Skip to main content
Cellular and Molecular Life Sciences: CMLS logoLink to Cellular and Molecular Life Sciences: CMLS
. 2014 Oct 22;72(4):673–689. doi: 10.1007/s00018-014-1767-0

General mechanisms of drought response and their application in drought resistance improvement in plants

Yujie Fang 1,2, Lizhong Xiong 1,
PMCID: PMC11113132  PMID: 25336153

Abstract

Plants often encounter unfavorable environmental conditions because of their sessile lifestyle. These adverse factors greatly affect the geographic distribution of plants, as well as their growth and productivity. Drought stress is one of the premier limitations to global agricultural production due to the complexity of the water-limiting environment and changing climate. Plants have evolved a series of mechanisms at the morphological, physiological, biochemical, cellular, and molecular levels to overcome water deficit or drought stress conditions. The drought resistance of plants can be divided into four basic types-drought avoidance, drought tolerance, drought escape, and drought recovery. Various drought-related traits, including root traits, leaf traits, osmotic adjustment capabilities, water potential, ABA content, and stability of the cell membrane, have been used as indicators to evaluate the drought resistance of plants. In the last decade, scientists have investigated the genetic and molecular mechanisms of drought resistance to enhance the drought resistance of various crops, and significant progress has been made with regard to drought avoidance and drought tolerance. With increasing knowledge to comprehensively decipher the complicated mechanisms of drought resistance in model plants, it still remains an enormous challenge to develop water-saving and drought-resistant crops to cope with the water shortage and increasing demand for food production in the future.

Keywords: Drought resistance breeding, Drought tolerance, Drought avoidance, Gene function, Genetic engineering

Introduction

The growth, development, and reproduction of plants require sufficient water. Approximately, one-third of the Earth’s land area is arid and semi-arid, while periodically unexpected climatic droughts often occur in most of the other land areas. Water scarcity can be lethal to plants and lead to enormous social problems and economic losses. The development of the modern science and technology revolution, on one hand, has largely increased our capabilities in exploring the natural resources which have dramatically improved human life; on the other hand, the continuously growing world population, together with widespread water pollution and unpredictable climatic change, further aggravates the shortage of water resources [13]. It is estimated that total global water consumption has tripled over the last 50 years according to the United Nation Water Development Report (WWDR-3; http://www.unesco.org/new/en/natural-sciences/environment/water/wwap/wwdr/wwdr3-2009/), and the agricultural water demand which accounts for more than two-thirds of the global water use dramatically increased in the past few decades (http://ceowatermandate.org/business-case/global-water-trends/increasing-water-demand/) [46]. Drought not only inflicts heavy agricultural production losses, but also contributes to ecological damage, land desertification, and soil erosion. Therefore, water scarcity has been considered as an urgent global and environmental problem.

Water shortage has attracted great concern and stimulated more and more research inputs on the fundamental science of the drought resistance of plants and the application of the acquired knowledge for developing water-saving and drought-resistant crops. In recent years, a number of countries and international organizations have launched research projects on exploring the drought-resistance and water-saving mechanisms of plants to identify key genes or tools for the improvement of plant drought resistance. The Consultative Group on International Agricultural Research (CGIAR) organized the Generation Challenge Programme (GCP) as a time-bound 10-year programme in 2003 to improve crop breeding for drought-prone conditions (http://www.generationcp.org). The CGIAR has initiated programs for reducing rural poverty, improving human nutrition and health, increasing global food security, and sustainably managing of natural resources through agricultural research (http://www.cgiar.org/our-research/). Early in 1970s, the CGIAR was well-known for the contribution to the Green Revolution [7]. In recent years, CGIAR scientists integrated the progress in multiple disciplines (genome sequencing of major crop species, functional genomics, genetic markers, and genetic engineering, etc.) and application of information and communication technologies (ICTs) in their research [8]. They also adopted open source approaches for accelerating crop improvement research and increased the application of improved varieties of important crops such as rice, wheat, maize, sorghum, millet, potatoes, and so forth [8, 9]. In 2005, European and African countries initiated the Improving Water Use Efficiency in Mediterranean agriculture (WUEMED) project for technical improvements in the field of agriculture and the management of water resources under drought conditions (http://www.distagenomics.unibo.it/wuemed/index.html). Many countries including China have also provided special funds to support the basic study and genetic improvement of drought-resistance in crops.

For more than two decades, scientists have conducted vast amounts of research, including studies on morphological traits, and the structural, physiological, biochemical, and molecular regulation of above-/below-ground parts, to unravel the mechanisms of drought responses in plants. In recent years, with the rapid development of the theory and technology of modern biology (such as molecular genetics, genomics, proteomics, and metabolomics), researchers have made great strides in elucidating the biochemical, genetic, and signaling networks involved in plant drought responses, however, the underlying sophisticated mechanisms that differentiate resistance from susceptibility within a species, especially for crops, remain largely unclear. When plants are confronted with drought or more commonly water deficit, they respond to the stress by integrating very diverse responses and adaptive mechanisms at the morphological, physiological, and molecular levels, but different plant species or genotypes within a species often have large variations in the utilization of these mechanisms. At the early stage of water deficit, plants usually have an ability to absorb water from the underground efficiently through root system, partially close stomata to reduce water loss from transpiration, and alter the metabolism to match with the available carbon resource [10, 11]. As stress conditions increase, some osmolytes such as prolines, soluble sugars, spermines, and betaine accumulate in plant cells to maintain the cell turgor pressure [12]. Variations in the activities of numerous oxidation-protective enzymes such superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), and glutathione reductase (GR) are also frequently observed in drought-stressed plants [10, 13]. More general responses of plants to drought stress include the altered expression of numerous genes, such as those related to stress signal transduction, and the transcription and regulation of thousands of functional proteins, which collectively contribute to the molecular control of drought resistance.

In this review, we mainly focus on the major morphological and physiological traits associated with the mechanisms underlying drought resistance in plants, and then we briefly summarize the advances in characterization of the effective genes for drought response in plants. Finally, we provide an overview on the progress on drought resistance improvement in crops. This review aims to provide general mechanisms of plant drought response and their potential application in drought resistance breeding to a broad spectrum of scientific communities, especially for plant biologists and breeders who are not working in this field.

Dissecting drought resistance in plants

With respect to adaptation to water-deficit, plants can be roughly divided into three types: xerophytes (mainly in arid areas), mesophytes (mainly in semi-arid and sub-humid areas), and hydrophytes which are mainly distributed in environments with sufficient moisture or water [14]. The majority of model plants and crops belong to the mesophytes. Meteorologically, drought refers to the condition caused by a sufficiently long period of dry weather which results in plant injury [15, 16]. May and Milthorpe defined drought resistance as “the ability of plants to grow satisfactorily when exposed to water deficits” [17, 18]. Later, crop breeders defined drought resistance in terms of “relative yield of genotypes” or “the ability of a crop plant to produce its economic product with minimum loss in a water-deficit environment relative to the water-constraint-free management” [16, 19]. Plants have evolved several different types of drought resistance strategies which allow them to adapt to specific habitats for the benefit of their growth and development. Drought resistance is defined as the integrated capability of plants in response and adaption to the harsh environment caused by drought stress conditions. This capability is a sophisticated trait and is related to the adaptations at different levels, ranging from plant morphology and anatomical structures to physiological and biochemical reactions [20].

Plant drought resistance involves four major mechanisms: drought avoidance (DA) (or “shoot dehydration avoidance” in some literature), drought tolerance (DT), drought escape (DE), and drought recovery [15, 2126]. Among the four components of drought resistance, DA and DT are the two major mechanisms for drought resistance conferred by plants [25]. DA is the capability of plants to maintain fundamental normal physiological processes under mild or moderate drought stress conditions by adjusting certain morphological structures or growth rates to avoid the negative effects caused by drought stress [23, 27, 28]. DA is principally characterized by the maintenance of high plant water potentials in the presence of a water shortage [16, 23]. Plants generally adopt three strategies to accomplish DA: (1) reducing water loss via rapid stomatal closure, leaf rolling [29], and increasing wax accumulation on the leaf surface in many plant species such as alfalfa, tobacco, and rice [3032]; (2) enhancing the water uptake ability through a well-developed root system (especially increased rooting depth, rooting density or root/shoot ratio) and enhancing the water storage abilities in specific organs (such as fleshy water-storing tissues of cacti, the truck of candlenut, earthnuts or tubers of some plants, etc.) [29, 33, 34]; (3) accelerating or decelerating the conversion from vegetative growth to reproductive growth to avoid complete abortion at the severe drought stress stage [16, 23]. DT refers to the ability of plants to sustain a certain level of physiological activities under severe drought stress conditions through the regulation of thousands of genes and series of metabolic pathways to reduce or repair the resulting stress damage [16, 23, 35]. Plants commonly exert protoplasmic tolerance by increasing osmoregulatory molecules in the cells to maintain the cell turgor pressure, and adjusting the activities of cell defense enzymes to reduce the accumulation of hazardous substances. The term DE refers to natural or artificial adjustment of the growth period, life cycle, or planting time of plants to prevent the growing season from encountering local seasonal or climatic drought [16, 36]. Farmers usually choose crop varieties with short life cycles which complete their life cycle by avoiding the seasonal drought stress in agricultural production. Drought recovery refers to the plant capability to resume growth and gain yield (for crops) after exposure to severe drought stress which causes a complete loss of turgor pressure and leaf dehydration [23].

The drought resistance of plants is quite complex. For a given plant species, plants often combine different categories of mechanisms to confer drought resistance at different developmental stages. At a particular developmental stage, plant drought resistance is associated with a series of events (such as stomatal movement, photosynthesis, cell osmotic regulation, synthesis of protective macromolecules and antioxidants, etc.) in every conceivable facet at the morphological, physiological, and molecular levels. In addition, natural drought stress is dynamic and unpredictable. Therefore, it is rather difficult to comprehensively and accurately evaluate the overall drought resistance of a given plant species. Levit [22] pointed out that the determination of drought resistance is much more difficult than that of other stress resistances. Researchers often use a specific trait or several combined indicators to assess plant drought resistance depending on the purposes of their study. These indicator traits can be generally classified into three types: DA-related, DT-related, and integrative indicators [11, 23].

The indicators associated with DA are usually related to the moisture maintenance and water uptake and use efficiency. DT-associated indicators mainly cover physiological parameters related to osmotic adjustment (OA) (such as osmotic potential, proline content, soluble sugar content, ABA content, etc.) and alleviation of drought damage (such as the activities of protective enzymes and chlorophyll content, etc.). Some complex traits relevant to biomass or economic yield under stress conditions are also used to evaluate the drought resistance of crops in agricultural production. These traits include fresh or dry weight, survival rate, stay-green capability, seed-setting rate, spikelet fertility, grain weight, and so on. These traits are more meaningful and effective in breeding for drought resistance, although they have seldom been used to unveil the mechanisms of drought resistance at the physiological and molecular levels.

Drought avoidance-associated mechanisms

Modulation of root system architecture

Plants constantly obtain water (and nutrients as well) from the soil through their roots. Hence, the root system plays a critical role in response to water deficit stress. Some plants have the robust ability to increase root growth at the early stage of drought stress to absorb the water in deep soil [11]. A positive correlation has been found between penetration ability of roots and the degree of drought resistance in Phaseolus acutifolius [37]. The length, weight, volume, and density of plant roots were also reported to be associated with the drought resistance in crops [3841]. Nevertheless, other research showed a lack of perceptible association between root traits with single plant and plot yield under reproductive stage stress in rice [42]. In dry areas, woody plant seedlings have vertical roots with ten times the length of the above ground height [43]. With this extensive root system and rooting depth, plants are able to maintain a higher water potential and a longer duration of transpiration under drought conditions, which provides further advantages for their growth and development [44]. Rooting depth, volume, and distribution are mainly influenced by the depth and range of soil moisture. In cases of soil water deficit, plants dynamically adapt and modify their root system architecture by changing their root growth in diverse manners depending on the species [4547]. It is evident that severe soil water deficit can reduce root elongation, branching, and the formation of the cambium layer, and root tips of plants growing in arid soil become suberified [48]. Root growth is also influenced by the water or nutrient status of the aerial portion of the plant [49]. Increased root/shoot ratios are often observed under water stress conditions [50, 51]. For a long time, the root/shoot ratio has been used as a criterion to describe the plant capacity for drought resistance [5255].

Leaf traits

The morphological and physiological responses of leaves to drought stress are crucial to reduce water loss and promote water use efficiency. When plants sense severe water deficiency, their leaves droop or roll because of the loss of cell turgor pressure and this phenomenon is called wilting [56]. High rates of transpiration temporarily induce an insufficient water supply, and some plants wilt around the middle of the day, while a decline in transpiration relieves the water deficit at night, and the rolled leaves slowly re-expand. Wilting is a passive movement of plant leaves to prevent excess water consumption under drought stress conditions. Apart from this, some plants can actively adjust the orientation of leaf blades to keep them parallel to the direction of incident solar radiation by rolling. The phototropic movement of plant leaves can regulate the interception of solar radiation [57, 58]. When the leaf blades expand in a direction perpendicular to the direction of solar radiation, the single blade receives the largest amount of radiation, while deviation from the vertical direction will reduce the amount of radiation. Upright leaves under water stress conditions will receive less radiation, resulting in reduced water loss and better overall water status, indicating that erect leaves are an effective mechanism of DA [5961]. Leaf rolling is a common response of plants to water deficit, and it is a mechanism to reduce water consumption when water stress is present [57]. Leaf rolling is a drought-adaptive trait induced by turgor pressure, and osmotic adjustment can delay leaf rolling [62]. Both passive and active leaf movements have a role in reducing incident solar radiation and thus reducing leaf surface temperature, protecting plants from excess water loss.

