990). Cold acclimation and freezing stress tolerance: Role of protein metabolism. Annual Review of Plant Physiology and Plant Molecular Biology 41, 187–223. Guy, C., Kaplan, F., Kopka, J., Selbig, J. and Hincha, D. K. (2008). Metabolomics of temperature stress. Plant Physiology 132(2), 220–235. Hannah, M. A., Dana, W., Susanne, F., Oliver, F., Heyer, A. G. Hincha, D. K. et al. (2006). Natural genetic variation of freezing tolerance in Arabidopsis. Plant Physiology 142(1), 98–112. Hanson, A. D. and Nelsen, C. E. (1978). Betaine accumulation and [C]formate metabolism in water-stressed barley leaves. Plant Physiology 62(2), 305–312. Hanson, A. D., Rhodes, D. and Tracer, C. (1983). Evidence for synthesis of choline and betaine via phosphoryl base intermediates in salinized sugarbeet leaves. Plant Physiology 71(3), 692–700. Hanson, A. D., Rathinasabapathi, B., Chamberlin, B. Gage, D. A. et al. (1991). Comparative physiological evidence that beta-Alanine Betaine and Choline-O-Sulfate act as compatible osmolytes in halophytic Limonium species. Plant Physiology 97(3), 1199–1205. Hanson, A. D., Rathinasabapathi, B., Rivoal, J., Burnet, M., Dillon, M. O. Gage, D. A. et al. (1994). Osmoprotective compounds in the Plumbaginaceae: A natural experiment in metabolic engineering of stress tolerance. Proceedings of the National Academy of Sciences of the United States of America 91(1), 306–310. Hanson, J. et al. (2008). The sucrose regulated transcription factor bZIP11 affects amino acid metabolism by regulating the expression of ASPARAGINE SYNTHETASE1 and PROLINE DEHYDROGENASE2. The Plant Journal 53(6), 935–949. Hare, P. and Cress, W. (1997). Metabolic implications of stress induced proline accumulation in plants. Plant Growth Regulation 21, 79–102. Hare, P., Cress, W. and van Staden, J. (1998). Dissecting the roles of osmolyte accumulation during stress. Plant, Cell & Environment 21, 535–553. Hasegawa, P. M., Hasegawa, P. M., Bressan, R. A., Zhu, J. K. Bohnert, H. J. et al. (2000). Plant cellular and molecular responses to high salinity. Annual Reviews of Plant Physiology and Plant Molecular Biology 51, 463–499. Haudecoeur, E. et al. (2009). Proline antagonizes GABA-induced quenching of quorum-sensing in Agrobacterium tumefaciens. Proceedings of the National Academy of Sciences of the United States of America 106(34), 14587–14592. Holmstrom, K. O., Welin, B., Mandal, A., Kristiansdottir, I., Teeri, T. H., Lamark, T., Strøm, A. R. Palva, E. T. et al. (1994). Production of the Escherichia coli betaine-aldehyde dehydrogenase, an enzyme required for the synthesis of the osmoprotectant glycine betaine, in transgenic plants. The Plant Journal 6(5), 749–758. Holthauzen, L. M. and Bolen, D. W. (2007). Mixed osmolytes: The degree to which one osmolyte affects the protein stabilizing ability of another. Protein Science 16(2), 293–298. Hong, Z., Lakkineni, K., Zhang, Z. Verma, D. P. et al. (2000). Removal of feedback inhibition of delta(1)-pyrroline-5-carboxylate synthetase results in increased

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

139

proline accumulation and protection of plants from osmotic stress. Plant Physiology 122(4), 1129–1136. Hoque, M. A. et al. (2008). Proline and glycinebetaine enhance antioxidant defense and methylglyoxal detoxification systems and reduce NaCl-induced damage in cultured tobacco cells. Journal of Plant Physiology 165(8), 813–824. Hu, C. A., Delauney, A. J. and Verma, D. P. (1992). A bifunctional enzyme (delta 1-pyrroline-5-carboxylate synthetase) catalyzes the first two steps in proline biosynthesis in plants. Proceedings of the National Academy of Sciences of the United States of America 89(19), 9354–9358. Hu, L. et al. (2005). Overexpression of mtlD gene in transgenic Populus tomentosa improves salt tolerance through accumulation of mannitol. Tree Physiology 25(10), 1273–1281. Huang, A. H. and Cavalieri, A. J. (1979). Proline oxidase and water stress-induced proline accumulation in spinach leaves. Plant Physiology 63(3), 531–535. Inan, G. et al. (2004). Salt cress. A halophyte and cryophyte Arabidopsis relative model system and its applicability to molecular genetic analyses of growth and development of extremophiles. Plant Physiology 135(3), 1718–1737. Ingram, J. and Bartels, D. (1996). The molecular basis of dehydration tolerance in plants. Annual Review of Plant Physiology and Plant Molecular Biology 47, 377–403. Iordachescu, M. and Imai, R. (2008). Trehalose biosynthesis in response to abiotic stresses. Journal of Integrative Plant Biology 50(10), 1223–1229. Ishitani, M. et al. (1996). Coordinate transcriptional induction of myo-inositol metabolism during environmental stress. The Plant Journal 9(4), 537–548. Islam, M. M. et al. (2009). Exogenous proline and glycinebetaine increase antioxidant enzyme activities and confer tolerance to cadmium stress in cultured tobacco cells. Journal of Plant Physiology 166(15), 1587–1597. Jaindl, M. and Popp, M. (2006). Cyclitols protect glutamine synthetase and malate dehydrogenase against heat induced deactivation and thermal denaturation. Biochemical and Biophysical Research Communications 345(2), 761–765. Jakoby, M. et al. (2002). bZIP transcription factors in Arabidopsis. Trends in Plant Science 7(3), 106–111. Jang, I. C., Oh, S. J., Seo, J. S., Choi, W. B., Song, S. I., Kim, C. H., Kim, Y. S., Seo, H. S., Choi, Y. D., Nahm, B. H. Kim, J. K. et al. (2003). Expression of a bifunctional fusion of the Escherichia coli genes for trehalose-6-phosphate synthase and trehalose-6-phosphate phosphatase in transgenic rice plants increases trehalose accumulation and abiotic stress tolerance without stunting growth. Plant Physiology 131(2), 516–524. Johnson, M. D. and Sussex, I. M. (1995). 1 L-myo-inositol 1-phosphate synthase from Arabidopsis thaliana. Plant Physiology 107, 613–619. Johnson, H. E., Broadhurst, D., Goodacre, R. Smith, A. R. et al. (2003). Metabolic fingerprinting of salt-stressed tomatoes. Phytochemistry 62(6), 919–928. Jolivet, Y., Larher, F. and Hamelin, J. (1982). Osmoregulation in halophytic higher plants: The protective effect of glycine betaine against the heat destabilization of membranes. Plant Science Letters 25, 193–201. Jones, G. P. et al. (2006). Occurrence and stress response of N-methylproline compounds in Tamarix species. Phytochemistry 67(2), 156–160. Jose-Estanyol, M., Gomis-Ruth, F. X. and Puigdomenech, P. (2004). The eightcysteine motif, a versatile structure in plant proteins. Plant Physiology and Biochemistry 42(5), 355–365. Kant, S. et al. (2006). Evidence that differential gene expression between the halophyte, Thellungiella halophila, and Arabidopsis thaliana is responsible for

