Cushman 2001

Embed Size (px)

Citation preview

  • 8/8/2019 Cushman 2001

    1/10

    Update on Crassulacean Acid Metabolism

    Crassulacean Acid Metabolism. A Plastic PhotosyntheticAdaptation to Arid Environments1

    John C. Cushman*

    Department of Biochemistry, University of Nevada, Reno, Nevada 895570014

    Crassulacean acid metabolism (CAM) is an impor-tant elaboration of photosynthetic carbon fixationthat allows chloroplast-containing cells to fix CO

    2

    initially at night using phosphoenolpyruvate carbox-ylase (PEPC) in the cytosol. This leads to the forma-tion of C4 organic acids (usually malate), which arestored in the vacuole. Subsequent daytime decarbox-ylation of these organic acids behind closed stomata

    creates an internal CO2 source that is reassimilatedby Rubisco in the chloroplast. The refixation of thisinternal CO

    2generates carbohydrates via the conven-

    tional photosynthetic carbon reduction cycle. Thus,CAM involves a temporal separation of carbon fixa-tion modes in contrast to the spatial separation foundin C

    4plants. The first recognition of the nocturnal

    acidification process can be traced to the Romans,who noted that certain succulent plants taste more

    bitter in the morning than in the evening (Rowley,1978). However, formal descriptions of the ability ofsucculent plants to conduct nocturnal CO2 fixation orto acidify photosynthetic tissues at night and deacid-

    ify them during the day did not appear until the early19th century (de Saussure, 1804; Heyne, 1815). Theterm CAM was coined to give credit to Heynesobservations that were made using Bryophyllum caly-cinum, a succulent member of the Crassulaceae.

    Since these early descriptions, a detailed account ofthe sequence of biochemical reactions of the CAMcycle (Ranson and Thomas, 1960), the complexity ofthe biochemical variations in the pathway amongdifferent CAM species, and its regulation by the en-vironment have been achieved (Osmond, 1978; Ting,1985). Initial nocturnal CO

    2fixation by PEPC occurs

    when stomata are open and transpirational waterlosses are low. CO2 release during the day promotesstomatal closure and concentrates CO2 aroundRubisco, suppressing its oxygenase activity, therebyminimizing photorespiration. The net effect of thisCO

    2-concentrating strategy is that CAM plants ex-

    hibit water use efficiency (WUE) rates severalfoldhigher than C3 and C4 plants under comparable con-ditions (Drennan and Nobel, 2000). Thus, CAM istypically, although not exclusively, associated with

    plants that inhabit extremely arid environments (e.g.deserts), semi-arid regions with seasonal water avail-ability (e.g. Mediterranean climates), or habitats withintermittent water supply (e.g. tropical epiphytichabitats). Most notable among these are commer-cially or horticulturally important plants such aspineapple ( Ananas comosus), agave (Agave subsp.),cacti (Cactaceae), and orchids (Orchidaceae). CAM is

    also correlated with various anatomical or morpho-logical features that minimize water loss, includingthick cuticles, low surface-to-volume ratios, largecells and vacuoles with enhanced water storage ca-pacity (i.e. succulence), and reduced stomatal sizeand/or frequency.

    The selective advantage of high WUE likely ac-counts for the extensive diversification and specia-tion among CAM plants principally in water-limitedenvironments. Intensive ecophysiological studiesover the last 20 years have documented that CAM ispresent in approximately 7% of vascular plant spe-cies, a much larger percentage than the percentage ofC4

    species (Winter and Smith, 1996a). The wide-

    spread distribution of CAM among 33 taxonomicallydiverse families (Smith and Winter, 1996) suggeststhat CAM most likely evolved independently on nu-merous occasions in different families and evenwithin individual families (Griffiths, 1989; Ehleringerand Monson, 1993; Pilon-Smits et al., 1996). Morerecent phylogenetic reconstructions using PEPC se-quence information have provided more convincingsupport for the polyphyletic origins of CAM (Gehriget al., 1998b, 2001). It is curious that CAM is alsofound in aquatic vascular plants where it presumablyenhances inorganic carbon acquisition in certainaquatic environments where CO2 availability can be-

    come rate limiting for photosynthesis (Keeley, 1996,1998). Thus, the daytime limitation of CO2 availabil-ity, brought about by water-conserving stomatal clo-sure in arid terrestrial habitats or by competitionfrom other species and the high diffusional resis-tances limiting access to CO2 in aquatic habitats,appears to be the common factor responsible for theevolution of CAM.