Plants with increased drought resistance often have xeromorphic structures such as smaller and thicker leaves, more epidermal trichomes, smaller and denser stomata, a thicker cuticle epidermis, thicker palisade tissue, a higher ratio of palisade to spongy parenchyma thickness, and a more developed vascular bundle sheath [63]. Leaf epidermal trichomes reduce plant transpiration under intense light conditions and help to reflect light [64]. Lipids accumulate in the epidermis to form wax and increase the reflectivity of sunlight to prevent plants from excessive transpiration and high leaf surface temperatures [65]. The fortified sclerenchymas (mechanical tissues) can reduce the damage from wilting and protect the plants from direct light irradiation [66]. Palisade tissues and vascular bundles ensure transportation and retention of water and nutrients [67]. These features effectively reduce excess water loss and enhance the water-holding ability to avoid damage from exposure to drought stress conditions.

Stomata are pores which formed in the leaves of terrestrial plants during a long-term evolutionary process. As the vital organs for exchanging gas and water between the plant and the external environment, stomata play critical roles in the activities of plant life by ensuring maximum absorption of CO2 for photosynthesis, and meanwhile controlling the optimal transpiration. The stomatal density and aperture are closely related to the plant drought resistance [68]. Guard cells which in pairs surround the stomatal pores are extremely sensitive to environmental conditions. After receiving the environmental stimuli, changes of water potential and turgor movement in guard cells control the opening and closure of stomata, and further regulate pivotal physiological processes in plants such as transpiration and photosynthesis. Under water-limiting conditions, the function of stomata in adjusting transpiration is particularly important. Stomata of water-saving plants (which avoid dehydration by reducing transpiration) are sensitive to water deficit, and the leaf stomata close before the leaf water status approaches wilting, thereby exerting a DA function. Stomata respond to water stress mainly in two ways: (1) as a direct response to the air humidity in which guard cells and adjacent epidermal cells directly evaporate moisture to induce stomatal closure and prevent leaf water deficit, and (2) stomata respond to the water potential changes in the leaves in which stomata close when the leaf water potential falls below a certain threshold.

Stomatal movement is controlled by osmotic potential changes in guard cells, and K+ is one of the major ions affecting the osmotic potential in guard cells [69]. The influx or efflux of K+ from the guard cells plays a critical role in changing the osmotic potential and the turgor pressure, which leads to the opening or closure of stomata [70]. The outward- and inward-K+ channels on the plasma membrane of guard cells are vital to the transmembrane transport of K+. There are various specific proteins such as substrate-binding proteins (including ABA-binding proteins, acetylcholine receptors, GTP-binding proteins, and light receptors), pumps, and channels on the plasma membrane of guard cells which are involved in the control of stomatal movement [7174]. These proteins are crucial for receiving and transducing stress signals in guard cells, and constitute the fundamental basis of the opening and closure of stomata under drought stress conditions. Furthermore, numerous studies suggest that stomatal movement is also controlled by the ABA signaling which is triggered by roots in drying soil profile [75, 76]. Stomatal closure is an important strategy to support water conservation by plants during drought stress. Thus, better understanding of the stomatal movement mechanisms is crucial for optimizing water use efficiency related to drought resistance in plants.

Besides, other leaf-associated traits such as epidermal hairs, cuticular wax, along with leaf water potential, relative water content, water loss rate, and canopy temperature are also used as criteria for appraisal of DA [11].

Photosynthesis

Plant production is mainly determined by photosynthesis, and plant photosynthesis is governed mainly by stomata for CO2/water exchange and photosynthetic activity in mesophyll cells. Water stress affects not only the light reactions, but also the assimilation efficiency of the dark reactions, thereby reducing the contents of the photosynthetic products [10, 7779]. Plants have evolved three photosynthetic pathways including C3, C4, and crassulacean acid metabolism (CAM) to assimilate atmospheric CO2. Generally, plants utilizing C4 and CAM photosynthetic mechanisms can better adapt to drought-prone climate [80]. C3 plants open their stomata during the day for CO2 absorption and fixation and close their stomata at night. This mechanism is deficient when C3 plants confront water limitation because it does not retain moisture under drought stress conditions. C4 plants have evolved a metabolic pump to concentrate CO2 in the bundle sheath cells, and perform the fixation of CO2 in mesophyll cells and the bundle sheath cells separately [81]. This particular mechanism contributes to higher water use efficiency than that of C3 plants and provides more chance for C4 plants to survive in arid areas [81]. In the CAM cycle photosynthetic pathway, plants open their stomata for CO2 absorption and fixation at night, and close their stomata to reduce transpiration water loss during the day. Therefore, CAM metabolism can dramatically increase the water use efficiency and is proposed to be a plastic photosynthetic adaptation to extremely arid environments [82]. When challenged by water stress, some plants considered as facultative CAM species are capable of switching their photosynthetic pathway from the C3 cycle to the CAM cycle mode [83, 84]. Researchers have found that the key enzyme in the CAM metabolic pathway, phosphoenolpyruvate carboxylase, is transcriptionally regulated by water stress conditions [85].

Drought tolerance-associated mechanisms

Osmotic adjustment

Plants accumulate a variety of organic and inorganic substances (such as sugars, polyols, amino acids, alkaloids, and inorganic ions) to increase their concentration in the cytochylema, reduce the osmotic potential, and improve cell water retention in response to water stress. This phenomenon is defined as osmotic adjustment (OA) [86, 87], a significant strategy for plant drought tolerance. OA has been documented to sustain cell structure and photosynthesis at low water potentials, and to delay leaf senescence and death and improve root growth as water deficits become severe [26]. Substances currently known to be involved in OA encompass several types of organic compounds such as mannitol, proline, glycine, betaine, trehalose, fructan, inositol, and inorganic ions [8892]. These organic substances can regulate the plasma osmotic potential, and protect the enzymes and plasma membranes. In addition, changes in the ion and water channels control the export and import of ions and moisture for plant cells, which also contributes to OA. The inorganic ions mainly regulate the osmotic potential of the vacuole to maintain turgor pressure; however, a high concentration of inorganic ions is likely to cause metabolism disorders in plant cells. Hydration of the membrane and the surface layer of intracellular proteins are also important for stabilizing the structures of biological macromolecules. The OA substances are capable of stabilizing the surface-bound water of and sustaining the spatial structure of biological macromolecules [9396].

Proline has very strong hydration abilities. Its hydrophobic part is able to bind to proteins while its hydrophilic part is able to bind to water molecules, allowing proteins to access more water to enhance their solubility and to prevent protein denaturation from dehydration under osmotic stress conditions [97]. Proline acts not only as a cytoplastic protective agent for enzymes and cell structure, but also for adjusting the redox potential and reducing cell acidity [98100]. Trehalose is a reducing disaccharide. Under drought stress conditions, the intercellular trehalose content increases rapidly to block the transformation of the phospholipids bilayer membrane from the liquid crystal state to the solid state, and stabilize the structure of proteins, nucleic acids, and other biological macromolecules [88]. Betaine is a metabolic intermediate belonging to the water-soluble alkaloid quaternary ammonium compounds, and functions as one of the nontoxic OA substances in higher plants [101]. Betaine also helps to stabilize the structures and activities of photosynthesis, including protective enzymes, and helps in the maintenance of membrane integrity against the pervasive damage under drought stress conditions [102104].

Late embryogenesis abundant (LEA) proteins, aquaporins (AQP), and molecular chaperones have also been demonstrated with crucial roles in OA in plant cells. LEA proteins are formed during the process of seed development. They are low-molecular weight proteins which are usually between 10 and 30 kDa, and consist of basic amino acids rich in lysine, glycine, and serine, and commonly lack cysteine and tyrosine residues [105]. LEAs are hyper-hydrophilic proteins with extremely high thermal stability, and they can remain in the aqueous state even under boiling conditions. LEA proteins can protect biological macromolecules, redirect intracellular water distribution, bind to inorganic ions to avoid the damage attributed to the accumulation of high concentrations of ions under drought stress conditions, prevent excessive dehydration of plant tissues, and control the expression of other genes by binding to nucleic acids [106]. Aquaporins (AQP) are channel proteins which are responsible for water transportation. AQPs form selective water transport channels in plant cells and regulate the rapid transmembrane transport of moisture during the processes of seed germination, cell elongation, stomatal movements, and abiotic stress responses [107, 108]. AQPs are divided into three groups according to their subcellular location: plasma-membrane-intrinsic proteins (PIPs), tonoplast-intrinsic proteins (TIPs), and nodulin-26-like major-intrinsic proteins (NLMs) [109]. AQPs mediate passive water transport along the osmotic pressure gradient inside and outside the membrane, and they are capable of regulating the moisture balance for the entire plant under drought stress conditions. This is achieved by maintaining the water potential balance between the xylem parenchyma cells and the transpiration current, regulating water transport across cells and tissues, as well as adjusting the cell turgor and volume [108]. Further characterization of key genes controlling the biosynthesis and metabolism pathways of these OA substances may provide useful candidate genes for improving drought tolerance by transgenic approach.

Antioxidant defense systems

Oxidative stress commonly occurs along with drought stress. Antioxidant defense system is one of the drought response mechanisms. Aerobic metabolism which provides energy for plant growth and development is often accompanied by the generation of reactive oxygen species (ROS) as by-products such as 1O2, H2O2, O·−2, and OH•. Under normal circumstances, the intracellular generation and removal of ROS is under dynamic equilibrium. When plants suffer from exposure to drought stress conditions, the dynamic equilibrium is broken and the excessive accumulation of ROS injures cells, and the oxidative deterioration may ultimately lead to cell death [110]. The membrane phospholipids and fatty acids which are sensitive to the over-accumulation of ROS are damaged, resulting in the peroxidation of membrane lipids [111]. Under ROS stress, the spatial configurations of various membrane proteins or enzymes are disturbed, leading to increased membrane permeability and ion leakage, chlorophyll destruction, metabolism perturbations, and even severe injury or death of plants [112].

Plants produce ROS in chloroplasts, peroxisomes, mitochondria, endoplasmic reticulum, plasma membrane, and the cell wall due to the imbalance between the generation and utilization of electrons under drought stress conditions [113]. ROS attack the most sensitive biological macromolecules in plant cells to induce lipid peroxidation, protein carbonylation, and DNA damage, and impair their functions to result in a catastrophic cascade of events [114]. To protect cells against the deleterious effects of excessive ROS, plants have evolved a series of sophisticated enzymatic and non-enzymatic antioxidant defense mechanisms to maintain the homeostasis of the intracellular redox state. The protective enzymes include superoxide dismutase (SOD), catalase (CAT), ascorbate peroxidase (APX), glutathione peroxidase (GPX), glutathione reductase (GR), glutathione S-transferase (GST), dehydroascorbate reductase (DHAR), monodehydroascorbate reductase (MDAR), thioredoxin peroxidase (TPX), alternative oxidase (AOX), peroxiredoxin (PrxR/POD), etc. [115, 116]. SOD, the H2O–H2O cycle, the AsA-GSH cycle, GPX, and CAT cooperatively build up major ROS-scavenging pathways [113, 117, 118]. The balances among the SOD, APX, and CAT activities are pivotal to maintaining the H2O2 homeostasis in plants. The non-enzymatic antioxidant system is comprised of several reducing substances, such as ascorbic acid (AsA), glutathione (GSH), carotenoids (CAR), α-tocopherol (vitamin E), cytochrome f (Cytf), flavanones, anthocyanins, and so on [112, 115, 119]. To date, quite a few instructive cases have been reported to achieve enhanced drought tolerance by eliminating the excessive accumulation of ROS in cells or promoting the ROS-mediated signal transduction in plants [110, 120, 121]. Further characterization of genes controlling the antioxidant defense system may also provide useful candidate genes for improving drought tolerance by transgenic approach.