Author's personal copy 140

L. SZABADOS ET AL.

higher levels of the compatible osmolyte proline and tight control of Naþ uptake in T. halophila. Plant, Cell & Environment 29(7), 1220–1234. Kaplan, F., Joachim, K., Haskell, D. W., Wei, Z., Cameron Schiller, K., Nicole, G., Sung, D. Y. Guy, C. L. et al. (2004). Exploring the temperature-stress metabolome of Arabidopsis. Plant Physiology 136(4), 4159–4168. Kaplan, F., Kopka, J., Sung, D. Y., Zhao, W., Popp, M., Porat, R. Guy, C. L. et al. (2007). Transcript and metabolite profiling during cold acclimation of Arabidopsis reveals an intricate relationship of cold-regulated gene expression with modifications in metabolite content. The Plant Journal 50(6), 967–981. Karim, S. et al. (2007). Improved drought tolerance without undesired side effects in transgenic plants producing trehalose. Plant Molecular Biology 64(4), 371–386. Kathuria, H., Giri, J., Nataraja, K. N., Murata, N., Udayakumar, M. Tyagi, A. K. et al. (2009). Glycinebetaine-induced water-stress tolerance in codA-expressing transgenic indica rice is associated with up-regulation of several stress responsive genes. Plant Biotechnology Journal 7(6), 512–526. Katori, T., Ikeda, A., Iuchi, S., Kobayashi, M., Shinozaki, K., Maehashi, K., Sakata, Y., Tanaka, S. Taji, T. et al. (2010). Dissecting the genetic control of natural variation in salt tolerance of Arabidopsis thaliana accessions. Journal of Experimental Botany 61(4), 1125–1138. Kavi Kishor, P. B. et al. (2005). Regulation of proline biosynthesis, degradation, uptake and transport in higher plants: Its implications in plant growth and abiotic stress tolerance. Current Science 88(3), 424–438. Kempa, S. et al. (2008). A central role of abscisic acid in stress-regulated carbohydrate metabolism. PLoS ONE 3(12), e3935. Kerepesi, I., Galiba, G. and Ba´nyai, E. (1998). Osmotic and salt stresses induced differential alteration in water-soluble carbohydrate content in wheat seedlings. Journal of Agricultural and Food Chemistry 46, 5347–5354. Khedr, A. H. et al. (2003). Proline induces the expression of salt-stress-responsive proteins and may improve the adaptation of Pancratium maritimum L. to salt-stress. Journal of Experimental Botany 54(392), 2553–2562. Kintisch, E. (2009). Global warming: Projections of climate change go from bad to worse, scientists report. Science 323(5921), 1546–1547. Kishor, P., Hong, Z., Miao, G. H., Hu, C. A. A. Verma, D. P. S. et al. (1995). Overexpression of [delta]-pyrroline-5-carboxylate synthetase increases proline production and confers osmotolerance in transgenic plants. Plant Physiology 108(4), 1387–1394. Kiyosue, T. et al. (1996). A nuclear gene encoding mitochondrial proline dehydrogenase, an enzyme involved in proline metabolism, is upregulated by proline but downregulated by dehydration in Arabidopsis. The Plant Cell 8(8), 1323–1335. Korn, M., Ga¨rtner, T., Erban, A., Kopka, J., Selbig, J. Hincha, D. K. et al. (2010). Predicting Arabidopsis freezing tolerance and heterosis in freezing tolerance from metabolite composition. Molecular Plant 3(1), 224–235. Kosmas, S. A., Argyrokastritis, A., Loukas, M. G., Eliopoulos, E., Tsakas, S. Kaltsikes, P. J. et al. (2006). Isolation and characterization of droughtrelated trehalose 6-phosphate-synthase gene from cultivated cotton (Gossypium hirsutum L.). Planta 223(2), 329–339. Kranner, I., Beckett, R. P., Wornik, S., Zorn, M. Pfeifhofer, H. W. et al. (2002). Revival of a resurrection plant correlates with its antioxidant status. The Plant Journal 31(1), 13–24.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

