. u . 4, {133...rszsh .. ;... as: .51.... . . , 1.: . 54;. p. In «2.1.1.... a... , 5:53.}! . Gina-ti). 37!. . 3...... . ‘ . .1 .. . 5397.3». s 2. n». a... «1.... ... . :r...v;........,.!t;; 3.. 1.! .u 4943.3}: .1. ' a...» .5. 1.7. {£312 1... 923......»3. J. . . 4.1“}. § . on. . . . l. , 11‘. 71.13.12. . :. .Lrv.xc..vl.m.1.lq..4‘aj..l , , H ”414?. .C‘IIIJ .. . ., Tam». Hawk, EEK}?fifimgggaafzfmf . ‘ . . gamut—£3 Efiégflégfir . an. . “Raga? .... .Ekfimm , @qfifli. _ ‘vl . . .7 . L .L J {:3 l ”1.12.: C-t‘. ..\I ’ C I . (23794 3.2% IIllgl‘ATE Uill'VERSITY CHIGAN S EligT LANSING, MICH 48824-1048 This is to certify that the dissertation entitled Genetic engineering for dehydration-stress tolerance in cucumber (Cucumis sativus L.) by expressing the Arabidopsis thaliana-transcriptional regulators CBF1 and CBF3 and the mannose-6—phosphate reductase gene M6PR from celery (Apium graveolens L.) presented by Mohamed Saleh Tawfik has been accepted towards fulfillment of the requirements for the Ph.D. ’ Horticulture Major Professor’s Signature 54 7 -c 5" Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 c:/Cli_=i-C/DateDuo.indd-p.15 Genetic engineering for dehydration-stress tolerance in cucumber (Cucumis sativus L.) by expressing the Arabidopsis thaliana-transcriptional regulators CBF1 and CBF3 and the mannose-6-phosphate reductase gene M6PR from celery (Apium graveolens L.) By Mohamed Saleh Tawfik A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Horticulture Department 2005 ABSTRACT Genetic engineering for dehydration-stress tolerance in cucumber (Cucumis sativus L.) by expressing the Arabidopsis thaliana-transcriptional regulators CBF1 and CBF 3 and the mannose-6-phosphate reductase gene M6PR from celery (Apium graveolens L.) By Mohamed Saleh Tawfik Salinity and drought conditions are major factors affecting plant productivity and distribution worldwide. To engineer resistance to dehydration stress in cucumber (Cucumis sativus L.), transgenic cucumber were generated with genes associated with enhanced abiotic stress tolerance: the mannose-6-phosphate reductase (M6PR) gene from celery for mannitol production, and the CBF1/BREE] b and CBF 3/DREBI a, abiotic stress-associated transcriptional regulators from Arabidopsis thaliana. To transgenic M6PR cucumbers produced detectable mannitol, indicating functionality of the M6PR gene in cucumber. However, mannitol accumulation in the T1 progeny was highly variable making this trait-difficult to work with. Eleven lines of cucumber were produced with the CBF genes, integration and expression was verified in the T0, T1 and T2 generation. Under greenhouse conditions, T1 and T2 CBF—cucumber plants accumulated elevated levels of proline and soluble sugars, a signature for CBF expression in Arabidopsis, indicating ability of the CBF gene to induce stress related responses in cucumber. Proline and soluble sugars accumulation were highly correlated, suggesting coordinated regulation in the transgenic plants. In the absence of salt or drought stress, the CBF cucumbers showed equivalent growth compared to the nontransgenic controls. In the presence of salt and drought stress, transgenic plants had less reduction in growth. Plant performance and fruit production was evaluated under field conditions. Prior to salinity-stress, transgenic and nontransgenic cucumber lines grew equivalently. CBF- cucumber plants accumulated significantly higher levels of compatible solutes in leaves (proline and soluble sugars) and roots (proline) compared to the nontransgenic controls. Transgenic plants also had elevated levels of K+ and CaH ions and a decreased NaVK+ ratio in root tissues, suggesting a wider range of adaptive responses in the transgenic plants than has been reported previously. In the absence of salinity, CBF lines had less fresh weight than the nontransgenic controls; however, dry weight and fruit yield were equivalent to the nontransgenics. In the presence of salinity stress, CBF-transgenic plants showed significantly less reduction in fresh weight, dry weight, fruit number and fruit weight. These results suggest that expression of the CBF/DREB in cucumber, a species known for sensitivity to salinity and drought conditions, may offer an effective approach to enhance salinity and drought tolerance. Acknowledgments I would like to thank my committee members for their endless support and encouragement during the last five years, Dr. Mike F. Thomashow, Dr. Wayen H. Loescher, and Dr. Mathieu N gouajio. I particularly would like to thank my major professor Dr. Rebecca Grumet, for her guidance; you were the best supervisor any student could ever ask for; your support carried me over all the difficulties I had. I also would like to thank the previous and the current members of the ABSP project for their unconditional support for the last decade. I particularly would like to thank Dr. John Dodds, Dr. Catherine Ives, Dr. Johan Brink, and Dr. Karim Maredia, your help and support is beyond my capability to express it in words. Also I would like to thank Dr. Colm Lawler, for his trust and friendship. I also‘would like to thank the long list of friends and colleagues in the plant science community at MSU, being part of this elite group is an honor, you enriched my life and made my last five years something to cherish and hold on to for the rest of my life. Thanks to all the current and previous members of the Grumet’s lab, your presence made the lab a fun place to be there every day. Finally, I would like to finish by thanking my wonderful family, my Dad, Mom and my wonderful sisters, for their everlasting love and support. iv Table of content List of Tables List of Figures Chapter I Literature review Innoducfion Causes and effects of plant water stress A- Drought effects on plants B- Salinity effects on plants Improvement of dehydration stress tolerance in plants A- Conventional breeding for drought and salt tolerance B- Genetic engineering for dehydration stress 1- Genes involved in Na” ion exclusion from the cyt0plasm 2- Genes that control osmolyte and compatible solute 3- Genes that encode stress induced proteins 4- Genes involved in transcriptional regulation of stress response mechanism in plants a- The CBF system b- Regulation of CBF c- Effect of overexpression of CBF/DREB on the plant transcriptome d- Effect of CBF/DREB-overexpression on plant metabolome e- Conservation of the CBF/DREB system among plant species f- Engineering dehydration-stress tolerance in plants using the CBF/DREB system The objectives of this dissertation Literature cited page viii ix \lQ-bwi—H 10 16 21 23 23 25 28 31 32 33 37 39 Chapter H Expression of the Arabidopsis thaliana transcriptional regulators, CBF1 and CBF 3, confers dehydration stress tolerance in cucumber (Cucumis sativus L.) plants. Introduction 55 Materials and methods 59 Plant constructs and transformation 59 DNA and RNA isolation, PCR, and Northern blot analysis 60 Salinity experiments 61 Drought experiments 62 Sugar and proline analyses 63 Chlorophyll fluorescence 63 Results 64 Discussion 80 Literature cited 86 Chapter III Cucumber (Cucumis sativus L.) plants expressing the Arabidopsis thaliana- transcriptional regulators, CBF] and CBF 3, are more tolerant to salinity stress under field conditions Introduction 91 Materials and methods 96 Plant materials 96 Field salinity Experiment 96 Soluble sugars, proline and mineral analysis 98 Results 99 Discussion 108 Literature cited 114 vi page Chapter IV Introduction of the celery mannose-6-phosphate reductase (M6PR) gene for mannitol production into cucumber (Cucumis sativus L). Introduction 118 Materials and methods 123 Plant constructs, transformation and seed production 123 DNA isolation and PCR analysis 123 ELISA analysis 123 Mannitol analysis 124 Results and discussion 125 Literature cited 131 Conclusions and future work 134 Appendix 139 vii List of Table Table l-lA. Genetic engineering for dehydration stress tolerance in plants: A. Ion exclusion/sequestration/balance Table l-lB. Genetic engineering for dehydration stress tolerance in plants: B. Osmotic adjustment/dehydration stress tolerance Table l-lC. Genetic engineering for dehydration stress tolerance in plants: C. Stressed induced proteins Table l-lD. Genetic engineering for dehydration stress tolerance in plants: D. Transcription factors and other proteins involved in gene regulation Table 2-1. Segregation analysis of T1 transgenic cucumber lines expressing the Arabidopsis CBF1 and CBF3 genes. Presence or absence of the gene was determined by PCR analysis Table 2-2. Levels of soluble sugar, free proline, above ground fresh weight and dry weight, and plant height in CBF-transgenic and control lines 15 days post initiation of N aCl treatment Table 3-1. Number of nodes, branches, male and female flowers of T2 355:CBF and non-transgenic cucumber lines prior to initiation of salinity treatment in the field, 3 weeks post transplanting. Table 3-2. Fresh weight, dry weight, fruit number and total fruit yield of CBF— expressing transgenic families and the nontransgenic control lines 24 days post initiation of salinity treatment. Table 4-1. Segregation analysis of T1 transgenic cucumber lines expressing the celery-M6PR gene. Presence or absence of the gene was determined by PCR analysis Table 4-2. Mannitol accumulation in segregating Tl-M6PR plants. viii 11 12 13 14 65 70 100 107 125 130 List of Figures Figure 2-1. Expression of A. thaliana-transcriptional regulators CBF1 and CBF3 genes in T2 transgenic cucumber lines. Figure 2-2. Free proline content (A-C) and total soluble sugars (D-F) of two transgenic lines (A4, B3), non-transformed control (C) and nontransgenic segregants (AZ). Figure 2-3. Total soluble sugar, proline content, above ground fresh and dry weight of ten transgenic lines (CBF1 and CBF 3), non-transformed control (C) and azygous segregants (AZ) 15 days post initiation of 100 mM NaCl treatment. Figure 2-4. A- Correlation between soluble sugar and proline content in cucumber plants 15 days post initiation of 100 mM NaCl treatment. Figure 2-5. Expression of A. thaliana-transcriptional regulators CBF1 and CBF 3 genes in four T2 cucumber expressing lines and non-transgenic lines (azygous) growing the presence or absence of 100 mM NaCl for 2 weeks. Figure 2-6. Total soluble sugar and free proline content of CBF] and CBF3-expressing cucumber lines and nontransgenic controls under well- irrigated and drought conditions. Figure 2-7. Chlorophyll fluorescence (F v/F m) in CBF-expressing cucumber plants and azygous non-transgenic plants, under well-irrigated and drought- stressed conditions. Figure 2-8. Effect of drought conditions on above ground fresh weight (A), dry weight (B) weight, and plant height (C) of CBF], and CBF3-transgenic lines and non-transformed controls C, and non-transgenic segregants AZ. Figure 3-1. Accumulation of proline ug/gdw (A, B) and soluble sugars (C, D) in cucumber plants growing in the field in the absence (A and C) or presence (B and D) of 100 mM NaCl. Figure 3-2 Accumulation of proline mg/gdw (A) and soluble sugars mg/gdw (B) in root tissues of cucumber plants growing in the field in the absence or presence of 100 mM salinity stress. Figure 3-3. Potassium (A), sodium (B), Na/K ratio (C) and calcium content (D) in roots of cucumber plants growing in the field for 24 days, in the presence or absence of salinity treatment. ix 66 67 69 72 73 75 77 78 101 103 104 Figure 3-4. Above ground fresh (A) and dry weight (B) accumulation, average fruit number (C) and total fruit yield (D) per plot in cucumber plants growing in the field in the presence or absence of salinity treatment for 24 days. 106 Figure 4-1. PCR analysis of the presence of the M6PR gene in To cucumber plants. 126 Figure 4-2. Gas chromatography for mannitol analysis in one of the cucumber To plants (M3) (top) and nontransgenic wild-type cucumber plants (bottom) 127 Figure 4-3. Gas chromatography for mannitol analysis in wild type (top) and T1 plant from family M3 (bottom). 129 Chapter I Literature Review Introduction Environmental factors that impose water-deficits stress place a major limitation on plant productivity (Bray, 1994; Bohnert and Jensen, 1996). Water deficit is intrinsic to most abiotic forms of stress, such as drought, salinity and freezing temperatures (Bohnert and Jensen, 1996). Deleterious effects can be manifested as a reduction in transpiration and photosynthesis, a reduction in growth rate due to reduced cell enlargement, and reduction in the synthesis of metabolites and structural compounds (Zhang et al., 1999). Cellular water deficit also disrupts membrane integrity, which causes loss of cellular water potential and denaturation of cellular proteins (Bray, 1997). To overcome these limitations and to improve production efficiency of plants, development of dehydration stress-tolerant crops is essential (Khush, 1999). In this literature review, I will summarize causes and effects of dehydration stresses in plants and efforts to develop dehydration stress tolerance in plants via conventional breeding and genetic engineering approaches. 1- Causes and effects of plant water stress . Water is the driving force of living organisms; it works as a medium for the biochemical activities in all living cells (Xiong and Zhu, 2002) and is involved in biosynthesis and assembly of molecules into organized structures (Tanford, 1978). In plant cells, water potential (WW), is responsible for generating the needed turgor pressure for cell expansion. The water potential of a given cell is composed of pressure potential. (Y’p, which reflects the physical pressure generated by cell wall) and osmotic potential ('I’s which is generated by the solute concentration inside the cell). Plant water deficit results fi'om inability of plants to acquire their water needs, resulting in loss of turgor and/or osmotic stress. This could be due to the unavailability of water under drought and freezing temperatures, or the presence of highly negative osmotic pressure, due to high salt concentrations in the growing environment. Cellular water deficit, if prolonged, could be lethal to plant cells, interfering with basic metabolic pathways and changes in membrane shape and integrity (Bray, 1997). The ability of plants to respond to and survive water deficit is a complex phenomenon, which requires adjustment at the molecular, cellular and whole plant level (Greenway and Munns, 1980; Ingram and Bartels, 1996; Zhu et al., 1997). At the molecular level, osmotic stress will trigger cascades of signals involving CaH and reactive oxygen molecules as primarily signals to activate pathways critical for plant survival under the stress conditions (Knight et al., 1997; Knight and'Knight, 2001). At the cellular level, responses include metabolic adjustment to produce compatible solutes (Cherry, 1989), activation of transporters at the plasma and vacuolar membranes for ion sequestration or exclusion (Blumwald and Poole, 1985; Shi et al., 2000), and activation of enzymes involved in detoxification of free radicals (Mittova et al., 2002; Bor et al., 2003; Mittova et al., 2004; Badavvi etal., 2004). At the whole plant-level, responses include closure of plant stomatal apparatus coupled with an inhibition of vegetative growth and increase in root growth (Maggio et al., 2003). I-A- Drought effects on plants Drought is a serious environmental factor that affects plant production worldwide. For example, it is estimated that about 25% of the total cultivated areas with rice in the world is under rain-fed; the shortage of water from one season to another is a serious threat to yield stability (Babu et al., 2001 ). Drought stress occurs when the rate of water uptake from the soil is less than plant transpiration rate (Bonhert and Jensen, 1996). One of the first responses of plants to dehydration stress is triggered by the increase of ABA concentration, which generates a cascade of signals that leads to a decline in stomatal conductance to minimize transpiration and to keep it in balance with water absorption from soil (Zeevaart and Creelrnan, 1988; Chandler and Robertson, 1994). At this stage the plant can still maintain turgor, and partial stomatal closure can occur several times on daily basis, especially during mid-day. If unfavorable conditions continue for a long time, then the stomatal apparatus loses the ability to compensate for the lack of water and stomatal conductance declines sharply (Quarrie, 1989). High abscisic acid (ABA) concentrations in plant tissues under drought conditions contribute to reduction of leaf area and plant height, and pollen abortion (Quarrie, 1989). Persistence of drought stress eventually causes a dramatic reduction in all processes contributing to plant yield and reduction in plant growth in general (Cherry, 1989). Persistence of drought conditions eventually forces the plant to concentrate on survival and water conservation mechanisms (Cherry, 1989). ABA application or an endogenous transient increase in ABA due to drought causes cytosolic pH changes and membrane depolarization; this increases the concentration of free cytosolic Ca2+ ions in guard cells in response to transient drought perception (Leung and Giraudat, 1998). Free cytosolic Ca” activates cyclic adenosine 5’-diphosphate ribose (cADPR), which plays a major role in ABA response (Allen and Schroeder, 1998). Another major player that is activated due to dehydration stress is phospholipase C (PLC), reSponsible for releasing inositol 1,4,5-triphosphate (1P3), which in turn mediates the release of Ca2+ ions into the cytosol (Hirayama et al., 1995). Recent reports conducted on Arabidopsis cell suspensions suggests that this transient increase in 1P3 is independent of ABA but still requires Ca2+ binding (Takahashi et al., 2001). I-B Salinity effects on plants Stress caused by high salinity in soil or in the irrigation water is a serious factor limiting the productivity of major agricultural crops as the majority of the agriculturally important plants species are sensitive to high salt concentrations (McWilliam, 1986; .. Zhang and Blumwald, 2001). Soil salinity affects about 5% of all cultivated land, approximately 77 million ha (Jain and Selvaraj, 1997; Tester and Davenport, 2003). Areas that are affected with salinity are increasing; for exertiple, 1/3 of the irrigated land worldwide is currently affected by salinity (Tester and Davenport, 2003). Salinization of soil is expected to reach up to 30% in the next 25 years and up to 50% of the arable land by the year 2050 (Wang et al., 2003). Plant response to salt-induced water deficit depends on several factors including genotype, length and severity of water loss, stage of deveIOpment, and environmental factors such as temperature and humidity (Bray, 1994). High salinity causes both hyperosrnotic and hyperionic stress effects, which if sufficiently severe, could result in plant death (Bohnert et al., 1999; Hasegawa et al., 2000). The plant cell membrane serves as an impermeable barrier to macromolecules and also most molecules of low molecular mass. Thus, high salt conditions can lead to increased extracellular solute concentration, which causes a flux of water out of the cells, resulting in a decrease in turgor pressure and an increase in concentrations of intracellular solutes (Lichtentaler, 1995). In addition to the lack of water, exposure to high salinity leads to “toxic sodium effect” where by excess toxic Na+ in the cytoplasm causes a deficiency of essential ions such as K+ and Ca+ (Bohnert and Jensen, 1996; Hasegawa et al., 2000). High concentration of Na+ ions in the cytosol causes metabolic toxicity; this is in part due to the ability of sodium ions to compete with K+ ions for binding sites for several enzymes (Tester and Davenport, 2003). High Na+ and Cl' concentrations also disrupt enzyme function, protein synthesis, structure and solubility and membrane structure and fimction (Blum, 1988). It also has been suggested that accumulation of salts in older leaves reduces supply of hormones to the growing tissues, which contribute to poor growth in salt stressed conditions (Munns, 1993). To survive such conditions, some plants have developed mechanisms to deal with excess sodium ions either by compartrnentation of the toxic ions into the vacuole (Apes at al., 1999; Blumwald et al., 2000; Zhang and Blumwald, 2001; Zhang et al., 2001) or by exclusion from the cell (Shi et al., 2000; Zhu, 2002; Shi et al., 2003). Interestingly, halophyte (plants that normally grow in saline areas and can tolerate up to 0.5 M NaCl concentration before suffering injuries) are able to use ions fiom the surrounding environment for osmotic adjustment by internally distributing them in a way to keep the Na+ ions away from the cytosol (Zhu, 2000). Alternatively, some plants tend to accumulate compatible solutes in order to overcome high salinity problem (Tarczynski et al. 1993; Bohnert and Jansen, 1996; Sakamoto and Murata, 2000), while other species have the ability to activate proteins that are involved in damage repair in plant cells (Ingram and Bartels, 1996; Campbell and Close, 1997). High salinity conditions are also responsible for generating reactive oxygen radicals which, if not dealt with, could lead to unbalance in cellular 0; processing (Rental and Knight, 2004). Strong correlation was found between the ability of plants to adapt to high salinity levels and the increased activity of antioxidant enzymes such as superoxide dismutase (SOD), peroxidase (POX), catalase (CAT), ascorbate peroxidase (APX) and glutathione reductase (GR) (Mittler and Zilinskas, 1994; Liu et al., 1999; Mittova et al., 2002; Bor et al., 2003; Mittova et al., 2004; Badawi et al., 2004). In general, the process of adaptation to salinity is coupled with activation of different signaling pathways in plants that lead to changes in gene expression (Hasegawa et al., 2000; Xiong et al., 2002; Zhu, 2002; Shinozaki et al, 2003; Seki et al., 2003). For example, imposing dehydration stress conditions on A. thaliana resulted in the activation of transcriptional factors that are involved in plant adaptation to salt-induced- dehydration-stress (Stockinger et al., 1997; J aglo-Ottosen et al., 1998; Liu et al., 1998; Kasuga et al., 1999; Haake et al., 2002; Chinnusamy et al., 2003). Microarray data from Arabidopsis plants grown under drought and salinity revealed the presence of 277 upregulated transcripts (5 fold) under drought conditions and 194 cDNAs that are induced under salinity conditions (Seki et al., 2002); the upregulated transcripts, 128 are strictly induced under drought conditions, 119 transcript are also induced under salinity conditions (Seki et al., 2002). Not surprisingly, it was found that about 51% of the drought induced transcripts are also induced under high-salinity conditions; similarly, about 72% of the high-salt induced transcripts are also induced under drought conditions (Seki et al., 2002). These results strongly indicate the correlation between drought and salinity signaling mechanisms in plants, which also could explain why many genes encoding for late embryonic abundant proteins (LEA), heat-shock proteins (HSp), osmoprotectant biosynthesis, carbohydrate metabolic enzymes, detoxification enzymes, transporters, ion channels and membrane modification enzymes are activated under both hi gh-salinity and drought stresses (Thomashow, 1999; Knight and Knight, 2001; Shinozaki and Yamaguchi-Shinozaki, 2003; Seki et al., 2002; Shinozaki et al., 2003). The same thing is also true for many genes coding for transcriptional regulators, mitogen activated protein kinases (MAPKs), and phosphatases that are involved in regulating plant response to high salinity and dehydration (Thomashow, 1999; Knight and Knight, 2001; Shinozaki et al., 2003) II- Improvement of dehydration stress tolerance in plants II-A Conventional breeding for drought and salt tolerance Drought and salt response, and apparent tolerance of a species, vary according to the type, concentration and distribution of salts in the root growing zone, duration of stress, and developmental stage of the plant (Jain and Selvaraj, 1997). These are also influenced by other environmental factors such as temperature, humidity, reduced oxygen in poorly drained or structured soil, and elevated C02 in the surrounding environment (Pasternak, 1987). Traditional breeding strategies to develop drought and salt tolerant plants has had only limited success, probably due to a combination of difficulties in establishing selection conditions and the complexity of the resistance mechanisms (Flowers and Yeo, 1995). For example, when evaluating yield performance of a crop under saline conditions, one should consider the variation in salinity levels within a field, the possibility of interaction between salinity level and other environmental factors such as soil fertility, drainage quality and water loss due to transpiration (Flowers, 2004). Thus, using yield components, as main criteria for selection requires a long period and multiple locations for testing, and evaluation (Blum, 1989). Quesada et al., (2002) attributed the lack of success in breeding for salt tolerance to the quantitative nature of most of the processes involved in salt tolerance. In maize, Ribaut et al. (1997) measured several yield components (grain yield, car number, kernel number and kernel weight) in plants growing in three different water-stress regimes. They found that the correlation between the grain yield under well-watered and severe stress conditions was very low (0.31). Therefore, selection based on yield component values would not be effective. They also reported that no major quantitative trait loci (QTL) expressing more than 13% of the phenotypic variance were detected for any of the studied traits. Ribaut et al. (1997) concluded that there were inconsistencies in the QTL genomic positions across the three different water regimes. Working with interspecific crosses between salt-sensitive tomato plants (Lycopersicon esculentum) and salt-tolerant L. pimpinellifolium plants, Foolad (1999) reported the presence of a weak correlation (0.22) between seed germination rate and the percentage of plant survival under salt stress. The overall results indicated that salt tolerance during seed germination was independent of salt tolerance during vegetative growth (F oolad, 1999). In general, QTL that are linked to salt and drought tolerance at one developmental stage are not necessarily linked to tolerance at other stages (Cushrnan and Bohnert, 2000). In spite of these complexities, a number of salt-tolerant varieties of crops such as wheat and rice have been developed (Shannon and Noble, 1990; Forster, 2001). For example, crossing durum wheat and the wild relative Aegilops tauschii, increased K+/Na+ discrimination in the synthetic hexaploid hybrid wheat (Pritchard et al., 2002). This increase in K/Na discrimination was significantly correlated with fresh weight accumulation in wheat plants under salt-stressed conditions. In rice, Senadhira et al., (2002), reported the production of dihaploid lines of rice from crossing of two lndica breeding lines, one of which is superior in yield, the other is superior in salinity tolerance. Field trials conducted for 5 years revealed that some of the dihaploid lines performed better than other cultivars grown in saline-prone lands. Some lines showed several desirable traits such as high yield, salinity tolerance, and early maturation. Interestingly, the use of in vitro tissue culture and somaclonal variation techniques has resulted in development of salt-tolerant plants (Safarnejad et al., 1996; Boscherini et al., 1999). In alfalfa, Safarnejad et al. (1996) isolated somatic clones of alfalfa, which showed increased salt tolerance, greater accumulation of proline, and a greater increase in glutathione reductase compared to the parental line. In tomato, Boscherini et al. (1999) identified a somatic clone that showed enhanced tolerance to salinity compared to wild type when tested at different NaCl concentrations (0, 75, 150, 300 mM NaCl). Leaf and flower necrosis was observed only in wild type plants. Plants coming from the somatic clone retained higher leaf turgor compared to the wild type when tested at 150 mM NaCl. II-B Genetic engineering for dehydration stress Differences in gene expression profiles among dehydration stress sensitive and tolerant plants indicate that the ability to withstand these unfavorable conditions is conferred by genetically encoded mechanisms (Bray, 1994). A variety of approaches have been used in order to engineer enhanced salinity and drought tolerance in different plant species (Table 1-1). II-B-l Genes involved in Na+ ion exclusion from the cytoplasm One way for plants to avoid Na+ ion toxicity is to exclude Na+ fiom the cytosol. Ion transport across membranes in plant cells (plasma membrane and tonoplast) is driven by proton gradients generated by proton pumps located at the different membranes (Sze et al., 1999). The main pump in the cell plasma membrane is the plasma membrane H+- ATPase (PM H+-ATPase), which is responsible for generating the gradient between the cytosol and the extracellular environment, making the cytosol more basic and the outer environment more acidic (Palrngren, 1998). At the vacuole, there are two major pumps that are responsible for generating an acidic pH inside the vacuole; those pumps are the vacuolar H+-ATPase (V-ATPase) and the vacuolar H+ pumping pyrophosphatase (H+- PPase) (Sze et al., 1999). Plant cells use the electrochemical gradient that is generated by the different pumps to load/unload different materials into or out of the cytosol. In the case of sodium compartrnentation, Na+ ions are loaded from the cytosol directlyth the vacuole before it reaches a critical toxic concentration. The presence of 10 Table 1. Genetic Engineering for dehydration stress tolerance in plants: A. Ion F Resistance content Canola Tomato Rice Wheat rates; grain yield no; fresh and dry weight; Conditions of stress treatment and with 200 mM NaCl for 4D Salinity: Hydrdponic-culture mM NaCl) Salinity: Germination supplemented with 0300 mM NaCl Salinity: Field trail (in soil with EC. of 1.2. 