Phytohormones and chlorophyll content

Endogenous phytohormones are also responsive to environmental cues, and several phytohormones have indispensable roles in coordinating the responses to drought stress through rather complicated crosstalk mechanisms [122, 123]. ABA was regarded as the phytohormone which is most closely related to drought stress responses in plants. Soil water deficit is perceived as a distant signal by root cells which then triggers a huge increase in the de novo synthesis of ABA [124]. ABA is transported mainly to leaves as an intercellular messenger and recognized by guard cells which trigger stomatal closure via intracellular signal transduction, and weakening the metabolic activities related to plant growth [125127]. The influences of ABA have multiple effects on the drought response, encompassing the regulation of stomatal closure, channel activities in guard cells, transcriptional levels of the calmodulin protein, and the expression of some ABA-responsive genes [128131]. As a key chemical messenger of drought signals, ABA is located at the central position in the regulatory network of stomatal closure [132, 133]. ABA has also been shown to promote the synthesis and accumulation of proline by affecting the activity of pyrroline-5-carboxylate reductase (P5CR) and enhancing the synthetic activity of glutamoyl-phosphate [134136]. The expression of Δ1-pyrroline-5-carboxylate synthetase (P5CS) can be induced by drought and ABA [137, 138]. ABA is also the most important signaling molecule for drought signal transduction. There are at least three relatively independent signal transduction pathways which function in drought stress responses in plants. Two pathways are ABA-dependent, and the other pathway is ABA-independent [139]. To date, ABA signaling in the drought response has been well-illustrated. Drought stress induces the biosynthesis of intracellular ABA which activates the corresponding transcription factors, and then promotes the expression of downstream drought-related genes [140].

Thylakoid membranes are the structural foundation for light absorption, transmission, and transformation in chloroplasts. Many types of pigment–protein complexes in charge of light absorption, transmission, and transformation are located on the thylakoid membranes. Under drought stress conditions, the membrane system in plant cells including the membrane structure associated with photosynthesis is destroyed due to the deficit of moisture, nutrients, and energy, resulting in the disruption of physiological processes. These factors are likely to directly or indirectly affect the chlorophyll content. It has been reported that the chlorophyll content significantly declines with decreasing soil water content, and the decrease of total chlorophyll content is mainly due to the decrease of chlorophyll a [141144]. The plants which can maintain higher chlorophyll contents under water stress conditions are considered to use light energy more efficiently, and therefore are thought to have increased drought resistance.

Molecular basis of drought resistance

With the growing knowledge of modern biology and power of biotechnical tools, the mystery of the molecular mechanism of plant drought resistance has been gradually unveiled in the past decade. The response of plants to drought stress is a complex process involving many genes and signaling pathways. The transcript abundance of a large number of genes with diverse functions alters under drought stress conditions [145]. The external drought stimuli are perceived and captured by sensors on the membrane which have not been well-characterized, and then the signals are passed down through multiple signal transduction pathways, resulting in the expression of drought-responsive genes and drought adaptation [146]. A variety of secondary messengers (such as Ca2+, ROS, phosphoglycerol, ABA, and diacylglycerol) and transcriptional regulators play significant roles in various signal transmitting pathways [147]. Emerging evidence suggests that drought-inducible genes are under complex governance, including the transcriptional cascades. The expression products of drought-responsive genes are mainly divided into three categories: (1) the proteins involved in the signaling cascades and transcriptional regulation (such as protein kinase, protein phosphatase, and transcription factors); (2) functional proteins that protect the cellular membranes and other proteins (such as late embryogenesis abundant protein, antioxidants, and osmotin); and (3) proteins associated with the uptake and transport of water and ions (such as aquaporins and sugar transporters) [139, 146, 148, 149].

Much effort has been justifiably dedicated to identifying the key components in elucidating the genetic and molecular bases of drought resistance by quantitative trait locus (QTL) mapping and cloning, mutant screening, expression profiling, and verification of candidate genes. Many QTLs for root traits, leaf traits, and physiological traits related to drought resistance have been mapped, and some of them have been validated for plant performance under drought stress conditions [11]. The research has largely accelerated the progress of understanding the complexity of genetic and environmental interactions under drought stress conditions, and provided valuable references for drought resistance improvement. Meanwhile, thousands of genes responsive to drought stress have been identified by Affymetrix GeneChip technology and RNA sequencing [150153].

Numerous candidate genes derived from the mutant screening or expression profiling studies have been further characterized for their functions in drought response. Regulatory proteins have been proven to play crucial roles in the responses of plants to drought stress conditions. The phosphorylation and dephosphorylation of proteins in plants are common events induced by drought stress conditions. Several types of kinases such as calcium-dependent protein kinases (CDPKs), CBL (calcineurin B-like) interacting protein kinase (CIPK), mitogen-activated protein kinases (MAPKs), and sucrose non-fermenting protein (SNF1)-related kinase 2 (SnRK2) have been reported to participate in drought response. An Arabidopsis the CDPK gene CPK10 was reported to mediate stomatal movement via the ABA and Ca2+ signaling pathways in response to drought stress conditions [154]. OsCDPK7 was shown to positively regulate drought and salt stress tolerance in rice [155]. OsCIPK23 was found to be induced by various abiotic stress conditions and phytohormones, and OsCIPK23 RNAi suppression transgenic plants were more sensitive to drought stress conditions [156]. MAP kinase signaling cascades in plants, such as OsMPK5 and the MAPK kinase kinase (MAPKKK) gene DSM1 in rice, have also been implicated with roles in regulating drought resistance [157159]. A SnRK2 gene SnRK2C is thought to confer drought tolerance by regulating the expression of stress-responsive genes in Arabidopsis [160]. Transcription factors are another indispensable group of regulatory proteins which modulate gene expression to respond to drought stress at the transcriptional level. Many members from diverse transcription factor families such as APETALA2/Ethylene-responsive element binding protein (AP2/EREBP), basic leucine zipper (bZIP), MYB, NAM-ATAF1/2-CUC2 (NAC), and zinc finger have been reported to be involved in the drought responses. An AP2/EREBP domain transcription factor SHN was reported to confer enhanced drought tolerance by activating wax biosynthesis to alter cuticle properties in Arabidopsis [161]. Some dehydration-responsive element-binding factors (DREB) with the AP2 domain such as OsDREBs and ARAG1 have also been reported to participate in the drought responses in rice [162, 163]. The ABA-responsive element-binding proteins/factors (AREBs/ABFs) belong to a subfamily of bZIP transcription factors, and they are well-known for their roles in ABRE-dependent ABA signaling under drought stress conditions. Three members from the AREB/ABF subfamily, AREB1, AREB2, and ABF3, are up-regulated by ABA and their full activation requires ABA [164]. The areb1 areb2 abf3 triple mutant displays enhanced ABA resistance and reduced drought tolerance, indicating that the three factors coordinately govern ABRE-dependent gene expression under water stress conditions [164]. Several rice bZIP proteins such as OsbZIP23 and the constitutive active form of OsbZIP46 were also characterized to have a high potential in drought resistance improvement in rice [165, 166]. Cominelli et al. [167] reported a R2R3-MYB gene AtMYB60 which is specifically expressed in guard cells to modulate the physiological responses of stomata. The T-DNA insertion mutant atmyb60-1 showed constitutively increased stomatal closure and was more resistant to water stress. OsMYB2, with a multiple stress-induced expression pattern, was found to be related to the increased tolerance to salt, cold, and dehydration stresses in transgenic plants by regulating the accumulated amount of H2O2 and malondialdehyde and the expression of genes encoding proline syhthase and transporters in rice [168]. NAC transcription factors are involved in almost every aspect of plant activities throughout the entire plant life cycle. A stress-induced rice NAC gene SNAC1 was proven to have high potential in engineering drought resistance improvement in rice, wheat, and cotton [169171]. Over-expression of OsNAC10 driven by a root-specific promoter RCc3 remarkably enlarged the root diameter in transgenic rice and consequently resulted in enhanced drought tolerance at the reproductive stage, and improved grain yield in the field under both normal and drought stress conditions [172]. Zinc finger factors belong to another super family of transcription factors. Constitutive over-expression of a Cys2/His2 (C2H2)-type zinc finger protein encoding the ZPT2-3 gene in transgenic petunia improved tolerance to dehydration stress [173]. Another zinc finer protein DST was documented to function as a negative regulator of drought and salt tolerance by controlling the genes involved in H2O2-mediated stomatal movement in rice [174]. Reduced stomatal density and increased stomatal closure was observed in the dst mutant, and the mutant showed enhanced tolerance to both drought and salt stress conditions [174].

Phytohormones are documented to have different roles in the response to drought stress. Among the various phytohormones, ABA is the one that is most closely related to drought stress [76]. In recent years, the ABA biosynthesis and metabolic pathways in higher plants have been clarified through the biochemical and genetic analyses of ABA biosynthesis-deficient and ABA-insensitive mutants [175, 176]. ABA is synthesized de novo from an indirect carotenoids (C40) pathway in which zeaxanthin epoxidase (ZEP), 9-cis-epoxycarotenoid dioxygenase (NCED), and ABA-aldehyde oxidase (AAO) are key enzymes for the ABA biosynthesis pathway in higher plants [177]. Notable progress has been made to improve drought tolerance in plants by manipulating the ABA biosynthesis and metabolism pathways. Over-expression of AtZEP in Arabidopsis led to more vigorous growth of transgenic plants under drought treatment conditions compared to wild-type plants [178]. Over-expression of the AtNCED3 and PvNCED1 genes resulted in significantly more accumulation of endogenous ABA in transgenic plants and enhanced drought resistance [179, 180]. The LOS5/ABA3 gene encodes a molybdenum cofactor (Moco) sulfurase which produces a cofactor (the sulfurylated MoCo) of AAO, which functions in the final step of ABA biosynthesis in Arabidopsis [181]. A los5 mutant showed more severe drought-induced damage compared to the wild-type, and over-expression of AtLOS5 increased ABA accumulation and drought resistance in both transgenic tobacco and cotton [181183]. Xiao et al. [184] overexpressed several well-characterized genes including NCED2 and LOS5 in rice for testing drought resistance under field conditions, and the results showed that some of these genes might be potential candidates for drought resistance breeding.

ROS generation and scavenging has been recognized to influence the drought resistance of plants [110, 185]. Some ROS genes have been utilized to engineer drought-resistant plants. Over-expression of a pea manganese superoxide dismutase gene (MnSOD) under the control of an oxidative stress-inducible promoter SWPA2 in rice chloroplasts enhanced the drought tolerance of the transgenic rice [186]. Cytosolic APX1 was revealed to play an important role in response to a combination of drought and heat stress conditions [187]. ATGPX3, a gene encoding an Arabidopsis thaliana glutathione peroxidase, was found to act as a scavenger and an oxidative signal transducer in ABA and drought stress signaling, playing a distinctive role in H2O2 homeostasis [188]. Several other genes such as OsSKIPa and OsSRO1c have been reported to modulate drought resistance by controlling ROS metabolism and regulating ROS homeostasis in plants [121, 189].

LEA proteins are expressed at specific stages of late embryonic development and play critical roles in desiccation tolerance by capturing water, stabilizing and protecting the structure and function of proteins and membranes, as well as acting as molecular chaperons and hydrophilic solutes to protect cells from the damage of water stress [190]. Xiao et al. [191] identified OsLEA3-1 and tested the performance of transgenic rice over-expressing OsLEA3-1 under drought stress conditions in the field. The results indicated that the transgenic families exhibited higher grain yields than the wild-type under drought stress conditions. Transporter proteins have been demonstrated to function as regulators of drought resistance [192]. Transgenic maize co-transformed with betA and TsVP genes showed significantly improved drought resistance [193]. A guard cell plasma membrane ABCC-type ABC transporter gene AtMRP4 was investigated for its role in stomatal aperture regulation in Arabidopsis [194]. The atmrp4 mutant plants exhibited larger stomatal apertures than the wild-type plants, and consequently also exhibited reduced drought resistance [194]. AQPs are involved in OA and regulation of water movement across cellular membranes in plants. Transgenic rice expressing an AQP gene RWC3 controlled by a stress-inducible SWPA2 promoter displayed enhanced root osmotic hydraulic conductivity, leaf potential, and relative cumulative transpiration under PEG treatment [195]. Over-expression of another AQP gene PgTIP1 increased biomass production and water-deficit tolerance in Arabidopsis [196].

Accumulation of osmoprotectants (such as trehalose, betaine, and proline) is a widespread adaptive strategy for plants to retain the water potential, cell turgor, and membrane stability to avoid drought stress-induced damage [197]. TPS1 encodes trehalose-6-phosphate synthase, which is a key enzyme for trehalose biosynthesis in yeast. Yeo et al. [198] reported that constitutively over-expressing TPS1 in potato significantly improved drought resistance. Over-expression of a betaine aldehyde dehydrogenase encoding gene BADH from spinach under the control of a stress-induced rd29 promoter enhanced the tolerance of transgenic potato to drought and salt stress conditions [199]. P5CS functions as a rate-limiting enzyme in proline biosynthesis. Kishor et al. [200] successfully raised proline production/yield and enhanced osmotolerance in transgenic plants by over-expression of the mothbean P5CS gene in tobacco.

Plenty of other metabolic-related proteins have also been demonstrated to be closely associated with the drought response in plants. The rice Drought-Induced Wax Accumulation 1 (DWA1) and Glossy 1 (GL1-2) genes were proposed to be involved in leaf wax deposition, and both the dwa1 and osgl1-2 mutants were impaired in cuticular wax accumulation under drought stress conditions and exhibited increased drought sensitivity [32, 201]. Ornithine δ-aminotransferase (δ-OAT) is a proline and arginine metabolism-related enzyme, and over-expression of OsOAT which is a target gene of SNAC2 resulted in improved resistance to drought, osmotic, and oxidative stresses, indicating that OsOAT conferred drought tolerance by enhancing pre-accumulation and ROS-scavenging in rice [202].

