Chemical root to shoot signaling under drought

Size: px
Start display at page:

Download "Chemical root to shoot signaling under drought"

Transcription

1 Review Chemical root to shoot signaling under drought Daniel P. Schachtman 1 and Jason Q.D. Goodger 2 1 Donald Danforth Plant Science Center, 975N. Warson Rd, St. Louis, MO 63132, USA 2 School of Botany, The University of Melbourne, VIC 3010 Australia Chemical signals are important for plant adaptation to water stress. As soils become dry, root-sourced signals are transported via the xylem to leaves and result in reduced water loss and decreased leaf growth. The presence of chemical signals in xylem sap is accepted, but the identity of these signals is controversial. Abscisic acid (ABA), ph, cytokinins, a precursor of ethylene, malate and other unidentified factors have all been implicated in root to shoot signaling under drought. This review describes current knowledge of, and advances in, research on chemical signals that are sent from roots under drought. The contribution of these different potential signals is discussed within the context of their role in stress signaling. Chemical and hydraulic signaling under drought stress Numerous studies have shown that plant roots can sense changes in abiotic factors such as soil water content [1,2], soil bulk density or compaction [3], soil oxygen content [4] and changes in the nutrient composition of soil (both enrichment and depletion) [5]. Root sensing of water deficit has been widely studied, and this review focuses on the outcomes of the early stage of water-deficit sensing: the transport through the xylem of chemical signals that ultimately reduce leaf transpiration and leaf growth. Chemical signals can be differentiated from hydraulic signals [6] but both are important because they reduce stomatal conductance and leaf growth under conditions of water deficit. Chemical signals most probably dominate during early stages of stress before hydraulic signals are produced [7], and become less important under severe drought when leaf water potential declines and leaves wilt. It is likely that the hydraulic signals that may trigger the production of the hormone abscisic acid (ABA) in leaves dominate as plants become more water stressed [8]. This review focuses on chemical signals because of their importance during the early stages of plant response to water deficit, and because their manipulation has applications in increasing water use efficiency (WUE) in agriculture. Agricultural practices involving chemical signals: deficit irrigation and partial root-zone drying Root to shoot signaling under conditions of both mild and severe drought is an important area for research because of its implications for agricultural production and the WUE of plants. Fresh water supplies are predominantly used for Corresponding author: Schachtman, D.P. (dschachtman@danforthcenter.org). agriculture, but this usage will need to be reduced in the future as supplies become limited. Where irrigation can be manipulated, researchers have experimented with, and successfully implemented, both deficit irrigation and partial root-zone drying (PRD) [9,10]. Both of these drying regimes reduce leaf transpiration and limit vegetative growth, thereby increasing WUE, and have demonstrated efficacy in both woody perennials [9] and herbaceous crop plants [11,12] (Box 1). Although these treatments can increase the WUE of certain crop species, knowledge of the role of chemical signals sent from root to shoot during such practices is still developing [13]. Some studies have shown that chemical signals such as ABA are generated by the reductions in soil water content and act on the leaves to reduce transpiration and growth [14]. Where PRD and deficit irrigation have been used in agriculture, the most success has come from applications to tree crops and to vines (in viticulture) [9,10]. With horticultural crops, the management of individual plants is more intensive and economic return is often linked to fruit quality. In one study, the use of PRD or deficit irrigation in viticulture increased WUE by about 40% while only decreasing yield by 15% [10]. PRD improved the quality of the grape berry by increasing the anthocyanin content [10] and the quality of wine produced [1]. In tomato (Lycopersicon esculentum), PRD has been shown to increase crop quality. For example, one PRD study found a 21% increase in the Brix value of fruit (a measurement of the mass ratio of dissolved sugar to water), and a 50% increase in WUE with only a 15% decrease in yield [15]. A physiological explanation for such a small reduction in yield has been proposed on the basis that the fruit receives water and minerals from the xylem early in development, but later, more than 90% of the water supplying the fruit is transported via the phloem [15]. Particularly in arid climates, the use of precision irrigation methods such as PRD and deficit irrigation puts into practice the principles developed from basic scientific studies on how root signals increase WUE while maintaining crop quality and yields. ABA is a key regulator of leaf stomatal conductance The production of ABA in roots and its transport to the leaves provides the plant with a mechanism for transmitting a chemical signal to report on the water status of the soil. This system could have evolved specifically for this purpose or the chemical signal could merely be a consequence of the increased production of ABA required to maintain root growth under water deficit [16]. A dominant /$ see front matter ß 2008 Elsevier Ltd. All rights reserved. doi: /j.tplants Available online 6 May

2 Review Trends in Plant Science Vol.13 No.6 Box 1. Partial root-zone drying and deficit irrigation in plants Partial root-zone drying (PRD) and deficit irrigation (DI) are examples of the manipulation of root signals to enhance the water use efficiency (WUE) of agriculture. PRD is an irrigation method where half the root zone is allowed to dry while the other half is kept wet. After a period of time, the irrigation is alternated so that the previously dry root zone receives water and the previously watered side is allowed to dry. DI is an irrigation method in which the amount of water supplied is less than that required to achieve maximum crop evapotranspiration. Both methods can reduce transpiration and leaf growth while increasing WUE with only small decreases in yield. Outstanding questions Is the source of ABA that results in stomatal closure under drought coming from the roots or the leaves? How much variation in root to shoot signaling and xylem sap composition is there within a species? How does the nutrient status of a plant interact with chemical signaling during water stress? In which region of the root and in which particular root cells is ABA synthesized? What are the different forms of cytokinins that are transported in xylem sap, and what roles do they play in root to shoot signaling? What causes changes in ph in sap under drought and why do ph changes appear inconsistent between certain species? role for ABA in root to shoot signaling under drought and in the control of stomatal conductance was demonstrated in early reports [17]. Several groups have also used a bioassay approach to show that ABA that is fed to leaves decreases transpiration [18,19]. It is also well known that ABA has a direct effect on guard cell closure [20]. However, the ABA that acts on guard cells may not originate entirely in the roots. Recent reports suggest that the ABA that acts on guard cells could be produced in the leaves of some species, such as tomato and sunflower [21 23]. In tobacco (Nicotiana plumbaginifolia), grafting experiments showed that ABA came from the roots under conditions of drought [24]. ABA is synthesized in both roots [25] and leaves, but not much is known about the precise location of this synthesis in roots, which may influence how plants perceive and monitor soil water content. ABA content in roots is well correlated with both soil moisture and the relative water content of roots in many plant species [26,27]. Some reports suggest that ABA is synthesized in root tips [28], but synthesis has been assessed mainly by measuring ABA content rather than synthetic activity, so conclusions related to the site of synthesis need to be re-examined more carefully. One study using detached roots of pea (Pisum sativum) and Asiatic dayflower (Commelina communis) found that ABA synthesis occurred somewhere between the root tip and a point 3 cm distal to the tip [29]. Recent studies have tried to locate enzymes that are involved in ABA biosynthesis, including aldehyde oxidase, which catalyzes the final steps in ABA biosynthesis; however, these studies have focused on ABA production under differing nitrogen nutrition rather than drought stress [30]. More work will be required to clarify exactly where the ABA is produced in roots under drought stress. Although the role of ABA in controlling stomatal conductance is strongly supported by many experiments and experimental approaches [25], there may be other substances in the sap that act in concert with ABA to control drought-induced stomatal closure. For example, recent findings have suggested an interaction between jasmonic acid and ABA in plants under drought stress [31]. In addition, the dominant role of ABA as a root to shoot signal has been challenged by experiments showing that the ABA concentrations of xylem sap from droughtstressed plants were much lower than the concentrations of exogenous ABA required to close stomata in detached leaves [32]. It may be that the use of exogenous hormones in bioassays may exclude important components of xylem sap that act synergistically with ABA in planta, thus accounting for the higher levels of exogenous ABA that are required to reduce transpiration. Grafting experiments have also been used to determine the source of ABA in drought-induced stomatal closure. Some of these experiments suggest that leaf-sourced ABA is important for stomatal closure [22], but the influence of ABA precursors that might be transported from roots to leaves was not measured, and therefore the role of roots in chemical signaling was not completely resolved [25]. Other studies show that production of ABA in the rootstock has a strong influence on and is negatively correlated with stomatal conductance [33]. Overall, ABA is a dominate signal in controlling growth and transpiration, but other factors could also be important. The importance and role of root-sourced ABA is still controversial, but some of the conflicting findings may be due to differences in the intensity of stress imposed and the time-course of the development of water deficit. Involvement of ph in signaling Changes in the ph of xylem sap commonly observed under drought stress can be an important component of root to shoot signaling and may act synergistically with ABA. In many plant species (e.g. sunflower [Helianthus annuus] [34], Phaseolus coccineus [35] and C. communis [36]), xylem sap ph becomes more alkaline when plants are water stressed and this leads to enhanced stomatal closure and even reduced growth. The potential effects of ph have been outlined previously [37] and include: i) changes in ABA metabolism resulting in increased leaf ABA concentration; ii) direct effects on leaf water status that could alter guard cell turgor or sensitivity to leaf ABA concentrations; iii) direct effects on ion fluxes through the guard cell plasma membrane; and iv) altered distribution of ABA in the leaf, specifically an increase the concentration of ABA in the apoplast outside of the guard cells (Figure 1). It is this change in apoplast ph that probably has the greatest role in signaling. As ABA is a weak acid it may become protonated or deprotonated in the ph range found in the apoplast of leaves. This was clearly shown in the early 1980s in experiments on mesophyll protoplasts that documented high rates of ABA uptake at acidic ph and almost no uptake at the more alkaline ph of 7.5 [38]. In certain plant species, the ph of the leaf apoplast increases as the soil dries because of the delivery of xylem sap that is of a more alkaline ph. The ABA that arrives from roots via the xylem will remain deprotonated under the more alkaline conditions and will not be taken up passively by mesophyll 282

3 Review Trends in Plant Science Vol.13 No.6 cells. The result is that less ABA is transported into the mesophyll cells and a build-up of ABA in the apoplast leads to enhanced stomatal closure. Under well-watered conditions, when the apoplast is more acidic, ABA would passively enter the symplast, and apoplastic concentrations would not increase as rapidly as they do under water stress. Therefore, the effect of ph is indirect in that it leads to the accumulation of ABA in the apoplast and enhances the effect of ABA on guard cells [35]. This mechanistic explanation is both attractive and intuitive. Nevertheless, the explanation sets up another hypothesis that ABA receptors are on the plasma membrane rather than inside the guard cells. Recently, three ABA receptors have been found [39 41]. Two of these ABA receptors reside inside the cell but a third was found on the cell surface [39]. Therefore, plant cells could sense both extracellular and intracellular ABA concentrations. Under conditions of drought, the increased stomatal closure that occurs because of the increased ph of sap suggests that extracellular ABA is sensed by guard cells via receptors on the plasma membrane. Several studies have shown that ph does not act alone on guard cells or in controlling leaf growth. For example, the use of tomato [42] and barley (Hordeum vulgare) [43] mutants that are deficient in ABA biosynthesis showed that ABA was necessary for both stomatal closure and the inhibition of leaf growth. In the tomato mutant flacca, transpiration increased as the ph of artificial sap increased, suggesting that ABA is also necessary to prevent stomatal opening and water loss. In the barley mutant Az34, ABA was necessary for ph to act as a signal under conditions of drought stress [43]. Many studies that show a role for ph in stomatal closure have used tomato as a model system. It is important to point out, however, that there may be differences in response to ABA and ph between species and perhaps even between genotypes within the same species. One study compared C. communis and Arabidopsis thaliana and found that the increased ph of external solutions led to a 27% increase in stomatal opening in C. communis but had no effect on A. thaliana [44]. Both species, however, responded similarly in the presence of ABA and only changes in the kinetics of response were observed. In soybean (Glycine max) [45], xylem sap ph did not change in response to mild drought and was not correlated with decreased stomatal conductance during the initial stages of drought. The xylem sap of soybean did, however, become more alkaline during extended drought, but this alkalization occurred well after a measured decrease in stomatal conductance. The response of xylem sap ph to drought stress does not appear to be consistent in all species or even in different experiments using the same species. One study showed that the sap ph of sunflower and C. communis did not change significantly as soil dried, whereas the xylem sap ph of tomato did increase as soil dried [46]. In maize (Zea mays) grown using ammonium as the nitrogen source, there was no change in sap ph during early [7] or even during later stages of drought (L. Ernst and D.P. Schachtman, unpublished). Other experiments using maize have been conducted in the field and have found that the sap ph of drought-stressed plants growing on a loam soil was more alkaline as the period of drought lengthened [47]. Such inconsistencies may be explained by the fact that solute concentrations within the xylem sap vary with flow rate in intact plants [48,49]. The sap solute concentrations and delivery rates that best represent those of xylem sap flow in intact plants are most accurately determined when sap is collected at rates equivalent to whole-plant transpiration. Therefore the rate of whole plant transpiration must be determined prior to detopping plants to harvest sap. External pressure must be applied to roots so that sap exudes at the flow rate of whole plant transpiration, a fact often overlooked by experiments in which only xylem sap is collected [49]. The mechanism of ph change might involve nitrate availability. As soils dry, nutrient availability is reduced because of physiochemical changes. Under conditions of reduced nitrate availability, nitrate reductase activity shifts to roots, causing changes in the ph of sap [50]. Changes in the activity of nitrate reductase that are caused by drought conditions lead to changes in organic acid production, which alter sap ph. In particular, malate concentration increases in xylem sap under drought, resulting in sap that is less acidic than that in well-watered conditions when more nitrate is loaded into the xylem. When the effects of a nitrate reductase inhibitor (sodium tungstate) were tested [51], the inhibitor was found to prevent the alkalization of the sap in tomato under drought conditions. Similarly, the addition of nitrate to tomato under water-stress conditions enhanced the alkalization of the xylem sap [46]. A synergistic effect of xylem nitrate and ABA on stomatal conductance in detached leaves of C. communis has also been shown [46]. Moreover, the results of work on maize suggest that, under some circumstances where neither compound is as effective alone, a combination of ABA and nitrate is required to elicit an effect on shoot physiology [52]. The effects of changes in the internal ph of the guard cells have been very well characterized [20]. Less is known about the effects of apoplastic changes in ph on guard cells, but one study [53] showed that stomatal opening is inhibited by acidification and that the plasma membrane potential is depolarized (which would facilitate the opening of outward-rectifying potassium channels) when the apoplast is acidified. More studies are needed to clarify the effects of changes in external ph on guard cells when the internal ph is held constant. If extracellular alkalinization also leads to an increased intracellular ph, it could be an important part of the signaling process for stomatal closure. Conjugates of ABA as potential signals Although ABA seems to play a dominant role in root to shoot signaling under drought, it also seems likely, on the basis of several studies, that other substances are also involved. ABA is present in xylem sap in conjugated forms, such as abscisic acid-glucose ester (ABA-GE). It has been suggested that a conjugated form of ABA could serve as a transported form of the hormone and, moreover, as a stress signal [32,54]. As many as six ABA conjugates have been found in xylem sap from well-watered and water-stressed sunflower [55], and a correlation was found in one study 283

4 Review Trends in Plant Science Vol.13 No.6 Figure 1. Changes in xylem sap composition under drought. Drought causes the alkalinization of xylem sap ph in certain plant species. (a) Well-watered plant with apoplastic ph 6.0. (b) Plant under drought conditions with apoplastic ph 7.0. In plants in which xylem sap ph increases when the soil becomes dry, ABA-induced stomatal closure is enhanced. This is thought to be due to increased apoplastic concentrations of ABA. Further changes in xylem sap composition under drought are also responsible 284

5 Review Trends in Plant Science Vol.13 No.6 between increased ABA-GE concentrations and drought [56]. The importance of conjugated forms of ABA in drought responses and the mechanisms of hydrolysis [57] of such forms of ABA in the apoplast are, however, still poorly understood. A next step in establishing the importance of conjugated forms of ABA might be to genetically vary the hydrolysis of ABA conjugates. To do this, it will be essential to identify the genes encoding the apoplastic b-glucosidases that are responsible for hydrolysis before subsequent manipulation of these enzymes can proceed. Cytokinins as signals Cytokinins could also be an important signal traveling from roots to the shoots. Root-produced cytokinins are clearly involved in responses to nutrient deprivation [5] and, as they are produced mainly in roots, could be important in drought responses [58]. Despite this, there have been few reports that provide information on the cytokinin content of xylem sap and how that content changes under drought conditions. In grapevines, a 50% reduction in zeatin (Z) and zeatin riboside (ZR) was found in plants that had been subjected to PRD [59]. In a more recent study in which Z, ZR and zeatin nucleotide (ZN) were measured, PRD of tomato reduced the ZN content of the xylem sap, but the magnitude of that change and the contribution of ZN to the total cytokinin content were not shown [60].Inat least two studies on sunflower, xylem sap, combined Z and ZR and combined isopentenyladenine and isopentenyladenosine concentrations in xylem sap decreased under drought-stressed conditions [61]. It is possible that the ABA:cytokinin ratios in xylem sap are important for stress signaling [55]. The effects of cytokinins are also supported by a study showing that transpiration was higher in transgenic plants that overexpressed the isopentenyltransferase, causing the plants to produce more cytokinins after a heat shock treatment [62]. Although recent data show decreased cytokinin concentrations in the xylem under drought stress, it is still not clear that all plant species respond in the same way to cytokinin at the concentrations found in the leaf and guard cells [63]. Strong conclusions about cytokinin content are premature because the complexity of the cytokinin profiles has not been fully explored [64]. In a recent study on maize [65], we found a decrease in Z and ZR concentrations in xylem sap coming from roots of drought stressed plants as compared to well watered controls. Surprisingly, we also found high concentrations of the aromatic cytokinin 6-benzylaminopurine (BAP) in maize xylem sap, the concentration of which increased significantly as a result of water stress [65]. With the availability of increasingly sensitive mass spectrometry methods, it should be possible to explore comprehensively the changes in concentrations of many of the cytokinins, and to identify those that play a role in root to shoot signaling. Other chemical signals Chemical and protein-based factors other than ABA, cytokinins or ph could also be involved in root to shoot signaling. Although it has now been established clearly that xylem sap from many plant species contains proteins [66,67], the presence of peptides in xylem sap has only recently been demonstrated [68]. The addition of xylem sap from tomato induced rapid alkalinization in a tomato cellsuspension bioassay within minutes. This approach has been used previously to identify peptides. In an experiment using maize suspension cells, alkalinization was induced by maize xylem sap, but the activity could not be purified or abolished with proteases [65]. MicroRNAs have also been implicated as potential signal molecules that move systemically [69], and recent work on Arabidopsis found that drought treatment induced the synthesis of a specific microrna: mir-169 g [70]. The functional significance of peptides, micrornas and proteins in the xylem sap remains largely unknown, but peptides have been shown to play important signaling roles in plants and have also been shown to move systemically via the phloem [71,72]. Many compounds, including hormones, inorganic ions, amino acids, sugars and organic acids have been identified in xylem sap [7,34], but among the organic acids found in abundance in the xylem sap, only malate has been implicated in the control of stomatal closure [73]. Stomatal opening in ash (Fraxinus excelsior) leaves could be prevented when these leaves were supplied with mm malate. Similarly, elevated extracellular concentrations of malate have also been shown to close stomates in Vicia faba [74]. Further work is needed to establish the role of malate in root to shoot signaling. Ethylene could be another important factor under drought conditions. It is known that a precursor of ethylene, 1-aminocyclopropane-1-carboxylic acid (ACC), moves in the xylem from root to shoots [75]. Under drought stress, ACC transported in the xylem sap might result in the measured increase of ethylene evolution in leaves [76]. A role for ethylene under drought was demonstrated by the use of an ACC oxidase (ACO) antisense in tomato. In these antisense plants, ethylene evolution was much lower than normal under both well-watered and drought conditions. Under soil-drying conditions, the stomatal response in the ACO antisense plants was the same as wild type, but a decrease in leaf growth was measured in wild type but not the ACO antisense plants in response to soil drying [76].In maize, ethylene evolution in leaves could not be correlated with reductions in leaf elongation under drought [77]. This study also showed that ABA does not play a role in reducing leaf elongation in maize [77]. This suggests that for reduced transpiration and inhibition of leaf growth. These changes include increases in malate and ACC (1-aminocyclopropane-1-carboxylate) concentrations, decreases in flow rates, and reductions in the concentrations of cytokinins, zeatin and zeatin riboside. After guard cells perceive ABA, changes in potassium (K + ) and anion (A ) fluxes result in stomatal closure. (c) Stomatal closure under drought. The affects of ABA on guard cells under drought suggest the importance of a plasma-membrane-bound receptor, such as the recently identified GCR2 (G-PROTEIN COUPLED RECEPTOR2), but the importance of intracellular receptors, such as CHLH (the H subunit of the magnesium protoporphyrin-ix chelatase that is localized in the chloroplast), cannot be ruled out. Other unknown and yet-to-be-identified plasma membrane receptors could also be important in transducing the increased ABA concentrations in the apoplast that cause stomatal closure. After guard cells perceive ABA, the efflux of potassium (K + ) and anions (A ) leads to stomatal closure. SLAC, slow anion channel; RAC, rapid anion channel. 285

6 Review Trends in Plant Science Vol.13 No.6 ethylene may play a role in decreased leaf growth in one plant species and that ACC may be one component of longdistance root-sourced signals under drought. Conclusions The production of root-sourced chemical signals under conditions of water deficit has been associated with reduced transpiration and/or leaf growth. However, the identity and relative contribution to signaling of these rootsourced chemicals remains controversial. This controversy may be due to differing responses between species, the different intensities of stress treatments applied, the time at which samples were collected during the imposition of drought, and/or the different methods used for xylem sap extraction [43]. Until recently [65], few studies have provided comprehensive information on the composition of xylem sap. Such comprehensive analyses will clarify the contribution of different chemicals to root to shoot signaling, and the complexity of constituents and their interactions. As we increase our knowledge of the precise identity and biosynthesis of the primary chemical signals that are produced by roots, it will become feasible to alter plant sensitivity to soil water deficit. This will provide new molecular breeding strategies for tailoring crops to maintain or even increase yields under specific water stress conditions, such as intermittent and terminal drought. Acknowledgements DPS was supported in part by NSF-Plant Genome # JQDG is the recipient of an Australian Research Council Post-doctoral Fellowship (Inductry linkage project #LP ) References 1 Davies, W.J. et al. (2002) Stomatal control by chemical signalling and the exploitation of this mechanism to increase water use efficiency in agriculture. New Phytol. 153, Wilkinson, S. and Davies, W.J. (2002) ABA-based chemical signalling: the co-ordination of responses to stress in plants. Plant Cell Environ. 25, Masle, J. and Farquhar, G.D. (1988) Effects of soil strength on the relation of water-use efficiency and growth to carbon isotope discrimination in wheat seedlings. Plant Physiol. 86, Drew, M.C. et al. (1990) Regulation of K + uptake and transport to the xylem in barley roots; K + distribution determined by electron probe X-ray microanalysis of frozen-hydrated cells. J. Exp. Bot. 41, Schachtman, D.P. and Shin, R. (2007) Nutrient sensing and signaling: NPKS. Annu. Rev. Plant Biol. 58, Gollan, T. et al. (1986) Soil water status affects the stomatal conductance of fully turgid wheat and sunflower leaves. Aust. J. Plant Physiol. 13, Goodger, J.Q.D. et al. (2005) Relationships between xylem sap constituents and leaf conductance of well-watered and waterstressed maize across three xylem sap sampling techniques. J. Exp. Bot. 56, Christmann, A. et al. (2007) A hydraulic signal in root-to-shoot signalling of water shortage. Plant J. 52, Fereres, E. and Soriano, M.A. (2007) Deficit irrigation for reducing agricultural water use. J. Exp. Bot. 58, dos Santos, T.P. et al. (2003) Partial rootzone drying: effects on growth and fruit quality of field-grown grapevines (Vitis vinifera). Funct. Plant Biol. 30, Kang, S. et al. (1998) Water use efficiency of controlled alternate irrigation on root-divided maize plants. Agric. Water Manag. 38, Tang, L.S. et al. (2005) Physiological and yield responses of cotton under partial rootzone irrigation. Field Crops Res. 94, Dodd, I.C. (2007) Soil moisture heterogeneity during deficit irrigation alters root-to-shoot signaling of abscisic acid. Funct. Plant Biol. 34, Dry, P.R. and Loveys, B.R. (1999) Grapevine shoot growth and stomatal conductance are reduced when part of the root system is dried. Vitis 38, Davies, W.J. et al. (2000) Regulation of leaf and fruit growth in plants growing in drying soil: exploitation of the plants chemical signaling system and hydraulic architecture to increase the efficiency of water use in agriculture. J. Exp. Bot. 51, Sharp, R.E. et al. (1994) Confirmation that abscisic-acid accumulation is required for maize primary root elongation at low water potentials. J. Exp. Bot. 45, Davies, W.J. and Zhang, J. (1991) Root signals and the regulation of growth and development of plants in drying soil. Annu. Rev. Plant Physiol. Mol. Biol. 42, Munns, R. (1992) A leaf elongation assay detects an unknown growth inhibitor in xylem sap from wheat and barley. Aust. J. Plant Physiol. 19, Zhang, J. and Davies, W.J. (1991) Antitranspirant activity in xylem sap of maize plants. J. Exp. Bot. 42, Schroeder, J.I. et al. (2001) Guard cell signal transduction. Annu. Rev. Plant Physiol. Mol. Biol. 52, Fambrini, M. et al. (1995) Characterization of a wilty sunflower (Helianthus annuus L.) mutant III. Phenotypic interaction in reciprocal grafts from wilty mutant and wild-type plants. J. Exp. Bot. 286, Holbrook, N.M. et al. (2002) Stomatal control in tomato with ABAdeficient roots: response of grafted plants to soil drying. J. Exp. Bot. 53, Jones, H.G. et al. (1987) Growth and water relations of wilty mutants of tomato (Lycopersicon esculentum Mill.). J. Exp. Bot. 38, Borel, C. et al. (2001) Does engineering abscisic acid biosynthesis in Nicotiana plumbaginifolia modify stomatal response to drought? Plant Cell Environ. 24, Thompson, A.J. et al. (2007) Regulation and manipulation of ABA biosynthesis in roots. Plant Cell Environ. 30, Zhang, J. and Davies, W.J. (1989) Abscisic acid produced in dehydrating roots may enable the plant to measure the water status of the soil. Plant Cell Environ. 12, Liang, Z. et al. (1997) How do roots control xylem sap ABA concentration in response to soil drying. Plant Cell Physiol. 38, Zhang, J. and Tardieu, F. (1996) Relative contribution of apices and mature tissues to ABA synthesis in droughted maize root systems. Plant Cell Physiol. 37, Zhang, J. and Davies, W.J. (1987) Increased synthesis of ABA in partially dehydrated root tips and ABA transport from roots to leaves. J. Exp. Bot. 38, Katalin Barabás, N. et al. (2000) Distribution of the Mo-enzymes aldehyde oxidase, xanthine dehydrogenase and nitrate reductase in maize (Zea mays L.) nodal roots as affected by nitrogen and salinity. Plant Sci. 155, Mahouachi, J. et al. (2007) Hormonal changes in papaya seedlings subjected to progressive water stress and re-watering. Plant Growth Regul. 53, Munns, R. and King, R.W. (1988) Abscisic acid is not the only stomatal inhibitor in the transpiration stream of wheat plants. Plant Physiol. 88, Soar, C.J. et al. (2006) Scion photosynthesis and leaf gas exchange in Vitis vinifera L. cv. Shiraz: Mediation of rootstock effects via xylem sap ABA. Aust. J. Grape Wine Res. 12, Gollan, T. et al. (1992) Stomatal response to drying soil in relation to changes in the xylem sap composition of Helianthus annus. I. The concentration of cations, anions, amino acids in, and ph of, the xylem sap. Plant Cell Environ. 15, Hartung, W. et al. (1998) Factors that regulate abscisic acid concentrations at the primary site of action at the guard cell. J. Exp. Bot. 49, Wilkinson, S. and Davies, W.J. (1997) Xylem sap ph increase: a drought signal received at the apoplastic face of the guard cell that involves the suppression of saturable abscisic acid uptake by the epidermal symplast. Plant Physiol. 113, Wilkinson, S. (1999) ph as a stress signal. Plant Growth Regul. 29,

7 Review Trends in Plant Science Vol.13 No.6 38 Kaiser, W.M. and Hartung, W. (1981) Uptake and release of abscisic acid by isolated photoautotrophic mesophyll cells, depending on ph gradients. Plant Physiol. 68, Liu, X. et al. (2007) A G protein-coupled receptor is a plasma membrane receptor for the plant hormone abscisic acid. Science 315, Razem, F.A. et al. (2006) The RNA-binding protein FCA is an abscisic acid receptor. Nature 439, Shen, Y.Y. et al. (2006) The Mg-chelatase H subunit is an abscisic acid receptor. Nature 443, Wilkinson, S. et al. (1998) Effects of xylem ph on transpiration from wild-type and flacca tomato leaves. A vital role for abscisic acid in preventing excessive water loss even from well-watered plants. Plant Physiol. 117, Bacon, M.A. et al. (1998) ph-regulated leaf cell expansion in droughted plants is abscisic acid dependent. Plant Physiol. 118, Prokic, L. et al. (2006) Species-dependent changes in stomatal sensitivity to abscisic acid mediated by external ph. J. Exp. Bot. 57, Liu, F.L. et al. (2003) Hydraulic and chemical signals in the control of leaf expansion and stomatal conductance in soybean exposed to drought stress. Funct. Plant Biol. 30, Jia, W. and Davies, W.J. (2007) Modification of leaf apoplast ph in relation to stomatal sensitivity to root-sourced abscisic acid signals. Plant Physiol. 143, Bahrun, A. et al. (2002) Drought-induced changes in xylem ph, ionic composition, and ABA concentration act as early signals in field-grown maize (Zea mays L.). J. Exp. Bot. 53, Else, M.A. et al. (1994) Concentrations of abscisic acid and other solutes in xylem sap from root systems of tomato and castor-oil plants are distorted by wounding and variable sap flow rates. J. Exp. Bot. 45, Jackson, M.B. (1997) Hormones from roots as signals for the shoots of stressed plants. Trends Plant Sci. 2, Liu, F.L. et al. (2005) A review of drought adaptation in crop plants: changes in vegetative and reproductive physiology induced by ABAbased chemical signals. Aust. J. Agric. Res. 56, Wilkinson, S. (2004) Water use efficiency and chemical signalling. In Water Use Efficiency in Plant Biology (Bacon, M.A., ed.), pp , Blackwell Publishing Ltd 52 Wilkinson, S. (2007) Nitrate signalling to stomata and growing leaves: interactions with soil drying, ABA, and xylem sap ph in maize. J. Exp. Bot. 58, Roelfsema, M.R. and Prins, H.B. (1998) The membrane potential of Arabidopsis thaliana guard cells; depolarizations induced by apoplastic acidification. Planta 205, Munns, R. and Sharp, R.E. (1993) Involvement of abscisic acid in controlling plant growth in soils of low water potential. Aust. J. Plant Physiol. 20, Hansen, H. and Dorffling, K. (1999) Changes of free and conjugated abscisic acid and phaseic acid in xylem sap of drought-stressed sunflower plants. J. Exp. Bot. 50, Sauter, A. et al. (2002) A possible stress physiological role of abscisic acid conjugates in root-to-shoot signalling. Plant Cell Environ. 25, Dietz, K.J. et al. (2000) Extracellular beta-glucosidase activity in barley involved in the hydrolysis of ABA glucose conjugate in leaves. J. Exp. Bot. 51, Sakakibara, H. (2006) Cytokinins: activity, biosynthesis, and translocation. Annu. Rev. Plant Biol. 57, Stoll, M. et al. (2000) Hormonal changes induced by partial rootzone drying of irrigated grapevine. J. Exp. Bot. 51, Kudoyarova, G. et al. (2007) Effect of partial rootzone drying on the concentration of zeatin-type cytokinins in tomato (Solanum lycopersicum L.) xylem sap and leaves. J. Exp. Bot. 2, Hansen, H. and Dörffling, K. (2003) Root-derived trans-zeatin riboside and abscisic acid in drought-stressed and rewatered sunflower plants: interaction in the control of leaf diffusive resistance? Funct. Plant Biol. 30, Teplova, I. et al. (2000) Response of tobacco plants transformed with the ipt gene to elevated temperature. Russ. J. Plant Physiol. 47, Dodd, I.C. (2003) Hormonal interactions and stomatal responses. J. Plant Growth Regul. 22, Davies, W.J. et al. (2005) Long-distance ABA signaling and its relation to other signaling pathways in the detection of the soil drying and the mediation of the plant s response to drought. J. Plant Growth Regul. 24, Alvarez, S. et al. (2008) Metabolomic and proteomic changes in the xylem sap of maize under drought. Plant Cell Environ. 31, Alvarez, S. et al. (2006) Characterization of the maize xylem sap proteome. J. Proteome Res. 5, Kehr, J. et al. (2005) Analysis of xylem sap proteins from Brassica napus. BMC Plant Biol. 5, Neumann, P.M. (2007) Evidence for long-distance xylem transport of signal peptide activity from tomato roots. J. Exp. Bot. 58, Sunkar, R. et al. (2007) Small RNAs as big players in plant abiotic stress responses and nutrient deprivation. Trends Plant Sci. 12, Zhao, B. et al. (2007) Identification of drought-induced micrornas in rice. Biochem. Biophys. Res. Commun. 354, Lindsey, K. et al. (2002) Peptides: new signalling molecules in plants. Trends Plant Sci. 7, Ryan, C.A. et al. (2002) Polypeptide hormones. Plant Cell 14 (Suppl), S251 S Patonnier, M.P. et al. (1999) Drought-induced increase in xylem malate and mannitol concentrations and closure of Fraxinus excelsior L. stomata. J. Exp. Bot. 50, Hedrich, R. et al. (1994) Malate-sensitive anion channels enable guardcells to sense changes in the ambient CO 2 concentration. Plant J. 6, Else, M.A. and Jackson, M.B. (1998) Transport of 1- aminocyclopropane-1-carboxylic acid (ACC) in the transpiration stream of tomato (Lycopersicon esculentum) in relation to foliar ethylene production and petiole epinasty. Aust. J. Plant Physiol. 25, Sobeih, W.Y. et al. (2004) Long-distance signals regulating stomatal conductance and leaf growth in tomato (Lycopersicon esculentum) plants subjected to partial root-zone drying. J. Exp. Bot. 55, Voisin, A.S. et al. (2006) Are ABA, ethylene or their interaction involved in the response of leaf growth to soil water deficit? An analysis using naturally occurring variation or genetic transformation of ABA production in maize. Plant Cell Environ. 29,

8 Ecology of Water Relations in Plants Yoseph Negusse Araya, The Open University, Milton Keynes, UK Water is an important resource for plant growth. Availability of water in the soil determines the niche, distribution and competitive interaction of plants in the environment. Introduction Advanced article Article Contents. Introduction. Water Uptake and Movement through Plants. Water Stress and Plants. Plant Sensing and Adaptation to Water Stress. Distribution of Plants in Response to Water Regime doi: / a Importance of water for plants Water typically constitutes 80 95% of the mass of growing plant tissues and plays a crucial role for plant growth (Taiz and Zeiger, 1998). Plants require water for a number of physiological processes (e.g. synthesis of carbohydrates) and for associated physical functions (e.g. keeping plants turgid). Water accomplishes its many functions because of its unique characteristics: the polarity of the molecule H 2 O (which makes it an excellent solvent), viscosity (which makes it capable of moving through plant tissues by capillary action) and thermal properties (which makes it capable of cooling plant tissues). Plants require water, soil nutrients, carbon dioxide, oxygen and solar radiation for growth. Of these, water is most often the most limiting: influencing productivity (Taiz and Zeiger, 1998) as well as the diversity of species (Rodriguez- Iturbe and Porporato, 2004) in both natural and agricultural ecosystems. This is illustrated graphically in Figure 1. How does water affect ecology of plants? In order to understand the ecology of plant water relations it is important to understand from where and how plants acquire water in their environment (the latter is discussed in the section on water uptake and movement through plants). Unlike animals, which are capable of wandering around to forage for resources, plants are for the most part stationary, depending on the availability of nutrients in their surrounding environment (soil and/or atmosphere). Of these two sources of resources, i.e. soil and atmosphere, the soil is by far the major and more accessible reservoir. Consequently, the soil is the primary store and regulator in the water flow of ecosystems, by intercepting precipitation input and controlling its use by organisms (Rodriguez-Iturbe and Porporato, 2004). Figure 2 summarizes the soil and plant water interrelationship. Soil moisture availability is dependent on the soil particle size distribution (also called soil texture) and arrangement of these particles (soil structure). The soil texture and structure influence the size of soil pores where water is held by Total net productivity g 1 m 2 year 1 Diversity/Shannon Wiener index Moisture index Moisture Forest Woodland Grassland Desert Total net productivity Forest Woodland Grassland Desert Plant species diversity Forest Woodland Grassland Desert Figure 1 Moisture, total net productivity and plant species diversity of selected vegetation communities, along an elevation gradient from Santa Catalina Mountains, Arizona (after Whittaker and Niering, 1975). The elevation gradient ranges from 1000 to 3000 m above sea level. The moisture index relates to precipitation ranges of 190 mm per annum (moisture index 8) and 850 mm per annum (moisture index 1). ENCYCLOPEDIA OF LIFE SCIENCES & 2007, John Wiley & Sons, Ltd. 1