10.6 and 13.7 dSm-i) soild were irrigated 4 were C G F References Table 1. B. Osmotic Resistance PSCR Tobacco (Pyrroline-SCarboxylateReducase) Rice PDHase (proline dehydrogenase) MtlD (mannitol-1phosphate Tobacco dehydrogenase) Wheat Eggplant CodA (Coline oxidase) BetB. Betaine aldehyde dehydrogenase BADH Betaine aldehyde dehydrogenase (Betaine aldehyde dehydrogenase) fluorescence; roots biomass; TPS1 (trehalose—1-phospho trasferase) T6PSIT6PP S-adenosylmethionine decarboxylase SAM-DC Sperrnidine biosynthesis mg..- 12 tolerance Conditions of stress treatment C G F References or Irrigation with 50 150 ml (non-stress) H20 every 4D grain yield Drought In assay with 10% (w/v) PEG Drought 21 ‘D water withheld at 10 leaf stage to 15% relative water content assay on for 2D In vitm on medium with 75mM NaCl for 35D Table 1. C. Stress induced | C G F References Resistance Conditions of stress treatment Rice recovery; on or and root weight Plants trays with 200mM NaCl 10D then with SOmM NaCl for 30D OT were Wheat every Tomato medium with 125 or mM NaCI for 28D assay on Oil for 100 tolerance Table 1. D- T kinase) factors and other involved in Resistance Conditions of stress treatment were grown same were seedling assay on agar plates supplemented with 500mM sorbitol for 011 were paper soaked with either 250 mM for 4D or mannitol (300 and 400 mM) C G F References X large, acidic membrane-bound vacuoles in plant cells allows cells to efficiently compartmentalize excess Na+ ions into the vacuole by the vacuolar Na+/H+ antiporter (Blumwald and Poole, 1985). The difference in the H+ is initially established by the H+- ATPase pump (Blumwald et al., 2000). In salt tolerant species, an increase in transcript level of Na+/H+ antiporters was observed upon exposing plants to high salt levels (Tester and Davenport, 2003). Apse et al. (1999) overexpressed the A. thaliana AtNHXI gene (coding for a vacuolar Na+/H+ antiporter) in Arabidopsis and showed that transgenic plants were able to tolerate up to 200 mM NaCl treatment (Table l-lA). Tomato plants overexpressing the AtNHXI gene accumulated 20-28 fold more sodium in their vegetative tissues compared to wild type plants (Zhang and Blumwald, 2001) and AtNHXI-overexpressing canola plants accumulated up to 6% of their dry weight as sodium compared to almost 0.01% in non-transgenic plants (Zhang et al., 2001). Salt exclusion can be facilitated by the use of the SOS (Salt Overly Sensitive) genes, which encode a plasma membrane Na+/11+ antiporter(SOS1), a serine/threonine protein kinase ($052) and a myristoylated Ca2+-binding protein (5053). Identification of those genes has also furthered our understanding of Ca2+ signaling in plant response pathways to salinity. Sudden change in Na+ ion concentration in the cytosol is immediately coupled with a transient change in the Ca2+ concentration in the cytosol; this transient change in cytosolic Ca2+ is known as the Ca signature (Knight et al., 1997). The calcium ions then bind to the myristoylated Ca2+-binding protein, encoded by SOS3, which then mediates downstream responses (Liu and Zhu, 1998; Ishitani et al., 2000; Zhu, 2002). S053 interacts with and activates the $052, which is a serine/threonine protein kinase (Halfter et al., 2000; Liu et al., 2000; Zhu, 2002). Both SOS3/SOSZ proteins are 15 responsible for regulating the plasma membrane Na+/H" antiporter, which is encoded by the SOS] gene in Arabidopsis (Shi et al., 2000). Overexpression of the plasma membrane Na+/IV antiporter gene SOS] in a mutated strain of yeast lacking the Na+/H+ antiporters had a slight but significant increase on yeast survival on medium supplemented with 100 mM NaCl. Co-expression of SOS3/SOS2 as well as SOS] genes in the same mutated strain of yeast had a dramatic effect on yeast survival on medium supplemented with NaCl (Zhu, 2002). Moreover, overexpression of SOS] gene in Arabidopsis significantly improved plant salt tolerance (Shi et al., 2003). Shi et al., (2003) reported that transgenic plants overexpressing SOS] gene accumulated less Na+ in their tissues than their non-transgenic counterparts (Table l-lA). Recently, there have reports indicating that the 8082 protein kinase may have multiple effects on other genes (Cheng etal., 2004; Qiu et al., 2004). Qiu et al., (2004) demonstrated that the tonoplast Na+/H+ transporter in Arabidopsis is also one of the targets of the SOS regulatory pathway, which means that there might be more branches to the SOS pathway. Interestingly, Cheng et a]. (2004) showed that 8082 protein kinase also regulates the vacuolar I-I‘L/Ca2+ antiporter CAXl. Co-expression of SOS2 specifically activated CAX] gene in yeast. Using the yeast two-hybrid assay, $082 was found to interact with the N terminus of CAXI. II-B—2 Genes that control osmolyte and compatible solute content in plants An important adaptation to osmotic stress is the ability to accumulate compatible solutes (e. g., proline, glycine-betaine, alcohol sugars, fructans and trehalose) in the cytoplasm under dehydration-stress conditions (Bohnert and Jansen, 1996). Compatible 16 solutes, which are sometimes called osmoprotectants, are non-toxic organic metabolites of low molecular weight that act to raise the osmotic potential of the cell, or to stabilize membranes or macromolecular structure (Bohnert and Jensen, 1996). Engineering strategies for developing salt stress tolerance has been performed with genes encoding production of proline, glycine betaine, mannitol, fructans, and trehalose (Holmberg and Bulow, 1998; Taiz and Zeiger, 1998). Several groups have demonstrated that proline accumulating in response to water or salt stress can act as an osmoprotectant in plant cells under salt stress (Kishor et al., 1995; Taiz and Zeiger, 1998; Zhu et al., 1998; Hong et al., 2000; Ain-Lhout et al., 2001). Mali and Mehta (1977) were among the first to report on the rapid accumulation of proline in rice varieties upon exposure to drought stress. Water stress-tolerant rice plants showed a 5.4 fold increase in free proline compared to 1.2 fold increases in sensitive varieties. Increased proline accumulation in response to dehydration stresses has been observed in numerous species, and has been reviewed extensively in the literature (Cherry, 1989). One of the first attempts to engineer plants to overproduce proline came from Kishor et al. (1995), which who overexpressed in tobacco the moth bean Al-pyrroline-S- carboxylate synthetase (PSCS), a bifunctional enzyme that catalyzes the conversion of glutamate to proline. The transgenic plants accumulated 10 to 18 fold more proline than the non-transgenics. Proline accumulation was accompanied by an increase in salinity tolerance measured as increased germination percentage and seedling fresh weight when grown in 200 mM NaCl (Table l-lB). Zhu et al. (1998) also reported similar results in rice (Oryza sativa L) when overexpressing the same enzyme under a stress inducible promoter. Moreover, Ronde et al. (2000) overexpressed the Al-pyrroline-S-carboxylate l7 reductase, (in antisense orientation) in soybean plants. Antisense-soybeans failed to survive a 6-day drought stress at 37 C° in contrast to wild type plants that survived the treatment. Similarly, removal of the feedback inhibition of PSCS enzyme resulted in an increased accumulation of proline in plant tissue that correlated with an increased tolerance to osmotic stresses (Hong et al., 2000). Glycinebetaine is a common compatible solute in many different organisms including certain plant species (Sakarnoto and Murata, 2000). Betaine in vitro acts as an osmoprotectant by stabilizing the structure of proteins and the highly ordered structure of membranes against the adverse effects of water deficit conditions such as high salinity and extreme temperature (Gorham, 1996). Many plant species grown in saline and arid areas accumulate glycinebetaine in response to drought and salinity (Grumet and Hanson, 1986; Saneoka et al., 1995; Sakarnoto and Murata, 2000). Hayashi et al. (1997) achieved enhanced salt-tolerance in lirabidopsis by overexpressing the soil bacterium Arthrobacten globiformis codA gene (choline oxidase, a key enzyme in glycine-betaine production). Overexpression of the E. coli betA gene (which encodes choline dehydrogenase) in tobacco confers salt-tolerance (Holmstrom et al., 2000). Growth of control plants was totally inhibited by watering with 200 mM NaCl solution, whereas transgenics were not affected (Table l—lB). J ia et al., (2002) transformed tomato plants with the Atriplex hortensis-BADH gene, which encodes betaine aldehyde dehydrogenase, to convert betaine aldehyde into glycine-betaine. Transgenic tomato plants grew normally under 120mM NaCl and exhibited enhanced root development compared to non-transgenic plants (Table 1B). More recently, cabbage plants overexpressing the bacterial betA gene, exhibited higher tolerance to NaCl stress compared to nontransformed plants 18 (Bhattacharya etal., 2004). Transgenic cabbage plants showed better growth response and greater stability in maintaining plant water relations at high levels of salinity. Sugar alcohols also work as osmolytes in plant cells exposed to salt stress (Cherry, 1989; T arczynski et al., 1993; Bohnert and Jensen, 1996; Bohnert et al., 1999) and can also serve as scavengers for reactive oxygen species (Halliwell et al., 1988). Tarczynski et al. (1993) first reported stress protection of transgenic tobacco plants by overexpressing the osmolyte sugar alcohol mannitol (Table 1-1 B). Since then several reports have been published indicating that overexpression of genes involved in production of sugar alcohols in plants can confer stress tolerance (Karakas et al., 1997; Liu et al., 1999; Zhifang and Loescher, 2003). Karakas et al. (1997) transformed tobacco plants with a gene encoding mannitol-l-phosphate dehydrogenase (MtlD). Transgenic plants were not affected under salt stressed conditions that caused a dry weight reduction of 44% in the nontransgenic controls. More recently, Tilahun et a1. (2003) showed that overexpression of the bacterial gene mtlD in wheat resulted in plants that were more tolerant to high salt stress. When subjected to 150 mM NaCl, transgenic T2 wheat plants showed 50% reduction in fresh weight and 30% reduction in dry weight compared to 77% and 73% reduction in fresh and dry weight respectively in non-transgenic wheat (Table 1B). In contrast to using the bacterial gene mtlD, Everard et a]. (1997) isolated a gene encoding mannose-6-phosphate reductase (M6PR), key enzyme for mannitol production from celery. Zhifang and Loescher (2003) introduced the M6PR into Arabidopsis plants and showed that transgenic Arabidopsis plants were able to grow normally and complete normal development and seed production in the presence of 300 mM NaCl. Similarly, expression of the E. coli GutD gene (encoding for gluctiol-6- 19 phosphate dehydrogenase) a key enzyme for biosynthesis of the sugar alcohol sorbitol, caused increased accumulation of sorbitol in transgenic maize, and enhanced salt- tolerance compared to non-transgenics (Liu et al., 1999). Fructans are soluble polymers of fructose that are produced by approximately 15% of flowering plant species. Accumulation of fi'uctans in the cell vacuole helps maintain water potential gradients of cells by raising the osmotic potential of the cell (Pilon-Smits et al., 1995). Pilon-Smits et a1. (1999) overexpressed a gene from Bacillus subtilis (SacB gene) to produce bacterial fi'uctans in sugar beet. The growth of transgenics was significantly enhanced under drought conditions compared to the nontransgenics, as measured by higher dry leaf and root weights. The disaccharide trehalose (a-D-glucopyranosyl-l , or -D-glucopyranoside) is present in a large variety of microorganisms and plants where it can serve as a reserve carbohydrate and as an osmoprotectant (Vogel et al., 2001). The occurrence of trehalose has also been documented in several desiccation-tolerant plants (Muller et al., 1995). It protects membranes and proteins in cells exposed to salt-stress induced dehydration (Penna, 2003). Pilon—Smits et al. (1998) showed that transgenic tobacco plants overexpressing the Escherichia coli OtsA and OtsB genes (encoding trehalose-6- phosphate-synthase and trehalose-6-phosphatase respectively) responded better to dehydration stresses compared to non—transgenics. Under drought conditions the transgenics yielded 30-39% more dry weight compared to non-transgenics. They also reported that detached leaves from transgenic tobacco plants had a higher capacity to retain water than the wild type. Using the OtsA and OtsB genes under salt inducible 20 promoter, Garg et al. (2002) reported successful production of transgenic rice plants with an elevated tolerance to 100 mM NaCl (Table l-lB). II-B-3 Genes that encode stress induced proteins such as the LEA (late embryogenesis abundant) and COR (cold regulated) proteins Another important group of genes that play a role in plant adaptation and resistance to dehydration-stress induced conditions are known as LEA (late embryo genesis abundant) and COR (cold regulated) genes. LEA proteins were first identified and characterized in cotton as a set of proteins that accumulate in embryos at the late stage of seed development when seeds are undergoing the dehydration process necessary for long term survival in a dormant state (Dure et al., 1981). Transcription of genes encoding LEA proteins is also activated in other tissues such as leaves subjected to osmotic stresses (Zhang et al., 2000). LEA proteins are divided into 5 major groups according to sequence homology (Swire-Clack and Marcotte, 1999). Group 1 proteins that might play a role in binding or replacement of water, group 2 and 4 proteins that may play a role in maintaining protein and membrane structure under severe dehydration, and finally group 3 and 5 that are thought to have a role in ion sequestration in plant cells (Swire-Clack and Marcotte, 1999). The class 2 LEA proteins (also known as the lea D11 family) are dehydrins that accumulate in plant cells in response to dehydration stresses and low temperatures. The dehydrins, which range from 82 to 575 amino acids in length, share several conserved domains (Close, 1997). The first, named the K domain, is an a-helix domain composed of 15 amino acids (EKKCIMDKIKEKLPG) which exists in single or multiple COpies. The second domain is the S-segment, consisting of a stretch of 6-10 Ser 21 residues. The third is the Y-segment (T/VDEYGNP), located at the N-terminus (Close, 1996). It has been hypothesized that LEA proteins play a role in desiccation tolerance during seed development and in response to dehydration and salinity stress (Hoekstra et al., 2001 and Close, 1997). This role is probably achieved through maintenance of protein or membrane structure, sequestration of ions, binding of water, and function as molecular chaperones (Bray, 1997). Two classes of LEA proteins have been shown to have a direct functional role in salt and dehydration tolerance in plants. Rice plants transformed with the HV A1 gene from barley (a group 3 LEA protein) showed an increased tolerance to dehydration and salinity (Table 1-1C), compared to the non-transgenics (Xu et al., 1996). Improved salinity tolerance was also reported in yeast cells expressing the tomato LEA protein LE25, a group4 LEA protein (Irnai et al., 1996). Another group of LEA proteins that has a role in water binding or replacement is the LEAI group; yeast cells overexpressing the LEA] protein, Em, had enhanced growth when subjected to medium with high osmolarity (Swire-Clark and Marcotte, 1999). Plants exposed to low non-freezing temperature undergo a phenomenon known as cold acclimation, a process that is necessary for many plant species to survive freezing temperatures. Thomashow and co-workers identified a group of genes, designated as COld Responsive (COR), that are induced upon cold acclimation (Gilmour et al., 1992; Lin and Thomashow, 1992; Hovarth et al., 1993). The COR genes were also identified by other groups and are also known as LT 1 (low temperature-induced), KIN (cold- inducible), RD (responsive to desiccation) and ERD (early dehydration-inducible). COR genes comprise four families, each of which is composed of two genes that are physically 22 linked in the genome in tandem array (Thomashow, 1999). COR15, 78, 6.6 and 47 encode hydrophilic polypeptides. The COR47 hydrophilic polypeptide belongs to group 2 LEA proteins (Thomashow, 1999). COR genes may help plant cells to tolerate potentially damaging effects of dehydration associated with freezing-induced drought injury (Steponkus et al., 1998). As temperatures drop below 0 °C, ice formation initiates in the intercellular spaces of plant tissues, resulting in a drop of water potential outside the cell. The water potential gradient will facilitate the movement of unfrozen water. At —1 0 °C, more than 90% of the osmotically active water moves out of the cell causing dehydration injuries (Thomashow, 1999). Over expression of the COR15a gene in Arabidopsis thaliana protoplasts resulted in a small, but significant increase in protoplast survival upon freezing over the range of — 4.5 to —7.0 °C (Artus et al., 1996). COR15a, which encodes a chloroplast-targeted polypeptide, enhances the freezing tolerance of chloroplasts by protecting membranes from freeze-induced dehydration. II-B-4 Genes involved in transcriptional regulation of stress response mechanism in plants II-B-4a The CBF system In general, plant responses to abiotic stresses and water deficits are multigenic, where cascades of biochemical and cellular changes are necessary for plants to adapt to environmental changes (Bohnert and Jensen, 1996). Thus induction of cascades of responses may be more effective in increasing stress tolerance than single gene changes. In some cases, individual changes had only a modest effect on stress tolerance. For 23 example, introduction of glycinebetaine into Arabidopsis, Brasica napus and tobacco (Huang et al., 2000) caused only a small increase in stress tolerance. Although transgenic plants accumulated 8-18 fold more glycinebetaine than the non-transgenic controls, there was only a moderate improvement in salt tolerance in tobacco plants as measured by fresh weight of shoots. In addition, only B. napus showed a slightly better photosynthetic rate in response to salinity (Huang et al., 2000). Overexpression of the cold induced CORI5a gene in Arabidopsis thaliana did not improve the freezing tolerance at the whole plant level (Artus et al., 1996). An alternative approach to the induction of a single gene is to induce an array of adaptive plant responses through the use of key transcription factors (Thomashow, 1999). The COR genes are characterized by the presence of a common cis-acting element within their promoter region that confers stress-induction. Baker et al. (1994) reported that the 5’ region of the Arabidopsis thaliana c0r15a gene includes a cis-acting element that confers cold-, drought-and ABA-regulated gene expression. Yarnaguchi-Shinozaki and Shinozaki (1994) identified a 9 bp cis-acting element (TACCGACAT) at the promoter region of the COR 78/RD29A gene (COR 78 and RDZ9A are alternative designations of the same gene); and named it drought responsive element (DRE). The DRE stimulated gene expression in response to low temperature, dehydration, and high salinity in Arabidopsis (Y amaguchi-Shinozaki and Shinozaki, 1994). Stockinger et al. (1997) identified a 5 bp DNA regulatory element (CCGAC) at the promoter of the COR gene family and designated it as the C-repeat (CRT). The CRT element, which also occurs in the DRE element, was found to be essential for COR gene transcription in response to low temperature (Stockinger et al., 1997). 24 Using the yeast one-hybrid system, Stockinger et al. (1997) isolated an A. thaliana cDNA clone encoding the transcription factor CBF] [CRT (C-repeat)/DRE (Drought Response Element) Binding Factor]. CBF1 is a 24 kDa protein, with a nuclear localization domain and an activation domain; it also has an AP2 domain that has a DNA-binding site. Transcripts of the CBF gene family increase dramatically within 15 minutes after transferring Arabidopsis plants to low non-freezing temperature. This increase is followed by an increase in COR gene transcripts (Gilmour et al., 1998), indicating that CBF gene expression is an early step in the COR gene transcriptional cascade. Similarly, the Shinozaki group cloned two cDNAs encoding for DREB [DRE (Drought Response Element) Binding] proteins DREB] a and DREBZb and showed that expression of DREB] a was activated by low temperature, while DREBZb transcript was activated by dehydration (Liu et al., 1998). CBF], CBFZ and CBF 3 are alternative designation of DREB] b, DREB] c and DREB] a. II-B-4b Regulation of CBF Gilmour et al. (1998) proposed that due to the rapid induction of CBF genes upon cold treatment (about 15 minutes) and the lack of a CRT/DRE element in the promoter of CBF/DREB] genes, another protein that regulates expression must be present in warm conditions. This protein would bind to the CBF promoter and induce CBF expression upon cold treatment. This hypothetical protein was designated as Inducer of CBF Expression, ICE (Gilmour et al., 1998; Thomashow, 2001; Thomashow et al., 2001). Upon exposure of plants to low non-freezing temperatures, a modification in ICE or in 25 another associated protein would occur, this would allow ICE to bind to CBF promoters and upregulate CBF expression. A breakthrough in understanding cold and freezing tolerance mechanisms was achieved when Chinnusamy et al. (2003) isolated and identified ICE], an upstream transcription activator that positively upregulates transcription of the CBF gene family. ICE] gene encodes a 53.5 kD nuclear localized MY C-likc basic-Helix-Loop-Helix (bHLH domain) transcription factor. ICEl has an acidic domain near the N-terminus with a typical bHLH DNA binding domain and a dirnerization domain near the C-terminus. ICEI binds to a cis-element (CANNTG) about 1 kb upstream of the CBF 3 promoter. Zarka et al., (2003) found that there are at least two regulatory elements in the promoter region of CBF2, Inducer of CBF Expression region 1 and 2 (ICErl and ICEr2 respectively), including the core sequence (CANNTG). Results from Chinnusamy et al. (2003) indicate that there are 5 MYC recognition sites at the promoter of CBF 3 . Knockout and overexpression experiments revealed the role of the ICE] gene in chilling and freezing tolerance in Arabidopsis plants. Mutated Arabidopsis plants (iceI) were impaired in their response to chilling and freezing temperatures, while ICE] - overexpressing plants were significantly more chilling and freezing tolerant than wild type (Chinnusamy et al., 2003). Interestingly, although I CE 1 -overexpressing plants were chilling and freezing tolerant, they did not show an elevation in CBF 3 gene expression level in warm temperature; an increase in CBF 3 expression was only observed under cold temperature (Chinnusamy et al., 2003). In ice] -mutants, the expression levels of CBF] and CBF 2 were similar to wild type plants, although there was some delay in expression at l and 3 26 hours of cold treatment (Chinnusamy et al., 2003). Results from Chinnusamy et al. (2003) indicate that expression of members the CBF/DREB] gene family might be regulated differently due to the observation that members of the CBF/DREB] family are differentially induced under different conditions (Haake et al., 2002) and that members of CBF/DREB] family are differentially expressed under the same stress condition (N ovillo et al., 2004). For example, screening of the Arabidopsis genome revealed the presence of another CBF homolog, CBF 4, which has a 63% amino acid similarity to the CBF gene farme (Haake et al., 2002). The expression of CBF4/DREBId is not induced under low temperatures, but is rapidly induced in response to drought and ABA treatment. Recently, Novillo et al. (2004) investigated the contribution of each CBF/DREB] member in cold adaptation. Using knockout mutants of CBF2, Novillo et al. (2004) found the surprising results that cbf2 plants had an elevated level of CBF] and CBF 3 transcript even under warm temperature. This resulted in induction of target genes and increased freezing, dehydration, and salinity tolerance (N ovillo etal., 2004). Northern blot analysis revealed a delay in expression between the different members of the CBF gene family, with CBF] and CBF 3 expression preceding the expression of CBF 2. These results are consistent with the fact that the transcript level of CBF 2 is almost 5 times higher than the transcript levels of CBF] and CBF 3 under warm condition (Fowler and Thomashow, 2002). Novillo et al., (2004) concluded that CBF2 might act as a negative regulator of both CBF] and CBF 3 . Under warm condition, the steady state of CBF2 transcript would negatively regulate the expression of CBF] and CBF 3, to make sure that their expression is tightly controlled. Upon cold exposure, ICE and other proteins would be activated to 27 rapidly increase the expression of CBF] and CBF 3, resulting in increased environmental- induced dehydration tolerance (Novillo et al., 2004). II-B-4c Effect of overexpression of CBF/DREB on the plant transcriptome Examination of gene expression changes allows for comparison between the effects of cold treatment and CBF overexpression. Transcriptome-profiling experiments in Arabidopsis, using Affymetrix-Gene—Chip (8297 genes) revealed changes in transcript level of 306 genes (about 4% of the total genes tested) in response to cold (Fowler and Thomashow, 2002). Of these 306 genes, 218 genes showed a 3-fold or greater transcript increase and 88 genes showed a two-fold transcript decrease in Arabidopsis plants treated at 4 C° (Fowler and Thomashow, 2002). Of the upregulated genes, more than 70% increased transiently, with expression reaching the highest level during the first 24 hours of cold treatment, followed by a dramatic decline in transcript level (Fowler and Thomashow, 2002). Transiently-upregulated genes were classified into several groups, transcription factors (CBF1 and 3, other AP2 domain containing proteins, zinc finger proteins, MYB proteins, MADS box containing proteins, and other ethylene responsive element binding factors); cell metabolic regulation (e.g., carbohydrate and osmolyte biosynthesis along with other genes involved in starch catabolism); transporters (water channels, ion and sugar transporters); cellular communication (protein kinases and proteases); cellular defense and detoxification mechanisms (LEA proteins and other enzymes involved in radicals detoxification); and finally a class of proteins with unknown function (45 genes). The remaining genes were categorized as long-term upregulated genes. Long-term upregulated genes continue to accumulate several fold 28 (compared to plants growing in warm temperature) even when tested after 7 days (Fowler and Thomashow, 2002). This group of genes also included transcription factors, genes involved in metabolic pathways, transporters, cell signaling, cell maintenance and detoxification as well as 20 genes of unknown function. Genes that were down regulated by cold included genes involved in photosynthesis and metabolism, signal transduction, heat shock and transcriptional regulation (Fowler and Thomashow, 2002). Seki et al. (2002) also reported that more than 40% of transcripts induced under cold temperature (53 transcript were induced under cold treatment in total) are also induced under both drought and high salinity, among those are different members of the COR genes (COR15a, COR4 7 and COR 78). Sequencing of the promoter region of the 22 genes that are induced under drought, cold and high salinity conditions indicated the presence of the CRT/DRE element in 16 genes and 8 of which also had the ABRE element in their promoter region (Seki et al., 2002). Comparison of transcriptome profiling between warm-grown Arabidopsis plants constitutively expressing CBF], CBF 2 or CBF 3 and control nontransgenic plants revealed that not all genes upregulated under cold temperature are also upregulated when overexpressing CBF. Thus, the CBF regulon represents only a part of the genes that are cold upregulated (Fowler and Thomashow, 2002). It was found that 60 genes (more than 80% of those genes are transiently upregulated) upregulated under cold were not upregulated in plants overexpressing any of the CBF genes; those were suggested to be CBF-independent. Of the 41 genes upregulated in CBF overexpressing plants, 30 were also upregulated by cold, which makes them the members of the CBF regulon (Fowler and Thomashow, 2002). Genes that were upregulated include genes were known to have 29 CDT /DRE elements, such as COR6. 