Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

Embed Size (px)

Citation preview

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    1/14

     (201), re4. [DOI: 10.1126/scisignal.2002164]4Science Signaling Archana Joshi-Saha, Christiane Valon and Jeffrey Leung (29 November 2011)Guard Cell

    A Brand New START: Abscisic Acid Perception and Transduction in the`

    This information is current as of 1 October 2012.The following resources related to this article are available online at http://stke.sciencemag.org.

    Article Tools http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4

    Visit the online version of this article to access the personalization and article tools:

    Related Content

     http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/173/ra32 http://stke.sciencemag.org/cgi/content/abstract/sigtrans;5/226/eg7

    's sites:Science The editors suggest related resources on

    References  http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBL 1 article(s) hosted by HighWire Press; see:cited byThis article has been

    http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#otherarticlesThis article cites 126 articles, 60 of which can be accessed for free:

    Glossary http://stke.sciencemag.org/glossary/

    Look up definitions for abbreviations and terms found in this article:

    Permissions http://www.sciencemag.org/about/permissions.dtl

    Obtain information about reproducing this article:

    Association for the Advancement of Science; all rights reserved.

    for the Advancement of Science, 1200 New York Avenue, NW, Washington, DC 20005. Copyright 2008 by the American(ISSN 1937-9145) is published weekly, except the last week in December, by the American AssociationScience Signaling 

    http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/173/ra32http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/173/ra32http://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/173/ra32http://stke.sciencemag.org/cgi/content/abstract/sigtrans;5/226/eg7http://stke.sciencemag.org/cgi/content/abstract/sigtrans;5/226/eg7http://stke.sciencemag.org/cgi/content/abstract/sigtrans;5/226/eg7http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#otherarticleshttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#otherarticleshttp://stke.sciencemag.org/glossary/http://stke.sciencemag.org/glossary/http://stke.sciencemag.org/glossary/http://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://www.sciencemag.org/about/permissions.dtlhttp://stke.sciencemag.org/glossary/http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#otherarticleshttp://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4#BIBLhttp://stke.sciencemag.org/cgi/content/abstract/sigtrans;4/173/ra32http://stke.sciencemag.org/cgi/content/abstract/sigtrans;5/226/eg7http://stke.sciencemag.org/cgi/content/full/sigtrans;4/201/re4http://oascentral.sciencemag.org/RealMedia/ads/click_lx.ads/sciencesignaling/cgi/reprint/L28/1815059444/Bottom1/AAAS/PDF-Signaling-CellSignalingTechnology-120808/CSTsponsor.raw/1?x

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    2/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 1

     IntroductionLand plants, being rooted in place, must

    sense and adapt to their incessantly fluctuat-

    ing surroundings. At one time or another, wehave all noticed a plant neglected in a shady

    corner exhibiting heliotropism: leaning out

    and orienting its leaves perpendicularly to-

    ward the Sun’s rays to maximize photosyn-

    thesis. In contrast to laboratory settings,

     plants in heterogeneous field conditions

    must continuously cope with the constraints

    in their environments to optimize growth.

    In a given day, the plant will have endured

    transient and local differences in light qual-

    ity and intensity, fluctuations in temperature,

    humidity, CO2, and possibly uneven water

    distribution in the soil. Carbon dioxide and

    water are among the most important ingre-dients contributing to plant biomass [6CO2 

    + 6H2O + light → sugar + 6O2]; in return,

     plants recycle CO2 and H2O, with O2 release

    as a photosynthetic by-product (1). Because

    land plants are protected by a waxy layer,

    transpirational water loss and gas exchange

    with the surrounding atmosphere are pos-sible almost exclusively through stomatal

     pores (Fig. 1A). In most plants, these pores

    are flanked by a pair of kidney-shaped guard

    cells, whose volumes can expand or contract

     by changes in turgor pressure to control the

    opening and closing of the pore, respective-

    ly (Fig. 1B). Sunlight stimulates stomatal

    opening to allow diffusion of atmospheric

    CO2  to photosynthetic tissues. This, how-

    ever, exacts a trade-off in water loss through

    transpiration, which will compromise growth

    (2). Stomates also close in response to el-

    evated CO2, the cause of global climate

    warming, and to high concentrations of pol-lutants, such as ozone (O3), presumably to

     prevent oxidative damage (3, 4). Guard cells

    are, therefore, multisensorial and integrate

    diverse cues in the leaf environment with

    endogenous growth signals to optimize the

     plant’s conflicting needs.

    The turgor pressure within the guard

    cells is modulated by the dynamic changes

    in intracellular concentrations of inorganic

    and organic ions (K +, Cl – , NO3 – , malate) and

    sugars. Depending on the nature of the in-

     put stimuli, coordinated ionic fluxes across

    membranous compartments of the guard

    cell will be assured by teams of different

    channels and transporters. The electrical

    signals generated by ion fluxes across the

     plasma membrane are then converted bythe cell into chemical messages to shape the

    final physiological output (in this case, the

     binary decision of stomatal opening or clos-

    ing). Because the guard cell is accessible

    to studies by pharmacological approaches,

    genetics, and molecular biology, it serves

    as an excellent higher plant cell model for

    unraveling signal integration between the

    environment and endogenous growth fac-

    tors. Furthermore, understanding the major

    mechanistic aspects of abscisic acid (ABA)

    signaling, which promotes stomatal closure

    in guard cells, will not be unique to this cell

    type but can provide a base to extend toother plant tissues or organs that respond to

    this hormone.

    A Retrospective on Abscisic AcidSignaling in the Guard Cell—TheCircuitry of Ion FluxesWater deficit stimulates the synthesis and,

    to a much lesser extent (~5%), release from

    storage of the hormone ABA to promote sto-

    matal closure. The early physiological intri-

    cacies of ABA signaling in guard cells were

    teased out largely by pharmacological and

     biophysical approaches that provided the

    first sketch of the circuitry (Fig. 1B). Thesestudies showed that an early detectable event

    triggered by ABA is the production of reac-

    tive oxygen species [(ROS), sometimes also

    known as oxidative burst], which stimulates

    Ca2+ release from internal stores and influx

    across the plasma membrane (5). The Ca2+-

    dependent release of anions (often simply

    referred to as Cl –  in the early days) into the

    apoplast, which is formed by the continuum

    of cell walls of adjacent cells as well as the

    extracellular spaces, causes depolarization

    of the plasma membrane (6  – 10). At the

    same time, Ca2+  also prevents membrane

    hyperpolarization by inhibiting H+ –ad-enosine triphosphatases [(ATPases), proton

     pumps coupled to ATP hydrolysis] required

    to drive stomatal opening (11). Two distinct

    types of anion efflux currents are detectable

    in the guard cell, designated as slow (S) or

    rapid (R) (12, 13). The S-type current is car-

    ried by a range of anions that include NO 3 – ,

    Cl – , and malate (14, 15). It was proposed

    that the S-type, and not the R-type, current

    was responsible for ABA-mediated stomatal

    P L A N T B I O L O G Y

    A Brand New START: Abscisic AcidPerception and Transduction in the

    Guard CellArchana Joshi-Saha,1 Christiane Valon,1 Jeffrey Leung1*

    *Corresponding author. E-mail, [email protected]

    1Institut des Sciences du Végétal, Centre Na-tional de la Recherche Scientifique, Unité Pro-pre de Recherche 2355, 1 Avenue de la Terrasse,Bâtiment 23, 91198 Gif-sur-Yvette, France.

    The soluble receptors of abscisic acid (ABA) have been identified in Arabidopsisthaliana . The 14 proteins in this family, bearing the double name of PYRABACTINRESISTANCE/PYRABACTIN-LIKE (PYR/PYL) or REGULATORY COMPONENTSOF ABA RECEPTOR (RCAR) (collectively referred to as PYR/PYL/RCAR), con-tain between 150 and 200 amino acids with homology to the steroidogenic acuteregulatory-related lipid transfer (START) protein. Structural studies of these re-ceptors have provided rich insights into the early mechanisms of ABA signaling.The binding of ABA to PYR/PYL/RCAR triggers the pathway by inducing struc-tural changes in the receptors that allows them to sequester members of theclade A negative regulating protein phosphatase 2Cs (PP2Cs). This liberates theclass III ABA-activated Snf1-related kinases (SnRK2s) to phosphorylate varioustargets. In guard cells, a specific SnRK2, OPEN STOMATA 1 (OST), stimulatesH2O2 production by NADPH oxidase respiratory burst oxidase protein F and in-hibits potassium ion influx by the inward-rectifying channel KAT1. OST1, the ki-nase CPK23, the calcium-dependent kinase CPK21, and the counteracting PP2Csmodulate the slow anion channel SLAC1, a pathway that contributes to stomatalresponses to diverse stimuli, including ABA and carbon dioxide. A minimal ABAresponse pathway that leads to activation of the SLAC1 homolog, SLAH3, andpresumably stomatal closure has been reconstituted in vitro. The identificationof the soluble receptors and core components of the ABA signaling pathway pro-vides promising targets for crop design with higher resilience to water deficitwhile maintaining biomass.

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    3/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 2

    closure. However, it was not clear at the time

    whether the two types of anion currents oc-

    curred through distinct channels or through

    modification of the same channel. ABA alsoinduces the net efflux of both K +  and Cl – 

    from the vacuole, which occupies 90% of

    the guard cell volume, to the cytoplasm and

    from the cytoplasm to outside of the guard

    cell. At least one other signaling branch of

    the pathway is independent of Ca2+, instead

    requiring cytoplasmic alkalinization as an

    intermediate (6 ). This pH-sensitive signal-

    ing branch modulates K +  efflux through

    outward-rectifying channels.