Genetic improvement of drought resistance in crops

Beside the importance of understanding the general mechanisms of drought response in plants, the exploration of genetic variations in diverse traits related to drought resistance within species is also important for genetic improvement of drought resistance in crops. Here, we briefly summarize some successful efforts or promising approaches related to drought resistance breeding in the last two decades. Delayed silking is a commonly observed phenomenon in maize when it is subjected to water stress at flowering stage [203]. The anthesis silking interval (ASI), a trait of intermediate heritability that generally exhibits a negative correlation with grain yield under drought conditions, provided a valuable and effective selection target for drought resistance in maize [204]. Root angle dramatically influences root distribution in the soil and is thus considered as a proxy for selecting deeper roots varieties to increase drought resistance [205]. Kato et al. [206] documented that root growth angle is useful for preliminary estimation of genotypic variation in root distribution in rice. Mace et al. [207] identified a putative association between root angle QTL and grain yield in sorghum. It has been demonstrated that modulation of the partitioning of metabolically important carbohydrates is a satisfying strategy for breeding drought-resistant cultivars in some crops [208, 209]. Storage and remobilisation of water soluble carbohydrates (WSC) in stems and leaves is an essential contributor to grain filling, especially under the water-limiting conditions [210]. Phenotype analysis was performed in three mapping populations for WSC mass per unit area (WSC-A) in wheat, indicating that WSC-A plays a significant role in assuring stable yield and grain size [211, 212]. Transpiration efficiency has been also reported to have a strong correlation with water use efficiency and yield under drought conditions, and therefore used as a selection criteria for drought resistance identification [213]. In maize, the stay-green trait was found to be closely related to grain yield, while in sorghum it is associated with maintenance of water status [214, 215]. Four QTLs which control stay-green have been mapped in sorghum [215]. Improved symbiotic nitrogen (N2) fixation was evidenced with effect in increasing the soybean yield by enhancing dry-matter accumulation under drought [216]. Sinclair et al. [217] screened one hundred lines of filial generation derived from a cross between Jackson and KS4895, and identified two soybean lines with decreased sensitivity of N2 fixation to water deficit and increased yield. In bermudagrass, rhizome production was proven to be an important trait which was closely linked to high productivity and drought resistance [218, 219]. Zhou et al. [219] comprehensively investigated the association between rhizomes and drought resistance by assessing eighteen bermudagrass genotypes (Cynodon spp.) from four climatic zones under drought. Canopy temperature depression (CTD), which can be measured by thermal imaging, was indicated to be a significant predictor of yield performance under drought stress in bread wheat [220, 221], while GA sensitivity was claimed to be implicated in drought resistance in wheat [222]. Probing and establishment of the morpho-physiological criteria which could precisely assess the degree of drought resistance is a crucial aspect in studying drought resistance in plants.

Conventional drought resistance breeding largely depends on genetic variation of the DR-related traits in a specific species, and it is a labor-intensive and time-consuming process [223, 224]. Genetic transformation or engineering provides an alternative or complementary approach for developing desired traits more efficiently [224]. Numerous genes, as introduced in the previous section of this review and other reviews [11], have shown to be effective in improving drought resistance based on experiments mostly conducted in green house. However, compared to other simple traits, genetic improvement of drought resistance in crops by the transgenic approach is still facing many challenges. The understanding of the general molecular mechanisms of drought responses and the genetic variation in a given plant species is rather limited, which is a major obstacle for the molecular breeding of drought resistance and/or the selection of appropriate candidate genes for transformation. Furthermore, some other stresses such as high temperature are usually co-incidental to drought stress. Therefore, it is foreseeable that the outcome of transferring one single gene to increase drought resistance is unsatisfactory in most instances. Present transgenic studies commonly use constitutive promoters including CaMV35S, ubiquitin, and actin to make overexpression constructs [191, 198, 225]. These promoters efficiently cause high expression levels of the target genes for improving drought resistance, but constitutive expression of some target genes frequently lead to adverse side effects on plant growth, development, or production under normal conditions, some of which are likely owing to the unnecessary consumption of energy or metabolic disturbance [226, 227].

To promote the application of transgenic technologies to drought resistance breeding, the following considerations may be taken into account in the future. First, multidisciplinary approaches may be adopted to decipher the elaborate genetic control of drought resistance mechanisms, which is also essential to unveil the genetic and molecular mechanisms of drought response and the crosstalk between different stresses. Second, compared to the single gene transformation, multi-gene transformation strategy that combines several major functional or regulatory genes or a series of genes in a signaling cascade contributing to drought resistance may be more reasonable or promising for improving drought resistance in plants. Third, attention should be paid to exploiting and identifying suitable promoters which could ensure stable and accurate transcription of genes in an anticipated pattern under specific stress conditions [223]. Drought-inducible promoters generally maintain low expression levels of the driven genes under unstressed conditions and activate the genes under water scarcity to minimize the potential adverse effect on normal growth [227]. Using promoters with specific temporal-spatial pattern to control the location and time of expression should also be considered. Furthermore, combination of traditional breeding (such as cross and/or recurrent backcross of wild relatives and elite cultivars) and transgenic technology may provide a great potential to pyramid the desired traits especially for improving resistance to multiple stresses.

Conclusion and perspective

Drought resistance in plants is an extremely complex trait. Plants adopt a suite of strategies encompassing morphological, physiological, cellular, and molecular changes to survive under drought stress conditions. Different plants adapt to drought stress via diverse and integrated mechanisms. In recent years, researchers have identified numerous genes related to the drought response of plants, and have made substantial progress in the genetic improvement of drought resistance. However, our understanding of the regulatory networks and the crosstalk between the signaling pathways under drought-prone conditions is still fragmentary. Current knowledge on the molecular basis of drought resistance in plants is merely the tip of the iceberg. We have a long way to go to comprehensively and intensively elucidate the significant biological functions involved in drought response, and place them in their accurate locations in the refined regulatory network to obtain a clear picture.

Drought is a major constraint of agricultural development. It is an enormous challenge to integrate modern genetics, genomics, molecular biology, and proteomics, along with the metabolomics approaches to investigate the crosstalk and fill the vacancies in the regulatory networks of plant drought response, and better apply the fundamental theories and experience into practice for plant drought resistance breeding. Integration of the molecular approaches, morpho-physiological analysis, and conventional breeding strategies will substantially accelerate the progress of cultivating drought-resistant plants for agricultural production.

Acknowledgments

Research on drought resistance in the authors’ laboratory has been supported by grants from the National Program for Basic Research of China (2012CB114305), the National Program on High Technology Development (2012AA10A303), the National Natural Science Foundation (31271316), and the Priority Academic Program Development of Jiangsu Higher Education Institutions.