141

Ksouri, R., Wided, M., Hans-Werner, K. and Abdelly, C. (2010). Responses of halophytes to environmental stresses with special emphasis to salinity. Advances in Botanical Research 53, 117–145. Kumar, R. (2009). Role of naturally occurring osmolytes in protein folding and stability. Archives of Biochemistry and Biophysics 491(1–2), 1–6. Kumar, V., Shriram, V., Kishor, P. B. K., Jawali, N. and Shitole, M. G. (2010). Enhanced proline accumulation and salt stress tolerance of transgenic indica rice by over-expressing P5CSF129A gene. Plant Biotechnology Reports 4, 37–48. Larher, F. R., Lugan, R., Gagneul, D., Guyot, S., Monnier, C., Lespinase, Y. and Bouchereau, A. (2009). A reassessment of the prevalent organic solutes constitutively accumulated and potentially involved in osmotic adjustment in pear leaves. Environmental and Experimental Botany 66, 230–241. Lehmann, S. et al. (2010). Proline metabolism and transport in plant development. Amino Acids 39, 949–962. Lisec, J., Meyer, R. C., Steinfath, M., Redestig, H., Becher, M., Witucka-Wall, H., Fiehn, O., To¨rje´k, O., Selbig, J., Altmann, T. Willmitzer, L. et al. (2008). Identification of metabolic and biomass QTL in Arabidopsis thaliana in a parallel analysis of RIL and IL populations. The Plant Journal 53(6), 960–972. Liu, Y. and Bolen, D. W. (1995). The peptide backbone plays a dominant role in protein stabilization by naturally occurring osmolytes. Biochemistry 34(39), 12884–12891. Liu, M. S., Chien, C. T. and Lin, T. P. (2008). Constitutive components and induced gene expression are involved in the desiccation tolerance of Selaginella tamariscina. Plant and Cell Physiology 49(4), 653–663. Loescher, W. H. (1987). Physiology and metabolism of sugar alcohols in higher plants. Physiologia Plantarum 70, 553–557. Loescher, W. H. et al. (1992). Mannitol synthesis in higher plants: Evidence for the role and characterization of a NADPH-dependent mannose 6-phosphate reductase. Plant Physiology 98(4), 1396–1402. Lopez, M., Tejera, N. A., Iribarne, C., Lluch, C. Herrera-Cervera, J. A. et al. (2008). Trehalose and trehalase in root nodules of Medicago truncatula and Phaseolus vulgaris in response to salt stress. Physiologia Plantarum 134(4), 575–582. Lugan, R., Niogret, M. F., Kervazo, L., Larher, F. R., Kopka, J. Bouchereau, A. et al. (2009). Metabolome and water status phenotyping of Arabidopsis under abiotic stress cues reveals new insight into ESK1 function. Plant, Cell & Environment 32(2), 95–108. Lugan, R., Niogret, M. F., Leport, L., Gue´gan, J. P., Larher, F., Savoure´, A., Kopka, J. and Bouchereau, A. (2010). Metabolome and water homeostasis analysis of Thellungiella salsuginea suggests that dehydration tolerance is a key response to osmotic stress in this halophyte. The Plant Journal 64, 215–229. Majee, M. et al. (2004). A novel salt-tolerant L-myo-inositol-1-phosphate synthase from Porteresia coarctata (Roxb.) Tateoka, a halophytic wild rice: Molecular cloning, bacterial overexpression, characterization, and functional introgression into tobacco-conferring salt tolerance phenotype. The Journal of Biological Chemistry 279(27), 28539–28552. Majumder, A. L., Johnson, M. D. and Henry, S. A. (1997). 1L-myo-inositol-1phosphate synthase. Biochimica et Biophysica Acta 1348(1–2), 245–256.

Author's personal copy 142

L. SZABADOS ET AL.

Martinez, J. P. et al. (2004). Is osmotic adjustment required for water stress resistance in the Mediterranean shrub Atriplex halimus L? Journal of Plant Physiology 161(9), 1041–1051. Maruyama, K., Takeda, M., Kidokoro, S., Yamada, K., Sakuma, Y., Urano, K., Fujita, M., Yoshiwara, K., Matsukura, S., Morishita, Y., Sasaki, R. Suzuki, H. et al. (2009). Metabolic pathways involved in cold acclimation identified by integrated analysis of metabolites and transcripts regulated by DREB1A and DREB2A. Plant Physiology 150(4), 1972–1980. Matysik, J., Alia, A., Bhalu, B. and Mohanty, P. (2002). Molecular mechanisms of quenching of reactive oxygen species by proline under stress in plants. Current Science 82(5), 525–532. McNeil, S. D., Nuccio, M. L., Ziemak, M. J. Hanson, A. D. et al. (2001). Enhanced synthesis of choline and glycine betaine in transgenic tobacco plants that overexpress phosphoethanolamine N-methyltransferase. Proceedings of the National Academy of Sciences of the United States of America 98(17), 10001–10005. Mehta, S. K. and Gaur, J. P. (1999). Heavy-metal-induced proline accumulation and its role in ameliorating metal toxicity in Chlorella vulgaris. The New Phytologist 143, 253–259. Meyer, R. C., Matthias, S., Jan, L., Martina, B., Hanna, W.-W., Otto´, T., Oliver, F., ¨ nne, E., Lothar, W., Joachim, S. Thomas, A. et al. (2007). The metabolic A signature related to high plant growth rate in Arabidopsis thaliana. Proceedings of the National Academy of Sciences of the United States of America 104 (11), 4759–4764. Miller, G. et al. (2005). Responsive modes of Medicago sativa proline dehydrogenase genes during salt stress and recovery dictate free proline accumulation. Planta 222(1), 70–79. Miller, G. et al. (2009). Unraveling delta1-pyrroline-5-carboxylate-proline cycle in plants by uncoupled expression of proline oxidation enzymes. The Journal of Biological Chemistry 284(39), 26482–26492. Mishra, S. and Dubey, R. S. (2006). Inhibition of ribonuclease and protease activities in arsenic exposed rice seedlings: Role of proline as enzyme protectant. Journal of Plant Physiology 163(9), 927–936. Mittler, R. and Blumwald, E. (2010). Genetic engineering for modern agriculture: Challenges and perspectives. Annual Review of Plant Biology 61, 443–462. Mohanty, A., Kathuria, H., Ferjani, A., Sakamoto, A., Mohanty, P., Murata, N. Tyagi, A. K. et al. (2002). Transgenics of an elite indica rice variety Pusa Basmati 1 harbouring the codA gene are highly tolerant to salt stress. Theoretical and Applied Genetics 106(1), 51–57. Moore, J. P., Hearshaw, M., Ravenscroft, N., Lindsey, G. G., Farrant, J. M. and Brandt, W. F. (2007). Desiccation-induced ultrastructural and biochemical changes in the leaves of the resurrection plant Myrothamnus flabellifolia. Australian Journal of Botany 55, 482–491. Moore, J. P., Le, N. T., Brandt, W. F., Driouich, A. Farrant, J. M. et al. (2009). Towards a systems-based understanding of plant desiccation tolerance. Trends in Plant Science 14(2), 110–117. Moradi, F. and Ismail, A. M. (2007). Responses of photosynthesis, chlorophyll fluorescence and ROS-scavenging systems to salt stress during seedling and reproductive stages in rice. Annals of Botany 99(6), 1161–1173. Muller, J. et al. (2001). Trehalose and trehalase in Arabidopsis. Plant Physiology 125 (2), 1086–1093. Munns, R. (2002). Comparative physiology of salt and water stress. Plant, Cell & Environment 25(2), 239–250.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