    In the early 1990s, indirect evidence

    was obtained for ABA perception sites on

    the “outside” (cell surface), as well as on

    the “inside” of the guard cell. Stomatal clo-sure can be evoked by exogenously applied

    ABA, hinting at an “outside” perception

    site; however, the protonated form of the

    weak acid ABA can readily permeate the

    lipid bilayer of the cell membrane, which

    could still suggest the possibility of cyto-

    solic receptors. Furthermore, the guard cells

    of Commelina communis  have substantial

    carrier-mediated uptake of ABA (16 , 17 ),

    which could deliver externally applied ABA

    to intracellular reception sites. The require-

    ment for an extracellular receptor was sug-

    gested by the ability of ABA at high pH,

    when it is charged and can no longer crossthe plasma membrane, to induce stomatal

    closure in Valerianella locusta (18) and the

    failure of ABA to inhibit stomatal opening

    when microinjected directly in the cytosol

    of Commelina  guard cells (19). Externally

    applied ABA to barley aleurone protoplasts

    (single cells without the cell wall derived

    from barley ovules) reversed the stimulation

    of α-amylase synthesis by gibberellic acid,

    whereas microinjecting up to 250 µM ABA

    was ineffective (20). A cell surface–local-

    ized receptor was also concordant with K +

    fluxes (21) and reporter gene expression in

    either  Arabidopsis or rice cell cultures thatwere stimulated by ABA coupled to carri-

    ers that could not penetrate membranes (21,

    22). Immunolocalization of presumptive ex-

    tracellular ABA binding sites has also been

    reported (23). On the other hand, there was

    accumulating evidence for internal ABA

    reception sites. In Vicia faba, externally ap-

     plied ABA was not effective in maintaining

    slow anion channel current in ATP-depleted

    V. faba guard cell protoplasts (24). In Com-

    melina, extracellular ABA was compara-

    tively less effective in regulating stomatal

    aperture at higher versus lower pH, the latter

    of which favors uptake by passive diffusionof the protonated form of ABA, providing

    indirect evidence for an intracellular recep-

    tor (17 , 19, 25, 26 ). The presence of an in-

    tracellular receptor was also supported by

    reports that stomatal closure was triggered

    in Commelina by release of caged ABA in

    the cytosol of guard cells (27 ) and by ABA

    directly microinjected into guard cells (17 ).

    The Essential Components of theCore Signaling Complex: SolubleReceptor, Protein Phosphatase 2C,and the Kinase SnRK2

    To dissect the underlying mechanisms bywhich ABA rapidly causes stomatal clos-

    ing, guard cell signaling came under joint

    assault in the early 1990s by genetics and

    molecular biology. Several putative and

    somewhat controversial ABA receptors

    have been proposed intermittently since the

    first report more than 25 years ago of bind-

    ing proteins in the plasmalemma of V. faba

    guard cells (28, 29). After many candidates

    that have been described as false starts (30),

    Fig. 1. Biophysics of stomatal movement. (A) One-week-old Ara- bidopsis   rosette leaf is shown with an image of a single stomate.The microscopic pores contoured by the two flanking guard cellsdefine a stomatal opening. The fluorescent round structures arechloroplasts. (B) Stomatal opening (left) and closing (right). Increas-ing turgor pressure inside the cells causes the two cells to swelland bow out from each other, resulting in the opening of the pore.Stomatal opening requires hyperpolarization of the plasma mem-brane and entry of K+. ABA accumulates in response to droughtand fosters stomatal closing. The earliest detectable signal is thepresence of reactive oxygen species (H2O2) and then a transient

    increase in Ca2+. A second signaling intermediate is sensitive to pHand stimulates K+ efflux through K+-outward rectifying channels. Be-cause 90% of the volume of the guard cell is the vacuole, the effluxof ions must first cross the vacuolar membrane (pale green oblongstructure), then the plasma membrane, and finally move into theapoplastic space. For simplicity, the left cell shows the plasma mem-brane proteins that are active during stomatal opening, and the rightcell shows the plasma membrane proteins that are active during sto-matal closure. Red, H+-ATPase; yellow, K+ inward-rectifying channel;light blue, Ca2+ permeant channel; dark blue, anion channels; lightgreen, K+ outward-rectifying channel.

    Ca2+

    A   B

    Low humidity

    ABA

    High CO2 

    Darkness

    ABA

    ABA

    Fusicoccin

    Light

    High humidity

    Low CO2

    Light / Low CO2

    A-

    Depolarization

    K +

    K +

    H+

    Hyperpolarization

    H2O2

    H2O

    2

    pH

       C   R   E   D   I   T   S  :   (   A   )   J   E   F   F   R   E   Y   L   E   U   N   G  ;   (   B   )   Y .   H   A   M   M   O   N   D   /   S   C   I   E   N   C   E   S   I   G   N   A   L   I   N   G

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    4/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 3

    those with properties matching

    the required physiological and

    molecular profiles were identi-

    fied in 2009. This momentous

    discovery, celebrated as one of

    the top 10 discoveries of the

    year (31), represents an awaken-ing in our understanding of the

    initial ABA signaling events.

    This soluble receptor family

    has 14 members in  Arabidopsis

    thaliana, and this high degree

    of functional redundancy may

    have cloaked their identity from

     being revealed by standard ge-

    netic screens (at least for the

    loss-of-function category of mu-

    tations). However, the discovery

    that the synthetic compound,

    4-bromo-N-[pyridin-2-yl meth-

    yl]naphthalene-1-sulfonamide, known as pyrabactin (Fig. 2), partially mimicked

    the inhibitory effects of ABA on seed ger-

    mination, early seedling growth, and gene

    expression program (32), presumably by

     binding and modifying the activities of sev-

    eral of the soluble ABA receptors simulta-

    neously, allowed Park et al . to side-step the

    functional redundancy barrier. By selecting

    for mutagenized seeds that germinated in

    the presence of inhibitory concentrations

    of pyrabactin, followed by map-based clon-

    ing of one such locus named PYRABACTIN

     RESISTANCE 1 ( PYR1), they succeeded in

    isolating a gene encoding a homolog to thesteroidogenic acute regulatory lipid trans-

    fer (START) proteins. These proteins are

    characterized by a structural scaffold that

    can accommodate a large spectrum of hy-

    drophobic ligands, such as lipids, antibiot-

    ics, and hormones (33). Importantly, PYR1

    interacted in a yeast two-hybrid screen, in

    an ABA-dependent manner, with several

     phosphatases of the protein phosphatase 2C

    (PP2C) family (namely, ABI1, ABI2, and

    HAB1) that had previously been established

    as key negative regulators of the ABA sig-

    naling cascade (34 – 40). Moreover, the pro-

    tein-protein interaction pattern established by the yeast two-hybrid system related to

    the phenotypes of the respective mutant

     plants. For example, the loss-of-function

    mutations Pro88  to Ser 88  (P88S) and Ser 152

    to Leu152  (S152L) in PYR1 that confer the

    receptor’s property of pyrabactin resistance

    and, conversely, the gain-of-function Gly168

    to Asp168 (G168D) mutation in ABI2, which

    defined this negative regulator and confers

    ABA resistance in the mutant plants (abi2-

    1), all reduced the interaction between thereceptor and PP2C. Independently, Ma et al .

    isolated a protein named REGULATORY

    COMPONENT OF ABA RECEPTOR 1

    (RCAR1), as a partner of ABI1 and ABI2

    (41). The mutation G168D in ABI2 (abi2-1)

    or the equivalent mutation G180D in ABI1

    (abi1-1) also abolished its interactions with

    RCAR1 (also known as PYL9).

    The crystal structures of several recep-

    tors were rapidly accomplished, and fine

    structural details were obtained in quick suc-

    cession for (i) ABA-bound, (ii) pyrabactin-

     bound, (iii) and ligand-free (PYR1, PYL1,

    PYL2) apo-receptors, as well as (iv) thetertiary complex ABA-PYL/RCAR-PP2C.

    These structural studies have uncovered

    a wealth of mechanistic insights into the

    early events of ABA signaling (42 – 50). The

    receptor protein has a central cradle that is

    formed by the alignment of seven-stranded

    antiparallel β  sheets wrapped around by along α helix from the C-terminal end of the

     protein. The bottom of the cradle is created

     by two other α helices situated between thefirst and second β sheets. The ABA moleculeis held inside the cavity by a combination

    of nonpolar and polar interactions. Among

    the charged interactions, the carboxylgroup of the ABA is plunged deep within

    the pocket, and it is in direct contact with

    Lys59 of PYR1 (or Lys86 of PYL1 and Lys64

    of PYL2) (Fig. 3, A, C, and D). This lysine

    is conserved in the gene family, with the ex-

    ception of PYL13 in which this residue is

    occupied by a glutamine. The access to the

    ABA molecule from outside the receptor is

    controlled by two important structures: The

    first is called the proline “gate” (with the

    signature amino acid motif SGLPA; A, Ala),

    which is conserved in all of the receptors

    except, again, PYL13, in which the leucine

    is replaced by phenylalanine. The second

    functionally important domain is called the

    leucine “latch” GG(E/D)HRL (where the

    slash means “or”; E, Glu; H, His; R, Arg),again with PYL13 as the outgroup having

    the E/D residue substituted by glutamine.

    The cyclohexane ring of the ABA molecule

    (Fig. 2A) extends toward the opening of the

     binding cavity and stabilizes the gate in the

    closed conformation by interactions with a

    number of hydrophobic amino acids, which

    are also conserved in all 14 receptor mem-

     bers (51). This closing of the gate is further

    secured by the positioning of the latch and

    the extension of an α-helical loop (“recoil”region) (50). This recoil region encompass-

    es 13 amino acids (Met147 to Phe159 in PYR,

    which align with Val177

     to Phe189

     in PYL1 inFig. 3) that, after ABA binding, coil into the

    C-terminal α helix of the receptor.In the absence of ABA, the receptor

    (PYR as the model) exists as an asymmet-

    ric dimer (50) with ~10°  deviation from atwofold (180°) symmetry. These mono-meric subunits are held together through

     bonds between their gates. The binding of

    ABA leads to conformational changes in the

    gate to allow the dimer to assume a perfect

    twofold symmetry, resulting in a more com-

     pact structure with a biconcave disc shape

    resembling a red blood cell. Coimmunopre-

    cipitation assays confirmed the existenceof dimer in vivo both with and without ex-

    ogenous ABA. It seems that ABA binding

    causes the dimer to dissociate into mono-

    mers and each monomer then binds a PP2C

    (52). Although the chemical structure of the

    agonist pyrabactin is very different from

    that of ABA (Fig. 2), pyrabactin binds (as

    a folded structure resembling π) PYR1 andat least several other member receptors to

    form a “productive” complex in which the

    gate is closed. However, pyrabactin does not

    activate PYL2; instead, it binds and forms

    a “nonproductive” complex. Thus, this syn-

    thetic compound could theoretically antago-nize activation of PYL2 by ABA (45, 49).

    In the productive ABA receptor–pyrabactin

    configurations [derived from the structural

    studies of PYL1-pyrabactin (45), PYR1-

     pyrabactin, and PYL1-pyrabactin-ABI1 (45,

    49)], the orientation of the bound pyrabactin

     provides the necessary van der Waals inter-

    actions to induce gate closure. In contrast,

    in the nonproductive mode [deduced from

    the PYL2-pyrabactin structure (45, 49)], the

    N

    O   OH

    OH

    OO=S=O

    Br

    NH

    O=S=O

    Br

    NH

    Fig. 2. ABA and pyrabactin used in classical and chemi-cal genetic screens, respectively, to identify signalingcomponents. (A) Chemical structure of ABA, highlightingits active carboxyl group that directly binds to a conservedLys residue (Lys86 in PYL1) deep in the pocket of the re-ceptors (except PYL13). (B) Structures of the syntheticchemical pyrabactin (left) and its analog apyrabactin(right). The pyradyl nitrogen (arrow) is important for thepyrabactin agonist effect, because apyrabactin is inactive.