References

  • 1.Trenberth KE, Dai A, van der Schrier G, Jones PD, Barichivich J, Briffa KR, Sheffield J. Global warming and changes in drought. Nat Clim Change. 2014;4:17–22. [Google Scholar]
  • 2.Woodward A, Smith KR, Campbell-Lendrum D, Chadee DD, Honda Y, Liu Q, Olwoch J, Revich B, Sauerborn R, Chafe Z, Confalonieri U, Haines A. Climate change and health: on the latest IPCC report. Lancet. 2014;383:1185–1189. doi: 10.1016/S0140-6736(14)60576-6. [DOI] [PubMed] [Google Scholar]
  • 3.Hussain M, Mumtaz S. Climate change and managing water crisis: Pakistan’s perspective. Rev Environ Health. 2014;29:71–77. doi: 10.1515/reveh-2014-0020. [DOI] [PubMed] [Google Scholar]
  • 4.Rost S, Gerten D, Bondeau A, Lucht W, Rohwer J, Schaphoff S. Agricultural green and blue water consumption and its influence on the global water system. Water Resour Res. 2008;44:W09405. [Google Scholar]
  • 5.Wada Y, Van Beek L, Bierkens MF. Modelling global water stress of the recent past: on the relative importance of trends in water demand and climate variability. Hydrol Earth Syst Sci. 2011;15:3785–3808. [Google Scholar]
  • 6.Döll P. Vulnerability to the impact of climate change on renewable groundwater resources: a global-scale assessment. Environ Res Lett. 2009;4:035006. [Google Scholar]
  • 7.Mann C. Reseeding the green revolution. Science. 1997;277:1038–1043. [Google Scholar]
  • 8.Renkow M, Byerlee D. The impacts of CGIAR research: a review of recent evidence. Food Pol. 2010;35:391–402. [Google Scholar]
  • 9.Byerlee D, Dubin HJ. Crop improvement in the CGIAR as a global success story of open access and international collaboration. Int J Commons. 2009;4:452–480. [Google Scholar]
  • 10.Reddy AR, Chaitanya KV, Vivekanandan M. Drought-induced responses of photosynthesis and antioxidant metabolism in higher plants. J Plant Physiol. 2004;161:1189–1202. doi: 10.1016/j.jplph.2004.01.013. [DOI] [PubMed] [Google Scholar]
  • 11.Hu H, Xiong L. Genetic engineering and breeding of drought-resistant crops. Annu Rev Plant Biol. 2014;65:715–741. doi: 10.1146/annurev-arplant-050213-040000. [DOI] [PubMed] [Google Scholar]
  • 12.Seki M, Umezawa T, Urano K, Shinozaki K. Regulatory metabolic networks in drought stress responses. Curr Opin Plant Biol. 2007;10:296–302. doi: 10.1016/j.pbi.2007.04.014. [DOI] [PubMed] [Google Scholar]
  • 13.Goswami A, Banerjee R, Raha S. Drought resistance in rice seedlings conferred by seed priming: role of the anti-oxidant defense mechanisms. Protoplasma. 2013;250:1115–1129. doi: 10.1007/s00709-013-0487-x. [DOI] [PubMed] [Google Scholar]
  • 14.Warming E, Balfour IB, Groom P, Vahl M. Oecology of plants. London: Oxford University Press; 1909. [Google Scholar]
  • 15.Kneebone W, Mancino C, Kopec D. Water requirements and irrigation. Turfgrass. 1992 [Google Scholar]
  • 16.Mitra J. Genetics and genetic improvement of drought resistance in crop plants. Curr Sci (Bangalore) 2001;80:758–763. [Google Scholar]
  • 17.May L, Milthorpe F. Drought resistance of crop plants. Proc Field Crop Abstr. 1962;15:171–179. [Google Scholar]
  • 18.Clarke JM, DePauw RM, Townley-Smith TF. Evaluation of methods for quantification of drought tolerance in wheat. Crop Sci. 1992;32:723–728. [Google Scholar]
  • 19.Fukai S, Cooper M. Development of drought-resistant cultivars using physiomorphological traits in rice. Field Crops Res. 1995;40:67–86. [Google Scholar]
  • 20.Blum A (2002) Drought stress and its impact. Available via DIALOG. http://www.plantstress.com/Articles/drought_i/drought_i.htm. Accessed 12 Oct 2014
  • 21.Lawlor DW. Genetic engineering to improve plant performance under drought: physiological evaluation of achievements, limitations, and possibilities. J Exp Bot. 2013;64:83–108. doi: 10.1093/jxb/ers326. [DOI] [PubMed] [Google Scholar]
  • 22.Levitt J. Responses of plants to environmental stresses, vol II. Water, radiation, salt, and other stresses. London: Academic Press; 1980. [Google Scholar]
  • 23.Luo LJ. Breeding for water-saving and drought-resistance rice (WDR) in China. J Exp Bot. 2010;61:3509–3517. doi: 10.1093/jxb/erq185. [DOI] [PubMed] [Google Scholar]
  • 24.Turner NC. Drought resistance and adaptation to water deficits in crop plants. In: Mussel H, Staples RC, editors. Stress physiology in crop plants. New York: Wiley; 1979. [Google Scholar]
  • 25.Yue B, Xue W, Xiong L, Yu X, Luo L, Cui K, Jin D, Xing Y, Zhang Q. Genetic basis of drought resistance at reproductive stage in rice: separation of drought tolerance from drought avoidance. Genetics. 2006;172:1213–1228. doi: 10.1534/genetics.105.045062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Turner NC, Wright GC, Siddique K. Adaptation of grain legumes (pulses) to water-limited environments. Adv Agron. 2001;71:194–233. [Google Scholar]
  • 27.Hall A, Schulze E. Drought effects on transpiration and leaf water status of cowpea in controlled environments. Funct Plant Biol. 1980;7:141–147. [Google Scholar]
  • 28.Blum A. Drought resistance, water-use efficiency, and yield potential-are they compatible, dissonant, or mutually exclusive? Aust J Agric Res. 2005;56:1159–1168. [Google Scholar]
  • 29.Tardieu F. Plant response to environmental conditions: assessing potential production, water demand, and negative effects of water deficit. Front Physiol. 2013;4:17. doi: 10.3389/fphys.2013.00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Zhang JY, Broeckling CD, Blancaflor EB, Sledge MK, Sumner LW, Wang ZY. Overexpression of WXP1, a putative Medicago truncatula AP2 domain-containing transcription factor gene, increases cuticular wax accumulation and enhances drought tolerance in transgenic alfalfa (Medicago sativa) Plant J. 2005;42:689–707. doi: 10.1111/j.1365-313X.2005.02405.x. [DOI] [PubMed] [Google Scholar]
  • 31.Cameron KD, Teece MA, Smart LB. Increased accumulation of cuticular wax and expression of lipid transfer protein in response to periodic drying events in leaves of tree tobacco. Plant Physiol. 2006;140:176–183. doi: 10.1104/pp.105.069724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Islam MA, Du H, Ning J, Ye H, Xiong L. Characterization of Glossy1-homologous genes in rice involved in leaf wax accumulation and drought resistance. Plant Mol Biol. 2009;70:443–456. doi: 10.1007/s11103-009-9483-0. [DOI] [PubMed] [Google Scholar]
  • 33.Sawidis T, Kalyva S, Delivopoulos S. The root-tuber anatomy of Asphodelus aestivus . Flora Morphol Distrib Funct Ecol Plants. 2005;200:332–338. [Google Scholar]
  • 34.Ogburn R, Edwards EJ. The ecological water-use strategies of succulent plants. Adv Bot Res. 2010;55:179–225. [Google Scholar]
  • 35.Passioura J. Drought and drought tolerance. Drought tolerance in higher plants: genetical, physiological and molecular biological analysis. Netherlands: Springer; 1997. [Google Scholar]
  • 36.Manavalan LP, Guttikonda SK, Tran L-SP, Nguyen HT. Physiological and molecular approaches to improve drought resistance in soybean. Plant Cell Physiol. 2009;50:1260–1276. doi: 10.1093/pcp/pcp082. [DOI] [PubMed] [Google Scholar]
  • 37.Mohamed MF, Keutgen N, Tawfika AA, Noga G. Dehydration-avoidance responses of tepary bean lines differing in drought resistance. J Plant Physiol. 2002;159:31–38. [Google Scholar]
  • 38.Price A, Steele K, Moore B, Jones R. Upland rice grown in soil-filled chambers and exposed to contrasting water-deficit regimes: II. Mapping quantitative trait loci for root morphology and distribution. Field Crops Res. 2002;76:25–43. [Google Scholar]
  • 39.Johnson W, Jackson L, Ochoa O, Van Wijk R, Peleman J, Clair DS, Michelmore R. Lettuce, a shallow-rooted crop, and Lactuca serriola, its wild progenitor, differ at QTL determining root architecture and deep soil water exploitation. Theor Appl Genet. 2000;101:1066–1073. [Google Scholar]
  • 40.Hammer GL, Dong Z, McLean G, Doherty A, Messina C, Schussler J, Zinselmeier C, Paszkiewicz S, Cooper M. Can changes in canopy and/or root system architecture explain historical maize yield trends in the US corn belt? Crop Sci. 2009;49:299–312. [Google Scholar]
  • 41.Forster B, Thomas W, Chloupek O. Genetic controls of barley root systems and their associations with plant performance. Aspects Appl Biol. 2005;73:199–204. [Google Scholar]
  • 42.Subashri M, Robin S, Vinod K, Rajeswari S, Mohanasundaram K, Raveendran T. Trait identification and QTL validation for reproductive stage drought resistance in rice using selective genotyping of near flowering RILs. Euphytica. 2009;166:291–305. [Google Scholar]
  • 43.Larcher W. Physiological plant ecology: ecophysiology and stress physiology of functional groups. Netherlands: Springer; 2003. [Google Scholar]
  • 44.Dixon R, Wright G, Behrns G, Teskey R, Hinckley T. Water deficits and root growth of ectomycorrhizal white oak seedlings. Can J For Res. 1980;10:545–548. [Google Scholar]
  • 45.Den Herder G, Van Isterdael G, Beeckman T, De Smet I. The roots of a new green revolution. Trends Plant Sci. 2010;15:600–607. doi: 10.1016/j.tplants.2010.08.009. [DOI] [PubMed] [Google Scholar]
  • 46.Smith S, De Smet I. Root system architecture: insights from Arabidopsis and cereal crops. Philos Trans R Soc Lond B Biol Sci. 2012;367:1441–1452. doi: 10.1098/rstb.2011.0234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Malamy J. Intrinsic and environmental response pathways that regulate root system architecture. Plant Cell Environ. 2005;28:67–77. doi: 10.1111/j.1365-3040.2005.01306.x. [DOI] [PubMed] [Google Scholar]
  • 48.Kramer PJ. Plant and soil water relationships: a modern synthesis. Plant and soil water relationships: a modern synthesis. New York: McGraw-Hill Book Company; 1969. [Google Scholar]
  • 49.Price AH, Cairns JE, Horton P, Jones HG, Griffiths H. Linking drought-resistance mechanisms to drought avoidance in upland rice using a QTL approach: progress and new opportunities to integrate stomatal and mesophyll responses. J Exp Bot. 2002;53:989–1004. doi: 10.1093/jexbot/53.371.989. [DOI] [PubMed] [Google Scholar]
  • 50.Wu Y, Cosgrove DJ. Adaptation of roots to low water potentials by changes in cell wall extensibility and cell wall proteins. J Exp Bot. 2000;51:1543–1553. doi: 10.1093/jexbot/51.350.1543. [DOI] [PubMed] [Google Scholar]
  • 51.Fulda S, Mikkat S, Stegmann H, Horn R. Physiology and proteomics of drought stress acclimation in sunflower (Helianthus annuus L.) Plant Biol (Stuttg) 2011;13:632–642. doi: 10.1111/j.1438-8677.2010.00426.x. [DOI] [PubMed] [Google Scholar]
  • 52.Pallardy SG. Physiology of woody plants. London: Academic Press; 2010. [Google Scholar]
  • 53.Tavakol E, Pakniyat H. Evaluation of some drought resistance criteria at seedling stage in wheat (Triticum aestivum L.) cultivars. Pak J Biol Sci. 2007;10:1113–1117. doi: 10.3923/pjbs.2007.1113.1117. [DOI] [PubMed] [Google Scholar]
  • 54.Champoux MC, Wang G, Sarkarung S, Mackill DJ, O’Toole JC, Huang N, McCouch SR. Locating genes associated with root morphology and drought avoidance in rice via linkage to molecular markers. Theor Appl Genet. 1995;90:969–981. doi: 10.1007/BF00222910. [DOI] [PubMed] [Google Scholar]
  • 55.Ali MA, Abbas A, Niaz S, Zulkiffal M, Ali S. Morpho-physiological criteria for drought tolerance in sorghum (Sorghum bicolor) at seedling and post-anthesis stages. Int J Agric Biol. 2009;11:674–680. [Google Scholar]
  • 56.Poorter L, Markesteijn L. Seedling traits determine drought tolerance of tropical tree species. Biotropica. 2008;40:321–331. [Google Scholar]
  • 57.Begg J, Turner N, Kramer P. Morphological adaptations of leaves to water stress. Adaptation of plants to water and high temperature stress. New York: Wiley; 1980. [Google Scholar]
  • 58.Ludlow MM, Bjorkman O. Paraheliotropic leaf movement in Siratro as a protective mechanism against drought-induced damage to primary photosynthetic reactions: damage by excessive light and heat. Planta. 1984;161:505–518. doi: 10.1007/BF00407082. [DOI] [PubMed] [Google Scholar]
  • 59.Stevenson K, Shaw R. Effects of leaf orientation on leaf resistance to water vapor diffusion in soybean (Glycine max L. Merr) leaves. Agron J. 1971;63:327–329. [Google Scholar]
  • 60.Meyer WS, Walker S. Leaflet orientation in water-stressed soybeans. Agron J. 1981;73:1071–1074. [Google Scholar]
  • 61.Oosterhuis DM, Walker S, Eastham J. Soybean leaflet movements as an indicator of crop water stress. Crop Sci. 1985;25:1101–1106. [Google Scholar]
  • 62.Hsiao TC, O’Toole JC, Yambao EB, Turner NC. Influence of osmotic adjustment on leaf rolling and tissue death in rice (Oryza sativa L.) Plant Physiol. 1984;75:338–341. doi: 10.1104/pp.75.2.338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Esau K. Anatomy of seed plants. Soil Sci. 1960;90:149. [Google Scholar]