143

Murakeozy, E. P., Nagy, Z., Duhaze´, C., Bouchereau, A. Tuba, Z. et al. (2003). Seasonal changes in the levels of compatible osmolytes in three halophytic species of inland saline vegetation in Hungary. Journal of Plant Physiology 160(4), 395–401. Nakamura, T., Yokota, S., Muramoto, Y., Tsutsui, K., Oguri, Y., Fukui, K. Takabe, T. et al. (1997). Expression of a betaine aldehyde dehydrogenase gene in rice, a glycinebetaine nonaccumulator, and possible localization of its protein in peroxisomes. The Plant Journal 11(5), 1115–1120. Nelson, D. E., Koukoumanos, M. and Bohnert, H. J. (1999). Myo-inositol-dependent sodium uptake in ice plant. Plant Physiology 119(1), 165–172. N’Guyen, A. and Lamant, A. (1988). Pinitol and myo-inositol accumulation in waterstressed seedling of maritime pine. Phytochemistry 27, 3423–3427. Noctor, G. (2006). Metabolic signalling in defence and stress: The central roles of soluble redox couples. Plant, Cell & Environment 29(3), 409–425. Nuccio, M. L., Russell, B. L., Nolte, K. D., Rathinasabapathi, B., Gage, D. A. Hanson, A. D. et al. (1998). The endogenous choline supply limits glycine betaine synthesis in transgenic tobacco expressing choline monooxygenase. The Plant Journal 16(4), 487–496. Ohnishi, N. and Murata, N. (2006). Glycinebetaine counteracts the inhibitory effects of salt stress on the degradation and synthesis of D1 protein during photoinhibition in Synechococcus sp. PCC 7942. Plant Physiology 141(2), 758–765. Oono, Y., Seki, M., Nanjo, T., Narusaka, M., Fujita, M., Satoh, R., Satou, M., Sakurai, T., Ishida, J., Akiyama, K., Iida, K. Maruyama, K. et al. (2003). Monitoring expression profiles of Arabidopsis gene expression during rehydration process after dehydration using ca 7000 full-length cDNA microarray. The Plant Journal 34(6), 868–887. Orsini, F., Matilde Paino, D. U., Gunsu, I., Sara, S., Dong-Ha, O., Mickelbart, M. V., Federica, C., Xia, L., Jae Cheol, J., Dae-Jin, Y., Bohnert, H. J. Bressan, R. A. et al. (2010). A comparative study of salt tolerance parameters in 11 wild relatives of Arabidopsis thaliana. Journal of Experimental Botany 61(13), 3787–3798. Papageorgiou, G. C. and Murata, N. (1995). The unusually strong stabilizing effects of glycine betaine on the structure and function of the oxygen-evolving Photosystem II complex. Photosynthesis Research 44, 243–252. Parida, A. K. and Das, A. B. (2005). Salt tolerance and salinity effects on plants: A review. Ecotoxicology and Environmental Safety 60(3), 324–349. Park, E. J., Jeknic´, Z., Sakamoto, A., DeNoma, J., Yuwansiri, R., Murata, N. Chen, T. H. et al. (2004). Genetic engineering of glycinebetaine synthesis in tomato protects seeds, plants, and flowers from chilling damage. The Plant Journal 40(4), 474–487. Park, E. J., Jeknic, Z. and Chen, T. H. (2006). Exogenous application of glycinebetaine increases chilling tolerance in tomato plants. Plant and Cell Physiology 47(6), 706–714. Park, E. J., Jeknic´, Z., Chen, T. H. Murata, N. et al. (2007). The codA transgene for glycinebetaine synthesis increases the size of flowers and fruits in tomato. Plant Biotechnology Journal 5(3), 422–430. Parvanova, D., Ivanov, S., Konstantinova, T., Karanov, E., Atanassov, A., Tsvetkov, T., Alexieva, V. Djilianov, D. et al. (2004). Transgenic tobacco plants accumulating osmolytes show reduced oxidative damage under freezing stress. Plant Physiology and Biochemistry 42(1), 57–63. Patra, B. et al. (2010). Enhanced salt tolerance of transgenic tobacco plants by coexpression of PcINO1 and McIMT1 is accompanied by increased level of myo-inositol and methylated inositol. Protoplasma 245(1–4), 143–152.