       C   R   E   D   I   T  :   Y .   H   A   M   M   O   N   D   /   S   C   I   E   N   C   E   S   I   G   N   A   L   I   N   G

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    5/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 4

    relative orientation of the pyrabactin along

    the length of the molecule is flipped by 180° 

    in PYL2. The pyrabactin in this inverted ori-

    entation can no longer supply the necessary

    van der Waals forces to maintain gate clo-

    sure. Thus, the binding of the ligand (ABA

    or pyrabactin) within the cavity is not itselfsufficient to trigger the downstream path-

    way. These structural comparisons revealed

    that the formation of a productive signaling

    complex also depends on the ability of the

     bound agonist to maintain the gate in the

    closed conformation. The maintenance of a

    closed gate and latch configuration is need-

    ed to create a binding surface to tether and

    inhibit the clade A PP2Cs.

    Comparisons of the hormone-bound

    receptor (ABA-PYL1) (44) to that of the

    tertiary structures ABA-PYL1-ABI1 and

    ABA-PYL2-HAB1 (48, 51, 53) revealed no

    structural difference in the receptor moietyof the complex, suggesting that the recep-

    tor is a rigid structure. An important insight

    informed by the tertiary complex is that

    amino acid Ser 112 of PYL1, or the equivalent

    Ser 89 of PYL2, make contact with Glu142 and

    Gly180 of ABI1 (Fig. 3). Thus, the Ser 112 or

    Ser 89  of the receptor functions like a plug

    and physically obstructs the entrance to the

    catalytic site of the phosphatase. In light

    of the data from structural (53), in vitro,

    and yeast two-hybrid analyses (32, 41), the

    G180D mutant abi1-1  and the analogous

    G168D abi2-1, with their glycine residues

    replaced by the bulkier and charged asparticacid, would permit escape from repression

     by receptor binding during ABA signaling,

    which would explain the dominant or con-

    stitutive nature of these mutations. Again,

    the structural data with these soluble recep-

    tors and the PP2Cs fit those from the genetic

    analysis of the mutants and molecular prop-

    erties of their proteins.

    The dissociation constants ( K d ) of four

    representative receptors (expressed as re-

    combinant proteins in bacteria) are unex-

     pectedly high, near or in the micromolar

    range. However, their affinity for ABA in-

    creases to the nanomolar range when com- patible PP2Cs are present (Table 1). This

    increase in receptor affinity for ABA in the

     presence of PP2C was also reflected by in

    vitro assays of the inhibition of protein phos-

     phatase activity by the receptors, which was

    sensitive to the ratios of the two components

    as well as the particular homolog of the clade

    A PP2C (Table 2). The efficiency of ABA-

    mediated inhibition of phosphatase activity

    in vitro was generally higher for ABI1 than

    for ABI2, and in terms of the receptors, ABA

    was more effective with RCAR3 than with

    RCAR1. For example, at the ratio of one

    PP2C to four receptor molecules, the me-

    dian inhibitory concentration (IC50) of either

    ABI1 or ABI2 by RCAR3 was between 15

    to 40 nM ABA; in comparison, RCAR1/ABI2 revealed a two- to threefold higher

    IC50 value of roughly 60 to 95 nM ABA (Ta-

     ble 2) (47 , 54). These observations suggest

    that the combination of particular RCARs

    and PP2Cs behaves as a coreceptor complex

    for ABA (although PP2Cs are not widely

    known to bind ABA, as would be expected

     by a classical coreceptor) and that together,

    different receptor-PP2C combinations might

    activate the drought adaptive response path-

    ways differently (54). The mechanistic basis

    of this enhanced ABA sensitivity displayed

     by the receptors in the presence of PP2Cs is

    not obvious, because ABA is cloistered deepwithin the cavity of the receptor. However,

    Trp300 of ABI1 (or Trp385 of HAB1), some-

    times referred to as the Trp lock (42), plays

    a unique structural role in the receptor-PP2C

    complex. It is the only amino acid residue in

    the phosphatase that bridges indirectly with

    the ABA molecule and the receptor simul-

    taneously through a combination of water-

    mediated and hydrophobic interactions,

    respectively (Fig. 3). Mutational analysis

    showed that this tryptophan is not essential

    for phosphatase activity, but only its affin-

    ity with the receptor and, as a consequence,

    ABA-dependent inhibition of ABI1 (53) orHAB1 (55) is affected when this residue

    is mutated. Conformational changes in the

    receptor induced by its interaction with this

    key tryptophan residue facilitate the fasten-

    ing of the receptor’s gate and latch into the

    closed configuration. Whether this could

     provide a structural rationale for the more

    than 10-fold increase in ABA binding affini-

    ty observed for the PYL-PP2C complexes as

    compared with the apo-receptors still needs

    to be confirmed (39, 41, 44, 51, 53, 54).

    Several research groups have indepen-

    dently, and by different experimental ap-

     proaches, identified an ABA-activated andcalcium-independent kinase in wheat (56 ),

    the broad bean V. faba (57 , 58), and its or-

    tholog in  Arabidopsis  (59, 60) that acts as

    a positive regulator of this stomatal closing

     pathway. It is this kinase that is muted by the

    PP2Cs when the ABA signaling pathway

    is off. The ABA-ACTIVATED PROTEIN

    KINASE was purified biochemically from

    V. faba  guard cells. When a catalytically

    dead variant of the kinase was expressed

    transiently in wild-type (WT) Vicia  guard

    cells, ABA-mediated activation of anion

    channels required to close stomates was

    abolished (58). Likewise, mutations in the

    homologous kinase SnRK2 in  Arabidop-

     sis, known variously as OPEN STOMATA

    1 (OST1), SRK2E, and SnRK2.6 (59, 60),also blocked the typical stomatal closing re-

    sponse to ABA and to progressive drought.

    Yoshida et al . demonstrated a direct interac-

    tion between ABI1 and OST1 by the yeast

    two-hybrid approach. Further, they delineat-

    ed a small amino acid motif, called domain

    II, at the noncatalytic C terminus of OST1

    as the direct docking site for ABI1 (61).

    This domain II is also found in the C termini

    of SnRK2.2/2D and SnRK2.3/2I, two other

    closely related ABA-activated SnRK2s in

    the same clade as OST1 (60, 62). Of the 10

    members in the entire family, these three

    SnRK2s seem to regulate all of the knownelementary ABA responses (63 – 65). Sev-

    eral serines within the activation loop of

    OST1 become phosphorylated in vivo in

    response to ABA (66 ). In vitro, ABI1 and its

    mutant counterpart abi1-1 dephosphoryl-

    ated the ABA-stimulated OST1 recovered

    from cell extracts (65, 66 ). Also, relative to

    WT plants, the ABA-activated kinase ac-

    tivity from plant extracts was lower in the

    dominant gain-of-function PP2C mutants

    (for example, abi1-1); conversely, it was

    higher in the PP2C loss-of-function mutants

    (65, 66 ), consistent with the notion that in

    vivo the three kinases are negatively regu-lated by these PP2Cs. Finally, OST1 (66 ),

    SnRK2.2, and SnRK2.3, along with 9 of the

    14 members of the soluble receptors (67 ),

    coimmunoprecipitate with ABI1 in  Arabi-

    dopsis protein extracts. The composition of

    the copurified proteins did not change re-

    gardless of whether or not exogneous ABA

    was added. This suggests that at least ABI1,

    the three ABA-activated SnRK2s, and at

    least nine members of the soluble receptors

    might be stable constituents of a core ABA

    signalosome (67 , 68). Because whole plants

    were used in the coimmunoprecipitation

    studies, all components may not be part ofthe same signalosome simultaneously, but

    different combinations of the constituents

    could exist in different tissues.

    Regulation of Ion Transport Acrossthe Plasma Membrane by the ABACore Signaling ComplexRelative to its closest homologs SnRK2.2 

    and SnRK2.3, mutations in the OST1 locus

    have the most negative impact on guard cell

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    6/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 5

    response to environmental stress (59, 60,

    69, 70). OST1 may thus be the key guard

    cell kinase regulating a large roster of tar-

    gets. A substantial fraction of the transcrip-

    tome responsive to ABA is regulated by

     b-ZIP transcription factors (71), which are

     potentially activated by OST1 (72, 73). The

    minimal pathway activated by ABA, which

     presumably leads to altered gene expression

    in vivo, has been reconsituted in vitro. The

    components consist of PYR1, ABI1, and

    OST1, which produced the ABA-dependent

     phosphorylation of the b-ZIP transcription

    factor ABA RESPONSIVE ELEMENT

    BINDING FACTOR 2 (ABF2) (more cor-

    Fig. 3. The mechanistic basis of ABA-mediated inhibition of PP2Cactivity. (A) Ribbon models of superimposed apo-PYL2 (gray) andABA-bound PYL2 (green) [generated using Protein Data Bank(PDB) accession codes 3KAZ and 3KBO (44 )]. The orientation of

    ABA (ball model) in the ligand binding pocket is shown. Magenta,apo-latch; blue, ABA-bound latch; pink, apo-gate; yellow, ABA-boundgate. (B) The PYL1-ABA-ABI1 tertiary complex [modified with per-mission from (53 )]. The Trp lock of ABI1 is shown as yellow spacefill.(Right) Close-up view of the intermolecular interactions that explainreceptor sequestration of PP2C upon ABA binding. (C) A general-ized scheme of the receptor-ABA-PP2C complex highlighting the es-sential serine residue (Ser112 in PYL1) in the receptor that tethers thePP2Cs by interacting with a Glu (Glu142 in ABI1) and a Gly residue(Gly180 in ABI1). A conserved Trp in the PP2Cs (Trp300 in ABI1) inter-acts with the ABA through a water molecule (blue dot). The carboxylgroup of ABA is in contact with a Lys residue (Lys86 in PYL1) deep in

    the pocket. N, N terminus. (D) In PYL1, amino acids participating inABA binding are underlined in black (86, 171, 116 to 121, and 143 to149). The START homology spans from amino acids 50 to 206. Reddots denote amino acids that dock onto the catalytic regions of ABI1.