  • 64.Abclulrahaman A, Oladele E. Response of trichomes to water stress in two species of Jatropha . Insight bot. 2011;1:15–21. [Google Scholar]
  • 65.Mohammadian MA, Watling JR, Hill RS. The impact of epicuticular wax on gas-exchange and photoinhibition in Leucadendron lanigerum (Proteaceae) Acta Oecol. 2007;31:93–101. [Google Scholar]
  • 66.Terashima I. Anatomy of non-uniform leaf photosynthesis. Photosyn Res. 1992;31:195–212. doi: 10.1007/BF00035537. [DOI] [PubMed] [Google Scholar]
  • 67.Guha A, Sengupta D, Kumar Rasineni G, Ramachandra Reddy A. An integrated diagnostic approach to understand drought tolerance in mulberry (Morus indica L.) Flora Morphol Distrib Funct Ecol Plants. 2010;205:144–151. [Google Scholar]
  • 68.Hetherington AM, Woodward FI. The role of stomata in sensing and driving environmental change. Nature. 2003;424:901–908. doi: 10.1038/nature01843. [DOI] [PubMed] [Google Scholar]
  • 69.Hosy E, Vavasseur A, Mouline K, Dreyer I, Gaymard F, Porée F, Boucherez J, Lebaudy A, Bouchez D, Véry A-A. The Arabidopsis outward K+ channel GORK is involved in regulation of stomatal movements and plant transpiration. Proc Natl Acad Sci USA. 2003;100:5549–5554. doi: 10.1073/pnas.0733970100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Li L, Kim B-G, Cheong YH, Pandey GK, Luan S. A Ca2+ signaling pathway regulates a K+ channel for low-K response in Arabidopsis . Proc Natl Acad Sci USA. 2006;103:12625–12630. doi: 10.1073/pnas.0605129103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Cousson A, Vavasseur A. Putative involvement of cytosolic Ca2+ and GTP-binding proteins in cyclic-GMP-mediated induction of stomatal opening by auxin in Commelina communis L. Planta. 1998;206:308–314. [Google Scholar]
  • 72.Wang H, Wang X, Zhang S, Lou C. Muscarinic acetylcholine receptor is involved in acetylcholine regulating stomatal movement. Chin Sci Bull. 2000;45:250–252. doi: 10.1007/BF02882908. [DOI] [PubMed] [Google Scholar]
  • 73.Shimazaki K-i, Doi M, Assmann SM, Kinoshita T. Light regulation of stomatal movement. Annu Rev Plant Biol. 2007;58:219–247. doi: 10.1146/annurev.arplant.57.032905.105434. [DOI] [PubMed] [Google Scholar]
  • 74.Yamazaki D, Yoshida S, Asami T, Kuchitsu K. Visualization of abscisic acid-perception sites on the plasma membrane of stomatal guard cells. Plant J. 2003;35:129–139. doi: 10.1046/j.1365-313x.2003.01782.x. [DOI] [PubMed] [Google Scholar]
  • 75.Schroeder JI, Kwak JM, Allen GJ. Guard cell abscisic acid signalling and engineering drought hardiness in plants. Nature. 2001;410:327–330. doi: 10.1038/35066500. [DOI] [PubMed] [Google Scholar]
  • 76.Zhang J, Jia W, Yang J, Ismail AM. Role of ABA in integrating plant responses to drought and salt stresses. Field Crop Res. 2006;97:111–119. [Google Scholar]
  • 77.Herppich WB, Peckmann K. Influence of drought on mitochondrial activity, photosynthesis, nocturnal acid accumulation and water relations in the CAM plants Prenia sladeniana (ME-type) and Crassula lycopodioides (PEPCK-type) Ann Bot. 2000;86:611–620. [Google Scholar]
  • 78.Pagter M, Bragato C, Brix H. Tolerance and physiological responses of Phragmites australis to water deficit. Aquat Bot. 2005;81:285–299. [Google Scholar]
  • 79.Basu P, Sharma A, Garg I, Sukumaran N. Tuber sink modifies photosynthetic response in potato under water stress. Environ Exp Bot. 1999;42:25–39. [Google Scholar]
  • 80.Ashraf M, Harris P. Photosynthesis under stressful environments: an overview. Photosynthetica. 2013;51:163–190. [Google Scholar]
  • 81.Chaves MM, Maroco JP, Pereira JS. Understanding plant responses to drought-from genes to the whole plant. Funct Plant Biol. 2003;30:239–264. doi: 10.1071/FP02076. [DOI] [PubMed] [Google Scholar]
  • 82.Cushman JC. Crassulacean acid metabolism. A plastic photosynthetic adaptation to arid environments. Plant Physiol. 2001;127:1439–1448. [PMC free article] [PubMed] [Google Scholar]
  • 83.Lin Z, Peng C, Lin G. Photooxidation in leaves of facultative CAM plant Sedum spectabile at C3 and CAM mode. Acta Bot Sin. 2003;45:301–306. [Google Scholar]
  • 84.Kerbauy GB, Takahashi CA, Lopez AM, Matsumura AT, Hamachi L, Félix LM, Pereira PN, Freschi L, Mercier H (2012) Crassulacean acid metabolism in epiphytic orchids: current knowledge, future perspectives. Appl Photosyn, pp 81–104
  • 85.Sanchez R, Flores A, Cejudo FJ. Arabidopsis phosphoenolpyruvate carboxylase genes encode immunologically unrelated polypeptides and are differentially expressed in response to drought and salt stress. Planta. 2006;223:901–909. doi: 10.1007/s00425-005-0144-5. [DOI] [PubMed] [Google Scholar]
  • 86.Morgan JM. Osmoregulation and water stress in higher plants. Annu Rev Plant Physiol. 1984;35:299–319. [Google Scholar]
  • 87.Rhodes D, Samaras Y (1994) Genetic control of osmoregulation in plants. Cellular and Molecular Physiology of Cell Volume Regulation, pp 347–361
  • 88.Crowe JH, Crowe LM, Chapman D. Preservation of membranes in anhydrobiotic organisms: the role of trehalose. Science. 1984;223:701–703. doi: 10.1126/science.223.4637.701. [DOI] [PubMed] [Google Scholar]
  • 89.Bianchi G, Gamba A, Limiroli R, Pozzi N, Elster R, Salamini F, Bartels D. The unusual sugar composition in leaves of the resurrection plant Myrothamnus flabellifolia . Physiol Plantarum. 1993;87:223–226. [Google Scholar]
  • 90.Ranieri A, Bernardi R, Lanese P, Soldatini GF. Changes in free amino acid content and protein pattern of maize seedlings under water stress. Environ Exp Bot. 1989;29:351–357. [Google Scholar]
  • 91.Thomas H, James A. Freezing tolerance and solute changes in contrasting genotypes of Lolium perenne L. acclimated to cold and drought. Ann Bot. 1993;72:249–254. [Google Scholar]
  • 92.Wang S, Wan C, Wang Y, Chen H, Zhou Z, Fu H, Sosebee RE. The characteristics of Na+, K+ and free proline distribution in several drought-resistant plants of the Alxa Desert, China. J Arid Environ. 2004;56:525–539. [Google Scholar]
  • 93.Schobert B, Tschesche H. Unusual solution properties of proline and its interaction with proteins. Biochim Biophys Acta Gen Subj. 1978;541:270–277. doi: 10.1016/0304-4165(78)90400-2. [DOI] [PubMed] [Google Scholar]
  • 94.Arakawa T, Timasheff S. The stabilization of proteins by osmolytes. Biophys J. 1985;47:411–414. doi: 10.1016/S0006-3495(85)83932-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Arakawa T, Timasheff SN. Stabilization of protein structure by sugars. Biochemistry. 1982;21:6536–6544. doi: 10.1021/bi00268a033. [DOI] [PubMed] [Google Scholar]
  • 96.Schobert B. Is there an osmotic regulatory mechanism in algae and higher plants? J Theor Biol. 1977;68:17–26. doi: 10.1016/0022-5193(77)90224-7. [DOI] [PubMed] [Google Scholar]
  • 97.Hoekstra FA, Golovina EA, Buitink J. Mechanisms of plant desiccation tolerance. Trends Plant Sci. 2001;6:431–438. doi: 10.1016/s1360-1385(01)02052-0. [DOI] [PubMed] [Google Scholar]
  • 98.Solomon A, Beer S, Waisel Y, Jones G, Paleg L. Effects of NaCl on the carboxylating activity of Rubisco from Tamarix jordanis in the presence and absence of proline-related compatible solutes. Physiol Plantarum. 1994;90:198–204. [Google Scholar]
  • 99.Venekamp J, Lampe J, Koot J. Organic acids as sources for drought-induced proline synthesis in field bean plants, Vicia faba L. J Plant Physiol. 1989;133:654–659. [Google Scholar]
  • 100.Saradhi PP. Proline accumulation under heavy metal stress. J Plant Physiol. 1991;138:554–558. [Google Scholar]
  • 101.McLean WF, Blunden G, Jewers K. Quaternary ammonium compounds in the Capparaceae . Biochem Syst Ecol. 1996;24:427–434. [Google Scholar]
  • 102.Sakamoto A, Murata N. The role of glycine betaine in the protection of plants from stress: clues from transgenic plants. Plant Cell Environ. 2002;25:163–171. doi: 10.1046/j.0016-8025.2001.00790.x. [DOI] [PubMed] [Google Scholar]
  • 103.Gorham J. Betaines in higher plants-biosynthesis and role in stress metabolism. Seminar series. Proc: Cambridge University Press, Cambridge; 1995. [Google Scholar]
  • 104.Ashraf M, Foolad M. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ Exp Bot. 2007;59:206–216. [Google Scholar]
  • 105.Shao HB, Liang ZS, Shao MA. LEA proteins in higher plants: structure, function, gene expression and regulation. Colloid Surf B. 2005;45:131–135. doi: 10.1016/j.colsurfb.2005.07.017. [DOI] [PubMed] [Google Scholar]
  • 106.Close TJ. Dehydrins: emergence of a biochemical role of a family of plant dehydration proteins. Physiol Plantarum. 1996;97:795–803. [Google Scholar]
  • 107.Maurel C. Aquaporins and water permeability of plant membranes. Annu Rev Plant Physiol Plant Mol Biol. 1997;48:399–429. doi: 10.1146/annurev.arplant.48.1.399. [DOI] [PubMed] [Google Scholar]
  • 108.Maurel C, Verdoucq L, Luu DT, Santoni V. Plant aquaporins: membrane channels with multiple integrated functions. Annu Rev Plant Biol. 2008;59:595–624. doi: 10.1146/annurev.arplant.59.032607.092734. [DOI] [PubMed] [Google Scholar]
  • 109.Santoni V, Gerbeau P, Javot H, Maurel C. The high diversity of aquaporins reveals novel facets of plant membrane functions. Curr Opin Plant Biol. 2000;3:476–481. doi: 10.1016/s1369-5266(00)00116-3. [DOI] [PubMed] [Google Scholar]
  • 110.Cruz dCM. Drought stress and reactive oxygen species: production, scavenging and signaling. Plant Signal Behav. 2008;3:156–165. doi: 10.4161/psb.3.3.5536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Moller IM. Plant mitochondria and oxidative stress: electron transport, NADPH turnover, and metabolism of reactive oxygen species. Annu Rev Plant Physiol Plant Mol Biol. 2001;52:561–591. doi: 10.1146/annurev.arplant.52.1.561. [DOI] [PubMed] [Google Scholar]
  • 112.Gill SS, Tuteja N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol Biochem. 2010;48:909–930. doi: 10.1016/j.plaphy.2010.08.016. [DOI] [PubMed] [Google Scholar]
  • 113.Mittler R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002;7:405–410. doi: 10.1016/s1360-1385(02)02312-9. [DOI] [PubMed] [Google Scholar]
  • 114.Møller IM, Jensen PE, Hansson A. Oxidative modifications to cellular components in plants. Annu Rev Plant Biol. 2007;58:459–481. doi: 10.1146/annurev.arplant.58.032806.103946. [DOI] [PubMed] [Google Scholar]
  • 115.Apel K, Hirt H. Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol. 2004;55:373–399. doi: 10.1146/annurev.arplant.55.031903.141701. [DOI] [PubMed] [Google Scholar]
  • 116.Mittler R, Vanderauwera S, Suzuki N, Miller G, Tognetti VB, Vandepoele K, Gollery M, Shulaev V, Van Breusegem F. ROS signaling: the new wave? Trends Plant Sci. 2011;16:300–309. doi: 10.1016/j.tplants.2011.03.007. [DOI] [PubMed] [Google Scholar]
  • 117.Miyake C. Alternative electron flows (water-water cycle and cyclic electron flow around PSI) in photosynthesis: molecular mechanisms and physiological functions. Plant Cell Physiol. 2010;51:1951–1963. doi: 10.1093/pcp/pcq173. [DOI] [PubMed] [Google Scholar]
  • 118.Dixon DP, Cummins L, Cole DJ, Edwards R. Glutathione-mediated detoxification systems in plants. Curr Opin Plant Biol. 1998;1:258–266. doi: 10.1016/s1369-5266(98)80114-3. [DOI] [PubMed] [Google Scholar]
  • 119.Blokhina O, Virolainen E, Fagerstedt KV. Antioxidants, oxidative damage and oxygen deprivation stress: a review. Ann Bot. 2003;91:179–194. doi: 10.1093/aob/mcf118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Reguera M, Peleg Z, Blumwald E. Targeting metabolic pathways for genetic engineering abiotic stress-tolerance in crops. Biochim Biophys Acta. 2012;1819:186–194. doi: 10.1016/j.bbagrm.2011.08.005. [DOI] [PubMed] [Google Scholar]
  • 121.Hou X, Xie K, Yao J, Qi Z, Xiong L. A homolog of human ski-interacting protein in rice positively regulates cell viability and stress tolerance. Proc Natl Acad Sci USA. 2009;106:6410–6415. doi: 10.1073/pnas.0901940106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Bano A, Hansen H, Dörffling K, Hahn H. Changes in the contents of free and conjugated abscisic acid, phaseic acid and cytokinins in xylem sap of drought stressed sunflower plants. Phytochemistry. 1994;37:345–347. [Google Scholar]
  • 123.Peleg Z, Blumwald E. Hormone balance and abiotic stress tolerance in crop plants. Curr Opin Plant Biol. 2011;14:290–295. doi: 10.1016/j.pbi.2011.02.001. [DOI] [PubMed] [Google Scholar]
  • 124.Sauter A, Davies WJ, Hartung W. The long-distance abscisic acid signal in the droughted plant: the fate of the hormone on its way from root to shoot. J Exp Bot. 2001;52:1991–1997. doi: 10.1093/jexbot/52.363.1991. [DOI] [PubMed] [Google Scholar]
  • 125.Campalans A, Messeguer R, Goday A, Pagès M. Plant responses to drought, from ABA signal transduction events to the action of the induced proteins. Plant Physiol Biochem. 1999;37:327–340. [Google Scholar]