9 Ecology of Water Relations in Plants Organic matter return Soil nutrients Mineralization and transport Plant uptake Plant uptake Plants Soil moisture Plant water loss Diffusion and gas exchange Soil aeration Air-filled pore space Figure 2 Schematic summary of the processes that influence the relationship between plants and soil water. Soil water potential (kpa) Soil water content (% volume of pore space) 100 Figure 3 Soil water availability and soil aeration availability for two representative sandy (solid line) and clayey (broken line) soils. Soil water contents on volume basis is shown against soil water potential (suction) and against air-filled pore space (volume of pore space not occupied by water). capillary forces. Soils with fine-sized particles, like clay hold more water than soils dominated by coarse grain particles of sand. However, this does not mean that all the water in fine-particle sized soils is available for plant uptake. This is because the capillary forces holding water in the pores of fine-textured soils are so powerful that the plants struggle to extract any water. Figure 3 shows soil water and soil aeration availability for different soil texture classes. Soil moisture availability primarily influences plants by two routes (see Figure 2), either by being directly limiting as a resource, or indirectly by filling pore spaces in soil and thereby excluding air, causing oxygen availability to become limiting for the activity of plant roots. This is explained further in the section on water stress and plants. Water Uptake and Movement through Plants Water is constantly moving from the soil, into plant roots, and through the xylem tissues of the stem to leaves where it Air-filled pore space (% volume) is ultimately lost to the atmosphere during transpiration. This cycle is referred to as the soil plant atmosphere continuum (SPAC). See also: Plant Water Relations When water moves through the SPAC, it travels through different mediums (including cell wall, cell membrane and air spaces) at different distances, which utilize different modes of transport. There are three principal modes of water transport: diffusion, mass flow and osmosis. In diffusion, water molecules move spontaneously from regions of high concentration to regions of lower concentration, i.e. along a concentration gradient. This movement is rapid over a short distance and thus drives short-distance transport, for example between cells and during the loss of water to the atmosphere from leaf stomata. In mass flow, groups of water molecules move under an external force, such as a build-up of pressure that forms a gradient. Mass flow of water is the predominant mode by which long-distance transport of water in stems is accomplished. It also accounts for much of the water flow though the soil and through the cell walls of plants. The third mode, called osmosis, is movement of water molecules through a semipermeable membrane, an example of which is the cytoplasmic membrane. Osmosis occurs spontaneously in both short- and long-distance transport as a response to driving forces of concentration (as in diffusion) and pressure gradient (as in mass flow). These driving forces of water movement of both osmotic (concentration) and mass flow (pressure) origin are collectively known as water potential. Water potential is measured in units of pressure or suction, i.e. force per unit area required to move a specific amount of water. The most common unit used for studying soil water potentials in the field is kilopascals (kpa). The movement of water in the SPAC is thus dependent on differences in water potential between surrounding soil and plant or atmosphere. Often, the water potential gradient is directed from the roots towards the shoot, as transpiring leaves exposed to the atmosphere have the lowest water potential. However, under situations when the soil is too dry this water potential gradient could be reversed, resulting in loss of water from plant roots to the soil. Also any environmental factors that influence the transpiration of the leaf stomata, e.g. wind or increase in temperature may further decrease the leaf water potential further, speeding up water loss. Water Stress and Plants In addition to an adequate level of water in their tissues, plants also require a continuous flux of water to perform vital processes such as photosynthesis and nutrient uptake. Water for these is not always available in the right quantity and quality at the right time. This imbalance in water 2

10 Ecology of Water Relations in Plants supply and plant requirements results in plants undergoing occasional or, in some cases acute, water stress. There are two types of water stresses that plants experience. One is when water is not available in sufficient quantity hence referred to as water-deficit, while the second one is that when water is available but in excess, called waterlogging. Water-deficit affects plants through decrease of leaf water potential, which in turn entails loss of cell turgor and stomatal closure. This results in decrease of transpiration and photosynthesis, which subsequently leads to reduced growth and if it persists, wilting. On the other hand, waterlogging occurs when a large proportion of the pore spaces in the soil are occupied by water. This means the diffusion of oxygen and gas exchange between the soil, plants and atmosphere is limited. The result of this is decreased root growth and functioning, which negatively affects plant growth and survival. Plants start suffering the consequences of water stress when certain thresholds for water-deficit and waterlogging are breached. Physiological plant studies have shown that soil water potentials approaching 5 kpa are sufficient to initiate plant stomatal closure, a classic response to water deficit (Hensen et al., 1989). On the other hand, waterlogging which cause 5 10% air-filled pore space in the soil (0% is achieved at soil saturation), result in hampering root activity, and hence induce aeration stress (Wesseling and van Wijk, 1957). Between these two thresholds of drying and waterlogging stress an optimal zone conducive to plant growth is achieved (Gowing et al., 2002). However, to be even more meaningful stress thresholds need to take account of time duration, over which the plant is subjected to the stress, i.e. as short periods of stress are less damaging than gentler but longer-term ones. One index that measures and cumulates the level of stress over the time duration it occurs, is called a sum exceedence value (SEV). SEVs are calculated separately for soil drying stress and for soil aeration stress, usually in unit of metre-weeks. SEVs were originally developed in the Netherlands by Sieben and colleagues in 1960s but later on successfully used in the UK by Gowing and colleagues (e.g. Silvertown et al., 1999) to integrate temporal variation in soil moisture at a scale relevant to the physiological response of plants. Moreover, SEVs take into account differences in soil type (as thresholds are specifically developed for each soil type under consideration) and are hence transferable between different sites. Plant Sensing and Adaptation to Water Stress Water stress is damaging to plants, so they have evolved a number of short-term responses as well as life history strategies that help them to cope. For this, a mechanism that senses water stress is crucial to the initiation of defensive processes. See also: Plant Stress Physiology Plant sensing of soil drying Water-deficit is the most common form of water stress studied in relation to sensing of impending soil drying by plant roots and the subsequent communication to shoots. In this connection, signals of a chemical nature have received a lot of attention, as they are suited for rapid communication between plant tissues. A well-known chemical signal of impending water stress originating from exposed roots is abscisic acid (ABA). ABA is synthesized by dehydrating roots in nongrowing tissues as well as in apices, and in the cortex (Hartung and Davies, 1991). An increase in ABA concentration in response to an increase in soil drying is known to initiate water-saving measures like reduction in transpiration rate and conductance (e.g. Hensen et al., 1989). A consequence of this sensing is that it determines the response of the plant and its competitive ability. For example, a plant which responds to the tiniest sign of stress will trade-off productivity for safety ( pessimist strategy ), while a plant that waits longer will trade safety for productivity ( optimist strategy ). Depending on the extent and duration of the actual stress, either of these two types of plants will emerge as the one having a better competitive advantage (Davies and Gowing, 1999). This will then influence, within the limits of physiological plasticity, their success in the plant community (see the section on Distribution of plants in response to water regime). Plant adaptations to water stress Plants respond to water stress in two ways: by avoidance of the stress or by tolerating it. Stress avoidance is accomplished when plants alter their growth schedule to escape the exposure to damaging stress. Well-known examples in this category include completing the life cycle while conditions are optimal, or using strategies to maximize water uptake from the environment and or conservation. On the other hand, the tolerance response to water stress occurs when plants develop certain characteristics, often of biochemical and or morphological nature, to minimize the potential damage from stress. An example of the latter could be additions to photosynthetic pathways such as crassulacean acid metabolism (Scott, 2000) for drying stress. Some morphological adaptations for flooding stress include development of air-space tissue (aerenchyma) within tissues and ventilation roots (pneumatophores). The development of the ability to metabolize products of anaerobic respiration and/or tolerate an accumulation of anaerobic metabolites is also another biochemical 3

11 Ecology of Water Relations in Plants 8 6 Carex nigra Carex riparia Carex hirta Carex disticha SEV aeration stress Carex panicea Carex flacca Carex acutiformis Carex acuta SEV drought stress Figure 4 Distributions of eight sedge species showing differentiation in niche space defined by hydrological axes on a fine-scale gradient. The x-axis depicts increasing soil drying stress, while the y-axis shows increasing flooding (i.e. aeration) stress (see the section on water stress and plants for explanation). The vertically hatched area in each graph shows the range of possible hydrological regimes and the solid area indicates the zone in which the species occurs significantly more frequently than by chance. Data are cumulated across 18 different meadow sites (after Gowing et al., 2002). Reproduced with permission from Her Majesty s Stationery Office. adaptation utilized by wetland plants. See also: Plant Response To Water-deficit Stress As a closing remark to this section, it is worth mentioning, an extreme form of adaptation to water-deficit by a group of plants known as poikilohydrics or resurrection plants. Poikilohydric plants show an ability of mature tissues such as the shoot, stem and leaves to tolerate almost complete dehydration of the tissues and then return as functional units very rapidly on rehydration, sometimes in as short as 24 h (Norwood et al., 2003). Obviously, such plants are native to and inhabit ecological niches that are subjected to lengthy periods of drought with brief periods of rain during the year, e.g. deserts of Southern Africa, Southern America and Western Australia (Scott, 2000). Ellenberg values, developed from field observations by the eminent German botanist, Heinz Ellenberg. However, only recently has the potent role of fine-scale heterogeneities in hydrology on a plot scale identified as principal driver for the defining structures in plant communities (Silvertown et al., 1999). These fine-scale differences in hydrological regime accomplish this structuring by creating realized niches, which are capable of being exploited by specific species. This is illustrated in Figure 4, with hydrological niches of eight species of sedge in UK wet meadows. The ecology of plant water relations thus can be elucidated by examination of the species hydrological niches within a community. These niches are a result of distinct plants differing physiological response to water stress, and the presence of other neighbouring plants. Distribution of Plants in Response to Water Regime Differences in water regime have been known to be behind existence of different vegetation types and ecosystems. Some widely known examples include global and regional distribution of plant communities. At global level examples include the major world biomes, such as tropical rainforest, deserts and tundra. At this level, precipitation differences as a result of latitude and incoming solar radiation define certain plants to prevail. A regional example is where plant communities are defined by precipitation differences associated with topographic features, such as elevation. A well known such example being a study as given in Figure 1.The distribution of plant species in relation to water regime at regional level, had also been examined using the subjective References Davies WJ and Gowing DJG (1999) Plant responses to small perturbations in soil water status. In: Press MC, Scholes JD and Barker MG (eds) Plant Physiological Ecology, Vol. 39, pp Oxford: Blackwell Science. Gowing D, Lawson C, Youngs E et al. (2002) The Water Regime Requirements and the Response to Hydrological Change of Grassland Plant Communities: DEFRA commissioned project BD1310, Final report to the Department for Environment, Food and Rural Affairs, Cranfield University, Silsoe. Hartung W and Davies WJ (1991) Drought-induced changes in physiology and ABA. In: Davies WJ and Jones HG (eds) Abscisic Acid Physiology and Biochemistry, pp Oxford: BIOS Science Publications Ltd. Hensen IE, Jensen CR and Turner NC (1989) Leaf gas exchange and water relations of lupins and wheat. I Shoot responses to soil water deficits. Australian Journal of Plant Physiology 16:

12 Ecology of Water Relations in Plants Norwood M, Toldi O, Richter A and Scott P (2003) Investigation into the ability of roots of the poikilohydric plant Craterostigma plantagineum to survive dehydration stress. Journal of Experimental Botany 54(391): Rodriguez-Iturbe I and Porporato A (2004) Ecohydrology of Water-Controlled Ecosystems: Soil Moisture and Plant Dynamics. Cambridge, UK: Cambridge University Press. Scott P (2000) Resurrection plants and the secrets of eternal leaf. Annals of Botany 85: Silvertown J, Dodd ME, Gowing DJG and Mountford JO (1999) Hydrologically defined niches reveal a basis for species richness in plant communities. Nature 400: Taiz L and Zeiger E (1998) Plant Physiology, 2nd edn. Sunderland, MA, USA: Sinauer Associates, Inc. Wesseling J and van Wijk WR (1957) In: Luthin JN (ed.) Drainage in Agricultural Lands, pp Madison, Wisconsin: American Society for Agronomy. Whittaker RH and Niering WA (1975) Vegetation of the Santa Catalina Mountains, Arizona. V. Biomass, production, and diversity along the elevation gradient. Ecology 56: Further Reading Archibold OW (1995) Ecology of World Vegetation. London, UK: Chapman & Hall. Ellenberg H (1988) Vegetation ecology of Central Europe, 4th edn. Cambridge, NY: Cambridge University Press. Etherington JR and William A (1976) Environment and Plant Ecology. London, UK: Wiley. Grime JP, Hodgson JG and Hunt R (1988) Comparative Plant Ecology: A Functional Approach to Common British Species. London, UK: Unwin Hyman. Lambers H, Stuart Chapin III F and Pons TL (1998) Plant Physiological Ecology. New York, USA: Springer New York, Inc. Proctor MCF and Tuba Z (2002) Poikilohydry and homoihydry: antithesis or spectrum of possibilities? New Phytologist 156: Smith JAC and Griffiths H (1993) Water Deficits: Plant Responses from Cell to Community. Oxford, UK: BIOS Scientific. 5

13 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Annu. Rev. Plant Biol : The Annual Review of Plant Biology is online at plant.annualreviews.org This article s doi: /annurev.arplant Copyright c 2008 by Annual Reviews. All rights reserved /08/ $20.00 Both authors contributed equally to this paper. Flooding Stress: Acclimations and Genetic Diversity J. Bailey-Serres 1, and L.A.C.J. Voesenek 2, 1 Center for Plant Cell Biology, University of California, Riverside, California 92521; serres@ucr.edu 2 Plant Ecophysiology, Institute of Environmental Biology, Utrecht University, NL-3584 CA Utrecht, The Netherlands Key Words aerenchyma, anoxia, response strategy, hypoxia, reactive oxygen species, submergence Abstract Flooding is an environmental stress for many natural and man-made ecosystems worldwide. Genetic diversity in the plant response to flooding includes alterations in architecture, metabolism, and elongation growth associated with a low O 2 escape strategy and an antithetical quiescence scheme that allows endurance of prolonged submergence. Flooding is frequently accompanied with a reduction of cellular O 2 content that is particularly severe when photosynthesis is limited or absent. This necessitates the production of ATP and regeneration of NAD + through anaerobic respiration. The examination of gene regulation and function in model systems provides insight into low-o 2 -sensing mechanisms and metabolic adjustments associated with controlled use of carbohydrate and ATP. At the developmental level, plants can escape the low-o 2 stress caused by flooding through multifaceted alterations in cellular and organ structure that promote access to and diffusion of O 2. These processes are driven by phytohormones, including ethylene, gibberellin, and abscisic acid. This exploration of natural variation in strategies that improve O 2 and carbohydrate status during flooding provides valuable resources for the improvement of crop endurance of an environmental adversity that is enhanced by global warming. 313

14 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Hypoxia (e.g., <20.9% and >0% O 2 at 20 C): characterized by increased anaerobic metabolism, increased ATP production via glycolysis owing to limited availability of O 2 for oxidative phosphorylation, and increased NAD + regeneration via lactate and ethanolic fermentation. Cellular ATP content is reduced Contents INTRODUCTION GENETIC DIVERSITY OF STRATEGIES TO SURVIVE FLOODING ACCLIMATION TO FLOODING AT THE CELLULAR LEVEL Overview of Cellular Adjustments to Oxygen Deprivation Low-Oxygen Sensing Management of the Energy Crisis THE LOW-OXYGEN ESCAPE SYNDROME Enhanced Growth Leading to the Emergence of Shoots Improvement of the Oxygen and Carbohydrate Status in Submerged Plants Improvement of Internal Gas Diffusion: Aerenchyma CONCLUSIONS AND FUTURE PERSPECTIVES INTRODUCTION Partial to complete flooding is detrimental for most terrestrial plants because it hampers growth and can result in premature death. Some plant species have a remarkable capacity to endure these conditions, and certain species can even grow vigorously in response to flooding. This interspecific variation has a strong impact on species abundance and distribution in flood-prone ecosystems worldwide (12, 31, 122, 138, 150, 151). Furthermore, flooding has a severe negative influence on the productivity of arable farmland because most crops are not selected to cope with flooding stress (121). The Intergovernmental Panel on Climate Change (IPCC) ( reported that the anthropogenically induced change of world climate increases the frequency of heavy precipitation and tropical cyclone activity. This is likely to engender more frequent flooding events in river flood plains and arable farmland, particularly affecting the world s poorest farmers (1). The observation that some plant species can cope with flooding stress and others cannot imposes the question of why a flooded environment is detrimental. The adversity is largely due to the dramatically reduced gas exchange between plants and their aerial environment during partial to complete submergence. Gases such as O 2,CO 2, and ethylene diffuse very slowly in water (46). Because of this tremendous barrier for gas diffusion, the cellular O 2 level can decline to concentrations that restrict aerobic respiration (39, 46). Depending on the tissue and light conditions, the cellular CO 2 level either increases in shoots in the dark and roots (47) or decreases in shoots in the light (83). The endogenous concentration of the gaseous plant hormone ethylene increases in tissues surrounded by water (59, 148). This accumulation activates adaptive signal transduction pathways, whereas similar concentrations hamper normal growth in many terrestrial plants (93). Furthermore, complete submergence decreases light intensity, dampening photosynthesis (141). A third major change in the flooded environment is the reduction of oxidized soil components to toxic concentrations (12). In summary, flooding is a compound stress in which the decline in molecular O 2 and thus the restriction of ATP synthesis and carbohydrate resources have major consequences for growth and survival. However, O 2 depletion is not the only active stress component, and often its impact is restricted to nonphotosynthesizing organs (84). O 2 shortage (hypoxia/anoxia) is not restricted to flooding stress. It is a frequent metabolic status of cells during normal development, particularly in tissues with high cell density, a high O 2 demand, and/or restricted O 2 entry, such as meristems, seeds, fruits, and storage organs (43). Fundamental insight into the low O 2 sensing mechanism, downstream signal transduction, and metabolic alterations that promote survival is key to increased crop 314 Bailey-Serres Voesenek

15 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org production in flood-prone environments and has wider implications for biologists (3, 43). Most studies on flooding stress have focused on relatively flood-tolerant species from genera such as Oryza, Rumex, and Echinochloa. Single species studies are valuable for an understanding of the regulation of various acclimations but less meaningful in an ecological perspective. Here genetic diversity in acclimations to flooding stress is discussed side by side with the molecular regulation of low-o 2 responses and flooding tolerance. Ultimately we aim to shed light on the genes, proteins, and processes controlling these phenotypes. GENETIC DIVERSITY OF STRATEGIES TO SURVIVE FLOODING Not all species in flood-prone environments are flood tolerant. Some species avoid flooding by completing their life cycle between Figure 1 Drained two subsequent flood events, whereas flooding periods are survived by dormant life stages [e.g., Chenopodium rubrum thrives in frequently flooded environments by timing its growth between floods and producing seeds that survive flooding (134)]. Established plants also use avoidance strategies through the development of anatomical and morphological traits. This amelioration response, here called the low oxygen escape syndrome (LOES) (Figure 1), facilitates the survival of submerged organs. Upon complete submergence several species from flood-prone environments have the capacity to stimulate the elongation rate of petioles, stems, or leaves. This fast elongation can restore contact between leaves and air but can also result in plant death if energy reserves are depleted before emergence. Concomitant with high elongation rates, the leaves also develop a thinner overall morphology, develop thinner cell walls and cuticles, and reorient Submerged Trait Low High Shoot elongation Low High Aerenchyma High Low Leaf thickness Around intercellular spaces Toward epidermis Chloroplast position Various species display the low-oxygen escape syndrome (LOES) when submerged. The syndrome includes enhanced elongation of internodes and petioles, the formation of aerenchyma in these organs (air spaces indicated by arrows labeled a), and increased gas exchange with the water layer through reduced leaf thickness and chloroplasts that lie directed toward the epidermis (indicated by arrows labeled b). Photographs are courtesy of Ronald Pierik, Liesje Mommer, Mieke Wolters-Arts, and Ankie Ammerlaan. a b Anoxia (e.g., 0% O 2 at 20 C): characterized by anaerobic metabolism, NAD + regeneration via lactate and ethanolic fermentation, and ATP production solely via glycolysis (2 4 mol ATP per mole hexose). Cellular ATP content is low, and ADP content is elevated LOES: low-oxygen escape syndrome Acclimation to Flooding Stress 315

16 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Sub1: Submergence1 polygenic locus of rice; determines submergence tolerance Oxygen deficiency/ deprivation: the natural and experimental conditions in which cellular oxygen content is reduced but metabolic status is not determined mtetc: mitochondrial electron transport chain ROS: reactive oxygen species chloroplasts toward the leaf surface. These traits reduce the resistance for diffusion of CO 2 and O 2, facilitating inward diffusion and thereby improving underwater photosynthesis and aerobic metabolism (82, 83). Thus, the LOES improves the aeration of the plant, which is further enhanced by the relatively low resistance for internal gas diffusion owing to a system of interconnected gas conduits called aerenchyma, a property typical of many wetland plants (24, 33). These conduits are constitutive, induced in existing tissues (roots, petioles, stems) (33) or formed during the development of adventitious roots that arise from the root shoot junction or stem nodes (115, 142). In specialized cases the longitudinal diffusion of O 2 to the root apex is further enhanced by the development of a barrier to radial oxygen loss to minimize escape of O 2 to the surrounding environment (24, 25). LOES is costly and will only be selected for in environments where the cost is outweighed by benefits such as improved O 2 and carbohydrate status, both contributing to a higher fitness (120). The flooding regime is an important determinant for selection in favor of or against LOES. A study on the distribution of species in the Rhine floodplains confirmed this hypothesis. Here LOES occurs predominantly in species from habitats characterized by prolonged, but relatively shallow, flooding events (150). However, the benefits of LOES do not outweigh the costs when the floods are too deep or ephemeral. These regimes favor a quiescence strategy characterized by limited underwater growth and conservation of energy and carbohydrates (39, 91). This strategy is a true tolerance mechanism, driven by adjustment of metabolism. With respect to low-o 2 stress, this includes the downregulation of respiration and limited stimulation of fermentation to create a positive energy budget when organ hypoxia starts (43, 148). The SUB1A gene of the polygenic rice (Oryza sativa L.) Submergence1 (Sub1) locus was shown to confer submergence tolerance through a quiescence strategy in which cell elongation and carbohydrate metabolism is repressed (41, 91, 159) (Figure 2). SUB1A, encodes an ethylene-responsive element (ERF) domain containing transcription factor (41). The lack of SUB1A-1 or the presence of a slightly modified allele is associated with reduced submergence tolerance and the induction of the LOES. This example demonstrates that environment-driven selection on a single locus can significantly alter survival strategy. ACCLIMATION TO FLOODING AT THE CELLULAR LEVEL Overview of Cellular Adjustments to Oxygen Deprivation During flooding, the onset of O 2 deprivation is rapid in the dark and in nonphotosynthetic cells. The reduced availability of O 2 as the final electron acceptor in the mitochondrial electron transport chain (mtetc) mediates a rapid reduction of the cellular ATP:ADP ratio and adenylate energy charge (AEC) ([ATP ADP]/[ATP+ADP+AMP]) (46). Cells cope with this energy crisis by relying primarily on glycolysis and fermentation to generate ATP and regenerate NAD +, respectively. Whether a LOES or a quiescence response to flooding is activated, cellular acclimation to transient O 2 deprivation requires tight regulation of ATP production and consumption, limited acidification of the cytosol, and amelioration of reactive oxygen species (ROS) produced either as O 2 levels fall during flooding or upon reoxygenation after withdrawal of the flood water. O 2 concentration is 20.95% at 20 C in air but ranges from 1 to 7% in the core of wellaerated roots, stems, tubers, and developing seeds (14, 44, 46, 107, 136, 137). Within a root, O 2 levels and consumption vary zonally; the highly metabolically active meristematic cells are in a continuous state of deficiency. Upon flooding, the 10,000-fold-slower diffusion of O 2 in water rapidly limits its availability for mitochondrial respiration. This deprivation is progressively more severe as 316 Bailey-Serres Voesenek

17 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Tolerant Lowland Rice Intolerant Deepwater Rice Strategy Quiescence LOES LOES Sub1 haplotype Carbohydrate consumption Fermentation capacity SUB1A-1, SUB1B, SUB1C Limited by SUB1A-1 High SUB1B, SUB1C or SUB1A-2, SUB1B, SUB1C High Moderate GA response Inhibited by SUB1A-1 Promoted by SUB1C Figure 2 Rice responds via different strategies to submergence. Flood-tolerant rice varieties invoke a quiescence strategy that is governed by the polygenic Submergence1 (Sub1) locus that encodes two or three ethylene-responsive factor proteins (41, 159). SUB1A is induced by ethylene under submergence and negatively regulates SUB1C mrna levels. Flood-intolerant varieties avoid submergence via the low oxygen escape syndrome (LOES). To this end SUB1C expression is promoted by gibberellic acid (GA) and is associated with rapid depletion of carbohydrate reserves and enhanced elongation of leaves and internodes. The LOES is unsuccessful when flooding is ephemeral and deep. Deepwater rice varieties survive flooding via a LOES, as long as the rise in depth is sufficiently gradual to allow aerial tissue to escape submergence (61). N.D., not determined. distance from the source increases and tissue porosity decreases. For example, the cortex of nonaerenchymatous maize (Zea mays L.) roots exposed to 10% O 2 becomes hypoxic, whereas the internal stele becomes anoxic. Even the apex of aerenchymatous roots encounters severe O 2 deprivation (46). In dense storage organs such as potato (Solanum tuberosum L.) tubers and developing plant seeds, exposure to 8% O 2 significantly reduces the endogenous O 2 level. However, the decrease in cellular O 2 is strikingly nonlinear from the exterior to the interior of the organ; cells at the interior of the tuber or endosperm maintain a hypoxic state (44, 136). This has led to the suggestion that an active mechanism may allow cells to avoid anoxia (43). Such a mechanism may include proactive limitation in the consumption of both ATP and O 2. The low K m for cytochrome c oxidase (COX) [140 nm ( 0.013%) O 2 ] should ensure that the activity of COX continues as long as O 2 is available (31, 46). However, a mechanism that inhibits the mtetc at or upstream of COX or inhibits SUB1B, SUB1C or SUB1A-2, SUB1B, SUB1C High N.D. High Acclimation to Flooding Stress 317

18 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Normoxia (e.g., 20.9% O 2 at 20 C): characterized by aerobic metabolism, NAD + regeneration primarily via the mitochondrial electron transport chain, and ATP production via mitochondrial oxidative phsophorylation (30 36 mol ATP per mol hexose consumed); cellular ATP content is normal O 2 consumption by other enzymes may allow cells to sustain hypoxia and avoid death. Low-Oxygen Sensing In animals the perception of O 2 deficit involves O 2 -binding proteins, ROS, and mitochondria. The O 2 -consuming prolyl hydroxylases (PHDs) are direct sensors of O 2 availability. Under normoxia, PHDs target the proteosomal degradation of hypoxia inducible factor 1α (HIF1α), a subunit of a heterodimeric transcription factor that regulates acclimation to hypoxia (51). The concomitant drop in PHD activity stimulates an elevation in HIF1α as O 2 declines. A paradox is that the production of ROS at the mitochondrial ubiquinone:cytochrome c reductase complex (Complex III) is necessary to initiate O 2 deficit responses (7, 51). There is limited understanding of the mechanisms by which plant cells sense and initiate signaling in response to O 2 deficit (3, 39, 43). Plants lack a HIF1α ortholog, although PHD mrnas are strongly induced by O 2 deficit in Arabidopsis thaliana and rice (67, 146). Furthermore, significant increases in mrnas encoding enzymes involved in ROS signaling and amelioration (16, 63, 67, 70, 71) and evidence of ROS production have been reported in several species upon transfer to low O 2 conditions. A challenge in monitoring ROS production during O 2 deficit is that ROS are produced readily upon reoxygenation. However, ethane, a product of membrane peroxidation by ROS, evolves from submerged rice seedlings in a closed system as levels of O 2 fall to as low as 1% (112), providing evidence that ROS form as O 2 levels decline. Blokhina and colleagues (11) demonstrated that in response to anoxia, H 2 O 2 accumulates to higher levels in the apoplast of root meristems of hypoxic wheat (Triticum aestivum) than in the more anoxia-tolerant rhizomes of Iris pseudacorus. InArabidopsis seedlings, H 2 O 2 levels increase in response to O 2 deprivation in a ROP GTPase dependent manner (6). Genotypes that limit ROP signaling under hypoxia display lower levels of H 2 O 2 accumulation and altered gene regulation in stressed seedlings. Indications that mitochondria are crucial to low-o 2 sensing in plants comes from the release of Ca 2+ from mitochondria of cultured maize cells within minutes of transfer to anoxia (127). This release may be activated by mitochondrial ROS production at Complex III of the mtetc (99). A rapid spike in cytosolic Ca 2+ was also observed in the cotyledons of Arabidopsis seedlings upon transfer to anoxia and again at higher amplitude upon reoxygenation (119). These Ca 2+ transients are required for alterations in gene expression that enhance ethanolic fermentation and ATP management during the stress (3, 66, 88, 119, 126, 128). Further studies are needed to confirm whether mitochondrionto-nucleus signaling, mediated by ROS production and Ca 2+ release from mitochondria, contributes to reconfiguration of metabolism under low O 2. Additional players in the acclimation response may be the reduction of ATP content and decline in cytosolic ph as well as change in levels of metabolites such as sucrose and pyruvate (3, 39). mrnas encoding mitochondrial alternative oxidase, (AOX) are strongly induced by low-o 2 stress (63, 67, 70, 71). AOX diverts ubiquinone from Complex III; if active as O 2 levels decrease, AOX would paradoxically reduce oxygen availability for COX and decrease ATP production. However, if active as O 2 levels rise upon reoxygenation, AOX may limit mitochondrial ROS production (99). Management of the Energy Crisis Within minutes of transfer to an O 2 -depleted environment, cells reliant on external O 2 limit processes that are highly energy consumptive and alter metabolism to increase anaerobic generation of ATP by cytosolic glycolysis (31). This shift is followed by fermentation of pyruvate to the major end products, ethanol or lactate, yielding NAD + to sustain anaerobic metabolism (Figure 3). A crisis in ATP availability ensues because glycolysis is inefficient, 318 Bailey-Serres Voesenek

19 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org yielding 2 to 4 mol ATP per mol hexose as compared with 30 to 36 mol ATP by the mtetc. Evaluation of gene transcripts, enzymes, and metabolites in a variety of species and genotypes demonstrated the production of minor metabolic end products that are also important for NAD + and NAD(P) + regeneration. Although mutant analyses with several species have demonstrated that glycolysis and fermentation are necessary for cell survival under O 2 deprivation, the enhancement of these processes is not well correlated with prolonged endurance of this stress (31, 46). The anaerobic energy crisis necessitates a blend of optimized ATP production with limited energy consumption. ATP-demanding processes such as DNA synthesis and cell division are curtailed (46), and the production of rrna is dramatically reduced (36). In Arabidopsis and other plants, low-o 2 stress markedly limits protein synthesis but maintains the initiation of translation of a subset of cellular mrnas, many of which encode enzymes involved in anaerobic metabolism and the amelioration of ROS (16, 36). Therefore, under O 2 deprivation, a mechanism operates that sequesters untranslated mrnas and lessens ATP expenditure, thereby allowing for the recovery of protein synthesis within minutes of reoxygenation. Carbohydrate mobilization and sucrose catabolism. The metabolic response to O 2 deprivation is orchestrated by the availability and mobilization of carbohydrates (31, 137). In some plants and tissues, the induction of amylases by low O 2 or flooding promotes the conversion of starch to glucose (Figure 3). However, the mobilization of starch during O 2 deprivation is not universal. Both the tubers of potatoes and rhizomes of the flood-tolerant marsh plant Acorus calamus L. have considerable carbohydrate reserves, but Acorus rhizomes are more capable of mobilizing starch into respirable sugars under anoxia (2). This slow consumption of starch allows the rhizomes to sustain a low level of metabolism that affords survival of long periods of submergence. Seeds of rice, rice weeds (e.g., some Echinochloa species), and tubers of Potamogeton pectinatus also mobilize starch under anoxia (29, 40, 50). In rice seeds, this starch mobilization requires the depletion of soluble carbohydrates, suggesting regulation by sugar sensing (50, 72). In organs lacking starch reserves or effective starch mobilization, the exhaustion of soluble sugars prior to reoxygenation is likely to result in cell death. Plants possess two independent routes for the catabolism of sucrose, the bidirectional UDP-dependent sucrose synthase (SUS) and the unidirectional invertase (INV) pathways (Figure 3). The net cost for entry into glycolysis is one mol pyrophosphate (PPi) per mol sucrose via the SUS route, if the UTP produced by UDP-glucose pyrophosphorylase (UGPPase) is utilized by fructokinase (FK) in the subsequent conversion of UDPglucose to glucose-6p or the ATP consumed by FK is recycled by nucleoside diphosphate (NDP) kinase. By contrast, the cost via the INV pathway is two mol ATP per mol sucrose. The SUS route is positively regulated under O 2 deprivation through opposing increases in SUS and the repression of INV gene expression and enzymatic activity (10, 14, 43, 44, 64, 67). The energetic disadvantage of the INV route was confirmed by the inability of transgenic potato tubers with elevated INV activity to maintain ATP levels under 8% O 2 (14). The SUS pathway is enhanced in a variety of species by rapid increases in transcription of SUS mrnas, which is most likely driven by sucrose starvation (64, 71). Other glycolytic reactions may utilize available PPi during O 2 deprivation, thereby improving the net yield of ATP per mol sucrose catabolized. The phosphorylation of fructose- 6P to fructose-1,6p 2 by the bidirectional PPi-dependent phosphofructokinase (PFP) is favored over the unidirectional ATPdependent phosphofructokinase (PFK), and a pyruvate Pi dikinase (PPDK) may substitute for cytosolic pyruvate kinase (PK) in O 2 - deprived rice seedlings (95). Acclimation to Flooding Stress 319

20 H 2 O Sucrose UDP Starch Invertase inhibitor Invertase SUS amylases Glucose Pi Starch Pase Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org UDP Fructose + Glucose UTP Lactate ADH Ethanol ATP ADP NDP kinase LDH NAD + PK HXK FK PFK ALDH Glucose-6P Fructose-6P ATP PPi PFP ADP Pi Glycolysis PDC NADH NADH NADH Glutamate 2-oxyglutarate Aspartate UDP-glucose + Fructose PPi Glucose-1P PGI PGM Fructose-1,6P 2 NAD + +CoASH UGPPase FK HCO PEPC 3 Pi Phosphoenolpyruvate PCK ADP AMP+PPi PPDK ATP ATP+Pi PDH Pyruvate CO 2 Acetaldehyde NAD + NADH ADP ATP ATP ADP UTP AlaAT NAD + NADH NAD + UDP x4 AspAT UTP CO 2 Alanine H 2 O Acetyl CoA H + CoASH NR NiR Glucose-1P NH 4 + GABA CO 2 GS GDC Glutamine Glutamate GOGAT GDH Citrate Aconitase Isocitrate 2-oxyglutarate CS ICDH CO Oxaloacetate acetate 2 2-oxyglutarate NAD MDH TCA cycle + NAD 2-OGDH + Malate CO Succinyl CoA 2 Fumarase Succinyl CoA Fumarate syn ADP SDH ATP FADH 2 Succinate NADH NAD(P) + FAD + γ-hydroxybutyrate ATP ADP NAD(P) + NAD(P)H NADH NAD + Succinate semialdehyde NAD(P)H GHBDH NO 3 NO 2 SSADH Glutamate NAD(P)H NAD(P) + NAD + NADH NH 4 + NADH 2-oxyglutarate GABA Alanine Pyruvate GABA-T NAD(P) + NAD(P)H 320 Bailey-Serres Voesenek