    Residues underlined in blue denote α-helical structures (aminoacids 34 to 47, 69 to 77, 82 to 84, and 183 to 208); orange underlinesindicate β strands (57 to 67, 89 to 93, 105 to 110, 117 to 122, 135to 137, and 148 to 175); and the red underline denotes the helicalturn at amino acids 128 to 131 (127 ). Various mutant alleles of thePYR1 locus have been transposed onto the equivalent amino acidsof PYL1. (E) ABI1; blue residues represent the catalytic region. Yel-low dots mark the amino acids that, when mutated, decreased PYL1binding. Orange overlines indicate E142, G180 (abi1-1), and W300(Trp lock), which define the entrance to the ABI1 catalytic center inthe crystal structure. Mutations around W300 show an eight–aminoacid insert specific to plant PP2Cs (42 ).

    ABA

     

         

    PYL1

    Gate pyr1-9 pyr1-3  

         

    Latch pyr1-6    pyr1-5 pyr1-8 

          

     

    1-4 pyr1-2 

    ABI1

    ABI1

    PYL1ABA receptor

    Clade A PP2C

     

     

       

    Apo-latch

    Apo-gate

    Latch

    Latch

    Recoil

    Recoil

    ABA-bound

    gate

    ABA

    Glu142

    Gly180

    Ser112

    gate

     Trp300

    GluGly

     Trp

    Catalytic

    center

    Lys

    SerGateLatch

    N

    N

       C   R   E   D   I   T   S  :   (   A   )   Y .   H   A   M   M   O   N   D   /   S   C   I   E   N   C   E   S   I   G   N   A   L   I   N   G  ;   (   B   )   M   O   D   I   F   I   E   D   W   I   T   H   P   E   R   M   I   S   S   I   O   N   F   R   O   M   N   A   T   U   R   E        4        6        2 ,   6   0   9  –   6   1   4   (   2   0   0   9   )

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    7/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 6

    rectly, a ~80–amino acid protein fragment)

    to represent the transcriptional output of

    this minimal pathway (68).OST1 also modifies membrane trans-

     port properties in the guard cells by phos-

     phorylating and inactivating one of the ma-

     jor potassium inward-rectifying channels,

    KAT1, which was shown in vitro and in

    the  Xenopus  oocyte heterologous expres-

    sion system (Fig. 4) (74). The ensuing re-

    duction in K +  influx is consistent with the

    known electrophysiological effect of ABA

    on modifying guard cell membrane trans-

     port (Fig. 1). ABA stimulates the produc-

    tion of H2O2 (Fig. 1) through plasma mem-

     brane–localized NADPH oxidases, which

    are encoded by 10 genes in the Arabidopsis genome. Although H2O2 is similar in struc-

    ture and chemical properties to H2O (75),

    unlike water, it is a powerful oxidant, or

    ROS. During signaling, the increases in

    concentration of H2O2 must be maintained

    above a certain threshold (estimated to be

     between 10 to 100 µM)—long enough tooxidize its effector molecules, but not so

    long as to cause cellular damage. Hence, to

    curb rampant H2O2 cytotoxicity, the activi-

    ties of the NADPH oxidases are thought to

     be tightly regulated by numerous cytosolic

    factors that include Ca2+, protein kinases,

    and small guanosine triphosphatases (76 ).Kwak et al . (77 ) identified mutations in two

    genes encoding NADPH oxidase catalytic

    subunits,  AtrbohD  and  AtrbohF   (respira-

    tory burst oxidase homolog D and F), that

    abolished ABA-induced stomatal closure,

    ABA-mediated promotion of ROS produc-

    tion, and ABA-induced increase in cyto-

    solic Ca2+. In vitro, OST1 phosphorylates

    AtrbohF, but not AtrbohD (Fig. 4) (78),

    which is also consistent with the lack of

    ABA-mediated ROS production in the ost1 

    mutant guard cell (60).

    The target of OST1 that has attracted mostof the attention is involved in the regulation

    of anion efflux critical for ABA-mediated

    stomatal closure. Anion efflux is controlled

     by a balance between phosphorylation and

    dephosphorylation events (79). Two groups

    independently converged on the locus

    SLOW ANION CHANNEL–ASSOCIATED

    1  (SLAC1) as the gene encoding the most

    likely S-type anion channel involved in sto-

    matal closure (80, 81) (Fig. 4). Both groups

    used impaired stomatal closure in response

    to either high CO2  (80) or hypersensitiv-

    ity to damage of photosynthetic tissues by

    ozone (O3) (81) as the phenotypic criterionfor the mutant screen. The putative SLAC1

     protein is a distant relative of bacterial and

    fungal C4-dicarboxylate transporters and a

    weak homolog (20% amino acid identity)

    of Mae1 of Schizosaccharomyces pombe,

    which has been functionally characterized

    as a malate uptake transporter. Guard cell

     protoplasts derived from slac1-2, compared

    with those from the WT plants, contain

    higher amounts of organic anions, notably

    malate and fumarate, and the inorganic ions

    Cl –   and K +, possibly as an indirect conse-

    quence of perturbed ionic homeostasis (80).

    There are three related SLAC1 homologs(SLAH) in the  Arabidopsis  genome. Re-

    verse transcriptase polymerase chain reac-

    tion (RT-PCR) coupled to histochemical

    analysis of the β-glucuronidase (GUS ) re- porter under the control of the individual

    SLAH   promoters revealed that all of them

    are expressed in various tissues besides

    guard cells in plants grown under the specif-

    ic conditions used for this analysis (80). De-

    spite the differences in their tissue-specific

    expression patterns, at least two of these are

    functionally interchangeable in guard cells,

     because ectopic expression of either SLAH1 

    and SLAH3 under the control of the SLAC1 

    guard cell–specific promoter complements

    the phenotypes of CO2 insensitivity and ac-

    cumulation of organic and inorganic ions of slac1-2 (80). The slac1 mutant is phenotypi-

    cally pleiotropic: The guard cells are only

    moderately sensitive to light and humidity,

    and they exhibit a pronounced indifference

    to ABA, NO, O3, and H2O2.

    These phenotypes associated with slac1 

    mutant plants and the rescue by SLAH1 or

    SLAH3 are important for several reasons.

    The first is that they provide genetic evi-

    dence that the CO2, O3, low humidity, and

    ABA perception pathways are interconnect-

    ed and that SLAC1 has an important role in

    integrating these environmental and endog-

    enous signals. Two of the transferred DNAinsertion mutant alleles of  slac1  that were

    studied by Saji et al . (82), which they called

    ozs, are identical to slac1-3 and slac1-4. The

    ozs  mutants were initially isolated on the

     basis of the appearance of necrotic lesions

    on leaves when exposed to O3  (~200 parts

     per billion or 0.2µl/l). At the stomatal level,however, the responses of the ozs  mutant

     plants to ABA, CO2, H2O2, and O3 measured

     by Saji et al . were indistinguishable from

    those of WT plants (82). The reasons for the

    discrepant observations are unknown.

    A second reason that the phenotypes of

    the  slac1  plants are biologically importantis that, in contrast to WT plants, the char-

    acteristic slow and sustained anion current,

    which is weakly voltage-dependent, is bare-

    ly detectable in the guard cell protoplasts

    derived from the  slac1  mutant. Only very

    weak background whole-cell membrane

    currents and patch-clamp seal currents were

    observed (81). Thus, SLAC1 most likely

    corresponds to the long-sought-after anion

    channel in the guard cell that is crucial for

    ABA-mediated stomatal closure.

    A final reason that the study of the slac1 

     plants was particularly important was that

    it provided strong evidence that the R-typeand the S-type anion currents are produced

     by different channel proteins. The R-type

    anion current, which is transient rather than

    sustained, is not affected in the  slac1  mu-

    tants (81), implying that the S- and R-type

    anion channels are unlikely due to post-

    translational modification of a single poly-

     peptide, which is consistent with previous

    suggestion based on physiological studies

    (12, 83). AtALMT12, a homolog of the

    Table 1. Dissociation constants of representative receptors in the presence and absenceof PP2Cs. The dissociation constants (extrapolation of ligand affinity to achieve half oc-cupancy of the receptor sites) were obtained by using receptors produced in Escherichiacoli  and calculated by isothermal titration calorimetry (ITC) or surface plasmon resonance(SPR). In the presence of PP2Cs (ABI1, ABI2, HAB1), the four PYR/PYL/RCAR recep-tors display substantially higher hormone affinity. N/A, not applicable.

    Receptors K d (µM) K d in the presence ofPP2C (nM)

    Techniques for K d measurements

    PYL9/RCAR1 0.7 64 (ABI2) ITC (41)

    PYL5/RCAR8 1.1 38 (HAB1) ITC (39 )

    PYR1/RCAR11 N/A 125 (0.8 HAB1)* ITC (32 )

    PYL8/RCAR3 1.0 18 (0.25 ABI1)* ITC (54 )

    PYL1/RCAR12 52 and 340 N/A ITC, SPR (53 )

    *Values estimated from IC50 using the indicated relative ratios of PP2C to the receptor.

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    8/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 7

    aluminum-activated malate trans- porter, fulfills the physiochemicalcharacteristics of the R-type an-ion channel (84, 85). These chan-nels were first studied in the rootsand are thought to play a role in

    releasing malate to chelate alumi-num in the rhizosphere (86 , 87 ).In mutants lacking the functionalAtALMT12, guard cells displayedreduced sensitivity to closingstimuli, such as the transition oflight to darkness, high concen-trations of CO2, and ABA. Thecharacteristic voltage-dependentR-type current was reduced by~40% in the mutant protoplastscompared with those preparedfrom WT plants. The gating of theR-channel is sensitive to malate.

    This malate-sensitivity was alsoobserved for AtALMT12 when itwas heterologously expressed andcharacterized in oocytes (85), sug-gesting that this property may beinherent to the transporter itself,unless  Xenopus  has conservedthe same regulatory machinery. Itis currently not known how ABAregulates AtALMT12.

    Guided by the structural stud-ies on the anion channel TehA, theSLAC1 homolog from the bac-terium  Haemophilus influenzae,

    SLAC1 would be a symmetricaltrimer composed of quasisym-metrical subunits, each having 10transmembrane helices arranged in pairsto form a central five-helix transmembrane

     pore (88). The pore is a relatively uniform passage of 5 Å in diameter lined with largelyhydrophobic residues, except for a constric-tion that is gated by an conserved phenylala-nine residue (Phe450). SLAC1 does not seemto have discrete anion binding sites in thechannel, compared, for example, with thoseof the CLC family of channels, which havediscrete ion binding sites with high field

    strength. The ion selectivity of SLAC1 isthought to be largely a function of the ener-getic cost of ion dehydration and thus repre-sents a unique pore structure for anion chan-nels. Despite the overall hydrophobicity ofthe ion-conducting tube, the electrostatic po-tential on the pore surface is polarized, andin particular, the electropositive nature of itscytoplasmic surface is thought to contributeto anion efflux.