6, COR15b, COR4 7, COR 78/RD29. Interestingly, genes that may be involved in osmolyte accumulation such as galactinol synthase for galactinol and raffinose production and P5 CS for proline production were also increased in response to CBF expression (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gilmour et al., 2000; Seki et al., 2001; Seki et al., 2002). It is known that many of the genes that are induced by the CBF genes in response to dehydration induced stress conditions are also induced by the application of abscisic acid ABA (Y amaguchi-Shinozaki and Shinozaki, 1994; Ishitani et al., 1997; Liu et al., 1998; Thomashow, 1999; Shinozaki and Yamaguchi-Shinozaki, 2000). Thomashow (1999) suggested that ABA concentration increases transiently in response to cold non-freezing temperature. This non-accumulative response helped establish the argument that ABA doesn’t play a role in cold acclimation, and that cold acclimation occurs via two separate, ABA-dependent and ABA-independent pathways. The ABA-independent pathway includes the CBF/DREB] genes (Liu et al., 1998; Thomashow, 1999; Shinozaki and Yamaguchi-Shinozaki, 2000; Shinozaki et al., 2003). More recently, Knight et al. (2004) showed that the activation of the CBF pathway could also occur via the ABA pathway; 100 uM ABA treatment for 1 hour was sufficient to increase the transcript and the protein level of the CBF], CBF 2 and CBF 3 genes. Knight et al. (2004) suggested that the CBF1- 3 genes, which are activated by cold and ABA treatment, might give them a distinctive role in plant adaptation under both fieezing and drought conditions, in contrast to CBF 4 which is strictly activated under drought and ABA treatment (Haake et al., 2002). This possibly adds another layer of complexity to signaling in plants under abiotic stress conditions. 30 II-B-4d Effect of CBF/DREB-overexpression on plant metabolome The CBF/DREB gene families are transcriptional activators that directly or indirectly work as master switches in regulating transcript in plants in response to dehydration- inducing conditions (Thomashow, 2001). Interests in exploring large scale changes in plant metabolome, as a direct result of CBF/DREB expression, has been recently investigated in Arabidopsis (Cook et al., 2004). Cook‘et al. (2004) included three ecotypes of Arabidopsis thaliana, differing in tolerance to freezing temperatures [Cape Verde Islands-1 (Cv-l), Wassilewskija-Z (W s- 1), and Columbia (CM]; the first is less freezing tolerant than the latter two. Using a GC— time—of-flight MS method to assess large scale changes in metabolic profile, at least 325 low molecular weight compounds (carbohydrates, amines, organic acids, and other polar molecules) increased 2-fold or more in Ws-plants in response to 14 days of cold treatment. Of those 325 compounds, 114 increased at least 5-fold higher compared to the control non-cold acclimated plants. On the other hand, in the Cv-l ecotype, only 269 compounds increased in response to cold acclimation, of which 244 were common with the Ws-l ecotype. The finding that only 53 out of the 269 compounds had at least 5-fold increase in Cv-l cold acclimated plants compared to 114 in the freezing tolerant Ws-l ecotype may explain why these ecotypes differ in their ability to withstand freezing temperatures. Cook et al. (2004) went on to compare the metabolome profile of CBF 3/DREBI a overexpressing Arabidopsis plants and cold acclimated-wild-type Col plants. These experiments revealed that of the 325 metabolites that increased in response to low temperature, 256 also significantly increased in CBF 3/DREBI a overexpressing plants. Of 31 those compounds, 102 increased at least 5-fold in the CBF3/DREBla overexpressing plants. These results clearly demonstrate the similarity between the metabolome of CBF3-expressing plants and the metabolome of cold-acclirnated wild-type plants. Cook et al. (2004) suggested that the dramatic increase in proline and low-molecular-weight soluble carbohydrates is a signature of the CBF regulon. II-B-4e Conservation of the CBF/DREB system among plant species CBF1/DREB] b homologs have been identified in Arabidopsis thaliana (Stockinger et al., 1997; Haake et al., 2002), canola (J aglo et al. 2001), barley (Choi et al., 2002), tomato (J aglo et al. 2001), rice (Choi et al., 2002; Dubouzet et al., 2003), strawberry and scur cherry (Owens et al., 2002), suggesting that the CBF/DREB system is highly conserved throughout the plant kingdom, including both dicots and monocots. J aglo et al. (2001) reported the presence of other CBF/DREB] homologs in Arabidopsis (DREBIe and DREBIf), and other plants. Canola (Brassica napus), which like Arabidopsis a member of the Brassica family, encodes two different CBF -1ike proteins that share approximately 76% homology to the Arabidopsis CBF] gene; similarly, BnCBF] and BnCBF 2, accumulated in canola plants within 30 minutes after transferring plants into cold nonfreezing temperature (J aglo et al., 2001). The expression of the BnCBF] and BnCBF 2 was followed by accumulation of the B. napus-CORI5a ortholog, B22115. J aglo et al. (2001) also reported the presence of CBF homologs in the more distantly related, chilling sensitive species tomato (Lycopersicon esculentum L). Similar to the Arabidopsis CBF], CBF 2 and CBF 3 genes, Zhang et al. (2004) found that LeCBF], 32 LeCBF2 and LeCBF 3, are organized in tandem and all lack introns. Interestingly, LeCBFI was induced only in response to cold non-fieezing conditions after about 30 minutes, reaching maximum induction after 2 hours, while LeCBF 2 and LeCBF 3 were not induced under cold. LeCBF] , LeCBF 2 and LeCBF 3 responded very weakly to mechanical agitation, drought, salinity and ABA treatment (Zhang et al., 2004). Owens et al. (2002), reported that CBF1-orthologs also exist in sour cherry (Prunus cerasus L.) and strawberry (F ragaria X ananassa Duchesne). The putative orthologs of sour cherry PcCBF] and strawberry F aCBF ] shared about 48% similarity at the amino acid level to the Arabidopsis CBF1 protein (Owens et al., 2002). Distantly related plants such the monocots rye (Secale cereale L.), wheat (T riticum aestivum L.), barley (Hordeum vulgare L.) and rice (Oryza sativa) also have CBF-like homologues (J aglo et al., 2001; Choi et al., 2002). The rye and wheat polypeptides had 30% and 34% sequence homology, respectively, to the Arabidopsis CBF1 polypeptide. Choi et al. (2002) identified a sequence, OsCBF 3, from the rice genome database similar to the CBF 3 gene fi'om Arabidopsis. This in turn was used to screen a barley BAC library, leading to identification of a barley CBF 3 ortholog. Expression of the barley CBF gene, HvCBF 3, was found to be cold induced (Choi et al., 2002). II-B-4f Engineering dehydration-stress tolerance in plants using the CBF/DREB system Jaglo-Ottosen et al. (1998) achieved a breakthrough by demonstrating that overexpression of a transcription factor in plants would result in activation a cascade of genes directly/indirectly involved in abiotic stresses tolerance. Transgenic, non- 33 acclimated Arabidopsis plants overexpressing the CBF] gene were more freezing tolerant than their non-acclimated control counterpart as determined by electrolyte leakage (a test used to determine membrane damage) and whole plant fieezing test assays. ELso values (the fieezing temperature that results in release of 50% of tissue electrolytes) indicated that CBF] overexpressing Arabidopsis plants were significantly more tolerant to freezing than wild type (3.3 °C difference). Similarly, Kasuga et al. (1999) overexpressed DREBIA in Arabidopsis and showed that DREBIA enhanced tolerance to drought, salt and fieezing in transgenic plants: 69.2% of the transgenics survived a 14 day dehydration treatment, whereas no wild type plants survived (Table 1-1 D). When exposed to —6 C° for 2 days, followed by 5 days exposure to 22 C°, less than 10% of the nonacclimated control plants survived while more than 75% of the transgenics survived. Moreover, 78.6% of the transgenics were able to survive a treatment of dipping in 600 mM NaCl solution for 2 h, before transplanting into pots, while only 17.9% of the wild-type plants survived this treatment (Kasuga et al., 1999). Arabidopsis plants overexpressing CBF 4 gene also were tolerant to freezing and drought stresses (Haake et al., 2002). Freezing tests of nonacclimated plants revealed that while only 1% of wild type plants survived freezing at —10°C for 20 hours, the range of survival was 52-100% in the transgenic plants, depending on the expression level of CBF 4. The same trends were achieved when plants were subjected to 9 days of dehydration; only 2% of the wild type plants survived compared to 87% of transgenic plants (Table 1-1D). Transgenic Arabidopsis plants overexpressing CBF] (J aglo-Ottosen et al., 1998) exhibited numerous physiological changes associated with CBF expression and increased 34 dehydration stress resistance including increased membrane stability, increased compatible solute accumulation and the ability to scavenge for reactive oxygen species (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gihnour et al., 2000; Hsieh et al. 2002a). Non-acclirnated transgenic Arabidopsis plants had lower EL50 values in response to freezing temperatures than their nontransgenic counterparts indicating that CBF/DREB- overexpressing plants suffered less membrane damage. CBF/DREB-overexpressing plants also accumulated higher levels of osmolytes e. g., (proline and soluble sugars) compared to the nontransgenic controls (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gilmour et al., 2000; Hsieh et al. 2002b). CBF3- overexpressing Arabidopsis lines had approximately 3-fold higher level of total soluble sugars and 5-fold higher proline levels compared to the nontransgenics (Gihnour et al., 2000). This increase in proline was accompanied with an increase in the transcript level of A-pyrroline-S-carboxylate synthetase (P5 CS), a key enzyme in proline synthesis. Haake et al. (2002) also reported similar results when overexpressing the CBF 4 gene in Arabidopsis, a gene that is only activated under drought and ABA treatment. In canola, non-cold acclimated plants overexpressing the Arabidopsis-CBF1, CBF2 or CBF 3 genes had EL50 value of —6 C° while their nontransgenic counterparts had a value of -3 C° (J aglo et al., 2001). The EL50 values were —l2.7 C° and -8.1 C°, for the transgenic and the nontransgenic controls, when plants were cold acclimated before performing the electro-leakage test (J aglo et al., 2001). Similarly, constitutive expression of the Arabidopsis-CBF genes in canola plants resulted in activation oan115 and B2128 (orthologs to the Arabidopsis COR] 5a and COR6.6 genes) without cold acclimation. Thus, canola appears to have a cold-response pathway that is very close to Arabidopsis 35 Arabidopsis CBF genes also can confer increased dehydration stress resistance to other less closely related species (Hseih et at., 2002a; Hseih et al., 2002b; Owens et al., 2002; Kasuga et al., 2004). Kasuga et al. (2004) used the stress-inducible rd29A promoter to drive the expression of DREBIA/CBF 3 gene in tobacco plants. Tobacco plants expressing DREBIA/CBF 3 under the dehydration inducible promoter rd29A were more drought tolerant compared to wild type plants and had higher photosynthetic activity under drought and cold-nonfieezing temperature. Similarly, tomato plants overexpressing the Arabidopsis CBF] gene were more dehydration stress tolerant than wild type plants; 83.3% of the transgenic plants survived a 4 week drought treatment while less than 6% of the nontransgenics survived this treatment (Hseih et al., 2002b). The transgenic plants had 3-4 fold more proline under nonstressed conditions compared to the nontransgenic controls. Proline levels increased under stress conditions in both transgenic and nontransgenic plants but were 30-60% higher in plants overexpressing the Arabidopsis-CBF] gene (Hseih et al., 2002b). Interestingly, whereas overexpression experiments of LeCBF] in Arabidopsis resulted in accumulation of COR gene transcripts and increased the EL50 values by —2.5 °C in the overexpressing lines, no differences in the electro-leakage were observed in transgenic tomato lines overexpressing AtCBF] , AtCBF 3 or LeCBF] (Zhang et al., 2004). Owens et al. (2002), overexpressed the A. thaliana-CBF] gene in strawberry and tested the plants for their ability to withstand freezing temperature compared to nontransgenic plants. The EL50 values indicated that CBF] expressing strawberry plants were significantly more tolerant to freezing temperatures than wild type (-10.3 °C in the transgenic compared to -6.4 °C in wild type). 36 III- The objectives of this dissertation The above examples clearly indicate that various types of genes can be used to engineer dehydration stress tolerance in plants. The objective of this work was to investigate the possibility to engineer enhanced dehydration stress tolerance in cucumber plants following two approaches. Cucumber plants are known to be sensitive to salinity and drought (Mass and Hoffinan, 1977). Salinity and drought has been reported to have strong negative effects on cucumber plants, especially on seed germination and seedling emergence, leaf expansion rate, photosynthesis, fruit set, as well as fruit growth rate and fruit quality (Chartzoulakis, 1994; Navazio and Staub, 1994; Ho and Adams, 1994; Tazuki, 1997; Serce et al., 1999; Drozdova et al., 2004). At the time I started this work, there was no published data on the evaluation of genetically engineered plants for enhanced salt or drought tolerance under field conditions. To our knowledge, there are only two published reports, both of which came in late 2004 that evaluated genetically engineered plants for enhanced dehydration stress tolerance under field conditions (Table 1-1). Quan et al. (2004) reported enhanced grain yield production by transgenic maize plants expressing a gene for betaine aldehyde dehydrogenase following drought stress period of 21 days, and Xue et al. (2004) tested wheat plants expressing the Arabidopsis tonoplast H‘L/Na+ antiporter gene for their ability to grow in saline soil and reported higher grain production in the transgenic plants compared to the nontransgenic controls. The aim of this thesis is to: l-Investi gate and test the possibility to improve dehydration-stress tolerance in cucumber (Cucumis sativus L.), plants by following two strategies 37 (a) Induction of mannitol production in cucumber plants by using the celery M6PR gene, or (b) Induction of an array of adaptive responsive mechanisms that are associated with dehydration-stress tolerance in plants by expressing the A. thaliana-transcriptional regulators CBF/DREB] genes in cucumber plants. 2- Evaluate and validate the performance and mu yield of transgenic cucumber plants for their ability to withstand dehydration stress under greenhouse and field conditions. 38 References Abe H, Urao T, Ito T, Seki M, Shinozaki K, Yarnaguchi-Shinozaki K (2002) Arabidopsis AtMYCZ (bHLH) and AtMYBZ (MyB) firnction as transcriptional activators in abscisic acid signaling. Plant Cell 15: 63-78. Abebe T, Guenzi AC, Martin B, Cushman J C (2003) Tolerance of mannitol-accumulating transgenic wheat to water stress and salinity. Plant Physiol. 131: 1748-1755. Ain-Lhout F, Zunzunegui FA, Diaz Barradas MC, Tirado R, Clavijio A, Garcia Novo F (2001) Comparison of proline accumulation in two Mediterranean shrubs subjected to natural and experimental water deficit. Plant Sci] 230: 175—183. Allen GJ, Schroeder J I (1998) Cyclic ADP-ribose and ABA signal transduction. Trends Plant Sci. 3: 123-125. Apse MP, Aharon GS, Snedden WA, Blumwald E (1999) Salt tolerance conferred by overexpression of a vacuolar Na+/H+ antiport in Arabidopsis. Science 285: 1256- 1258. Arrillaga I, Mascarell RG, Gisbert C, Sales E, Montesinos C, Serrano R, Moreno V (1998) Expression of the yeast HAL 2 gene in tomato increases the in vitro salt tolerance of transgenic progenies. Plant Sci. 136: 219—226. Artus NN, Uemura M, Steponkus PL, Gilmour SJ, Lin CT, Thomashow MF (1996) Constitutive expression of the cold-regulated Arabidopsis thaliana COR15a gene affects both chloroplast and protoplast freezing tolerance. Proc Natl Acad Sci USA 93: 701 -713. Babu RC, Shashidhar HE, Lilley JM, Thanh ND, Ray JD, Sadasivam S, Sarkarung S, O’Toole JC, Nguyen HT (2001) Variation in root penetration ability, osmotic adjustment and dehydration tolerance among accessions of rice adapted to rainfed lowland and upland ecosystems. Plant Breeding 120: 233-238. Badawi GH, Yamauchi Y, Shimada E, Sasaki R, Kawano N, Tanaka K, Tanaka K (2004) Enhanced tolerance to salt stress and water deficit by overexpressing superoxide dismutase in tobacco (Nicotiana tabacum) chloroplasts. Plant Sci. 166: 919-928. Baker SS, Wilhelm KS, Thomashow MF (1994) The 5’ region of the Arabidopsis thaliana cor15a has cis-acting element that confer cold-, drought-and ABA- regulated gene expression. Plant M01. Bio. 24: 701-713. Bhattacharya RC, Maheswari M, Dineshkumar V, Kirti PB, Bhat SR, Chopra VL (2004) Transformation of Brassica oleracea var. capitata with bacterial betA gene enhances tolerance to salt stress. Scientia Hort. 100: 215-227. Bierhuizen JF (1969) Water quality and yield depression. I C W No: 163-173. 39 Blum A (1988) Salinity resistance. In: Plant breeding for stress environment. Boca Raton, CRC Press 163-179. Blum A (1989) Breeding methods for drought resistance. In Jones HG, Flowers T], Jones MB eds. Plant under stress: Biochemistry, physiology and ecology and their application to plant “improvement. Cambridge University Press, 217-234. Blumwald E, Aharon GS, Apse MP (2000) Sodium transport in plant cells. Biochem. Biophys. Acta 1465: 140-151. Blumwald E, Poole RJ (1985) Na+/Hi antiport in isolated tonoplast vesicles from storage tissue of Beta vulgaris. Plant Physiol. 78: 163-167. Bohnert HJ, Jensen RG (1996) Strategies for engineering water-stress tolerance in plants. Trends Biotech. 14: 89-97. ' Bohnert HJ, Su H, Shen B (1999) Molecular mechanisms of salinity tolerance. In: Molecular responses to cold, drought, heat and salt stress in higher plants. Landes Company 29-62. Bor M, Ozdernir F, Turkan I (2003) The effect of salt stress on lipid peroxidation and antioxidants in leaves of sugar beet Beta vulgaris L. and wild beet Beta maritima L. Plant Sci. 164: 77-84. Bordas M, Montesinos C, Dabauza M, Salvador A, Roig LA, Serrano R, Moreno V (1997) Transfer of the yeast salt tolerance gene HALl to Cucurnis melo L cultivars and in vitro evaluation of salt tolerance. Transgenic Res. 6: 41-50. Boscherini G, Muleo R, Montagni G, Cinelli F, Pellegrini MG, Bemardini M, Buiatti M (1999) Characterization of salt tolerant plants derived from a Lycopersicon esculentum Mill. somaclone. J Plant Physiol. 155: 613-619. Bouchereau A, Aziz A, Larher F, Tanguy J M (1999) Polyarnines and environmental challenges: recent development. Plant Sci. 140: 103-125. Bray EA (1994) Molecular responses to water deficit. Plant Physiol. 103: 1035-1040. Bray EA (1997) Plant responses to water deficit. Trends Plant Sci. 2: 48-54. Campbell SA, Close TJ (1997) Dehydrins: genes, proteins and association with phenotypic traits. New Phytol 137: 61-74. Capell T, Bassie L, Christou P (2004) Modulation of the polyamine biosynthetic pathway in transgenic rice confers tolerance to drought stress. Proc Natl Acad Sci. USA. 101: 9909-9914. Chandler PM, Robertson M (1994) Gene expression regulated by abscisic acid and its relationship to stress tolerance. Annu Rev Plant Physiol Plant Mol Bio. 45: 113-141. Chartzoulakis KS (1994) Photosynthesis, water relation and leaf growth of cucumber exposed to salt stress. Scientia Hort. 59: 27-35. 40 Cheng NH, Liu JZ, Nelson RS, Hirschi KD (2004) Characterization of CXIP4, a novel Arabidopsis protein that activates the H‘L/Ca2+ antiporter, CAX] . FEBS Letters 559: 99-106. Cherry,JI-I, eds (1989) Environmental Stress in Plants: Biochemical and physiological mechanisms associated with environmental stress tolerance in plants. NATO ASI series vol 619. Springer-Verlag Berlin Heidelberg 27. Chinnusamy V, Ohta M, Kanrar S, Lee B, Hong X, Agarwal, M, Zhu J-K (2003) ICEl: a regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev 17: 1-12. Choi DW, Rodrigues EM, Close TJ (2002) Barley CBF 3 gene identification, expression pattern, and map location. Plant Physiol 129: 1781-1787. Close TJ (1996) Dehydrins: Emergence of a biochemical role of a family of plant dehydration proteins. Physiol Planta 97: 795-803. Close TJ (1997) A commonality in response of plants to dehydration and low temperature. Physiol Plant 100: 291-296. Cook D, Fowler S, Fiehn O, Thomashow MF (2004) A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. Proc Natl Acad Sci USA. 101: 15423-15248. Cushman JC, Bohnert H (2000) Genomic approaches to plant stress tolerance. Curr Opin Plant Bio 3: 117-124. De-Ronde J A, Spreeth MH, Cress WA (2000) Effect of antisense L-A-pyrroline-S- carboxylate reductase transgenic soybean plants subjected to osmotic and drought stress. Plant Growth Reg 32: 13-26. Drosdova IS, Pustovoitova TN, Dzhibladdze TG, Barabnshchikova NS, Zhdanova NE, Maevskaya SN, Bukhov NG (2004) Endogenous control of photosynthetic activity during progressive drought: influence of final products of photosynthesis. Russ J Plant Physiol 51: 742-750. DubouzetlJ G, Sakurna Y, Ito Y, Kasuga M, Dubouzet, EG, Miura S, Seki M, Shinozaki K, Yamaguchi-Shinozaki K (2003) OsDREB genes in rice, Oryza sativa L., encode transcription activators that function in drought-, high-salt- and cold-responsive gene expression. Plant J 33: 751—763. Dure L III, Greenway SC, Galau GA (1981) Developmental biochemistry of cotton seed embryogenesis and germination. XIV. Changing mRNA populations as shown by in vitro and in vivo protein synthesis. Biochemistry 20: 4162-4168. Everard JD, Cantini C, Grumet R, Plummer J, Loescher WH (1997) Molecular cloning of mannose-6-phosphate reductase and its developmental expression in celery. Plant Physiol 113: 1427-1435. 41 Everes D, Ovemey S, Simon P, Greppin H, Hausman JF (1999). Salt tolerance of solanum tuberosum L. overexpressing the heterologous osmotin-like protein. Biol Planta 42: 105-112. Flowers TJ (2004) Improving crop salt tolerance. J Exp. Botany 55: 307-319. Flowers TJ, Yeo AR (1995) Breeding for salinity tolerance in crop plants: where next? Aust J Plant Physiol 22: 875-884. Foolad MR (1999) Comparison of salt tolerance during seed germination and vegetative growth in tomato by QTL mapping. Genome 42: 727-734. Forster BP (2001) Mutation genetics of salt tolerance in barley: An assessment of Golden Promise and other semi-dwarf mutants. Euphytica 120: 317-328. Fowler S, Thomashow MF (2002) Arabidopsis transcriptome profiling indicated that multiple regulatory pathway are activated during cold acclimation in addition to the CBF cold response pathway. Plant Cell 14: 1675-1690. Fukushima E, Arata Y, Endo T, Sonnewald U, Sato F (2001) Improved salt tolerance of transgenic tobacco expressing apoplastic yeast- derived invertase. Plant Cell Physiol 42: 245—249. Garg AK, Kim J-K, Owens TG, Ranwala AP, Choi YD, Kochian LV, Wu RL (2002) Trehalose accumulation in rice plants confers high tolerance levels to different abiotic stresses. Proc Natl Acad Sci USA 99: 15898—15903. Gaxiola RA, Li J, Undurraga S, Dang LM, Allen GJ, Alper SL, Fink GR (2001) Drought- and salt-tolerant plants result from overexpression of the A VP] Hl-purnp. Proc Natl Acad Sci. USA. 98: 11444—11449. Gilmour S, Sebolt AM, Slazar MP, Everard JD, Thomashow MF (2000) Overexpression of the Arabidopsis CBF3 transcriptional activator mimics multiple biochemical changes associated with cold acclimation. Plant Physiol 124: 1854-1865. Gilmour SJ, Artus NN, Thomashow MF (1992) cDNA sequence analysis and expression of two cold-regulated genes of Arabidopsis thaliana. Plant Mol Biol 18: 13-21. Gilmour SJ, Zarka DG, Stockinger EJ, Salazar MP, Houghton JM, Thomashow MF (1998) Low temperature regulation of the Arabidopsis CBF family of AP2 transcriptional activators as an early step in cold-induced COR gene expression. Plant J 16: 433-442. Gorham J (1996) Glycinebetaine is a major nitrogen-containing solute in the malvaceae. Phytochemistry 43: 367-3 69. Greenway H, Munns R (1980) Mechanisms of salt tolerance in non-halophytes. Annu Rev Plant Physiol 31: 149-190. 42 Grumet R, Hanson AD (1986) Genetic evidence for an osmoregulatory function of glycinebetaine accumulation in barley. Aust J Plant Physiol 13: 353-364. Guy C1, Niemi KJ, Brambl R (1985) Altered gene expression during cold acclimation of spinach. Proc Natl Acad Sci USA 82: 3673-3677. Haake, V, Cook D, Riechmann J L, Pineda O, Thomashow MF, Zhang JZ (2002) Transcription factor CBF 4 is a regulator of drought adaptation in Arabidopsis. Plant Physiol 130: 639-649. Halfter U, Ishitani M, Zhu J -K (2000) The Arabidopsis SOSZ protein kinase physically interacts with and is activated by the calcium-binding protein SOS3. Proc Natl Acad Sci USA 97: 3730-3734. Halliwell B, Grootveld, M and Gutteridge, JMC (1988). Methods for the measurement of hydroxyl radicals in biochemical systems: deoxyribose degradation and aromatic hydroxylation. Methods Biochem. Anal. 33: 59—90. Hasegawa PM, Bressan RA, Zhu J K, Bohnert HJ (2000) Plant cellular and molecular responses to high salinity. Annu Rev Plant Physiol Plant Mol Biol 51: 463-499. Hayashi H, Alia I, Mustardy L, Deshniurn P, Ida M, Murata N (1997) Transformation of Arabidopsis thaliana‘ with the codA gene for choline oxidase; accumulation of glycinebetaine and enhanced tolerance to salt and cold stress. Plant J 12: 133-142. Hirayarna T, Ohto C, Mizoguchi T, Shinozaki K (1995) A gene encoding a phosphatidyl inositol-specific phospholipase C is induced by dehydration and salt stress in Arabidopsis thaliana. Proc Natl Acad Sci USA 92: 3903-3907. Ho LC, Adams P (1994) regulation of partitioning of dry matter and calcium in cucumber in relation to fiuit growth and salinity. Ann Bot 73: 539-545. Hoekstra FA, Golovina EA, Buitink J (2001) Mechanisms of plant desiccation tolerance. Trends Plant Sci 6: 431-438. Holmberg N, Bulow L (1998) Improving stress tolerance in plants by gene transfer. Trends Plant Sci 3: 61-66. Holmstrom KO, Somersalo S, Manda] A, Palva TE, Welin B (2000). Improved tolerance to salinity and low temperature in transgenic tobacco producing glycine betaine. J Exp Bot 51: 177-185. Holmstrom KO, Welin B, Manda] A, Kristiansdottir I, Teeri TH, Larnark T, Strom AR, Palva ET (1994) Production of the Escherichia coil betaine-aldehyde dehydrogenase, an enzyme required for the synthesis of the osmoprotectant glycine betaine in transgenic plants. Plant J 6: 749-758. 43 Hong Z, Lakkineni K, Zhang K, Verrna DPS (2000) Removal of feedback inhibition of Al-pyn'oline-S-carboxylate synthetase resulted in increased proline accumulation and protection of plants from osmotic stress. Plant Physiol 122: 1129-1136. Horvath DP, McLarney BK, Thomashow MF (1993) Regulation of Arabidopsis thaliana L. (Heyn.) cor78 in response to low temperature. Plant Physiol 103: 1047-1053. Hoshida H, Tanaka Y, Hibino T, Hayashi Y, Tanaka A, Takabe T, Takabe T (2000) Enhanced tolerance to salt stress in transgenic rice that overexpresses chloroplast glutarrrine synthetase. Plant Mol Biol 43: 103-111. Hsieh T-H, Lee J -T, Chamg YY, Chan M-T (2002b) Tomato plants ectopically expressing Arabidopsis CBF] show enhanced resistance to water deficit stress. Plant Physiol 130: 618-626. Hsieh T-H, Lee J -T, Yang P-Z, Chiu LH, Chamg YY, Wang Y-C, Chan M-T (2002a) Heterologous expression of the Arabidopsis C-repeat/Dehydration Response Element Binding factor 1 gene confers elevated tolerance to chilling and oxidative stresses in transgenic tomato. Plant Physiol 129: 1086-1094. Huang J, Hirji R, Adam L, Rozwadowski KL, Harnmerlindl JK, Keller WA, Selvaraj G (2000) Genetic engineering of glycinebetaine production toward enhancing stress tolerance in plants: metabolic limitations. Plant Physiol 122: 747-756. Imai R, Chang L, Ohta A, Bray EA, Takagi M (1996) A lea-class gene of tomato confers salt and freezing tolerance when overexpressed in Saccharomyces cerevisiae. Gene 170: 243-248. Ingram, J Bartels D (1996) The molecular basis of dehydration tolerance in plants. Annu Rev Plant Physiol Plant Mol Bio 47: 377-403. Ishitani M, Liu J, Halfter U, Kim CS, Wei M, Zhu J-K (2000) S083 function in plant salt tolerance requires myristoylation and calcium-binding. Plant Cell 12: 1667-1677. Ishitani M, Xiong LM, Stevenson B, Zhu J K (1997) Genetic analysis of osmotic and cold stress signal transduction in Arabidopsis: Interactions and convergence of abscisic acid-dependent and abscisic acid-independent pathways. Plant Cell 9: 1935-1949. J aglo-Ottosen KR, Gilmour SJ, Zarka DG, Schabenberger O, Thomashow MF (1998) Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 280: 104-106. J aglo-Ottosen KR, Kleff S, Amundsen KL, Zhang X, Haake V, Zhang JZ, Deits T, Thomashow MF (2001). Component of the Arabidopsis C-repeat/dehydration- responsive element binding factor cold-response pathway are conserved in Brassica napus and other plant species. Plant Physiol 127: 910-917. Jain RK, Selvaraj G (1997) Molecular genetic improvement of salt tolerance in plant. Biotech Ann Rev 3: 245-267. Jang IC, Oh SJ, Seo J S, Choi WB, Song SI, Kim CH, Kim WS, Seo HS, Choi YD, Nahm BH, Kim J -K (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 Physiol 131: 516—524. Jia G-X, Zhu Z-Q, Chang F-Q, Li YX (2002) Transformation of tomato with the BADH gene from Atriplex improves salt tolerance. Plant Cell Rep 21 :141—146. Kang JY, Choi H, Irn MY, Kim SY (2002) Arabidopsis basic leucine zipper proteins that mediate stress-responsive abscisic acid signaling. Plant Cell 14: 343—357. Karakas B, Ozias-Akins P, Stuchnoff C, Suefferheld M, Rieger M (1997) Salinity and drought tolerance of mannitol-accumulating transgenic tobacco. Plant Cell Environ 20: 609-616. Kasuga M, Liu Q, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1999) Improving plant drought, salt, and fieezing tolerance by gene transfer of a single-inducible transcription factor. Nature Biotech 17: 287-291. Kasuga M, Miura S, Shinozaki K, Yamaguchi-Shinozaki K (2004) A combination of the Arabidopsis DREBIA gene and stress-inducible rd29A promoter Improved drought- and low-temperature stress tolerance in tobacco by gene transfer. Plant Cell Physiol 45: 346-350. Kasukabe Y, He L, Nada K, Misawa S, Ihara I, Tachibana S (2004) Overexpression of sperrnidine synthase enhances tolerance to multiple environmental stresses and up- regulates the expression of various stress- regulated genes in transgenic Arabidopsis thaliana. Plant Cell Physiol 45: 712—722. Khush GS (1999) Green revolution: preparing for the 21St century. 