    A REMARKABLE PLASTICITY

    One of most striking themes to emerge in recentyears is the extent to which the phylogenetic and

    ecological diversity of CAM plants is also reflected in

    1 This work was supported by the National Science Foundation(Integrative Plant Biology, Plant Genome Programs) and by theNevada Agricultural Experiment Station.

    * E-mail [email protected]; fax 7757841650.

    www.plantphysiol.org/cgi/doi/10.1104/pp.010818.

    Plant Physiology, December 2001, Vol. 127, pp. 14391448, www.plantphysiol.org 2001 American Society of Plant Biologists 1439

  • 8/8/2019 Cushman 2001

    2/10

    a remarkable plasticity of the basic metabolic schemedescribed above. Genotypic, ontogenetic, and envi-ronmental factors such as light intensity, relative hu-midity, and water availability combine to govern theextent to which the biochemical and physiological

    attributes of CAM are expressed (Cushman and Bor-land, 2001). The photosynthetic plasticity of CAMoccurs within a continuum of diel gas exchange pat-terns that fall into four phases as defined by Osmond(1978). The nocturnal uptake of atmospheric and re-spiratory CO2 via PEPC to form C4 acids (phase I)and daytime organic acid decarboxylation to gener-ate elevated Ci and stomatal closure (phase III) areinterspersed with transitional periods of net CO

    2up-

    take at the start (phase II) and end of the day (phaseIV) when both PEPC- and Rubisco-mediated carbox-ylation can contribute to CO2 assimilation. The pro-portion of CO

    2taken up via PEPC at night or directly

    during the day by Rubisco (net CO2 assimilation) is

    dictated by the integration of stomatal behavior, fluc-tuations in organic acid and storage carbohydrateaccumulation, and the abundance and activity of pri-mary (PEPC) and secondary (Rubisco) carboxylatingand decarboxylating enzymes (e.g. malic enzyme orPEP carboxykinase), as well as gluconeogenic/glyco-lytic enzymes responsible for the synthesis and

    breakdown of C3 carbon skeletons.Depending on developmental and/or environmen-

    tal influences, a variety of CO2 assimilation, acid flux,and stomatal behavior characteristics may be ob-served outside the conventional pattern of the four

    phases (Table I). Nearly-C3

    or CAM cyclingspecies display daytime net CO2 uptake with refix-ation of respiratory CO

    2at night accompanied by

    only small diel C4 acid fluctuations. In plants grow-ing in thin soils or rock outcrops, this nocturnal

    recapture of respiratory CO2

    is thought to help main-tain a positive carbon balance during frequent epi-sodes of drought (Martin, 1996). However, the poten-tial conservation of water resulting from theinduction of CAM cycling varies widely (5%70%) invarious species (Borland, 1996; Martin, 1996). In C3-CAM intermediate species, such improvements inWUE are not always associated with CAM induction(Eller and Ferrari, 1997; Cushman and Borland, 2001).In obligate or constitutive CAM species, net CO2uptake occurs almost exclusively at night (phase I),with some net CO2 assimilation occurring duringphases II and IV, even under well-watered condi-tions, accompanied by large diel C4 acid fluctuations.