    When the channel is heterologously ex-

     pressed in the  Xenopus oocyte system, theactivity of SLAC1 was detected only whencoexpressed with any of the following threekinases: OST1 (89), CPK23, or CPK21(90). The major site of phosphorylation byOST1 is Ser 120  in the N-terminal cytosolicdomain of SLAC1 (4, 89, 91). This N-ter-minal cytosolic domain of SLAC1 is phos-

     phorylated by the CPKs at other unspecifiedmotifs. There are several other SLAC1 sitesthat are phosphorylated in vitro by OST1,

    and whether they have in vivo relevanceis not clear (4). In  Xenopus, the activatedSLAC1 displays higher permeability to

     NO3 –   compared with Cl –   and malate (89).

    The mutant slac1 can be complemented bythe ectopic expression of either SLAH1  orSLAH3  driven by the guard cell–specific

     promoter of SLAC1. However, SLAH1 doesnot contain any extended N- or C-terminalcytosolic domains that could constitutethe targets of phosphorylation. Thus, the

    molecular consequences of phosphorylation on the overallSLAC1 structure are not im-mediately obvious.

    The protein phosphatasesABI1, ABI2 (89, 90), and

    PP2CA (91) block SLAC1-mediated anion efflux in the Xenopus  expression system.The other homologous pro-tein phosphatases, such asHAB1 and HAB2, were lesseffective (90), at least in the

     Xenopus  system. Neither theWT catalytic activity of the ki-nase OST1 (89) nor that of the

     phosphatase AtPP2CA (91)is required for these proteinsto interact. An inactive formof AtPP2CA blocked phos-

     phorylation of SLAC1 by WTOST1 (91), indicating that theactivity of OST1 can be inhib-ited by physical entrapmentin addition to dephosphoryla-tion by the PP2Cs. ABI1 andABI2 interact with CPK21and CPK23, which was detect-ed with bimolecular fluores-cence complementation whentagged forms of these partnerkinase-phosphatase pairs werecoexpressed in the  Xenopus oocytes (90) or in mesophyll

    cells (92). Although CPK23can phosphorylate SLAC1 inheterologous systems, such

    as  Xenopus  oocytes, and although the an-ion current is reduced (by 70%) in guardcell protoplasts derived from the knockoutcpk23, its functional importance, if any, in

     planta is not clear (90). Despite the reducedanion current, no stoma phenotype wasnoted in the cpk23 knockout mutant in thesestudies (90). The opposite phenotypes ofreduced and increased stomatal apertures,respectively, were observed by others in theknockout mutant and in plants overexpress-

    ing  AtCPK23  (93). CPK21 functions as anegative regulator of abiotic stress responses

     because the cpk21 knockout is more toler-ant, rather than having the expected height-ened sensitivity, to prolonged osmotic stressas compared with WT plants (94). Theseapparently contradictory results suggest thatin the guard cell, there may be other targetsof these CPKs missing in the oocyte assaysin which only single targets were tested.There are also two other calcium-dependent

    Table 2. Receptor affinity to ABA depends on the presence of proteinphosphatases and their relative ratios. The IC50  values indicate theconcentration of ABA required to cause 50% inhibition of the PP2Cactivity by the receptor. The ratio of PP2C to receptor has a pro-nounced impact on IC50, which is commensurate with the amount ofinput PP2C. Because of this, the PP2Cs have been regarded by someas coreceptors of ABA. Note that PYR1 with 0.6 HAB1 yielded an IC50 

    of 390 nM, whereas a value of 125 nM was obtained in Table 1 using0.8 HAB1. These variable results for the same combination of PP2Cand receptors could be due to the different experimental conditions.

    Receptors Ratio PP2C:receptor IC50 (nM) Reference

    PYL9/RCAR1 0.25 ABI1

      0.50 ABI2

      0.25 ABI2

    35

      95

      60

    (54 )

      (54 )

      (54 )

    PYL5/RCAR8 0.60 HAB1

      2.00 ABI2

      2.00 ABI1

    35

      115

      123

    (39 )

      (39 )

      (39 )

    PYR1/RCAR11 0.60 HAB1  0.60 ABI2

      2.00 ABI1

    390*  360

      330

    (39 

    )  (39 )

      (39 )

    PYL8/RCAR3 0.25 ABI1

      0.50 ABI1

      2.00 ABI1

      0.25 ABI2

      2.00 ABI2

      0.6 HAB1

    18*

      23

      75

      30

      118

      135

    (54 )

      (54 )

      (39 )

      (54 )

      (39 )

      (39 )

    PYL4/RCAR10 2.00 ABI1 272 (39 )

      2.00 ABI2 110 (39 )

      0.60 HAB1 188 (39 )

    *Values can be compared to those in Table 1.

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    9/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 8

    kinases, CPK3 and CPK6, which have been

    implicated in the regulation of anion chan-

    nels; however, it is not known whether these

    kinases directly phosphorylate SLAC1 aswell, or whether they modify anion channel

    activity through an indirect effect (95).

    Slow anion conductance, particularly

     permeability to NO3 – , is not completely

    abolished in the  slac1 mutant. For example,

    the  slac1 mutant can still close the stomates

    in response to the transition from light to

    dark (80, 81), which requires anion efflux.

    This fueled the motivation to trawl the  Ara-

    bidopsis  genome for other anion channels

     behind the residual activities in the guard

    cell. Ectopically expressed SLAH3  can

    functionally restore the mutant phenotypes

    of  slac1-2, but unlike the original reporton the tissue-specific expression pattern of

    the SLAC   family (80), later investigations

    revealed that SLAH3  itself is expressed in

    the guard cell at ~50% of the amount of

    SLAC1, as estimated by quantitative PCR

    (92). In addition, the abundance of the

    SLAH3  transcript in guard cell protoplasts

    increased ~twofold in the  slac1 background

    compared with its abundance in WT plants,

    suggesting that these two anion channels

    can compensate for each other by feedback.

    Patch-clamp measurements of  slah3  guard

    cells revealed reduced current in nitrate-

     based medium, suggesting that SLAH3 isthe likely channel responsible for the re-

    sidual anion activities in  slac1; however, the

     slah3 mutant has no growth phenotype (92).

    The SLAH3 activity was also slowly deac-

    tivated by negative membrane potentials,

    reminiscent of the characteristics of the

    S-type anion channel. Like SLAC1, SLAH3

    was also phosphorylated at the N-termi-

    nal cytosolic segment by CPK21, which

    was blocked in the  Xenopus  experimental

    Fig. 4. Current model of the ABA signaling pathway in the guard cell. Inthe absence of ABA, the activities of the three kinases CPK21, CPK23,and OST1 are muted by the upstream PP2Cs (ABI1, ABI2). Light acti-vates H+-ATPases (for example, OST2), which in turn drive secondarytransporters, such as K+ influx channels (probably consisting of KAT1, aheterotetrameric complex, with its closest homolog KAT2). The bindingof ABA to the receptor leads to retention of the PP2Cs, thereby liberat-ing the kinases to phosphorylate the downstream targets. The OST1phosphorylates and inhibits the inward-rectifying K+ channels to pre-vent entry of K+ into the guard cell necessary for stomatal opening (A).This same kinase, however, phosphorylates and activates the NADPH

    oxidase AtrbohF to generate the second messenger H2O2, which islinked to Ca2+ release. OST1 and CPK21 or CPK23 phosphorylate andactivate the S-type anion channel SLAC1. OST1 also integrates theCO2 stimulus, but the intermediates (marked as ?) in this pathway havenot been determined. CPK21, but not OST1, phosphorylates and ac-tivates SLAH3 in Xenpous  oocytes. Ca2+  inhibits the proton pumpingactivity (for example, OST2), probably through the action of anotherCa2+-dependent kinase PKS5 (128 ). GORK is the major K+-outwardrectifying channel that is sensitive to cytosolic alkalinization and expelsK+ needed for stomatal closing (B) (97 ). Upward arrows denote stimula-tion of activity; downward arrows indicate inhibition of activity.

    CDPK SnRK2SnRK2 CDPK  CDPK CDPK CDPK  

    (CPK23?) (OST1) (CPK23?) (CPK21?)(OST1)

    Ca2+

    Stomata open Stomata closed

    SLAH3

    (Anion channel opens

    anions exit cells)

    KAT1 and

    KAT2

    KAT1

    and KAT2

     (K + influx

    channel closed)

     (K + channel

    open)

    OST2OST2

    (H+-ATPase

    inactive)

    SLAC1

    (Anion channel opens

    anions exit cells)

    ?

    ? ?

     

    GORK1

    (K + efflux

    channel opens)

    (PKS5)

    O COOHOH

    (CPK21?)

    K + H+

    CO2

    (AtrbohF)

    Cytosol

    H2O

    2

    Anions

    (H+-ATPase,

    H+ ions exported)

       C   R   E   D   I   T  :   Y .   H   A   M   M   O   N   D   /   S   C   I   E   N   C   E   S   I   G   N   A   L   I   N   G

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    10/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 9

    system by coexpression of ABI1 and ABI2.

    However, there are also some important dif-

    ferences between these homologous anion

    channels. Unlike SLAC1, SLAH3 is not

     phosphorylated by OST1. Compared with

    SLAC1, SLAH3 is twice as permeable to

     NO3 – 

    , and this anion has been proposed to be a physiological activator of this channel

    (92). In contrast, SLAC1 might be activated

     by bicarbonate (96 ), which is blocked in the

    knockout mutant ost1. Whether these appar-

    ent differences are physiologically relevant

    or the consequences of different experi-

    mental approaches needs more exploration.

    The fact that the activities of SLAC1 and

    SLAH3 are regulated by OST1 or CPKs or

     both is consistent with the Ca2+-dependent

    and -independent nature of the anion ef-

    flux critical for stomatal closure. However,

    questions remain concerning how these

    CPKs fit into the early steps of ABA sig-naling in planta. Like the ABA-dependent

    transcriptional pathway (68), the posttrans-

    lational regulation of SLAH3 that presum-

    ably contributes to stomatal closure has also

     been reconstructed in vitro (92). Binding of

    ABA to the receptor RCAR1 (also known

    as PYL9) blocks ABI1 phosphatase activity,

    freeing CPK21 to phosphorylate SLAH3

    (more correctly, its N-terminal cytosolic

    domain, which was used in the assay to

    represent the physiological endpoint) (Fig.