42: 646-655. Kim J -B, Kang JY, Kim SY (2004) Over-expression of a transcription factor regulating ABA responsive gene expression confers multiple stress tolerance. Plant Biotech J 2: 459-466. Kishitani S, Takanami T, Suzuki M, Oikawa M, Yokoi S, Ishitani M, Alvarez-Nakase AM, Takabe T, Takabe T (2000) Compatibility of glycinebetaine in rice plants: evaluation using rice plants with a gene for peroxisomal betaine aldehyde dehydrogenase fi'om barley. Plant Cell Environ 23: 107-114. Kishor PBR, Hog Z, Miao GH, Hu CAA, Verrna DPS (1995) Overexpression of Al- pyrroline—S-carboxylate synthetase increases proline production and confers osmotolerance in transgenic plants. Plant Physiol 108: 1387-1394. Knight H, Knight MR (2001) Abiotic stress signaling pathways: specificity and cross- talk. Trends Plant Sci 6: 262-267. Knight H, Trewavas AJ, Knight MR (1997) Calcium signaling in Arabidopsis thaliana responding to drought and salinity. Plant J 12: 1067-1078. 45 Knight H, Zarka DG, Okamoto H, Thomashow MF, Knight MR (2004) Abscisic acid induces CBF gene transcription and subsequent induction of cold-regulated genes via the CRT promoter element. Plant Physiol 135: 1710-1717. Kumar S, Dhingra A, Daniell H (2004) Plastid-expressed betaine aldehyde dehydrogenase gene in carrot cultured cells, roots, and leaves confers enhanced salt tolerance. Plant Physiol 136: 2843—2854. Laurie S, Feeney KA, Maathuis FJM, Heard PJ, Brown SJ, Leigh RA (2002) A role for HKTl in sodium uptake by wheat roots. Plant J 32: 139—149. Lee JH, Van Montagu M, Verbruggen N (1999) A highly conserved kinase is an essential component for stress tolerance in yeast and plant cells (Arabidopsis thaliana) stress signaling transcriptional regulation. Proc Natl Acad Sci USA 96: 5873—5877. Lee JT, Prasad V, Yang PT, Wu JF, Ho THD, Chang YY, Chan MT (2003). Expression of Arabidopsis CBF] regulated by an ABA/stress inducible promoter in transgenic tomato confers stress tolerance without affecting yield. Plant Cell Environ 26: 1181— 1190. Lee S-B, Kwon H-B, Kwon S-J, Park S-C, J eong M-J, Han SE, Byun M-O, Daniell H (2003) Accumulation of trehalose within transgenic chloroplasts confers drought tolerance. Mol Breeding 11: 1—13. Leung J and Giraudat J (1998) Abscisic acid signal transduction. Annu Rev Plant Physiol Plant Mol Biol 49: 199-222. Lichtentaler HK (1995) Vegetation stress: an introduction to the stress concept in plants. J Plant Physiol 148: 4-14. Lin C, Thomashow MF (1992) DNA sequence analysis of a complementary DNA for cold-regulated Arabidopsis gene cor15 and characterization of the COR15 polypeptide. Plant Physiol 99: 519-525. Liu J, Ishitani M, Halfler U, Kim CS, Zhu J -K (2000) The Arabidopsis thaliana SOS2 gene encodes a protein kinase that is required for salt tolerance. Proc Natl Acad Sci USA 97: 3735-3740. Liu J, Zhu J -K (1998) A calcium sensor homolog required for plant salt tolerance. Science 280: 1943-1945. Liu Q, Kasuga M, Salcuma Y, Abe H, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1998) Two transcription factors, DREB] and DREBZ, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low- temperature-responsive gene expression, respectively, in Arabidopsis. Plant «Cell 10: 1391-1406. Liu Y, Wang GY, Liu JJ, Peng XX, Xie YJ, Dai JR, Guo SW, Zhang FS (1999) Transfer of E. coli gutD gene into maize and regeneration of salt-tolerant transgenic plants. Sci China Ser 42: 90-95. 46 Maggie A, Bressan RA, Ruggiero C, Xiong L, Grillo S (2003) Salt tolerance: Placing advances in molecular genetics into a physiological and agronomic context. In Abiotic Stresses in Plants. Toppi LSD and Skowronska P (eds.). Kluwer Acad. Publ. 53-69. Mali PC, Mehta SL (1977) Effect of drought on enzymes and free proline in rice varieties. Phytochemistry 16: 1355-1357. Mass EV, Hoffman GJ (1977) Crop salt tolerance-current. J Irri Drain Div 103: 115-134. Mazel A, Leshem Y, Tiwari BS, Levine A (2004) Induction of salt and osmotic stress tolerance by overexpression of an intracellular vesicle trafficking protein AtRab7 (AtRabG3e). Plant Physiol 134: 118-128. McWilliam JR (1986) The national and international importance of drought and salinity effects on agricultural production. Aust J Plant Physiol 13: 1-13. Mittler R, Zilinskas BA (1994) Regulation of pea cytosolic ascorbate peroxidase and other antioxidant enzymes during the progression of drought stress and following recovery from drought. Plant J 5: 397-405. Mittova V, Guy M, Tal M, Volokita M (2002) Response of the cultivated tomato and its wild salt-tolerant relative Lycopersicon pennellii to salt-dependent oxidative stress: increased activity of enzymes in root plastids. Free Radic Res 36: 195-202. Mittova V, Guy M, Tal M, Volokita M (2004) Salinity up-regulates the antioxidative system in root mitochondria and peroxisomes of the wild salt-tolerant tomato species Lycopersicon pennellii. J Exp Bot 55: 1105-1112. Mukhopadhyay A, Vij S, Tyagi AK (2004) Overexpression of a zinc-finger protein gene from rice confers tolerance to cold, dehydration, and salt stress in transgenic tobacco. Proc Natl Acad Sci USA 101: 6309—6314. Muller J, Boller T, Wiernken A (1995) Trehalose and trehalase in plants: recent developments. Plant Sci 112: 1-9. Munns R (1993) Physiological processes limiting plant growth in saline soils: some dogmas and hypotheses. Plant Cell Environ 16: 15-24. Nakamura T, Yokota S, Muramoto Y, Tsutsui K, Oguri Y, Fukui K, Takabe T (1997) Expression of a betaine aldehyde dehydrogenase gene in rice, a glycinebetaine non accumulator, and possible localization of its protein in peroxisomes. Plant J 11: l 1 15-1 120. Nanjo T, Kobayashi M, Yoshiba Y, Kakubari Y, Yamaguchi-Shinozaki K, Shinozaki K (1999) Antisense suppression of proline degradation improves tolerance to freezing and salinity in Arabidopsis thaliana. FEBS Let 461: 205-210. Navazio JP, Staub J E (1994) Effects of soil-moisture, cultivar, and postharvest handling on pillowy fruit disorder in cucumber. J Amer Soc Hort Sci 119: 1234-1242. 47 Novillo F, Alonso J M, Ecker JR, Salinas J (2004) CBF2/DREB1C is a negative regulator of CBF1/DREBIB and CBF3/DREBIA expression and plays a central role in stress tolerance in Arabidopsis. Proc Natl Acad Sci USA 101: 3985-3990. Ohta M, Hayashi Y, Nakashima A, Harnada A, Tanaka A, Nakamura T, Hayakawa T (2002) Introduction of a Na+/H+ antiporter gene fiom Atriplex gmelini confers salt tolerance to rice. FEBS Let 532: 279-282. Owens CL, Thomashow MF, Hancock JF, Iezzoni AF (2002) CBF] orthologs in sour cherry and strawberry and the heterologous expression of CBF] in strawberry. J Amer Soc Hort Sci 127: 489-494. Palrngren MG (1998) Proton gradients and plant growth: Role of the plasma membrane H+-ATPase. Adv Bot Res 28: 1-70. Pandey GK, Reddy VS, Reddy MK, Deswal R, Bhattacharya A, Sopory SK (2002) Transgenic tobacco expressing Entamoeba histolytica calcium binding protein exhibits enhanced growth and tolerance to salt stress. Plant Sci 162: 41—47. Pardo J M, Reddy MP, Yang S, Maggio A, Huh G-H, Matsurnoto T, Coca MA, Paino- D’Urzo M, Koiwa H, Yun D-J, Watad AA, Bressan RA, Hasegawa PM (1998) Stress signaling through Ca21calmodulin-dependent protein phosphatase calcineurin mediates salt adaptation in plants. Proc Natl Acad Sci USA 95: 99681-9686. Park J M, Park 0], Lee S-B, Ham B-K, Shin R, Pack K-H (2001). Overexpression of the tobacco T si 1 gene encoding an EREBP/APZ—type transcription factor enhances resistance against pathogen attack and osmotic stress in tobacco. Plant Cell 13: 1035-1046. Pasternak D (1987) Salt tolerance and crop production- a comprehensive approach. Ann Rev Phytopathol 25: 271-291. Pellegrineschi A, Reynolds M, Pacheco M, Brito RM, Almeraya R, Yamaguchi- Shinozaki K, Hoisington D (2004) Stress-induced expression in wheat of the Arabidopsis thaliana DREBIA gene delays water stress symptoms under greenhouse conditions. Genome 47: 493—500. Penna S (2003) Building stress tolerance through over-producing trehalose in transgenic plants. Trends Plant Sci 8: 355-358. Piao HL, Lirn JH, Kim SJ, Cheong G-W, Hwang I (2001) Constitutive over-expression of AtGSK] induces NaCl stress responses in the absence of NaCl stress and results in enhanced NaCl tolerance in Arabidopsis. Plant J 27: 305-314. Pilon-Smits EAH, Ebskamp MJM, Paul MJ, Jeuken MJW, Weisbeek PJ, Smeekens SCM (1995) Improved performance of transgenic fructan-accumulating tobacco under drought stress. Plant Physiol 107: 631-638. Pilon-Smits EAH, Terry N, Sears T, Kim H, Zayed A, Hwang SB, Van-Dun K, Voogd E, Verwoerd TC, Krutwagen RWH, Goddijn OJ M (1998) Trehalose-producing 48 transgenic tobacco plants show improved growth performance under drought stress. J Plant Physiol 152: 525-532. Pilon-Smits EAH, Terrya N, Sears T, Van-Dunb K (1999) Enhanced drought resistance in fi'uctan-producing sugar beet. Plant Physiol Biochern 37: 313-317. Prabhavathi V, Yadav J S, Kumar PA, Rajarn MV (2002) Abiotic stress tolerance in transgenic eggplant (Solanum melongena L.) by introduction of bacterial mannitol phosphor-dehydrogenase gene. M01 Breeding 9: 137—147. Prasad KVSK, Sharmila P, Kumar PA, Pardha-Saradhi P (2000) Transformation of Brassicajuncea (L.) Czem with bacterial codA gene enhances its tolerance to salt stress. Mol Breeding 6: 489-499. Pritchard DJ, Hollington PA, Davies WP, Gorham J, DeLeon J LD, Mujeeb-Kazi, A (2002). K+/Na+ discrimination in synthetic hexaploid wheat lines: Transfer of the trait for Kli/Na+ discrimination from Aegilops tauschii into a T riticum turgidum background. Cereal Res Commun 30: 261- 267. Qiu QS, Guo Y, Quintero FJ, Pardo JM, Schurnaker KS, Zhu JK (2004) Regulation of vacuolar Na+/H+ exchange in Arabidopsis thaliana by the salt-overly-sensitive (SOS) pathway. J Biol Chem 279: 207-215. Quan R, Shang M, Zhang H, Zhao Y, Zhang J (2004) Engineering of enhanced glycine betaine synthesis improves drought tolerance in maize. Plant Biotech J 2: 477-486. Quarrie SA (1989) Abscisic acid as a factor in modeling drought resistance. In Cherry J H (ed.) Environmental Stress in Plants. NATO ASI series vol G19. Springer-Verlag Berlin Heidelberg 27-37. Quesada V, Garcia-Martinez S, Piqueras P, Ponce MR, Micol J L (2002) Genetic architecture of NaCl tolerance in Arabidopsis. Plant Physiol 130: 951-963. Rental MC, Knight MR (2004) Oxidative stress-induced calcium signaling in Arabidopsis. Plant Physiol 135: 1471-1479. Ribaut JM, Jiang C, Gonzalez-de-Leon D, Edmeades GO, Hoisington DA (1997). Identification of quantitative trait loci under drought conditions in tropical maize.2 Yield components and marker-assisted selection strategies. Theor Appl Gen 94: 887- 896. Rohila J S, Jain RK, Wua R (2002) Genetic improvement of Basmati rice for salt and drought tolerance by regulated expression of a barley Hva] cDNA. Plant Sci 163: 525-532. Ronde J A, Spreeth MH, Cress WA (2000). Effect of antisense L- A-l-pyrroline-S- carboxylate reductase in transgenic soybean plants subjected to osmotic and drought stress. Plant Growth Reg 32: 13-26. 49 Roy M, Wu R (2001) Arginine decarboxylase transgene expression and analysis of environmental stress tolerance in transgenic rice. Plant Sci 160: 869—875. Roy M, Wu R (2002) Overexpression of S-adenosylmethionine decarboxylase gene in rice increases polyamine level and enhances sodium chloride-stress tolerance. Plant Sci 163: 987- 992. Rus AM, Estan, MT, Gisbert C, Garcia-Sogo B, Serrano R, Caro M, Moreno V, Bolarin MC (2001) Expressing the yeast HAL] gene in tomato increases fruit yield and enhances K+fNa+ selectivity under salt stress. Plant Cell Environ 24: 875-880. Safarnejad A, Collin HA, Bruce KD, McNeilly T (1996). Characterization of alfalfa (Medicago sativa L) following in vitro selection for salt tolerance. Euphytica 92: 55- 61. Saijo Y, Hata S, Kyozuka J, Shimamoto K, Izui K (2000) Overexpression of a Ca2+- dependent protein kinase confers both cold and salt/drought stress tolerance in rice plants. Plant J 23: 319-327. Saijo Y, Kinoshita N, Ishiyama K, Hata S, Kyozuka J, Hayakawa K, Nakarnura T, Shimamoto K, Yarnaya T, Izui K (2001) A Cay-dependent protein kinase that endows rice plants with cold-and-salt-stress tolerance firnctions in vascular bundeles. Plant Cell Physiol 42: 1228—1233. Sakamoto A, Alia I, Murata N (1998). Metabolic engineering of rice leading to biosynthesis of glycinebetaine and tolerance to salt and cold. Plant Mol Biol 38: 1 01 1—1 01 9. Sakamoto A, Murata N (2000) Genetic engineering of glycinebetaine synthesis in plants: current status and implications for enhancement of stress tolerance. J Exp Bot 51: 81-88. Sakamoto H, Maruyama K, Sakuma Y, Meshi T, Iwabuchi M, Shinozaki K, Yamaguchi- Shinozaki K (2004) Arabidopsis CyleHisZ-Type Zinc-Finger Proteins function as transcription repressors under drought, cold, and high-salinity stress conditions. Plant Physiol 136: 2734—2746. Saneoka H, Nagasaka C, Hahn DT, Yang WJ Prernachandra GS, Joly RJ, Rhodes D (1995) Salt tolerance of glycinebetaine-deficient and containing maize lines. Plant Physiol 107: 631-638. Seki M, Kamei A, Yamaguchi-Shinozaki K, Shinozaki K (2003) Molecular responses to drought, salinity and different paths for plant protection. Curr Opin Biochem 14: 194-199. Seki M, Narusaka M, Abe H, Kasuga M, Yamaguchi-Shinozaki K, Caminci P, Hayashizaki Y, Shinozaki K (2001) Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a firll-length cDNA microarray. Plant Cell 13:61-72. 50 Seki M, Narusaka M, Ishida J, Nanjo T, Fujita M, Oono Y, Kamiya A, Nakajima M, Enju A, Sakurai T, Satou M, Akiyama K, Taji T, Yamaguchi-Shinozaki K, Caminci P, Kawai J, Hayashizaki Y, Shinozaki K (2002) Monitoring the expression profile of 7000 Arabidopsis genes under drought, cold and high-salinity using full length cDNA microarray. Plant J 31: 279—292. Senadhira D, Zapata-Arias FJ, Gregorio GB, Alejar MS, de la Cruz HC, Padolina TF, Galvez AM (2002) Development of the first salt-tolerant rice cultivar through indica/indica anther culture. Field Crop Res 76: 103-110. Serce S, Navazio JP, Gokce AF, Staub J E (1999) Nearly isogenic cucumber genotypes differing in leaf size and plant habit exhibit differential response to water stress. J Amer Soc Hort Sci 124: 358-365. Shannon MC, Noble CL (1990) Genetic approach for developing economic salt-tolerant crops. In Tanji KK ed. Agricultural salinity assessment and management. New York, Amer Soc Civ Engin 161-185. Shen Y-G, Du B-X, Zhang WK, Zhang J S, Chen SY (2002) AhCMO, regulated by stresses in Atriplex hortensis, can improve drought tolerance in transgenic tobacco. Theor Appl Genet 105:815—821. Shi H, Lee B-H, Wu SJ, Zhu J-K (2003). Overexpression of a plasma membrane Na+/H+ antiporter gene improves salt tolerance in Arabidopsis thaliana. Nature Biotech. 21: 81-85 . Shi HZ, Ishitani M, Kim C, Zhu J -K (2000) The Arabidopsis thaliana salt tolerance gene SOSl gene encodes a putative Na+/H+ antiporter. Proc Natl Acad Sci USA 97: 6896-6901. Shi HZ, Lee BH, Wu SJ, Zhu J -K (2003) Overexpression of a plasma membrane Na+/H+ antiporter gene improves salt tolerance in Arabidopsis thaliana. Nature Biotech 21: 81-85. Shinozaki K, Yamaguchi-Shinozaki K (2000). Molecular responses to dehydration and low temperature: differences and cross-talk between two stress signaling pathways. Curr Opin Plant Biol 3: 217-223. Shinozaki K, Yarnaguchi-Shinozaki K, Seki M (2003) Regulatory network of gene expression in the drought and cold stress responses. Curr Opin Plant Biol 6: 410-417. Sivamani E, Bahieldinb A, Wraith JM, Al-Niemi T, Dyer WE, Ho THD, Qu R (2000) Improved biomass productivity and water use efficiency under water deficit conditions in transgenic wheat constitutively expressing the barley HVAl gene. Plant Sci 155: 1—9. Srivastava S, Fristensky B, Kav NNV (2004) Constitutive expression of a PR10 Protein enhances the germination of Brassica napus under saline conditions. Plant Cell Physiol 45: 1320-1324. 51 St 311 Steponkus PL, Uemura M, Joseph RA, Gilmour SJ, Thomashow MF (1998) Mode of action of the COR15a gene on the freezing tolerance of Arabidopsis thaliana. Proc Natl Acad Sci USA 95: 14570-14575 Stockinger EJ, Gillmour SJ, Thomashow, MF (1997) Arabidopsis thaliana CBF1 encodes an AP2 domain-containing transcriptional activator that binds to the C- repeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in response to low temperature and water deficit. Proc Natl Acad Sci USA 94: 1035- 1040. Sugino M, Hibino T, Tanaka Y, Nii N, Takabe T, Takabe T (1999) Overexpression of DnaK from a halotolerant cyanobacteriurn Aphanothece halophytica acquires resistance to salt stress in transgenic tobacco plants. Plant Sci 137: 81—88. Swire-Clark GA, Marcotte WRJ (1999) The wheat LEA protein EM function as an osmoprotective molecule in Saccharomyces cerevisiae. Plant Mol Biol 39: 117-128. Sze H, Li X, Palmgren MG (1999) Energization of plant cell membranes by H+-pumping ATPases: regulation and biosynthesis. Plant Cell 11: 677-689. Taiz L, Zeiger E (1998) Growth and development: Stress physiology. In: Plant Physiology, 2“‘1 Edition Sinauer Associates, Inc., publishers. 725-757. Takahashi S, Katagiri T, Hirayarna T, Yamaguchi-Shinozaki K, Shinozaki K (2001) Hyperosmotic stress induces a rapid transient increase in inositol 1,4,5 trisphosphate independent of abscisic acid in Arabidopsis cell culture. Plant Cell Physiol 42: 214- 222. Tanaka Y, Hibino T, Hayashi Y, Tanaka A, Kishitani S, Takabe T, Yokota S, Takabe T (1999) Salt tolerance of transgenic rice overexpressing yeast mitochondrial-Mn-SOD in chloroplasts. Plant Sci 148: 131—138. Tanford C (1978) The hydrophobic effect and the organization of living matter. Science 200: 1012-1018. Tarczynski MC, Jensen RG, Bohnert HJ (1993) Stress protection of transgenic tobacco by production of the osmolyte mannitol. Science 259: 508-510. Tazuki A (1997) Growth of cucumber fruit as affected by the addition of NaCl to nutrient solution. J Jap Soc Hort Sci 66: 519-526. Tester M, Davenport R (2003) Na + tolerance and Na+ transport in higher plants. Ann Bot 91: 503-527. Thomashow MF (1999) Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu Rev Plant Physiol Plant Mol Biol 50: 571-599. Thomashow MF (2001) So what's new in the field of plant cold acclimation? Lots. Plant Physiol. 125: 89-93. 52 Thomashow MF, Gilmour SJ, Stockinger EJ, J aglo-Ottosen KR, Zarka DG (2001) Role of the Arabidopsis CBF transcriptional activators in cold acclimation. Physiol Plant 112:171-175. Tilahun A, Guenzi AC, Martin B, Cushman J C (2003) Tolerance of mannitol- accumulating transgenic wheat to water stress and salinity. Plant Physiol 13121748- 1 755. Vogel G, Fiehn O, Bressel JLR, Boller T, Wiemken A, Aeschbacher RA, Wingler A (2001) Trehalose metabolism in Arabidopsis: occurrence of trehalose and molecular cloning and characterization of trehalose-6-phosphate synthase homologues. J Exp Bot 52: 1817-1826. Wang W, Vinocur B, Altman A (2003) Plant responses to drought, salinity and extreme temperatures: towards genetic engineering for stress tolerance. Planta 218: 1-14. Winicov I (2000) Alfinl transcription factor overexpression enhances plant root growth under normal and saline conditions and improves salt tolerance in alfalfa. Planta 210: 416-422. Winicov I, Bastola DR (1999) Transgenic overexpression of the transcription factor Alfin] enhances expression of the endogenous MsPRPZ gene in alfalfa and improves salinity tolerance of the plants. Plant Physiol 120: 473—480. Xiong L, Zhu J -K (2002) Molecular and genetic aspects of plant responses to osmotic stress Plant Cell Environ 25: 131-139. Xu D, Duan X Wang B, Hong B, Ho TD, Wu R (1996) Expression of a late embryogenesis abundant protein gene, HVAl , from barley confers tolerance to water deficit and salt stress in rice. Plant Physiol 110: 249-257. Xue Z-Y, Zhi D-Y, Xueb G-B, Zhang H, Zhao Z-Y, Xia G-M (2004) Enhanced salt tolerance of transgenic wheat (T ritivum aestivum L.) expressing a vacuolar Na+/H+ antiporter gene with improved grain yields in saline soils in the field and a reduced level ofleafNa+. Plant Sci 167: 849—859. Yamaguchi-Shinozaki K, Shinozaki K (1994) A novel cis-acting element in an Arabidopsis gene is involved in responsiveness to drought, Low-ternperature, or high-salt stress. Plant Cell 6: 251-264. Zarka, DG, Vogel JT, Cook D, Thomashow MF (2003) Cold induction of Arabidopsis CBF genes involves multiple ICE (Inducer of CBF Expression) promoter elements and cold-regulatory circuit that is desensitized by low temperature. Plant Physiol 133: 910-918. Zeevaart, JAD, Creelrnan RA (1988) Metabolism and physiology of abscisic acid. Annu Rev Plant Physiol Plant Mol Biol 39: 439-473. Zhang HX, Blumwald E (2001) Transgenic salt-tolerance tomato plants accumulate salt in foliage but not in fruit. Nature Biotech 19: 765-768. 53 Zhang HX, Hodson JN, Williams JP, Blumwald E (2001) Engineering salt-tolerant Brassica plants: characterization of yield and seed oil quality in transgenic plants with increased vacuolar sodium accumulation. Proc Natl Acad Sci USA 98: 12832- 12836. Zhang X, Fowler SG, Cheng H, Lou Y, Rhee SY, Stockinger EJ, Thomashow MF (2004) Freezing-sensitive tomato has a functional CBF cold response pathway, but a CBF regulon that differs from that of freezing-tolerant Arabidopsis. Plant J 39: 1-15. Zhifang G, Loescher WH (2003) Expression of celery mannose-6-phosphate reductase in Arabidopsis thaliana enhances salt tolrance and induces biosynthesis of both mannitol and glucosyl-mannitol dimmer. Plant Cell Environ 26: 275-283. Zhu BC, Su J, Chan MC, Verrna DPS, Fan YL, Wu R (1998) Overexpression of a A1- pyrroline—S-carboxylate synthetase gene and analysis of tolerance to water- and salt- stress in transgenic rice. Plant Sci 139: 41—48. Zhu J -K (2000) Genetic analysis of plant salt tolerance using Arabidopsis. Plant Physiol 124: 941-948. Zhu J -K (2002) Salt and drought stress signal transduction in plants. Ann Rev Plant Biol 53: 247-273. Zhu J -K, Hasegawa PM, Bressan RA (1997). Molecular aspects of osmotic stress in plants. Crit Rev Plant Sci 16: 253-277. 54 Chapter II Expression of the Arabidopsis thaliana transcriptional regulators, CBF] and CBF3, confers dehydration stress tolerance in cucumber (Cucumis ! . sativus L.) plants. Introduction Environmental factors that impose dehydration stress affect species distribution and plant productivity (Bray, 1994; Bohnert and Jensen, 1996). As the driving force of living organisms, water serves as a medium for the biochemical activities in living cells, is involved in biosynthesis and assembly of molecules into organized structures, and is responsible for generating the needed turgor pressure for plant cell expansion (Tanford, 1978; Xiong and Zhu, 2002). When plants fail to acquire sufficient water, the resultant water deficit causes loss of turgor and/or osmotic stress. Dehydration stress can be caused by several abiotic stresses including drought, freezing temperatures and salinity. Soil salinity is estimated to affect about 77 million ha (Jain and Selvaraj, 1997; Tester and Davenport, 2003) and may reach up to 50% of the total arable land by the year 2050 (Wang et al., 2003), making it a priority to find ways to alleviate this problem. Similarly, due to climate changes worldwide, a long-term trend of higher temperatures with a decrease in rainfall is expected to negatively impact agricultural production, especially in arid and semiarid regions (Hillel and Rosenzweig, 2002). These trends, along with the need to increase food production for a continuously growing world population will increase drought- and salinity-affected areas. 55 In general, plant responses to changing environmental conditions are mediated by alterations in gene expression (Guy et al., 1985; Greenway and Munns, 1980; Bohnert and Jensen, 1996; Ingram and Bartels, 1996; Zhu et al., 1997). The first group of genes whose expression is altered in response to dehydration-inducing conditions include those encoding transcriptional factors (TF), mitogen activated protein kinases (MAPKs), dephosphorylation enzymes, and chromatin remodeling proteins (Thomashow, 1999; Knight and Knight, 2001; Shinozaki et al., 2003). Regulation of the first group of genes generates signals that lead to adaptive responses including the induction of genes involved in biosynthesis of osmolytes and compatible solutes, late embryogenesis abundant (LEA) proteins, transporters, and detoxification enzymes (Shinozaki and Yamaguchi-Shinozaki, 1997; Seki et al., 2001; Fowler and Thomashow, 2002; Seki et al., 2002). Thus, induction of response-cascades has been suggested as an effective approach to enhance dehydration stress tolerance (Thomashow, 1999). Study of processes that lead to freezing and drought tolerance in Arabidopsis thaliana revealed critical information about plant response to dehydration stress conditions. Exposure of Arabidopsis plants to low, non-fieezing temperatures or to drought stress conditions leads to increased expression of a group of genes, the COld Responsive (COR) and Responsive to Drought (RD) genes that are induced under both cold acclimation and drought conditions (Gihnour et al., 1992; Hovarth et al., 1993; Lin and Thomashow, 1992; Thomashow, 1999). The COR and RD genes are characterized by the presence of a cis-acting element (CCGAC) within their promoter regions known as the CRT/DRE element [CRT (C-repeat)/DRE (Drought Response Element)], which confers responsiveness to low temperature, dehydration and salinity (Baker et al., 1994; 56 Yarnaguchi-Shinozaki and Shinozaki, 1994; Stockinger et al., 1997). The CRT/DRE element is bound by the CBF/DREB (CRT Binding Factor/DRE binding) transcription factors which are characterized by a putative nuclear localization domain, an activation domain and an AP2 DNA binding domain (Stockinger et al., 1997; Liu et al., 1998). Transgenic Arabidopsis plants overexpressing CBF] showed elevated expression of target COR transcripts and were more tolerant to freezing stress than their non-transgenic counterparts as determined by electrolyte leakage and whole plant assays (J aglo-Ottosen et al., 1998). Similarly, transgenic Arabidopsis plants overexpressing DREB] a had enhanced drought and salinity resistance (Kasuga et al., 1999). Transgenic plants also exhibited several physiological changes associated with increased dehydration stress resistance including an increase in proline and total soluble sugars accumulation compared to their nontransgenic counterparts (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gilmour et al., 2000; Jaglo et al., 2001; Seki et al., 2001; Seki et al., 2002; Haake et al. 2002; Kasuga et a]. 2004). More recently, elevated expression of CBF was shown to cause large-scale-metabolome changes in Arabidopsis including marked increase in soluble sugars and amino acids leading to the suggestion that increased accumulation of some soluble carbohydrates (galactinol, glucose, raffinose and fructose) and proline is a signature of the CBF regulon (Cook et al., 2004). Induction of CBF genes also confers increased dehydration stress resistance to other less closely related species. Tobacco plants expressing CBF 3/DREB1A had higher photosynthetic activity under drought conditions compared to wild type plants (Kasuga et al., 2004); tomato plants overexpressing the Arabidopsis-CBF I/DREB] B gene showed elevated tolerance to drought (Hseih et al., 2002a, b) and salinity (Lee et al., 2003); and 57 strawberry plants expressing the Arabidopsis CBF1/DREBIB gene had less membrane damage in response to freezing temperatures than their nontransgenic counterparts (Owens et al., 2002). In the present work, we sought to investigate and test the possibility to improve dehydration—stress tolerance in cucumber (Cucumis sativus L.) by expressing the A. thaliana transcriptional activators CBF] and CBF 3. Cucumber is known to be sensitive to salinity and drought especially in arid and semi-arid areas (Mass and Hoffman, 1977). 