Author's personal copy 144

L. SZABADOS ET AL.

Paul, M. J. et al. (2010). Upregulation of biosynthetic processes associated with growth by trehalose 6-phosphate. Plant Signaling and Behavior 5(4), 386–392. Pedras, M. S. and Zheng, Q. A. (2010). Metabolic responses of Thellungiella halophila/salsuginea to biotic and abiotic stresses: Metabolite profiles and quantitative analyses. Phytochemistry 71(5–6), 581–589. Peng, Z., Lu, Q. and Verma, D. P. (1996). Reciprocal regulation of delta 1-pyrroline5-carboxylate synthetase and proline dehydrogenase genes controls proline levels during and after osmotic stress in plants. Molecular & General Genetics 253(3), 334–341. Perez-Arellano, I. et al. (2010). Pyrroline-5-carboxylate synthase and proline biosynthesis: From osmotolerance to rare metabolic disease. Protein Science 19(3), 372–382. Phang, J. M., Liu, W. and Zabirnyk, O. (2010). Proline metabolism and microenvironmental stress. Annual Review of Nutrition 30, 441–463. Pinheiro, C., Passarinho, J. A. and Ricardo, C. P. (2004). Effect of drought and rewatering on the metabolism of Lupinus albus organs. Journal of Plant Physiology 161(11), 1203–1210. Pramanik, M. H. and Imai, R. (2005). Functional identification of a trehalose 6-phosphate phosphatase gene that is involved in transient induction of trehalose biosynthesis during chilling stress in rice. Plant Molecular Biology 58(6), 751–762. Premachandra, G. S., Hahn, G. T., Rhodes, D. and Joly, R. J. (1995). Leaf water relations and solute accumulation in two grain sorghum lines exhibiting contrasting drought tolerance. Journal of Experimental Botany 46, 1833–1841. Rammesmayer, G. et al. (1995). Characterization of IMT1, myo-inositol O-methyltransferase, from Mesembryanthemum crystallinum. Archives of Biochemistry and Biophysics 322(1), 183–188. Ratcliffe, R. G. and Shachar-Hill, Y. (2005). Revealing metabolic phenotypes in plants: Inputs from NMR analysis. Biological Reviews of the Cambridge Philosophical Society 80(1), 27–43. Rayapati, P. J. and Stewart, C. R. (1991). Solubilization of a proline dehydrogenase from maize (Zea mays L.) mitochondria. Plant Physiology 95(3), 787–791. Rayapati, P. J., Stewart, C. R. and Hack, E. (1989). Pyrroline-5-carboxylate reductase is in Pea (Pisum sativum L.) leaf chloroplasts. Plant Physiology 91(2), 581–586. Rhodes, D., Handa, S. and Bressan, R. A. (1986). Metabolic changes associated with adaptation of plant cells to water stress. Plant Physiology 82(4), 890–903. Ribarits, A. et al. (2007). Two tobacco proline dehydrogenases are differentially regulated and play a role in early plant development. Planta 225(5), 1313–1324. Rolland, F., Baena-Gonzalez, E. and Sheen, J. (2006). Sugar sensing and signaling in plants: Conserved and novel mechanisms. Annual Review of Plant Biology 57, 675–709. Romero, C., Belle´s, J. M., Vaya´, J. L., Serrano, R. Culia´n˜ez-Macia`, F. A. et al. (1997). Expression of the yeast trehalose-6-phosphate synthase gene in transgenic tobacco plants: Pleiotropic phenotypes include drought tolerance. Planta 201(3), 293–297. Rontein, D., Basset, G. and Hanson, A. D. (2002). Metabolic engineering of osmoprotectant accumulation in plants. Metabolic Engineering 4(1), 49–56. Roosens, N. H. et al. (1998). Isolation of the ornithine-delta-aminotransferase cDNA and effect of salt stress on its expression in Arabidopsis thaliana. Plant Physiology 117(1), 263–271.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

145

Rosa, M., Prado, C., Podazza, G., Interdonato, R., Gonza´lez, J. A., Hilal, M. Prado, F. E. et al. (2009). Soluble sugars—Metabolism, sensing and abiotic stress: A complex network in the life of plants. Plant Signaling and Behavior 4(5), 388–393. Rosgen, J. (2007). Molecular basis of osmolyte effects on protein and metabolites. Methods in Enzymology 428, 459–486. Roychoudhury, A. et al. (2008). Comparative physiological and molecular responses of a common aromatic indica rice cultivar to high salinity with non-aromatic indica rice cultivars. Plant Cell Reports 27(8), 1395–1410. Rumpho, M. E., Edwards, G. E. and Loescher, W. H. (1983). A pathway for photosynthetic carbon flow to mannitol in celery leaves: Activity and localization of key enzymes. Plant Physiology 73(4), 869–873. Saito, K. and Matsuda, F. (2010). Metabolomics for functional genomics, systems biology, and biotechnology. Annual Review of Plant Biology 61, 463–489. Sakamoto, A. and Murata, N. (2000). Genetic engineering of glycinebetaine synthesis in plants: Current status and implications for enhancement of stress tolerance. Journal of Experimental Botany 51(342), 81–88. Sakamoto, A., Alia, and Murata, N. (1998). Metabolic engineering of rice leading to biosynthesis of glycinebetaine and tolerance to salt and cold. Plant Molecular Biology 38(6), 1011–1019. Sanchez, D. H., Siahpoosh, M. R., Roessner, U., Udvardi, M. Kopka, J. et al. (2008). Plant metabolomics reveals conserved and divergent metabolic responses to salinity. Plant Physiology 132(2), 209–219. Saradhi, P. P. et al. (1995). Proline accumulates in plants exposed to UV radiation and protects them against UV induced peroxidation. Biochemical and Biophysical Research Communications 209(1), 1–5. Satoh, R. et al. (2002). ACTCAT, a novel cis-acting element for proline- and hypoosmolarity-responsive expression of the ProDH gene encoding proline dehydrogenase in Arabidopsis. Plant Physiology 130(2), 709–719. Satoh, R., Fujita, Y., Nakashima, K., Shinozaki, K. Yamaguchi-Shinozaki, K. et al. (2004). A novel subgroup of bZIP proteins functions as transcriptional activators in hypoosmolarity-responsive expression of the ProDH gene in Arabidopsis. Plant and Cell Physiology 45(3), 309–317. Schat, H., Sharma, S. S. and Vooijs, R. (1997). Heavy metal-induced accumulation of free proline in a metal-tolerant and a nontolerant ecotype of Silene vulgaris. Physiologia Plantarum 101, 477–482. Schauer, N. and Fernie, A. R. (2006). Plant metabolomics: Towards biological function and mechanism. Trends in Plant Science 11(10), 508–516. Schluepmann, H. et al. (2004). Trehalose mediated growth inhibition of Arabidopsis seedlings is due to trehalose-6-phosphate accumulation. Plant Physiology 135(2), 879–890. Seki, M., Umezawa, T., Urano, K. Shinozaki, K. et al. (2007). Regulatory metabolic networks in drought stress responses. Current Opinion in Plant Biology 10 (3), 296–302. Semel, Y., Schauer, N., Roessner, U., Zamir, D. and Fernie, A. R. (2007). Metabolite analysis for the comparison of irrigated and non-irrigated field grown tomato of varying genotype. Metabolomics 3, 289–295. Sengupta, S. and Majumder, A. L. (2009). Insight into the salt tolerance factors of a wild halophytic rice, Porteresia coarctata: A physiological and proteomic approach. Planta 229(4), 911–929. Sengupta, S. et al. (2008). Inositol methyl tranferase from a halophytic wild rice, Porteresia coarctata Roxb. (Tateoka): Regulation of pinitol synthesis under abiotic stress. Plant, Cell & Environment 31(10), 1442–1459.