    Under severe drought conditions, many CAM spe-cies will undergo CAM-idling wherein stomata re-main closed day and night, preventing net CO

    2up-

    take, yet the plants will continue to conduct dielfluctuations in organic acids. Other modes of CAMsuch as latent CAM, indicated by organic acid con-centrations elevated above those normally present inC3 plants but without diel fluctuation, may representa nascent C

    3-to-CAM progression in some species

    (Schuber and Kluge, 1981). A hypothetical variationof CAM called rapid-cycling CAM has also beenproposed in which the CO2 acquisition and reduction

    Table I. Plasticity of CAM modes in relationship to environmental and developmental influences

    CAM VariationNet CO2Uptake

    RespiratoryCO2

    Refixation?

    Stomatal Behavior(Open)

    C4 Acid FluxDevelopmental/Environmental

    Impact

    Proposed Functions

    C3 Day No Day a /

    CAM cycling Day Yes Day / Maintain positive carbonbalance; improved

    WUE?b; reduced photo-respiration?

    C3-CAM (faculta-tive CAM)

    Day/night Yes Day/night / Maintain positive carbonbalance; improved

    WUE?; reduced photore-spiration

    CAM (obligateCAM)

    Phases I, II, and IV Yes Day/night / Improved WUE; reducedphotorespiration

    Phase II CAM Phases I and II Yes Morning/night / Improved WUE; reducedphotorespiration

    Phase I CAM Phase I only Yes Night only / Improved WUE; reducedphotorespiration

    CAM idling None Yes Always closed / Protection of photosyn-thetic apparatus from

    photoinibition; maintaina positive carbon bal-

    anceLatent CAM Day Yes Day (elevated) ?/? C3 to CAM progression?Rapid-cycling CAM

    (theoretical)All ? ? (rapid) ?/? ?

    a

    Dashes indicate no substantial occurrence or effect.b

    Question marks indicate that no information is available.

    Cushman

    1440 Plant Physiol. Vol. 127, 2001

  • 8/8/2019 Cushman 2001

    3/10

    phases of CAM may occur over time periods shorterthan the normal diel cycle (Cockburn, 1998).

    The best examples of CAM plasticity are the C3-

    CAM intermediate species found predominantlyamong the Aizoaceae, Crassulaceae, Portulaceae, and

    Vitaceae (Smith and Winter, 1996). These facultativeor inducible CAM species use the C3

    pathway tomaximize growth when water is abundant, but thenthey undergo a gradual C

    3-to-CAM transition oftencoincident with seasonal moisture availability (Win-ter et al., 1978). The C3-to-CAM transition reduceswater loss and maintains photosynthetic integrityunder water-limited conditions that ultimately trans-lates into reproductive success (Winter and Ziegler,1992). Among facultative CAM species, the commonice plant, Mesembryanthemum crystallinum, has beenmost intensively studied (Adams et al., 1998; Bohnertand Cushman, 2001). This model species undergoes agradual, largely irreversible, and partially develop-

    mentally regulated transition into CAM followingwater stress (Cushman et al., 1990; Herppich et al.,1992). In contrast, other inducible CAM species (e.g.Clusiaceae and Bromeliaceae) display more rapidand reversible shifts between C

    3photosynthesis and

    CAM in response to changes in water deficit, regard-less of leaf or plant ontogeny (Schmitt et al., 1988;Zotz and Winter, 1993; Luttge, 1996; Borland et al.,1998). The magnitude of CAM induction in faculta-tive CAM plants tends not only to be influenced bywater deficit, but also by associated environmentalconditions such as temperature, light intensity, andhumidity (Luttge, 2000). For example, it is well es-

    tablished that high light intensity or light quality canenhance CAM induction in the ice plant in the pres-ence or absence of salinity stress (McElwain et al.,1992; Cockburn et al., 1996; Miszalski et al., 2001).