58 Materials and methods Plant constructs and transformation The Agrobacterium plant transformation constructs containing CBF] or CBF 3, under control of the CaMV 35S promoter were kindly provided by M. F. Thomashow (J aglo- Ottosen et al., 1998). Cucumber transformation was performed using the following procedure derived from the methods of Tabei et al. (1998). Cucumber seeds, cv. Straight 8 (Hollar Seeds, Rocky Ford, Co.) were decoated and surface sterilized for 15 min with 15 % Clorox solution (1% sodium hypochlorite solution) with 1-2 drops of Tween 20 (Sigma-Aldrich, St. Louis, MO). Seeds were rinsed several times with sterilized distilled water and placed overnight in the dark on honnone-free MS basal salt mixture (Murashige and Skoog, 1962) supplemented with 30g/l sucrose and 2.5 g/l Phytagel (Sigma-Aldrich, St. Louis, MO). Cotyledon explants were prepared by separating from the embryo, removing the outer edges, and cutting into 4-6 explants. Explants were co- cultivated with Agrobacterium tumefaciens strain C58 (Deblaere et al., 1985) , by dipping in a 1/20 diluted overnight culture for 10 min. Explants were blotted on sterilized filter paper then cultured onto C1 shoot induction medium [MS basal salt mixture medium (Sigma-Aldrich, St. Louis, MO) supplemented with 2 mg/l benzyl amino purine (BAP) (Sigma-Aldrich, St. Louis, MO), 1 mg/l abscisic acid (ABA) (Sigma-Aldrich, St. Louis, MO), 30 g/l sucrose and 2.5 g/l Phytagel]. Plates were wrapped in aluminum foil and incubated in the dark on culture shelves at 25 C°. Three days later, explants were rinsed several times with sterilized H20, blotted on filter paper and cultured onto C2 medium [Cl medium supplemented with 400 mg/l Timentin® (GlaxoSmithKline, Research Triangle Park, NC)]. A week later, explants were transferred on to a selection 59 medium C3 [C2 medium supplemented with 75 mg/l kanamycin (Sigma-Aldrich, St. Louis, MO)] and were monitored for 3-4 weeks for shoot growth and elongation. Emerging shoots were then placed onto selection medium C4 [C3 medium supplemented with 8 g/l agar (Sigma-Aldrich, St. Louis, MO)]. Shoots were continuously grown on media supplemented with 75 mg/l kanamycin (Sigma-Aldrich, St. Louis, M0) for selection. Shoots were rooted in 250 ml Magenta boxes (Magenta Corp, Chicago, IL) containing 50 ml of MS medium with 30 g/l sucrose and 7 g/l agar. PCR-verified transgenic plants were transferred into 10 cm pots filled with Baccto soil mix (Michigan Peat Co., Houston TX) and kept in the growth chamber for 2-3 weeks (16/8 h light, temperature was maintained at 22 C°, Relative humidity was maintained around 50%). Acclimated shoots were transferred to the greenhouse for 2-3 weeks before transferring into 3.6 1 pots filled with Baccto soil mix for seed production. Transgenic CBF1 or CBF3-cucumber plants were self pollinated to produce fruits in the greenhouse; 6-8 weeks old fruits were harvested and seeds were extracted, dried and stored in a dry area. DNA and RNA isolation, PCR, and Northern blot analysis DNA was extracted fiom 200 mg young leaf samples using the Wizard DNA purification kit (Promega, Cat # A7951, Madison, WI). For RNA isolation, young leaf tissues were collected and immediately frozen and ground in liquid nitrogen and extracted using Concert Plant RNA Reagent (Invitrogen, Carlsbad, CA). RNA quantity and quality was determined by measuring absorbance at 260 nm and by gel electrophoresis (Sarnbrook and Russell, 2001). Quantitation of RNA was performed with 60 a Molecular Imager FX Pro multi-imager system (Bio-Rad® Laboratories, Hercules, CA). PCR was carried out using CBF]- and CBF3-specific primers (Stockinger et al., 1997). Northern hybridization analysis was conducted using 32P labeled CBF] and CBF 3 cDNA firll length probes following the procedure of Sarnbrook and Russell (2001). Salinig experiments T1 segregating progeny of transgenic To plants were planted in the greenhouse and screened by PCR for the presence of the introduced CBF genes at the 2-3 -leaf stage. Transgenic seedlings, non-transgenic T1 segregant progeny, (Azygous) and wild type ‘Straight 8’ (WT) plants were transplanted into 15cm clay pots filled with vermiculite (Therm-o-Rock East Inc, Grade no. 3A, Washington, PA). To minimize evaporation, the soil surface was covered with plastic disks that fitted at the top of the clay pots. In the first greenhouse experiment, two families (A4 and B3) were tested with three levels of NaCl (0 mM, 100 mM and 200 mM) in a randomized complete block design with three replications. Salt treatment was initiated seven days after seedling transplant with a stepwise increase of 50mM NaCl at two day intervals to reach 200mM NaCl. Plants were fertilized with 300 ppm 20:20:20 fertilizer once a week throughout the experiment. Young leaves (3-4 cm diameter) were collected every four days for sugar and proline analysis as described below. In the second and third greenhouse experiments, 10 and 9 CBF-families were tested (Table, 1), at three NaCl levels (0, 50 and 100 mM), in a randomized complete block design with six replications. Salt treatment was initiated 15 days after transplanting, with a stepwise increase of 50 mM NaCl at three day intervals to reach 100mM NaCl. Total 61 soluble sugars and proline were measured at time 0 (immediately before the salt treatment) and 15 days post initiation of salt treatment (dps). Growth parameters (total above ground fresh and dry weight, plant height and number of leaves) were measured at 15 dps. Statistical analysis was performed using Microsoft Windows® EXCEL and SAS® 9.1.2 programs. Drought experiments T2 segregating progeny of transgenic families were planted in the greenhouse and screened by PCR for the presence of the CBF gene at the 2-3-leaf stage. Transgenic seedlings, non-transgenic segregants, and wild type ‘Straight 8’ plants were transplanted into 3.6 1 pots filled with Baccto soil mix (Michigan Peat Co., Houston TX). A split plot design with water as the main effect was used with three replications in experiment 1 and four replications in experiments 2 and 3. Genotypes were assigned randomly within each main plot. Drought treatment was carried out by withholding water until the first sign of wilting (about nine days in experiments 1 and 2 and 12 days in experiment 3) followed by one day of irrigation to full soil saturation. In the first experiment, this cycle was repeated three times, in the second experiment this cycle was repeated twice and in the third experiment this cycle was performed one time. Plant heights were determined after each cycle of drought; total above-ground flesh and dry weight was determined at the end of the experiment. Leaf tissue samples were collected immediately before the beginning of the treatment, as well as after each drought cycle for proline and sugar analysis. 62 Sugar and proline analyses One young leaf (3-4 cm diameter) was collected from each plant at each sample date, freeze-dried for 48 hr, ground, and split into two aliquots, one for proline and one for sugar analysis. Proline analysis was carried out using the procedure described by Troll and Lindsey, (1955) (Appendix A). Total soluble sugar analysis was performed by the phenol/sulfuric acid method of Dubois et al., (1956) (Appendix A). Chlorophyll fluorescence Components of chlorophyll fluorescence were quantified using a portable Plant Efficiency Analyzer fluorometer (Hansatech, Norfolk, UK). Measurements were performed in the greenhouse, using attached leaves. Three leaves, 7-9 cm in diameter, and well-exposed to sun light, were chosen per plant for sequential measurements. After 30-40 min dark adaptation period, minimal fluorescence (F0), maximal fluorescence (Fm), variable fluorescence (Fv) and fluorescence efficiency (Fv/Fm) were measured immediately before the beginning of the drought treatment, and after 2, 6 and 12 days of drought. 63 Results Introduction and expression of the Arabidopsis CBF] and CBF3 transgenes in cucumber plants Thirteen transgenic cucumber plants were produced, six with CBF] and seven with CBF 3 . Presence and expression of CBF genes was verified in the To plants by PCR and northern blot analysis, respectively (data not shown). Successful transfer of the CBF] and CBF 3 genes into T1 and T2 progeny was verified using PCR analysis; x2 analysis of segregating progeny was consistent with single gene insertion in each case (Table 2-1). Analysis of CBF expression in transgenic T2 plants showed varying transcription levels among the different lines (Figure 2-1). Transgenic cucumber plants expressing CBF genes have elevated levels of proline and total soluble sugars in leaves compared to the non-transgenic controls Greenhouse experiments were performed to evaluate the response of CBF] and CBF3-transgenic cucumber plants to salt stress. Seeds were only available for two lines, A3 and B4. In the absence of salt stress, CBF-expressing cucmnber plants accumulated significantly higher levels of both fi'ee proline (up to 5 fold higher) and total soluble sugars (2 fold higher), compared to the nontransgenic controls (Figure 2-2A, D; ANOVA, P< 0.01 and P<0.05 for proline and soluble sugars, respectively). Under salt stress conditions, CBF-cucumber plants accumulated higher levels of proline and soluble sugars compared to the non-stressed conditions (ANOVA, P<0.01). The azygous (non- 64 Table 2-1. Segregation analysis of T1 transgenic cucumber lines expressing the Arabidopsis CBF] and CBF 3 genes. Presence or absence of the gene was determined by PCR analysis Line Gene construct T1 segregation 12 (3:1) (trans: non) - A1 35S.‘.'CBFI 46 I 11 0.71 ns A3 35S.'.'CBFI 49 Z 13 0.30 ns A4 35S.’.'CBFI 51 2 16 0.09 ns A5 35S.'.'CBFI 54 I 14 0.15 as A6 35S.‘.°CBFI 55 I 19 0.01 ns Bl 35S.’.‘CBF3 44 I 12 0.21 ns B4 35S.’.'CBF3 46 2 10 1.14 ns BS 35S.‘.'CBF3 49 I 10 1.63 ns B6 35S.’.'CBF3 47 I 9 0.69 ns B7 35S.'.'CBF3 40 I 12 0.02 ns 65 ._o.=:oo 9..qu a ma SIB 5.3 cos-Etta»: gonna Eozom .onoE Emu 5.3 cozuufitgc ".28.. :3. 42m 9520». 283 :0 v as .8: n .2 twang—cum 200.. an? 35:00 03:3: gains .< 2: 6.5.552: 05 2:0 “.0303 >302? mm; =_mon-¢om .58 *0 some Eat vogue—.00 we; use. 950» 2.0 62.: Lon—5.8.8 3:325... N... E mecca ammo use Emu 9.25.32 Eco—«2.63m: 2:235 .< *0 :o_mmoaxw TN 239... “:00 um mm mm vm mm _.m n< w< m< v< m< ..< amu< .3833 m2... .onanaoékmo nos: LonEzoaoélmo 256:3 66 422.53.: :3 .0 20:33:. amen «>3 NF new a .v “.2023". 225 wanna: :3.— .mm H “5.53.2032323.53 n 5.? 3.3.32 n no :38 2: Eu Ema .mcoEucoo “.83.: gun can 33:3 60:: ANmo Nw m V m>mD N— a v rli--l _ ilrl. e -l- 1-11-314 O 1 _Tl.|..l.ll_l!l.l..lfll.oulnllll_lllII.- iflo ll \ln _ mm W . wiiuununuun until... u 6 . \\ _. 1 To W / ll won .m ,- mh M ‘8 Rnuu. Te _ M 3 La .82 .25 8w u. .82 2:. 8a 0 m>mD 2. till a v .o l lllilollll“VIIIIHHMLHUOIII'IHIIIHHH...” 11 M e ..... L _ .8 m a m 6 ,, e, r. 2: m. T “.8 mMP -... _ w .82 a _ «z _ :E 2: m 2:. 2: m . a 980 Q. w v 930 N— w v , l i -l .- i l - i r- l,-.l.4o l- l.-rlilihmfliiilwwiruununmullrl .;o W\ l o... M 2.... Y1 *3 m r mm AW 2-... _ Rnw. M . :5 M -82 FWU _ . mm? _ 8 flaw-um .052 SE 0 D cam—Guam _Uw2 SE a < 67 transgenic segregant) and wild-type plants had equivalent levels of proline and soluble sugars accumulation, indicating the observed differences from CBF-plants were due to the presence of CBF genes. Exposure of cucumber plants to 200 mM NaCl did not cause an increase in either proline or sugar concentration (Figure 2-2C, F), compared to the 100 mM NaCl level (Figure 2-2 B, E). In addition, plants irrigated with 200 mM NaCl exhibited severe leaf discoloration and necrosis; therefore, 100 mM NaCl was the maximum concentration used in subsequent experiments. Additional CBF families (10 and 9 families) were tested in the second and third experiments, equivalent results were obtained in both experiments. Significant differences were observed between transgenic and nontransgenic families for proline and soluble sugar accumulation prior to the initiation of salt treatment; on average, CBF- expressing cucumber plants had 24.5 a 3.1 mg/gdw soluble sugar and 8.9 a 0.4 ug/gdw proline, compared to an average of 15.8 a 1.2 mg/gdw soluble sugars and 2.3 :t 0.3 pg/gdw proline in the nontransgenic controls (ANOVA, P<0.01 and P<0.05 for proline and soluble sugar, respectively). Fifteen days post initiation of salinity treatment, proline and total soluble sugar accumulation was significantly greater in the transgenic families compared to the nontransgenic controls at all salt levels (Figure 2-3A and B; Table 2-2). There were no significant differences between the CBF] and CBF 3 lines (Table 2-2). The transgenic cucumber plants showed a progressive increase in proline and soluble sugar accumulation at both 50 and 100 mM NaCl levels, while proline and sugars levels in the nontransgenic controls did not increase above 50 mM NaCl (Figure 2-3). The 68 .m_0:E>m :0:0 .3 mac: 0.:0mmc0bc0: .m_0:E>m 030.0 >0 02020:. 9.0 mac: 2:03:00... .m.:0E...0n.x0 £0: :. 0050.00 0.63 3.3mm: .:o_0>.:0m .r .ceEcmcxm :. 02mm. 3:0 m0; :023 Km 0:088 m.:0E...0:x0 £0: :. 02mm. 0.03 3:30:00 =< 0:2... L : . : . ..t . . ... 025 5.? 0:0 302.0030: £802.00: m 0:0 m 5.3 35.5.2.6 05 E0... 00.00: 00 0.0a. .EmEfimc. .002 SE 8.1.0 on .0 6000.5 «m0: m>00 mr .Ns $500.53 30030 0:0 .0. .0...:00 00.50.30.000: .fim 0:0 .wm .mm in .3 mac: ammo 0:0 w< 0:0 .m< .v< .n< ...< mmc: Emu. mac: 2:03:05 :3 .0 £0.03 30 0:0 £0.03 :3... 0:305 06:0 €2.50 0::05 .093 0.03.3 .20... .n.~ 050.". SE 03 SE om 5:5 .288? ES om 2:5 lllll llAN0 0.02, mm: - «cm—500.: 3.5.0.0 «mo: m>00 mr 0:903 in. -0 >0. ._ a 0 my E . : u. o SE 2: .25 8 .25.. .1 ill. lhll .. llllrlllll. l o O lira V : willful . 1 m a 2 M: a. , ol ov m. w. 6 9 L a m m 5/ . ow M on W .W t .M.. c. 3 €05.00... 3.5.0.0 “mo: ,2 ONF( .5532. >2.028.. A.26.. 3 5:035:80 05.2.. .m £60 3 60035380 camam 030.8 .000... -< 69 30000:: 0:000:80 0.:0w0:0:..:0: 0:0 0 530:5 0:... 0:3 ..0 0003. -m .00. 0s. .00. .00. .00. .5. 000.. 0:00 02.. :0 53.2 -0 .00. 000 .2 .3. .0... .2. 000.. E000 02: .0 000.). -. 00.03.000.00: . .0... v0. 0:0 no... v0. 0:000:00 0:000:00 <>OZ< .3 8:: 0.000000: 0:. 80: 0:08.50 0.3000330 0:0 8:: 0800800002 2. .0. .mm “0 Gave 05 mm 0..—9» £00m— .0000800000 500 0. 00.00.00 0:03 0:000: 05.03000. .. .0258qu :. 00.8. :30 00B :03? 6:: 0m. 00.0000 3008.80.00 .000. 0. 500. 0. 00.8. 0:03 8050:0w ..< .m 0:08.898 :0.. 80508» .. 0:0 0:000:00: 08.00.30: 0%. 0:0 N 0:08.898 :0.. 8050:0w N. 0:0 8080005800230: «.0 5.3 3:08.898 03. 0:0: 00.00.. 0:0 0.0Q 0.0 .0 0.0.: 0... a 0...: 0.: 0 0.0: 0.0. 0 0..... 0.0 0. 0.0... 0.0 a 0.00. 0.0 0. 0.00. 0.0 a 0.00. .0. a 0.00.. 000.0: 3 0.0.00.0 ...«0.0. 0.. 0.0.0. 2.00.0. 001.0. 0.. «0.: 0.0.1.00 0.30.0 0.30.00 0:30; .05 0 0.0 a 0.00 0.0 0 0.00 0.0 a 0.0.. 0.0 0. 0.00. 0.0 a 0.00 0.0 0. 0.00. 0.0 .0 0.00. 0.0 0 0.... 0.0 .0 0.00. 0:90.: :8...— t. 0.0 a 0.0 0.0 a 0.00 0.. a 0.00 «0.0.0 a 0.0 0.0 a 0.0. 0.. 0. 0.0. $0.0 .0 0.0 0.0 .0 0.0 0.0 a 0.0 05.0.5 $0.0 « 0.. 0 0.0 a 0.: 0.0 “0 0.00 «0.0 0. 0.00 0.0 0. 0.0.0 0.0 0. 0.00 30.0 a 0.: 0.0 « 0.00 0.0 0. 0.00 0:00.30 0:00-007. FED Emu 0:20-007. 205 Emu 00:00:07. 0 0.03.0 . EMU .007. 2:. 0... .007. 2:. 00 .007. 2.: 0 0:08.000 .002 ..0 8:00.... .000. 00.00 m. 8:: .0008 0:0 0.:0w0000im0 :. .030: .00... 0:0 .2303 :0 0:0 03.03 :00: 05.0% 0>0..0 00:20. 00: .:0w..0 0.0.0.00 :0 0.0.001. .N..~ 0.00... 70 amount of accumulation of the two types of compounds within a given genotype was highly correlated (r2=0.89; Figure 2—4A), indicating that they are part of a coordinated response. Salinity treatment caused a marked increase in soluble sugar and proline accumulation in CBF-plants beyond what would be expected from an additive effect of the two separate components, CBF in the absence of stress, and salt stress in the absence of CBF (Figure 48, C arrows). Thus, the effect of the CBF gene on solute accumulation increased in response to salt stress. Given the increased accumulation of compatible solutes in 3SS:CBF plants in response to salinity stress, CBF transcript levels were compared in the presence or absence of salinity stress (Figure 2-5). The lack of differences is consistent with expected results for expression driven by the constitutive CaMV3SS promoter, and this also indicates that there were not differences in CBF mRNA stability associated with salt StI'CSS. CBF-expressing cucumber plants showed less reduction in fresh and gg weight under salt stress conditions Growth data were collected at the end of experiments 2 and 3. At 0 mM NaCl, there were no significant differences between CBF— transgenic families and the nontransgenic controls for height, and above ground flesh and dry weight (Figure 2—3 C, D; Table 2-2). At 50 mM NaCl, neither the CBF nor the nontransgenic lines showed significant differences in growth. At 100 mM NaCl, nontransgenic controls exhibited a significant reduCtion in fresh weight (48%), dry weight (56%), and height (16%) relative to non-salt stressed plants. While 100 mM NaCl also affected growth of the transgenic cucumber 71 dicta .3 boot—2: m. 3.5.3 .0 cocoons a... :. “.mo .0 5.32.523 2.... dons—.8303 .0. Euuzm 2.3.3 new 3. «5.0:. 8 “.mo .o 3:35 2.. 2. 3.5.3 .o 20.33350 «83.—cc. when 2...; .umo can 3.5.3 .0 20.33.5200 3:23. .0 new m 3...: contomou mm a new N 358.398 ES. 2a Sun .32.. 0.53:8. .5358 23:...» 33.0 6.03:8 0.53:8.-20: .5352 £3.53 :30 ace—53.. Ga: 2E car new cm .0 533:5 32. get 3 35:. BnEzoao :. 22:8 «5.2.. new .895 033.8 .3232. 22.22.60 -< .1. 959.... ammo Emu 2:59 25% .7 2:5 2:5: 25% 2 . _ =5 (Mpfi/fiw) SJean alqnlos mun—mo Emu 2:59 25% 2:5 2:59 28% 2.5 m... m .3 ”.8- . on .2639: 25:9“. ov on cm or o aouwm _ _ o d . m. om @o N] O I 0.? 5 .W ‘0 o my I 8 o o . cm a w . cop «55.00 05.0... new Ewan» 233:5 (Mpfilfiw) 9:2an elqnlos leiol < 72 .00 0E0» .0200 E330 0.... .000... Emu 5.; ...0..0n.u..n>.. 6.2.0.. 00.. 0..... .05.. :000 5 00.000. 002" =.000.~.on. 00.... .0 2000 50.. 00.02.00 00.5 “.00. .950» 0:0 .300... N .0.— .002 EE 2: .0 00:00.... .0 02.000... 05 E 9:305 00.... 0:02am 0.:0ua:0.....0.. 0:0 00.... 9.300.008 .0nE3030 N .. 58 :. 00:0u 930 0:0 Emu 0.0.2.50. .0..0=....00..0...0:0:05 .< .0 :0.000.0xm_ .m.N 0.33.“. 73 Jan Semi lieu Sea»: 8653.: 5.3 anaemia» 305.3 n85u¢n2fln Ears mangle—fin un—Ofla—g lines (a reduction of 30%, 16%, and 22% in fresh weight, dry weight, and height . respectively), growth inhibition was significantly less severe relative to the nontransgenic controls for fresh weight (ANOVA, P< 0.05) and dry wei ght (AN OVA, P< 0.01). Plants expressing CBF1 or CBF 3 genes had almost twice the dry weight as the nontransgenic controls. CBF] and CBF 3 lines performed equivalently (Table 2—2). Grovvth in the presence of 100 mM NaCl appears to be correlated with levels of soluble sugars and proline levels (r2 = 0.79 and 0.78, respectively). Transgenic CBF-cucumber plants have elevated level of compatible solutes and less reduction in photosvnthetic capacity in response to drouth stress CBF1 and CBF3-expressing cucumber plants were also tested under drought stress conditions. The same trends were observed in all three experiments. Under well~irri gated conditions, significant differences in proline levels between the CBF~expressing families and the nontransgenic controls were observed while significant differences in soluble sugar were observed later in the experiment (Figure 2-6A, C; AN OVA, P<0.01). After the first cycle of drought (until the first symptoms of wilting of lower leaves), CBF- expressing cucumber plants accumulated approximately twice the amount of soluble sugar and approximately 5-fold higher levels of proline than their nontransgenic counterparts (Figure 2-68, D; ANOVA, P<0.01). Although afier a second cycle of drought proline and soluble sugar levels remained higher in the CBF-plants than the non- transgenics, soluble sugar level in the transgenic plants did not increase compared to the first cycle and levels of proline declined (Figure 6D). Moreover, the wilted lower leaves did not recover after the second re-irrigation. No differences in proline or soluble sugar 74 ..000..00.. .0 00000.0. .000 0.00 mm 000 N. .0 00.02.00 0.0.5 00000.. .00.. ..00.0.00...0.00.0.2.0.00... n 0..... 00.02.00. 0 .0 mm H 0000. 0... 0.0 0.00 .000...0000 2000.0 000 00.00.0232. .0000 0.0..000 0.000000... .000 000 000.. 30.00000 00.000.008.200 000 Emu .0 .0303 05.0.0 00.. 000 .0000 0.00.00 .0.0... 0.0 0.00... 9.00 9.00 r o d L o M w annnnnnnuwu”HUME. m. w .0 n O O m w m .8 m .u. n n m ,0 m w 0300......0; .0 M 8. 9.00 mm m. o 00-? N<-m.| mmlbl .3 5+ m<1II w<+ guflfiET-i? u< 0.0.65 .0 Mpfilfiw 912an qun|os moi Mpfilfiw swans elqn|os [2301 8 75 levels were observed between the azygous and the wild-type plants. Chlorophyll fluorescence was used to determine the maximum photochemical efficiency in cucumber plants under drought stress conditions. Measurements were taken immediately before the beginning of drought treatment and after 2, 6 and 12 days of drought in experiment 3. There were no significant differences between the CBF-transgenic cucumber plants and the nontransgenic controls under well irrigated conditions. Significant differences were detected after 12 days of drought treatment (Figure 2-7), at which time stressed plants began to show the first obvious signs of wilting in the lower leaves for both the transgenic and the nontransgenic controls. The photosynthetic activity of the upper, non- wilted leaves of CBF-expressing plants had 50% higher fluorescence value than the nontransgenic controls, reflecting greater stability of photosystem II (PSII) under drought stress conditions (Figure 2-7). CBF-expressing cucumber plants showed less ggowth reduction under drought stress conditions Growth parameters (plant height, above ground flesh and dry weight) were collected at the termination of the drought experiments. Under well-irrigated conditions, no significant differences were observed between CBF-transgenic families and the nontransgenic controls for height, above ground flesh weight and dry weight (Figure 2- 8). In response to a cycle of drought, transgenic-CBF lines showed significantly less reduction in plant height and dry weight than did the nontransgenic controls (average reduction of 45% and 46% for height and dry weight vs. 25% and 15%, for CBF lines; ANOVA, P< 0.05). The CBF-transgenic lines did not show significantly less reduction in 76 Chlorophyll fluorescence under well-irrigated conditions 800 1 ‘E l i l «:9 700 .. l a E . o l 8 600-} 8 3 i cg» 5004. r. ‘ __“_ LL ~. !OA1 IAs A81 085 OAzl 400 .g__:--,.._T.-_--__. ..--..4::...-__...-_..- . . -1 0 2 4 6 8 10 12 Days post initiation of drought Chlorophyll fluorescence under drought-stressed conditions i ii 5‘ 5 29 :2: 0 9.3 8 0 O 3 u. 0 2 4 , 6 8 10 12 Days post initiation of drought Figure 2-7. Chlorophyll fluorescence (Fv/Fm) in CBF-expressing cucumber plants and azygous non-transgenic plants, under well-irrigated and drought- stressed conditions. Each data point represents the mean of four replicates with three plants/replicateltreatment1- SE. ' 77 250 w — Emmi-emu E, 200 l :“flwh‘smifl 3.. .g 150 1 J: 100 m E 50 -‘ o l Controls Dry weight (g) Controls Height (cm) Transgenic lines Controls Figure 2-8. Effect of drought conditions on above ground fresh weight (A), dry weight (B) weight, and plant height (C) of CBF1, and CBF3-transgenic lines, non-transformed controls (CC), and non-transgenic segregants (AZ). Measurements were recorded after one cycle of drought stress. Light bars: well-irrigated-plants; dark bars: drought stressed-plants. Values are the mean 1 SE of three replicates/treatment with 3 plants/replicate. 78 flesh weight (39%) than the nontransgenic controls (50%) (ANOVA, P< 0.12). Equivalent trends were observed after two cycles of droughts; with significantly less reduction in height (48% vs. 29%) and dry weight (45% vs. 24%) for the CBF lines compared to the nontransgenic controls. 79 Discussion Several lines of transgenic cucumber plants expressing CBF1/DREBb and CBF3/DREBa genes under the control of the constitutive CaMV3SS promoter were produced and tested for physiological changes and response to dehydration stresses. Previous studies with the CBF gene family demonstrated that overexpression of CBF in Arabidopsis caused significant increase in accumulation of compatible solutes, especially soluble sugars and proline (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gihnour et al., 2000; Seki et al., 2001; Seki et al., 2002; Haake et al. 2002; Cook et al., 2004). Under nonstressed conditions, cucumber plants expressing CBF1 and CBF 3 genes had higher soluble sugar and proline levels than the nontransgenic controls; these levels continued to increase throughout the experiment, indicating that the heterologous-CBF1 and CBF 3 genes induce pathways in cucumber similar to those in A. thaliana. The levels of increase in soluble sugar accumulation (2—3 fold) and proline (5-fold) were comparable to that observed in CBF-overexpressing Arabidopsis plants (Gilmour et al., 2000; Gilmour et al., 2004). Similar findings were also observed when overexpressing CBF/DREB genes in the heterologous species, tomato and tobacco (Hsieh et al., 2002a; Hsieh et al., 2002b; Lee et al., 2003; Kasuga et al., 2004). Accumulation of soluble sugar was highly correlated with proline accumulation, indicating that syntheses of these compounds are coordinately regulated in the CBF- ‘ expressing plants. In Arabidopsis, the promoters of genes for key enzymes involved in proline and sugar biosynthesis (e.g., P5 CS and galactinol synthase) have binding sites for CBF and have been shown to be upregulated in response to CBF expression (Seki et al., 2001; Seki et al., 2002; Fowler and Thomashow, 2002; Gilmour et al., 2000; Vogel et al., 80 2005). While orthologous genes in cucumber may also include CBF-binding sites, direct transcriptional regulation by CBF would not fully explain the observed increase in proline and sugars in CBF-cucumbers in response to salt stress. Exposure to salinity increased the levels of soluble sugar and proline in CBF-plants beyond a simple additive effect of the individual contributions of salinity and CBF. The increase in proline and sugar accumulation in response to salt stress was not a direct result of salt-induced increase in CBF-transcript levels, indicating that the enhanced accumulation, is at least in part downstream of CBF per se. It is possible that post- transcriptional or translational regulation of key enzymes that are critical for production and accumulation of sugars and proline is affected. Alternatively, the initial induction of dehydration stress-related responses by CBF may result in cascades of signals that indirectly, lead to changes in transcription rate of key proline and sugar biosynthetic genes. Transcriptional profiling in CBF-overexpressing Arabidopsis plants indicated activation of several classes of genes including genes encoding transcriptional regulators (e.g., AP2 containing proteins, zinc-finger containing proteins, MYB-family transcriptional activators), genes involved in stress-signaling (MAP-kinases, calcineurin and calcineurin-like proteins), and genes involved in metabolism and catabolism (Fowler and Thomashow, 2002, Sike et al., 2001; Maruyama et al., 2004; Vogel et al., 2005). The responding genes including transcriptional factors could be clustered into groups whose expression increased or decreased at different time periods following transfer to the cold, suggesting sequential induction (Fowler and Thomashow, 2002, Sike et al., 2001; Maruyama et al., 2004; Vogel et al., 2005). These observations indicate involvement of multiple regulatory systems which can be initially triggered by CBF. Indeed, not all of 81 the upregulated genes include the CDT/DRE element in their promoters, suggesting that these genes may be secondarily regulated by one of the CBF-induced transcription factors (Fowler and Thomashow, 2002; Vogel et al., 2005). The up—regulation of these genes could eventually affect other pathways in plants that could also induce production of compatible solutes. Thus, CBF-induced pathways may influence subsequent responses, perhaps priming the CBF-cucumbers to allow for enhanced response to stresses. There appears to be a limitation, however, to the increase in proline and sugar in response to stress, even in the CBF -p1ants. In our conditions, the transgenic cucumber plants did not accumulate higher levels of soluble sugars and proline when treated with 200 mM NaCl or when imposing a second drought cycle. This may indicate a limitation in the adaptive responses that can be induced by the CBF/DREB genes, or limitation of the ability of a species to respond to the unfavorable stress conditions. In non-transgenic plants, accumulation of soluble sugar and proline did not increase beyond levels obtained with the 50 mM salt treatment, or with imposition of drought, while levels of soluble sugars and proline continued to increase in the CBF-expressing cucumber plants (at 100 mM NaCl and after 2 cycles of drought), suggesting that the CBF increased the range of response in cucumber plants to a higher limit. These differences in range of response were also reflected in plant growth. In the presence of 50 mM NaCl, no significant differences in above ground flesh and dry weight and plant height were observed among CBF-transgenic and nontransgenic lines, indicating that at 50 mM NaCl, wild-type cucumber plants were able to adapt and adjust to this salinity level. In the presence of 100 mM NaCl, the cucumber plants expressing the Arabidopsis-CBF1 and CBF 3 genes showed significantly less reduction in above- 82 ground flesh and dry weight, compared to the nontransgenic controls. These results coupled with the drought stress results, indicate that CBF allowed for increased salt and drought stress resistance in cucumber. Moreover, at the time of first wilt of lower leaves, CBF—expressing cucumber plants also exhibited less reduction in F v/Fm (a measure of stability of photosystem II) in response to drought stress, than did the nontransgenic controls, indicating an additional physiological effect of the CBF transgene. Similar effects were also reported in tomato and tobacco plants expressing CBF1/DREBb (Hseih et al., 2002; Lee et al., 2003; Kasuga et al., 2004). Increased photosynthetic stability may result flom altered expression of CBF-target genes, or an indirect effect of increased osmoprotectant. The presence of higher levels of proline has been reported to correlate with higher protection of photosystem II (De-Ronde eta1., 2004). Similarly, greater stability of PS II was also reported with the production of the compatible solute glycine betaine (Hayashi et al., 1997; Sakamoto et al., 1998; Kishitany et al., 2000; Holrnstom et al., 2000; Quan et al., 2004), trehalose (Garg et al., 2002; J ang et al., 2003), and mannitol (Loescher et al., personal communication), or by overexpressing the yeast invertase gene in tobacco plants, which results in accumulation of glucose and fluctose up to 8-fold (Fukushima et aL,2001) Several studies with CBF-overexpressing Arabidopsis plants have reported negative impacts on plant growth (Lui et al., 1998; Kasuga et al., 1999; Gilmour et al., 2000). Severity of growth retardation was positively correlated with CBF expression levels (Liu et al., 1998; Gilmour et al., 2000), and was minimized the by use of a stress-inducible promoter (Lee et al., 2003, Kasuga et al., 2004; Pellegrineschi et al., 2004). Additional 83 phenotypic differences in plants constitutively overexpressing CBF 3 , included darker leaves, shorter petioles, and delayed bolting in Arabidopsis (Gilmour et al., 2000), and shorter intemodes and less fluit and seed production in CBF-overexpressing tomatoes (Hsieh et al., 2002). Expression of the CBF gene in cucumber did not have visible retard growth under the conditions tested. The lack of negative effects in this study could be due to differences in expression levels of CBF] and CBF 3 genes in 3SS:CBF-cucumber plants compared to the levels 35S:CBF Arabidopsis, or to differences in the nature of the downstream responses in cucumber in response to CBF expression. Despite enhanced stress resistance in tomato, efforts to clone COR homologs flom tomato were not successful, even under low stringency conditions, indicating the absence of COR genes as a CBF target in that species (Hsieh et al., 2002). Furthermore, microarray analysis of Arabidopsis and tomato plants expressing CBF genes, revealed the presence of marked differences in induced gene expression between these species, including the failure to observe predicted orthologs to Arabidopsis CBF-regulon genes. These results further suggests that responses may differ in heterologous systems, and may have differential impacts on growth and development. In conclusion, the present study demonstrates that expression of CBF in cucumber, a species known for sensitivity to salinity and drought conditions, may offer an effective approach to enhance salinity and drought tolerance. Our results shows that 35$:CBF— expressing cucumber plants were able to adapt to a higher range of dehydration-induced stresses than did their nontransgenic counterparts without apparent costs on plant growth. This increase in stress resistance was also accompanied by coordinated physiological responses including accumulation of compatible solutes and maintenance of photosystem 84 II stability. Further studies to evaluate plant performance as well as fluit production under field conditions are needed to begin to assess agricultural potentials. 