Author's personal copy 146

L. SZABADOS ET AL.

Sharma, P. and Dubey, R. S. (2005). Modulation of nitrate reductase activity in rice seedlings under aluminium toxicity and water stress: Role of osmolytes as enzyme protectant. Journal of Plant Physiology 162(8), 854–864. Sharma, S. and Verslues, P. E. (2010). Mechanisms independent of abscisic acid (ABA) or proline feedback have a predominant role in transcriptional regulation of proline metabolism during low water potential and stress recovery. Plant, Cell & Environment 33(11), 1838–1851. Sharma, S. S., Schat, H. and Vooijs, R. (1998). In vitro alleviation of heavy metalinduced enzyme inhibition by proline. Phytochemistry 49(6), 1531–1535. Shen, Y. G. et al. (2002). AhCMO, regulated by stresses in Atriplex hortensis, can improve drought tolerance in transgenic tobacco. Theoretical and Applied Genetics 105(6–7), 815–821. Shirasawa, K., Takabe, T. and Kishitani, S. (2006). Accumulation of glycinebetaine in rice plants that overexpress choline monooxygenase from spinach and evaluation of their tolerance to abiotic stress. Annals of Botany 98(3), 565–571. Showalter, A. M. et al. (2010). A bioinformatics approach to the identification, classification, and analysis of hydroxyproline-rich glycoproteins. Plant Physiology 153(2), 485–513. Shulaev, V., Diego, C., Gad, M. Ron, M. et al. (2008). Metabolomics for plant stress response. Plant Physiology 132(2), 199–208. Shuman, J. L., Mittler, R. and Shulaev, V. (2006). Metabolite profiling in heat- and drought stressed Arabidopsis. Hortscience 41, 1057. Singh, V. et al. (2010). Proline improves copper tolerance in chickpea (Cicer arietinum). Protoplasma 245(1–4), 173–181. Siripornadulsil, S., Traina, S., Verma, D. P. Sayre, R. T. et al. (2002). Molecular mechanisms of proline-mediated tolerance to toxic heavy metals in transgenic microalgae. Plant Cell 14(11), 2837–2847. Smirnoff, N. and Cumbes, Q. J. (1989). Hydroxyl radical scavenging activity of compatible solutes. Phytochemistry 28(4), 1057–1060. Song, G. Q. et al. (2010). A novel mannose-based selection system for plant transformation using celery mannose-6-phosphate reductase gene. Plant Cell Reports 29(2), 163–172. Stiller, I. et al. (2008). Effects of drought on water content and photosynthetic parameters in potato plants expressing the trehalose-6-phosphate synthase gene of Saccharomyces cerevisiae. Planta 227(2), 299–308. Stitt, M. and Hurry, V. (2002). A plant for all seasons: Alterations in photosynthetic carbon metabolism during cold acclimation in Arabidopsis. Current Opinion in Plant Biology 5(3), 199–206. Stitt, M., Sulpice, R. and Keurentjes, J. (2010). Metabolic networks: How to identify key components in the regulation of metabolism and growth. Plant Physiology 152(2), 428–444. Street, T. O., Bolen, D. W. and Rose, G. D. (2006). A molecular mechanism for osmolyte-induced protein stability. Proceedings of the National Academy of Sciences of the United States of America 103(38), 13997–14002. Strizhov, N. et al. (1997). Differential expression of two P5CS genes controlling proline accumulation during salt-stress requires ABA and is regulated by ABA1, ABI1 and AXR2 in Arabidopsis. The Plant Journal 12(3), 557–569. Su, J., Hirji, R., Zhang, L., He, C., Selvaraj, G. Ray, W. et al. (2006). Evaluation of the stress-inducible production of choline oxidase in transgenic rice as a strategy for producing the stress-protectant glycine betaine. Journal of Experimental Botany 57(5), 1129–1135.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