21 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Metabolic end products. During O 2 deprivation, pyruvate decarboxylase (PDC) converts pyruvate to acetaldehyde, which is metabolized by alcohol dehydrogenase (ADH) to ethanol, with the regeneration of NAD + to sustain glycolysis. PDC- and ADH-deficient genotypes confirm the essentiality of ethanolic fermentation in the acclimation to flooding and low-o 2 stress (6, 31, 46, 65). In Arabidopsis seedlings, the level of induction of ADH is controlled by the activation of a ROP GTPase (5). O 2 deprivation promotes an increase in active ROP, which leads to the elevation of transcripts that encode ADH and ROPGAP4, a GTPase that inactivates ROP. In a ropgap4 null mutant, ADH mrna and ROS are significantly elevated under hypoxia, and seedling survival is reduced. This led to the proposal that a ROP rheostat controls the temporal regulation of ADH expression under low O 2 (3, 39). The production of ethanol is benign owing to its rapid diffusion out of cells, whereas the intermediate acetaldehyde is toxic. Acetaldehyde dehydrogenase (ALDH) catalyzes the conversion of acetaldehyde to acetate, with the concomitant reduction of NAD + to NADH. A mitochondrial ALDH is significantly induced by anoxia in coleoptiles of rice (67, 87), but not in seedlings of Arabidopsis (65). ALDH activity correlates with anaerobic germination capability of Echinochloa crus-galli under strict anoxia (40). Under O 2 -limiting conditions, ALDH consumes NAD + and may thereby limit glycolysis, whereas upon reoxygenation acetaldehyde converted to acetate by mitochondrial ALDH enters the tricarboxylic acid (TCA) cycle (Figure 3). In addition to ethanol, lactate is produced in plant cells under O 2 deprivation. The accumulation of lactate under low-o 2 stress has garnered considerable interest (31, 35, 48, 98) ever since the demonstration that its transient appearance precedes that of ethanol in the root tips of maize seedlings (105). The ph of the cytosol of maize root tips declines from 7.5 to a new equilibrium at ph 6.8 followingtransfer to anoxia. It is posited that the transition from lactic to ethanolic fermentation is controlled by a ph-stat. The 0.6 unit decrease in cytosolic ph favors the catalytic optimum of PDC and thereby limits lactate and promotes ethanol production. Anoxic ADHdeficient root tips continue to produce lactate and fail to stabilize the cytosolic ph, resulting in rapid cytosolic acidification and cell death (106). Thus, the switch from lactic to ethanolic fermentation is critical for the maintenance of cytosolic ph. An alternative proposal is that this switch, under conditions of O 2 deprivation and in aerobic cells in which ethanol is produced, is driven by a rise in pyruvate rather than the increase in lactate or reduction of cytosolic ph (130). When Figure 3 Metabolic acclimations under O 2 deprivation. Plants have multiple routes of sucrose catabolism, ATP production, and NAD + and NAD(P) + regeneration. Blue arrows indicate reactions that are promoted during the stress. Metabolites indicated in bold font are major or minor end products of metabolism under hypoxia. Abbreviations are as follows: 2-OGDH, 2-oxyglutarate dehydrogenase; ADH, alcohol dehydrogenase; AlaAT, alanine aminotransferase; ALDH, acetaldehyde dehydrogenase; AspAT, aspartate aminotransferase; CoASH, coenzyme A; CS, citrate synthase; FK, fructokinase; GABA-T, GABA transaminase; GDC, glutamate decarboxylase; GDH, glutamate dehydrogenase; GHBDH, γ-aminobutyrase dehydrogenase; GOGAT, NADPH-dependent glutamine: 2-oxoglutarate aminotransferase; GS, glutamine synthase; HXK, hexokinase; ICDH, isocitrate dehydrogenase; LDH, lactate dehydrogenase; MDH, malate dehydrogenase; NDP kinase, nucleoside diphosphate kinase; NiR, nitrite reductase; NR, nitrate reductase; PCK, phosphenolpyruvate carboxylase kinase; PDC, pyruvate decarboxylase; PDH, pyruvate dehydrogenase; PEPC, phosphenolpyruvate carboxylase; PFK, ATP-dependent phosphofructokinase; PFP, PPi-dependent phosphofructokinase; PGI, phosphoglucoisomerase; PGM, phosphoglucomutase; PK, pyruvate kinase; PPDK, pyruvate Pi dikinase; SDH, succinate dehydrogenase; SSADH, succinate semialdehyde dehydrogenase; Starch Pase, starch phosphorylase; SUS, sucrose synthase; UGPPase, UDP-glucose pyrophosphorylase. Acclimation to Flooding Stress 321

22 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org pyruvate levels increase, the low K m of mitochondrial pyruvate dehydrogenase (PDH) and high K m of PDC serve to limit carbon entry into the TCA cycle and promote ethanolic fermentation. Flooding stress is likely to involve a gradual transition from normoxia to hypoxia, allowing cells to initiate processes that favor survival. Plants exposed to a period of hypoxia for 2 to 4 h prior to transfer to an anoxic environment are more capable of avoiding cell death than those that undergo an abrupt anoxic shock (31). The preexposure to 3% or 4% O 2 reduces the severity of ATP depletion, allows the synthesis of stress-induced and normal cellular proteins (19), and activates a lactate efflux mechanism (158). Lactate removal from the cytoplasm may be accomplished by the hypoxia-induced nodulin intrinsic protein (NIP2;1), which was identified in Arabidopsis as a plasma membrane associated protein capable of driving lactate transport in Xenopus oocytes (23). Most likely, a decline in cytosolic ph of 0.2 to 0.5 units under O 2 shortfall establishes a new ph set point that influences multiple aspects of metabolism (35, 48, 95). The management of this ph decline involves ethanolic fermentation and is benefited by the availability of a lactate efflux mechanism and proton ATPase activity. However, some species or organ systems, such as the tuber shoots of Potamogeton pectinatus, do not show an adjustment in cytosolic ph during O 2 deprivation. The stem elongation in these shoots under anoxia results from cell expansion that occurs in the absence of an adjustment in cytosolic ph and appears to be maintained by tight constraints on ATP production and consumption (29). Besides the major fermentation end products, lactate and pyruvate, O 2 deficiency is associated with the elevation of alanine, γ- aminobutyric acid (GABA), succinate, and occasionally malate (29, 31, 46, 113, 137, 139). Strong induction of cytosolic and mitochondrial alanine aminotransferase (AlaAT), aspartate aminotransferase, mitochondrial glutamate dehydrogenase (GDH), and mitochondrial Ca 2+ /calmodulin-regulated glutamate decarboxylase (GDH) mrna and/or enzymatic activity is consistent with pyruvate conversion to alanine or GABA (Figure 3) (63, 67, 70, 71, 100, 139). GABA may be further metabolized via the mitochondrial GABA shunt to γ-hydroxylbutyrate with the regeneration of NAD(P) + (17). Upon reoxygenation, alanine can be recycled back to pyruvate, and GABA can be converted to succinate. Amino acid oxidation may thereby minimize the decline in cytosolic ph and reduce carbon loss via ethanol or lactate. An appreciation of the relative significance of the major and minor pathways of anaerobic metabolism will require metabolite profiling and flux studies that resolve organ specific and temporal aspects of production in relationship to changes in redox and energy status. Nitrite, nitric oxide, mitochondria, and hemoglobin. Nitrate and nitrite are also implicated in cellular adjustment to O 2 deprivation. Nitrate is assimilated and reduced to ammonia via nitrate reductase (NR) and nitrite reductase (NiR) (Figure 3). NR but not NiR mrnas increase significantly in response to hypoxia/anoxia in Arabidopsis and rice (67, 70, 71). Even without an increase in NR levels, a reduction of cytosolic ph may increase nitrite production because of the low ph optimum of this enzyme (57). Roots of tobacco plants engineered to have reduced NR levels display several metabolic anomalies under anoxia, including higher levels of soluble hexoses and ATP, enhanced ethanol and lactate production, and increased acidification of the cytosol (125). By contrast, maize seedling roots supplied with nitrate during anoxia maintain a slightly higher cytosolic ph than do control seedlings (69). Notably, the provision of micromolar levels of nitrite to seedling roots had a similar effect on the adjustment of cytosolic ph. This unexpected benefit of low levels of nitrite is unlikely to be due to a direct effect on NAD(P) + regeneration and may indicate a role of nitrite in a regulatory mechanism that augments homeostasis under low O Bailey-Serres Voesenek

23 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org A plant-specific association has surfaced between nitrate/nitrite metabolism, mitochondrial ATP synthesis, and a low-o 2 - induced nonsymbiotic Class 1 hemoglobin (HB). Plant mitochondria provided with micromolar levels of nitrite under anoxia have the capacity to coordinate the oxidation of NADH and NAD(P)H with low levels of ATP production (124). This nitritepromoted process involves the evolution of nitric oxide (NO) via a pathway that requires the activity of rotenone-insensitive NAD(P)H dehydrogenases, mtetc Complex III (ubiquinone:cytochrome c reductase), and Complex IV (COX). In the proposed pathway (124), NAD(P)H produced during O 2 deficit is oxidized by Ca 2+ -sensitive NAD(P)H dehydrogenases on the inner mitochondrial membrane surface, providing electrons to the ubiquinone pool. In the absence of O 2, nitrite may serve as an electron acceptor at Complex III or IV, yielding NO, which may activate signal transduction by promoting mitochondrial ROS production and Ca 2+ release. The cytosolic HB that accumulates under O 2 deprivation, however, scavenges and detoxifies NO in planta by converting it into nitrate in an NAD(P)H-consuming reaction over a broad ph optimum (30, 57). The coupled activities of HB and cytosolic NR regenerate nitrite that may enter the mitochondrion, where it continues the cycle of NO and ATP production (94, 124). A major challenge is to confirm in planta that nitrite conversion to NO functions as a surrogate final electron acceptor. Nonetheless, the scenario is consistent with reports that overexpression of HB in several species decreases rates of ethanolic fermentation, augments ATP maintenance, and fosters NO production under hypoxia. By contrast, the inhibition of HB expression increases NAD(P)H:NAD(P) + ratios and reduces cytosolic ph (30, 56, 57, 123). Notably, NO inhibits COX activity and thereby reduces ATP production under normoxia. Might NO formed during the transition from normoxia to hypoxia be the factor that dampens O 2 consumption to avoid cellular anoxia (43)? If so, the production of NO prior to the synthesis of HB may allow the cell to transition slowly from normoxia to hypoxia, providing a segue that augments energy management. THE LOW-OXYGEN ESCAPE SYNDROME Enhanced Growth Leading to the Emergence of Shoots Plants forage for limiting resources by adjusting carbon allocation and overall plant architecture such that the capture of resources is consolidated (93, 96). As O 2 and CO 2 become limiting for plants in flood-prone environments, species from widely dispersed families that share the capacity to survive in flood-intense environments initiate signaling pathways that lead to fast extension growth of shoot organs (101, 147). These leaves, when reaching the water surface, function as snorkels to facilitate the entrance of O 2 and the outward ventilation of gases such as ethylene and methane trapped in roots (24, 145). Another benefit of the emergence of leaf blades is a higher rate of carbon gain from aerial photosynthesis (82). Fast shoot elongation under water is not restricted only to species occurring in environments with periodic floods (e.g., deepwater rice, Rumex palustris, Ranunculus sceleratus) (61, 148, 150). It persists in true aquatics that develop floating leaves or flowers [e.g., Nymphoides peltata (82)] and in species that germinate in anaerobic mud followed by an extension growth phase to reach better-aerated water/air layers (e.g., seedlings of Oryza sativa, Potamogeton pectinatus, P. distinctus) (58, 113, 129). The explored mechanism of shoot elongation in Marsh dock (R. palustris) and deepwater rice (92, 115, 149) can be used to shed light on the mechanistic backbone of genetic diversity in flooding-induced shoot elongation. The shoot elongation response can occur in petioles or internodes, depending on the Acclimation to Flooding Stress 323

24 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org developmental stage or predominant growth form of the plant. Interestingly, petiole elongation in rosette plants is accompanied by hyponastic growth that changes the orientation of the petiole from prostrate to erect. This directional growth brings the leaf in such a position that enhanced petiole elongation will result in leaf blade emergence in the shortest possible time. Accordingly, petiole elongation lagged behind hyponastic growth in R. palustris rosettes (26). It is generally accepted that the submergence signal for enhanced shoot elongation is the gaseous phytohormone ethylene (76, 101, 147). Ethylene is biosynthesized via an O 2 - dependent pathway, and the endogenous concentration of this hormone is determined predominantly by production rate and outward diffusion. Both aspects are affected by submergence. Several biosynthetic genes [e.g., those encoding ACC synthase (ACS) and ACC oxidase (ACO)] are upregulated by submergence (102, 135, 154), whereas diffusion of ethylene to the outside environment is strongly hampered. As a result, the endogenous concentration rises to a new, higher equilibrium. Ethylene production persists in submerged shoots as O 2 continues to diffuse from the water into the shoot, guaranteeing relatively high endogenous O 2 concentrations in shoot cells even in the dark (80). Submergence or low oxygen also upregulates the expression of ethylene receptor genes, including RpERS1 in R. palustris (155), OsERL1 in deepwater rice (156), and ETR2 in Arabidopsis (16, 63, 70, 72). An elevation of ethylene receptor levels following submergence is counterintuitive because these molecules are negative regulators of ethylene signaling. However, this increase would allow rapid cessation of ethylene signaling as the plants emerge from the water and vent off the accumulated ethylene. Ethylene is the input signal for several parallel pathways required for fast elongation under water (Figure 4). Under fully submerged conditions the accumulated ethylene downregulates abscisic acid (ABA) levels via an inhibition of 9-cis-epoxycarotenoid dioxygenase (NCED) expression, a family of ratelimiting enzymes in ABA biosynthesis that belongs to the carotenoid cleavage dioxygenases, (CCDs) and via an activation of ABA breakdown to phaseic acid (9, 61, 110). The decline of the endogenous ABA concentration in R. palustris is required to stimulate the expression of gibberellin (GA) 3-oxidase, an enzyme that catalyzes the conversion to bioactive gibberellin (GA 1 ) (8), and in deepwater rice to sensitize internodes to GA (61). Downstream of GA, three sets of genes play a role in submergence-induced shoot elongation. The first group encodes proteins involved in cell wall loosening; the second, those involved in the cell cycle; and the third, those involved in starch breakdown. Additional genes with putative regulatory roles in enhanced internode elongation have been identified in flooded deepwater rice (22, 108, 117, 132, 133). The rigid cell wall constrains the rate and direction of turgor-driven cell growth. Significant increases in acid-induced cell wall extension upon submergence were observed in rice (20), R. palustris (152), and Regnellidium diphyllum (62). This could be reversed even when R. palustris petioles were desubmerged, emphasizing the correlation between extensibility and submergence-induced elongation (152). Cell wall extensibility is thought to be associated with cell-wall-loosening proteins, such as expansins (EXPs) and xyloglucan endotransglycosylase/hydrolases (XTHs) (27). Submergence-induced elongation is strongly correlated with increases in mrnas encoding expansins A (EXPA) and B (EXPB), along with EXP protein abundance and activity (21, 62, 68, 89, 152, 153). Interestingly, in some species ethylene directly regulates EXP expression (62, 152, 153) (Figure 4). In submerged R. palustris petioles, ethylene not only enhances EXP expression but also stimulates proton efflux into the apoplast (153), which is essential for EXP action. The second group of GA-regulated genes is involved in cell cycle regulation. In very young petioles of the fringed waterlily (Nymphoides peltata) and the youngest 324 Bailey-Serres Voesenek

25 Submergence OsACS1 OsACS2 OsACS5 RpACS1 OsACO1 RpACO1 RpERS1 Ethylene RpEXPA1 RdEXPA1 RpNCED1-4 RpNCED6-10 OsABA8ox1 OsUSP1 ABA RpGA3OX1 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Figure 4 Apoplastic acidification OsSUB1A cyc2os1 cyc2os2 cdc2os2 HistoneH3 OsRPA1 OsTMK OsGRF1-3 OsGRF7,8,10,12 OsDD3-4 OsSBF1 OsGRF9 GA Cell elongation and/or division OsEXPA2 OsEXPA4 OSEXPB3 OsEXPB4 OsEXPB6 OsEXPB11 OsXTR1 OsXTR3 OsSUB1C OsAMY3D Schematic model of the plant processes, hormones, and genes involved in submergence-induced shoot elongation (blue signifies upregulated genes and red signifies downregulated genes). Gene abbreviations are as follows: CYC2Os, cyclin; CDC2Os, cyclin-dependent kinase; OsACO and RpACO, ACC oxidase; OsACS and RpACS, ACC synthase; OsDD, differentially displayed (61); OsAMY, amylase (41); OsEXP, RdEXP, and RpEXP, expansins; OsGRF, growth-regulating factor (22); OsRPA, replication protein A1; OsSBF, sodium/bile acid symporter family (108); OsSUB1, submergence1; OsTMK, transmembrane protein kinase (133); OsUSP, universal stress protein (117); RpERS1, ethylene receptor (155); RpNCED, 9-cis-epoxycarotenoid dioxygenase; RpGA3ox, gibberellin 3-oxidase (8); OsXTR, xyloglucan endotransglucosylase-related (27); OsABA8ox, ABA 8 -hydroxylase (110). Os indicates Oryza sativa, Rd indicates Regnellidium diphyllum, and Rp indicates Rumex palustris. internode of deepwater rice, ethylene promotes not only cell elongation but also cell division. Consistent with this increase in cell division is the observed upregulation of cyclin (CYC2Os1, CYC2Os2), cyclin-dependent kinase (CDC2Os2), HistoneH3, and replication protein A1 (OsRPA1) (114, 115, 131). The third group of GA-regulated genes is involved in starch breakdown. R. palustris plants depleted of soluble sugars and starch show a very restricted underwater elongation response (49). Carbohydrates are required to deliver energy and the building blocks for new cell wall synthesis (115, 148). The requirement of carbohydrates can be fulfilled by the translocation of photosynthates and by the degradation of starch reserves via an increase in α-amylase activity (115). Fukao and colleagues (41) reported that α-amylase gene expression (OsAmy3D) in leaves of submerged Acclimation to Flooding Stress 325

26 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org rice is regulated by SUB1C, an ethyleneresponsive factor (ERF)-domain-containing protein of the polygenic Sub1 locus. This gene is regulated positively by GA and negatively by a related ERF in the Sub1 locus, SUB1A- 1, which is present in some rice accessions. These results imply that carbohydrate levels in submerged plants are also under hormonal control. There is considerable genetic variation between and within species in submergenceinduced elongation capacity. The closely related species Rumex acetosa and R. palustris show inhibition and stimulation of petiole elongation upon exposure to ethylene, respectively. Both accumulate significant amounts of ethylene when submerged (4), but R. acetosa lacks ABA downregulation (9), GA upregulation (104), and increased EXP expression (153). However, when R. acetosa is exposed to elevated GA levels without enhanced ethylene or when ABA levels are reduced with fluridone in submerged plants, petiole elongation is strongly stimulated (9, 104). This demonstrates that signal transduction components required for elongation growth downstream of ABA and GA are present in this species and can be activated. It also shows that in R. acetosa, contrary to R. palustris, ethylene cannot switch on this cascade. Most likely, elements downstream of ethylene but upstream of ABA/GA explain differences in ethylene-induced elongation between Rumex species. Rice cultivars also show variation in elongation capacity during submergence (28, 41, 120). The Sub1 locus controls underwater elongation through genetic distinctions in the two to three ERF proteins it encodes (41, 159) (Figure 2). SUB1A-1 is present in the Sub1 locus only in submergence-tolerant lines and is induced by ethylene. SUB1C is present in all rice lines and is induced by GA. The between-cultivar variation in elongation correlates with genotypic variation and expression of ERFs of the Sub1 locus. The slowly elongating rice varieties of indica rapidly and strongly induce SUB1A-1 upon submergence, whereas all elongating indica and japonica varieties lack either the SUB1A gene or the SUB1A-1 allele (159). Transformation of an elongating japonica variety with a SUB1A-1 full-length cdna under the control of the maize Ubiquitin1 promoter resulted in a significant repression of underwater elongation (159). The expression of SUB1A-1 coincides with repressed accumulation of transcripts for EXPs and reduced expression of SUB1C (41), suggesting that SUB1A acts upstream of GA regulation of EXPs and SUB1C. Improvement of the Oxygen and Carbohydrate Status in Submerged Plants At the whole-plant level, complete submergence leads to a dramatic shift in the carbon budget and energy status, potentially resulting in death. Some relief of this problem, with the leaves still submerged, is underwater photosynthesis (83). The significance of this was exposed by studies showing that light availability enhances survival under water in both flood-tolerant and intolerant species (55, 81, 86, 141) and that O 2 levels in submerged plants are affected by light intensity (90). Improved survival of submergence in the light is correlated with a higher carbohydrate status (97) and internal O 2 concentrations (80, 84, 103). However, underwater photosynthesis can be limited by low light and CO 2 availability. Consistent with these findings are studies showing that illumination can maintain sugar transport and leaf ATP content at near-normoxic levels under strict O 2 deprivation in rice and wheat leaves (85). True aquatics develop specialized leaves characterized by an overall thin leaf and cuticle, a high degree of dissection, and epidermal cell chloroplasts. These traits reduce the diffusion barriers and shorten the diffusion pathways, thus enhancing carbon input per leaf area and unit time (111). Other strategies, developed by true aquatics to enhance carbon gain, are the utilization of HCO 3 as carbon source, C4 or CAM metabolism, or hydrosoil CO 2 consumption 326 Bailey-Serres Voesenek

27 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org (75, 83). Very little information is available about the occurrence of these last strategies in terrestrial plants from flood-prone environments. Leaf acclimations to submergence have been characterized for R. palustris (82) and other amphibious species (18, 37, 157). Leaves developed under water are 20% thinner with an increased specific leaf area (SLA) (m 2 g 1 ), indicating a large surface area relative to mass. The higher SLA is related not only to the lower leaf thickness of aquatic leaves but also to their tenfold-lower starch content. Furthermore, aquatic leaves have thinner epidermal cell walls and cuticles, and their chloroplasts lie close to the epidermis rather than toward the intercellular spaces as is typical for aerial leaves (82). These acclimations are consistent with the view that CO 2 directly enters the mesophyll cells of these leaves via diffusion through the epidermis and not via stomata and intercellular gas diffusion. This diffusion pathway under water has a much higher diffusion resistance for gases than does intercellular diffusion. Calculations for R. palustris indicate a 15,000-fold-higher resistance to CO 2 diffusion in leaves under submergence than when in air (81). However, the morphological and anatomical changes decrease gas diffusion resistance for CO 2 (38). In R. palustris these acclimations result in a dramatic reduction of the diffusion resistance between submerged leaves and leaves in air to a factor of less than 400 (81). Functional consequences of these acclimations in R. palustris include higher rates of net underwater assimilation and lower CO 2 compensation points (81). Similar effects are also described for amphibious species (13, 55, 140). The relatively low diffusion resistance in aquatic leaves also permits increased inward diffusion of O 2 from the water layer into the shoot. This results, in the dark, in an internal O 2 concentration of 17% in acclimated petioles of R. palustris when submerged in airsaturated water, whereas nonacclimated petioles reach only 9% (80). These observations demonstrate that the water column can function as an important source of O 2 for terrestrial plants when they are exposed to submergence and that O 2 levels in leaves, stems, and petioles below the critical O 2 pressure (0.8%; 31) are rare and probably restricted to densely packed tissues or to aquatic environments that are extremely stagnant or have low O 2 levels. Although root systems will likely benefit from these shoot acclimations, O 2 pressures in the roots will still be much lower than the values mentioned here for shoots, especially at night, when there is no photosynthesis (90). It is therefore expected that even with LOES acclimations, roots will also rely on the metabolic cellular adjustments to O 2 deprivation for survival. Plants in frequently flooded environments are expected to display these traits at a higher frequency than do those in rarely flooded areas. Consistently, Ranunculus repens populations in temporary lakes are characterized by constitutively dissected leaves. This morphology allows for a relatively large leaf surface and an improved gas exchange and results in relatively high rates of underwater photosynthesis. Plants from more terrestrial populations have less-dissected leaves and relatively low rates of underwater photosynthesis (73, 74). However, a comparative study of nine species, both flooding tolerant and intolerant, showed that gas exchange acclimations under water are not restricted to flood-tolerant species (84). In this study all but one species developed aquatic leaves that were thinner and had thinner outer cell walls and cuticles and a higher SLA. These responses were independent of the species flooding tolerances. Furthermore, leaf plasticity upon submergence resulted in increased O 2 levels in all species. Therefore, between-species variation in inducible leaf acclimations in terrestrial plants, to optimize gas exchange when submerged, is not related to the variation in flooding tolerance of the species investigated (84). This conclusion hints toward a limited role of submergence signals, such as elevated ethylene, in inducing leaf acclimations that enhance gas exchange under water. Plants that are not exposed to flooding Acclimation to Flooding Stress 327

28 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org throughout their life are not expected to use these signals to switch on signaling cascades that lead to altered leaf anatomy and morphology. More likely, signals associated with changed rates of photosynthesis and/or reduced levels of carbohydrates induce these leaf acclimations. This hypothesis is consistent with observations that shade-acclimated plants with reduced rates of photosynthesis develop thinner leaves with higher SLA (78). Consistently, transgenic tobacco plants with substantially reduced Rubisco levels have reduced photosynthesis and increased SLA (34). Improvement of Internal Gas Diffusion: Aerenchyma Important traits for survival in flooded environments are those that reduce the resistance for diffusion of O 2 and CO 2 from the environment to the plant. Equally significant, however, is the resistance that hampers gas diffusion within organs. Fast gas diffusion can be accomplished only in a gaseous diffusion medium, over short distances, by limited loss of the gas along the diffusion path, and by restricted tortuosity of the diffusion route. These requirements are met in aerenchymatous tissue, characterized by longitudinally interconnected gas spaces in roots and shoots. Aerenchyma is either constitutively present and/or induced upon flooding (116, 144) and develops in existing tissues or concomitant with the development of new roots (32). Distinct physiological processes are at the basis of aerenchyma formation. This led to the discrimination of two aerenchyma types: (a) lysigenous aerenchyma formed by cell death and (b) schizogenous aerenchyma in which gas spaces develop through the separation of previously connected cells (23, 33, 60). A third type, termed expansigenous aerenchyma, is characterized by intercellular gas spaces that develop through cell division and cell enlargement, without cell separation or collapse/death (118). Combinations of these aerenchyma types also exist (118), and within one plant species different types can be present in different organs (32, 33). The mechanism of schizogenous aerenchyma formation is largely unknown as compared with that of lysigenous aerenchyma. Low O 2 and elevated ethylene can induce lysigenous aerenchyma development in roots of maize in a manner that is phenotypically similar to the process promoted by flooding (32). Under flooded conditions, subambient O 2 concentrations stimulate the production of ethylene, which accumulates in roots surrounded by water and induces programmed cell death (PCD) in the cortex tissue (53). Accordingly, hypoxic roots, exposed to inhibitors of ethylene biosynthesis or action, form no gas spaces (53). Downstream components of this regulatory route include protein kinases, protein phosphatases, G proteins, Ca 2+, and inositol phospholipids (54). The targets of these signaling routes include proteins associated with cell wall breakdown. The activity of cellulase increases in roots upon exposure to low O 2 or ethylene (52). Furthermore, increases in pectinase and xylanase activity (15) and the induction of XTH mrnas occur in diverse species in response to flooding or hypoxia (67, 70, 109). Large data collections are available on genetic diversity in traits that contribute to the delivery of O 2 to root tips. Justin & Armstrong (60) compared 91 species from wetland, intermediate, and nonwetland habitats. Nearly all the species from nonwetland environments had low root porosities, whereas high constitutive and increased porosities upon flooding were associated with species from wetland environments. Also, other studies confirmed the strong correlation between high root porosities and occurrence in wet environments (45, 77, 143). Interestingly, a comparative study on 35 wild Hordeum accessions from environments that differ in flooding intensity showed that this correlation does not always exist and that aerenchyma development can be constrained by phylogeny (42). Aerenchyma is also formed in shoot organs, providing a system of interconnected 328 Bailey-Serres Voesenek

29 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org channels from leaf to root tip. In a study with 14 species divided over seven families, the aerenchyma content of petioles strongly correlated with plant survival during complete submergence. This robust correlation persisted in environmental conditions with (light) and without (dark) underwater photosynthesis (79, 84). These observations suggest that aerenchyma is important not only for survival during partial flooding but also during complete submergence. Petiole aerenchyma likely facilitates the diffusion of O 2 from shoot organs to the roots. The O 2 involved can be photosynthetically derived during the light period or obtained from the water layer by the shoot during the dark. CONCLUSIONS AND FUTURE PERSPECTIVES The growing understanding of the molecular basis and genetic diversity in submergence and flooding acclimations provides opportunities to breed and engineer crops tolerant of these conditions that would benefit the world s farmers. The evaluation of diversity exposes plasticity in metabolic and developmental acclimations that enable distinct strategies that increase fitness in a flooded environment. Natural variation in acclimation schemes provides opportunities for development of crops with combinations of submergence tolerance traits that are optimal at specific developmental stages and under particular flooding regimes, which vary substantially worldwide. The first example of this is the use of marker-assisted breeding to introduce the submergence-tolerance conferring Sub1 genotype to selected rice cultivars (159), which may appreciably benefit rice production in flood-prone lands in the Third World. The further exploration of the molecular basis of genetic diversity in flooding tolerances is critical given the global climate change scenarios that predict heavy precipitation in regions of our planet. SUMMARY POINTS 1. Evaluation of diversity exposes the remarkable plasticity in metabolic and developmental acclimations that enable increased fitness in a flooded environment. 2. Plants employing an escape strategy develop a suite of traits collectively called the low-oxygen escape syndrome (LOES). 3. A consequence of low-o 2 stress is a requirement for energy conservation that is invoked through adjustments in gene expression, carbohydrate catabolism, NAD(P) + regeneration, and ATP production. 4. Energy conservation is influenced by a low-o 2 -induced nonsymbiotic hemoglobin that regulates cytosolic and mitochondrial processes, including rates of fermentation, NO, and ATP production. 5. Enhanced shoot elongation upon submergence requires the action of at least three hormones (ethylene, ABA, and GA) that regulate processes such as apoplastic acidification, cell wall loosening, cell division, and starch breakdown. 6. Anatomical and biochemical leaf acclimations upon submergence facilitate underwater photosynthesis as well as the inward diffusion of O 2 from the floodwater. 7. Aerenchyma in root and shoot tissue not only is important for survival during partial submergence but also facilitates O 2 diffusion from shoot to root while the plant is completely submerged. Acclimation to Flooding Stress 329

30 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org FUTURE ISSUES 1. Plant species differ in their growth response to ethylene during submergence. This is far from understood but probably involves signal transduction components upstream of ABA and GA. The characterization of SUB1A is an important finding in this respect. More work is needed because this is probably an important selection point to differentiate survival strategies. 2. A quiescence strategy (carbohydrate conservation) in rice is associated with submergence tolerance. However, not all plants that fail to elongate under water are tolerant. The question arises as to whether cells of these plants are metabolically inactive or simply lack other aspects also needed for tolerance (e.g., the ability to manage ATP, cytosolic ph, or cellular O 2 content; protection against ROS; or aerenchyma development). 3. Characterization of the Sub1 locus provides an opportunity to breed or engineer submergence-tolerant rice that could benefit farmers in flood-prone areas. Studies are needed to determine if submergence and salt tolerance can be combined because floodwaters can be saline. 4. The development of rice cultivars with improved underwater germination and low-o 2 escape capabilities may reduce herbicide use. DISCLOSURE STATEMENT The authors are not aware of any biases that might be perceived as affecting the objectivity of this review. ACKNOWLEDGMENTS The authors thank Ronald Pierik for his critical comments and his efforts to condense this review. Furthermore, we thank Tim Colmer, Takeshi Fukao, Robert Hill, Liesje Mommer, Angelika Mustroph, Pierdomenico Perata, and Eric Visser for comments and discussion. We regret that many publications were not cited owing to space limitations. Submergence and low- O 2 stress research in the Bailey-Serres lab is currently supported by the NSF (IBN ) and USDA ( ). Flooding research in the Voesenek lab has been continuously supported by the Netherlands Organization for Scientific Research. LITERATURE CITED 1. Arnell N, Liu C Climate change 2001: hydrology and water resources. In Report Intergovernmental Panel on Climate Change Arpagaus S, Braendle R The significance of α-amylase under anoxia stress in tolerant rhizomes (Acorus calamus L.) and nontolerant tubers (Solanum tuberosum L., var. Désirée). J. Exp. Bot. 51: Bailey-Serres J, Chang R Sensing and signalling in response to oxygen deprivation in plants and other organisms. Ann. Bot. 96: Banga M, Slaa EJ, Blom CWPM, Voesenek LACJ Ethylene biosynthesis and accumulation under drained and submerged conditions: a comparative study of two Rumex species. Plant Physiol. 112: Bailey-Serres Voesenek

31 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 5. Baxter-Burrell A, Chang R, Springer P, Bailey-Serres J Gene and enhancer trap transposable elements reveal oxygen deprivation-regulated genes and their complex patterns of expression in Arabidopsis. Ann. Bot. 91: Baxter-Burrell A, Yang Z, Springer PS, Bailey-Serres J RopGAP4- dependent Rop GTPase rheostat control of Arabidopsis oxygen deprivation tolerance. Science 296: Bell EL, Klimova TA, Eisenbart J, Moraes CT, Murphy MP, et al The Q o site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production. J. Cell Biol. 177: Benschop JJ, Bou J, Peeters AJM, Wagemaker N, Gühl K, et al Long-term submergence-induced elongation in Rumex palustris requires abscisic acid-dependent biosynthesis of gibberellin. Plant Physiol. 141: Benschop JJ, Jackson MB, Gühl K, Vreeburg RAM, Croker SJ, et al Contrasting interactions between ethylene and abscisic acid in Rumex species differing in submergence tolerance. Plant J. 44: Bieniawska Z, Paul Barratt DH, Garlick AP, Thole V, Kruger NJ, et al Analysis of the sucrose synthase gene family in Arabidopsis. Plant J. 49: Blokhina OB, Chirkova TV, Fagerstedt KV Anoxic stress leads to hydrogen peroxide formation in plant cells. J. Exp Bot. 52: Blom CWPM, Voesenek LACJ Flooding: the survival strategies of plants. Trends Ecol. Evol. 11: Boeger MRT, Poulson ME Morphological adaptations and photosynthesis rates of amphibious Veronica anagallis-aquatica L. (Scrophulariaceae) under different flow regimes. Aquat. Bot. 75: Bologa KL, Fernie AR, Leisse A, Loureiro ME, Geigenberger P A bypass of sucrose synthase leads to low internal oxygen and impaired metabolic performance in growing potato tubers. Plant Physiol. 132: Bragina TV, Rodionova NA, Grinieva GM Ethylene production and activation of hydrolytic enzymes during acclimation of maize seedlings to partial flooding. Russ. J. Plant Physiol. 50: Branco-Price C, Kawaguchi R, Ferreira RB, Bailey-Serres J Genome-wide analysis of transcript abundance and translation in Arabidopsis seedlings subjected to oxygen deprivation. Ann. Bot. 96: Breitkreuz KE, Allan WL, Van Cauwenberghe OR, Jakobs C, Talibi D, et al A novel γ-hydroxybutyrate dehydrogenase: identification and expression of an Arabidopsis cdna and potential role under oxygen deficiency. J. Biol. Chem. 278: Bruni NC, Young JP, Dengler NC Leaf developmental plasticity of Ranunculus flabellaris in response to terrestrial and submerged environments. Can. J. Bot. 74: Chang WW, Huang L, Shen M, Webster C, Burlingame AL, Roberts JK Patterns of protein synthesis and tolerance of anoxia in root tips of maize seedlings acclimated to a low-oxygen environment, and identification of proteins by mass spectrometry. Plant Physiol. 122: Cho HT, Kende H Expansins and internodal growth of deepwater rice. Plant Physiol. 113: Cho HT, Kende H Expression of expansin genes is correlated with growth in deepwater rice. Plant Cell 9: Identifies signal transduction pathway components that regulate ADH and ROS production under oxygen deprivation. Acclimation to Flooding Stress 331