85 References Baker SS, Wilhelm KS, Thomashow MP (1994) The 5’ region of the Arabidopsis thaliana cor15a has cis-acting element that confer cold-, drought-and ABA- regulated gene expression. Plant M01 Biol 24: 701-713. Bohnert HJ, Jensen RG (1996) Strategies for engineering water-stress tolerance in plants. Trends Biotch 14: 89-97. Bray EA (1994) Molecular responses to water deficit. Plant Physiol 103: 1035-1040. Cook D, Fowler S, Fiehn O, Thomashow MF (2004) A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. Proc Natl Acad Sci USA 101: 15423-15248. De Ronde J A, Cress WA, Kruger GHJ, Strasser R], Van Staden J (2004) Photosynthetic response of transgenic soybean plants, containing an Arabidopsis P5 CR gene, during heat and drought stress. J Plant Physiol 161: 1211-1224. Deblaere R, Byteriebier B, Degreve H, Deboeck F, Schell J, Van Montagu M, Leemans J (1985) Efficient octopine Ti plasmid-derived vectors for Agrobacterium —mediated gene transfer to plants. Nuc Acid Res 13: 4777-4788. Dubois M, Gilles KA, Hamilton J K, Rebers PA, Smith F (1956) Colorimetric methods for determination of sugars and related substances. Anal Chem 28: 350-356. Fowler S, Thomashow MF (2002) Arabidopsis transcriptome profiling indicated that multiple regulatory pathway, are activated during cold acclimation in addition to the CBF cold response pathway. Plant Cell 14: 1675: 1690. Fukushima E, Arata Y, Endo T, Sonnewald U, Sato F (2001) Improved salt tolerance of transgenic tobacco expressing apoplastic yeast- derived invertase. Plant Cell Physiol 42: 245—249. Garg AK, Kim J -K, Owens TG, Ranwala AP, Choi YD, Kochian LV, Wu RL (2002) Trehalose accumulation in rice plants confers high tolerance levels to different abiotic stresses. Proc Natl Acad Sci USA 99: 15898—15903. Gilmour S, Sebolt AM, Slazar MP, Evarard JD, Thomashow MF (2000) Overexpression of the Arabidopsis CBF 3 transcriptional activator mimics multiple biochemical changes associated with cold acclimation. Plant Physiol 124: 1854-1865. Gihnour SJ, Artus NN, Thomashow MF (1992) cDNA sequence analysis and expression of two cold-regulated genes of Arabidopsis thaliana. Plant Mol Biol 18: 13-21. 86 Gilmour SJ, Fowler SG, Thomashow MF (2004) Arabidopsis transcriptional activators CBF1, CBF2, and CBF 3 have matching functional activities. Plant Mol Biol 54: 767—781. Greenway H, Munns R (1980) Mechanisms of salt tolerance in non-halophytes. Ann Rev Plant Physiol 31: 149-190. Guy Cl, Nierni KJ, Brambl R (1985) Altered gene expression during cold acclimation of spinach. Proc Natl Acad Sci USA 82: 3673-3677. Haake, V, Cook D, Riechmann J L, Pineda O, Thomashow MF, Zhang JZ (2002) Transcription factor CBF 4 is a regulator of drought adaptation in Arabidopsis. Plant Physiol 130: 639-649. Hayashi H, Alia I, Mustardy L, Deshnium P, Ida M, Murata N (1997) Transformation of Arabidopsis thaliana with the codA gene for choline oxidase; accumulation of glycine-betaine and enhanced tolerance to salt and cold stress. Plant J 12: 133-142. Hillel D, Rosenzweig C (2002) Desertification in relation to climate variability and change. Advances Agronomy 77: 1-38 2002 Holmstrom KO, Somersalo S, Mandal A, Palva TE, Welin B (2000) Improved tolerance to salinity and low temperatures in transgenic tobacco producing glycine-betaine. J Exp Bot 51: 177-185. Horvath DP, McLarney BK, Thomashow MF (1993) Regulation of Arabidopsis thaliana L. (Heyn.) cor78 in response to low temperature. Plant Physiol 103: 1047-1053. Hsieh T-H, Lee J -T, Chamg YY, Chan M-T (2002b) Tomato plants ectopically expressing Arabidopsis CBF1 show enhanced resistance to water deficit stress. Plant Physiol 130: 618-626. Hsieh T-H, Lee J -T, Yang P-Z, Chiu LH, Chamg YY, Wang Y-C, Chan M-T (2002a) Heterology expression of the Arabidopsis C-repeat/Dehydration Response Element Binding factor 1 gene confer elevated tolerance to chilling and oxidative stresses in transgenic tomato. Plant Physiol 129: 1086-1094. Ingram, J Bartels D (1996) The Molecular basis of dehydration tolerance in plants. Annu Rev Plant Physiol Plant Mol Biol 47: 377-403. J aglo KR, Kleff S, Amundsen KL, Zhang X, Haake V, Zhang JZ, Deits T, Thomashow MF (2001). Component of the Arabidopsis C-repeat/dehydration-responsive element binding factor cold-response pathway are conserved in Brassica napus and other plant species. Plant Physiol 127: 910-917. 87 J aglo-Ottosen KR, Gilmour SJ, Zarka DG, Schabenberger O, Thomashow MP (1998) Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 280: 104-106. Jain RK, Selvaraj G (1997) Molecular genetic improvement of salt tolerance in plant. Biotech Ann Rev 3: 245-267. Jang IC, Oh SJ, Seo JS, Choi WB, Song SI, Kim CH, Kim WS, Seo HS, Choi YD, Nahm BH, Kim J -K (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 Physiol 131: 516—524. Kasuga M, Liu Q, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1999) Improving plant drought, salt, and fieezing tolerance by gene transfer of a single-inducible transcription factor. Nature Biotech 17: 287-291. Kasuga M, Miura S, Shinozaki K, Yamaguchi-Shinozaki K (2004) A combination of the Arabidopsis DREBIA gene and stress-inducible rd29A promoter Improved drought-and low-temperature stress tolerance in tobacco by gene transfer. Plant Cell Physiol 45: 346-350. Kishitani S, Takanami T, Suzuki M, Oikawa M, Yokoi S, Ishitani M, Alvarez-Nakase AM, Takabe T, Takabe T (2000) Compatibility of glycinebetaine in rice plants: evaluation using rice plants with a gene for peroxisomal betaine aldehyde dehydrogenase fi'om barley. Plant Cell Environ 23: 107-114. Knight H, Knight MR (2001) Abiotic stress signaling pathways: specificity and cross- talk. Trends Plant Sci 6: 262-267. Lee JT, Prasad V, Yang PT, Wu J F, Ho T H D, Charng YY, Chan MT (2003) Expression of Arabidopsis CBF1 regulated by an ABA/stress inducible promoter in transgenic tomato confers stress tolerance without affecting yield. Plant Cell Environ 26: 1181-1190. Lin C, Thomashow MF (1992) DNA sequence analysis of a complementary DNA for cold-regulated Arabidopsis gene cor15 and characterization of the COR15 polypeptide. Plant Physiol 99: 519-525. Liu Q, Kasuga M, Sakuma Y, Abe H, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1998) Two transcription factors, DREBl and DREBZ, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low- temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10: 1391-1406. 88 Maruyama K, Sakuma Y, Kasuga M, Ito Y, Seki M, Goda H, Shimada Y, Yoshida S, Shinozaki K, Yamaguchi-Shinozaki K (2004) Title: Identification of cold-inducible downstream genes of the Arabidopsis DREBIA/ CBF 3 transcriptional factor using two microarray systems. Plant J 38: 982-993. Mass EV, Hoffinan GJ (1977) Crop salt tolerance-current. J Irri Drain Div 103: 115-134. Murashige T, Skoog F (1962) A revised medium for rapid grow and bioassays with tobacco tissue culture. Physiol Plant 15: 473-497. Owens CL, Thomashow MF, Hancock JF, Iezzoni AF (2002) CBF1 orthologs in sour cherry and strawberry and the heterologous expression of CBF1 in strawberry. J Amer Soc Hort Sci 127: 489—494. Pellegrineschi A, Reynolds M, Pacheco M, Brito RM, Almeraya R, Yamaguchi- Shinozaki K, Hoisington D (2004) Stress-induced expression in wheat of the Arabidopsis thaliana DREBIA gene delays water stress symptoms under greenhouse conditions. Genome 47. 493—500. Quan R, Shang M, Zhang H, Zhao Y, Zhang J (2004) Engineering of enhanced glycine betaine synthesis improves drought tolerance in maize. Plant Biotech J 2: 477-486. Sakamoto A, Alia I, Murata N (1998). Metabolic engineering of rice leading to biosynthesis of glycine-betaine and tolerance to salt and cold. Plant Mol Biol 38: 1011—1019. Sambrook'J, Russel D W (2001) Molecular Cloning: A Laboratory Manual. 3rd Edition. Cold Spring Harbor Press. Seki M, Narusaka M, Abe H, Kasuga M, Yarnaguchi-Shinozaki K, Caminci P, Hayashizaki Y, Shinozaki K (2001) Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a full-length cDNA microarray. Plant Cell 13:61-72. Seki M, Narusaka M, Ishida J, Nanjo T, Fujita M, Oono Y, Kamiya A, Nakajima M, Enju A, Sakurai T, Satou M, Akiyama K, Taji T, Yamaguchi-Shinozaki K, Caminci P, Kawai J, Hayashizaki Y, Shinozaki K (2002) Monitoring the expression profile of 7000 Arabidopsis genes under drought, cold and high-salinity using fiill length cDNA microarray. Plant J 31: 279-292. Shinozaki K, Yamaguchi-Shinozaki K (1997) Gene expression and signal transduction in water-stress response. Plant Physiol 115: 327-334. Shinozaki K, Yamaguchi-Shinozaki K, Seki M (2003) Regulatory network of gene expression in the drought and cold stress responses. Curr Opin Plant Biol 6: 410- 41 7. 89 Stockinger EJ, Gillmour SJ, Thomashow, MF (1997) Arabidopsis thaliana CBF1 encodes an AP2 domain-containing transcriptional activator that binds to the C- repeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in response to low temperature and water deficit. Proc Natl Acad Sci USA 94: 1035- 1040. Tabei Y, Kitada S, Nishizawa Y, Kikuchi N, Kayano T, Hibi T and Akutsu K (1998). Transgenic cucumber plants harboring a rice chitinase gene exhibit enhanced resistance to gray mold (Battytis cinerea). Plant Cell Rep 17: 159-164. Tanford C (1978) The hydrophobic effect and the organization of living matter. Science 200: 1012-1018. Tester M, Davenport R (2003) Na+ tolerance and Na+ transport in higher plants. Ann Bot 91: 503-527. Thomashow MF (1999) Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu Rev Plant Physiol Plant Mol Bio 50: 571-599. Troll W, Lindsley J (1955) A photometric method for the determination of proline. J Biol Chem 215: 655-660. Vogel JT, Zarka DG, Van Buskirk HA, Fowler SG, Thomashow MF (2005) Roles of the CBF 2 and ZAT12 transcription factors in configuring the low temperature transcriptome of Arabidopsis. Plant J 41: 195—21 1. Wang W, Vinocur B, Altman A (2003) Plant responses to drought, salinity and extreme temperatures: towards genetic engineering for stress tolerance. Planta 218: 1-14. Xiong L, Zhu J -K (2002) Molecular and genetic aspects of plant responses to osmotic stress Plant Cell Environ 25: 131-139. Yamaguchi-Shinozaki K, Shinozaki K (1994) A novel cis-acting element in an Arabidopsis gene is involved in responsiveness to drought, low-temperature, or high-salt stress. Plant Cell 6: 251-264. Zhang X, Fowler SG, Cheng H, Lou Y, Rhee SY, Stockinger EJ, Thomashow MF (2004) Freezing-sensitive tomato has a functional CBF cold response pathway, but a CBF regulon that differs from that of fi'eezing-tolerant Arabidopsis. Plant J 39: 1-15. Zhu J -K, Hasegawa PM, Bressan RA (1997). Molecular aspects of osmotic stress in plants. Crit Rev Plant Sci 16: 253-277. 90 Chapter III Cucumber (Cucumis sativus L.) plants expressing the Arabidopsis thaliana-transcriptional regulators, CBF] and CBF3, are more tolerant to salinity stress under field conditions. Introduction Stress caused by high concentrations of NaCl in soil or irrigation water negatively influences productivity of major agricultural crops (McWilliam, 1986; Zang and Blumwald, 2001). Statistical assessments of naturally salt-affected areas worldwide vary, but in general it is estimated that close to 1 billion hectares (approximately 7% of the world’s land area) have naturally saline soil (Ghassemi et al., 1995). In addition to the naturally affected areas,it is estimated that 77 million hectares have become salt affected due to extensive agricultural practices, mainly in irrigated areas worldwide. It has been estimated that these numbers will increase to affect up to 50% of the total arable land by the year 2050 (Wang et al., 2003). Moreover, the demand to increase food production for the continuously grong world population will result in the need for more land for agricultural production, leading to an increase in salinity affected areas (Ghasserrri et al., 1995). Economic damage due to soil salinization has been difficult to assess, but Ghassemi et al. (1995) estimated that economic damage due to soil salinization at the Colorado River Basin is about 750 million US dollars/year and was about 330 and 208 million US dollars/year for the Punjab area and the Murray—Darling Basin in Australia. 91 Excess NaCl in soil and irrigation water causes hyperosmotic and hyperionic stress effects, which if sufficiently severe can result in plant death (Bohnert et al., 1999; Hasegawa et al., 2000). The hyperosmotic effect results from concentration of extracellular solutes, which causes a flux of water out of the cell, a decrease in the osmotic potential within the cell, and in the cellular turgor pressure (Lichtentaler, 1995). The hyperionic effect resulting from exposure to high salinity leads to “toxic sodium effect”, whereby excess Na+ in the cytoplasm causes an unbalance of other essential ions such as K+ and Ca+ (Bohnert and Jensen, 1996; Hasegawa et al., 2000). High concentration of Na+ ion in the cytosol causes metabolic toxicity, in part due to the competition between K+ and Na+ for binding sites for several enzymes (Tester and Davenport, 2003). High Na+ and Cl' concentrations interfere with enzyme function, protein synthesis, structure and solubility, and membrane fluidity and function (Blum, 1988) Traditional breeding programs to develop salinity tolerance in plants have had modest success due to difficulties in establishing selection criteria, limited availability of sources of genetic resistance, quantitative nature of resistance, and the variety of mechanisms involved in salinity tolerance (Flowers and Yeo, 1995; Quesada et al., 2002). For example, when evaluating yield performance of a crop under saline conditions, many factors can influence performance, including variation in salinity levels within a field, or possibility of interaction between salinity level and other environmental factors such as soil fertility, drainage quality, and water loss due to transpiration (Flowers, 2004). Thus, using yield components as main criteria for selection requires a long period and multiple locations for testing, and evaluation (Blum, 1989). Furthermore, results from several 92 groups indicate that QTL linked to salinity tolerance at a given developmental stage may differ fiom those linked to tolerance at another developmental stage (Greenway and Munns, 1980; Foolad, 1999; Cushman and Bohnert, 2000; Quesada et al., 2002; FoOlad, 2004). Physiological and molecular studies aimed toward understanding plant response to salinity stress have indicated the complexity of this phenomenon, where an entire cascade of biochemical and cellular changes is necessary to adapt to high salinity stress (Bohnert and Jensen, 1996). Gene expression analysis, in the model plant A. thaliana under different dehydration inducing conditions (drought, salinity and fi'eezing temperatures), revealed changes in expression patterns of several groups of genes. One of the first groups that shows immediate changes are those that encode transcription factors (TFs), mitogen activated protein kinases (MAPKs), dephosphorylation enzymes, and chromatin remodeling proteins (Thomashow, 1999; Knight and Knight, 2001; Hasegawa et al., 2000; Xiong et al., 2002; Zhu, 2002; Shinozaki et al, 2003; Seki et al., 2003; Vogel et al., 2005). This primary response is followed by activation of multiple mechanisms that are essential for plant adaptation to dehydration stresses (Bohnert and Jensen, 1996; Ingram and Bartels 1996; Campbell and Close, 1997). The multigenic nature of plant response to salinity suggests that induction of multiple adaptive mechanisms at the same time might be a good strategy to engineer salinity tolerance in plant species. The CBF/DREB [CRT (C-repeat) Binding Factor /DRE (Drought Response Element Binding)] gene family encodes transcriptional activators that work as master switches in regulating plant response to dehydration-inducing conditions (Thomashow, 1999; Shinozaki et al., 2003). Expression of the CBF gene family is 93 activated in response to low temperatures, drought or high salinity (Stockinger et al., 1997; J aglo-Ottosen et al., 1998; Liu et al., 1998; Kasuga et al., 1999; Haake et al. 2002). Transgenic Arabidopsis plants overexpressing CBF/DREB genes showed elevated levels of resistance to dehydration stresses relative to their nontransgenic counterpart, as determined by electrolyte leakage and whole plant test assays (J aglo-Ottosen et al., 1998; Kasuga et al., 1999). Similarly, in growth chamber trials, transgenic Arabidopsis plants overexpressing CBF3/DREBIa had enhanced resistance to drought and salinity stresses (Kasuga et al., 1999; Kasuga et al., 2004); expression of CBF1/DREB] b gene in tomato increased the resistance levels to salinity and drought (Hsieh et al., 2002; Lee et al., 2003); and expression of either CBF1/DREB] b or CBF 3/DREBI a genes in cucumber plants reduced their susceptibility to salinity and drought stress compared to their nontransgenic control counterparts (Tawfik and Grumet, 2001 and 2003). In general, the increase in dehydration stress resistance is accompanied by increased membrane stability and/or accumulation of compatible solutes, especially proline and soluble carbohydrates (J aglo-Ottosen et al., 1998; Kasuga et al., 1999; Gilmour et al., 2000; J aglo et al., 2001; Seki et al., 2001; Seki et al., 2002; Haake et al. 2002; Hseih et al., 2002b; Tawfik and Grumet, 2003; Kasuga et al. 2004). This elevation in compatible solute accumulation in CBF/DREB expressing transgenic plants, especially in soluble sugars and proline, has been described as a signature for CBF/DREB expression (Cook et al., 2004). Despite these successful examples which clearly demonstrate the potential value of the CBF/DREB system in increasing dehydration stress tolerance in plants, enhanced tolerance has not yet been demonstrated in field conditions. Indeed, Flowers (2003) stated that “after years of research using transgenic plants to alter salt tolerance, the value of this 94 approach has yet to be established in the field”. To our knowledge, there are only two published reports that evaluated genetically engineered plants for enhanced dehydration stress tolerance under field conditions. Quan et al. (2004) reported enhanced grain yield production by transgenic maize plants expressing a gene for betaine aldehyde dehydrogenase following drought stress period of 21 days. Xue et- al. (2004) tested wheat plants expressing the Arabidopsis tonoplast H’L/Na+ antiporter gene for their ability to grow in saline soil and reported higher grain production in the transgenic plants compared to the nontransgenic controls. In this work we sought to evaluate the performance of CBF-expressing cucumber plants under salinity stress in field conditions (Tawfik and Grumet, 2001; Tawfik and Grumet, 2003). Cucumber is known to be sensitive to salt (Mass and Hoffman, 1977). Salinity delays seed germination and seedling emergence, decreases leaf expansion rate and water potential, and decreases plant photosynthesis and yield (Chartzoulakis, 1994; Tazuki, 1997). Previously, we demonstrated that cucumber plants expressing CBF] and CBF 3 genes had elevated levels of resistance to salinity and drought conditions in greenhouse conditions. In the current study, CBF] and CBF3-expressing cucumber plants were tested for their ability to withstand continuous irrigation with 100 mM NaCl for 25 days under field conditions. Transgenic 35S:CBF cucumber plants had higher levels of proline and soluble sugars in leaves and accumulated higher levels of K+ and Ca4+ ions in roots relative to non-transgenic controls in the presence or absence of salinity stress. In response to salinity, the CBF-cucumber lines exhibited significantly less reduction in growth and fi'uit yield than did the non-transgenic controls, demonstrating the potential effectiveness of CBF in conferring salt stress resistance in the field. 95 Materials and methods Plant materials The transgenic 35S-CBFI and 35S-CBF 3 cucumber lines were produced as described in chapter 11. Of the 13 CBF-expressing lines, four lines were selected for the field trial; two lines expressing CBF1, A1 (high level of CBF1 expression) and A5 (low level of CBF1 expression); and two lines expressing CBF 3 gene, Bl (low level of CBF 3 expression) and, B5 (high level of CBF 3 expression) (Chapter 11). Two types of controls were included: wild-type parental cultivar “Straight 8” (Hollar Seeds, Rocky Ford, Co.) and azygous sibling progeny of the transgenic CBF-cucumber lines. T2 segregating azygous progeny and commercial “Straight 8” seeds were planted in trays (51 x 40 x 6.5 cm, 32 cells/tray, Hummert "‘ International, Earth City, Mo) in the greenhouse and screened by PCR for the presence of the CBF genes at the 2-3-leaf stage. DNA was extracted from 200 mg young leaf samples of seedlings using the Wizard DNA purification kit (Promega, Cat # A7951, Madison, WI). PCR was carried out using CBF1- and CBF3-specific primers (Stockinger et al., 1997). Non-transgenic segregants from the different CBF-lines were pooled for the azygous control plots. Field salinig experiment The field experiment was arranged in a split plot design, with four replications with salt treatment as the main plot and genotype as the sub-plot. Six genotypes were tested [four T2 CBF-families, parental Straight 8 (wild-type WT) and azygous segregant T2 progeny (Az)], with 6 plants per subplot. To allow for regulation of salinity levels, seedlings were transplanted into 50 x 30 x 15 cm plastic bags filled with 22.5 kg 96 playground sand (Sandastic Co., IL) and perforated with drainage holes along the edges. Plants were spaced 60 cm apart along rows of 1.5 m wide black plastic mulch; between row spacing was 2.1 m. Two control “Straight 8” plants separated each plot. The experimental plots were surrounded by two bOrder rows of control “Straight 8” plants on all sides. Border plants were directly transplanted into the soil. Irrigation was applied manually every other day using a 750 liter water tank connected to a tractor. The plants were allowed to acclimate for three weeks before starting salt stress treatment. Two salinity levels were used (0 mM and 100 mM NaCl), and stepwise salt application was carried out with an increase of 25 mM every other day until reaching the 100 mM NaCl level. Once the desired salinity level was reached, irrigation was applied daily, until run through, (approximately 2.0 l/day). Plants were fertilized with 150 ppm 20:20:20 (N: P: K) fertilizer twice a week throughout the experiment. All measurements and leaf sampling were conducted using the middle four plants of each plot. Sampling for sugar and proline content was done four times during the experiment, at one-week intervals starting just before the initiation of salt application. Two 3-4 cm diameter leaves were collected from the main stem of each plant; sampled leaves of a given plot were combined for proline and sugar analysis. Samples were taken early in the morning and placed immediately in liquid nitrogen. Growth measurements (number of nodes on the main stem, number of branches, and number of male and female flowers) were recorded just before initiation of the salinity treatment. Fruit were harvested three times, at 12, 18, and 24 days post initiation of salinity treatment. Twenty four days post initiation of salt treatment, vines were harvested and above ground fresh 97 and dry weight was measured. After removing the above ground parts, the sand surrounding the roots was washed away with water; roots were removed and then rinsed several additional times with water before being placed in plastic bags on ice. Once in the lab, roots were washed several times with deionized water, blotted on paper towels, and then stored at -80 C° until firrther analysis. Soluble sugars, proline and mineral analysis Leaf samples collected from the field were placed in 13x100 mm glass tubes and freeze-dried for 48 hr. Samples from a given plot were pooled, ground, and split into two aliquots for proline and sugar analyses. Root samples were freeze dried for 48 hr. Samples of a given plot were pooled, ground with a morter and pestel in liquid nitrogen, and split into aliquots for proline, sugar and mineral analysis. Proline and soluble sugars analyses were carried out following the procedure described by Troll and Lindsley (1955) and Dubois et al., (1956), respectively. Mineral analyses were performed on ashed root samples prepared by placing 2.0 g of freeze dried roots into ceramic crucibles and incinerating at 500 °C for 16 hrs to insure complete ashing. Afier cooling, ash weight was determined. Weighed samples of approximately 100 mg ash were added to 25 ml of 3N HNO3 digestion solution. The samples were incubated for 1 hour at room temperature and the solution filtered through 90mm x 100, No.2, Whatrnan® filter paper (Whatrnan® International Ltd, Maidstone, England) into labeled vials. The concentration of Na+, K+ and Ca” ions were determined by the MSU soil analysis laboratory using a DC. Plasma Emission atomic analyzer (Pye Unicam SP9). 98 Results CBF-expressing cucumber plants and the nontransgenic controls had guivalent in growth under field conditions before salinig treatment Six genotypes were tested [two T2 CBF1 -fami1ies (A1 and A5), two CBF3-families (B1 and B5), parental “Straight 8” (wild-type WT) and azygous segregant T2 progeny (Az)]. Prior to initiation of salinity treatment, several grth parameters were measured (Table 3-1). All lines performed equivalently; no significant differences were observed in vine length, number of nodes, nrnnber of branches, or number of male and female flowers between transgenic cucumber plants and the nontransgenic controls (Table 3-1). Transgenic cucumber plants expressing CBF genes have elevated level of proline and total soluble sugars compare to the non-transgenic controls In the absence of salinity treatment, transgenic CBF- cucrnnber plants accumulated significantly higher levels of proline (5-10 fold) compared to the nontransgenic controls, and then levels increased throughout the growing season (Figure 3-1). As the season progressed levels of soluble sugars also were significantly higher in the non-salt stressed CBF-cucumber plants compared to their nontransgenic counterparts. When irrigated with 100 mM NaCl CBF-cucumber plants accumulated significantly higher levels of proline and total soluble sugars compared to both non salt stressed conditions. CBF lines accumulated higher levels of proline and soluble sugars than salt stressed nontransgenic controls (ANOVA, P< 0.01). Twenty one days following initiation of salinity, the CBF1 and CBF3-transgenic cucumber families, on average, had more than 15 fold higher levels of proline and 4 times higher levels of total soluble sugars compared to the nontransgenic 99 4293.53 ©2335 v 53> 303 23:93 is mm H :35 2: Be 32.3 NdH _.m v 1.1113334}. ( I.tt:. Nd H d.m Nd H ..m Nd H wN - _d N NN markers: 25:?