  • 126.Wilkinson S, Davies WJ. ABA-based chemical signalling: the co-ordination of responses to stress in plants. Plant09 Cell Environ. 2002;25:195–210. doi: 10.1046/j.0016-8025.2001.00824.x. [DOI] [PubMed] [Google Scholar]
  • 127.Boursiac Y, Léran S, Corratgé-Faillie C, Gojon A, Krouk G, Lacombe B. ABA transport and transporters. Trends Plant Sci. 2013;18:325–333. doi: 10.1016/j.tplants.2013.01.007. [DOI] [PubMed] [Google Scholar]
  • 128.Bacon MA, Wilkinson S, Davies WJ. pH-regulated leaf cell expansion in droughted plants is abscisic acid dependent. Plant Physiol. 1998;118:1507–1515. doi: 10.1104/pp.118.4.1507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Rabbani MA, Maruyama K, Abe H, Khan MA, Katsura K, Ito Y, Yoshiwara K, Seki M, Shinozaki K, Yamaguchi-Shinozaki K. Monitoring expression profiles of rice genes under cold, drought, and high-salinity stresses and abscisic acid application using cDNA microarray and RNA gel-blot analyses. Plant Physiol. 2003;133:1755–1767. doi: 10.1104/pp.103.025742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Cocucci M, Negrini N. Changes in the levels of calmodulin and of a calmodulin inhibitor in the early phases of radish (Raphanus sativus L.) seed germination: effects of ABA and Fusicoccin. Plant Physiol. 1988;88:910–914. doi: 10.1104/pp.88.3.910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Mori IC, Murata Y, Yang Y, Munemasa S, Wang YF, Andreoli S, Tiriac H, Alonso JM, Harper JF, Ecker JR. CDPKs CPK6 and CPK3 function in ABA regulation of guard cell S-type anion-and Ca2+-permeable channels and stomatal closure. PLoS Biol. 2006 doi: 10.1371/journal.pbio.0040327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Harris MJ, Outlaw WH. Rapid adjustment of guard-cell abscisic acid levels to current leaf-water status. Plant Physiol. 1991;95:171–173. doi: 10.1104/pp.95.1.171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Tardieu F, Davies WJ. Stomatal response to abscisic acid is a function of current plant water status. Plant Physiol. 1992;98:540–545. doi: 10.1104/pp.98.2.540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Delauney AJ, Verma DPS. Proline biosynthesis and osmoregulation in plants. Plant J. 1993;4:215–223. [Google Scholar]
  • 135.Verbruggen N, Hermans C. Proline accumulation in plants: a review. Amino Acids. 2008;35:753–759. doi: 10.1007/s00726-008-0061-6. [DOI] [PubMed] [Google Scholar]
  • 136.Verslues PE, Bray EA. Role of abscisic acid (ABA) and Arabidopsis thaliana ABA-insensitive loci in low water potential-induced ABA and proline accumulation. J Exp Bot. 2006;57:201–212. doi: 10.1093/jxb/erj026. [DOI] [PubMed] [Google Scholar]
  • 137.Strizhov N, Abraham E, Ökrész L, Blickling S, Zilberstein A, Schell J, Koncz C, Szabados L. Differential expression of two P5CS genes controlling proline accumulation during salt-stress requires ABA and is regulated by ABA1, ABI1 and AXR2 in Arabidopsis . Plant J. 1997;12:557–569. doi: 10.1046/j.1365-313x.1997.00557.x. [DOI] [PubMed] [Google Scholar]
  • 138.Abrahám E, Rigó G, Székely G, Nagy R, Koncz C, Szabados L. Light-dependent induction of proline biosynthesis by abscisic acid and salt stress is inhibited by brassinosteroid in Arabidopsis . Plant Mol Biol. 2003;51:363–372. doi: 10.1023/a:1022043000516. [DOI] [PubMed] [Google Scholar]
  • 139.Shinozaki K, Yamaguchi-Shinozaki K. Gene networks involved in drought stress response and tolerance. J Exp Bot. 2007;58:221–227. doi: 10.1093/jxb/erl164. [DOI] [PubMed] [Google Scholar]
  • 140.Choi H, Hong J, Ha J, Kang J, Kim SY. ABFs, a family of ABA-responsive element binding factors. J Biol Chem. 2000;275:1723–1730. doi: 10.1074/jbc.275.3.1723. [DOI] [PubMed] [Google Scholar]
  • 141.Sayed O. Chlorophyll fluorescence as a tool in cereal crop research. Photosynthetica. 2003;41:321–330. [Google Scholar]
  • 142.Li RH, Guo PG, Michael B, Stefania G, Salvatore C. Evaluation of chlorophyll content and fluorescence parameters as indicators of drought tolerance in barley. Agric Sci China. 2006;5:751–757. [Google Scholar]
  • 143.Kocheva K, Lambrev P, Georgiev G, Goltsev V, Karabaliev M. Evaluation of chlorophyll fluorescence and membrane injury in the leaves of barley cultivars under osmotic stress. Bioelectrochemistry. 2004;63:121–124. doi: 10.1016/j.bioelechem.2003.09.020. [DOI] [PubMed] [Google Scholar]
  • 144.Guo P, Baum M, Varshney RK, Graner A, Grando S, Ceccarelli S. QTLs for chlorophyll and chlorophyll fluorescence parameters in barley under post-flowering drought. Euphytica. 2008;163:203–214. [Google Scholar]
  • 145.Bartels D, Sunkar R. Drought and salt tolerance in plants. Crit Rev Plant Sci. 2005;24:23–58. [Google Scholar]
  • 146.Zhu JK. Salt and drought stress signal transduction in plants. Annu Rev Plant Biol. 2002;53:247–273. doi: 10.1146/annurev.arplant.53.091401.143329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Xiong L, Schumaker KS, Zhu JK. Cell signaling during cold, drought, and salt stress. Plant Cell. 2002;14:165–183. doi: 10.1105/tpc.000596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Shinozaki K, Yamaguchi-Shinozaki K, Seki M. Regulatory network of gene expression in the drought and cold stress responses. Curr Opin Plant Biol. 2003;6:410–417. doi: 10.1016/s1369-5266(03)00092-x. [DOI] [PubMed] [Google Scholar]
  • 149.Hirayama T, Shinozaki K. Research on plant abiotic stress responses in the post-genome era: past, present and future. Plant J. 2010;61:1041–1052. doi: 10.1111/j.1365-313X.2010.04124.x. [DOI] [PubMed] [Google Scholar]
  • 150.Wang D, Pan Y, Zhao X, Zhu L, Fu B, Li Z. Genome-wide temporal-spatial gene expression profiling of drought responsiveness in rice. BMC Genom. 2011;12:149. doi: 10.1186/1471-2164-12-149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Malone JH, Oliver B. Microarrays, deep sequencing and the true measure of the transcriptome. BMC Biol. 2011;9:34. doi: 10.1186/1741-7007-9-34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Dugas DV, Monaco MK, Olson A, Klein RR, Kumari S, Ware D, Klein PE. Functional annotation of the transcriptome of Sorghum bicolor in response to osmotic stress and abscisic acid. BMC Genom. 2011;12:514. doi: 10.1186/1471-2164-12-514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Seki M, Narusaka M, Ishida J, Nanjo T, Fujita M, Oono Y, Kamiya A, Nakajima M, Enju A, Sakurai T. Monitoring the expression profiles of 7000 Arabidopsis genes under drought, cold and high-salinity stresses using a full-length cDNA microarray. Plant J. 2002;31:279–292. doi: 10.1046/j.1365-313x.2002.01359.x. [DOI] [PubMed] [Google Scholar]
  • 154.Zou J-J, Wei F-J, Wang C, Wu J-J, Ratnasekera D, Liu W-X, Wu W-H. Arabidopsis calcium-dependent protein kinase CPK10 functions in abscisic acid-and Ca2+-mediated stomatal regulation in response to drought stress. Plant Physiol. 2010;154:1232–1243. doi: 10.1104/pp.110.157545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Saijo Y, Hata S, Kyozuka J, Shimamoto K, Izui K. Over-expression of a single Ca2+-dependent protein kinase confers both cold and salt/drought tolerance on rice plants. Plant J. 2000;23:319–327. doi: 10.1046/j.1365-313x.2000.00787.x. [DOI] [PubMed] [Google Scholar]
  • 156.Yang W, Kong Z, Omo-Ikerodah E, Xu W, Li Q, Xue Y. Calcineurin B-like interacting protein kinase OsCIPK23 functions in pollination and drought stress responses in rice (Oryza sativa L.) J Genet Genomics. 2008;35:S531–S532. doi: 10.1016/S1673-8527(08)60073-9. [DOI] [PubMed] [Google Scholar]
  • 157.Ning J, Li X, Hicks LM, Xiong L. A Raf-like MAPKKK gene DSM1 mediates drought resistance through reactive oxygen species scavenging in rice. Plant Physiol. 2010;152:876–890. doi: 10.1104/pp.109.149856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Xiong L, Yang Y. Disease resistance and abiotic stress tolerance in rice are inversely modulated by an abscisic acid-inducible mitogen-activated protein kinase. Plant Cell. 2003;15:745–759. doi: 10.1105/tpc.008714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Sinha AK, Jaggi M, Raghuram B, Tuteja N. Mitogen-activated protein kinase signaling in plants under abiotic stress. Plant Signal Behav. 2011;6:196–203. doi: 10.4161/psb.6.2.14701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Umezawa T, Yoshida R, Maruyama K, Yamaguchi-Shinozaki K, Shinozaki K. SRK2C, a SNF1-related protein kinase 2, improves drought tolerance by controlling stress-responsive gene expression in Arabidopsis thaliana . Proc Natl Acad Sci USA. 2004;101:17306–17311. doi: 10.1073/pnas.0407758101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Aharoni A, Dixit S, Jetter R, Thoenes E, van Arkel G, Pereira A. The SHINE clade of AP2 domain transcription factors activates wax biosynthesis, alters cuticle properties, and confers drought tolerance when overexpressed in Arabidopsis . Plant Cell. 2004;16:2463–2480. doi: 10.1105/tpc.104.022897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Chen JQ, Meng XP, Zhang Y, Xia M, Wang XP. Over-expression of OsDREB genes lead to enhanced drought tolerance in rice. Biotechnol Lett. 2008;30:2191–2198. doi: 10.1007/s10529-008-9811-5. [DOI] [PubMed] [Google Scholar]
  • 163.Zhao L, Hu Y, Chong K, Wang T. ARAG1, an ABA-responsive DREB gene, plays a role in seed germination and drought tolerance of rice. Ann Bot. 2010;105:401–409. doi: 10.1093/aob/mcp303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Yoshida T, Fujita Y, Sayama H, Kidokoro S, Maruyama K, Mizoi J, Shinozaki K, Yamaguchi-Shinozaki K. AREB1, AREB2, and ABF3 are master transcription factors that cooperatively regulate ABRE-dependent ABA signaling involved in drought stress tolerance and require ABA for full activation. Plant J. 2010;61:672–685. doi: 10.1111/j.1365-313X.2009.04092.x. [DOI] [PubMed] [Google Scholar]
  • 165.Xiang Y, Tang N, Du H, Ye H, Xiong L. Characterization of OsbZIP23 as a key player of the basic leucine zipper transcription factor family for conferring abscisic acid sensitivity and salinity and drought tolerance in rice. Plant Physiol. 2008;148:1938–1952. doi: 10.1104/pp.108.128199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Tang N, Zhang H, Li X, Xiao J, Xiong L. Constitutive activation of transcription factor OsbZIP46 improves drought tolerance in rice. Plant Physiol. 2012;158:1755–1768. doi: 10.1104/pp.111.190389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Cominelli E, Galbiati M, Vavasseur A, Conti L, Sala T, Vuylsteke M, Leonhardt N, Dellaporta SL, Tonelli C. A guard-cell-specific MYB transcription factor regulates stomatal movements and plant drought tolerance. Curr Biol. 2005;15:1196–1200. doi: 10.1016/j.cub.2005.05.048. [DOI] [PubMed] [Google Scholar]
  • 168.Yang A, Dai X, Zhang W-H. A R2R3-type MYB gene, OsMYB2, is involved in salt, cold, and dehydration tolerance in rice. J Exp Bot. 2012;63:2541–2556. doi: 10.1093/jxb/err431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Hu H, Dai M, Yao J, Xiao B, Li X, Zhang Q, Xiong L. Overexpressing a NAM, ATAF, and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in rice. Proc Natl Acad Sci USA. 2006;103:12987–12992. doi: 10.1073/pnas.0604882103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Saad AS, Li X, Li HP, Huang T, Gao CS, Guo MW, Cheng W, Zhao GY, Liao YC. A rice stress-responsive NAC gene enhances tolerance of transgenic wheat to drought and salt stresses. Plant Sci. 2013;203–204:33–40. doi: 10.1016/j.plantsci.2012.12.016. [DOI] [PubMed] [Google Scholar]
  • 171.Liu G, Li X, Jin S, Liu X, Zhu L, Nie Y, Zhang X. Overexpression of rice NAC gene SNAC1 improves drought and salt tolerance by enhancing root development and reducing transpiration rate in transgenic cotton. PLoS ONE. 2014;9:e86895. doi: 10.1371/journal.pone.0086895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Jeong JS, Kim YS, Baek KH, Jung H, Ha SH, Do Choi Y, Kim M, Reuzeau C, Kim JK. Root-specific expression of OsNAC10 improves drought tolerance and grain yield in rice under field drought conditions. Plant Physiol. 2010;153:185–197. doi: 10.1104/pp.110.154773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Sugano S, Kaminaka H, Rybka Z, Catala R, Salinas J, Matsui K, Ohme-Takagi M, Takatsuji H. Stress-responsive zinc finger gene ZPT2-3 plays a role in drought tolerance in petunia. Plant J. 2003;36:830–841. doi: 10.1046/j.1365-313x.2003.01924.x. [DOI] [PubMed] [Google Scholar]