32 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 25. The first report that a barrier to radial oxygen loss can be induced in existing roots. 26. Describes the requirement of a petiole angle signal for submergenceinduced petiole elongation. 41. Identified physiological aspects controlled by genetic variation at the Sub1 locus conferring distinctions between escape and quiescence survival strategies under submergence. 22. Choi D, Kim JH, Kende H Whole genome analysis of the OsGRF gene family encoding plant-specific putative transcription activators in rice. Plant Cell Physiol. 45: Choi WG, Roberts DM Arabidopsis NIP2;1: a major intrinsic protein transporter of lactic acid induced by anoxic stress. J. Biol. Chem. 282: Colmer TD Long-distance transport of gases in plants: a perspective on internal aeration and radial oxygen loss from roots. Plant Cell Environ. 26: Colmer TD, Gibberd MR, Wiengweera A, Tinh TK The barrier to radial oxygen loss from roots of rice (Oryza sativa L.) is induced by growth in stagnant solutions. J. Exp. Bot. 49: Cox MCH, Millenaar FF, de Jong vanberkel YEM, Peeters AJM, Voesenek LACJ Plant movement. Submergence-induced petiole elongation in Rumex palustris depends on hyponastic growth. Plant Physiol. 132: Darley CP, Forrester AM, McQueen-Mason SJ The molecular basis of plant cell wall extension. Plant Mol. Biol. 47: Das KK, Sarkar RK, Ismail AM Elongation ability and nonstructural carbohydrate levels in relation to submergence tolerance in rice. Plant Sci. 168: Dixon MH, Hill SA, Jackson MB, Ratcliffe RG, Sweetlove LJ Physiological and metabolic adaptations of Potamogeton pectinatus L. tubers support rapid elongation of stem tissue in the absence of oxygen. Plant Cell Physiol. 47: Dordas C, Rivoal J, Hill RD Plant haemoglobins, nitric oxide and hypoxic stress. Ann. Bot. 91: Drew MC Oxygen deficiency and root metabolism: injury and acclimation under hypoxia and anoxia. Annu. Rev. Plant Physiol. Plant Mol. Biol. 48: Drew MC, He CJ, Morgan PW Programmed cell death and aerenchyma formation in roots. Trends Plant Sci. 5: Evans DE Aerenchyma formation. New Phytol. 161: Evans JR, Von Caemmerer S, Setchell BA, Hudson GS The relationship between CO 2 transfer conductance and leaf anatomy in transgenic tobacco with a reduced content of rubisco. Aust. J. Plant Physiol. 21: Felle HH ph regulation in anoxic plants. Ann. Bot. 96: Fennoy SL, Nong T, Bailey-Serres J Transcriptional and post-transcriptional processes regulate gene expression in oxygen-deprived roots of maize. Plant J. 15: Frost-Christensen H, Bolt Jorgensen L, Floto F Species specificity of resistance to oxygen diffusion in thin cuticular membranes from amphibious plants. Plant Cell Environ. 26: Frost-Christensen H, Floto F Resistance to CO 2 diffusion in cuticular membranes of amphibious plants and the implication for CO 2 acquisition. Plant Cell Environ. 30: Fukao T, Bailey-Serres J Plant responses to hypoxia. Is survival a balancing act? Trends Plant Sci. 9: Fukao T, Kennedy RA, Yamasue Y, Rumpho ME Genetic and biochemical analysis of anaerobically-induced enzymes during seed germination of Echinochloa crus-galli varieties tolerant and intolerant of anoxia. J. Exp. Bot. 54: Fukao T, Xu K, Ronald PC, Bailey-Serres J A variable cluster of ethylene response factor-like genes regulates metabolic and developmental acclimation responses to submergence in rice. Plant Cell 18: Bailey-Serres Voesenek

33 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 42. Garthwaite AJ, Von Bothmer R, Colmer TD Diversity in root aeration traits associated with waterlogging tolerance in the genus Hordeum. Funct. Plant Biol. 30: Geigenberger P Response of plant metabolism to too little oxygen. Curr. Opin. Plant Biol. 6: Geigenberger P, Fernie AR, Gibon Y, Christ M, Stitt M Metabolic activity decreases as an adaptive response to low internal oxygen in growing potato tubers. Biol. Chem. 381: Gibberd MR, Gray JD, Cocks PS, Colmer TD Waterlogging tolerance among a diverse range of Trifolium accessions is related to root porosity, lateral root formation and aerotropic rooting. Ann. Bot. 88: Gibbs J, Greenway H Review: Mechanisms of anoxia tolerance in plants. I. Growth, survival and anaerobic catabolism. Funct. Plant Biol. 30: Greenway H, Armstrong W, Colmer TD Conditions leading to high CO 2 (>5 kpa) in waterlogged-flooded soils and possible effects on root growth and metabolism. Ann. Bot. 98: Greenway H, Gibbs J Review: mechanisms of anoxia tolerance in plants. II. Energy requirements for maintenance and energy distribution to essential processes. Funct. Plant Biol. 30: Groenveld HW, Voesenek LACJ Submergence-induced petiole elongation in Rumex palustris is controlled by developmental stage and storage compounds. Plant Soil 253: Guglielminetti L, Yamaguchi J, Perata P, Alpi A Amylolytic activities in cereal seeds under aerobic and anaerobic conditions. Plant Physiol. 109: Guzy RD, Schumacker PT Oxygen sensing by mitochondria at complex III: the paradox of increased reactive oxygen species during hypoxia. Exp. Physiol. 91: He CJ, Drew MC, Morgan PW Induction of enzymes associated with lysigenous aerenchyma formation in roots of Zea mays during hypoxia or nitrogen-starvation. Plant Physiol. 105: He CJ, Finlayson SA, Drew MC, Jordan WR, Morgan PW Ethylene biosynthesis during aerenchyma formation in roots of maize subjected to mechanical impedance and hypoxia. Plant Physiol. 112: He CJ, Morgan PW, Drew MC Transduction of an ethylene signal is required for cell death and lysis in the root cortex of maize during aerenchyma formation induced by hypoxia. Plant Physiol. 112: He JB, Bögemann GM, Van der Steeg HM, Rijnders JHGM, Voesenek LACJ, Blom CWPM Survival tactics of Ranunculus species in river floodplains. Oecologia 118: Hunt PW, Klok EJ, Trevaskis B, Watts RA, Ellis MH, et al Increased level of hemoglobin 1 enhances survival of hypoxic stress and promotes early growth in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 99: Igamberdiev AU, Hill RD Nitrate, NO and haemoglobin in plant adaptation to hypoxia: an alternative to classic fermentation pathways. J. Exp. Bot. 55: Ishizawa K, Murakami S, Kawakami Y, Kuramochi H Growth and energy status of arrowhead tubers, pondweed turions and rice seedlings under anoxic conditions. Plant Cell Environ. 22: Jackson MB Ethylene-promoted elongation: an adaptation to submergence stress. Ann. Bot. 101: A hemoglobin-like protein that is expressed in response to hypoxia confers increased tolerance to the stress. Acclimation to Flooding Stress 333

34 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 60. Comprehensive overview of genetic diversity in aerenchyma architecture and development. 60. Justin SHFW, Armstrong W The anatomical characteristics of roots and plant response to soil flooding. New Phytol. 106: Kende H, Van Der Knaap E, Cho HT Deepwater rice: a model plant to study stem elongation. Plant Physiol. 118: Kim JH, Cho HT, Kende H α-expansins in the semiaquatic ferns Marsilea quadrifolia and Regnellidium diphyllum: evolutionary aspects and physiological role in rachis elongation. Planta 212: Klok EJ, Wilson IW, Wilson D, Chapman SC, Ewing RM, et al Expression profile analysis of the low-oxygen response in Arabidopsis root cultures. Plant Cell 14: Koch K Sucrose metabolism: regulatory mechanisms and pivotal roles in sugar sensing and plant development. Curr. Opin. Plant Biol. 7: Kürsteiner O, Dupuis I, Kuhlemeier C The pyruvate decarboxylase1 gene of Arabidopsis is required during anoxia but not other environmental stresses. Plant Physiol. 132: Kuzmin EV, Karpova OV, Elthon TE, Newton KJ Mitochondrial respiratory deficiencies signal up-regulation of genes for heat shock proteins. J. Biol. Chem. 279: Lasanthi-Kudahettige R, Magneschi L, Loreti E, Gonzali S, Licausi F, et al Transcript profiling of the anoxic rice coleoptile. Plant Physiol. 144: Lee Y, Kende H Expression of α-expansins is correlated with internodal elongation in deepwater rice. Plant Physiol. 127: Libourel IG, van Bodegom PM, Fricker MD, Ratcliffe RG Nitrite reduces cytoplasmic acidosis under anoxia. Plant Physiol. 142: Liu F, Vantoai T, Moy L, Bock G, Linford LD, Quackenbush J Global transcription profiling reveals novel insights into hypoxic response in Arabidopsis. Plant Physiol. 137: Loreti E, Poggi A, Novi G, Alpi A, Perata P Genome-wide analysis of gene expression in Arabidopsis seedlings under anoxia. Plant Physiol. 137: Loreti E, Yamaguchi J, Alpi A, Perata P Sugar modulation of α-amylase genes under anoxia. Ann. Bot. 91: Lynn DE, Waldren S Morphological variation in populations of Ranunculus repens from the temporary limestone lakes (turloughs) in the West of Ireland. Ann. Bot. 87: Lynn DE, Waldren S Physiological variation in populations of Ranunculus repens L. (creeping buttercup) from the temporary limestone lakes (turloughs) in the west of Ireland. Ann. Bot. 89: Maberly SC, Madsen TV Affinity for CO 2 in relation to the ability of freshwater macrophytes to use HCO 3. Funct. Ecol. 12: Malone M, Ridge I Ethylene-induced growth and proton excretion in the aquatic plant Nymphoides peltata. Planta 157: McDonald MP, Galwey NW, Colmer TD Similarity and diversity in adventitious root anatomy as related to root aeration among a range of wetland and dryland grass species. Plant Cell Environ. 25: Mommer L, De Kroon H, Pierik R, Bögemann GM, Visser EJW A functional comparison of acclimation to shade and submergence in two terrestrial plant species. New Phytol. 167: Mommer L, Lenssen JPM, Huber H, Visser EJW, De Kroon H Ecophysiological determinants of plant performance under flooding: a comparative study of seven plant families. J. Ecol. 94: Bailey-Serres Voesenek

35 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 80. Mommer L, Pedersen O, Visser EJW Acclimation of a terrestrial plant to submergence facilitates gas exchange under water. Plant Cell Environ. 27: Mommer L, Pons TL, Visser EJW Photosynthetic consequences of phenotypic plasticity in response to submergence: Rumex palustris as a case study. J. Exp. Bot. 57: Mommer L, Pons TL, Wolters-Arts M, Venema JH, Visser EJW Submergence-induced morphological, anatomical, and biochemical responses in a terrestrial species affects gas diffusion resistance and photosynthetic performance. Plant Physiol. 139: Mommer L, Visser EJW Underwater photosynthesis in flooded terrestrial plants: a matter of leaf plasticity. Ann. Bot. 96: Mommer L, Wolters-Arts M, Andersen C, Visser EJW, Pedersen O Submergenceinduced leaf acclimation in terrestrial species varying in flooding tolerance. New Phytol. 176: Mustroph A, Boamfa EI, Laarhoven LJ, Harren FJ, Pörs Y, Grimm B Organ specific analysis of the anaerobic primary metabolism in rice and wheat seedlings II: Light exposure reduces needs for fermentation and extends survival during anaerobiosis. Planta 225: Nabben RHM, Blom CWPM, Voesenek LACJ Resistance to complete submergence in Rumex species with different life histories: the influence of plant size and light. New Phytol. 144: Nakazono M, Tsuji H, Li Y, Saisho D, Arimura S, Tsutsumi N, Hirai A Expression of a gene encoding mitochondrial aldehyde dehydrogenase in rice increases under submerged conditions. Plant Physiol. 124: Nie X, Durnin DC, Igamberdiev AU, Hill RD Cytosolic calcium is involved in the regulation of barley hemoglobin gene expression. Planta 223: Ookawara R, Satoh S, Yoshihito T, Ishizawa K Expression of α-expansin and xyloglucan endotransglucosylase/hydrolase genes associated with shoot elongation enhanced by anoxia, ethylene and carbon dioxide in arrowhead (Sagittaria pygmaea Miq.) tubers. Ann. Bot. 96: Pedersen O, Vos H, Colmer TD Oxygen dynamics during submergence in the halophytic stem succulent Halosarcia pergranulata. Plant Cell Environ. 29: Perata P, Voesenek LACJ Submergence tolerance in rice requires Sub1A, an ethylene-response-factor-like gene. Trends Plant Sci. 12: Pierik R, Millenaar FF, Peeters AJM, Voesenek LACJ New perspectives in flooding research: the use of shade avoidance and Arabidopsis thaliana. Ann. Bot. 96: Pierik R, Tholen D, Poorter H, Visser EJW, Voesenek LACJ The Janus face of ethylene: growth inhibition and stimulation. Trends Plant Sci. 11: Planchet E, Jagadis Gupta K, Sonoda M, Kaiser WM Nitric oxide emission from tobacco leaves and cell suspensions: rate limiting factors and evidence for the involvement of mitochondrial electron transport. Plant J. 41: Plaxton WC, Podesta FE The functional organization and control of plant respiration. Crit. Rev. Plant Sci. 25: Potters G, Pasternak TP, Guisez Y, Palme KJ, Jansen MAK Stress-induced morphogenic responses: growing out of trouble? Trends Plant Sci. 12: Ram PC, Singh BB, Singh AK, Ram P, Singh PN, et al Submergence tolerance in rainfed lowland rice: physiological basis and prospects for cultivar improvement through marker-aided breeding. Field Crops. Res. 76: Describes anatomical and biochemical changes upon submergence that affect underwater photosynthesis in Rumex palustris. 93. Presents an integrated view on the growthinhibitory and -stimulating effects of the gaseous plant hormone ethylene. Acclimation to Flooding Stress 335

36 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 98. Ratcliffe RG Intracellular ph regulation in plants under anoxia. In Regulation of Tissue ph in Plants and Animals: A Reappraisal of Current Techniques, ed. S Egginton, EW Taylor, JA Raven, pp Cambridge, UK: Cambridge Univ. Press 99. Rhoads DM, Umbach AL, Subbaiah CC, Siedow JN Mitochondrial reactive oxygen species. Contribution to oxidative stress and interorganellar signaling. Plant Physiol. 141: Ricoult C, Echeverria LO, Cliquet JB, Limami AM Characterization of alanine aminotransferase (AlaAT) multigene family and hypoxic response in young seedlings of the model legume Medicago truncatula. J Exp. Bot. 57: Ridge I Ethylene and growth control in amphibious plants. In Plant Life in Aquatic and Amphibious habitats, ed RMM Crawford, pp Oxford, UK: Blackwell Sci. Publ Rieu I, Cristescu SM, Harren FJM, Huibers W, Voesenek LACJ, et al RP-ACS1, a flooding-induced 1-aminocyclopropane-1-carboxylate synthase gene of Rumex palustris, is involved in rhythmic ethylene production. J. Exp. Bot. 56: Rijnders JGHM, Armstrong W, Darwent MJ, Blom CWPM, Voesenek LACJ The role of oxygen in submergence-induced petiole elongation in Rumex palustris: in situ measurements of oxygen in petioles of intact plants using microelectrodes. New Phytol. 147: Rijnders JGHM, Yang YY, Kamiya Y, Takahashi N, Barendse GWM, et al Ethylene enhances gibberellin levels and petiole sensitivity in flooding-tolerant Rumex palustris but not in flooding-intolerant R. acetosa. Planta 20: Roberts JKM, Callis J, Wemmer D, Walbot V, Jardetzky O Mechanisms of cytoplasmic ph regulation in hypoxic maize root tips and its role in survival under anoxia. Proc. Natl. Acad. Sci. USA 81: Roberts JKM, Chang K, Webster C, Callis J, Walbot V Dependence of ethanolic fermentation, cytoplasmic ph regulation, and viability on the activity of alcohol dehydrogenase in hypoxic maize root tips. Plant Physiol. 89: Rolletschek H, Weschke W, Weber H, Wobus U, Borisjuk L Energy state and its control on seed development: Starch accumulation is associated with high ATP and steep oxygen gradients within barley grains. J. Exp. Bot. 55: Rzewuski G, Sauter M The novel rice (Oryza sativa L.) gene OsSbf1 encodes a putative member of the Na + /bile acid symporter family. J. Exp. Bot. 53: Saab IN, Sachs MM A flooding-induced xyloglucan endo-transglycosylase homolog in maize is responsive to ethylene and associated with aerenchyma. Plant Physiol. 112: Saika H, Okamoto M, Miyoshi K, Kushiro T, Shinoda S, et al Ethylene promotes submergence-induced expression of OsABA8ox1, a gene that encodes ABA 8 -hydroxylase in rice. Plant Cell Physiol. 48: Sand-Jensen K, Frost-Christensen H Plant growth and photosynthesis in the transient zone between land and stream. Aquat. Bot. 63: Santosa IE, Ram PC, Boamfa EI, Laarhoven LJ, Reuss J, et al Patterns of peroxidative ethane emission from submerged rice seedlings indicate that damage from reactive oxygen species takes place during submergence and is not necessarily a postanoxic phenomenon. Planta 226: Sato T, Harada T, Ishizawa K Stimulation of glycolysis in anaerobic elongation of pondweed (Potamogeton distinctus) turion. J. Exp. Bot. 53: Sauter M Differential expression of CAK (cdc2-activating kinase)-like protein kinase, cyclins and cdc2 genes from rice during the cell cycle and in response to gibberellin. Plant J. 11: Bailey-Serres Voesenek

37 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 115. Sauter M Rice in deep water: How to take heed against a sea of trouble. Naturwissenschaften 87: Sauter M, Mekhedov SL, Kende H Gibberellin promotes histone H1 kinase activity and the expression of cdc2 and cyclin genes during the induction of rapid growth in deepwater rice internodes. Plant J. 7: Sauter M, Rzewuski G, Marwedel T, Lorbiecke R The novel ethylene-regulated gene OsUsp1 from rice encodes a member of a plant protein family related to prokaryotic universal stress proteins. J. Exp. Bot. 53: Seago JL, Marsh LC, Stevens KJ, Soukup A, Votrubova O, Enstone DE A reexamination of the root cortex in wetland flowering plants with respect to aerenchyma. Ann. Bot. 96: Sedbrook JC, Kronebusch PJ, Borisy GG, Trewavas AJ, Masson PH Transgenic AEQUORIN reveals organ specific cytosolic Ca2 + responses to anoxia in Arabidopsis thaliana seedlings. Plant Physiol. 111: Setter TL, Laureles EV The beneficial effect of reduced elongation growth on submergence tolerance of rice. J. Exp. Bot. 47: Setter TL, Waters I Review of prospects for germplasm improvement for waterlogging tolerance in wheat, barley and oats. Plant Soil. 253: Silvertown J, Dodd ME, Gowing DJG, Mountford JO Hydrologically defined niches reveal a basis for species richness in plant communities. Nature 400: Sowa AW, Duff SM, Guy PA, Hill RD Altering hemoglobin levels changes energy status in maize cells under hypoxia. Proc. Natl Acad. Sci. USA 95: Stoimenova M, Igamberdiev AU, Gupta KJ, Hill RD Nitrite-driven anaerobic ATP synthesis in barley and rice root mitochondria. Planta 226: Stoimenova M, Libourel IGL, Ratcliffe RG, Kaiser WM The role of nitrate reduction in the anoxic metabolism of roots II. Anoxic metabolism of tobacco roots with or without nitrate reductase activity. Plant Soil 253: Subbaiah CC, Bush DS, Sachs MM Elevation of cytosolic calcium precedes anoxia gene expression in maize suspension-cultured cells. Plant Cell 6: Subbaiah CC, Bush DS, Sachs MM Mitochondrial contribution to the anoxic Ca 2+ signal suspension cultured cells. Plant Physiol. 118: Subbaiah CC, Zhang J, Sachs MM Involvement of intracellular calcium in anaerobic gene expression and survival of maize seedlings. Plant Physiol. 105: Summers JE, Ratcliffe RG, Jackson MB Anoxia tolerance in the aquatic monocot Potamogeton pectinatus: absence of oxygen stimulates elongation in association with an unusually large Pasteur effect. J. Exp. Bot. 51: Tadege M, Dupuis I, Kuhlemeier C Ethanolic fermentation: new functions for an old pathway. Trends Plant Sci. 4: Van der Knaap E, Jagoueix S, Kende H Expression of an ortholog of replication protein A1 (RPA1) is induced by gibberellin in deepwater rice. Proc. Natl. Acad. Sci. USA 94: Van der Knaap E, Kim JH, Kende H A novel gibberellin-induced gene from rice and its potential regulatory role in stem growth. Plant Physiol. 122: Van der Knaap E, Song WY, Ruan DL, Sauter M, Ronald PC, Kende H Expression of a gibberellin-induced leucine-rich repeat receptor-like protein kinase in deepwater rice and its interaction with kinase-associated protein phosphatase. Plant Physiol. 120: The first paper that experimentally demonstrates that costs are associated with stimulated shoot elongation underwater Demonstrates that ATP production can be sustained under anoxia by provision of succinate and nitrite. Acclimation to Flooding Stress 337

38 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 134. Van der Sman AJM, Blom CWPM, van de Steeg HM Phenology and seed production in Chenopodium rubrum, Rumex maritimus, and Rumex palustris as related to photoperiod in river forelands. Can. J. Bot. 70: Van der Straeten D, Zhou Z, Prinsen E, Van Onckelen HA, van Montagu MC A comparative molecular-physiological study of submergence response in lowland and deepwater rice. Plant Physiol. 125: van Dongen JT, Roeb GW, Dautzenberg M, Froehlich A, Vigeolas H, et al Phloem import and storage metabolism are highly coordinated by the low oxygen concentrations within developing wheat seeds. Plant Physiol. 135: van Dongen JT, Schurr U, Pfister M, Geigenberger P Phloem metabolism and function have to cope with low internal oxygen. Plant Physiol. 131: Van Eck WHJM, Van der Steeg HM, Blom CWPM, De Kroon H Is tolerance to summer flooding correlated with distribution patterns in river floodplains? A comparative study of 20 terrestrial grassland species. Oikos 107: Vanlerberghe GC, Feil R, Turpin DH Anaerobic metabolism in the N-limited green alga Selenastrum minutum: I. Regulation of carbon metabolism and succinate as a fermentation product. Plant Physiol. 94: Vervuren PJA, Beurskens SMJH, Blom CWPM Light acclimation, CO 2 response and long-term capacity of underwater photosynthesis in three terrestrial plant species. Plant Cell Environ. 22: Vervuren PJA, Blom CWPM, de Kroon H Extreme flooding events on the Rhine and the survival and distribution of riparian plant species. J. Ecol. 91: Visser EJW, Blom CWPM, Voesenek LACJ Flooding-induced adventitious rooting in Rumex: morphology and development in an ecological perspective. Acta Bot. Neerl. 45: Visser EJW, Bögemann GM, Van de Steeg HM, Pierik R, Blom CWPM Flooding tolerance of Carex species in relation to field distribution and aerenchyma formation. New Phytol. 148: Visser EJW, Colmer TD, Blom CWPM, Voesenek LACJ Changes in growth, porosity, and radial oxygen loss from adventitious roots of selected mono- and dicotyledonous wetland species with contrasting types of aerenchyma. Plant Cell Environ. 23: Visser EJW, Nabben RHM, Blom CWPM, Voesenek LACJ Elongation by primary lateral roots and adventitious roots during conditions of hypoxia and high ethylene concentrations. Plant Cell Environ. 20: Vlad F, Spano T, Vlad D, Daher FB, Ouelhadj A, Kalaitzis P Arabidopsis prolyl 4-hydroxylases are differentially expressed in response to hypoxia, anoxia and mechanical wounding. Physiol. Plantarium 130: Voesenek LACJ, Blom CWPM Stimulated shoot elongation: a mechanism of semiaquatic plants to avoid submergence stress. In Plant Responses to Environmental Stress: From Phytohormones to Genome Reorganization, ed. HR Lerner, pp New York: Marcel Dekker 148. Voesenek LACJ, Colmer TD, Pierik R, Millenaar FF, Peeters AJM How plants cope with complete submergence. New Phytol. 170: Voesenek LACJ, Jackson MB, Toebes AHW, Vriezen WH, Colmer TD Desubmergence-induced ethylene production in Rumex palustris: regulation and ecophysiological significance. Plant J. 33: Bailey-Serres Voesenek

39 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 150. Voesenek LACJ, Rijnders JHGM, Peeters AJM, van de Steeg HM, de Kroon H Plant hormones regulate fast shoot elongation under water: from genes to communities. Ecology 85: Voesenek LACJ, van der Sman AJM, Harren FJM, Blom CWPM An amalgamation between hormone physiology and plant ecology: a review on flooding resistance and ethylene. J. Plant Growth Regul. 11: Vreeburg RAM, Benschop JJ, Peeters AJM, Colmer TD, Ammerlaan AHM, et al Ethylene regulates fast apoplastic acidification and expansin A transcription during submergence-induced petiole elongation in Rumex palustris. Plant J. 43: Vriezen WH, De Graaf B, Mariani C, Voesenek LACJ Submergence induces expansin gene expression in flooding-tolerant Rumex palustris and not in flooding-intolerant R. acetosa. Planta 210: Vriezen WH, Hulzink R, Mariani C, Voesenek LACJ Aminocyclopropane-1- carboxylate oxidase activity limits ethylene biosynthesis in Rumex palustris during submergence. Plant Physiol. 121: Vriezen WH, van Rijn CPE, Voesenek LACJ, Mariani C A homolog of the Arabidopsis thaliana ERS gene is actively regulated in Rumex palustris upon flooding. Plant J. 11: Watanabe H, Saigusa M, Hase S, Hayakawa T, Satoh S Cloning of a cdna encoding an ETR2-like protein (Os-ERL1) from deepwater rice (Oryza sativa L.) and increase in its mrna level by submergence, ethylene, and gibberellin treatments. J. Exp. Bot. 55: Wells CL, Pigliucci M Adapative phenotypic plasticity: the case of heterophylly in aquatic plants. Perspect. Plant Ecol. Evol. Syst. 3: Xia JH, Roberts J Improved cytoplasmic ph regulation, increased lactate efflux, and reduced cytoplasmic lactate levels are biochemical traits expressed in root tips of whole maize seedlings acclimated to a low-oxygen environment. Plant Physiol. 105: Xu K, Xu X, Fukao T, Canlas P, Marghirang-Rodriguez R, et al Sub1A is an ethylene-response-factor-like gene that confers submergence tolerance to rice. Nature 442: Discusses that submergenceinduced shoot elongation is predominantly a trait associated with habitats characterized by shallow and prolonged flooding events Landmark map-based cloning of a polygenic locus that encodes a determinant of submergence tolerance in rice. Acclimation to Flooding Stress 339

40 Contents Annual Review of Plant Biology Volume 59, 2008 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Our Work with Cyanogenic Plants Eric E. Conn 1 New Insights into Nitric Oxide Signaling in Plants Angélique Besson-Bard, Alain Pugin, and David Wendehenne 21 Plant Immunity to Insect Herbivores Gregg A. Howe and Georg Jander 41 Patterning and Polarity in Seed Plant Shoots John L. Bowman and Sandra K. Floyd 67 Chlorophyll Fluorescence: A Probe of Photosynthesis In Vivo Neil R. Baker 89 Seed Storage Oil Mobilization Ian A. Graham 115 The Role of Glutathione in Photosynthetic Organisms: Emerging Functions for Glutaredoxins and Glutathionylation Nicolas Rouhier, Stéphane D. Lemaire, and Jean-Pierre Jacquot 143 Algal Sensory Photoreceptors Peter Hegemann 167 Plant Proteases: From Phenotypes to Molecular Mechanisms Renier A.L. van der Hoorn 191 Gibberellin Metabolism and its Regulation Shinjiro Yamaguchi 225 Molecular Basis of Plant Architecture Yonghong Wang and Jiayang Li 253 Decoding of Light Signals by Plant Phytochromes and Their Interacting Proteins Gabyong Bae and Giltsu Choi 281 Flooding Stress: Acclimations and Genetic Diversity J. Bailey-Serres and L.A.C.J. Voesenek 313 v

41 Roots, Nitrogen Transformations, and Ecosystem Services Louise E. Jackson, Martin Burger, and Timothy R. Cavagnaro 341 A Genetic Regulatory Network in the Development of Trichomes and Root Hairs Tetsuya Ishida, Tetsuya Kurata, Kiyotaka Okada, and Takuji Wada 365 Molecular Aspects of Seed Dormancy Ruth Finkelstein, Wendy Reeves, Tohru Ariizumi, and Camille Steber 387 Trehalose Metabolism and Signaling Matthew J. Paul, Lucia F. Primavesi, Deveraj Jhurreea, and Yuhua Zhang 417 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Auxin: The Looping Star in Plant Development René Benjamins and Ben Scheres 443 Regulation of Cullin RING Ligases Sara K. Hotton and Judy Callis 467 Plastid Evolution Sven B. Gould, Ross F. Waller, and Geoffrey I. McFadden 491 Coordinating Nodule Morphogenesis with Rhizobial Infection in Legumes Giles E.D. Oldroyd and J. Allan Downie 519 Structural and Signaling Networks for the Polar Cell Growth Machinery in Pollen Tubes Alice Y. Cheung and Hen-ming Wu 547 Regulation and Identity of Florigen: FLOWERING LOCUS T Moves Center Stage Franziska Turck, Fabio Fornara, and George Coupland 573 Plant Aquaporins: Membrane Channels with Multiple Integrated Functions Christophe Maurel, Lionel Verdoucq, Doan-Trung Luu, and Véronique Santoni 595 Metabolic Flux Analysis in Plants: From Intelligent Design to Rational Engineering Igor G.L. Libourel and Yair Shachar-Hill 625 Mechanisms of Salinity Tolerance Rana Munns and Mark Tester 651 Sealing Plant Surfaces: Cuticular Wax Formation by Epidermal Cells Lacey Samuels, Ljerka Kunst, and Reinhard Jetter 683 Ionomics and the Study of the Plant Ionome David E. Salt, Ivan Baxter, and Brett Lahner 709 vi Contents

42 Alkaloid Biosynthesis: Metabolism and Trafficking Jörg Ziegler and Peter J. Facchini 735 Genetically Engineered Plants and Foods: A Scientist s Analysis of the Issues (Part I) Peggy G. Lemaux 771 Indexes Cumulative Index of Contributing Authors, Volumes Cumulative Index of Chapter Titles, Volumes Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Errata An online log of corrections to Annual Review of Plant Biology articles may be found at Contents vii

43 Flooding-Stress in Plants Martin M Sachs, Agricultural Research Service, United States Department of Agriculture, Urbana, Illinois, USA Flooding is a stress that crop plants frequently encounter and results in huge economic losses. Studies on the effects of flooding at the molecular level may lead to improvements that will result in flood-tolerant crop plants.. Introduction Advanced article. The Anaerobic Response Article Contents. Role of the ANPs in Anaerobic Metabolism. Other Responses to Anoxia and Flooding Tolerance Mechanisms in Plants. Future Research. Biography Information Online posting date: 15 th December 2008 Introduction When heavy rains occur shortly after planting a crop, leaving standing water in the field, it can lead to unacceptable losses and may necessitate replanting. Waterlogged soils are rapidly depleted of oxygen by microorganisms, leaving the germinating seedlings in an anoxic environment. During the 1993 flooding in the midwestern United States about 20 million acres of corn and soybean were inundated, leading to heavy economic losses. Measures to enhance a crop plant s tolerance to this type of flooding would be very beneficial to farmers. The response of plants (especially maize; Zea mays ssp. mays) to anoxia has been the subject of extensive genetic, physiological, biochemical and molecular analyses that may eventually lead to such improvements. See also: Plant Stress Physiology The Anaerobic Response Anaerobic treatment of maize seedlings drastically alters the profile of total protein synthesis. Initially during anoxia, there is an immediate and complete repression of preexisting (aerobic) protein synthesis along with a rapid increase in the synthesis of a small class of polypeptides ( 33 kda, the transition polypeptides; TPs). After 90 min of anoxia, the synthesis of a second class of 20 anaerobic proteins (ANPs) is induced. After 5 h these 20 ANPs are selectively synthesized, and account for more than 70% of the total translation, whereas TP synthesis becomes negligible (Figure 1). Regulation Regulation of both transcription and translation is involved in this response. At the translation level, anaerobic ELS subject area: Plant Science How to cite: Sachs, Martin M (December 2008) Flooding-Stress in Plants. In: Encyclopedia of Life Sciences (ELS). John Wiley & Sons, Ltd: Chichester. DOI: / a pub2 treatment of maize seedlings disrupts most polysomes and leads to a redirection of protein synthesis. After 72 h, protein synthesis decreases concurrently with the start of cell death. At the transcription level, genes encoding the ANPs are activated by anoxia, whereas other genes are repressed. Putative regulatory sequences have been found in the promotors of a number of anaerobically induced genes. Ca 2+ appears to be a signal used to transfer information that a plant is being flooded from the mitochondria to the nucleus and starts the cascade of events leading to the anaerobic response. See also: Calcium Signalling and Regulation of Cell Function Other tissues In the presence of air, each maize organ examined synthesizes a tissue-specific spectrum of polypeptides. Most organs (e.g. the scutellum and endosperm of the developing kernel) actually synthesize many or all of the ANPs constitutively, along with many other proteins under aerobic conditions. Only the root tissue of mature plants and seedlings and other pre-emergent seedling organs (e.g. the coleoptile) show a dramatic increase in the level of synthesis of the ANPs by anaerobic induction. However, under anoxia, most other organs examined also selectively synthesize the ANPs. Moreover, except for a few characteristic qualitative and quantitative differences, the patterns of ANP synthesis in diverse organs are remarkably similar. This response has also been observed in maize tissue culture. A similar shift in the pattern of protein synthesis due to anaerobiosis has been observed in root tissue of many other plant species. A notable exception is maize leaves, which cannot synthesize protein under anoxia, and do not survive even a brief exposure to anaerobic conditions. Hypoxia versus anoxia Hypoxia (low oxygen) has also been found to induce the synthesis of the ANPs in maize primary roots, but does not cause the complete repression of preexisting protein synthesis observed during anoxia. Significant levels of aerobic protein synthesis are still observed at oxygen concentrations as low as 0.2%. In addition, a novel set of polypeptides, not normally observed under This is a US Government work and is in the public domain in the United States of America. ENCYCLOPEDIA OF LIFE SCIENCES. John Wiley & Sons, Ltd 1

44 Flooding-Stress in Plants Figure 1 Protein synthesis in a maize primary root during anaerobic treatment. Fluorographs of native sodium dodecyl sulfate (SDS) 2-D polyacrylamide gels that were loaded with extracts from maize primary roots after the following treatments. (a) One hour pulse labelling with [3 H]-leucine under aerobic conditions. (b e) Pulse labelling with [3 H]-leucine during the specified times under anaerobic conditions. The arrow labelled TPs indicates the position of the transition polypeptides. The unlabelled arrow indicates the position of alcohol dehydrogenase-1 (ADH1). MW, molecular mass. aerobic or anaerobic conditions, is synthesized under hypoxic conditions. Additionally, a set of ethylene responses is induced by hypoxia, but not anoxia. See also: Ethylene As with leaves, it has been found that the maize primary root tip is also very sensitive to anoxic treatment, surviving only a few hours in the absence of oxygen. However, it has been shown that a hypoxic pretreatment allows survival of the root tip for a few days under subsequent anoxia. In contrast, the more proximal part of the primary root and the pre-emergent shoot (i.e. with an intact coleoptile) can survive anoxia for a few days without a hypoxic pretreatment. Besides this reprogramming of gene expression in a coordinated fashion, metabolic (e.g. switch to a fermentative pathway) and structural (e.g. aerenchyma formation) changes also occur during flooding. See also: Anaerobic Respiration Hypoxia also occurs in plant organs that are bulky and/ or have fast metabolic rates, under otherwise well aerated conditions. Role of the ANPs in Anaerobic Metabolism Most of the ANPs identified were found to be enzymes of glycolysis or sugar-phosphate metabolism such as aldolase enolase, glucose-6-phosphate isomerase, glyceraldehyde- 3-phosphate dehydrogenase, sucrose synthase and alcohol dehydrogenase (ADH). In addition, the induction of transcription and enzyme activity of pyruvate decarboxylase (PDC) in maize, and lactate dehydrogenase (LDH) in barley, have been reported during anaerobic stress, indicating that they may represent other ANPs. Synthesis of these enzymes appears to be an adaptation allowing some adenosine triphosphate (ATP) production when oxygen is limited. Alcohol dehydrogenase Molecular studies of the maize anaerobic response stemmed from the extensive analysis of maize ADH genetics. ADH activity is essential for the survival of maize 2 This is a US Government work and is in the public domain in the United States of America. ENCYCLOPEDIA OF LIFE SCIENCES. John Wiley & Sons, Ltd