    4) (92). In parallel, it is expected that the

    depolarization of the plasma membrane

    evoked by SLAC1 and SLAH3 would leadto cytosolic alkalinization and activation

     by a pH-sensitive pathway activating the

    outward-rectifying K +  current, which has

     been identified as GORK (97 ) (Fig. 4). K + 

    efflux, therefore, has a twin role: to restore

    the charge unbalance due to expulsion of

    the anions by SLAC1 and SLAH3, and to

    relieve guard cell turgor pressure requisite

    for stomatal closure.

    OST1, an Integrator of ABA andCO2 SignalsThe work on the functional relation between

    SLAC1 and OST1 in guard cell signalinghas parlayed into fresh insight into how

    ABA signaling is integrated with the re-

    sponse to CO2, the other signal besides H2O

    with direct relevance to accumulation of

     biomass and climate change (96 ). Plants re-

    spond to increased CO2 [800 parts per mil-

    lion (ppm), compared with ambient CO2 of

    ~350 ppm] by closing the stomates, which

    requires carbonic anhydrase activity (98),

     but the early signaling events have not been

    entirely clear (99). CO2  is thought to dif-

    fuse passively across the plasma membrane

    during photosynthesis, but pharmacologi-

    cal studies and reverse genetic studies have

    suggested that certain aquaporins present in

    the plasma membrane and chloroplast enve-

    lope might actively transport CO2, at leastin the mesophyll of tobacco (100 – 102). The

    identification of the  slac1  mutant brings a

    genetic proof that the guard cell itself is

    equipped with CO2 sensors and signal trans-

    duction pathways. In fact, increased CO2 

    activates anion currents in the guard cells

    (98). High and low bicarbonate concentra-

    tions that promoted either stomatal closing

    and opening, respectively, had also been

    observed decades ago (103). Guard cells of

    the  slac1  mutant are insensitive to several

    environmental stimuli, including changes

    in CO2  concentration (4, 80), and guard

    cell–derived protoplasts of  slac1  plants donot produce anion currents when exposed

    to high concentrations of CO2 and carbonic

    acid (CO2/HCO3 – ; a mixture of CO2 and car-

     bonic acid was used in these experiments);

    HCO3 –  is condensed from CO2 and water, a

    reaction catalyzed by the CO2-binding pro-

    teins carbonic anhydrases (96 ). These stud-

    ies also suggested that HCO3 – , more than

    CO2, might be the intracellular activator

    of anion channels (96 ). It appears that one

    of the consequences of CO2/HCO3 –   is the

     priming or enhancement of the Ca2+  sensi-

    tivity of SLAC1. SLAC1 is phosphorylated

     by OST1, and mutant ost1 plants exhibited anormal stomatal opening response to low at-

    mospheric CO2 (60). Thus, it was surprising

    to find that, compared with WT guard cell

     protoplasts, the anion currents from those

    of ost1 were also not triggered by increased

    CO2/HCO3 – , and stomatal closing was im-

     paired. In contrast, the stomatal opening

    response to bicarbonate was normal, albeit

    slower, in the quadruple mutant for the ABA

    receptors  pyr1,  pyl1,  pyl2, and   pyl4  (96 ).

    Together, these results suggest that OST1 is

    a convergent point for both ABA and CO2 in

    the stomatal closure pathways.

    Concluding Remarks andFuture ProspectsWe have come a long way since the first

     biochemical identification of ABA-bind-

    ing proteins in the plasmalemma of the V.

     faba  guard cell (28). The discovery of the

    cytosolic ABA receptors, characterized by

    the presence of the START domain, has

    led to elucidation of the early steps in the

    ABA signaling pathway. The accessibility

    of ABA to this family of inside receptors

    is probably partially modulated by ATP-

     binding cassette transporters (104, 105),

    which is reminiscent of the carrier-medi-

    ated ABA uptake reported for Commelina 

    (16 , 17 ). A parsimonious ABA signaling

     pathway, as defined by reconstitution invitro, is composed of a soluble ABA re-

    ceptor (PYR1), a PP2C (ABI1), a SnRK2

    (OST1), and a transcription activator b-ZIP

    (ABF2) that binds ABA-regulated promot-

    ers (68). Phosphorylation of a fragment of

    ABF2 by OST1 in this in vitro reconstitu-

    tion was shown to be ABA-dependent (68).

    Likewise, a similar minimal pathway was

    reconstructed for the ABA response in the

    guard cell. Because SLAH3 is functionally

    equivalent to SLAC1 and the  slac1 mutant

    stomata are insensitive to a battery of clos-

    ing signals (80), the prototypical members

    in the stomatal-closing pathway consist ofRCAR1 (PYL9) (the cytosolic ABA recep-

    tor), AB11 and ABI2 (the negative regulato-

    ry PP2Cs), OST1 and possibly CPK23 and

    CPK21 (the positively regulating kinases),

    and SLAC1 and SLAH3 (as the S-type an-

    ion channels initiating the depolarization of

    the plasma membrane prerequisite to sto-

    matal closure) (92). Because the receptors,

    PP2Cs, CPKs, and SnRK2s are all encoded

     by large gene families, tremendous combi-

    natorial possibilities are possible, enabling

     plants to finely modulate the intensities of

    the output. Just one example of this inher-

    ent flexibility is the combination of apo-receptor and PP2C, which affects the bind-

    ing constants to ABA (39, 54) (Table 1).

    Immunoprecipitation of ABI1 (tagged

    with yellow fluorescent protein) from plant

    extracts recovered a large number of solu-

     ble receptors (9 of 14) (67 ). Other proteins

    that coprecipitated with ABI1 included the

    REGULATORY PARTICLE NONATPASE

    10 (RPN10), a subunit of the 19S   regula-

    tory complex of the 26S  proteasome (106 ).

    Whether this indicates that some of these

    signaling components might be regulated

     by protein stability in the guard cell is not

    known (67 ). However, in germinating seeds,the b-ZIP transcription regulator ABI5 ac-

    cumulates in the mutant rpn10  (106 , 107 ).

    Additional proteins that coprecipitated with

    ABI1 included the sucrose-phosphate syn-

    thase 1F and the ribosomal protein PRL12B.

    The H+-ATPase, OST2 (108), whose activ-

    ity is suppressed by ABA before stomatal

    closing, was also recovered along with the

    receptors in the ABI1 immunoprecipita-

    tions. AHA2, which shares overlapping

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    11/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 10

    functions with OST2 (109), was not recov-

    ered, however. The proton pumps are usu-

    ally considered to be the endpoints of sig-

    naling pathways; thus, the direct association

    of OST2 with the ABA receptor complex

    hints at the possibility that the pathway is

    much more complex in composition in the plant context. How the functions of these

    other “accessory” proteins are integrated

    into the “core” ABA signaling components

    established in vitro remains to be investi-

    gated. Besides PYR/PYL/RCAR, there are

    also membrane-associated candidate re-

    ceptors—GTG1, GTG2 (110), and ABAR

    (111, 112)—although some have been un-

    able to reproduce the binding of ABA to

    ABAR (113).

    It is still too early to pronounce wheth-

    er GTG1 and GTG2 might fit the profile

    of the outside ABA perception site. Also,

    where or how the GTG1 and GTG2 relateto the PYR/PYL/RCAR-mediated pathway

    or whether they represent part of parallel

    and independent signaling cascades remain

    fascinating questions. Structural studies

    of PYR1 bound to both S-(+)-ABA and

    R-(–)-ABA stereoisomers showed that the

    differences in the chirality of both isomers

    are accommodated within the soluble re-

    ceptors by the rotation of the ABA ring by

    ~180°  (50). Neither ABAR nor GTG1 or

    GTG2 can bind the nonnatural R-(–)-ABA

    isomer (110, 114); Nonetheless, the R-(–)-

    ABA isomer induces long-term responses,

    such as seed germination (115). SLAC1is phosphorylated by OST1 and the Ca2+-

    dependent CPK21, at least when assayed

    in the  Xenopus  oocyte system. Reverse

    genetics and electrophysiological studies

    have also implicated CPK3 and CPK6 in

    the regulation of the S-type anion efflux

    in response to ABA (95). The precise rela-

    tion between these two Ca2+-dependent ki-

    nases and S-type anion transporters is not

    yet clear, but does reinforce the importance

    of Ca2+ in “priming” or accentuating the re-

    sponse to ABA (43). The calcium-binding

     protein NpSCS exerts a suppressing effect

    on all SnRK2s tested in vitro, and this in-hibition is calcium-dependent (116 ). If so,

    this suggests that SLAC1 may be regulated

     by two mutually exclusive phosphorylation

     pathways in guard cells.

    ABA also seems to play developmen-

    tal roles other than in stress signaling and

    drought protection. Studies carried out in

    tomato and  Arabidopsis  suggest that ABA

    is required to limit ethylene production

    during the course of normal plant growth

    (117 , 118). The dose-dependent effect of

    ABA is evident in roots, where elongation

    in  Arabidopsis  is stimulated by exogenous

    ABA at 0.1 µM and is inhibited when the

    hormone is applied at concentrations above

    1.0 µM (119). Suppressing ABA production

     by mutations or in transgenic plants resultsin developmental defects, such as altered

    organization of the mesophyll and disrupted

    stomatal morphogenesis (120, 121). Plants

    in unstressed conditions contain a function-

    ally relevant basal amount of ABA. Careful

    liquid chromatography–mass spectrometry

    measurements of ABA content in 4-week-

    old Arabidopsis seedlings detected between

    10 to 40 nM of the hormone (122). How-

    ever, ABA is unequally distributed in vari-

    ous cells in plants. Using an ABA-sensitive

     promoter driving the expression of a lucifer-

    ase gene, the hormone is more concentrated

    in guard cells and in the root tips, with anestimated detectable threshold of 0.3 µM

    (123). In V. faba  and on a per-guard-cell

     basis, ABA in the concentration range of 0.7

    to 1.6 fg of ABA (equivalent to ~0.7 to 1.6

    µM) has been calculated (29, 124, 125). In

    vitro, the IC50 of PP2C activity by ABA act-

    ing through several members of the PYR1/

    PYL1/RCAR family occurs in the nanomo-

    lar range (39, 41, 54). Thus, the amounts of

    ABA required to inhibit PP2Cs and activate

    the signaling pathway are near the basal

    concentrations of ABA in guard cells. Be-

    cause the synthetic ABA agonist, pyrabac-

    tin, can bind PYL2 in two orientations byan induced fit mechanism to produce either

    a productive or nonproductive conforma-

    tion, it has been suggested that there might

     be naturally existing antagonists of ABA

    receptors in plants that could lock the ABA

    receptors in nonproductive orientations,

     perhaps as a safety mechanism against aber-

    rant ABA signaling (45, 126 ). The structural

    insight gained from PYL2 complexed with

    either ABA or pyrabactin offers a tangible

     possibility to embark on rational design of

    chemical modulators of drought resilience

    for crops.