147

Sumner, L. W., Mendes, P. and Dixon, R. A. (2003). Plant metabolomics: Large-scale phytochemistry in the functional genomics era. Phytochemistry 62(6), 817–836. Szabados, L. and Savoure, A. (2010). Proline: A multifunctional amino acid. Trends in Plant Science 15(2), 89–97. Sze´kely, G., Abraha´m, E., Cse´plo, A., Rigo´, G., Zsigmond, L., Csisza´r, J., Ayaydin, F., Strizhov, N., Ja´sik, J., Schmelzer, E., Koncz, C. Szabados, L. et al. (2008). Duplicated P5CS genes of Arabidopsis play distinct roles in stress regulation and developmental control of proline biosynthesis. The Plant Journal 53(1), 11–28. Szoke, A. et al. (1992). Subcellular location of delta-pyrroline-5-carboxylate reductase in root/nodule and leaf of soybean. Plant Physiology 99(4), 1642–1649. Taji, T. et al. (2004). Comparative genomics in salt tolerance between Arabidopsis and Arabidopsis-related halophyte salt cress using Arabidopsis microarray. Plant Physiology 135(3), 1697–1709. Takagi, H. (2008). Proline as a stress protectant in yeast: Physiological functions, metabolic regulations, and biotechnological applications. Applied Microbiology and Biotechnology 81(2), 211–223. Tang, W., Peng, X. and Newton, R. J. (2005). Enhanced tolerance to salt stress in transgenic loblolly pine simultaneously expressing two genes encoding mannitol-1-phosphate dehydrogenase and glucitol-6-phosphate dehydrogenase. Plant Physiology and Biochemistry 43(2), 139–146. Tarczynski, M. C., Jensen, R. G. and Bohnert, H. J. (1993). Stress protection of transgenic tobacco by production of the osmolyte mannitol. Science 259 (5094), 508–510. Tari, I. et al. (2008). Changes in chlorophyll fluorescence parameters and oxidative stress responses of bush bean genotypes for selecting contrasting acclimation strategies under water stress. Acta Biologica Hungarica 59(3), 335–345. Tester, M. and Bacic, A. (2005). Abiotic stress tolerance in grasses. From model plants to crop plants. Plant Physiology 137(3), 791–793. ¨ zkum, D. Tipirdamaz, R., Gagneul, D., Duhaze, C., Aı¨nouche, A., Monnier, C., O and Larher, F. R. (2006). Clustering of halophytes from an inland salt marsh in Turkey according to their ability to accumulate sodium and nitrogenous osmolytes. Environmental and Experimental Botany 57, 139–153. Urano, K., Kurihara, Y., Seki, M. Shinozaki, K. et al. (2010). ’Omics’ analyses of regulatory networks in plant abiotic stress responses. Current Opinion in Plant Biology 13(2), 132–138. Verbruggen, N. and Hermans, C. (2008). Proline accumulation in plants: A review. Amino Acids 35(4), 753–759. Verbruggen, N., Villarroel, R. and Van Montagu, M. (1993). Osmoregulation of a pyrroline-5-carboxylate reductase gene in Arabidopsis thaliana. Plant Physiology 103(3), 771–781. Verbruggen, N. et al. (1996). Environmental and developmental signals modulate proline homeostasis: Evidence for a negative transcriptional regulator. Proceedings of the National Academy of Sciences of the United States of America 93(16), 8787–8791. Verslues, P. E., Kim, Y. S. and Zhu, J. K. (2007). Altered ABA, proline and hydrogen peroxide in an Arabidopsis glutamate:Glyoxylate aminotransferase mutant. Plant Molecular Biology 64(1–2), 205–217. Voetberg, G. and Stewart, C. R. (1984). Steady state proline levels in salt-shocked barley leaves. Plant Physiology 76(3), 567–570.

Author's personal copy 148

L. SZABADOS ET AL.

Vogel, G., Aeschbacher, R. A., Mu¨ller, J., Boller, T. Wiemken, A. et al. (1998). Trehalose-6-phosphate phosphatases from Arabidopsis thaliana: Identification by functional complementation of the yeast tps2 mutant. The Plant Journal 13(5), 673–683. Vogel, G., Fiehn, O., Jean-Richard-dit-Bressel, L., Boller, T., Wiemken, A., Aeschbacher, R. A. Wingler, A. et al. (2001). Trehalose metabolism in Arabidopsis: Occurrence of trehalose and molecular cloning and characterization of trehalose-6-phosphate synthase homologues. Journal of Experimental Botany 52(362), 1817–1826. Wang, L. W. and Showalter, A. M. (2004). Cloning and salt-induced, ABA-independent expression of choline mono-oxygenase in Atriplex prostrata. Plant Physiology 120(3), 405–412. Wang, X., Li, W., Li, M. and Welti, R. (2006). Profiling lipid changes in plant response to low temperature. Physiologia Plantarum 126, 90–96. Wang, F., Zeng, B., Sun, Z. Zhu, C. et al. (2009). Relationship between proline and Hg2þ-induced oxidative stress in a tolerant rice mutant. Archives of Environmental Contamination and Toxicology 56(4), 723–731. Weber, A. P., Horst, R. J., Barbier, G. G. Oesterhelt, C. et al. (2007). Metabolism and metabolomics of eukaryotes living under extreme conditions. International Review of Cytology 256, 1–34. Weckwerth, W. (2003). Metabolomics in systems biology. Annual Review of Plant Biology 54, 669–689. Wei, W. et al. (2010a). Cloning and expression analysis of 1 L-myo-inositol-1-phosphate synthase gene from Ricinus communis L. Zeitschrift fur Naturforschung. Section C 65(7–8), 501–507. Wei, A. et al. (2010b). The pyramid of transgenes TsVP and BetA effectively enhances the drought tolerance of maize plants. Plant Biotechnology Journal 9, 216–229. Weigel, P., Weretilnyk, E. A. and Hanson, A. D. (1986). Betaine aldehyde oxidation by spinach chloroplasts. Plant Physiology 82(3), 753–759. Welti, R., Li, W., Li, M., Sang, Y., Biesiada, H., Zhou, H. E., Rajashekar, C. B., Williams, T. D. Wang, X. et al. (2002). Profiling membrane lipids in plant stress responses. Role of phospholipase D alpha in freezing-induced lipid changes in Arabidopsis. The Journal of Biological Chemistry 277(35), 31994–32002. Weltmeier, F., Ehlert, A., Mayer, C. S., Dietrich, K., Wang, X., Schu¨tze, K., Alonso, R., Harter, K., Vicente-Carbajosa, J. Dro¨ge-Laser, W. et al. (2006). Combinatorial control of Arabidopsis proline dehydrogenase transcription by specific heterodimerisation of bZIP transcription factors. The EMBO Journal 25(13), 3133–3143. Whittaker, A., Bochicchio, A., Vazzana, C., Lindsey, G. Farrant, J. et al. (2001). Changes in leaf hexokinase activity and metabolite levels in response to drying in the desiccation-tolerant species Sporobolus stapfianus and Xerophyta viscosa. Journal of Experimental Botany 52(358), 961–969. Widodo, et al. (2009). Metabolic responses to salt stress of barley (Hordeum vulgare L.) cultivars, Sahara and Clipper, which differ in salinity tolerance. Journal of Experimental Botany 60, 4089–4103. Wingler, A. and Roitsch, T. (2008). Metabolic regulation of leaf senescence: Interactions of sugar signalling with biotic and abiotic stress responses. Plant Biology 10(Suppl. 1), 50–62. Wood, A. J. et al. (1996). Betaine aldehyde dehydrogenase in sorghum. Plant Physiology 110(4), 1301–1308.