45 Flooding-Stress in Plants seeds and seedlings during flooding. Normal seedlings survive 3 days of anoxia, while mutants that lack ADH survive only a few hours of flooding. ADH is the major terminal enzyme of fermentation in plants and is responsible for recycling NAD + during anoxia. Most animal cells are very sensitive to anoxia. One possible reason for this is that animals rely solely on lactic acid fermentation, since they lack the enzyme PDC that catalyses a necessary intermediate step in ethanolic fermentation. A model proposed by Davies (1980) postulates that there is a tight regulation of ph to prevent cytoplasmic acidosis during anoxia in plants. Initially upon anaerobiosis, LDH is active, causing a drop in cytoplasmic ph with the accumulation of lactic acid. However, as the ph drops LDH (with a relatively high ph optimum) activity decreases while PDC (with a relatively low ph optimum) activity increases. Thus, it is proposed that after a short period of oxygen deprivation, ethanolic fermentation would predominate in plants. Thus, as in animal cells, maize ADH-null mutants might rely only on LDH for recycling NAD + and succumb to cytoplasmic acidosis. Alternatively, a plant that is ADH-null may accumulate toxic levels of acetaldehyde, the substrate for ADH produced by the enzymatic reaction catalysed by PDC. A mutation (atn1), which allows ADH-null seedlings to survive 24 h of anoxia, may allow further understanding of the role of ADH in flooding tolerance. See also: Oxidative Phosphorylation Although some minimum level of ADH activity appears to be required for any flooding tolerance, natural variation observed in long-term seedling survival of anaerobic stress does not appear to be correlated with variation in the levels of ADH activity. Variation in flooding tolerance has been found among maize lines. When maize lines are compared, survival during anoxia ranges from 2 to 6 days, and this tolerance appears to be inherited as a simple dominant trait. The lines showing variation in flooding tolerance also vary in induced ADH activity levels over a fivefold range, but there is no apparent relationship between these two traits. Other Responses to Anoxia and Flooding Tolerance Mechanisms in Plants Most genes characterized in plants that are induced by oxygen deprivation (anoxia or hypoxia) encode enzymes of glucose-phosphate metabolism (mostly glycolysis and fermentation), and thus function to allow some energy production in the face of limited oxygen supply. In fact, most research on molecular aspects of oxygen deprivation responses in plants has focused on the enzymology of energy production that aids in short-term survival of flooding. However, flooding also induces structural modifications, such as the formation of cortical intercellular air spaces (aerenchyma). These modifications can facilitate oxygen diffusion to submerged tissues and, as such, are generally proposed to aid in prolonging survival during flooding. Other anoxia-induced maize genes have been identified that are not involved in glucose-phosphate metabolism. These genes, with unique kinetics of mrna (messenger ribonucleic acid) induction under anaerobic conditions compared to other anoxia-induced genes, were not induced by heat, cold or salt stress, or by seedling death. Thus, their induction appears to be specific to oxygen deprivation. The function of one of these (xet1) has been identified as encoding a xyloglucan endotransglycosylase, a putative cell wall loosening and degradation enzyme that is proposed to play a role in wall metabolism during germination, cell expansion and fruit ripening. This enzyme may be involved in aerenchyma formation during flooding. The mechanism of aerenchyma formation is through selective degradation of some cells in the cortex and is termed, lysigeny, i.e. programmed cell death (PCD) and it is mainly induced by low oxygen concentrations in the soil, coordinated with induced ethylene accumulation, during such events as excessive rain, irrigation or flooding. Specific cellulase and protease isoforms are also induced, in maize roots, by oxygen deprivation. These may also be involved in morphological changes that occur in plants during flooding, such as root tip death. See also: Plant Salt Stress; Plant Temperature Stress Another mechanism of aerenchyma formation called schizogeny prevails in various plant species where gas-filled spaces are formed by controlled cell division and expansion. This is more characteristic of plants inhabiting excessively wet and flooded soils and under such circumstances the process is predominantly a constitutive event. Adventitious root formation at the soil surface is another adaptation to flooding conditions. This characteristic allows the root system to obtain oxygen directly from the air because the adventitious roots develop at or immediately below the waterlogged soil surface. Other mechanisms such as use of alternative electron acceptors, changes in mitochondrial untrastructure and enhanced tolerance to oxidative stress during recovery from anoxia are also being investigated in roles that confer increased flooding tolerance to plants. Future Research In addition to variation found in maize, some plants have been found to be extremely tolerant to long-term flooding. Natural tolerance to flooding is found in vegetation growing in flood plains, swamps and other areas that are frequently waterlogged. Z. mays ssp. huehuetenangensis, Z. luxurians and especially Z. nicaraguensis, close relatives of maize, all show excellent tolerance to flooding. These teosintes differ from cultivated maize in having constitutive schizogenous aerenchyma, adventitious root formation at the soil surface and show tolerance to Fe 2+,H 2 S in waterlogged soils (Mano and Omori, 2007). Other flooding tolerant grass species include rice (Oryza sativa) and barnyard grass (Echinochloa phyllopogon) whose seeds can germinate under anaerobic conditions. A variant allele This is a US Government work and is in the public domain in the United States of America. ENCYCLOPEDIA OF LIFE SCIENCES. John Wiley & Sons, Ltd 3

46 Flooding-Stress in Plants (Sub1A-1) in rice enhances submergence tolerance (Xu et al., 2006). Other grasses, such as reed mannagrass (Glyceria maxima), common reed (Phragmites australis) and plants such as American cranberry (Vaccinium macrocarpon), sweet flag (Acorus calamus) and pondweeds (in the genus Potamogeton), have also developed mechanisms for long-term survival in standing water. Studies of the mechanisms that flooding tolerant plants possess are underway in order to gain an understanding as to how these plants cope with extended periods of oxygen deprivation. Additional studies are underway to understand how some plants can alter their metabolism so that ATP consumption is minimized at oxygen concentrations well above those actually limiting for respiration (either during flooding or in cases of localized hypoxia). Coupled with further knowledge of how other more susceptible plants respond to anoxia, understanding mechanisms involved in long-term survival during anaerobiosis will ultimately allow molecular breeders to produce flooding tolerant varieties of crops such as maize. See also: Plant Breeding and Crop Improvement Biography Information Martin M. Sachs, PhD Genetics 1981, University of California, Berkeley. Discovered the anaerobic response (selective translation of ANPs) in plants. Research on maize flooding-stress has focused on understanding the genes involved in this response and signal transduction events allowing plants to perceive oxygen deprivation and leading to the redirection in transcription and translation that results and related adaptive morphological changes that occur. References Davies DD (1980) Anaerobic metabolism and production of organic acids. In: Stumpf PK and Conn EE (eds) The Biochemistry of Plants: A Comprehensive Treatise, vol. 2, pp New York: Academic Press. Mano Y and Omori F (2007) Breeding for flooding tolerant maize using teosinte as a germplasm resource. Plant Root 1: Xu K, Xu K, Fukao T et al. (2006) Sub1A is an ethylene responsive-factor-like gene that confers submergence tolerance to rice. Nature 442: Further Reading Bailey-Serres J and Chang R (2005) Sensing and signalling in response to oxygen deprivation in plants and other organisms. Annals of Botany 96: Bailey-Serres J and Voesenek LACJ (2008) Flooding stress: acclimations and genetic diversity. Annual Review of Plant Biology 59: Geigenberger P (2003) Response of plant metabolism to too little oxygen. Current Opinion in Plant Biology 6: Jackson MB (2008) Ethylene-promoted elongation: an adaptation to submergence stress. Annals of Botany 101: Sachs MM and Vartapetian BB (2007) Plant anaerobic stress I. Metabolic adaptation to oxygen deficiency. Plant Stress 1: Subbaiah CC and Sachs MM (2008) Responses to oxygen deprivation and potential for enhanced flooding tolerance in maize. In: Bennetzen JL and Hake SC (eds) Handbook of Maize: Its Biology. New York: Springer (in press), Vartapetian BB and Crawford RMM (2007) The international society for plant anaerobiosis: history, functions and activity. Plant Stress 1: 1 3. Vartapetian BB, Sachs MM and Fagerstedt KV (2008) Plant anaerobic stress II. Strategy of avoidance of anaerobiosis and other aspects of plant life under hypoxia and anoxia. Plant Stress 2: Voesenek LACJ, Colmer TD, Pierik R, Millenaar FF and Peeters AJM (2006) How plants cope with complete submergence. New Phytologist 170: This is a US Government work and is in the public domain in the United States of America. ENCYCLOPEDIA OF LIFE SCIENCES. John Wiley & Sons, Ltd

47 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Annu. Rev. Plant Biol : The Annual Review of Plant Biology is online at plant.annualreviews.org This article s doi: /annurev.arplant Copyright c 2008 by Annual Reviews. All rights reserved /08/ $20.00 Plant Aquaporins: Membrane Channels with Multiple Integrated Functions Christophe Maurel, Lionel Verdoucq, Doan-Trung Luu, and Véronique Santoni Biochimie et Physiologie Moléculaire des Plantes, SupAgro/INRA/CNRS/UM2 UMR 5004, F Montpellier Cedex 1, France; maurel@supagro.inra.fr, verdoucq@supagro.inra.fr, luu@supagro.inra.fr, santoniv@supagro.inra.fr Key Words environmental stress, gating, major intrinsic protein, nutrient, transport selectivity, water relations Abstract Aquaporins are channel proteins present in the plasma and intracellular membranes of plant cells, where they facilitate the transport of water and/or small neutral solutes (urea, boric acid, silicic acid) or gases (ammonia, carbon dioxide). Recent progress was made in understanding the molecular bases of aquaporin transport selectivity and gating. The present review examines how a wide range of selectivity profiles and regulation properties allows aquaporins to be integrated in numerous functions, throughout plant development, and during adaptations to variable living conditions. Although they play a central role in water relations of roots, leaves, seeds, and flowers, aquaporins have also been linked to plant mineral nutrition and carbon and nitrogen fixation. 595

48 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org MIP: major intrinsic protein PIP: plasma membrane intrinsic protein TIP: tonoplast intrinsic protein NIP: nodulin- 26 like intrinsic protein SIP: small basic intrinsic protein Contents INTRODUCTION MOLECULAR AND CELLULAR PROPERTIES The Plant Aquaporin Family Mechanisms of Transport Molecular Mechanisms of Regulation AQUAPORIN FUNCTIONS THROUGHOUT PLANT GROWTH AND DEVELOPMENT Water Transport Nitrogen, Carbon, and Micronutrient Acquisition AQUAPORINS IN A VARIABLE ENVIRONMENT Changes in Irradiance Water, Salt, and Nutrient Stresses. 608 Cold Stress Anoxia Biotic Interactions CONCLUSIONS INTRODUCTION Aquaporins are small integral membrane proteins that belong to the ancient family of major intrinsic proteins (MIPs), with members in animals, microbes, and plants. Fifteen years after their discovery in plants, it now appears that studies on aquaporins have provided unique perspectives into multiple integrated aspects of plant biology. Aquaporins first raised considerable interest because of their water channel activity. This finding was unexpected in plants and, although it may not have induced a real paradigm shift in understanding of membrane water transport (124, 140), it led researchers to revisit many aspects of plant water relations and to link these aspects to novel physiological contexts. More recently, MIPs were proved to be more than water channels (141), and other transport substrates of great physiological significance have been identified. Although the term aquaporin was initially restricted to water-transporting MIPs, we now use this term in a broader sense, referring to all plant MIPs as aquaporins. The molecular and cellular properties of aquaporins were recently reviewed in detail (18, 91). The aim of the present review is to examine how a wide range of selectivity profiles and regulation properties allows aquaporins to be integrated in numerous functions throughout plant development and during adaptations to variable environmental conditions. MOLECULAR AND CELLULAR PROPERTIES The Plant Aquaporin Family Subfamilies. Plant aquaporins show a high multiplicity of isoforms, with 35 and 33 homologs in Arabidopsis and rice, respectively (62, 113, 120). On the basis of sequence homology, aquaporins in most plant species can be divided into four subgroups. The plasma membrane intrinsic proteins (PIP) (with two phylogenic subgroups, PIP1 and PIP2, and 13 isoforms in Arabidopsis) and the tonoplast intrinsic proteins (TIP) (10 homologs in Arabidopsis) are the most abundant aquaporins in the plasma membrane and vacuolar membrane (tonoplast), respectively (62, 113). The third subfamily comprises the nodulin- 26 like intrinsic membrane proteins (NIPs), which were named after soybean (Glycine max) nodulin-26 (GmNOD26), an abundant aquaporin expressed in the peribacteroid membrane of N 2 -fixing symbiotic root nodules. NIPs are also present in nonlegume plant species (9 homologs in Arabidopsis) (149). A fourth class comprises small basic intrinsic proteins (SIPs) (3 homologs in Arabidopsis) (56, 62, 113). Although these four classes are conserved among all plant species, the aquaporin gene family shows signs of rapid and recent evolution and orthologs cannot necessarily be distinguished between species (120). In addition, some plant species have acquired additional, novel types of aquaporins. 596 Maurel et al.

49 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org For instance, a homolog of the bacterial glycerol facilitator GlpF has been acquired by the moss Physcomitrella patens by horizontal gene transfer (45), and the genome of this organism and some higher plants (such as poplar) encodes a fifth class of aquaporins, which are closely related to but yet clearly distinct from PIPs (139; U. Johanson, personal communication). Subcellular localization. Plant aquaporins localize in all subcellular compartments forming or derived from the secretory pathway. This broad localization pattern reflects the high degree of compartmentation of the plant cell and the need for the cell to control water and solute transport not only across the plasma membrane but also across intracellular membranes. Similar to PIPs, some NIPs localize in the plasma membrane (82, 134). By contrast, the three Arabidopsis SIP homologs reside mainly in the endoplasmic reticulum (56). However, aquaporins cannot simply be assigned to homogeneous subcellular compartments. For instance, immunocytochemical studies using isoform-specific anti-tip antibodies revealed that distinct types of vacuole that can coexist in the same cell are equipped with specific combinations of TIP isoforms; TIP1 and TIP2 isoforms are preferentially associated with the large lytic vacuoles and vacuoles accumulating vegetative storage proteins, respectively (59). More recently, Arabidopsis thaliana AtTIP1;1 was shown to accumulate in spherical structures named bulbs, tentatively identified as intravacuolar invaginations made of a double tonoplast membrane (118). Preferential expression of PIPs in plasmalemmasomes (convoluted plasma membrane invaginations that dip into the vacuole) has also been observed in Arabidopsis leaves (116). Finally, preferential expression of a PIP and a NIP homolog on the distal side of root exo- and endodermal cells has been described in maize and rice, respectively (46, 82). Such cell polarization is consistent with the uptake and centripetal transport of water and solute in roots (see below). A future challenge is to understand how aquaporins can be specifically targeted to membrane subdomains in the plant cell and how targeting contributes to their functional specialization. Mechanisms of Transport Pore structure and transport mechanisms. X-ray crystallography determination of atomic structures of microbial, animal, and plant homologs points to highly conserved structural features in the aquaporin family (38, 137). Aquaporins are kda proteins comprising six membrane-spanning domains tilted along the plane of the membrane and linked by five loops (A to E ) located on the intra- (B, D) or extracytoplasmic (A, C, E ) side of the membrane. The N- and C-terminal extremities are both exposed to the cytosol (Figure 1). A central aqueous pore is delineated by the transmembrane domains and loops B and E, which both carry a conserved Asn-Pro-Ala (NPA) motif and dip from either side of the membrane into the center of the molecule. Projection structures determined by cryo-electron microscopy indicate that, similar to their animal and microbial counterparts, PIPs and TIPs occur as tetramers in their native membranes (24, 34). X-ray structures have confirmed this type of assembly (38, 137) and in combination with molecular dynamics simulations have provided critical insights into the fundamental principles of aquaporin transport selectivity (38, 133) (Figure 1). In brief, the substrate specificity of aquaporins can be explained by several mechanisms, including size exclusion at two main pore constrictions [aromatic/arg (Ar/R) and NPA] and stereospecific recognition of the substrate mediated by spatially defined H-bonding and hydrophobic interactions within the pore. The remarkable impermeability of aquaporins to protons is explained by electrostatic repulsion, dipole orientation, and transient isolation of the water molecule as it passes within a single Plasmalemmasome: convoluted plasma membrane invagination Plant Aquaporins 597

50 a b Phe81 Asn101 His210 Arg225 Asn222 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Figure 1 Representative atomic structure of a plant aquaporin (a) and general molecular mechanisms of transport selectivity (b). (a) Structure of the open conformation of Spinacia oleracea plasma membrane intrinsic protein 2;1 (SoPIP2;1) [Protein Data Bank (PDB) ID 2B5F] (137) showing a typical tetrameric arrangement. Each monomer is composed of six tilted transmembrane helices; the N-terminal (red ) and C-terminal ( green) helices of the top left monomer are shown. The pores of individual monomers are emphasized by the space-filling representation of the three other monomers. (b) The two highly conserved Asn-Pro-Ala (NPA) motifs (represented by Asn101 and Asn222, green) are in close proximity to form one of the main pore constrictions. Another constriction called Ar/R (red ) is formed on the extracytoplasmic side of the membrane by a spatial arrangement of aromatic (Ar) residues, such as Phe81 and His210, facing an Arg (R) residue, here Arg225. Proton transport is blocked by electrostatic repulsion in the Ar/R constriction and the dipole orientation of the water molecule by the two Asn residues of the NPA motifs. This results in a transient isolation of the water molecule within the single file of water molecules that fills the pore (orange spheres). file of water molecules through the center of the pore (11, 38, 133) (Figure 1). The molecular basis of plant aquaporin selectivity has been investigated more specifically by homology modeling of pore structures at the Ar/R constriction (8, 150). Analysis of all 35 Arabidopsis homologs yielded up to nine pore types (150) and additional types exist in maize and rice (8). Whereas all PIPs exhibit a narrow pore structure typical of orthodox, water-selective aquaporins, larger substrate specificity was predicted for other plant homologs. According to this analysis, AtNIP6;1 belongs to one of two NIP subgroups and as such exhibits a low and high permeability to water and urea, respectively (151). An Ala119Trp substitution, made to mimic the pore configuration of members of the other NIP subgroup, also confers novel permeability properties, i.e., higher permeability to water and failure to transport urea. This result and other examples in animal aquaporins (11) show that point mutations can drastically alter transport specificity and that these proteins may be engineered to accommodate novel substrates of interest. Transport assays and aquaporin substrates. Functional expression in Xenopus oocytes or yeast was essential to show that plant MIP homologs of all four subclasses can function as water channels (56, 66, 94, 115). Enhanced water permeability of proteoliposomes containing a purified aquaporin provides the ultimate proof of water channel activity. Such functional reconstitution has been performed with GmNOD26 purified 598 Maurel et al.

51 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org from native peribacteroid membranes (27) or after production of Spinacia oleracea SoPIP2;1 in Pichia pastoris (67). Although strict comparative measurements have not been performed in plants, plant aquaporins may, similar to their animal homologs, exhibit marked differences (up to 30-fold) in intrinsic water transport activity (154). Expression studies in Xenopus oocytes also show that, similar to animal and bacterial aquaglyceroporins, some plant aquaporin isoforms can transport small neutral solutes such as glycerol (12), urea (42), formamide, acetamide (115), methylammonium (53), boric acid (134), silicic acid (82), or lactic acid (20). Ammonia (NH 3 ) and CO 2 transport is detected using substrate-induced extra- and intracellular acidification, respectively, whereas ammonium (NH 4 + ) transport by Triticum aestivum TaTIP2;1 results in inward currents (53, 144). Finally, expression in yeast cells deficient in endogenous systems responsible for urea or hydrogen peroxide uptake has proved efficient to screen, on the basis of a survival assay, aquaporin isoforms that possibly transport these molecules; these properties are subsequently confirmed by true transport assays (13, 77). Several approaches have established that aquaporins contribute significantly to the permeability of plant membranes to water and small neutral solutes. In most studies, mercury derivatives, which act through oxidation and binding to Cys residues, were used as common aquaporin blockers. Plant aquaporins do not have Cys residues at conserved positions and various residues may be involved in plant aquaporin inhibition (23). We also note that mercury-resistant PIPs have been described in Arabidopsis and tobacco (12, 25). In some studies, the permeability profiles of the vacuolar, peribacteroid, and plasma membranes were characterized by stopped-flow spectrophotometry on purified membrane vesicles, and mercury induced a marked (50% 90%) inhibition of water transport in the first two types of membranes (42, 95, 104, 105, 115). In addition, a good parallel was established between the high permeability of the tobacco tonoplast and soybean peribacteroid membrane to urea and formamide, respectively, and the capacity of Nicotiana tabacum NtTIPa and GmNOD26 to transport these solutes (42, 115). In other studies, the respective water permeabilities of the plasma membrane and the tonoplast and their sensitivity to mercury were inferred from independent osmotic swelling assays on protoplasts and isolated vacuoles and calculations using a three-compartment model (92, 99, 102). Figure 2 summarizes the contribution of plant aquaporins to water and solute transport in multiple subcellular compartments. Molecular Mechanisms of Regulation Cotranslational and posttranslational modifications. Because of their high abundance in plant membranes, and despite their high hydrophobicity, some aquaporins have proved to be particularly amenable to biochemical analysis, in comparison with other membrane proteins (34, 48, 63). Proteomics, and mass spectrometry techniques in particular, have recently been added to more classical techniques to produce a thorough description of aquaporin co- and posttranslational modifications (26, 121, 122). For instance, N-terminal maturation of PIP1s and PIP2s occurs through N-α-acetylation or cleavage of the initiating residue, respectively (121). In vivo and in vitro labeling studies, experiments with antiphosphopeptide antibodies, and mass spectrometry analyses have provided direct evidence for phosphorylation of Ser residues in the N-terminal and C- terminal tails of Phaseolus vulgaris PvTIP3;1, GmNOD26, and SoPIP2;1 (26, 43, 63, 64, 96). PIPs show a conserved phosphorylation site in loop B and multiple (up to three) and interdependent phosphorylations occur in adjacent sites of their C-terminal tail (63, 64; S. Prak, S. Hem, J. Boudet, N. Sommerer, G. Viennois, M. Rossignol, C. Maurel & V. Santoni, unpublished results). Purification of calcium-dependent protein Plant Aquaporins 599

52 Multivesicular body / late endosome / prevacuolar compartment Protein storage vacuole Cell wall Plasma membrane Vegetative protein storage vacuole Peribacteroid membrane CO 2 Nucleus and ER Chloroplast Golgi apparatus Early endosome Ammonia Water Hydrogen peroxide Water Lytic/central vacuole Glycerol Urea Water Ammonia? Water Vacuolar bulb Water Water Lactic acid Boric acid Silicic acid Plasmalemmasome PIP1s TIP1s AtNIP2;1 SIPs PIP2s TIP2s Lsi1/OsNIP2;1

53 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org the first plant membrane proteins found to be methylated (121). For instance, AtPIP2;1 can carry one or two methyl groups on its Lys3 and Glu6 residues, respectively. These data show that, in addition to a high isoform multiplicity, plant aquaporins occur in a large variety of modified forms, which suggests intricate coand posttranslational regulation mechanisms. Gating. The gating of aquaporins, i.e., the opening and closing of the pore, can be regulated by multiple factors. A role for phosphorylation in gating PvTIP3;1, GmNOD26, and SoPIP2;1 was first deduced from functional expression in oocytes of these aquaporins, either wild-type or with point mutations at their phosphorylation sites (43, 63, 93), and by using pharmacological alterations of endogenous protein phosphatases and kinases. A role for phosphorylation in GmNOD26 gating has been unambiguously established by stopped-flow measurements in purified peribacteroid membranes, showing that alkaline phosphatase-mediated dephosphorylation leads to reduced water permeability (43). Water transport measurements in plasma membrane vesicles purified from Arabidopsis suspension cells or Beta vulgaris roots also suggest that PIPs can be gated from the cytosolic side by protons and divalent cations (4, 41). A half-inhibition of water transport is observed at ph 7.5 and for free Ca 2+ concentrations in the 100 μm range (4, 41). Beet plasma membranes exhibit an additional affinity component in the 10 nm range (4). The molecular bases of aquaporin gating have been elucidated from structure-function analyses in Xenopus oocytes and more recently from the atomic structures of SoPIP2;1 in its open and closed conformations (137, 138). These studies established that protons are sensed by a His residue that is perfectly conserved in loop D of all PIPs (138). The molecular mechanisms that lead to a conformational change of loop D and occlusion of the pore upon protonation of the His residue or binding of divalent cations are detailed in Figure 3. The atomic structure of SoPIP2;1 also indicates how phosphorylation of loop B would unlock loop D to allow the open conformation. By contrast, phosphorylation of the C-terminal tail would act in trans to prevent loop D of an adjacent monomer from adopting a closed-pore conformation (137). A role for solutes in gating aquaporins has been proposed, based mainly on pressure probe measurements in Chara cells (155). Inhibition of cell water permeability is linked to the presence of the solute on either side of the membrane and is strongly dependent on solute molecular size. A tension/cohesion model Figure 2 The multiple cellular functions of plant aquaporins. The figure illustrates the variety of transport functions achieved by aquaporins in various subcellular compartments. The different aquaporin subclasses or isoforms are identified below the illustration in distinct colors. Isoforms of the plasma membrane intrinsic protein 1 (PIP1) and PIP2 subfamilies are thought to follow the secretory pathway, which carries cargo from the endoplasmic reticulum (ER) toward the plasma membrane through the Golgi apparatus. PIPs also undergo repeated cycles of endocytosis and recycling through endosomal compartments before being eventually targeted to the lytic vacuole through the multivesicular body. In Arabidopsis leaves, PIP1s label plasmalemmasomes (116). Tonoplast intrinsic protein 1s (TIP1s) are found in the lytic vacuole membrane. AtTIP1;1 localizes in spherical structures named bulbs in epidermal cells of young cotyledons or salt-treated roots (15, 118). TIP2s and TIP3s are preferentially associated with vacuoles that accumulate vegetative storage proteins and seed protein storage vacuoles, respectively. Nodulin-26 like intrinsic membrane proteins (NIPs) show a broad range of subcellular localization patterns. AtNIP2;1 is localized in the endoplasmic reticulum and the plasma membrane (20, 97), the Oryza sativa silicon influx transporter low silicon rice 1 (Lsi1, also namedosnip2;1) and the Arabidopsis thaliana boric acid channel AtNIP5;1 are localized in the plasma membrane, whereas Glycine max nodulin-26 (GmNOD26) is exclusively expressed in the peribacteroid membrane. Gating: opening and closing of a membrane channel pore Plant Aquaporins 601

54 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Arg118 Ser115 Glu31 Asp28 Arg118 Ser115 Glu31 Asp28 Figure 3 a b Loop B His193 Loop D Loop D Leu197 His193 Leu197 Molecular mechanisms of plasma membrane intrinsic protein (PIP) gating. The Spinacia oleracea SoPIP2;1 structure was solved in an open (a) [Protein Data Bank (PDB) ID 2B5F] and in a closed (b) (PDB ID 1Z98) conformation (137). The His193 residue ( green) is perfectly conserved in loop D of all PIPs. In the open conformation (a), His193 is not protonated and loop D is distal from the other cytoplasmic loop B. By contrast, the protonation of His193 (b) allows interaction with an acidic residue of the N terminus, Asp28 ( purple). This in turn drives a conformational change of loop D and occlusion of the pore by displacement of the hydrophobic side chain of Leu197 ( yellow) into the cytosolic pore mouth. Binding of divalent cations [Cd 2+ in the atomic structure, ( purple sphere)] would also involve Asp28 and an adjacent acidic residue (Glu31). Loop D would then be stabilized in the closed pore conformation through a network of H-bond and hydrostatic interactions, involving Arg118. In this model, phosphorylation of loop B, at Ser115 ( pink), would disrupt this network of interactions and unlock loop D to allow the open conformation. was proposed in which exclusion of the solute from the narrow vestibule of the pore would result in osmotic forces and tensions, which in turn would collapse the pore (155). Hydroxyl radicals also induce a marked ( 90%) and reversible inhibition of water transport in Chara cells, which was interpreted in terms of direct oxidative gating of aquaporins (50). By contrast, the inhibition of aquaporins by reactive oxygen species (ROS) in the Arabidopsis root seems to involve cell signaling mechanisms (Y. Boursiac, J. Boudet, O. Postaire, D.-T. Luu, C. Tournaire-Roux & C. Maurel, unpublished results). Tetramer assembly and cellular trafficking of PIPs. Recent studies have pointed to aquaporin trafficking as a critical point for regulating aquaporin expression and function. The inability of some PIP1 isoforms to be functionally expressed in Xenopus oocytes has been reported by several laboratories. Fetter and coworkers (31) explained this inability by a failure of these aquaporins to traffic to the oocyte plasma membrane, and showed that coexpression of maize PIP1s in Xenopus oocytes with reduced amounts of PIP2 isoforms could alleviate this defect. Affinity copurification and coimmunopurification studies provided the first biochemical evidence that PIP1s and PIP2s physically interact, both in oocytes and plants (31, 158). The results of fluorescence resonance energy transfer (FRET) imaging in living maize protoplasts coexpressing PIP1s and PIP2s further support a model in which aquaporins of the two classes directly interact, very likely by heterotetramerization, to facilitate PIP1 trafficking (158). A possible role for PIP1 phosphorylation was recently added to this model (135). Phosphorylation on loop B of a Mimosa PIP1 is not necessary for aquaporin interaction but enhances the overall water transport activity of PIP1-PIP2 complexes in oocytes (135). Whereas interaction-dependent trafficking of PIP1s and PIP2s offers a broad range of combinatorial regulations, a future challenge is to determine to what extent this 602 Maurel et al.

55 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org process can dominate the expression of PIP1 or PIP2 homotetramers. Antisense inhibition experiments in Arabidopsis of PIP1s and PIP2s, alone or in combination, have suggested that the two classes of aquaporins contribute to the same functional water transport units (90). Similar to other membrane proteins, PIP2 aquaporins are subjected to constitutive cycling. Their endocytosis is clathrindependent (28) and reduced by auxin (108). Export of PIP2 aquaporins from the endoplasmic reticulum is also critically controlled and the role of a di-acidic motif contained in the N-terminal tail of PIP2s was recently uncovered in maize and Arabidopsis (F. Chaumont, personal communication; M. Sorieul, D.-T. Luu, V. Santoni & C. Maurel, unpublished results). The cellular mechanisms that determine aquaporin trafficking and their subcellular relocalization in response to stimuli will surely fuel intense investigations in the coming years. AQUAPORIN FUNCTIONS THROUGHOUT PLANT GROWTH AND DEVELOPMENT Water Transport Principles of plant water transport. A wide range of cell water permeabilities can be observed between distinct cell types and throughout plant development. For instance, cell pressure probe measurements indicate that, in growing epicotyls of pea, cortical cells have a 30-fold higher hydraulic conductivity than epidermal cells (130). Also, swelling assays on isolated protoplasts from rape roots indicate that their osmotic water permeability coefficient increases from 10 to 500 μm sec 1 within less than two days (114). Although the contribution of the lipid membranes should be taken into account, one challenge will be to determine how the aquaporin equipment of each individual cell can determine such strikingly different water transport properties. In these respects, attempts have been made to re- late the cell-specific expression of aquaporin isoforms in radish taproots and maize primary roots and the water permeability of protoplasts derived from the various cell types of these materials (46, 132). In the whole plant, long-distance transport of fluids occurs mostly through vascular tissues, which do not present significant membrane barriers. Yet, living tissues can be the site of intense flows of water during transpiration or expansion growth. For this, water can flow along various paths: (i ) the apoplastic path, i.e., within the cell wall continuum, (ii ) the symplastic route through cytoplasmic continuities and plasmodesmata, and (iii ) the transcellular path across cell membranes (mainly plasma membranes), which in many tissues is mostly mediated by aquaporins. Although it is not very specific and is inactive on certain aquaporins (see above), mercury represents one of the very few tools available to evaluate the contribution of aquaporins to water transport in plant tissues. The general toxicity of this compound in vivo must be carefully evaluated (reviewed in 61), and researchers checked that mercury does not perturb xylem solute transport and respiration in aspen roots. By contrast, mercury depolarizes wheat root cells in parallel to inhibition of water transport (61). A reversibility of mercury effects by reducing agents is also required to show that the blocking effects are due to oxidation mechanisms and not to irreversible damage of the cells. Despite all these restrictions, the effects of mercury on water transport have been characterized in a large variety of physiological contexts. Overall, these studies provide a consistent picture of the role of aquaporins, in particular during root water transport (61). Yet, new specific aquaporin blockers are critically needed. Gold and silver ions have been described as potent aquaporin blockers in vitro but the use of these compounds in vivo seems to be problematic and their mode of action is as yet unknown (106). Transpiration. Because it induces an intense renewal of water throughout the plant, Hydraulic conductivity: water permeability (i.e., intrinsic capacity to transport water) of a membrane, cell, or tissue Plant Aquaporins 603

56 a Maize Epidermis Exodermis Rice b Endodermis Cuticle Epidermis i Palisade mesophyll ii Spongy mesophyll Xylem vessel Phloem sieve tube Bundle sheath Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org ZmPIP2;1/2 ZmPIP2;5 Lp r : root hydraulic conductivity Movement of water Uptake of silicic acid Lsi1/OsNIP2;1 Lsi2 Movement of CO 2 Movement of water Guard cell PIP1s TIP1s Figure 4 Aquaporin-mediated transport of water and solutes in roots (a) and leaves (b). Schematic cross sections with representations of the tissue-specific expression patterns of aquaporins and paths of transport are shown. Aquaporin expression and water transport in maize roots is summarized according to Reference 46, whereas uptake of silicic acid in rice roots by Oryza sativa Lsi1 (OsNIP2;1) in combination with the efflux transporter Low silicon rice 2 (Lsi2) is drawn according to References 82 and 83. Expression of plasma membrane intrinsic protein 1s (PIP1s) and tonoplast intrinsic protein 1s (TIP1s) in Brassica napus leaves was summarized according to Reference 35. The movement of water can follow the cell-to-cell (symplastic and transcellular) (i ) or apoplastic (ii ) path. transpiration represents an obvious context in which to investigate aquaporin function in roots and leaves. A thorough description of cell-specific expression of aquaporins in roots was recently performed in maize using in situ and quantitative RT-PCR and immunolocalization (46). Strong expression of ZmPIP2;1 and ZmPIP2;5 is observed in both the exodermis and endodermis of the mature zone, suggesting that because of the presence of Casparian strips a bypass of the apoplastic path may be necessary in these cell layers (46) (Figure 4a). Also, a strong expression in the stele and vascular tissues is consistently observed in roots of several plant species (46, 124, 132). Several lines of functional evidence also show that aquaporins significantly contribute to water uptake by roots. Firstly, mercury inhibits the root hydraulic conductivity (Lp r ) by 30% 90% in more than ten plant species (61). In addition, antisense inhibition of aquaporins of the PIP1 and/or PIP2 subclasses reduced Lp r by approximately 50% in tobacco or Arabidopsis (90, 128). Finally, two allelic Arabidopsis pip2;2 knockout mutants show, with respect to wild-type plants, a reduction of 25% 30% in hydraulic conductivity of root cortex cells (60). In addition, the osmotic hydraulic conductivity of entire roots, as derived from free exudation measurements, is decreased by 14% in the mutants, pointing to PIP2;2 as an aquaporin specialized in osmotic fluid transport in the Arabidopsis root (60). Further analysis of mutants corresponding to aquaporin genes with distinct cell-specific expression patterns may help 604 Maurel et al.