    References and Notes

      1. A. M. Hetherington, F. I. Woodward, The role of

    stomata in sensing and driving environmentalchange. Nature  424, 901–908 (2003).

    2. A. Lebaudy, A. Vavasseur, E. Hosy, I. Dreyer, N.

    Leonhardt, J.-B. Thibaud, A.-A. Véry, T. Simon-neau, H. Sentenac, Plant adaptation to fluctu-

    ating environment and biomass production are

    strongly dependent on guard cell potassiumchannels. Proc. Natl. Acad. Sci. U.S.A.  105,

    5271–5276 (2008).

    3. T.-H. Kim, M. Böhmer, H. Hu, N. Nishimura, J.I. Schroeder, Guard cell signal transduction net-

    work: Advances in understanding abscisic acid,CO2, and Ca

    2+  signaling. Annu. Rev. Plant Biol. 

    61, 561–591 (2010).

    4. T. Vahisalu, I. Puzõrjova, M. Brosché, E. Valk,M. Lepiku, H. Moldau, P. Pechter, Y.-S. Wang,

    O. Lindgren, J. Salojärvi, M. Loog, J. Kangas-

     järvi, H. Kollist, Ozone-triggered rapid stomatalresponse involves the production of reactive

    oxygen species, and is controlled by SLAC1 andOST1. Plant J. 62, 442–453 (2010).

    5. D. Cho, D. Shin, B. W. Jeon, J. M. Kwak, ROS-

    mediated ABA signaling. J. Plant Biol. 52, 102–113 (2009).

    6. M. R. Blatt, Cellular signaling and volume control

    in stomatal movements in plants.Annu. Rev. CellDev. Biol. 16, 221–241 (2000).

    7. J. I. Schroeder, G. J. Allen, V. Hugouvieux, J. M.

    Kwak, D. Waner, Guard cell signal transduction.Annu. Rev. Plant Physiol. Plant Mol. Biol.  52,

    627–658 (2001).

    8. M. R. G. Roelfsema, R. Hedrich, In the light ofstomatal opening: New insights into ‘the Water-

    gate’. New Phytol. 167, 665–691 (2005).9. S. Pandey, W. Zhang, S. M. Assmann, Roles

    of ion channels and transporters in guard cell

    signal transduction. FEBS Lett. 581, 2325–2336

    (2007).10. C. Sirichandra, A. Wasilewska, F. Vlad, C. Valon,

    J. Leung, The guard cell as a single-cell modeltowards understanding drought tolerance and

    abscisic acid action. J. Exp. Bot. 60, 1439–1463

    (2009).11. T. Kinoshita, M. Nishimura, K.-i. Shimazaki,

    Cytosolic concentration of Ca2+  regulates theplasma membrane H+-ATPase in guard cells of

    fava bean. Plant Cell  7, 1333–1342 (1995).

    12. J. I. Schroeder, B. U. Keller, Two types of anionchannel currents in guard cells with distinct volt-

    age regulation. Proc. Natl. Acad. Sci. U.S.A. 89,

    5025–5029 (1992).13. M. R. Roelfsema, V. Levchenko, R. Hedrich, ABA

    depolarizes guard cells in intact plants, through

    a transient activation of R- and S-type anionchannels. Plant J. 37, 578–588 (2004).

    14. C. Schmidt, J. I. Schroeder, Anion selectivity ofslow anion channels in the plasma membrane

    of guard cells (large nitrate permeability). Plant

    Physiol. 106, 383–391 (1994).15. B. U. Keller, R. Hedrich, K. Raschke, Voltage-de-

    pendent anion channels in the plasma membrane

    of guard cells. Nature  341, 450–453 (1989).16. E. A. C. MacRobbie, ABA-induced ion efflux in

    stomatal guard cells: Multiple actions of ABA

    inside and outside the cell. Plant J. 7, 565–576(1995).

    17. A. Schwartz, W.-H. Wu, E. B. Tucker, S. M. Ass-

    mann, Inhibition of inward K+ channels and sto-matal response by abscisic acid: An intracellular

    locus of phytohormone action. Proc. Natl. Acad.Sci. U.S.A. 91, 4019–4023 (1994).

    18. W. Hartung, The site of action of abscisic acid

    at the guard cell plasmalemma of Valerianella

    locusta . Plant Cell Environ. 6, 427–428 (1983).19. B. E. Anderson, J. M. Ward, J. I. Schroeder,

    Evidence for an extracellular reception site forabscisic acid in Commelina   guard cells. Plant

    Physiol. 104, 1177–1183 (1994).

    20. S. Gilroy, R. L. Jones, Perception of gibberel-lin and abscisic acid at the external face of the

    plasma membrane of barley (Hordeum vulgare  L.) aleurone protoplasts. Plant Physiol.  104,

    1185–1192 (1994).

    21. E. Jeannette, J. P. Rona, F. Bardat, D. Cornel, B.Sotta, E. Miginiac, Induction of RAB18  gene ex-

    pression and activation of K+ outward rectifying

    channels depend on an extracellular perception

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    12/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 11

    of ABA in Arabidopsis thaliana  suspension cells.

    Plant J. 18, 13–22 (1999).

    22. T. F. Schultz, R. S. Quatrano, Evidence for sur-face perception of abscisic acid by rice suspen-

    sion cells as assayed by Em gene expression.

    Plant Sci. 130, 63–71 (1997).

    23. D. Yamazaki, S. Yoshida, T. Asami, K. Kuchitsu,

    Visualization of abscisic acid-perception sites on

    the plasma membrane of stomatal guard cells.Plant J. 35, 129–139 (2003).

    24. M. Schwarz, J. I. Schroeder, Abscisic acidmaintains S-type anion channel activity in ATP-

    depleted Vicia faba  guard cells. FEBS Lett. 428,

    177–182 (1998).25. A. B. Ogunkanmi, D. J. Tucker, T. A. Mansfield,

    An improved bioassay for abscisic acid andother antitranspirants. New Phytol. 72, 277–282

    (1973).

    26. N. W. Paterson, J. D. B. Weyers, R. A. Brook, Theeffect of pH on stomatal sensitivity to abscisic

    acid. Plant Cell Environ. 11, 83–89 (1988).

    27. A. C. Allan, M. D. Fricker, J. L. Ward, M. H. Beale,A. J. Trewavas, Two transduction pathways medi-

    ate rapid effects of abscisic acid in Commelina  

    guard cells. Plant Cell  6, 1319–1328 (1994).28. C. Hornberg, E. W. Weiler, High-affini ty binding

    sites for abscisic acid on the plasmalemma ofVicia faba   guard cells. Nature   310, 321–324(1984).

    29. P. McCourt, R. Creelman, The ABA receptors —We report you decide. Curr. Opin. Plant Biol. 11,

    474–478 (2008).

    30. L. B. Sheard, N. Zheng, Plant biology: Signaladvance for abscisic acid. Nature  462, 575–576

    (2009).

    31. The News Staff, Breakthrough of the year: Therunners-up. Science  326, 1600–1607 (2009).

    32. S.-Y. Park, P. Fung, N. Nishimura, D. R. Jensen,

    H. Fujii, Y. Zhao, S. Lumba, J. Santiago, A. Ro-drigues, T.-F. Chow, S. E. Alfred, D. Bonetta, R.

    Finkelstein, N. J. Provart, D. Desveaux, P. L. Ro-driguez, P. McCourt, J.-K. Zhu, J. I. Schroeder,

    B. F. Volkman, S. R. Cutler, Abscisic acid inhibits

    type 2C protein phosphatases via the PYR/PYL

    family of START proteins. Science  324

    , 1068–1071 (2009).

    33. C. Radauer, P. Lackner, H. Breiteneder, The Betv 1 fold: An ancient, versatile scaffold for binding

    of large, hydrophobic ligands. BMC Evol. Biol. 8,

    286 (2008).34. J. Leung, M. Bouvier-Durand, P.-C. Morris, D.

    Guerrier, F. Chefdor, J. Giraudat, Arabidopsis  ABA response gene ABI1: Features of a calci-

    um-modulated protein phosphatase. Science  

    264, 1448–1452 (1994).35. K. Meyer, M. P. Leube, E. Grill, A protein phos-

    phatase 2C involved in ABA signal transduction

    in Arabidopsis thaliana . Science   264, 1452–1455 (1994).

    36. J. Leung, S. Merlot, J. Giraudat, The Arabidop- 

    sis ABSCISIC ACID-INSENSITIVE2  (ABI2 ) and

    ABI1 genes encode homologous protein phos-

    phatases 2C involved in abscisic acid signaltransduction. Plant Cell  9, 759–771 (1997).37. P. L. Rodriguez, G. Benning, E. Grill, ABI2, a sec-

    ond protein phosphatase 2C involved in abscisicacid signal transduction in Arabidopsis . FEBS

    Lett. 421, 185–190 (1998).

    38. A. Saez, N. Apostolova, M. Gonzalez-Guzman,M. P. Gonzalez-Garcia, C. Nicolas, O. Lorenzo,

    P. L. Rodriguez, Gain-of-function and loss-of-

    function phenotypes of the protein phosphatase2C HAB1 reveal its role as a negative regulator

    of abscisic acid signalling. Plant J. 37, 354–369

    (2004).39. J. Santiago, A. Rodrigues, A. Saez, S. Rubio, R.

    Antoni, F. Dupeux, S.-Y. Park, J. A. Márquez, S.

    R. Cutler, P. L. Rodriguez, Modulation of drought

    resistance by the abscisic acid receptor PYL5through inhibition of clade A PP2Cs. Plant J. 60,

    575–588 (2009).40. N. Leonhardt, J. M. Kwak, N. Robert, D. Waner,

    G. Leonhardt, J. I. Schroeder, Microarray expres-

    sion analyses of Arabidopsis   guard cells and

    isolation of a recessive abscisic acid hypersen-sitive protein phosphatase 2C mutant. Plant Cell  

    16, 596–615 (2004).41. Y. Ma, I. Szostkiewicz, A. Korte, D. Moes, Y. Yang,

    A. Christmann, E. Grill, Regulators of PP2C

    phosphatase activity function as abscisic acidsensors. Science  324, 1064–1068 (2009).

    42. S. R. Cutler, P. L. Rodriguez, R. R. Finkelstein, S.R. Abrams, Abscisic acid: Emergence of a core

    signaling network. Annu. Rev. Plant Biol.  61,

    651–679 (2010).43. K. E. Hubbard, N. Nishimura, K. Hitomi, E. D.