  • 174.Huang XY, Chao DY, Gao JP, Zhu MZ, Shi M, Lin HX. A previously unknown zinc finger protein, DST, regulates drought and salt tolerance in rice via stomatal aperture control. Genes Dev. 2009;23:1805–1817. doi: 10.1101/gad.1812409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Nambara E, Marion-Poll A. Abscisic acid biosynthesis and catabolism. Annu Rev Plant Biol. 2005;56:165–185. doi: 10.1146/annurev.arplant.56.032604.144046. [DOI] [PubMed] [Google Scholar]
  • 176.Oritani T, Kiyota H. Biosynthesis and metabolism of abscisic acid and related compounds. Nat Prod Rep. 2003;20:414–425. doi: 10.1039/b109859b. [DOI] [PubMed] [Google Scholar]
  • 177.Xiong L, Zhu JK. Regulation of abscisic acid biosynthesis. Plant Physiol. 2003;133:29–36. doi: 10.1104/pp.103.025395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Park H-Y, Seok H-Y, Park B-K, Kim S-H, Goh C-H, B-h Lee, Lee C-H, Moon Y-H. Overexpression of Arabidopsis ZEP enhances tolerance to osmotic stress. Biochem Biophys Res Commun. 2008;375:80–85. doi: 10.1016/j.bbrc.2008.07.128. [DOI] [PubMed] [Google Scholar]
  • 179.Qin X, Zeevaart JA. Overexpression of a 9-cis-epoxycarotenoid dioxygenase gene in Nicotiana plumbaginifolia increases abscisic acid and phaseic acid levels and enhances drought tolerance. Plant Physiol. 2002;128:544–551. doi: 10.1104/pp.010663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Iuchi S, Kobayashi M, Taji T, Naramoto M, Seki M, Kato T, Tabata S, Kakubari Y, Yamaguchi-Shinozaki K, Shinozaki K. Regulation of drought tolerance by gene manipulation of 9-cis-epoxycarotenoid dioxygenase, a key enzyme in abscisic acid biosynthesis in Arabidopsis . Plant J. 2001;27:325–333. doi: 10.1046/j.1365-313x.2001.01096.x. [DOI] [PubMed] [Google Scholar]
  • 181.Xiong L, Ishitani M, Lee H, Zhu JK. The Arabidopsis LOS5/ABA3 locus encodes a molybdenum cofactor sulfurase and modulates cold stress- and osmotic stress-responsive gene expression. Plant Cell. 2001;13:2063–2083. doi: 10.1105/TPC.010101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Yue Y, Zhang M, Zhang J, Duan L, Li Z. Arabidopsis LOS5/ABA3 overexpression in transgenic tobacco (Nicotiana tabacum cv. Xanthi-nc) results in enhanced drought tolerance. Plant Sci. 2011;181:405–411. doi: 10.1016/j.plantsci.2011.06.010. [DOI] [PubMed] [Google Scholar]
  • 183.Yue Y, Zhang M, Zhang J, Tian X, Duan L, Li Z. Overexpression of the AtLOS5 gene increased abscisic acid level and drought tolerance in transgenic cotton. J Exp Bot. 2012;63:3741–3748. doi: 10.1093/jxb/ers069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Xiao B-Z, Chen X, Xiang C-B, Tang N, Zhang Q-F, Xiong L-Z. Evaluation of seven function-known candidate genes for their effects on improving drought resistance of transgenic rice under field conditions. Mol Plant. 2009;2:73–83. doi: 10.1093/mp/ssn068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Bowler C, Mv Montagu, Inze D. Superoxide dismutase and stress tolerance. Annu Rev Plant Biol. 1992;43:83–116. [Google Scholar]
  • 186.Wang F-Z, Wang Q-B, Kwon S-Y, Kwak S-S, Su W-A. Enhanced drought tolerance of transgenic rice plants expressing a pea manganese superoxide dismutase. J Plant Physiol. 2005;162:465–472. doi: 10.1016/j.jplph.2004.09.009. [DOI] [PubMed] [Google Scholar]
  • 187.Koussevitzky S, Suzuki N, Huntington S, Armijo L, Sha W, Cortes D, Shulaev V, Mittler R. Ascorbate peroxidase 1 plays a key role in the response of Arabidopsis thaliana to stress combination. J Biol Chem. 2008;283:34197–34203. doi: 10.1074/jbc.M806337200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Miao Y, Lv D, Wang P, Wang XC, Chen J, Miao C, Song CP. An Arabidopsis glutathione peroxidase functions as both a redox transducer and a scavenger in abscisic acid and drought stress responses. Plant Cell. 2006;18:2749–2766. [Google Scholar]
  • 189.You J, Zong W, Li X, Ning J, Hu H, Xiao J, Xiong L. The SNAC1-targeted gene OsSRO1c modulates stomatal closure and oxidative stress tolerance by regulating hydrogen peroxide in rice. J Exp Bot. 2013;64:569–583. doi: 10.1093/jxb/ers349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Hand SC, Menze MA, Toner M, Boswell L, Moore D. LEA proteins during water stress: not just for plants anymore. Annu Rev Physiol. 2011;73:115–134. doi: 10.1146/annurev-physiol-012110-142203. [DOI] [PubMed] [Google Scholar]
  • 191.Xiao B, Huang Y, Tang N, Xiong L. Over-expression of a LEA gene in rice improves drought resistance under the field conditions. Theor Appl Genet. 2007;115:35–46. doi: 10.1007/s00122-007-0538-9. [DOI] [PubMed] [Google Scholar]
  • 192.Rai V, Tuteja N, Takabe T. Improving crop resistance to abiotic stress. New York: Wiley; 2012. [Google Scholar]
  • 193.Wei A, He C, Li B, Li N, Zhang J. The pyramid of transgenes TsVP and BetA effectively enhances the drought tolerance of maize plants. Plant Biotechnol J. 2011;9:216–229. doi: 10.1111/j.1467-7652.2010.00548.x. [DOI] [PubMed] [Google Scholar]
  • 194.Klein M, Geisler M, Suh SJ, Kolukisaoglu HÜ, Azevedo L, Plaza S, Curtis MD, Richter A, Weder B, Schulz B. Disruption of AtMRP4, a guard cell plasma membrane ABCC-type ABC transporter, leads to deregulation of stomatal opening and increased drought susceptibility. Plant J. 2004;39:219–236. doi: 10.1111/j.1365-313X.2004.02125.x. [DOI] [PubMed] [Google Scholar]
  • 195.Lian H-L, Yu X, Ye Q, Ding X-S, Kitagawa Y, Kwak S-S, Su W-A, Tang Z-C. The role of aquaporin RWC3 in drought avoidance in rice. Plant Cell Physiol. 2004;45:481–489. doi: 10.1093/pcp/pch058. [DOI] [PubMed] [Google Scholar]
  • 196.Peng Y, Lin W, Cai W, Arora R. Overexpression of a Panax ginseng tonoplast aquaporin alters salt tolerance, drought tolerance and cold acclimation ability in transgenic Arabidopsis plants. Planta. 2007;226:729–740. doi: 10.1007/s00425-007-0520-4. [DOI] [PubMed] [Google Scholar]
  • 197.Chen TH, Murata N. Enhancement of tolerance of abiotic stress by metabolic engineering of betaines and other compatible solutes. Curr Opin Plant Biol. 2002;5:250–257. doi: 10.1016/s1369-5266(02)00255-8. [DOI] [PubMed] [Google Scholar]
  • 198.Yeo ET, Kwon HB, Han SE, Lee JT, Ryu JC, Byu M. Genetic engineering of drought resistant potato plants by introduction of the trehalose-6-phosphate synthase (TPS1) gene from Saccharomyces cerevisiae . Mol Cells. 2000;10:263–268. [PubMed] [Google Scholar]
  • 199.Zhang N, Si H-J, Wen G, Du HH, Liu BL, Wang D. Enhanced drought and salinity tolerance in transgenic potato plants with a BADH gene from spinach. Plant Biotechnol Rep. 2011;5:71–77. [Google Scholar]
  • 200.Kishor PK, Hong Z, Miao GH, Hu CAA, Verma DPS. Overexpression of Δ1-pyrroline-5-carboxylate synthetase increases proline production and confers osmotolerance in transgenic plants. Plant Physiol. 1995;108:1387–1394. doi: 10.1104/pp.108.4.1387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zhu X, Xiong L. Putative megaenzyme DWA1 plays essential roles in drought resistance by regulating stress-induced wax deposition in rice. Proc Natl Acad Sci USA. 2013;110:17790–17795. doi: 10.1073/pnas.1316412110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.You J, Hu H, Xiong L. An ornithine delta-aminotransferase gene OsOAT confers drought and oxidative stress tolerance in rice. Plant Sci. 2012;197:59–69. doi: 10.1016/j.plantsci.2012.09.002. [DOI] [PubMed] [Google Scholar]
  • 203.Bolanos J, Edmeades G. The importance of the anthesis-silking interval in breeding for drought tolerance in tropical maize. Field Crops Res. 1996;48:65–80. [Google Scholar]
  • 204.Monneveux P, Ribaut JM (2006) Secondary traits for drought tolerance improvement in cereals. Drought adaptation in cereals, pp 97–143
  • 205.Wasson AP, Richards RA, Chatrath R, Misra SC, Prasad SV, Rebetzke GJ, Kirkegaard JA, Christopher J, Watt M. Traits and selection strategies to improve root systems and water uptake in water-limited wheat crops. J Exp Bot. 2012;63:3485–3498. doi: 10.1093/jxb/ers111. [DOI] [PubMed] [Google Scholar]
  • 206.Kato Y, Abe J, Kamoshita A, Yamagishi J. Genotypic variation in root growth angle in rice (Oryza sativa L.) and its association with deep root development in upland fields with different water regimes. Plant Soil. 2006;287:117–129. [Google Scholar]
  • 207.Mace ES, Singh V, Van Oosterom EJ, Hammer GL, Hunt CH, Jordan DR. QTL for nodal root angle in sorghum (Sorghum bicolor L. Moench) co-locate with QTL for traits associated with drought adaptation. Theor Appl Genet. 2012;124:97–109. doi: 10.1007/s00122-011-1690-9. [DOI] [PubMed] [Google Scholar]
  • 208.Cuellar-Ortiz SM, De La Paz Arrieta-Montiel M, Acosta-Gallegos J, Covarrubias AA. Relationship between carbohydrate partitioning and drought resistance in common bean. Plant Cell Environ. 2008;31:1399–1409. doi: 10.1111/j.1365-3040.2008.01853.x. [DOI] [PubMed] [Google Scholar]
  • 209.Hossain A, Sears R, Cox TS, Paulsen G. Desiccation tolerance and its relationship to assimilate partitioning in winter wheat. Crop Sci. 1990;30:622–627. [Google Scholar]
  • 210.Araus JL, Slafer GA, Royo C, Serret MD. Breeding for yield potential and stress adaptation in cereals. Crit Rev Plant Sci. 2008;27:377–412. [Google Scholar]
  • 211.Van Herwaarden A, Richards R, Angus J (2003) Water-soluble carbohydrates and yield in wheat. In: Proceedings of the 11th Australian agronomy conference. The Australian Society of Agronomy, Geelong
  • 212.Rebetzke G, Van Herwaarden A, Jenkins C, Weiss M, Lewis D, Ruuska S, Tabe L, Fettell N, Richards R. Quantitative trait loci for water-soluble carbohydrates and associations with agronomic traits in wheat. Crop Pasture Sci. 2008;59:891–905. [Google Scholar]
  • 213.Richards R. Defining selection criteria to improve yield under drought. Plant Growth Regul. 1996;20:157–166. [Google Scholar]
  • 214.Zheng H, Wu A, Zheng C, Wang Y, Cai R, Shen X, Xu R, Liu P, Kong L, Dong S. QTL mapping of maize (Zea mays) stay-green traits and their relationship to yield. Plant Breeding. 2009;128:54–62. [Google Scholar]
  • 215.Harris K, Subudhi PK, Borrell A, Jordan D, Rosenow D, Nguyen H, Klein P, Klein R, Mullet J. Sorghum stay-green QTL individually reduce post-flowering drought-induced leaf senescence. J Exp Bot. 2007;58:327–338. doi: 10.1093/jxb/erl225. [DOI] [PubMed] [Google Scholar]
  • 216.King CA, Purcell LC. Soybean nodule size and relationship to nitrogen fixation response to water deficit. Crop Sci. 2001;41:1099–1107. [Google Scholar]
  • 217.Sinclair TR, Purcell LC, King CA, Sneller CH, Chen P, Vadez V. Drought tolerance and yield increase of soybean resulting from improved symbiotic N2 fixation. Field Crops Res. 2007;101:68–71. [Google Scholar]
  • 218.Zhou Y. Drought resistance of turf bermudagrasses (Cynodon spp.) collected from Australian biodiversity. Australia: The University of Queensland; 2013. [Google Scholar]
  • 219.Zhou Y, Lambrides CJ, Fukai S. Drought resistance and soil water extraction of a perennial C4 grass: contributions of root and rhizome traits. Funct Plant Biol. 2014;41:505–519. doi: 10.1071/FP13249. [DOI] [PubMed] [Google Scholar]
  • 220.van Ginkel M, Ogbonnaya F. Novel genetic diversity from synthetic wheats in breeding cultivars for changing production conditions. Field Crops Res. 2007;104:86–94. [Google Scholar]
  • 221.Olivares-Villegas JJ, Reynolds MP, McDonald GK. Drought-adaptive attributes in the Seri/Babax hexaploid wheat population. Funct Plant Biol. 2007;34:189–203. doi: 10.1071/FP06148. [DOI] [PubMed] [Google Scholar]
  • 222.Flintham J, Börner A, Worland A, Gale M. Optimizing wheat grain yield: effects of Rht (gibberellin-insensitive) dwarfing genes. J Agric Sci. 1997;128:11–25. [Google Scholar]
  • 223.Tester M, Langridge P. Breeding technologies to increase crop production in a changing world. Science. 2010;327:818–822. doi: 10.1126/science.1183700. [DOI] [PubMed] [Google Scholar]
  • 224.Ashraf M. Inducing drought tolerance in plants: recent advances. Biotechnol Adv. 2010;28:169–183. doi: 10.1016/j.biotechadv.2009.11.005. [DOI] [PubMed] [Google Scholar]
  • 225.Shou H, Bordallo P, Wang K. Expression of the Nicotiana protein kinase (NPK1) enhanced drought tolerance in transgenic maize. J Exp Bot. 2004;55:1013–1019. doi: 10.1093/jxb/erh129. [DOI] [PubMed] [Google Scholar]
  • 226.Yoon HK, Kim SG, Kim SY, Park CM. Regulation of leaf senescence by NTL9-mediated osmotic stress signaling in Arabidopsis . Mol Cells. 2008;25:438–445. [PubMed] [Google Scholar]
  • 227.Nakashima K, Tran LS, Van Nguyen D, Fujita M, Maruyama K, Todaka D, Ito Y, Hayashi N, Shinozaki K, Yamaguchi-Shinozaki K. Functional analysis of a NAC-type transcription factor OsNAC6 involved in abiotic and biotic stress-responsive gene expression in rice. Plant J. 2007;51:617–630. doi: 10.1111/j.1365-313X.2007.03168.x. [DOI] [PubMed] [Google Scholar]

Articles from Cellular and Molecular Life Sciences: CMLS are provided here courtesy of Springer

RESOURCES