— dewd. wdHNZ vdeA: mdflm... dedN. ded... NW.532.— 252 Nd H dd Nd H NS md H d.\. Nd H v.5 Nd H Z. Nd H To moan—3.5 he hon—:52 deoN_ de mem— mdfl dd. md w. N4; deNm. mdfl mN_ woes—...eaoafiaz vN H Nahm ...N H. m.mw 0.. .4... ddm m; H on «... H mug o; H dim A53 5w=£ 03..) 0906:; mzewhu< -swmir I}! a: I. I .. l w<- l- ; - .1< -..: It -8 u. l , . £2.59 umeowmeabnez mflzfiflumkflb II Nd r... _.m J muEEfi—ékflb I... I. 1111;!!! £298.23,— 100 deuce—awash “mom 8.83 m five 2: 5 E253: 5:53 we :23me 8 coma A699 23> dew 3%»me mos: 82833 omeowmawbéoe use .Amm use 3: mos: Mkmbfinm NH .G< Ba 3: was: $5qu NH .3 235: ease as use .3555 .38: .8 8:32 .3 033 dc: baton new £03.53 :30 3 525359. 2m c.2250 £535.35: 6:: 3.8 new c.0953 2.3 5.3 32323. 2a mmcom ammo ..o Emu mEmmEaxo woe: 35:50:“. gnome—3P 65332322.“ .58 5:5 ace—Snob 333:3. v .0 mm H 53.: En San ..omz 2:. 2: no 3 new my 8535 no 8 new 3 3:33 9: 5 2o: 9: c. @5265 mace-n Lon—cause E E .8 £33 033.3 new 3 .3 269m: 2595 he :o_uu_:E:eo< .7” 0.59.... Eases. €58 :8 98 .5258 v.8 98 E 3 N o _ illrilr O - - ...... h o m n n 8 m - on ...-w m m .2: w r 02 m w m -02 mm o? m m w :3... o - 8N ( o _ oom ( E E N o r1! - Ill 0 I O .- com we 4 com M w. m.. .- coop we - door w W $5-413 83 W 82 w 8.? Furl w _ :58 2+ 2.”; g 88 :84, m :3. < 101 counterparts. Salinity did not cause a significant increase in levels of compatible solute accumulation in the azygous or the parental “Straight 8” wild-type controls. No significant differences were found between CBF1 vs. CBF3-cucumber lines (ANOVA, P< 0.76). Azygous plants did not differ from parental “Straight 8” plants. Accumulation of compatible solutes was also tested in root samples at the termination of the experiments, 24 days post initiation of salinity stress treatment. In the absence and presence of salinity stress, CBF-cucumber roots accumulated higher levels of proline than their nontransgenic counterparts (ANOVA, P<0.05; Figure 3-2 A). No significant differences in accumulation of soluble sugar were detected between CBF-transgenic and the nontransgenic cucumber plants (Figure 3-2 B). Transgenic cucumber plants expressing CBF genes have elevated level of potassium ions in their roots compared to the non-transgenic controls Analysis of ashed root samples revealed differences in ion composition between the transgenic and the nontransgenic controls. CBF -1ines exhibited 10-15 fold higher levels of K+ ions than the nontransgenic controls in the presence and absence of salinity stress (Figure 3-3A). On the other hand, Na+ levels did not differ significantly between the transgenic and nontransgenic plants and did not increase significantly in response to salinity treatment (Figure 3-3B). Transgenic cucumber plants had markedly lower Na+/K+ ratio in roots under both non—salt stress, and salt stress conditions compared to the nontransgenic controls. The average Na+/K+ ratio did not change in the transgenic lines in response to salinity stress (Figure 3-3C). Calcium levels also differed in the CBF cucumber transgenic lines with approximately 2-fold higher levels than in the 102 1° 1 minim L_l£alt-stressed 1 A 8 1 .9 3' l »‘d ”I :52: l ‘3’ ‘l 1 5;? 3.2% 3 6 ‘ i: :3 r g 4 “‘ :w 3 it E l El ::. 2 “ if . i 5: v.: ‘ : o 3 . t: A1 A5 B1 10 :3 3 '0 2’ 5’ l V J l 1‘: a. 25 . ': £13 . g 7 g 9.41 w . g i 3: 2 i , . “5i . :3: , g,- . g o : ,‘ , . 030 ~ E i 33; B1 35 Figure 3-2. Accumulation of proline mglgdw (A) and soluble sugars mglgdw (B) in root tissues of cucumber plants growing in the field in the absence (light bars) or presence of 100 mM salinity stress (dark bars). Data represents mean 1 SE of four replicates/treatment with four plants/replicate. 103 A 0 mM NaCI 100 mM NaCl Mg/gdw 4 A1 A5 3135 A2 cc A1 A5 B1 BSAz cc B Transgenic Controls Transgenic Controls Mg/gdw cc A1 A5 B1 85 Az cc A1 A5 B1 BS Transgenic Controls Transgenic Controls C 3 ‘ NalK .9 "5 i x 2 g 1 ‘ i ii 1 OJ 1143:: ‘ 1 ....— A1 A5 B135Azcc A1 A5 B1 BSAzCC Transgenic Controls Transgenic Controls D 41 Ca , 3 i g , r: 1;) 2 ‘ 1 ; iri i r 9 i l i . c» 1 . l i 2 0 ‘ . . , 31‘ A1 A5 B1 BSAZCC A1 A5 BlBSAzCC Transgenic Controls Transgenic Controls Figure 3-3. Potassium (A), sodium (B), NalK ratio (C) and calcium content (D) in roots of cucumber plants growing in the field for 24 days, in the presence or absence of salinity treatment. Each value is the mean 1 SE of four replicates samples composed of ashed root tissue pooled from four plants/replicate. 104 nontransgenic lines under both salt stressed, and non-salt stressed conditions (Figure 3- 3D). CBF-cucumber plants had less reduction in fresh and dry weight under salt stress conditions Under non-salt stressed conditions, CBF-expressing cucumber families had lower fresh weight than the nontransgenic controls at the time of harvest; dry weight was equivalent for the CBF and non-transgenic controls (Figure 3—4; Table 3-2). Salinity treated CBF-cucumber plants on average did not show a significant reduction in fresh weight (690 g vs. 646 g in the absence or presence of salinity), while nontransgenic cucumber plants had an average reduction of 38% in fresh weight in response to salinity treatment. Similarly, on average, salinity treated CBF-expressing plants did not show a significant reduction in dry weight, while azygous and wild type Straight 8 plants had an average reduction of more than 50%. Transgenic cucumber plants had higper field under salinity conditions compared to the nontransgenic controls In the absence of salinity stress, no significant differences were observed in number of fruits or fi'uit weight between the CBF-expressing cucumber plants and the nontransgenic controls (Figure 3-4; Table 3-2). Salinity treated CBF-expressing lines did not show significant reduction in fruit number or weight, compared to a 35% reduction in fruit number and 50% reduction in fruit weight for the nontransgenic controls. Salt stressed CBF1 lines (Al and A5) had higher yields than the CBF3 lines (Bland B5). 105 6:035:00 commohm gum 3:32am. 23 {an .mcoEucoo ummmwbm gameo: 232%. 2.3 EB... 6550 VN ..8 «co—5.8.: 3228 do cocoons .o 3:395 23 5 2o: 05 E 9:255 manna .onanao E “o... an R: 22> 23.: «9295 can 8. 33:5: :3: mafia; .cozflaEzoou Amv £995 in “Eu 9: snot 2595 o>on< .Yn 2:2". 8 2 mm a 2 E #0 r r T . .L ..Mwl. l1 . l". c r O . F . . a , .m .. ,, 09 N H a , . m h 8N V J.— 4.. w m. m m w \l q M m _ w . _ 8m m C r o m Bu .3 0 Id 2 m m M com M B m. m. U. 6 1 N cm W 02 .m fl. . - illm /\ commobm gum 5 mm 89 nommmzmufiméozm kW 0 106 Sgwoa Emmouwom oEom§::o: :5 w Emmabm Egamgbéo: :8: :28: 8: 033 N .2: :5 a 8:: :8: :28: 8: 8:: 8:55-me0 :8 m2 :5 :< 8:: :8: :28: 2: 85:54:30 :8 8:: _ Abofifiommo: . 5d .36 V5 58:8 _::owo:to .3 $52 $05: :8 a 55:5 £058.: $39 80.: 320:8 b.:ao:_:w_m 8: mom—:9: FED :o 60:: 3:035: 2: :8: 880:6 >_:::o:_:w_m 2: anabogw oEowmgbéoZ 9., ... .889: 8:: :o>o 3.32 6:93.63 v 5:: 5253532: 28:3: :6: .3 mm ”m :38 2: 8: 20:» :9: :8 San 333:2: v 5:? 30:53:22: 28:3: :5: :0 mm H :88 2: 2: Emma? E: :5 :8: 8: Sam 25H: .nga: 2.3m: mmoHNE is H :0: 2.: H ..S a: H a: S H v: a: BBQ: :3.— 3:9 2 H ..2 .33”: I H32 .... H3: :2 H :3 «N H of M: H 5.: .... H 2.. .8::=.: ...—...: ..oHfiE :2 H32: . a: H :8 3: H33. :3 H 2.2: 0.: H 32 o... H 23: ...N. H 32 E :39: .3: ..mm H 3% S: H 2% :3 H :3 3: H 28 I: H S. 5 0.2. H 0.2.: . ”.8 H :22 2: H 2:: a. 3303 :8; mmmU mum—U "mu—GU .—'.:—U 8:: uioua:a.=-:ez 3:: ofiuumfizh 3:: quwm:a.:-:eZ 8:: uaowmgfi. fivaZ Eu: 2: fivaz :5 : 30:58: 36:3 mo :o:m:_:m “mo: 93: Va 8:: 35:8 o_:owm::::o: 2: :5 8:38.: o_:owm::: mEmmoExoimU .«o 22% :9: :88 :5 898:: :9: Emmy.» b: 4&6? :men: .34: 2:5. 107 Discussion The CBF gene has been demonstrated to confer dehydration stress resistance in several species in growth chamber and greenhouse studies (Kasuga et al., 1999; Hsieh et al., 2002; Lee et al., 2003; Pellegrineschi et al., 2004; Kasuga et al., 2004). This study tested the performance of the 35$:CBF-cucumber plants under field conditions. Consistent with greenhouse experiments (Chapter II), field grown transgenic 35S:CBF- cucumber plants had elevated levels of proline and soluble sugars relative to the nontransgenic controls. CBF-cucumber plants progressively accumulated higher levels of the compatible solutes throughout the growing season, even in the absence of salinity stress. Irrigation with 100 mM NaCl caused significant elevation in soluble sugars and proline levels in the leaves of CBF-cucumber plants while levels of sugar and proline did not change in the nontransgenic controls, indicating enhanced ability of the CBF-lines to respond to salinity stress. The higher levels of proline in roots of the transgenic CBF- lines indicate that CBF caused osmotic adjustment throughout the plant. In addition to the increases in compatible solutes, marked changes in ion composition in root tissues also were observed. Levels of K+ ion in the transgenic lines were about 10- 15 fold higher than the nontransgenic controls, in the presence or absence of salt stress; levels of Ca++ were approximately two-fold higher. To our knowledge this is the first report that shows the influence of CBF expression on ion composition in plants, suggesting an effect on ion transport properties in roots. K+ and CaH are major nutrients and are the two most abundantly distributed cations in plant tissue (Devlin and Witham, 1983). In addition to serving as a primary contributor to cell turgor, K” is also involved in 108 many physiological processes in plants (Zhu et al., 1997). Ca++ is critical for stabilizing and maintaining cell walls and is involved in various signal transduction pathways (Knight and Knight, 2001). Exposure of plants to high salinity leads to a “toxic sodium effect”, whereby excess Na+ in the cytoplasm causes a deficiency of essential ions such as K)“ and Ca2+, and competes with K+ ions for binding sites for several enzymes (Bohnert and Jensen, 1996; Hasegawa et al., 2000; Tester and Davenport, 2003). Living cells tend to accumulate K+ and exclude Na+, to maintain sufficient levels of K+ to perform essential functions that sodium either cannot fulfill or actively inhibits (Epstien, 1998). The increase in K+ content in CBF-cucumber roots was comparable in the presence or absence of salinity stress, suggesting that excess Na+ ions did not prevent the elevated K+ accumulation by the CBF-cucumber roots. Accumulation of Na+ did not differ between transgenic and nontransgenic plants, and was not significantly increased in the presence of salinity stress, suggesting selectivity in K+ ion uptake. A possible explanation for the enhanced in K+ and Cal+ accumulation by the CBF- cucumber roots may be due to differential expression of different transporters and channels responsible for ion uptake from the soil or subsequent movement through the plant. Recent transcriptional profiling ofArabidopsis plants overexpressing CBF/DREB genes (Seki et al., 2001; Fowler and Thomashow, 2002; Vogel et al., 2005) provided an Opportunity to examine global gene expression profile changes due to CBF/DREB. Among the upregulated genes are putative and known transporter proteins and membrane channel proteins that might play a role in preferential selectivity to K“ and Ca++ ions over the toxic Na+ ions (Fowler and Thomashow, 2002; Vogel et al., 2005). 109 One of the genes that was upregulated in CBF2 overexpressing Arabidopsis plants encodes a CaH/ATPase transporter, which contains a CRT element in its promoter region (V ogel et al., 2005). Calcium has been shown to maintain or enhance the selective absorption of potassium by plants at high concentration of of sodium (Epstien, 1998), thus such a transporter might facilitate enhanced Ca“ and K+ uptake. Whether a similar CBF-inducible transporter is present in cucumber roots, or whether other transporters may be affected, remains to be determined. In general, the samples used for the Arabidopsis transcriptional analyses performed to date have been primarily composed of shoot tissue (V ogel et al., 2005). More comprehensive analysis of root tissue may lead to identification of additional relevant genes as possible targets for CBF/DREB. Another possible explanation for K+ accumulation may be related to the elevated levels of proline accumulation in CBF-cucumber roots. It has been observed that in response to several abiotic and biotic stimuli in plants, that there is a correlation between compatible solute accumulation and KJr ion content (Garcia et al., 1993; Hare and Cress, 1997; Backor et al., 2004). Garcia et al. (1993) found that lower Na/K ratios were obtained upon treating rice roots with several osmoprotectant, including proline. Backor et al. (2004) recently found that in heavy-metal resistant strains of lichen photobionts (T rebouxia erici) levels of proline correlated with ability to block K+ ion efflux. Garg et al. (2002) observed that transgenic rice engineered for trehalose accumulation were able to maintain a higher level of selectivity for K+ over Na+ uptake in the roots. Thus expression of CBF genes, either directly (through altered gene expression) or indirectly (by changing levels of compatible solutes), might affect processes involved in ion homeostasis in plants. 110 Prior to initiation of salinity treatment, no significant differences were observed between the transgenic and the nontransgenic plants as measured by vine length, number of nodes, number of branches, female and male flowers. At the end of the experiment, in the absence of salinity, CBF lines had less fi'esh weight than the nontransgenic controls; however, dry weight and fruit number and fi'uit weight were equivalent to the nontransgenics. These results suggest the possibility to obtain positive effects of CBF for stress resistance without a negative impact on fruit yield, although this would need to be verified in more extensive testing situations. In the presence of salinity stress, CBF-transgenic plants showed enhanced tolerance to stress conditions compared to the nontransgenic lines as measured by fresh weight, dry weight, fiuit number, and fiuit weight. CBF1-cucumber lines did not show reduction in yield, compared to 34% reduction in fi'uit number and more than 50% reduction in fruit weight in the nontransgenic lines. Thus the CBF genes conferred increased salt stress tolerance to cucumber plants. The growing demand to increase food production worldwide, requires a multi- disciplinary approach that will include adding new land to the agricultural production area, the use of low quality saline water and the reuse of drainage waters, as well as developing new salinity tolerant plants capable of adapting to a wider range of dehydration-inducing stresses. Introduction of the CBF gene into cucumber activated a variety of salt adaptive responses including increased in compatible solute accumulation and maintenance of higher I(+/Na+ balance. The CBF-transgenic cucumbers showed enhanced resistance to salinity stress, with minimal or no reduction in growth and fruit 11] yield in the absence of salt stress, making this approach very promising to engineer dehydration resistance in crops. 112 References Backor M, Fahselt D, Wu CT (2004) Free proline content is positively correlated with copper tolerance of the lichen photobiont T rebowcia ericz' (Chlorophyta). Plant Sci 167: 151-157. Blum A (1988) Salinity resistance. In: Plant breeding for stress environment. Boca Raton, CRC Press 163-179. Blum A (1989) Breeding methods for drought resistance. In Jones HG, Flowers TJ, Jones MB eds. Plant under stress: Biochemistry, physiology and ecology and their application to plant improvement. Cambridge University Press, 217-234. Bohnert HJ, Jensen RG (1996) Strategies for engineering water-stress tolerance in plants. Trends Biotch 14: 89-97. Bohnert HJ, Su H, Shen B (1999) Molecular mechanisms of salinity tolerance. In: Molecular responses to cold, drought, heat and salt stress in higher plants. Landes Company 29-62. Campbell'SA, Close TJ (1997) Dehydrins: genes, proteins and association with phenotypic traits. New Phytol 137: 61-74. Chartzoulakis KS (1994) Photosynthesis, water relation and leaf growth of cucumber exposed to salt stress. Scientia Hort. 59: 27- 35. Cook D, Fowler S, Fiehn O, Thomashow MF (2004) A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. Proc Natl Acad Sci USA 101: 15423-15248. Cushman J C, Bohnert H (2000) Genomic approaches to plant stress tolerance. Curr Opin Plant Biol 3: 117-124. Dubois M, Gilles KA, Hamilton JK, Rebers PA, Smith F (1956) Colorimetric methods for determination of sugars and related substances. Anal Chem 28: 350-356. Epstein E (1998) How calcium enhances plant salt tolerance (1998) Science 280: 1906- 1907. Flowers TJ (2004) Improving crop salt tolerance. J Exp Bot 55: 307-319. Flowers TJ, Yeo AR (1995) Breeding for salinity tolerance in crop plants: where next? Aust J Plant Physiol 22: 875-884. 113 Foolad MR (1999) Comparison of salt tolerance during seed germination and vegetative growth in tomato by QTL mapping. Genome 42: 727-734. Foolad MR (2004) recent advances in genetics of salt tolerance in tomato. Plant Cell Tissue Organ Cult 76: 101-119. Forster BP (2001) Mutation genetics of salt tolerance in barley: An assessment of Golden Promise and other semi-dwarf mutants. Euphytica 120: 317-328. Fowler S, Thomashow MF (2002) Arabidopsis transcriptome profiling indicated that multiple regulatory pathway are activated during cold acclimation in addition to the CBF cold response pathway. Plant Cell 14: 1675-1690. Garcia AB, Engler AJ, Van Montagu M, Caplan A (1993) Effect of NaCl and osmoprotectants on growth inhibition, ion accumulation, and gene expression in rice (0ryza sativa). In: Close TJ and Bray EA (eds) Plant responses to cellular dehydration during environmental stress, pp 272—273. Garg AK, Kim JK, Owens TG, Ranwala AP, Do Choi Y, Kochian LV, Wu RJ (2002) Trehalose accumulation in rice plants confers high tolerance levels to different abiotic stresses. Proc Natl Acad Sci USA 99: 15898-15903. Ghassirni F, J akeman AJ, Nix HA (1995). Salinisation of land and water resources: human causes, extent, management and case studies. Canberra, Australia: The Australian National University, Wallingford, Oxon, UK: CAB International. Gilmour S, Sebolt AM, Slazar MP, Evarard JD, Thomashow MF (2000) Overexpression of the Arabidopsis CBF 3 transcriptional activator mimics multiple biochemical changes associated with cold acclimation. Plant Physiol 124: 1854-1865. Greenway H, Munns R (1980) Mechanisms of salt tolerance in non-halophytes. Ann Rev Plant Physiol Plant Mol Biol. 31: 149-190. Haake, V, Cook D, Riechmann J L, Pineda O, Thomashow MF, Zhang JZ (2002) Transcription factor CBF4 is a regulator of drought adaptation in Arabidopsis. Plant Physiol 130: 639-649. Hare PD, Cress WA (1997) Metabolic implications of stress-induced proline accumulation in plants. Plant Grth Reg 21: 79—102. Hasegawa PM, Bressan RA, Zhu J K, Bohnert HJ (2000) Plant cellular and molecular responses to high salinity. Ann Rev Plant Physiol Plant Mol Biol 51: 463-499. Hsieh T-H, Lee J -T, Charng YY, Chan M-T (2002b) Tomato plants ectopically expressing Arabidopsis CBF1 show enhanced resistance to water deficit stress. Plant Physiol 130: 618-626. 114 Hsieh T-H, Lee J-T, Yang P-Z, Chiu LH, Chamg YY, Wang Y-C, Chan M-T (2002a) Heterology expression of the Arabidopsis C-repeat/Dehydration Response Element Binding factor 1 gene confer elevated tolerance to chilling and oxidative stresses in transgenic tomato. Plant Physiol 129: 1086-1094. Ingram, J Bartels D (1996) The Molecular basis of dehydration tolerance in plants. Ann Rev Plant Physiol Plant Mol Biol 47:377-403. J aglo KR, Kleff S, Amundsen KL, Zhang X, Haake V, Zhang JZ, Deits T, Thomashow MF (2001). Component of the Arabidopsis C-repeat/dehydration-responsive element binding factor cold-response pathway are conserved in Brassica napus and other plant species. Plant Physiol 127: 910-917. Jaglo-Ottosen KR, Gihnour SJ, Zarka DG, Schabenberger O, Thomashow MF (1998) Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 280: 104-106. Kasuga M, Liu Q, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1999) Improving plant drought, salt, and freezing tolerance by gene transfer of a single-inducible transcription factor. Nature Biotech 17: 287-291. Kasuga M, Miura S, Shinozaki K, Yamaguchi-Shinozaki K (2004) A combination of the Arabidopsis DREBIA gene and stress-inducible rd29A promoter Improved drought- and low-temperature stress tolerance in tobacco by gene transfer. Plant Cell Physiol 45: 346-3 50. Knight H, Knight MR (2001) Abiotic stress signaling pathways: specificity and cross- talk. Trends Plant Sci. 6: 262-267. Lee JT, Prasad V, Yang PT, Wu J F , Ho THD, Chang YY, Chan MT (2003). Expression of Arabidopsis CBF1 regulated by an ABA/stress inducible promoter in transgenic tomato confers stress tolerance without affecting yield. Plant Cell Environ 26: 1181— 1190. Lichtentaler HK (1995) Vegetation stress: an introduction to the stress concept in plants. J Plant Physiol 148: 4-14. Liu Q, Kasuga M, Sakuma Y, Abe H, Miura S, Yamaguchi-Shinozaki K, Shinozaki K (1998) Two transcription factors, DREBl and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low- temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10: 1391-1406. Mass EV, Hoffman GJ (1977) Crop salt tolerance-current. J Irri Drain Div 103: 115-134. 115 McWilliam JR (1986) The national and international importance of drought and salinity effects on agricultural production. Aust J Plant Physiol 13: 1-13. Pellegrineschi A, Reynolds M, Pacheco M, Brito RM, Almeraya R, Yarnaguchi- Shinozaki K, Hoisington D (2004) Stress-induced expression in wheat of the Arabidopsis thaliana DREBIA gene delays water stress symptoms under greenhouse conditions. Genome 47: 493—500. Quan R, Shang M, Zhang H, Zhao Y, Zhang J (2004) Engineering of enhanced glycine betaine synthesis improves drought tolerance in maize. Plant Biotech J 2: 477-486. Quesada V, Garcia-Martinez S, Piqueras P, Ponce MR, Micol J L (2002) Genetic architecture of NaCl tolerance in Arabidopsis Plant Physiol 130: 951-963. Sambrook J, Russel D W (2001) Molecular Cloning: A Laboratory Manual. 3rd Edition. Cold Spring Harbor Press. Seki M, Kamei A, Yamaguchi-Shinozaki K, Shinozaki K (2003) Molecular responses to drought, salinity and different paths for plant protection. Curr Opin Biochem 14: 194-199. Seki M, Narusaka M, Abe H, Kasuga M, Yamaguchi-Shinozaki K, Carninci P, Hayashizaki Y, Shinozaki K (2001) Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a full-length cDNA microarray. Plant Cell 13:61-72. Seki M, Narusaka M, Ishida J, Nanjo T, Fujita M, Oono Y, Kamiya A, Nakajima M, Enju A, Sakurai T, Satou M, Akiyama K, Taji T, Yamaguchi-Shinozaki K, Caminci P, Kawai J, Hayashizaki Y, Shinozaki K (2002) Monitoring the expression profile of 7000 Arabidopsis genes under drought, cold and high-salinity using full length cDNA microarray. Plant J 31 : 279-292. Shinozaki K, Yamaguchi-Shinozaki K, Seki M (2003) Regulatory network of gene expression in the drought and cold stress responses. Curr Opin Plant Biol 6: 410-417. Stockinger EJ, Gillmour SJ, Thomashow, MF (1997) Arabidopsis thaliana CBF1 encodes an AP2 domain-containing transcriptional activator that binds to the C- repeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in response to low temperature and water deficit. Proc Natl Acad Sci USA 94: 1035- 1040. Tawfik M, Grumet R (2001) Production of transgenic cucumber plants (Cuucumis sativus L.) plants expressing the Arabidopsis cold stress related transcription factors, CBF1 and CBF3. Hortscience 36: 455. 116 Tawfik M. and Grumet R (2003) Production of transgenic cucumber plants over- expressing the Arabidopsis thaliana transcriptional factor CBF show enhanced tolerance to salinity. Hortscience 38: 814. Tazuki A (1997) Growth of cucumber fi'uit as affected by the addition of NaCl to nutrient solution. J Jpn Soc Hort Sci 66: 519—526. Tester M, Davenport R (2003) Na + tolerance and Na+ transport in higher plants. Annals Bot. 91: 503-527. Thomashow MF (1999) Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Ann Rev Plant Physiol Plant Mol Biol 50: 571-599. Troll W, Lindsley J (1955) A photometric method for the determination of proline. J Biol Chem 215: 655-660. Vogel G, Fiehn O, Bressel J LR, Boller T, Wiemken A, Aeschbacher RA, Wingler A (2001) Trehalose metabolism in Arabidopsis: occurrence of trehalose and molecular cloning and characterization of trehalose-6-phosphate synthase homologues. J Exp Bot 52: 1817-1826. Wang W, Vinocur B, Altman A (2003) Plant responses to drought, salinity and extreme temperatures: towards genetic engineering for stress tolerance. Planta 218: 1-14. Xiong L, Zhu J -K (2002) Molecular and genetic aspects of plant responses to osmotic stress Plant Cell Environ 25: 131-139. Xue Z-Y, Zhi D-Y, Xueb G-B, Zhang H, Zhao Z-Y, Xia G-M (2004) Enhanced salt tolerance of transgenic wheat (T ritz'vum aestivum L.) expressing a vacuolar Na+/H+ antiporter gene with improved grain yields in saline soils in the field and a reduced level of 1eafNa+. Plant Sci 167: 849—859. Zhang HX, Blumwald E (2001) Transgenic salt-tolerance tomato plants accumulate salt in foliage but not in fruit. Nature Biotech 19: 765-768. Zhang X, Fowler SG, Cheng H, Lou Y, Rhee SY, Stockinger EJ, Thomashow MF (2004) Freezing-sensitive tomato has a functional CBF cold response pathway, but a CBF regulon that differs from that of freezing-tolerant Arabidopsis. Plant J 39: 1-15. Zhu J -K (2002) Salt and drought stress signal transduction in plants. Ann Rev Plant Biol 53: 247-273. Zhu J -K, Hasegawa PM, Bressan RA (1997). Molecular aspects of osmotic stress in plants. Crit Rev Plant Sci 16: 253-277. 