57 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org dissect the contribution of the various cell layers to distinct modes (i.e., osmotic or hydrostatic) of water transport within the root. The preferential expression of some aquaporin isoforms in vascular tissues and cell types such as tracheary elements, xylem parenchyma cells, and phloem-associated cells suggests a general role of aquaporins in sap transport throughout the plant body (reviewed in 92). In the leaf, the water supply at the evaporating sites is sustained by a flow of liquid water from the vascular system through the extravascular compartment, including the vascular bundles and the mesophyll (Figure 4b). A role for aquaporins in mediating water transfer from the veins to the stomatal chamber has been proposed on the basis of two lines of evidence. Firstly, mercury can inhibit leaf hydraulic conductance (K leaf ) in sunflower and in six temperate deciduous trees (1, 103). Secondly, lightdependent changes in K leaf in walnut occur within one hour, are associated with changes in expression of PIP2 aquaporin transcripts, and both are inhibited by 100 μm cycloheximide, indicating a role for protein synthesis in K leaf regulation (22). In a recent study, mesophyll protoplasts were isolated from Arabidopsis genotypes differing in stomatal aperture or from plants grown at varying relative air humidity (98). Surprisingly, researchers observed an inverse relationship between the rate of transpiration in the plant and the water permeability of the isolated protoplasts. These variations occur without any alteration in the leaf PIP content, suggesting that aquaporin function is controlled at the posttranslational level. The physiological significance of this control and its occurrence in other plant species are as yet unknown. Expression of aquaporins has also been reported in stomatal guard cells (36, 123). Despite the critical role played by these cells in maintaining the whole plant water status, the function of aquaporins in stomatal movements is as yet unclear and deserves more attention for future research. Tissue expansion. Although some aquaporin isoforms seem to be specific to dividing cells (9), a strong link between PIP and TIP aquaporin expression and cell expansion has been observed in numerous plant materials. For instance, expression of the AtTIP1;1 promoter is associated with cell enlargement in Arabidopsis roots, hypocotyls, leaves, and flower stems (81), and transcript accumulation is enhanced by the growth-promoting hormone gibberellic acid (GA3) (109), suggesting that AtTIP1;1 may contribute to the differentiation of a large central vacuole in fully elongated cells (81). Because most plant cells have short halftimes of water exchange (130), water influx into a single plant cell can hardly be limiting during expansion growth. Membrane water transport in growing tissues should rather be envisioned in the context of a transcellular delivery of water from vascular tissues toward peripheral expanding tissues (148). The presence, in certain physiological conditions, of significant water potential gradients within growing tissues supports the idea that this type of water transport can be limiting. Accordingly, mercury blocks tissue growth in maize roots, but exclusively in the older cells, distal to the apex (55). These cells are characterized, with respect to younger cells, by a high, mercury-sensitive cell hydraulic conductivity and reduced symplastic connections with the phloem, suggesting that aquaporin-mediated transcellular water transport is necessary for delivery of water from the phloem into the cells. In castor bean hypocotyls, the transcript abundance of a specific PIP2 isoform and a high hydraulic conductivity of cortical cells both show the same light- and spatialdependence as that of tissue expansion (30). Finally, diurnal epinastic movements (unfolding) of tobacco leaves can be accounted for by a differential growth of the upper and lower surfaces of the petiole. A limiting role for tobacco aquaporin 1 (NtAQP1) in this process was proposed on the basis of observations that expression of NtAQP1 in the petiole shows a diurnal rhythm that coincides K leaf : leaf hydraulic conductance Plant Aquaporins 605

58 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org g m : mesophyll conductance to CO 2 with leaf unfolding and that antisense inhibition of NtAQP1 impairs leaf movement (127). These studies draw a convincing, but qualitative, picture of aquaporin function in expansion growth. Future studies will have to quantitatively integrate growth rates, cell and tissue water relation parameters, and aquaporin expression in time and space (148). Fruit development and ripening, which involve cell enlargement and cotransport of sugar and water, will also represent important processes in which to investigate aquaporin function (19, 110). Tissue desiccation and imbibition. Plant reproduction requires the intense desiccation of certain organs, which then acquire specific dissemination and resistance properties. TIP isoforms specific to pollen grains have been reported in Arabidopsis but their function during pollen desiccation and/or pollen tube growth is as yet unclear (111). By contrast, antisense inhibition of PIP2 aquaporins in tobacco delays anther dehydration and dehiscence, suggesting that these aquaporins are involved in water flow out of the anther via the vascular bundle and/or evaporation (14). Seed germination represents another remarkable context, during early tissue imbibition and subsequent embryo growth, in which aquaporins may mediate a fine temporal and spatial control of water transport; evidence for aquaporin function in seeds is emerging. In Brassica napus, the germination rate of seeds that have gone through various priming treatments is strongly correlated with the transcript abundance of a PIP2 aquaporin (39). In pea and Arabidopsis, mercurials reduce the speed of seed imbibition and seed germination, respectively (145, 147). Finally, expression in tobacco and rice of sense and antisense PIP1 transgenes shows that the speed and extent of germination of seeds in normal and/or water stress conditions is positively correlated with aquaporin expression (76, 156). Liu and coworkers (76) suggested a role for PIP1s in stimulating seed germination by nitric oxide (NO) in rice. Nitrogen, Carbon, and Micronutrient Acquisition Nitrogen fixation. A first link between aquaporins and nitrogen (N) assimilation came from the observation that expression of some aquaporin genes is dependent on N compounds; some genes, such as ZmPIP1;5b, are strongly induced by nitrate (40), whereas others, such as AtTIP2;1, are induced under long-term N starvation or short term NH 4 + supply (79). This type of regulation was first interpreted as a reflection of well-known connections between water relations and N metabolism. However, evidence was recently presented that aquaporins of the PIP, NIP, and TIP subfamilies can transport N compounds. Transport of urea by TIPs (42, 70, 140) may contribute to urea equilibration within the cell and storage into and remobilization from the vacuole (70). Wheat and Arabidopsis TIP2 homologs also show a remarkable permeability to NH 3 and may therefore contribute to significant loading of this compound and acidtrapping of the protonated form (NH 4 + )in the vacuole (53, 57, 79). However, studies of transgenic Arabidopsis that overexpresses AtTIP2;1 failed to establish any significant role for this aquaporin in NH 4 + accumulation (79). CO 2 transport and carbon metabolism. CO 2 transport by aquaporins in planta was first evaluated by treating Vicia faba and Phaseolus vulgaris leaves with mercury (136). In these two experiments, mercury altered the dependency of photosynthesis on intercellular but not on chloroplastic CO 2. This effect was interpreted to mean that CO 2 diffusion into the chloroplast [i.e., mesophyll conductance to CO 2 ( g m )] is blocked and therefore involves proteins, possibly aquaporins. Transgenic plants with altered aquaporin expression provide systems in which to explore this issue further (47, 144). In tobacco plants with antisense inhibition or an antibiotic-inducible overexpression of NtAQP1, g m positively correlates with the NtAQP1 expression level 606 Maurel et al.

59 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org (33, 144). In view of the CO 2 transport activity of NtAQP1 in oocytes, this result was interpreted as evidence that this PIP1 homolog serves as a CO 2 pore in tobacco leaves (33, 144). Interestingly, in various genetic and physiological contexts g m is positively correlated with maximal stomatal conductance and CO 2 assimilation capacity (see 33 and references therein). Up to now, it was unclear whether changes in g m can be accounted for by changes in leaf anatomy or by changes in cell permeability to CO 2 (47). Following evidence that PIP contributes to both g m and leaf hydraulic conductance, the hypothesis that aquaporins coregulate CO 2 and H 2 O transport in the mesophyll emerged (Figure 4b). A role for aquaporins in carbohydrate storage and compartmentation has also been suggested. In a first study, tomato fruits with antisense inhibition of a PIP homolog showed increases in organic acid content and decreases in sugar content; these defects are associated with a marked alteration of the ripening process (19). In another study, metabolomic analysis of Arabidopsis plants lacking AtTIP1;1 expression revealed complex alterations in the accumulation of various sugars, organic acids, and starch (84). Although this provocative hypothesis remains to be confirmed, reduced expression of AtTIP1;1 was proposed to alter vesicle trafficking and therefore carbohydrate compartmentation. Nutrient uptake. Two recent studies unraveled novel functions of aquaporins in plant nutrient uptake. Transcriptome analysis of Arabidopsis roots revealed striking upregulation of AtNIP5;1 in response to boron (B) deficiency (134). Interestingly, AtNIP5;1 transports boric acid in Xenopus oocytes and significantly contributes to root B uptake, as shown in two independent nip5;1 insertion lines. The physiological significance of AtNIP5;1 was further underscored by mutant plants that, under B limitation, display a striking growth retardation of shoot and roots and an inhibition of flower and silique formation (134). We note that a role for membrane channels in B transport was first proposed in an early study on squash roots (29). However, the role of aquaporins had remained uncertain owing to the low B transport activity of the PIPs and NIPs investigated in that study. Silicon is a major mineral component of certain plants, such as Gramineae, and more generally, helps plants withstand abiotic stresses and pathogen attacks. Molecular characterization of low silicon rice1 (lsi1), a mutant of rice defective in silicon uptake into roots, led to the unexpected finding that Lsi1 encodes a NIP homolog (82). Lsi1 transports silicon after heterologous expression in oocytes. Lsi1 is expressed on the distal side of exodermal and endodermal root cells and may contribute, in combination with the efflux transporter Lsi2 (83), to a vectorial transport of silicic acid from the soil solution into the xylem, a limiting step for translocation of silicon to the aerial parts (Figure 4a). A similar function can be expected in other important crops such as maize, which also accumulates significant amounts of silicon and has close Lsi1 homologs (82). AQUAPORINS IN A VARIABLE ENVIRONMENT Changes in Irradiance Light is a key environmental parameter that, besides its long-term effects on plant growth and development, diurnally affects the plant metabolic regime and therefore affects water relations. A primary effect of light is to control stomatal aperture, and therefore transpiration. In guard cells of sunflower leaves, the transcript abundance of a TIP homolog (Sun- TIP7) is under diurnal regulation and is maximal at the end of the day, during stomatal closure, suggesting that this aquaporin contributes to water efflux from the guard cell (123). Light also enhances K leaf in many plant species (22, 103, 142). For instance in sunflower, the 50% increase in K leaf induced by light is fully sensitive to mercury inhibition, Plant Aquaporins 607

60 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org suggesting that the variations in K leaf are due to changes in aquaporin activity (103). This was confirmed in walnut twigs, where K leaf shows a very tight kinetic correlation with the abundance of two major PIP2 transcripts during a transition from dark to high light (22). Enhanced activity of leaf aquaporins during the day, and therefore increased K leaf, may favor water transport into the inner leaf tissues when transpiration is maximal. This process would avoid excessive drops in leaf water potential, reduce xylem tensions, and therefore prevent possible xylem embolization. Light interception is optimized by diurnal movements of leaves, a process in which aquaporins also participate. In the Mimosaceae, the movement of leaves and leaflets is determined by coordinate swelling and shrinking of cells on opposing sides of a motor organ, the pulvinus. In Samanea saman, the osmotic water permeability of protoplasts isolated from the pulvinus shows diurnal regulation and is maximal in the mornings and evenings, concomitant with leaf movement (100). Accumulation of a PIP2 homolog in these cells is under circadian control and in phase with these rhythmic changes in water transport. In Mimosa pudica, motor cells harbor both a tannin and a central aqueous vacuole. Immunocytochemistry experiments indicate that the latter type of vacuoles shows an approximate tenfold higher density of TIP1 aquaporins as compared with the former type, in agreement with the major contribution of the central vacuole to water exchanges during cell volume regulation (32). Diurnal variations of root Lp r, with a twoto threefold increase during the day, have been observed in many plant species and may contribute, together with light-dependent regulation of K leaf, to reducing xylem tension under conditions of high transpiration demand (16, 49, 78). In Lotus japonicus and maize, for instance, root water transport is maximally enhanced around midday and is matched or slightly preceded by an increase in the abundance of PIP1 and PIP2 transcripts (49, 78). In maize, ZmPIP1;5 transcripts are detected in all root cell types during the day but restricted to the epidermis during the night (40). However, the abundance of PIP proteins in maize roots shows a more complex diurnal variation profile than that of transcripts, suggesting a role for posttranslational regulation (78). The mechanisms that allow light perception and long-distance control of aquaporins in the plant root deserve more precise investigation. Water, Salt, and Nutrient Stresses Regulation of turgor and intracellular water movements. Because it is central for plant water relations, the regulation of water transport during water deficit has been the object of extensive research. Understanding the role of aquaporins in this context now requires integration of numerous observations made at the molecular, cell, and tissue levels. Fundamental regulation properties that explain the remarkable ability of plant cells to withstand water deprivation have emerged from basic knowledge of plant cell water relations and from more recent research on aquaporins. Water deficit induces primarily an efflux of water, which can result in a marked drop in cell turgor and ultimately, but more rarely, in cell plasmolysis or cytorrhysis. In this context, the cytosol, which contributes to a minor fraction of the plant cell volume, may be very sensitive to differential flow of water across the plasma membrane and the tonoplast. Abrupt changes in cytosolic volume can theoretically be avoided if mobilization of water from or into the vacuole is nonlimiting (92, 140). Studies with membranes purified from wheat and tobacco have confirmed that, in these preparations at least, the tonoplast shows much higher water permeability and aquaporin activity than the plasma membrane (95, 104). Osmotic stress also requires an adjustment of plasma membrane surface area; recent results link this process to the regulation of membrane water permeability. In Vicia faba guard cells, a pretreatment with inhibitors 608 Maurel et al.

61 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org of membrane trafficking (wortmannin, cytochalasin D) slows down cell shrinkage in response to hypertonicity, suggesting that a reduction of cell hydraulic conductivity and possibly aquaporin downregulation is induced (126). When protoplasts isolated from maize suspension cells are hypotonically challenged, a subset of these protoplasts exhibits a retarded swelling, which was interpreted to mean that their initial water permeability is extremely low and is dynamically adjusted during the course of cell swelling and mobilization of membrane material at the cell surface (101). In agreement with these functional data, dynamic changes in aquaporin subcellular localization were observed in osmotically challenged cells. These processes may also reflect transfer of aquaporins to subcellular compartments devoted to protein degradation. For instance, mannitol-induced osmotic stress in ice plant suspension cells induces the relocalization of McTIP1;2 from the tonoplast to a putative endosomal compartment (146). This process is dependent on aquaporin glycosylation and a camp-dependent pathway. In salt-treated Arabidopsis roots, AtTIP1;1, but not the AtTIP2;1 homolog, is relocalized in vacuolar bulbs (15). In addition, redistribution of PIPs from the plasma membrane to internal compartments contributes to the downregulation of root water uptake (15). The regulation of water transport in Chara cells in response to changing osmotic or hydrostatic pressures has been interpreted as the result of a direct gating of aquaporins by these factors (152, 155). In higher plant cells, aquaporins are more likely under the control of osmo- and pressure-sensing molecules and downstream signaling cascades. For instance, the downregulation of water permeability in melon protoplasts by salt can be counteracted by okadaic acid, a protein phosphatase inhibitor (89). More specifically, phosphorylation of SoPIP2;1 in spinach leaf fragments decreases in response to a hyperosmotic treatment (64). A general model of cell osmoregulation involving stretch-activated Ca 2+ channel and Ca 2+ -dependent phosphorylation of SoPIP2;1 was proposed to explain that aquaporin phosphorylation, and therefore cell hydraulic conductivity, would be maximal at high water potential to favor water entry in fully turgid cells (63). Finally, several authors proposed that aquaporins themselves may function as osmosensors, but the molecular and cellular mechanisms involved remain elusive (51, 86). Whole plant level. At the whole plant level, a major effect of drought is to reduce transpiration through stomatal closure. Yet, in extreme drought conditions, high tension in the xylem can lead to vessel occlusion by embolization. A specific role for aquaporins in embolism refilling and recovery of stem axial conductance after drought was proposed in grapevine on the basis of mercury inhibition experiments (80). In roots of most plant species investigated, drought or salt stresses also result in a marked decrease in Lp r (61). In Arabidopsis for instance, exposure to 100 mm NaCl reduces Lp r by 70% with a half-time of approximately 45 min. The fact that residual Lp r of salt-stressed Arabidopsis or paprika roots becomes resistant to mercury was interpreted to mean that aquaporin activity is downregulated in these conditions (17, 88). During the day, this early response may provide a hydraulic signal to the leaf to trigger stomatal closure, whereas during the night, it may avoid a backflow of water to the drying soil. Numerous early studies reported on water stress dependent expression of aquaporin genes and a large variety of individual regulation profiles were described (92). A more comprehensive understanding of the processes involved has emerged from recent studies in maize, rice, radish, and Arabidopsis, in which expression of the whole aquaporin family was considered (3, 15, 44, 58, 73, 85, 131, 159). In salt-stressed roots of Arabidopsis and maize, a coordinated downregulation of most aquaporin transcripts occurs, which over the first 24 h of stress can contribute to Lp r downregulation (3, 15, 85). A recovery toward initial Plant Aquaporins 609

62 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org transcript abundance occurs over longer term stresses (85, 159). Transcriptional control of aquaporins in drought-stressed leaves appears to be more complex and, although a tendency to overall aquaporin gene downregulation is also observed, specific upregulation of certain PIP transcripts occurs in rice and Arabidopsis leaves (3, 44). Interestingly, the two transcripts that are upregulated in Arabidopsis are specifically expressed in aerial parts (3). Although their tissue expression pattern is as yet unknown, these isoforms may facilitate water flow toward critical cell types. Studies in Hordeum vulgare (barley) leaves suggest that increased abundance of HvPIP1;6 transcripts in response to salt may reflect a role for this aquaporin in promoting residual growth of the leaf under stress (37). In agreement with its central role in plant responses to water stress, abscisic acid (ABA) seems to mediate, at least in part, droughtand salt stress-induced aquaporin regulation. For instance, treatment of maize roots with ABA results over 1 2 h in a transient increase in hydraulic conductivity of the whole organ and of cortical cells, by factors of 3 4 and 7 27, respectively (54). Consistent with these effects, ABA also rapidly enhances the expression of some PIP isoforms (159). In rice roots, a strong induction of several PIPs is observed in response to water deficit, specifically in an upland cultivar that shows an enhanced production of ABA (73). This response may optimize uptake of residual soil water at the onset of soil drying. Genetic approaches have also been useful to investigate the function of aquaporins during drought. Transgenic tobacco plants with antisense inhibition of PIP1 and transgenic Arabidopsis plants with antisense inhibition of PIP1 expression and PIP2 expression showed lower leaf water potentials than wild-type plants under drought stress conditions (90, 128). Most strikingly, the recovery following rewatering of leaf wilting in tobacco and leaf water potential and plant hydraulic conductance in Arabidopsis is significantly delayed in the antisense plants (90, 128). Therefore, PIP aquaporins contribute to adaptation of the plants to drought by mechanisms that remain to be determined, and even more significantly, contribute to rehydration of the whole plant body after drought. Another genetic strategy is to enhance aquaporin expression in transgenic plants. Although spectacular phenotypes are observed in most studies, aquaporin overexpression has either beneficial (44, 72, 156) or adverse (2, 68) effects on drought tolerance, depending on the aquaporin gene or the plant species investigated. Therefore, the relevance of this approach for biotechnological improvement of plant tolerance to water stress remains uncertain. One reason may be that many studies relied on overexpression of an aquaporin in a heterologous plant species (2, 44, 68, 156). Inadequate regulation of the foreign aquaporin may disturb the endogenous stress response. In these respects, more relevant insights were provided by a study showing that Oryza sativa OsPIP1;3 is specifically induced by water stress in an upland, drought-avoidant cultivar of rice (72). Furthermore, the performance of a lowland cultivar under drought can be significantly enhanced by expression of this aquaporin under a stress-responsive promoter (72). In future studies it will be important to evaluate the capacity to recover from water stress after rewatering in transgenic plants that ectopically express an aquaporin. Responses to nutrient stress. Strong interactions exist between the nutrient and water status of plants; integration of these two aspects seems to be critical for a deeper understanding of plant stress responses. For instance, deprivation of N, phosphorus (P), or sulfur (S) in plants results after a few days in a significant inhibition of water transport in whole roots or individual root cells; initial root water transport properties can be restored in the 24 h following nutrient resupply (16, 21, 125). A downregulation of water channels under N and P deprivation is invoked on the basis of the insensitivity of residual Lp r to mercury (16). The adverse effects of nutrient starvation 610 Maurel et al.

63 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org on plant water relations have also been studied in sorghum under drought stress. As compared with replete conditions, P starvation enhances the inhibition of Lp r by a polyethylene glycol treatment and slows down its recovery after water resupply (125). The molecular and cellular mechanisms involved in these regulations remain unclear. Stimulation of maize Lp r by NO 3 is blocked in the presence of tungstate, an inhibitor of nitrate reductase, suggesting that products of the N assimilation pathway are required for activation of aquaporin functions (10). A general transcriptional control of aquaporins by nutrient stress is also observed in the Arabidopsis root and, for instance, calcium deprivation results in an overall transcriptional downregulation of aquaporins (85). The effects of potassium (K) starvation are more moderate but a downregulation of Arabidopsis PIPs is also induced in the long term. By contrast, K deprivation in rice induces a twofold stimulation of Lp r after 4 6 h and enhances expression of some PIP isoforms, in parallel to expression of K channels (75). Coregulation of aquaporins and K transport systems has also been observed in roots treated with CsCl, which in addition to blocking K transport, reduces Lp r and aquaporin expression (75). These data suggest tight interactions between water and K transport during cell turgor regulation. Cold Stress Chilling of plant roots (i.e., exposure to 4 C 8 C) reduces root pressure, sap flow, and Lp r in a few hours (5, 71, 157). These effects in turn induce water deficit symptoms in shoots, such as decreased leaf water potential and stomatal closure (157). In maize, both a chilling-tolerant and a chilling-sensitive variety show an initial, >80% decline in Lp r but the tolerant variety shows a unique capacity to spontaneously overcompensate Lp r upon prolonged (>3 d) chilling (5). The speed and reversibility of inhibition of Lp r by chilling in cucumber and rice and a concomitant six- to ninefold reduction in cortex cell hydraulic conductivity in cucumber roots suggest that inhibition of aquaporin activity is involved (71, 157). Comprehensive gene expression analyses in roots and shoots of rice, maize, and Arabidopsis show that chilling induces a marked (two- to fourfold) decrease in abundance for most PIP transcripts (5, 58, 120, 157). Normal gene expression is restored in the 24 h following the return to permissive temperature. However, in maize and rice roots under cold stress or recovery, the abundance of aquaporin transcripts and proteins is not always correlated, suggesting the occurrence of posttranscriptional regulation (5, 157). In addition, the abundance of PIP1s and of phosphorylated PIP2 forms (as monitored by immunodetection) increases during prolonged chilling in roots of both a chilling-tolerant and a chillingsensitive variety of maize; intriguingly, this response is not correlated to their differential Lp r regulation (5). Because the sensitivities to chilling and H 2 O 2 are correlated in the tolerant and sensitive varieties, it was proposed that ROS-induced damage probably dominates the aquaporin response and determines the poor performance of the sensitive variety during stress (5). A more direct relationship between aquaporins and ROS was determined in cucumber (71). In this species, H 2 O 2 accumulates in response to chilling and treatment of roots by exogenous H 2 O 2 inhibits Lp r to the same extent as chilling. Flowers can also perceive temperature. In tulip, diurnal movements of petals are controlled in part by changes in temperature (7). Petal opening and an associated water retention can be induced at 20 C and are linked to phosphorylation in a microsomal membrane fraction of a 31-kDa protein, tentatively identified as a PIP isoform. Petal closure is induced at 5 C and is associated with decreased phosphorylation of the putative aquaporin. The effects of a calcium chelator [1,2-bis(o-aminophenoxy)ethane- N,N,N,N -tetraacetic acid, or BAPTA] and a calcium channel blocker on petal Plant Aquaporins 611

64 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org AM: arbuscular mycorrhiza movement and the associated phosphorylation of the aquaporin suggest a role for calcium signaling in this process (7). Finally, freeze-thaw cycles occurring during winter can, similar to severe drought stress, embolize xylem vessels of woody plants (119). Embolism repair may be achieved by hydrolysis of starch in adjacent parenchyma cells, exudation of the resulting sugars in the vessel, and concomitant water influx to chase out the air bubble. In walnut twigs, the transcripts and proteins corresponding to two PIP2 homologs show seasonal variations and preferentially accumulate in the xylem parenchyma during winter, suggesting a role for these aquaporins in embolism repair (119). Anoxia Flooding or compaction of soils results in acute oxygen deprivation (anoxia) of plant roots, which is a major stress for cultivated plants. Most plant species investigated show a rapid inhibition of Lp r in response to anoxia. Tournaire-Roux and coworkers (138) delineated the organ and cell bases of this process in Arabidopsis roots. They showed that anoxic stress results in acidosis of root cells and that the Lp r of excised roots diminishes in parallel to cytosolic ph. These observations can be linked to the molecular mechanism of PIP aquaporin gating by cytosolic protons, which is conserved in PIPs of all plants and therefore can explain how an anoxic stress results in a massive inhibition of root water uptake. This regulation may avoid excessive dilution of xylem sap after flooding or, on a longer term, favor ethylene accumulation, which in turn induces aerenchyma differentiation (52). Ethylene, which enhances Lp r in hypoxic aspen roots, may also compensate for the initial inhibition of water transport in response to oxygen deprivation (65). Consistent with the physiological inhibition of Lp r, a general downregulation of PIP and TIP genes in response to hypoxia occurs in Arabidopsis (74). By contrast, expression of AtNIP2;1 is markedly induced upon flooding stress and hypoxia (20, 74). AtNIP2;1 transports lactic acid and may therefore provide a path for release of this fermentation product from root cells, to contribute to cytosolic ph regulation and metabolic adaptation to long-term hypoxia (20). Biotic Interactions Rhizobia-legume symbiosis. Interactions of plants with soil microorganisms, which have long been known as central for plant mineral nutrition and metabolism, more recently appear to play an important role in plant water relations and tolerance to environmental stresses (6, 43, 143). Notably, GmNOD26, the first plant aquaporin to be identified, is specifically expressed in symbiotic nitrogen-fixing nodules formed after infection of soybean by Rhizobiaceae bacteria (149). Similar nodulins have been identified in other legumes. GmNOD26 is a major component of the peribacteroid membrane, a membrane of plant origin that surrounds the bacteroid and mediates exchanges with the root cell. Antibodies raised against either the native or phosphorylated form of GmNOD26 reveal that maximal expression of the protein and its subsequent phosphorylation coincide with bacteroid maturation (43). Because of its solute transporting activity, GmNOD26 has been tentatively linked to a channel-mediated import of NH 3 from the bacteroid, but unequivocal evidence for such function is still lacking (105, 149). The water transport activity of GmNOD26 may also help the plant cell to couple osmoregulation of the plant cytosol and peribacteroid space. Accordingly, drought and salt stress result in a threefold increase in GmNOD26 phosphorylation, suggesting that enhanced water permeability is required for nodule osmoregulation and adaptation to water stress. (43). Mycorrhiza. Arbuscular mycorrhizas (AM) represent the most common form of symbiosis between land plants and soil fungi. Similar to the Rhizobia-legume symbiosis, this 612 Maurel et al.

65 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org interaction results in deep anatomical changes of root cells, involving in this case the differentiation of convoluted, periarbuscular membrane structures that are the site of extensive exchanges of mineral nutrient (phosphate), carbohydrates (photosynthates), and water with the fungus. This membrane specialization results in marked changes in PIP and TIP gene expression, with a specific profile depending on the plant host or the symbiotic fungus (reviewed in 143). In poplar, mycorrhized plants show, with respect to nonmycorrhized plants, a 55% increase in Lp r ; besides changes in root anatomy (internal surface), this increase can be accounted for by enhanced expression of most of the PIPs expressed in roots (87). By contrast, the Lp r of mycorrhized Phaseolus vulgaris plants is reduced approximately threefold and this reduction is associated with a decreased abundance of PIP2s and their phosphorylated forms (6). Plant aquaporins expressed during AM symbiosis may also contribute to NH 3 import from the fungus (143). AM symbiosis also exerts beneficial effects on tolerance of plants to water stress, whether induced by drought, salinity, or chilling (6, 143). These effects may be mediated through alteration of both root water uptake and transpiration to promote water economy. A specific role for aquaporins was deduced in a study on transgenic tobacco, showing that antisense inhibition of NtAQP1 reduces the positive effects of mycorrhiza on root and leaf growth under drought (112). Two recent studies in lettuce, soybean, and tomato drew an interesting parallel between the effects of AM and drought stress, which synergistically regulate PIP genes in roots and leaves (reviewed in 143). Nematode and other infections. Specific aquaporins also seem to be involved in plantpathogen interactions as an adaptive response to infection-induced changes in plant cell morphology. For instance, infection of roots by nematodes leads to the differentiation of giant cells that serve as feeding sites for the parasite. Regulatory sequences that are specifically responsive to nematode infection were identified in the promoter sequence of a tobacco TIP gene (107). Enhanced expression of this aquaporin might be necessary to achieve extensive delivery of water and solutes to the parasite, together with proper osmotic regulation of the giant cells. In tomato, incompatible interaction with the parasite Cuscuta reflexa induces the expression of a PIP1 homolog, probably in relation to the auxin-dependent elongation of hypodermal cells induced after pathogen attachment (153). CONCLUSIONS Aquaporins have provided a unique molecular entry into plant water relations and their study has significantly improved our understanding of integrated mechanisms of water transport, in roots in particular. Yet, in view of the complex expression and regulation profiles of aquaporins, their role in regulating water transport in many other physiological and developmental contexts, including seed germination, stomatal regulation, and leaf water transport, deserves further investigation (Figure 5). In most studies, aquaporin function is experimentally monitored through water flow intensities and kinetics. However, in the whole plant the overall flow of water across plant tissues is determined by stomatal regulation and/or solute transport. Therefore, it will be important to consider how, in this context, aquaporins may critically determine local water potential gradients rather than water flow intensities. These new considerations may lead to a better understanding of the role of aquaporins during cell elongation and water stress. Studies on aquaporins have also led us far beyond membrane water transport. The transport of solutes of great physiological significance, such as CO 2,H 2 O 2, B, or silicic acid, is now well established and has linked aquaporins to many functions, including carbon metabolism, oxidative stress responses, and plant mineral nutrition (Figure 5). Yet, Plant Aquaporins 613

66 Anther dehiscence Cell expansion Light GA3 CO 2 transport Stomatal movement Extra-vascular transport Leaf movement Circadian & diurnal controls Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Micro-nutrient uptake Embolism repair Seed germination Water uptake Biotic interaction Drought Freezing Nitric oxide Circadian & diurnal controls Drought Salt Nutrient availability Cold Anoxia Pathogen-interactions... Drought Salt Nutrient availability Figure 5 Integrated functions of plant aquaporins. The endogenous and environmental factors acting on each of the indicated aquaporin functions are shown. GA3, gibberellin. novel putative substrates of plant aquaporins such as arsenate, NO, and NO 3 await further investigation. We also note that several plant aquaporins, in the NIP and SIP subgroups in particular, have unknown functions and that new aquaporin subclasses are being discovered (139). Finally, aquaporin functions need to be further integrated in the whole plant physiology. This will first require a better understanding of how the various transport activities of aquaporins are coupled with those of other transport proteins (75, 83, 134). The chains of events that lead to control of aquaporin functions by local or long-distance signals, during development or in response to biotic or abiotic signals, will also have to be elucidated. Finally, and although the field of aquaporin research has already enlarged considerably, we may not be at the end of our surprises because novel primary functions as diverse as cell proliferation (117) or virus replication (69) might be anticipated for plant aquaporins. SUMMARY POINTS 1. Aquaporins are membrane channels that have a conserved structure and facilitate the transport of water and/or small neutral solutes (urea, boric acid, silicic acid) or gases (ammonia, carbon dioxide). 614 Maurel et al.

67 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 2. Aquaporins exhibit a high isoform multiplicity that reflects distinct transport specificities and subcellular localizations. 3. Aquaporin transport activity can be regulated by multiple mechanisms, including regulation of transcript or protein abundance, subcellular trafficking, or gating by phosphorylation or cytosolic protons. 4. Aquaporins play a central role in plant water relations. They mediate the regulation of root water transport in response to a variety of environmental stimuli. They facilitate water transport through inner leaf tissues during transpiration and in expanding tissues. 5. Multiple integrated roles of aquaporins in carbon and nitrogen assimilation and micronutrient uptake are being uncovered. FUTURE ISSUES 1. The transport specificity of aquaporins lacking function and, in particular, of novel classes of aquaporins recently discovered in certain plant species should be investigated. 2. The mechanisms governing aquaporin subcellular trafficking should be investigated, and in particular it will be important to evaluate in planta how the functional expression of aquaporins of the plasma membrane intrinsic proteins 1 and 2 (PIP1 and PIP2) subclasses is determined by mutual physical interactions. 3. Investigation of the function and regulation of aquaporins in poorly explored physiological contexts, such as stomatal regulation or seed germination, will be required. 4. The mechanisms that determine regulation of aquaporins by light in inner leaf tissues should be dissected, and the role of aquaporins in coregulating CO 2 and H 2 O transport in these tissues should be deciphered. 5. The relevance of altered aquaporin expression for biotechnological improvement of plant tolerance to water stress must be explored. The role of aquaporins during water stress recovery and in conditions similar to those experienced by crops in the field will have to be specified. DISCLOSURE STATEMENT The authors are not aware of any biases that might be perceived as affecting the objectivity of this review. ACKNOWLEDGMENTS Work in our laboratory is supported by grants from INRA (AgroBI AIP300) and Génoplante (ANR-05-GPLA ). We thank members of our group for fruitful discussions and apologize to all colleagues whose valuable work could not be cited owing to space limitations. Plant Aquaporins 615

68 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org LITERATURE CITED 1. Aasamaa K, Sober A Seasonal courses of maximum hydraulic conductance in shoots of six temperate deciduous tree species. Funct. Plant Biol. 32: Aharon R, Shahak Y, Wininger S, Bendov R, Kapulnik Y, Galili G Overexpression of a plasma membrane aquaporin in transgenic tobacco improves plant vigor under favorable growth conditions but not under drought or salt stress. Plant Cell 15: Alexandersson E, Fraysse L, Sjövall-Larsen S, Gustavsson S, Fellert M, et al Whole gene family expression and drought stress regulation of aquaporins. Plant Mol. Biol. 59: Alleva K, Niemietz CM, Sutka M, Maurel C, Parisi M, et al Plasma membrane of Beta vulgaris storage root shows high water channel activity regulated by cytoplasmic ph and a dual range of calcium concentrations. J. Exp. Bot. 57: Aroca R, Amodeo G, Fernández-Illescas S, Herman EM, Chaumont F, Chrispeels MJ The role of aquaporins and membrane damage in chilling and hydrogen peroxide induced changes in the hydraulic conductance of maize roots. Plant Physiol. 137: Aroca R, Porcel R, Ruiz-Lozano JM How does arbuscular mycorrhizal symbiosis regulate root hydraulic properties and plasma membrane aquaporins in Phaseolus vulgaris under drought, cold or salinity stresses? New Phytol. 173: Azad AK, Sawa Y, Ishikawa T, Shibata H Phosphorylation of plasma membrane aquaporin regulates temperature-dependent opening of tulip petals. Plant Cell Physiol. 45: Bansal A, Sankararamakrishnan R Homology modeling of major intrinsic proteins in rice, maize and Arabidopsis: comparative analysis of transmembrane helix association and aromatic/arginine selectivity filters. BMC Struct. Biol. 7:27 9. Barrieu F, Thomas D, Marty-Mazars D, Charbonnier M, Marty F Tonoplast intrinsic proteins from cauliflower (Brassica oleracea L. var. botrytis): immunological analysis, cdna cloning and evidence for expression in meristematic tissues. Planta 204: Barthes L, Deléens E, Bousser A, Hoarau J, Prioul J-L Xylem exudation is related to nitrate assimilation pathway in detopped maize seedlings: use of nitrate reductase and glutamine synthetase inhibitors as tools. J. Exp. Bot. 47: Beitz E, Wu B, Holm LM, Schultz JE, Zeuthen T Point mutations in the aromatic/arginine region in aquaporin 1 allow passage of urea, glycerol, ammonia, and protons. Proc. Natl. Acad. Sci. USA 103: Biela A, Grote K, Otto B, Hoth S, Hedrich R, Kaldenhoff R The Nicotiana tabacum plasma membrane aquaporin NtAQP1 is mercury-insensitive and permeable for glycerol. Plant J. 18: Bienert GP, Møller AL, Kristiansen KA, Schulz A, Møller IM, et al Specific aquaporins facilitate the diffusion of hydrogen peroxide across membranes. J. Biol. Chem. 282: Bots M, Vergeldt F, Wolters-Arts M, Weterings K, van As H, Mariani C Aquaporins of the PIP2 class are required for efficient anther dehiscence in tobacco. Plant Physiol. 137: Boursiac Y, Chen S, Luu D-T, Sorieul M, van den Dries N, Maurel C Early effects of salinity on water transport in Arabidopsis roots. Molecular and cellular features of aquaporin expression. Plant Physiol. 139: Carvajal M, Cooke DT, Clarkson DT Responses of wheat plants to nutrient deprivation may involve the regulation of water-channel function. Planta 199: Maurel et al.