Author's personal copy PLANTS IN EXTREME ENVIRONMENTS

149

Xue, X., Liu, A. and Hua, X. (2009). Proline accumulation and transcriptional regulation of proline biosynthesis and degradation in Brassica napus. BMB Reports 42(1), 28–34. Yancey, P. H., Clark, M. E., Hand, S. C., Bowlus, R. D. and Somero, G. N. (1982). Living with water stress: Evolution of osmolyte systems. Science 217(4566), 1214–1222. Yang, X., Liang, Z. and Lu, C. (2005). Genetic engineering of the biosynthesis of glycinebetaine enhances photosynthesis against high temperature stress in transgenic tobacco plants. Plant Physiology 138(4), 2299–2309. Yang, X., Wen, X., Gong, H., Lu, Q., Yang, Z., Tang, Y., Liang, Z. Lu, C. et al. (2007). Genetic engineering of the biosynthesis of glycinebetaine enhances thermotolerance of photosystem II in tobacco plants. Planta 225(3), 719–733. Yang, X., Liang, Z., Wen, X. Lu, C. et al. (2008). Genetic engineering of the biosynthesis of glycinebetaine leads to increased tolerance of photosynthesis to salt stress in transgenic tobacco plants. Plant Molecular Biology 66(1–2), 73–86. Yang, S. L., Lan, S. S. and Gong, M. (2009). Hydrogen peroxide-induced proline and metabolic pathway of its accumulation in maize seedlings. Journal of Plant Physiology 166, 1694–1699. Yeo, E. T., Kwon, H. B., Han, S. E., Lee, J. T., Ryu, J. C. Byu, M. O. et al. (2000). Genetic engineering of drought resistant potato plants by introduction of the trehalose-6-phosphate synthase (TPS1) gene from Saccharomyces cerevisiae. Molecules and Cells 10(3), 263–268. Yoo, J. H. et al. (2005). Direct interaction of a divergent CaM isoform and the transcription factor, MYB2, enhances salt tolerance in arabidopsis. The Journal of Biological Chemistry 280(5), 3697–3706. Yoon, K. A., Nakamura, Y. and Arakawa, H. (2004). Identification of ALDH4 as a p53-inducible gene and its protective role in cellular stresses. Journal of Human Genetics 49(3), 134–140. Yoshiba, Y. et al. (1995). Correlation between the induction of a gene for delta 1-pyrroline-5-carboxylate synthetase and the accumulation of proline in Arabidopsis thaliana under osmotic stress. The Plant Journal 7(5), 751–760. Yoshida, K. T. et al. (1999). Temporal and spatial patterns of accumulation of the transcript of Myo-inositol-1-phosphate synthase and phytin-containing particles during seed development in rice. Plant Physiology 119(1), 65–72. Yoshida, K. T., Fujiwara, T. and Naito, S. (2002). The synergistic effects of sugar and abscisic acid on myo-inositol-1-phosphate synthase expression. Physiologia Plantarum 114(4), 581–587. Yousfi, N. et al. (2010). Effects of water deficit stress on growth, water relations and osmolyte accumulation in Medicago truncatula and M. laciniata populations. C R Biol 333(3), 205–213. Zhang, J., Tan, W., Yang, X. H. Zhang, H. X. et al. (2008). Plastid-expressed choline monooxygenase gene improves salt and drought tolerance through accumulation of glycine betaine in tobacco. Plant Cell Reports 27(6), 1113–1124. Zhang, Y. et al. (2009). Inhibition of SNF1-related protein kinase1 activity and regulation of metabolic pathways by trehalose-6-phosphate. Plant Physiology 149(4), 1860–1871. Zhao, Y., Aspinal, D. and Paleg, L. G. (1992). Protection of membrane integrity in Medicago sativa L. by glycine betaine against effects of freezing. Journal of Plant Physiology 140, 541–543. Zhifang, G. and Loescher, W. H. (2003). Expression of a celery mannose 6-phosphate reductase in Arabidopsis thaliana enhances salt tolerance and indices

Author's personal copy 150

L. SZABADOS ET AL.

biosynthesis of both mannitol and a glucosyl-mannitol dimer. Plant, Cell & Environment 26, 275–283. Zuther, E., Koehl, K. and Kopka, J. (2007). Comparative metabolome analysis of the salt response in breeding cultivars of rice. In Advances in Molecular Breeding Toward Drought and Salt Tolerance Crops, (M. A. Jenks, P. M. Hasegawa and S. M. Jain, eds.), pp. 285–315. Springer-Verlag, Berlin, Heidelberg, New York.