117 Chapter IV Introduction of the celery mannose—6-phosphate reductase (M6PR) gene for mannitol production into cucumber (Cucumis sativus L). Introduction Plant response to unfavorable conditions requires adjustment at the molecular, cellular and whole plant level (Greenway and Munns, 1980; Ingram and Bartels, 1996 and Zhu et al., 1997). One of the many mechanisms that plants have developed to overcome the low osmotic potential associated with salinity and drought conditions, is the ability-to accumulate compatible solutes in the cytoplasm (Tarczynski et al., 1993; Bohnert and Jansen, 1996; Shen et al., 1997; Sakamoto and Murata, 2000). Compatible solutes (e.g., proline, sugar alcohols, fi'uctans, trehalose, quaternary ammonia compounds, and tertiary sulfonic compounds), are non-toxic organic metabolites of low molecular weight that can decrease the osmotic potential of cells without interfering with cellular metabolism. The compounds can also serve as osmoprotectants to help stabilize membranes and macromolecular structures (Bohnert and Jensen, 1996, Stoop et al., 1996, Zhang et al., 1999). Thus, attempts to engineer enhanced salinity tolerance in plants have included the use of genes encoding key enzymes for biosynthesis of compatible solutes such as marmitol (Tarczynski et al. 1993; Karakas et al., 1997; Abebe et al., 2003; Zhifang and Loescher, 2003), proline (Kishor et al., 1995; Zhu et al., 1998; Ronde et al., 2000), and glycine-betaine (Holmstrom et al., 1994; Hayashi et al., 199 7; Sakamoto et al., 1998; Holmstrom et al., 2000; Kishitani et al., 2000; Jia et al., 2002) 118 Sugar alcohols such as mannitol, galactitol and sorbitol represent the chemically reduced form of aldoses or ketose sugars (Loescher and Everard, 1996, Loescher and Everard, 2000). it also has been suggested that sugar alcohols may play an important role in scavenging active oxygen species and preventing peroxidation of lipids, which can lead to membrane damage (Halliwell et al., 1988; Smirnoff and Cumbes, 1989; Tarczynski et al., 1993; Bohnert and Jensen, 1996; Stoop et al., 1996; Bohnert et al., 1999). Synthesis of mannitol also was suggested to work as a supplemental mechanism to dissipate reducing power (N ADPH) accumulated during the light reactions of photosynthesis (Loescher, 1987). Mannitol is the most common form of sugar alcohol in nature, and it has been reported in numerous plant species, including many crops such as carrot, parsley, celery, green beans, cabbage, pumpkins, coffee and olive trees (Loescher et al., 1992; Stoop et al., 1996). Plant species that produce mannitol as one of their primary photosynthetic products tend to have a substantial dehydration stress tolerance as in celery, coffee and olive trees (Loescher et al., 1992; Stoop et al., 1996). A role for mannitol in adaptation to dehydration stress is supported by changes in mannitol production in response to salinity and drought stress. Exposure of celery plants to 300 mM NaCl resulted in a shift in the pool size of sucrose, mannitol and starch towards nearly exclusive accumulation of mannitol (Everard et al., 1994). Similar results were observed when testing celery plants grown in saline hydroponic culture equivalent to 30% sea water (Stoop and Pharr, 1994). No differences in dry weight were observed between salt stressed and non stressed plants, suggesting that the total carbohydrate assimilation was not affected by salinity. Furthermore, salt stress induces expression of 119 key mannitol biosynthetic enzymes and down regulates mannitol eatabolic enzymes ( Zamski et al., 1996; Loescher and Everard, 2000; Zamski et al., 2001; Zhifang and Loescher, 2003). Several studies reported enhanced stress protection of transgenic plants by introducing bacterial genes for sugar alcohol production. Expression of the Echerichz'a coli mannitol-l-phosphate dehydrogenase gene (mtID) in tobacco resulted in the accumulation of mannitol in leaves and roots of transgenic tobacco plants as detected by NMR and mass spectroscopy (Tarczynski et al. 1992; Tarczynski et al., 1993; Karakas et al., 1997) The mannitol—accumulating tobacco plants had an elevated level of tolerance to 100 mM NaCl, as indicated by flesh weight, plant height and root biomass. In addition to salt stress resistance, mtID-expressing tobacco plants had higher relative water content in their leaf tissues, in response to drought stress (Karakas etal., 1997). Eggplant (Solanum melongena L.) seedlings expressing the mt] D gene exhibited increased tolerance to salinity stresses as measured by increased germination rate and higher fresh and dry weight compared to the nontransgenic controls at 200 mM NaCl (Prabhavathi etal., 2002). Transgenic T2 mtID-wheat plants subjected to 150 mM NaCl showed less reduction in fresh and dry weight than did the non-transgenic wheat (Abebe et al., 2003). Similarly, expression of the E. coli GutD gene encoding glucitol-6-phosphate dehydrogenase, a key enzyme for biosynthesis of the sugar alcohol sorbitol in maize plants, also increased sorbitol accumulation and enhanced salt tolerance compared to the nontransgenic controls (Liu et al., 1999). Rice plants (Oryza sativa L.) expressing the E. coli GutD and the mtID genes, were able to accumulate both sorbitol and mannitol in 120 their vegetative tissue and were more salt tolerant than their nontransgenic counterparts (T ilahon et al, 2003). In addition to the use of the bacterial genes, mannitol biosynthetic genes from celery also have been used to engineer mannitol accumulation (Zhifang and Loescher, 2003). In celery, mannitol is synthesized from fi'uctose-6-phosphate in three steps:- Fructose-G-P <—> mannose-S-P __, mannitol-1-P —> mannitol PMI M6PR Pase Fructose-6-P (fructose-6-phosphate), mannose—6-P (mannose-6-phosphate), mannitol-l -P (mannitol-l-phosphate), PMI (phosphomannose isomerase), M6PR (mannose-6-phosphate reductase), Pase (phosphatase) In celery, fi'uctose-6-phosphate and mannose-6-phosphate are in an equilibrium state; fructose-6-phosphate is converted into mannose-6-phosphate by phosphomannose isomerase (PMI). Conversion of mannose-6-phosphate into mannitol-l-phosphate which is performed by mannose-6-phosphate reductase (M6PR) is the first committed step in the pathway and appears to be the primary site of regulation of mannitol biosynthesis (Everard and Loescher, 1997; Zhifang and Loescher, 2003). Mannitol-l-phosphate is then converted into mannitol by phosphatase enzyme (Pase). The M6PR gene was fist cloned by Everard et al. (1997) fi'om celery plants (Apium graveolens L.). Expression of the M6PR gene, in Arabidopsis thaliana, a non-mannitol accumulating species, enabled plants to grow and complete their normal life cycle (including seed production) in the presence of 300 mM NaCl (Zhifang and Loescher, 2003). 121 In this work, I investigated the possible use of the M6PR gene to engineer enhanced dehydration stress tolerance in cucumber plants. Transgenic M6PR-cucumber plants were produced in our laboratory, analyzed for the presence of M6PR gene by PCR and tested for accumulation of mannitol in leaf tissue. 122 Materials and methods Plant constructs, transformation and seed production The Agrobacterium plant transformation construct containing the M6PR gene, under the control of the CaMV 35S promoter (Zhifang and Leoscher, 2003) was kindly provided by W. H. Leoscher (Michigan State University). Cucumber transformation was performed based as described in Chapter II. DNA isolation and PCR analysis DNA was isolated from young leaf samples; 200 mg leaf tissue was ground in liquid nitrogen and extracted using the Wizard DNA purification kit (Promega, Cat # A7951, Madison, WI). DNA quality was determined by gel electrophoresis (Sambrook and Russell, 2001). PCR was carried out according to the procedure of Sambrook and Russell (2001), using M6PR specific primers (RG278, forward, CACAGCACACACACCAC and RG279, reverse CACACATTCCCCTCCACA). ELISA analysis ELISA test was used to test for the presence of the NPT 11 protein in the To transgenic plants using NPTII-ELISA® kit (Agdia Inc., Elkhart, IN). One leaf disc of each plant was collected and placed in 96 well ELISA plate and stored in -80 C° until further analysis. Just before starting the ELISA procedure, leaf discs were left at room temperature followed by re-fi'eezing for 2-3 times to ensure cell wall leakage of tissues. 123 Mannitol analysis One leaf (6-8 cm diameter; approximate fresh weight 1.5 g) was collected from To and T1 plants, freeze-dried for 48 hr, ground and placed in 13x100 mm disposable glass tubes (Cat No. 47729-572, VWR international, West Chester, PA). Total soluble carbohydrates were extracted according to the procedure of Loescher et al. (1997) (Appendix A). Standards were previously prepared using 0.1 g of fructose, glucose, sucrose, rafflnose, mannitol, myo-inositol, all mixed together and dissolved in 100 ml H20 (this should give 1000 ppm) to establish elution peak time as discussed in more details at appendix A. 124 Results and discussion Introduction and expression of the celery (Apium gaveolens}M6PR gene in cucumber plants Six transgenic cucumber plants were produced, and the presence of the M6PR gene was verified in the To plants by PCR (Figure 4-1). Successful transfer of the M6PR gene into T1 progeny was verified using PCR analysis (Table 4-1). )8 analysis of segregating progenies was consistent with a single gene insertion for the six families. Table 4-1. Segregation analysis of T1 transgenic cucumber lines expressing the celery- M6PR gene. Presence or absence of the gene was determined by PCR analysis Line T1 segregation (trans: non) 12 @1) M1 29 : 8 0.22 ns M2 31 :18 5.30 ns M3 42:9 4.6 ns M4 38:12 0.03 ns M5 34:8 3.86 ns M6 22:11 2.75 ns 125 H20 4. 12M1M2M3M4 7 8 9M511§E 415 bp Figure 4-1. PCR analysis of the presence of the M6PR gene in To cucumber plants. Lane 1-11: putative transgenic cucumber plants. Lane 12-14: PCR mix, plasmid control and H20. The arrow indicates the expected 415 bp product size. 126 mm A. (61mm MOM) Mannitol 8.495 21.258 -" 25.080 M6PR M3 3. 8 r; J 1} 29.227 35000 30000‘ 25000 150001 100001 FlDt A. (mammal) %om 11.145 12.455 21.259 Wild-type 7‘ or .5 O ... .N G ... U 31 7 25.164 27.5 m 127 Figure 4-2. Gas chromatography for mannitol analysis in one of the cucumber To Plants (M3) GOP) and nontransgenic wild-type cucumber plant (bottom). The arrows indicate the 8.5 elution time which represents the expected mannitol peak based on elution time of the mannitol standard. The peak at 8.5 is absent in the wild type cucumber plant. Transgenic cucumber plants expressing the M6PR gene have elevated level of mannitol in leaves compared to the non-transgenic controls Mannitol accumulation in the To progeny was tested using gas chromatography (Figure 4-2). All six putative PCR-verified transgenic cucumber plant showed the expected mannitol peak (elution time 8.5 min) indicating marmitol presence. Mannitol also was observed occasionally in the T1 M6PR plants (Figure 4-3; for example, M3), the M6PR T1 plants did not consistently show mannitol accumulation (Table 4-2). Mannitol accumulation in four families was further investigated. To test whether the inconsistency in observed mannitol accumulation was due to catabolism of mannitol or translocation into sink tissue, leaf and sink tissues were collected from greenhouse growing cucumber plants at two different time points; early in the day (before 10:00 am) and later in the day (between 3-5:00 pm). Mannitol was not detectable in majority of samples, and no obvious pattern could be determine relative to old vs. young tissue, leaves vs. flowers or time of the day. Due to the inconsistence in mannitol accumulation of the T1 progeny, this project was not pursued further. 128 counts 1 FlD1A.(o1oacm1eom.uoh1) g ‘ .-. Control 35000 1 1 “°°°° 1 4 25000 4 1 20000 '1 15000 § 10000 1'9 L o I '1“ T"'+T T 1'" I' '1 n 1‘ 1‘ '“1 75 10 12.5 15 17.5 20 22.5 25 27.5 .... K 3 counts F101 A. mmsmm ..N; M6PR M3 35000 3 1 30000 1 1 25000 '1 ID .. 1 *0. 1 C 20000 1 g 1 E 15000 1 .1 10000 1 l 4 8 ._ g ‘ v- 0 h - a ..- § 3 ... .. o 5 _ Q Q q M—a—A—n—e—M— o 1 1'0 13 23 23 m' Figure 4-3. Gas chromatography for mannitol analysis in wild type (top) and T1 plant fi'om family M3 (bottom). The arrows indicate the 8.5 elution time which represent the expected mannitol peak based on elution time of the mannitol standard. The first peak which represent mannitol is absent in the wild type plants and indicated by the black arrow. 129 Table 4-2. Mannitol accumulation in segregating Tl-M6PR plants. Different tissues were analyzed (young leaves, fully expanded leaves and female flowers) were collected at two time points (before 10:00 am and after 3:00 pm) from M6PR plants growing in the greenhouse. Number of plants showing mannitol accumulation PCR $16121 positive Young leaves fully expanded leaves Flowers . - Early Late Early Late Early Late Ml 12:4 2:14 1:14 0:14 1:14 3:10 1:13 M3 9:5 0:14 0:14 4:10 2:12 7:4 2:12 M4 14:3 4:13 2:15 6:11 2:15 0:17 -—--- M5 10:8 2:16 1:17 3:15 6:12 1:16 3:15 130 References Abebe T, Guenzi AC, Martin B, Cushman J C (2003) Tolerance of mannitol-accumulating transgenic wheat to water stress and salinity. Plant Physiol 131: 1748-1755. Bohnert HJ, Jensen RG (1996) Strategies for engineering water-stress tolerance in plants. Trends Biotech 14: 89-97. Bohnert HJ, Su H, Shen B (1999) Molecular mechanisms of salinity tolerance. In: Molecular responses to cold, drought, heat and salt stress in higher plants. Landes Company 29-62. Everard JD, Cantini C, Grumet R, Plummer J, Loescher WH (1997) Molecular cloning of mannose—6-phosphate reductase and its developmental expression in celery. Plant Physiol 113: 1427-1435. Everard JD, Gucci R, Kann SC, Flore J A, Loescher WH (1994) Gas exchange and carbon partitioning in the leaves of celery (Apium graveolens L.) at various levels of root zone salinity. Plant Physiol 106: 281-292. Greenway H, Munns R (1980) Mechanisms of salt tolerance in non-halophytes. Annul Rev Plant Physiol 31: 149-190. Halliwell B, Grootveld, M and Gutteridge, JMC (1988). Methods for the measurement of hydroxyl radicals in biochemical systems: deoxyribose degradation and aromatic hydroxylation. Methods Biochem Anal 33: 59—90. Hayashi H, Alia, Mustardy L, Deshnium P, Ida M, Murata N (1997) Transformation of Arabidopsis thaliana with the codA gene for choline oxidase; accumulation of glycinebetaine and enhanced tolerance to salt and cold stress. Plant J 12: 133-142. Holmstrom KO, Somersalo S, Mandal A, Palva TE, Welin B (2000). Improved tolerance to salinity and low temperature in transgenic tobacco producing glycine betaine. J Exptl Bot 51: 177-185. Holmstrom KO, Welin B, Mandal A, Kristiansdottir I, Teeri TH, Lamark T, Strom AR, Palva ET (1994) Production of the Escherichia coil betaine-aldehyde dehydrogenase, an enzyme required for the synthesis of the osmoprotectant glycine betaine, in transgenic plants. Plant J. 6: 749-758. Ingram, J Bartels D (1996) The molecular basis of dehydration tolerance in plants. Annul Rev Plant Physiol Plant Mol Biol 47: 377-403. J ia GX, Zhu ZQ, Chang FQ, Li YX (2002) Transformation of tomato with the BADH gene fi'om Atriplex improves salt tolerance. Plant Cell Rep 21: 141-146. 131 Karakas B, OziasAkins P, Stushnoff C, Suefferheld M, Rieger M (1997) Salinity and drought tolerance of mannitol-accumulating transgenic tobacco. Plant Cell Env 20: 609-616. Kishitani S, Takanami T, Suzuki M, Oikawa M, Yokoi S, Ishitani M, Alvarez-Nakase AM, Takabe T, Takabe T (2000) Compatibility of glycinebetaine in rice plants: evaluation using rice plants with a gene for peroxisomal betaine aldehyde dehydrogenase from barley. Plant Cell Env. 23: 107-114. Kishor PBR, Hog Z, Miao GH, Hu CAA, Verma DPS (1995) Overexpression of Al- pyrroline-S-carboxylate synthetase increases proline production and confers osmotolerance in transgenic plants. Plant Physiol 108: 1387-1394. Liu Y, Wang GY, Liu JJ, Peng XX, Xie YJ, Dai JR, Guo SW, Zhang FS (1999) Transfer of E. coli gutD gene into maize and regeneration of salt-tolerant transgenic plants. Sci China Ser 42: 90-95. Loescher WH (1987) Physiology and metabolism of sugar alcohols in higher plants. Physiol Plant 70: 553-557. Loescher WH, Everard JD (1997) Sugar alcohol metabolism in sinks and sources. In E Zamski, AA Schaffer, eds, Photoassimilate Distribution in Plant and Crops: Source- Sink Relationships. Marcel Dekker, New York, pp 185—207. Loescher WH, Everard JD (2000) Regulation of sugar alcohol biosynthesis. In RC Leegood, TD Sharkey, S von Caemmerer, eds, Photosynthesis: Physiology and Metabolism. Kluwer Academic Publisher, Dordrecht, The Netherlands, pp 275—299. Loescher WH, Tyson RH, Everard JD, Redgwell RJ, Bieleski RL (1992) Mannitol synthesis in higher plants: evidence for the role and characterization of a NADPH- dependent mannose 6-phosphate reductase. Plant Physiol 98: 1396-1402. Prabhavathi V, Yadav J S, Kumar PA, Rajam MV (2002) Abiotic stress tolerance in transgenic eggplant (Solanum melongena L.) by introduction of bacterial mannitol phosphor-dehydrogenase gene. Mol Breeding 9: 137—147. Ronde J A, Spreeth MH, Cress WA (2000). Effect of antisense L- A-l-pyrroline-S- carboxylate reductase in transgenic soybean plants subjected to osmotic and drought stress. Plant Grow Reg 32: 13-26. Sakamoto A, Murata N (2000) Genetic engineering of glycinebetaine synthesis in plants: current status and implications for enhancement of stress tolerance. J Exp Bot 51: 81-88. 132 Sakamoto A, Murata N (2000) Genetic engineering of glycinebetaine synthesis in plants: current status and implications for enhancement of stress tolerance. J Exptl Bot 51: 81-88. Shen B, Jensen R, Bohnert, HJ (1997) Increased resistance to oxidative stress in transgenic plants by targeting mannitol biosynthesis to chloroplasts. Plant Physiol 113:1177—1183. Smirnoff N, Cumbes Q] (1989) Hydroxyl radical scavenging activity of compatible solutes. Phytochemistry 28: 1057-1060. Stoop J MH, Pharr DM (1994) Mannitol metabolism in celery stressed by excess macronutrients. Plant Physiol 106: 503—51 1. Stoop J MH, Williamson JD, Pharr DM (1996) Mannitol metabolism in plants: a method for coping with stress. Trends Plant Sci 1: 139—144. Stoop, J .M.H. and Pharr, D.M. (1994) Mannitol metabolism in celery stressed by excess macronutrients. Plant Physiol 106: 503-51 1. Tarczynski MC, Jensen RG, Bohnert HJ (1993) Stress protection of transgenic tobacco by production of the osmolyte mannitol. Science 259: 508-510. Tarczynski MC, Jensen RG; Bohnert HJ (1992) Expression of bacterial thD gene in transgenic tobacco leads to production and accumulation of mannitol. Proc Natl Acad Sci USA 89: 2600-2604. Tilahun A, Guenzi AC, Martin B, Cushman J C (2003) Tolerance of mannitol- accumulating transgenic wheat to water stress and salinity. Plant Physiol 13121748- 1755. Zhang J, Nguyen HT, Blum A (1999) Genetic analysis of osmotic adjustment in crop plants. J Exp Bot 50: 291-302. Zhifang G, Loescher WH (2003) Expression of celery mannose-6-phosphate reductase in Arabidopsis thaliana enhances salt tolerance and induces biosynthesis of both mannitol and glucosyl-mannitol dimer. Plant Cell Environ 26: 275-283. Zhu BC, Su J, Chan MC, Verma DPS, Fan YL, Wu R (1998) Overexpression of a A1- pyrroline-S-carboxylate synthetase gene and analysis of tolerance to water- and salt- stress in transgenic rice. Plant Sci 139: 41-48. Zhu J -K, Hasegawa PM, Bressan RA (1997). Molecular aspects of osmotic stress in plants. Crit Rev Plant Sci 16: 253-277. 133 Conclusions and future work In the current work, I produced several lines of transgenic cucumber plants expressing the Arabidopsis thaliana transcriptional regulators CBF1/DREBb and CBF3/DREBa genes under the control of the constitutive CaMV35S promoter (Chapter 2). Greenhouse experiments indicated that CBF1 and CBF 3 cucumber plants showed array of adaptive responses to the imposed salinity and drought stress, including the accumulation of compatible solutes (e. g., proline and soluble sugar), less reduction in photochemical efficiency as measured by chlorophyll fluorescence, less growth reduction, and accumulation of higher above ground dry matter compared to non- transgenic controls. These observations are similar to other reports that introduced CBF genes into tomato, tobacco and rice; however, CBF homologues have been identified in all of those species. This raises the question of whether cucumber has CBF homologues. I tried to look for CBF homologues in cucumber by following more than one approach; for example, trying southern hybridization using CBF] and CBF 3 as probes with low stringency washing conditions, designing degenerate primers for conserved common motifs among the known CBF genes from different species, and finally using the cucumber genomic library to screen for CBF homologues. The fact that I could not clone any homologues does not role out the possibility that cucumber may have CBF homologues. One thing to pursue in the future would be expand efforts to identify CBF- homologs from cucumber and ask if they are true functional homologues of the Arabidopsis-CBF gene family, by testing for expression, phenotypic and physiological changes in Arabidopsis plants overexpressing these homologues. It would also be 134 beneficial to understand how these CBF-homologues are regulated in response to different stresses. Despite the enhanced stress resistance in tomato, expression of predicted COR homologs did not increase in response to overexpression of CBF. Given the fact that there are marked differences between Arabidopsis and tomato, in gene expression in response to CBF-overexpression, and the current failure to observe CBF-regulation of predicted tomato orthologs to Arabidopsis CBF-regulon genes, further suggests that responses may differ in heterologous systems, which in turn may make finding of CBF- target genes in cucumber more challenging. One way to answer this would be by using microarray analysis to compare gene transcription profiles of CBF-expressing and nontransgenic control cucumber plants. The one drawback in this approach is that the microarray chip would be from a different species (Arabidopsis). Another alternative could be the generation of subtractive libraries from CBF-transgenic and nontransgenic cucumber plants. I Salinity treatment (50 and 100 mM NaCl) caused a marked increase in soluble sugar and proline accumulation in CBF-plants beyond what would be expected from an additive effect of CBF and salt stress. The same trend was also observed in some of the work on CBF-expressing plants, although it was less pronounced than in our study. The lack of differences in CBF levels in the transgenic lines in the presence or absence of salinity indicated the presence of other levels of regulation downstream of CBF in cucumber. Recently, Cook et al. (2004) showed metabolic changes in Arabidopsis in response to cold acclimation and overexpression of CBF genes. The idea that CBF genes, which encode transcriptional regulators, could activate waves of responses makes it of 135 interest to monitor metabolomic changes in cucumber plants under different stress conditions. Thus questions here would be what are the metabolic changes in cucumber plants in the presence or absence of salinity? What are the changes in the CBF-cucumber line in the presence or absence of salinity? How would these differ at different stress levels? Informations from these experiments might also help identify possible CBF target genes. These types of studies could provide insights into the secondary regulation in plants and how this is relate to gene expression profiling. We also reported for the first time that expression of the CBF system in cucumber plants caused an increase in ion composition of cucumber roots. Cucumber roots accumulated higher levels of K+ and Ca++ ions in the presence or absence of salinity stress, suggesting that excess Na+ ions did not prevent the accumulation of K" to higher levels in the roots of CBF-cucumber plants and maybe an increased selectivity in K+ uptake or due to differential expression of different transporters and channels responsible for ion uptake or movement through the plant. The fact that CaH/ATPase transporter was one of the upregulated genes in CBF-overexpressing plants raises questions about ion transport regulation in CBF-expressing plants. Calcium has been shown to maintain or enhance the selective absorption of potassium over sodium in plants under salt stress (Epstein, 1998). This may be through activation of Na+/H+ antiporters at the plant plasma membrane or maybe by activating some high K selective channels. Another question is whether changes in Na+/1C ratio is an indirect side effect of accumulation of compatible solutes in plant tissue? If so then the phenomenon would be similar to that in previous reports (Chapter III). One way to 136 answer these questions would be by using tissue specific microarray analyses, or more comprehensive analysis of root tissue may lead to identification of additional genes as possible targets for CBF/DREB. Another possibility would be to generate subtractive libraries from roots of CBF-cucumber plants and compare it with nontransgenic control cucumber plants. Finally, when we first started this work we had few questions to ask; can CBF confer dehydration stress tolerance in a heterologous system? What are the limitation of the transgenic cucumber plants in response to salinity and drought? What are the possible effects of the CBF gene on growth and yield data of cucumber plants in the presence or absence of salinity stress? Although my current work was an attempt to answer many of these questions, our findings also raised new questions: 1- What are the metabolic changes in cucumber plants in the presence or absence of salinity and drought? How does it differ in CBF cucumber lines in the presence or absence of dehydration stresses? 2- How would these differ at different stress levels? Can these metabolic changes be used to predict possible CBF-target genes in cucumber? 3- What cause the accumulation of higher K+ levels in roots of CBF-cucumber plants; it is due to changes in K+ uptake, K+ selectivity, or in ion movement in plant tissue? 4- The changes in the Na+/K+ ratio an indirect side effect of accumulation of compatible solutes in plant tissue? In summary, results demonstrate the potential usefulness of the CBF/DREB system in providing elevated levels of dehydration stress tolerance in salt sensitive species such 137 as cucumber and also raise new questions about the function of CBF in regulating stress tolerance related phenotypes. 138 Appendix A: Proline, soluble sugar and mannitol protocols Proline analysis One leaf (3-4 cm diameters) was collected from each plant, and freeze-dried for 48 hrs and ground to assure uniformity. Each sample was split into two equal weights for proline and sugar analysis. Proline analysis was carried out using the ninhydrin reagent procedure described by Troll and Lindsley, (1955). Tissue samples (0.1-0.2 g dry weight) were placed in 13x100 mm disposable glass tubes (Cat No. 47729-572, VWR international, West Chester, PA) and proline extracted fiom dried tissue by overnight soaking in 3.0 m1 of distilled H20 followed by heating at 80 C° for 30 min. The supernatant was then transferred into a fresh 13x 100 glass tube and the heating step was repeated with fresh 3 ml of water H20 for 30 min. Tissues and debris were removed from the samples by centrifuging for 1 min at 2500 g. the supernatant was then transferred into fresh glass tubes. 400 pl of the extract solution was transferred to microcentrifuge tubes containing 400 pl of the ninhydrin reagent (prepared as described below) and 400 pl of glacial acetic acid. One gram of Ninhydren reagent (Sigma-Aldrich, St. Louis, MO) was prepared by dissolved in 16 m1 of concentrated phosphoric acid in the dark; then bringing the volume to 40.0 ml with glacial acetic acid (MERCK KGaA, Darmstadt, Germany). The sample tubes were closed and placed in a boiling water bath for an hour. The tubes were removed from the water bath, and cooled to room temperature for 10-15 min. The samples were transferred into a new 13x100 mm disposable glass tube, and 2.5 m1 of toluene added to each tube. The tubes were periodically gently shaken by hand for a few second to allow the red color to dissolve in the toluene phase. Absorbance of the toluene phase was measured spectophotometrically at A515. The standard curve was prepared by 139 dissolving 10 mg of LProline (Sigma-Aldrich, St. Louis, MO) into 10 ddeO m1 (lmg/ml, lOOOppm). l, 2, 4, 6, 8, and 10 pl of the standard was added into a fresh microcentrifuge tube then brought up to 400 pl with water and processed in as described above. Total soluble sugar analysis Total soluble sugar analysis was performed by the phenol/sulfuric acid method as described by Dubois et al., (1956). Three ml of 80% ethanol was added to 0.1-0.2 g of the dried samples in 13x100 mm glass tubes and incubated overnight at 4 C°. The next day, total soluble sugars were extracted by boiling the samples for 30 min; the supernatant was transferred into a fresh 13x 100 glass tube and the heating step was repeated with fresh 3 ml of 80% ethanol. Tissues and debris were removed from the samples by centrifuging for 1 min at 2500 g. The supernatant was then transferred into fresh glass tubes followed by adding 3 m1 of chloroform (MERCK KGaA, Darmstadt, Germany); samples were mixed by inverting several times and centrifuged for 5 min at 2500 g. The clear upper aqueous phase was transferred into a fresh 13x100 mm disposable glass tubes. 100 pl of the extract solution was added to 1.9 ml HzO, followed by 50 pl 80% phenol (Sigma- Aldrich, St. Louis, MO); 5 ml of concentrated sulfuric acid was quickly added. Samples were lefi for 10 min before gently vortexing at low speed. Absorbance of the mix was measured at A485. 140 Mannitolggalvsis One leaf (6-8 cm diameter; approximate fresh weight 1.5 g) was collected from To and T1 plants, freeze-dried for 48 hr, ground and placed in 13x100 mm disposable glass tubes (Cat No. 47729-5 72, VWR international, West Chester, PA). To total extract soluble carbohydrates, samples were boiled in 3 ml 80% ethanol for 30 min and the supernatant transferred into fresh 13x100 mm glass tubes. The pellet was re-extracted by boiling again with 2 ml 80% ethanol for another 30 min. Afier the extraction, 5 ml water was added to each sample, and samples were transferred into 15 ml disposable Corning tubes (Corning Inc., Corning, NY). Five ml of chloroform (MERCK KGaA, Darmstadt, Germany) was added to each sample; the samples were capped and shaken by hand and then centrifuged for 5 min at 2500 g. The clear upper aqueous phase was transferred into a fresh 13x100 mm disposable glass tubes and placed in heating blocks at 50 C° in the fume-hood until the samples were completely dried (approximately 2-3 hrs.). One ml of derivatization solution [pyridine (Sigma-Aldrich, St. Louis, MO) kept over NaOH pellets (Sigma-Aldrich, St. Louis, MO)] was added to each tube. The tubes were capped tightly and vortexed until all the dried sugar crystal were dissolved. The tubes were placed in a heating block at 70-80 C° for an hour; samples were vortexed every 30 min, then transferred to a tube rack and allowed to cool to room temperature for 10-20 min. One ml of room temperature hexamethyldisilazane (HMDS) (Sigma-Aldrich, St. Louis, MO) was added to each sample. Samples were allowed to stand for 20-30 seconds before adding 100 pl of trifluoroacetic acid (TFAA) (Sigma-Aldrich, St. Louis, MO), followed by brief vortexing. Samples were incubated for an hour at room temperature, then 1.0 m1 of clear upper aqueous phase was transferred into GC Vials (Alltech Corporation, Deerfield, IL), 141 capped tightly and placed in the auto-sampler tray for GC-Series H analysis (Hewlett Packard 5890-11 gas chromatography, Palo Alto, CA) which was fitted with a DB-17 capillary column (J &W Scientific, Folsom, Ca, USA). Standards were previously prepared using 0.1 g of fructose, glucose, sucrose, raffinose, inositol, all mixed together and dissolved in 100 ml HzO (this should give 1000 ppm). Different volumes of the standard mix were transferred into 13x100 mm glass tubes (1.0, 0.5, 0.1, 0.05, 0.025 and 0.01 ml) and were placed in heating blocks at 50 C° in the fume-hood until the samples were completely dried (approximately 2-3 hrs.) followed by the same derivatization steps that was described earlier. 142 1113111311111111111le