69 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 17. Carvajal M, Martinez V, Alcaraz CF Physiological function of water channels as affected by salinity in roots of paprika pepper. Physiol. Plant. 105: Chaumont F, Moshelion M, Daniels MJ Regulation of plant aquaporin activity. Biol. Cell 97: Chen GP, Wilson ID, Kim SH, Grierson D Inhibiting expression of a tomato ripening-associated membrane protein increases organic acids and reduces sugar levels of fruit. Planta 212: Choi WG, Roberts DM Arabidopsis NIP2;1, a major intrinsic protein transporter of lactic acid induced by anoxic stress. J. Biol. Chem. 282: Clarkson DT, Carvajal M, Henzler T, Waterhouse RN, Smyth AJ, et al Root hydraulic conductance: diurnal aquaporin expression and the effects of nutrient stress. J. Exp. Bot. 51: Cochard H, Venisse JS, Barigah TS, Brunel N, Herbette S, et al New insights into the understanding of variable hydraulic conductances in leaves. Evidence for a possible implication of plasma membrane aquaporins. Plant Physiol. 143: Daniels MJ, Chaumont F, Mirkov TE, Chrispeels MJ Characterization of a new vacuolar membrane aquaporin sensitive to mercury at a unique site. Plant Cell 8: Daniels MJ, Chrispeels MJ, Yeager M Projection structure of a plant vacuole membrane aquaporin by electron cryo-crystallography. J. Mol. Biol. 294: Daniels MJ, Mirkov TE, Chrispeels MJ The plasma membrane of Arabidopsis thaliana contains a mercury-insensitive aquaporin that is a homolog of the tonoplast water channel protein TIP. Plant Physiol. 106: Daniels MJ, Yeager M Phosphorylation of aquaporin PvTIP3;1 defined by mass spectrometry and molecular modeling. Biochemistry 44: Dean RM, Rivers RL, Zeidel ML, Roberts DM Purification and functional reconstitution of soybean Nodulin 26. An aquaporin with water and glycerol transport properties. Biochemistry 38: Dhonukshe P, Aniento F, Hwang I, Robinson DG, Mravec J, et al Clathrinmediated constitutive endocytosis of PIN auxin efflux carriers in Arabidopsis. Curr. Biol. 17: Dordas C, Chrispeels MJ, Brown PH Permeability and channel-mediated transport of boric acid across membrane vesicles isolated from squash roots. Plant Physiol. 124: Eisenbarth DA, Weig AR Dynamics of aquaporins and water relations during hypocotyl elongation in Ricinus communis L. seedlings. J. Exp. Bot. 56: Fetter K, Van Wilder V, Moshelion M, Chaumont F Interactions between plasma membrane aquaporins modulate their water channel activity. Plant Cell 16: Fleurat-Lessard P, Frangne N, Maeshima M, Ratajczak R, Bonnemain JL, Martinoia E Increased expression of vacuolar aquaporin and H + -ATPase related to motor cell function in Mimosa pudica L. Plant Physiol. 114: Flexas J, Ribas-Carbó M, Hanson DT, Bota J, Otto B, et al Tobacco aquaporin NtAQP1 is involved in mesophyll conductance to CO 2 in vivo. Plant J. 48: Fotiadis D, Jenö P, Mini T, Wirtz S, Müller SA, et al Structural characterization of two aquaporins isolated from native spinach leaf plasma membranes. J. Biol. Chem. 276: Frangne N, Maeshima M, Schaffner AR, Mandel T, Martinoia E, Bonnemain JL Expression and distribution of a vacuolar aquaporin in young and mature leaf tissues of Brassica napus in relation to water fluxes. Planta 212: Plant Aquaporins 617

70 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 36. Fraysse L, Wells B, McCann MC, Kjellbom P Specific plasma membrane aquaporins of the PIP1 subfamily are expressed in sieve elements and guard cells. Biol. Cell. 97: Fricke W, Akhiyarova G, Wei W, Alexandersson E, Miller A, et al The short-term growth response to salt of the developing barley leaf. J. Exp. Bot. 57: Fujiyoshi Y, Mitsuoka K, de Groot BL, Philippsen A, Grubmüller H, et al Structure and function of water channels. Curr. Opin. Struct. Biol. 12: Gao YP, Young L, Bonham-Smith P, Gusta LV Characterization and expression of plasma and tonoplast membrane aquaporins in primed seed of Brassica napus during germination under stress conditions. Plant Mol. Biol. 40: Gaspar M, Bousser A, Sissoëff I, Roche O, Hoarau J, Mahe A Cloning and characterization of ZmPIP1 5b, an aquaporin transporting water and urea. Plant Sci. 165: Gerbeau P, Amodeo G, Henzler T, Santoni V, Ripoche P, Maurel C The water permeability of Arabidopsis plasma membrane is regulated by divalent cations and ph. Plant J. 30: Gerbeau P, Güclü J, Ripoche P, Maurel C Aquaporin Nt-TIPa can account for the high permeability of tobacco cell vacuolar membrane to small neutral solutes. Plant J. 18: Guenther JF, Chanmanivone N, Galetovic MP, Wallace IS, Cobb JA, Roberts DM Phosphorylation of soybean nodulin 26 on serine 262 enhances water permeability and is regulated developmentally and by osmotic signals. Plant Cell 15: Guo L, Wang ZY, Lin H, Cui WE, Chen J, et al Expression and functional analysis of the rice plasma-membrane intrinsic protein gene family. Cell Res. 16: Gustavsson S, Lebrun A, Norden K, Chaumont F, Johanson U A novel plant major intrinsic protein in Physcomitrella patens most similar to bacterial glycerol channels. Plant Physiol. 139: Hachez C, Moshelion M, Zelazny E, Cavez D, Chaumont F Localization and quantification of plasma membrane aquaporin expression in maize primary root: a clue to understanding their role as cellular plumbers. Plant Mol. Biol. 62: Hanba YT, Shibasaka M, Hayashi Y, Hayakawa T, Kasamo K, et al Overexpression of the barley aquaporin HvPIP2;1 increases internal CO 2 conductance and CO 2 assimilation in the leaves of transgenic rice plants. Plant Cell Physiol. 45: Harvengt P, Vlerick A, Fuks B, Wattiez R, Ruysschaert JM, Homble F Lentil seed aquaporins form a hetero-oligomer which is phosphorylated by a Mg 2+ -dependent and Ca 2+ -regulated kinase. Biochem. J. 352: Henzler T, Waterhouse RN, Smyth AJ, Carvajal M, Cooke DT, et al Diurnal variations in hydraulic conductivity and root pressure can be correlated with the expression of putative aquaporins in the roots of Lotus japonicus. Planta 210: Henzler T, Ye Q, Steudle E Oxidative gating of water channels (aquaporins) in Chara by hydroxyl radicals. Plant Cell Environ. 27: Hill AE, Shachar-Hill B, Shachar-Hill Y What are aquaporins for? J. Membr. Biol. 197: Holbrook NM, Zwieniecki MA Plant biology: water gate. Nature 425: Holm LM, Jahn TP, Møller AL, Schjoerring JK, Ferri D, et al NH 3 and NH 4 + permeability in aquaporin-expressing Xenopus oocytes. Pflügers Arch. 450: Hose E, Steudle E, Hartung W Abscisic acid and hydraulic conductivity of maize roots: a study using cell- and root-pressure probes. Planta 211: Maurel et al.

71 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 55. Hukin D, Doering-Saad C, Thomas CR, Pritchard J Sensitivity of cell hydraulic conductivity to mercury is coincident with symplasmic isolation and expression of plasmalemma aquaporin genes in growing maize roots. Planta 215: Ishikawa F, Suga S, Uemura T, Sato MH, Maeshima M Novel type aquaporin SIPs are mainly localized to the ER membrane and show cell-specific expression in Arabidopsis thaliana. FEBS Lett. 579: Jahn TP, Møller AL, Zeuthen T, Holm LM, Klaerke DA, et al Aquaporin homologues in plants and mammals transport ammonia. FEBS Lett. 574: Jang JY, Kim DG, Kim YO, Kim JS, Kang H An expression analysis of a gene family encoding plasma membrane aquaporins in response to abiotic stresses in Arabidopsis thaliana. Plant Mol. Biol. 54: Jauh GY, Fischer AM, Grimes HD, Ryan CA Jr, Rogers JC δ-tonoplast intrinsic protein defines unique plant vacuole functions. Proc. Natl. Acad. Sci. USA 95: Javot H, Lauvergeat V, Santoni V, Martin-Laurent F, Güclü J, et al Role of a single aquaporin isoform in root water uptake. Plant Cell 15: Javot H, Maurel C The role of aquaporins in root water uptake. Ann. Bot. 90: Johanson U, Karlsson M, Johansson I, Gustavsson S, Sjövall S, et al The complete set of genes encoding major intrinsic proteins in Arabidopsis provides a framework for a new nomenclature for major intrinsic proteins in plants. Plant Physiol. 126: Johansson I, Karlsson M, Shukla VK, Chrispeels MJ, Larsson C, Kjellbom P Water transport activity of the plasma membrane aquaporin PM28A is regulated by phosphorylation. Plant Cell 10: Johansson I, Larsson C, Ek B, Kjellbom P The major integral proteins of spinach leaf plasma membranes are putative aquaporins and are phosphorylated in response to Ca 2+ and apoplastic water potential. Plant Cell 8: Kamaluddin M, Zwiazek JJ Ethylene enhances water transport in hypoxic aspen. Plant Physiol. 128: Kammerloher W, Fischer U, Piechottka GP, Schäffner AR Water channels in the plant plasma membrane cloned by immunoselection from a mammalian expression system. Plant J. 6: Karlsson M, Fotiadis D, Sjövall S, Johansson I, Hedfalk K, et al Reconstitution of water channel function of an aquaporin overexpressed and purified from Pichia pastoris. FEBS Lett. 537: Katsuhara M, Koshio K, Shibasaka M, Hayashi Y, Hayakawa T, Kasamo K Overexpression of a barley aquaporin increased the shoot/root ratio and raised salt sensitivity in transgenic rice plants. Plant Cell Physiol. 44: Kim MJ, Kim HR, Paek KH Arabidopsis tonoplast proteins TIP1 and TIP2 interact with the cucumber mosaic virus 1a replication protein. J. Gen. Virol. 87: Kojima S, Bohner A, von Wirén N Molecular mechanisms of urea transport in plants. J. Membr. Biol. 212: Lee SH, Singh AP, Chung GC Rapid accumulation of hydrogen peroxide in cucumber roots due to exposure to low temperature appears to mediate decreases in water transport. J. Exp. Bot. 55: Lian HL, Yu X, Lane D, Sun WN, Tang ZC, Su W Upland rice and lowland rice exhibited different PIP expression under water deficit and ABA treatment. Cell Res. 16: Lian HL, Yu X, Ye Q, Ding XS, Kitagawa Y, et al The role of aquaporin RWC3 in drought avoidance in rice. Plant Cell Physiol. 45: Plant Aquaporins 619

72 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 74. Liu F, Vantoai T, Moy LP, Bock G, Linford LD, Quackenbush J Global transcription profiling reveals comprehensive insights into hypoxic response in Arabidopsis. Plant Physiol. 137: Liu HY, Sun WN, Su WA, Tang ZC Co-regulation of water channels and potassium channels in rice. Physiol. Plant. 128: Liu HY, Yu X, Cui DY, Sun MH, Sun WN, et al The role of water channel proteins and nitric oxide signaling in rice seed germination. Cell Res. 17: Liu LH, Ludewig U, Gassert B, Frommer WB, von Wirén N Urea transport by nitrogen-regulated tonoplast intrinsic proteins in Arabidopsis. Plant Physiol. 133: Lopez F, Bousser A, Sissoëff I, Gaspar M, Lachaise B, et al Diurnal regulation of water transport and aquaporin gene expression in maize roots: contribution of PIP2 proteins. Plant Cell Physiol. 44: Loqué D, Ludewig U, Yuan L, von Wirén N Tonoplast intrinsic proteins AtTIP2;1 and AtTIP2;3 facilitate NH 3 transport into the vacuole. Plant Physiol. 137: Lovisolo C, Schubert A Mercury hinders recovery of shoot hydraulic conductivity during grapevine rehydration: evidence from a whole-plant approach. New Phytol. 172: Ludevid D, Höfte H, Himelblau E, Chrispeels MJ The expression pattern of the tonoplast intrinsic protein γ-tip in Arabidopsis thaliana is correlated with cell enlargement. Plant Physiol. 100: Ma JF, Tamai K, Yamaji N, Mitani N, Konishi S, et al A silicon transporter in rice. Nature 440: Ma JF, Yamaji N, Mitani N, Tamai K, Konishi S, et al An efflux transporter of silicon in rice. Nature 448: Ma S, Quist TM, Ulanov A, Joly R, Bohnert HJ Loss of TIP1;1 aquaporin in Arabidopsis leads to cell and plant death. Plant J. 40: Maathuis FJ, Filatov V, Herzyk P, Krijger GC, Axelsen KB, et al Transcriptome analysis of root transporters reveals participation of multiple gene families in the response to cation stress. Plant J. 35: MacRobbie EA Osmotic effects on vacuolar ion release in guard cells. Proc. Natl. Acad. Sci. USA 103: Marjanovic Z, Uehlein N, Kaldenhoff R, Zwiazek JJ, Weiss M, et al Aquaporins in poplar: what a difference a symbiont makes! Planta 222: Martínez-Ballesta MC, Aparicio F, Pallás V, Martínez V, Carvajal M Influence of saline stress on root hydraulic conductance and PIP expression in Arabidopsis. J. Plant Physiol. 160: Martínez-Ballesta MC, Martínez V, Carvajal M Regulation of water channel activity in whole roots and in protoplasts from roots of melon plants grown under saline conditions. Aust. J. Plant Physiol. 27: Martre P, Morillon R, Barrieu F, North GB, Nobel PS, Chrispeels MJ Plasma membrane aquaporins play a significant role during recovery from water deficit. Plant Physiol. 130: Maurel C Plant aquaporins: novel functions and regulation properties. FEBS Lett. 581: Maurel C, Javot H, Lauvergeat V, Gerbeau P, Tournaire C, et al Molecular physiology of aquaporins in plants. Int. Rev. Cytol. 215: Maurel C, Kado RT, Guern J, Chrispeels MJ Phosphorylation regulates the water channel activity of the seed-specific aquaporin α-tip. EMBO J. 14: Maurel et al.

73 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 94. Maurel C, Reizer J, Schroeder JI, Chrispeels MJ The vacuolar membrane protein γ-tip creates water specific channels in Xenopus oocytes. EMBO J. 12: Maurel C, Tacnet F, Güclü J, Guern J, Ripoche P Purified vesicles of tobacco cell vacuolar and plasma membranes exhibit dramatically different water permeability and water channel activity. Proc. Natl. Acad. Sci. USA 94: Miao GH, Hong Z, Verma DP Topology and phosphorylation of soybean nodulin- 26, an intrinsic protein of the peribacteroid membrane. J. Cell Biol. 118: Mizutani M, Watanabe S, Nakagawa T, Maeshima M Aquaporin NIP2;1 is mainly localized to the ER membrane and shows root-specific accumulation in Arabidopsis thaliana. Plant Cell Physiol. 47: Morillon R, Chrispeels MJ The role of ABA and the transpiration stream in the regulation of the osmotic water permeability of leaf cells. Proc. Natl. Acad. Sci. USA 98: Morillon R, Lassalles JP Osmotic water permeability of isolated vacuoles. Planta 210: Moshelion M, Becker D, Biela A, Uehlein N, Hedrich R, et al Plasma membrane aquaporins in the motor cells of Samanea saman: diurnal and circadian regulation. Plant Cell 14: Moshelion M, Moran N, Chaumont F Dynamic changes in the osmotic water permeability of protoplast plasma membrane. Plant Physiol. 135: Murai-Hatano M, Kuwagata T Osmotic water permeability of plasma and vacuolar membranes in protoplasts I: high osmotic water permeability in radish (Raphanus sativus) root cells as measured by a new method. J. Plant Res. 120: Nardini A, Salleo S, Andri S Circadian regulation of leaf hydraulic conductance in sunflower (Helianthus annuus L. cv Margot). Plant Cell Environ. 28: Niemietz CM, Tyerman SD Characterization of water channels in wheat root membrane vesicles. Plant Physiol. 115: Niemietz CM, Tyerman SD Channel-mediated permeation of ammonia gas through the peribacteroid membrane of soybean nodules. FEBS Lett. 465: Niemietz CM, Tyerman SD New potent inhibitors of aquaporins: silver and gold compounds inhibit aquaporins of plant and human origin. FEBS Lett. 531: Opperman CH, Taylor CG, Conkling MA Root-knot nematode-directed expression of a plant root-specific gene. Science 263: Paciorek T, Zazímalová E, Ruthardt N, Petrásek J, Stierhof YD, et al Auxin inhibits endocytosis and promotes its own efflux from cells. Nature 435: Phillips AL, Huttly AK Cloning of two gibberellin-regulated cdnas from Arabidopsis thaliana by subtractive hybridization: expression of the tonoplast water channel, γ-tip, is increased by GA 3. Plant Mol. Biol. 24: Picaud S, Becq F, Dédaldéchamp F, Ageorges A, Delrot S Cloning and expression of two plasma membrane aquaporins expressed during the ripening of grape berry. Funct. Plant Biol. 30: Pina C, Pinto F, Feijó JA, Becker JD Gene family analysis of the Arabidopsis pollen transcriptome reveals biological implications for cell growth, division control, and gene expression regulation. Plant Physiol. 138: Porcel R, Gómez M, Kaldenhoff R, Ruiz-Lozano JM Impairment of NtAQP1 gene expression in tobacco plants does not affect root colonisation pattern by arbuscular mycorrhizal fungi but decreases their symbiotic efficiency under drought. Mycorrhiza 15: Plant Aquaporins 621

74 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 113. Quigley F, Rosenberg JM, Shachar-Hill Y, Bohnert HJ From genome to function: the Arabidopsis aquaporins. Genome Biol. 3: Ramahaleo T, Morillon R, Alexandre J, Lassalles J-P Osmotic water permeability of isolated protoplasts. Modifications during development. Plant Physiol. 119: Rivers RL, Dean RM, Chandy G, Hall JE, Roberts DM, Zeidel ML Functional analysis of Nodulin 26, an aquaporin in soybean root symbiosomes. J. Biol. Chem. 272: Robinson DG, Sieber H, Kammerloher W, Schäffner AR PIP1 aquaporins are concentrated in plasmalemmasomes of Arabidopsis thaliana mesophyll. Plant Physiol. 111: Saadoun S, Papadopoulos MC, Hara-Chikuma M, Verkman AS Impairment of angiogenesis and cell migration by targeted aquaporin-1 gene disruption. Nature 434: Saito C, Ueda T, Abe H, Wada Y, Kuroiwa T, et al A complex and mobile structure forms a distinct subregion within the continuous vacuolar membrane in young cotyledons of Arabidopsis. Plant J. 29: Sakr S, Alves G, Morillon R, Maurel K, Decourteix M, et al Plasma membrane aquaporins are involved in winter embolism recovery in walnut tree. Plant Physiol. 133: Sakurai J, Ishikawa F, Yamaguchi T, Uemura M, Maeshima M Identification of 33 rice aquaporin genes and analysis of their expression and function. Plant Cell Physiol. 46: Santoni V, Verdoucq L, Sommerer N, Vinh J, Pflieger D, Maurel C Methylation of aquaporins in plant plasma membrane. Biochem. J. 400: Santoni V, Vinh J, Pflieger D, Sommerer N, Maurel C A proteomic study reveals novel insights into the diversity of aquaporin forms expressed in the plasma membrane of plant roots. Biochem. J. 372: Sarda X, Tousch D, Ferrare K, Legrand E, Dupuis JM, et al Two TIP-like genes encoding aquaporins are expressed in sunflower guard cells. Plant J. 12: Schäffner AR Aquaporin function, structure, and expression: are there more surprises to surface in water relations? Planta 204: Shangguan Z-P, Lei T-W, Shao M-A, Xue Q-W Effects of phosphorus nutrient on the hydraulic conductivity of sorghum (Sorghum vulgare Pers.) seedling roots under water deficiency. J. Integr. Plant Biol. 47: Shope JC, Mott KA Membrane trafficking and osmotically induced volume changes in guard cells. J. Exp. Bot. 57: Siefritz F, Otto B, Bienert GP, van der Krol A, Kaldenhoff R The plasma membrane aquaporin NtAQP1 is a key component of the leaf unfolding mechanism in tobacco. Plant J. 37: Siefritz F, Tyree MT, Lovisolo C, Schubert A, Kaldenhoff R PIP1 plasma membrane aquaporins in tobacco: from cellular effects to function in plants. Plant Cell 14: Sjövall-Larsen S, Alexandersson E, Johansson I, Karlsson M, Johanson U, Kjellbom P Purification and characterization of two protein kinases acting on the aquaporin SoPIP2;1. Biochim. Biophys. Acta 1758: Steudle E Water flow in plants and its coupling to other processes: an overview. Methods Enzymol. 174: Suga S, Komatsu S, Maeshima M Aquaporin isoforms responsive to salt and water stresses and phytohormones in radish seedlings. Plant Cell Physiol. 43: Maurel et al.

75 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 132. Suga S, Murai M, Kuwagata T, Maeshima M Differences in aquaporin levels among cell types of radish and measurement of osmotic water permeability of individual protoplasts. Plant Cell Physiol. 44: Tajkhorshid E, Nollert P, Jensen MØ, Miercke LJ, O Connell J, et al Control of the selectivity of the aquaporin water channel family by global orientational tuning. Science 296: Takano J, Wada M, Ludewig U, Schaaf G, von Wirén N, Fujiwara T The Arabidopsis major intrinsic protein NIP5;1 is essential for efficient boron uptake and plant development under boron limitation. Plant Cell 18: Temmei Y, Uchida S, Hoshino D, Kanzawa N, Kuwahara M, et al Water channel activities of Mimosa pudica plasma membrane intrinsic proteins are regulated by direct interaction and phosphorylation. FEBS Lett. 579: Terashima I, Ono K Effects of HgCl 2 on CO 2 dependence of leaf photosynthesis: Evidence indicating involvement of aquaporins in CO 2 diffusion across the plasma membrane. Plant Cell Physiol. 43: Törnroth-Horsefield S, Wang Y, Hedfalk K, Johanson U, Karlsson M, et al Structural mechanism of plant aquaporin gating. Nature 439: Tournaire-Roux C, Sutka M, Javot H, Gout E, Gerbeau P, et al Cytosolic ph regulates root water transport during anoxic stress through gating of aquaporins. Nature 425: Tuskan GA, Difazio S, Jansson S, Bohlmann J, Grigoriev I, et al The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science 313: Tyerman SD, Bohnert HJ, Maurel C, Steudle E, Smith JA Plant aquaporins: their molecular biology, biophysics and significance for plant water relations. J. Exp. Bot. 50: Tyerman SD, Niemietz CM Plant aquaporins: multifunctional water and solute channels with expanding roles. Plant Cell Environ. 25: Tyree MT, Nardini A, Salleo S, Sack L, El Omari B The dependence of leaf hydraulic conductance on irradiance during HPFM measurements: any role for stomatal response? J. Exp. Bot. 56: Uehlein N, Fileschi K, Eckert M, Bienert G, Bertl A, Kaldenhoff R Arbuscular mycorrhizal symbiosis and plant aquaporin expression. Phytochemistry 68: Uehlein N, Lovisolo C, Siefritz F, Kaldenhoff R The tobacco aquaporin NtAQP1 is a membrane CO 2 pore with physiological functions. Nature 425: Vander Willigen C, Postaire O, Tournaire-Roux C, Boursiac Y, Maurel C Expression and inhibition of aquaporins in germinating Arabidopsis seeds. Plant Cell Physiol. 47: Vera-Estrella R, Barkla BJ, Bohnert HJ, Pantoja O Novel regulation of aquaporins during osmotic stress. Plant Physiol. 135: Veselova TV, Veselovskii VA, Usmanov PD, Usmanova OV, Kozar VI Hypoxia and imbibition injuries to aging seeds. Russ. J. Plant. Physiol. 50: Volkov V, Hachez C, Moshelion M, Draye X, Chaumont F, Fricke W Water permeability differs between growing and nongrowing barley leaf tissues. J. Exp. Bot. 58: Wallace IS, Choi WG, Roberts DM The structure, function and regulation of the nodulin 26-like intrinsic protein family of plant aquaglyceroporins. Biochim. Biophys. Acta 1758: Plant Aquaporins 623

76 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org 150. Wallace IS, Roberts DM Homology modeling of representative subfamilies of Arabidopsis major intrinsic proteins. Classification based on the aromatic/arginine selectivity filter. Plant Physiol. 135: Wallace IS, Roberts DM Distinct transport selectivity of two structural subclasses of the nodulin-like intrinsic protein family of plant aquaglyceroporin channels. Biochemistry 44: Wan X, Steudle E, Hartung W Gating of water channels (aquaporins) in cortical cells of young corn roots by mechanical stimuli (pressure pulses): effects of ABA and of HgCl 2. J. Exp. Bot. 55: Werner M, Uehlein N, Proksch P, Kaldenhoff R Characterization of two tomato aquaporins and expression during the incompatible interaction of tomato with the plant parasite Cuscuta reflexa. Planta 213: Yang BX, Verkman AS Water and glycerol permeabilities of aquaporins 1 5 and MIP determined quantitatively by expression of epitope-tagged constructs in Xenopus oocytes. J. Biol. Chem. 272: Ye Q, Wiera B, Steudle E A cohesion/tension mechanism explains the gating of water channels (aquaporins) in Chara internodes by high concentration. J. Exp. Bot. 55: Yu QJ, Hu YL, Li JF, Wu Q, Lin ZP Sense and antisense expression of plasma membrane aquaporin BnPIP1 from Brassica napus in tobacco and its effects on plant drought tolerance. Plant Sci. 169: Yu X, Peng YH, Zhang MH, Shao YJ, Su WA, Tang ZC Water relations and expression analysis of plasma membrane intrinsic proteins in sensitive and tolerant rice during chiling and recovery. Cell Res. 16: Zelazny E, Borst JW, Muylaert M, Batoko H, Hemminga MA, Chaumont F FRET imaging in living maize cells reveals that plasma membrane aquaporins interact to regulate their subcellular localization. Proc. Natl. Acad. Sci. USA 104: Zhu C, Schraut D, Hartung W, Schäffner AR Differential responses of maize MIP genes to salt stress and ABA. J. Exp. Bot. 56: Maurel et al.

77 Contents Annual Review of Plant Biology Volume 59, 2008 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Our Work with Cyanogenic Plants Eric E. Conn 1 New Insights into Nitric Oxide Signaling in Plants Angélique Besson-Bard, Alain Pugin, and David Wendehenne 21 Plant Immunity to Insect Herbivores Gregg A. Howe and Georg Jander 41 Patterning and Polarity in Seed Plant Shoots John L. Bowman and Sandra K. Floyd 67 Chlorophyll Fluorescence: A Probe of Photosynthesis In Vivo Neil R. Baker 89 Seed Storage Oil Mobilization Ian A. Graham 115 The Role of Glutathione in Photosynthetic Organisms: Emerging Functions for Glutaredoxins and Glutathionylation Nicolas Rouhier, Stéphane D. Lemaire, and Jean-Pierre Jacquot 143 Algal Sensory Photoreceptors Peter Hegemann 167 Plant Proteases: From Phenotypes to Molecular Mechanisms Renier A.L. van der Hoorn 191 Gibberellin Metabolism and its Regulation Shinjiro Yamaguchi 225 Molecular Basis of Plant Architecture Yonghong Wang and Jiayang Li 253 Decoding of Light Signals by Plant Phytochromes and Their Interacting Proteins Gabyong Bae and Giltsu Choi 281 Flooding Stress: Acclimations and Genetic Diversity J. Bailey-Serres and L.A.C.J. Voesenek 313 v

78 Roots, Nitrogen Transformations, and Ecosystem Services Louise E. Jackson, Martin Burger, and Timothy R. Cavagnaro 341 A Genetic Regulatory Network in the Development of Trichomes and Root Hairs Tetsuya Ishida, Tetsuya Kurata, Kiyotaka Okada, and Takuji Wada 365 Molecular Aspects of Seed Dormancy Ruth Finkelstein, Wendy Reeves, Tohru Ariizumi, and Camille Steber 387 Trehalose Metabolism and Signaling Matthew J. Paul, Lucia F. Primavesi, Deveraj Jhurreea, and Yuhua Zhang 417 Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Auxin: The Looping Star in Plant Development René Benjamins and Ben Scheres 443 Regulation of Cullin RING Ligases Sara K. Hotton and Judy Callis 467 Plastid Evolution Sven B. Gould, Ross F. Waller, and Geoffrey I. McFadden 491 Coordinating Nodule Morphogenesis with Rhizobial Infection in Legumes Giles E.D. Oldroyd and J. Allan Downie 519 Structural and Signaling Networks for the Polar Cell Growth Machinery in Pollen Tubes Alice Y. Cheung and Hen-ming Wu 547 Regulation and Identity of Florigen: FLOWERING LOCUS T Moves Center Stage Franziska Turck, Fabio Fornara, and George Coupland 573 Plant Aquaporins: Membrane Channels with Multiple Integrated Functions Christophe Maurel, Lionel Verdoucq, Doan-Trung Luu, and Véronique Santoni 595 Metabolic Flux Analysis in Plants: From Intelligent Design to Rational Engineering Igor G.L. Libourel and Yair Shachar-Hill 625 Mechanisms of Salinity Tolerance Rana Munns and Mark Tester 651 Sealing Plant Surfaces: Cuticular Wax Formation by Epidermal Cells Lacey Samuels, Ljerka Kunst, and Reinhard Jetter 683 Ionomics and the Study of the Plant Ionome David E. Salt, Ivan Baxter, and Brett Lahner 709 vi Contents

79 Alkaloid Biosynthesis: Metabolism and Trafficking Jörg Ziegler and Peter J. Facchini 735 Genetically Engineered Plants and Foods: A Scientist s Analysis of the Issues (Part I) Peggy G. Lemaux 771 Indexes Cumulative Index of Contributing Authors, Volumes Cumulative Index of Chapter Titles, Volumes Annu. Rev. Plant Biol : Downloaded from arjournals.annualreviews.org Errata An online log of corrections to Annual Review of Plant Biology articles may be found at Contents vii

80 Plant Response to Waterdeficit Stress Elizabeth A Bray, University of Chicago, Chicago, Illinois, USA When plants do not receive sufficient water they are subjected to a stress called water deficit. Water deficit in the plant disrupts many cellular and whole plant functions, having a negative impact on plant growth and reproduction. Plants have evolved many different mechanisms to deal with the occurrence of this stress as it occurs in their environments. Availability of water is the most important factor in the environment that reduces the production of our crops. Effects of Water Deficit on Plant Growth and Development Drought, a period of abnormally dry weather, results in soil-water deficit and subsequently plant-water deficit. The lack of water in the environment constitutes a stress when it induces an injury in the plant. Water deficit in the plant disrupts many cellular and whole plant functions, having a negative impact on plant growth and reproduction. Crop yields are reduced by 69% on average when plants are exposed to unfavourable conditions in the field (Boyer, 1982). Availability of water is the most important factor in the environment that reduces the production of our crops. As water is increasingly needed for human populations and prime agricultural lands are used for housing, the availability of water will have a greater impact on our ability to produce crops. See also: Agricultural Production In nature, certain species are adapted to dry environments. The genotype determines the ability of the plant to survive and thrive in environments with low water availability. In addition, the duration of the water-deficit stress, the rate of stress imposition and the developmental stage of the plant at the time of stress imposition also affect plant growth and the ability of the plant to produce a crop. Whether the amount of water present in the environment is a stress is different for each species. Sensing mechanisms, yet to be identified, initiate the responses to water deficit, which occur at the molecular, metabolic, cellular, physiological and developmental levels. Many of these responses are driven by changes in gene expression. See also: Hot Deserts Resistance can Occur through Avoidance or Tolerance of Water-deficit Stress Resistance to water deficit may arise from the ability to tolerate water deficit or from mechanisms that allow Avoidance of water deficit Escape by completing life cycle in a wet season Avoid water deficit using strategies of maximum water acquisition Avoid water deficit by maintaining water in the cells Advanced article Article Contents. Effects of Water Deficit on Plant Growth and Development. Resistance can Occur through Avoidance or Tolerance of Water-deficit Stress. Water Dynamics of the Plant Cell. Whole Plant and Physiological Responses. Osmotic Adjustment Permits Water Uptake and Turgor Recovery. The Role of the Signalling Molecule Abscisic Acid. Genes Regulated by Water Deficit Affect Water-deficit Resistance. Mechanisms of Gene Regulation. Deduction of Gene Function through Overexpression in Transgenic Plants. Bringing our Knowledge to Crop Plants in the Field doi: / a pub2 Resistance to water-deficit stress Tolerance of water deficit Tolerate low water potential while maintaining turgor pressure (ψ p ) Tolerate cellular dehydration using biochemical and morphological strategies to protect cells from injury Figure 1 Resistance to water-deficit stress can arise from mechanisms involving avoidance or tolerance of the water deficit. avoidance of the water deficit (Figure 1). Some species, such as desert ephemerals, are able to escape drought by completing their life cycle when water is plentiful. Others avoid water deficit with the development of a large root system that permits improved extraction of water from the soil. Avoidance of water deficit may also be achieved by using mechanisms that save water as in succulents. Some plants may also have improved water-use efficiency, such as found in crassulacean acid metabolism plants, in which stomata are open at night and an alternative form of carbon assimilation promotes the use of less water. However, these plants do not tolerate water deficit. Other plants have biochemical and morphological mechanisms such as found in mosses and resurrection ENCYCLOPEDIA OF LIFE SCIENCES & 2007, John Wiley & Sons, Ltd. 1

81 Plant Response to Water-deficit Stress plants that permit the plants to withstand dehydration. See also: Crassulacean Acid Metabolism; Plant Stress Physiology Morphological characteristics of some species permit continued survival in arid environments, while other plants must acclimate to the stress to permit survival. Some responses may promote physiological adaptation to water deficit; other responses may indicate that an injury has occurred. It is difficult to distinguish functions of the responses to water-deficit stress Which responses are a direct result of water deficit and are effective in preparing the cell to function in a water-deficit condition? Some mechanisms promote survival and limit injury, but may do this using a mechanism that slows crop production, and are therefore not useful strategies to utilize in crop plants. It is important to identify mechanisms that permit continued growth during periods of water deficit to promote crop production. Water Dynamics of the Plant Cell The water within the cell is defined in terms of its free energy content or ability to do work. The free energy per unit volume of water is the water potential (c w ). Water is taken into the plant if the water potential is less than that of the environment surrounding the cell (Figure 2), since water moves down a chemical gradient. If the water potential in the soil solution is higher than that of the cells, water can be transported into the cells of the root. The water potential of the cell is dependent upon two important parameters: the osmotic potential (c s ) and the turgor pressure (c p ). The content of solutes in the water of the cell (c s ) and the pressure of the cellular contents against the cell wall (c p ) decrease the water potential. An additional component, ψ w outside > ψ w inside H 2 O H 2 O Osmotic adjustment ψ w outside < ψ w inside H 2 O H 2 O Figure 2 Cellular and soil-water potential control water uptake into the cell. Osmotic adjustment, a lowering of cellular osmotic potential, can permit water uptake and restore cellular turgor. the matric potential (c m ), or the binding of water to surfaces, also reduces the cell water potential. Equation (1) is used to describe cell water potential. c w ¼ c s þ c p þ c m Changes in cellular water relations trigger further events that are manifested in plant responses at the molecular, metabolic, cellular, physiological and developmental levels. See also: Plant Water Relations Whole Plant and Physiological Responses The water of the plant can also be viewed in the context of a soil plant air continuum: the plant is a column of water between the soil and the air. Transpiration, release of water from the plant, will continue when water is available. Under drought conditions in the field, soilwater content drops, which does not favour water movement into the root cells. Water will be lost through transpiration and will not be fully replaced, causing a loss in turgor in the plant cells. As a defence against water loss, transpiration decreases. In the leaf, the pore of the stomatal complex closes in response to soil-water deficit. As turgor decreases in the guard cells surrounding the stomatal pore, the cells fill the pore, thus reducing the stomatal aperture; this is the main cause of reduced transpiration. An Arabidopsis leucine-rich repeat receptor-like kinase (ERECTA) regulates leaf transpiration efficiency through a combined effect on stomatal density and photosynthetic capacity (Masle et al., 2005). When the stomata are closed, uptake of carbon dioxide is also reduced, reducing the carbon assimilation rate of the plant. Depending upon the duration of the water deficit, this may reduce crop production and cause injury to the chloroplasts through the process of photoinhibition. There may also be an interaction with other stresses, such as heat stress, when transpiration is reduced that will also contribute to the strain on the plant. See also: Forest Ecosystems; Photosynthesis: Ecology; Plant Stress Physiology; Plant Temperature Stress As cellular turgor approaches zero, growth will be inhibited. Cellular expansion or growth depends upon cellular turgor. The pressure of the cellular contents against the cell wall is the driving force for cell expansion and turgor is dependent upon uptake of water. The stage of development of the plant at the time of the occurrence of water deficit will alter the outcome. Growth will return if turgor is restored as the plant acclimatizes to water-deficit stress. See also: Plant Cell Growth and Elongation; Turgor Pressure ½1Š 2

Ecology of Water Relations in Plants

Ecology of Water Relations in Plants Ecology of Water Relations in s Yoseph Negusse Araya, The Open University, Milton Keynes, UK Water is an important resource for plant growth. Availability of water in the soil determines the niche, distribution

More information

Open Research Online The Open University s repository of research publications and other research outputs

Open Research Online The Open University s repository of research publications and other research outputs Open Research Online The Open University s repository of research publications and other research outputs Ecology of Water Relations in Plants Book Section How to cite: Araya, Yoseph Negusse (2007). Ecology

More information

Water Relations in Viticulture BRIANNA HOGE AND JIM KAMAS

Water Relations in Viticulture BRIANNA HOGE AND JIM KAMAS Water Relations in Viticulture BRIANNA HOGE AND JIM KAMAS Overview Introduction Important Concepts for Understanding water Movement through Vines Osmosis Water Potential Cell Expansion and the Acid Growth

More information

Hormonal and other chemical effects on plant growth and functioning. Bill Davies Lancaster Environment Centre, UK

Hormonal and other chemical effects on plant growth and functioning. Bill Davies Lancaster Environment Centre, UK Hormonal and other chemical effects on plant growth and functioning Bill Davies Lancaster Environment Centre, UK Integrating the impacts of soil drought and atmospheric stress High radiant load Reduced

More information

POTASSIUM IN PLANT GROWTH AND YIELD. by Ismail Cakmak Sabanci University Istanbul, Turkey

POTASSIUM IN PLANT GROWTH AND YIELD. by Ismail Cakmak Sabanci University Istanbul, Turkey POTASSIUM IN PLANT GROWTH AND YIELD by Ismail Cakmak Sabanci University Istanbul, Turkey Low K High K High K Low K Low K High K Low K High K Control K Deficiency Cakmak et al., 1994, J. Experimental Bot.