    Getzoff, J. I. Schroeder, Early abscisic acid sig-

    nal transduction mechanisms: Newly discoveredcomponents and newly emerging questions.

    Genes Dev. 24, 1695–1708 (2010).

    44. K. Melcher, L.-M. Ng, X. E. Zhou, F.-F. Soon, Y.Xu, K. M. Suino-Powell, S.-Y. Park, J. J. Weiner,

    H. Fujii, V. Chinnusamy, A. Kovach, J. Li , Y. Wang,J. Li, F. C. Peterson, D. R. Jensen, E.-L. Yong,B. F. Volkman, S. R. Cutler, J.-K. Zhu, H. E. Xu,

    A gate-latch-lock mechanism for hormone sig-nalling by abscisic acid receptors. Nature  462,

    602–608 (2009).

    45. K. Melcher, Y. Xu, L.-M. Ng, X. E. Zhou, F.-F.Soon, V. Chinnusamy, K. M. Suino-Powell, A. Ko-

    vach, F. S. Tham, S. R. Cutler, J. Li, E.-L. Yong, J.-

    K. Zhu, H. E. Xu, Identification and mechanismof ABA receptor antagonism. Nat. Struct. Mol.

    Biol. 17, 1102–1108 (2010).

    46. K. Melcher, X.-E. Zhou, H. E. Xu, Thirsty plantsand beyond: Structural mechanisms of abscisic

    acid perception and signaling.Curr. Opin. Struct.Biol. 20, 722–729 (2010).

    47. A. S. Raghavendra, V. K. Gonugunta, A. Christ-

    mann, E. Grill, ABA perception and signalling.

    Trends Plant Sci. 15

    , 395–401 (2010).48. T. Umezawa, K. Nakashima, T. Miyakawa, T.

    Kuromori, M. Tanokura, K. Shinozaki, K. Yama-guchi-Shinozaki, Molecular basis of the core

    regulatory network in ABA responses: Sensing,

    signaling and transport. Plant Cell Physiol. 51,1821–1839 (2010).

    49. F. C. Peterson, E. S. Burgie, S.-Y. Park, D. R. Jen-sen, J. J. Weiner, C. A. Bingman, C.-E. A. Chang,

    S. R. Cutler, G. N. Phillips Jr., B. F. Volkman,

    Structural basis for selective activation of ABAreceptors. Nat. Struct. Mol. Biol. 17, 1109–1113

    (2010).

    50. N. Nishimura, K. Hitomi, A. S. Arvai, R. P. Rambo,C. Hitomi, S. R. Cutler, J. I. Schroeder, E. D. Get-

    zoff, Structural mechanism of abscisic acid bind-

    ing and signaling by dimeric PYR1.Science  326,1373–1379 (2009).

    51. P. Yin, H. Fan, Q. Hao, X. Yuan, D. Wu, Y. Pang,C. Yan, W. Li, J. Wang, N. Yan, Structural insightsinto the mechanism of abscisic acid signaling by

    PYL proteins. Nat. Struct. Mol. Biol. 16, 1230–1236 (2009).

    52. F. Dupeux, J. Santiago, K. Betz, J. Twycross, S.-

    Y. Park, L. Rodriguez, M. Gonzalez-Guzman,M. R. Jensen, N. Krasnogor, M. Blackledge, M.

    Holdsworth, S. R. Cutler, P. L. Rodriguez, J. A.

    Márquez, A thermodynamic switch modulatesabscisic acid receptor sensitivity. EMBO J. 30,

    4171–4184 (2011).

    53. K.-i. Miyazono, T. Miyakawa, Y. Sawano, K. Kubo-ta, H.-J. Kang, A. Asano, Y. Miyauchi, M. Taka-

    hashi, Y. Zhi, Y. Fujita, T. Yoshida, K. S. Kodaira,

    K. Yamaguchi-Shinozaki, M. Tanokura, Structur-

    al basis of abscisic acid signalling. Nature  462,609–614 (2009).

    54. I. Szostkiewicz, K. Richter, M. Kepka, S. Dem-mel, Y. Ma, A. Korte, F. F. Assaad, A. Christmann,

    E. Grill, Closely related receptor complexes dif-

    fer in their ABA selectivity and sensitivity. Plant

    J. 61

    , 25–35 (2010).55. F. Dupeux, R. Antoni, K. Betz, J. Santiago, M.

    Gonzalez-Guzman, L. Rodriguez, S. Rubio, S.-Y.Park, S. R. Cutler, P. L. Rodriguez, J. A. Márquez,

    Modulation of abscisic acid signaling in vivo

    by an engineered receptor-insensitive proteinphosphatase type 2C allele. Plant Physiol. 156,

    106–116 (2011).56. R. R. Johnson, R. L. Wagner, S. D. Verhey, M. K.

    Walker-Simmons, The abscisic acid-responsive

    kinase PKABA1 interacts with a seed-specificabscisic acid response element-binding factor,

    TaABF, and phosphorylates TaABF peptide se-

    quences. Plant Physiol. 130, 837–846 (2002).57. J. Li, S. M. Assmann, An abscisic acid-act ivated

    and calcium-independent protein kinase from

    guard cells of fava bean. Plant Cell   8, 2359–2368 (1996).

    58. J. Li, X.-Q. Wang, M. B. Watson, S. M. Assmann,Regulation of abscisic acid-induced stomatalclosure and anion channels by guard cell AAPK

    kinase. Science  287, 300–303 (2000).59. R. Yoshida, T. Hobo, K. Ichimura, T. Mizoguchi, F.

    Takahashi, J. Aronso, J. R. Ecker, K. Shinozaki,

    ABA-activated SnRK2 protein kinase is requiredfor dehydration stress signaling in Arabidopsis .

    Plant Cell Physiol. 43, 1473–1483 (2002).

    60. A.-C. Mustilli, S. Merlot, A. Vavasseur, F. Fenzi, J.Giraudat, Arabidopsis  OST1 protein kinase me-

    diates the regulation of stomatal aperture by ab-

    scisic acid and acts upstream of reactive oxygenspecies production. Plant Cell   14, 3089–3099

    (2002).61. R. Yoshida, T. Umezawa, T. Mizoguchi, S. Taka-

    hashi, F. Takahashi, K. Shinozaki, The regulatory

    domain of SRK2E/OST1/SnRK2.6 interacts with

    ABI1 and integrates abscisic acid (ABA) and os-motic stress signals controlling stomatal closure

    in Arabidopsis . J. Biol. Chem. 281, 5310–5318(2006).

    62. M. Boudsocq, H. Barbier-Brygoo, C. Laurière,

    Identification of nine sucrose nonfermenting1-related protein kinases 2 activated by hyper-

    osmotic and saline stresses in Arabidopsis thali- ana . J. Biol. Chem. 279, 41758–41766 (2004).

    63. H. Fujii, P. E. Verslues, J.-K. Zhu, Arabidop- 

    sis   decuple mutant reveals the importance ofSnRK2 kinases in osmotic stress responses in

    vivo. Proc. Natl. Acad. Sci. U.S.A.  108, 1717–

    1722 (2011).64. H. Fujii, J.-K. Zhu, Arabidopsis  mutant deficient

    in 3 abscisic acid-activated protein kinases re-

    veals critical roles in growth, reproduction, andstress. Proc. Natl. Acad. Sci. U.S.A. 106, 8380–

    8385 (2009).65. T. Umezawa, N. Sugiyama, M. Mizoguchi, S.Hayashi, F. Myouga, K. Yamaguchi-Shinozaki,

    Y. Ishihama, T. Hirayama, K. Shinozaki, Type 2Cprotein phosphatases directly regulate abscisic

    acid-activated protein kinases in Arabidopsis .

    Proc. Natl. Acad. Sci. U.S.A. 106, 17588–17593(2009).

    66. F. Vlad, S. Rubio, A. Rodrigues, C. Sirichandra,

    C. Belin, N. Robert, J. Leung, P. L. Rodriguez,C. Laurière, S. Merlot, Protein phosphatases 2C

    regulate the activation of the Snf1-related kinase

    OST1 by abscisic acid in Arabidopsis . Plant Cell  

    21, 3170–3184 (2009).

  • 8/20/2019 Joshi-saha2011_A Brand New START_Abscisic Acid Perception and Transduction in the Guard Cell

    13/14

    REV IEW

    www.SCIENCESIGNALING.org 29 November 2011 Vol 4 Issue 201 re4 12

      67. N. Nishimura, A. Sarkeshik, K. Nito, S. Y. Park,

    A. Wang, P. C. Carvalho, S. Lee, D. F. Caddell,

    S. R. Cutler, J. Chory, J. R. Yates, J. I. Schroeder,PYR/PYL/RCAR family members are major in-

    vivo ABI1 protein phosphatase 2C-interactingproteins in Arabidopsis . Plant J.  61, 290–299

    (2010).

    68. H. Fujii, V. Chinnusamy, A. Rodrigues, S. Rubio,

    R. Antoni, S.-Y. Park, S. R. Cutler, J. Sheen, P.L. Rodriguez, J.-K. Zhu, In vitro reconstitution of

    an abscisic acid signalling pathway. Nature  462,660–664 (2009).

    69. X. Xie, Y. Wang, L. Williamson, G. H. Holroyd, C.

    Tagliavia, E. Murchie, J. Theobald, M. R. Knight,W. J. Davies, H. M. O. Leyser, A. M. Hethering-

    ton, The identification of genes involved in thestomatal response to reduced atmospheric rela-

    tive humidity. Curr. Biol. 16, 882–887 (2006).

    70. H. Fujii, P. E. Verslues, J.-K. Zhu, Identificationof two protein kinases required for abscisic acid

    regulation of seed germination, root growth, and

    gene expression in Arabidopsis . Plant Cell  19,485–494 (2007).

    71. T. Yoshida, Y. Fujita, H. Sayama, S. Kidokoro, K.

    Maruyama, J. Mizoi, K. Shinozaki, K. Yamaguchi-Shinozaki, AREB1, AREB2, and ABF3 are mas-

    ter transcription factors that cooperatively regu-late ABRE-dependent ABA signaling involved indrought stress tolerance and require ABA for full

    activation. Plant J. 61, 672–685 (2010).72. T. Furihata, K. Maruyama, Y. Fujita, T. Umeza-

    wa, R. Yoshida, K. Shinozaki, K. Yamaguchi-

    Shinozaki, Abscisic acid-dependent multisitephosphorylation regulates the activity of a tran-

    scription activator AREB1. Proc. Natl. Acad. Sci.