More information

Cytokinin. Fig Cytokinin needed for growth of shoot apical meristem. F Cytokinin stimulates chloroplast development in the dark

Cytokinin. Fig Cytokinin needed for growth of shoot apical meristem. F Cytokinin stimulates chloroplast development in the dark Cytokinin Abundant in young, dividing cells Shoot apical meristem Root apical meristem Synthesized in root tip, developing embryos, young leaves, fruits Transported passively via xylem into shoots from

More information

CHAPTER TRANSPORT

CHAPTER TRANSPORT CHAPTER 2 2.4 TRANSPORT Uptake of CO2 FOCUS: Uptake and transport of water and mineral salts Transport of organic substances Physical forces drive the transport of materials in plants over a range of distances

More information

Common Effects of Abiotic Stress Factors on Plants

Common Effects of Abiotic Stress Factors on Plants Common Effects of Abiotic Stress Factors on Plants Plants are living organisms which lack ability of locomotion. Animals can move easily from one location to other. Immovable property of plants makes it

More information

NOTES: CH 36 - Transport in Plants

NOTES: CH 36 - Transport in Plants NOTES: CH 36 - Transport in Plants Recall that transport across the cell membrane of plant cells occurs by: -diffusion -facilitated diffusion -osmosis (diffusion of water) -active transport (done by transport

More information

Stomata and water fluxes through plants

Stomata and water fluxes through plants Stomata and water fluxes through plants Bill Davies The Lancaster Environment Centre, UK Summary Stomata and responses to the environment Conductance, a function of frequency and aperture Measuring/estimating

More information

Resource acquisition and transport in vascular plants

Resource acquisition and transport in vascular plants Resource acquisition and transport in vascular plants Overview of what a plant does Chapter 36 CO 2 O 2 O 2 and and CO 2 CO 2 O 2 Sugar Light Shoots are optimized to capture light and reduce water loss

More information

Understanding how vines deal with heat and water deficit

Understanding how vines deal with heat and water deficit Understanding how vines deal with heat and water deficit Everard Edwards CSIRO AGRICULTURE & FOOD How hot is too hot? Cell death will occur in any vine tissue beyond a threshold (lethal) temperature cell

More information

AP Biology Chapter 36

AP Biology Chapter 36 Chapter 36 Chapter 36 Transport in Plants 2006-2007 Transport in plants - Overview H2O & minerals transport in xylem transpiration evaporation, adhesion & cohesion negative pressure Sugars transport in

More information

Chapter 12 & 13 Transport, Soil and Mineral Nutrition

Chapter 12 & 13 Transport, Soil and Mineral Nutrition Chapter 12 & 13 Transport, Soil and Mineral Nutrition Topics Methods of transport Xylem transport Phloem transport Soils properties and nutrient absorption Macro and micro essential nutrient elements Too

More information

Ch. 36 Transport in Vascular Plants

Ch. 36 Transport in Vascular Plants Ch. 36 Transport in Vascular Plants Feb 4 1:32 PM 1 Essential Question: How does a tall tree get the water from its roots to the top of the tree? Feb 4 1:38 PM 2 Shoot architecture and Light Capture: Phyllotaxy

More information

Movement of water and solutes in plants Chapter 4 and 30

Movement of water and solutes in plants Chapter 4 and 30 Movement of water and solutes in plants Chapter 4 and 30 Molecular Movement Diffusion Molecules or ions moving in the opposite direction = movement against a diffusion gradient. Rates of diffusion are

More information

Lecture notes on stomatal conductance. Agron 516: Crop physiology. Dr. Mark Westgate.

Lecture notes on stomatal conductance. Agron 516: Crop physiology. Dr. Mark Westgate. Lecture notes on stomatal conductance. Agron 516: Crop physiology. Dr. Mark Westgate. Diurnal variation of stomatal conductance has direct consequences for leaf and canopy gas exchange Measure diurnal

More information

Chapter 36: Transport in Vascular Plants - Pathways for Survival

Chapter 36: Transport in Vascular Plants - Pathways for Survival Chapter 36: Transport in Vascular Plants - Pathways for Survival For vascular plants, the evolutionary journey onto land involved differentiation into roots and shoots Vascular tissue transports nutrients

More information

C MPETENC EN I C ES LECT EC UR U E R

C MPETENC EN I C ES LECT EC UR U E R LECTURE 7: SUGAR TRANSPORT COMPETENCIES Students, after mastering the materials of Plant Physiology course, should be able to: 1. To explain the pathway of sugar transport in plants 2. To explain the mechanism

More information

Chapter 36~ Transport in Plants

Chapter 36~ Transport in Plants Chapter 36~ Transport in Plants Structural Features Used for Resource Acquistion Roots and stems to do transport of resources Diffusion, active transport, and bulk flow Work in vascular plants to transport

More information

Macro- and Microscopic Aspects of Fruit Water Relations Influencing Growth and Quality in Tomato

Macro- and Microscopic Aspects of Fruit Water Relations Influencing Growth and Quality in Tomato Macro- and Microscopic Aspects of Fruit Water Relations Influencing Growth and Quality in Tomato W. van Ieperen, U. van Meeteren, J. Oosterkamp and G. Trouwborst Department of Plant Sciences, Horticultural

More information

in angiosperms 10/29/08 Roots take up water via roots Large surface area is needed Roots branch and have root hairs Cortex structure also helps uptake

in angiosperms 10/29/08 Roots take up water via roots Large surface area is needed Roots branch and have root hairs Cortex structure also helps uptake in angiosperms A. Root System Roots take up water via roots Large surface area is needed Roots branch and have root hairs Cortex structure also helps uptake 1 B. Minerals Nitrogen (NO 3-,NH 4+ ) Potassium

More information

Water use efficiency in agriculture

Water use efficiency in agriculture Water use efficiency in agriculture Bill Davies The Lancaster Environment Centre, UK Summary Introduction and definitions Impacts of stomata, environment and leaf metabolism on WUE Estimating WUE and modifications

More information

Transport in Plants Notes AP Biology Mrs. Laux 3 levels of transport occur in plants: 1. Uptake of water and solutes by individual cells -for

Transport in Plants Notes AP Biology Mrs. Laux 3 levels of transport occur in plants: 1. Uptake of water and solutes by individual cells -for 3 levels of transport occur in plants: 1. Uptake of water and solutes by individual cells -for photosynthesis and respiration -ex: absorption of H 2 O /minerals by root hairs 2. Short distance cell-to-cell

More information

Organs and leaf structure

Organs and leaf structure Organs and leaf structure Different types of tissues are arranged together to form organs. Structure: 2 parts (Petiole and Leaf Blade) Thin flat blade, large surface area Leaves contain all 3 types of

More information

Introduction to Plant Transport

Introduction to Plant Transport Introduction to Plant Transport The algal ancestors of plants were completely immersed in water and dissolved minerals. The adaptation to land involved the differentiation of the plant body into roots,

More information

Transport of substances in plants

Transport of substances in plants Transport of substances in plants We have already looked at why many organisms need transport systems with special reference to surface area and volume. The larger the volume : surface area ratio, the

More information

Chapter 35 Regulation and Transport in Plants

Chapter 35 Regulation and Transport in Plants Chapter 35 Regulation and Remember what plants need Photosynthesis light reactions Calvin cycle light sun H 2 O ground CO 2 air What structures have plants evolved to supply these needs? Interdependent

More information

Translocation 11/30/2010. Translocation is the transport of products of photosynthesis, mainly sugars, from mature leaves to areas of growth and

Translocation 11/30/2010. Translocation is the transport of products of photosynthesis, mainly sugars, from mature leaves to areas of growth and Translocation Translocation is the transport of products of photosynthesis, mainly sugars, from mature leaves to areas of growth and storage. Phloem is the tissue through which translocation occurs. Sieve

More information

Plant Transport and Nutrition

Plant Transport and Nutrition Plant Transport and Nutrition Chapter 36: Transport in Plants H 2 O & Minerals o Transport in xylem o Transpiration Evaporation, adhesion & cohesion Negative pressure. Sugars o Transport in phloem. o Bulk

More information

BioWash as an Adjuvant, Translocation Promoter, and Cationic Exchange Stimulator Overview of Processes within the Plant

BioWash as an Adjuvant, Translocation Promoter, and Cationic Exchange Stimulator Overview of Processes within the Plant BioWash as an Adjuvant, Translocation Promoter, and Cationic Exchange Stimulator Overview of Processes within the Plant Photosynthesis is the primary driver of the plant. Through a series of complex steps,

More information

TREES. Functions, structure, physiology

TREES. Functions, structure, physiology TREES Functions, structure, physiology Trees in Agroecosystems - 1 Microclimate effects lower soil temperature alter soil moisture reduce temperature fluctuations Maintain or increase soil fertility biological

More information

Abiotic Stress in Crop Plants

Abiotic Stress in Crop Plants 1 Abiotic Stress in Crop Plants Mirza Hasanuzzaman, PhD Professor Department of Agronomy Sher-e-Bangla Agricultural University E-mail: mhzsauag@yahoo.com Stress Stress is usually defined as an external

More information

Question 1: What are the factors affecting the rate of diffusion? Diffusion is the passive movement of substances from a region of higher concentration to a region of lower concentration. Diffusion of

More information

of water unless it is moving via the symplast Water moves into the xylem for transport up the plant Water that does not cross the

of water unless it is moving via the symplast Water moves into the xylem for transport up the plant Water that does not cross the Uptake of water The through Casparian Strip blocks root epidermis by passage osmosis of water unless it is moving via the symplast Water moves into the xylem for transport up the plant Water that does

More information

Plant Ecophysiology in a Restoration Context

Plant Ecophysiology in a Restoration Context Objectives: How can the foundations of and theory in plant ecophysiological restoration ecology ecological restoration? Light and energy relations Photosynthesis Microclimate Belowground resource availability

More information

Transport in Vascular Plants

Transport in Vascular Plants Chapter 36 Transport in Vascular Plants PowerPoint Lectures for Biology, Seventh Edition Neil Campbell and Jane Reece Lectures by Chris Romero Vascular tissue Transports nutrients throughout a plant; such

More information

Introduction to Plant Transport

Introduction to Plant Transport Introduction to Plant Transport The algal ancestors of plants were completely immersed in water and dissolved minerals. The adaptation to land involved the differentiation of the plant body into roots,

More information

OCR (A) Biology A-level

OCR (A) Biology A-level OCR (A) Biology A-level Topic 3.3: Transport in plants Notes Plants require a transport system to ensure that all the cells of a plant receive a sufficient amount of nutrients. This is achieved through

More information

Importance. The Reaction of Life : The conversion of the sun s energy into a form man and other living creatures can use.

Importance. The Reaction of Life : The conversion of the sun s energy into a form man and other living creatures can use. PLANT PROCESSES Photosynthesis Importance The Reaction of Life : The conversion of the sun s energy into a form man and other living creatures can use. Photo light Synthesis to put together 3 Important

More information

CONTROL OF PLANT GROWTH AND DEVELOPMENT BI-2232 RIZKITA R E

CONTROL OF PLANT GROWTH AND DEVELOPMENT BI-2232 RIZKITA R E CONTROL OF PLANT GROWTH AND DEVELOPMENT BI-2232 RIZKITA R E The development of a plant the series of progressive changes that take place throughout its life is regulated in complex ways. Factors take part

More information

Transport in Plants. Transport in plants. Transport across Membranes. Water potential 10/9/2016

Transport in Plants. Transport in plants. Transport across Membranes. Water potential 10/9/2016 Transport in Plants Transport in plants How is a plant able to move water and nutrients from roots to the rest of the plant body? Especially tall trees? Sequoia can be over 300 feet tall! Transport across

More information

Chapter 4-2. Transpiration diffusion of water vapor

Chapter 4-2. Transpiration diffusion of water vapor Chapter 4-2 Transpiration diffusion of water vapor Transpiration diffusion of water vapor Diffusion is primary means of any further movement of the water out of the leaf. That is water movement is controlled

More information

Other Metabolic Functions of Water in Grapevines

Other Metabolic Functions of Water in Grapevines Other Metabolic Functions of Water in Grapevines Jim Kamas Assoc. Professor & Extension Specialist Texas A&M Agrilife Extension Viticulture & Fruit Lab Fredericksburg, TX Water is. 80 90% of the fresh

More information

PLANT SCIENCE. 9.2 Transport in Angiospermophytes

PLANT SCIENCE. 9.2 Transport in Angiospermophytes PLANT SCIENCE 9.2 Transport in Angiospermophytes Support of terrestrial plants Support of terrestrial plants comes through: Thickened cellulose in cell walls Turgor pressure of cells Lignified xylem Xylem

More information

Transport, Storage and Gas Exchange in Flowering Plants

Transport, Storage and Gas Exchange in Flowering Plants Sixth Year Biology Transport, Storage and Gas Exchange in Flowering Plants Miss Rochford In this topic: Uptake and transport of: Water and minerals Carbon dioxide Gas exchange Transport of photosynthesis

More information

Trees are: woody complex, large, long-lived self-feeding shedding generating systems compartmented, self optimizing

Trees are: woody complex, large, long-lived self-feeding shedding generating systems compartmented, self optimizing BASIC TREE BIOLOGY Trees are: woody complex, large, long-lived self-feeding shedding generating systems compartmented, self optimizing Roots: absorb water and minerals store energy support and anchor

More information

The three principal organs of seed plants are roots, stems, and leaves.

The three principal organs of seed plants are roots, stems, and leaves. 23 1 Specialized Tissues in Plants Seed Plant Structure The three principal organs of seed plants are roots, stems, and leaves. 1 of 34 23 1 Specialized Tissues in Plants Seed Plant Structure Roots: absorb

More information

CASE STUDY WATER ABSORPTION AND TRANSPORT IN PLANTS

CASE STUDY WATER ABSORPTION AND TRANSPORT IN PLANTS CASE STUDY WATER ABSORPTION AND TRANSPORT IN PLANTS Presentation of the problem: We need a pump to uplift water to a tank. The requirement of a pump is to pull water against the gravity. Look at the human

More information

CBSE Quick Revision Notes (Class-11 Biology) CHAPTER-11 TRANSPORT IN PLANTS

CBSE Quick Revision Notes (Class-11 Biology) CHAPTER-11 TRANSPORT IN PLANTS CBSE Quick Revision Notes (Class-11 Biology) CHAPTER-11 TRANSPORT IN PLANTS Plant transport various substance like gases, minerals, water, hormones, photosynthetes and organic solutes to short distance

More information

Biology 213 Exam 3 Practice Key

Biology 213 Exam 3 Practice Key Biology 213 Practice Key 1. (4) Explain the difference between a macronutrient and a micronutrient and cite two examples of each category? Macronutrients are the minerals needed by the plant in greater

More information

2018 Version. Photosynthesis Junior Science

2018 Version. Photosynthesis Junior Science 2018 Version Photosynthesis Junior Science 1 Plants fill the role of Producers in a community Plants are special because they have leaves and are able to produce their own food by the process of photosynthesis

More information

Hort Chapter 10 MENGEL et al, 5th Ed

Hort Chapter 10 MENGEL et al, 5th Ed Hort 5504 POTASSIUM Chapter 10 MENGEL et al, 5th Ed POTASSIUM CYCLE 1 SOIL POTASSIUM K minerals and K release ~2-3% of earth s crust is K K tied to clay particles (< 2 µm size) Frequently soils high in

More information

Biology 1030 Winter 2009

Biology 1030 Winter 2009 Meeting Tissue Needs II Chapter 36 (738-755) Chapter 37 (756-770) Cellular Currency Plants harvest solar energy Photosynthesis Produces sugars Proteins, nucleic acids, lipids? H 2 O CO 2 Plants cells still

More information

Interactions between ozone and drought stress in plants: mechanisms and implications. Sally Wilkinson and William J. Davies, Lancaster University

Interactions between ozone and drought stress in plants: mechanisms and implications. Sally Wilkinson and William J. Davies, Lancaster University Interactions between ozone and drought stress in plants: mechanisms and implications Sally Wilkinson and William J. Davies, Lancaster University STOMATA: At the leaf surface water is lost to the atmosphere

More information

ABSORPTION OF WATER MODE OF WATER ABSORPTION ACTIVE AND PASSIVE ABSORPTION AND FACTORS AFFECTING ABSORPTION.

ABSORPTION OF WATER MODE OF WATER ABSORPTION ACTIVE AND PASSIVE ABSORPTION AND FACTORS AFFECTING ABSORPTION. ABSORPTION OF WATER MODE OF WATER ABSORPTION ACTIVE AND PASSIVE ABSORPTION AND FACTORS AFFECTING ABSORPTION. PRELUDE OF WATER POTENTIAL Most organisms are comprised of at least 70% or more water. Some

More information

Describe plant meristems. Where are they located? perpetually embryonic cells found at tips of roots and shoots (apical vs.

Describe plant meristems. Where are they located? perpetually embryonic cells found at tips of roots and shoots (apical vs. Which conditions have the higher rate of transpiration? Light or dark: Humid or dry: Breezy or still air: Hot or warm: light (need CO 2 for photosyn.) dry (lower H 2 O potential out) breezy (greater evaporation)

More information

Bio Factsheet. Transport in Plants. Number 342

Bio Factsheet. Transport in Plants.   Number 342 Number 342 Transport in Plants This Factsheet: Explains why plants need a transport system Describes what plants transport Describes the tissues which carry out transport Outlines the position of the xylem

More information

Questions for Biology IIB (SS 2006) Wilhelm Gruissem

Questions for Biology IIB (SS 2006) Wilhelm Gruissem Questions for Biology IIB (SS 2006) Plant biology Wilhelm Gruissem The questions for my part of Biology IIB, Plant Biology, are provided for self-study and as material for the exam. Please note that the

More information

Transport in Plant (IGCSE Biology Syllabus )

Transport in Plant (IGCSE Biology Syllabus ) Transport in Plant (IGCSE Biology Syllabus 2016-2018) Plants have transport systems to move food, water and minerals around. These systems use continuous tubes called xylem and phloem: - Xylem vessels

More information

Chapter 30: Plant Nutrition & Transport

Chapter 30: Plant Nutrition & Transport Chapter 30: Plant Nutrition & Transport Carnivorous Plants Capture animals to supplement their nutrient intake Venus flytrap lures insects with sugary bait; closes on victim Cobra lily lures insects down

More information

REVERSAL OF ABSCISIC ACID INDUCED STOMATAL CLOSURE BY BENZYL ADENINE

REVERSAL OF ABSCISIC ACID INDUCED STOMATAL CLOSURE BY BENZYL ADENINE (1976) 7 REVERSAL OF ABSCISIC ACID INDUCED STOMATAL CLOSURE BY BENZYL ADENINE BY V. S. R. DAS, I. M. RAO AND A. S. RAGHAVENDRA* Department of Botany, Sri Venkateswara University, Tirupati 51702, India

More information

Chapter 25 Plant Processes. Biology II

Chapter 25 Plant Processes. Biology II Chapter 25 Plant Processes Biology II 25.1 Nutrients and Transport Plants grow by adding new cells through cell division Must have steady supply of raw materials to build new cells Nutrients (most) Plants

More information

HORMONAL MODIFICATION OF PLANT RESPONSE TO WATER STRESS By Y. MIZRAHI* and A. E. RWHMOND* [Manuscript received 8 November 1971] Abstract

HORMONAL MODIFICATION OF PLANT RESPONSE TO WATER STRESS By Y. MIZRAHI* and A. E. RWHMOND* [Manuscript received 8 November 1971] Abstract HORMONAL MODIFICATION OF PLANT RESPONSE TO WATER STRESS By Y. MIZRAHI* and A. E. RWHMOND* [Manuscript received 8 November 1971] Abstract Addition of either kinetin or abscisic acid (ABA) to the root medium

More information

STOLLER ENTERPRISES, INC. World leader in crop nutrition

STOLLER ENTERPRISES, INC. World leader in crop nutrition A new paradigm for crop production - Page 1 of 6 A NEW PARADIGM FOR CROP PRODUCTION Most agronomists are taught about the chemical process of manufacturing photosynthates (PS). The plants breathe in carbon

More information

THE EFFECT OF ABSCISIC ACID ON STOMATAL BEHAVIOUR IN FLACCA, A V\^ILTY MUTANT OF TOMATO, IN DARKNESS

THE EFFECT OF ABSCISIC ACID ON STOMATAL BEHAVIOUR IN FLACCA, A V\^ILTY MUTANT OF TOMATO, IN DARKNESS New Phytol (1972) 71, 81-84. THE EFFECT OF ABSCISIC ACID ON STOMATAL BEHAVIOUR IN FLACCA, A V\^ILTY MUTANT OF TOMATO, IN DARKNESS BY M. T A L AND D O R O T I M B E R Division of Life Sciences, Negev Institute

More information

MULTIPLE CHOICE QUESTIONS

MULTIPLE CHOICE QUESTIONS 54 BIOLOGY, EXEMPLAR PROBLEMS CHAPTER 11 TRANSPORT IN PLANTS MULTIPLE CHOICE QUESTIONS 1. Which of the following statements does not apply to reverse osmosis? a. it is used for water purification. b. In

More information

References. 1 Introduction

References. 1 Introduction 1 Introduction 3 tion, conservation of soil water may result in greater soil evaporation, especially if the top soil layers remain wetter, and the full benefit of sustained plant physiological activity

More information

EFFECTS OF ATMOSPHERIC CO 2 ENRICHMENT ON PLANT HORMONES

EFFECTS OF ATMOSPHERIC CO 2 ENRICHMENT ON PLANT HORMONES EFFECTS OF ATMOSPHERIC CO 2 ENRICHMENT ON PLANT HORMONES SPPI & CO2SCIENCE ORIGINAL PAPER August 29, 2012 EFFECTS OF ATMOSPHERIC CO 2 ENRICHMENT ON PLANT HORMONES Citation: Center for the Study of Carbon

More information

The Science of Plants in Agriculture Pl.Sci 102. Getting to Know Plants

The Science of Plants in Agriculture Pl.Sci 102. Getting to Know Plants The Science of Plants in Agriculture Pl.Sci 102 Getting to Know Plants Growth and Development of Plants Growth and Development of Plants Why it s important to have knowledge about plant development. What

More information

Two major categories. BIOLOGY 189 Fundamentals of Life Sciences. Spring 2004 Plant Structure and Function. Plant Structure and Function

Two major categories. BIOLOGY 189 Fundamentals of Life Sciences. Spring 2004 Plant Structure and Function. Plant Structure and Function BIOLOGY 189 Fundamentals of Life Sciences Spring 2004 Plant Structure and Function 18 16 14 12 10 8 6 Examination #1 Class Average: 33/60 for 55% 4 Chapters 31-32 32 2 0 6 10 15 20 25 30 35 40 45 50 55

More information

Major Plant Hormones 1.Auxins 2.Cytokinins 3.Gibberelins 4.Ethylene 5.Abscisic acid

Major Plant Hormones 1.Auxins 2.Cytokinins 3.Gibberelins 4.Ethylene 5.Abscisic acid Plant Hormones Lecture 9: Control Systems in Plants What is a Plant Hormone? Compound produced by one part of an organism that is translocated to other parts where it triggers a response in target cells

More information

TRANSPIRATION. An important regulator of transpiration is the stomatal complex composed of the opening or

TRANSPIRATION. An important regulator of transpiration is the stomatal complex composed of the opening or BIOL 1134 1 TRANSPIRATION LEARNING OBJECTIVES After completing this exercise, students should be able to: Describe the process of and principles behind transpiration. Describe how stomata, guard cells,

More information

Waterlogging tolerance of trees

Waterlogging tolerance of trees Waterlogging tolerance of trees Tapani Repo, Metla Silviculture in Changing Environment, Nov. 24-25, 2014 Contents Motivation Background concerning waterlogging tolerance An example of dormancy waterlogging

More information

Chapter III Inorganic Nutrients of Plants The main teaching content Focus and difficulties Teaching methods Section I water metabolism in plant

Chapter III Inorganic Nutrients of Plants The main teaching content Focus and difficulties Teaching methods Section I water metabolism in plant Chapter III Inorganic Nutrients of Plants The main teaching content: The importance of water in the plant vital movement, water potential,water transport and distribution in plants, physiological significance

More information

AP Biology. Transport in plants. Chapter 36. Transport in Plants. Transport in plants. Transport in plants. Transport in plants. Transport in plants

AP Biology. Transport in plants. Chapter 36. Transport in Plants. Transport in plants. Transport in plants. Transport in plants. Transport in plants Chapter 36. Transport in Plants evaporation, adhesion & cohesion negative pressure evaporation, adhesion & cohesion negative pressure transport in phloem bulk flow Calvin cycle in leaves loads sucrose

More information

Chapter 36. Transport in Vascular Plants

Chapter 36. Transport in Vascular Plants Chapter 36 Transport in Vascular Plants Overview: Pathways for Survival For vascular plants The evolutionary journey onto land involved the differentiation of the plant body into roots and shoots Vascular

More information

Level 2 Plant Growth and Development Part I Toby Day MSU Extension Horticulture Associate Specialist

Level 2 Plant Growth and Development Part I Toby Day MSU Extension Horticulture Associate Specialist Level 2 Plant Growth and Development Part I Toby Day MSU Extension Horticulture Associate Specialist Pages 24-38 Montana Master Gardener Handbook Plant Growth and Development Whole Plant Organs Tissues

More information

BIOL 221 Concepts of Botany Water Relations, Osmosis and Transpiration:

BIOL 221 Concepts of Botany Water Relations, Osmosis and Transpiration: BIOL 221 Concepts of Botany Topic 12: Water Relations, Osmosis and Transpiration: A. Water Relations Water plays a critical role in plants. Water is the universal solvent that allows biochemical reactions

More information

PH as a stress signal

PH as a stress signal Plant Growth Regulation 29: 87 99, 1999. 1999 Kluwer Academic Publishers. Printed in the Netherlands. 87 PH as a stress signal Sally Wilkinson Department of Biological Sciences, IENS, Lancaster University,

More information

Plant Development. Chapter 31 Part 1

Plant Development. Chapter 31 Part 1 Plant Development Chapter 31 Part 1 Impacts, Issues Foolish Seedlings, Gorgeous Grapes Gibberellin and other plant hormones control the growth and development of plants environmental cues influence hormone

More information

Increasing Processing Tomato Fruit Soluble Solids

Increasing Processing Tomato Fruit Soluble Solids Increasing Processing Tomato Fruit Soluble Solids Diane M Beckles Department of Plant Sciences dmbeckles@ucdavis.edu Processing Tomato Conference @ UC Davis December 13 th 2018 Soil Micronutrients Cultivar

More information

CHAPTER 32 TRANSPORT IN PLANTS OUTLINE OBJECTIVES

CHAPTER 32 TRANSPORT IN PLANTS OUTLINE OBJECTIVES CHAPTER 32 TRANSPORT IN PLANTS OUTLINE I. The traffic of water and solutes occurs on cellular, organ, and whole-plant levels: an overview of transport in plants A. Transport at the Cellular Level B. Short

More information

13. The diagram below shows two different kinds of substances, A and B, entering a cell.

13. The diagram below shows two different kinds of substances, A and B, entering a cell. Name 1. In the binomial system of nomenclature, which two classification groups provide the scientific name of an organism? A) kingdom and phylum B) phylum and species C) kingdom and genus D) genus and

More information

Discovery of compounds that keep plants fresh ~ Controlling plant pore openings for drought tolerance and delay in leaf withering ~

Discovery of compounds that keep plants fresh ~ Controlling plant pore openings for drought tolerance and delay in leaf withering ~ Discovery of compounds that keep plants fresh ~ Controlling plant pore openings for drought tolerance and delay in leaf withering ~ April 9, 2018 A team of scientists at Nagoya University has discovered

More information

BIOL 1030 Introduction to Biology: Organismal Biology. Fall 2009 Sections B & D. Steve Thompson:

BIOL 1030 Introduction to Biology: Organismal Biology. Fall 2009 Sections B & D. Steve Thompson: BIOL 1030 Introduction to Biology: Organismal Biology. Fall 2009 Sections B & D Steve Thompson: stthompson@valdosta.edu http://www.bioinfo4u.net 1 How plants get the stuff they need Feed me... feed me...

More information

NAME ONE THING we have in common with plants. If

NAME ONE THING we have in common with plants. If Cellular Respiration NAME ONE THING we have in common with plants. If you said cellular respiration, you are right. That is one thing we have in common with plants, slugs, slime mold, and spiders. Living

More information

Plant Growth and Development

Plant Growth and Development Plant Growth and Development Concept 26.1 Plants Develop in Response to the Environment Factors involved in regulating plant growth and development: 1. Environmental cues (e.g., day length) 2. Receptors

More information

Bio 102 Chapter 32 Transport in Plants

Bio 102 Chapter 32 Transport in Plants Bio 102 Chapter 32 Transport in Plants 2006-2007 Passive Water & Mineral Absorption Water absorption from soil OSMOSIS = transport of WATER across cell membrane WATER POTENTIAL determines direction of

More information

Why Calcium is So Important

Why Calcium is So Important Why Calcium is So Important Calcium - A Transportation Problem By Dr. Lynette Morgan As hydroponic growers we like to think that by supplying our plants with all the nutrients they need in the right ratios,

More information

Exchanging Materials in Plants

Exchanging Materials in Plants Exchanging Materials in Plants 1 of 23 Boardworks Ltd 2012 2 of 23 Boardworks Ltd 2012 3 of 23 Boardworks Ltd 2012 All living things need to exchange materials Plants need to obtain certain materials for

More information

35 Transport in Plants

35 Transport in Plants Transport in Plants 35 Transport in Plants 35.1 How Do Plants Take Up Water and Solutes? 35.2 How Are Water and Minerals Transported in the Xylem? 35.3 How Do Stomata Control the Loss of Water and the

More information

BIOL 221 Concepts of Botany Spring Water Relations, Osmosis and Transpiration

BIOL 221 Concepts of Botany Spring Water Relations, Osmosis and Transpiration BIOL 221 Concepts of Botany Spring 2008 Topic 07: Water Relations, Osmosis and Transpiration A. Water Relations Water plays a critical role in plants. Water is the universal solvent that allows biochemical

More information

Title Allantoin by Inosine in Nutrient So. Author(s) Toshihiro; Yokoi, Daisuke; Osaki, M

Title Allantoin by Inosine in Nutrient So. Author(s) Toshihiro; Yokoi, Daisuke; Osaki, M Title Rice Root Growth with Increasing in Allantoin by Inosine in Nutrient So Author(s) Tokuhisa, Dai; Okazaki, Keiki; Shin Toshihiro; Yokoi, Daisuke; Osaki, M Citation The Proceedings of the Internationa

More information

B2 Quick Revision Questions. B2 for AQA GCSE examination 2018 onwards

B2 Quick Revision Questions. B2 for AQA GCSE examination 2018 onwards B2 Quick Revision Questions Question 1 Which raw materials are used in photosynthesis and what are the products of the reaction? Answer 1 Carbon dioxide Water Glucose Oxygen Question 2 What type of reaction

More information

Synoptic Biology: Water Potential

Synoptic Biology: Water Potential Number 225 Synoptic Biology: Water Potential This Factsheet reviews the very wide range of exam questions that test your understanding of water potential. Osmosis: the diffusion of water through a partially

More information

Drought Signaling in Plants

Drought Signaling in Plants Drought Signaling in Plants G Sivakumar Swamy Plants are itrmly anchored in the soil, and they cannot evade the vagaries of nature by moving towards a convenient location as we normally encounter in case

More information

Transport in Plants (Ch. 23.5)

Transport in Plants (Ch. 23.5) Transport in Plants (Ch. 23.5) Transport in plants H 2 O & minerals transport in xylem Transpiration Adhesion, cohesion & Evaporation Sugars transport in phloem bulk flow Gas exchange photosynthesis CO

More information

Modulation of Root Signals in Relation to Stomatal Sensitivity to Root-sourced Abscisic Acid in Drought-affected Plants

Modulation of Root Signals in Relation to Stomatal Sensitivity to Root-sourced Abscisic Acid in Drought-affected Plants Journal of Integrative Plant Biology 2007, 49 (10): 1410 1420 Invited Review Modulation of Root Signals in Relation to Stomatal Sensitivity to Root-sourced Abscisic Acid in Drought-affected Plants Huibo

More information

Chapter C3: Multicellular Organisms Plants

Chapter C3: Multicellular Organisms Plants Chapter C3: Multicellular Organisms Plants Multicellular Organisms Multicellular organisms have specialized cells of many different types that allow them to grow to a larger size than single-celled organisms.

More information