    MOLECULAR GENETICS OF CAM

    Since the first molecular characterization of thecommon ice plant Ppc1 gene encoding a CAM-specific isoform of PEPC more than a decade ago(Cushman et al., 1989), a large number of enzymes,transporters, and regulatory proteins required forCAM have been identified and characterized (forreview, see Cushman and Bohnert, 1999, 2001; Cush-man and Borland, 2001). Most studies have beenrestricted to inducible C

    3-CAM models (e.g. common

    ice plant and Kalanchoe sp.) because the differentialexpression of genes induced in response to waterdeficit serves as a convenient and reliable indicator oftheir potential functional role(s) in CAM. Greaterinvestments have been made in establishing molecu-lar genetic resources for common ice plant than otherCAM models because this species is also a halophyteand has been extensively investigated to understandsalinity stress tolerance mechanisms (Bohnert andCushman, 2001; Bohnert et al., 2001). CAM induction

    in response to salinity, water deficit, osmotic stress,

    or abscisic acid treatment is controlled primarily bytranscriptional activation (Cushman et al., 1989,2000b) initiated through a signaling cascade withapparent requirements for calcium and calcium-dependent protein kinase activities (Taybi and Cush-

    man, 1999; Golldack and Dietz, 2001). In general,transcript and protein accumulation patterns are wellcorrelated; however, discrepancies between tran-script and protein abundance have suggested thatchanges in mRNA stability and utilization or trans-lational efficiency are also likely to govern gene ex-pression changes during the C

    3-to-CAM transition

    (Cushman et al., 1990; DeRocher and Bohnert, 1993).Detailed analysis of the PEPC gene families from

    facultative and obligate CAM species including pine-apple ( Ananas comosus), K. blossfeldiana, K. daigremon-tiana, common ice plant, and Vanilla planifolia has in-dicated that a single member of a four- to six-memberPEPC gene family is typically recruited to fulfill theprimary carboxylation and carbon flux requirementsof CAM, as demonstrated by its enhanced expressionin CAM-performing leaves (Cushman et al., 1989; Ge-hrig et al., 1995, 1998a). Remaining isoforms, whichpresumably fulfill anapleurotic housekeeping ortissue-specific functional roles, generally show lowertranscript or protein abundance and remain unaf-fected in their expression following CAM induction.This gene recruitment paradigm likely pertains toother gene families as well. Enhanced expression ofenzymes for C

    4acid metabolism is accompanied by

    corresponding increases in carbohydrate-forming and-degrading enzymes and transcripts (Holtum andWinter, 1982; Paul et al., 1993; Hausler et al., 2000).Elevated organellar PEP (Kore-eda et al., 1996) andtriose and hexose phosphate transport activities (Neu-haus and Schulte, 1996; Kore-eda and Kanai, 1997)associated with CAM induction in common ice plantare matched by light-enhanced increases in transcriptabundance and diurnal gene expression patterns of aPEP phosphate translocator and a Glc-6-P phosphatetranslocator (Hausler et al., 2000). However, the ex-pression of a chloroplast Glc transporter and a triosephosphate transporter remain largely unchanged(Hausler et al., 2000; S. Kore-eda and J.C. Cushman,unpublished data). Tonoplast H-translocating

    ATPase transport activity and expression of corre-sponding tonoplast H-translocating ATPase subunitgenes for energizing vacuolar malate storage is en-hanced during the C3-CAM transition in common iceplant (Rockel et al., 1998a, 1998b; Barkla et al., 1999;Golldack and Dietz, 2001). Molecular characteriza-tion of the vacuolar malate transporters, carriers, andchannels for malate influx and efflux has remained achallenge (Luttge et al., 2000). Recent measurementsof vacuolar malate transport activities demonstratean approximate 3-fold increase following CAM in-duction in common ice plant (Luttge et al., 2000). Astrategy to analyze differences in polypeptide expres-

    sion patterns in C3- versus CAM-performing leaves

    Crassulacean Acid Metabolism

    Plant Physiol. Vol. 127, 2001 1441

  • 8/8/2019 Cushman 2001

    4/10

    of common ice plant is being used to identify candi-date vacuolar malate transporters. Antisera raisedagainst affinity chromatography-purified tonoplastvesicle fractions from K. daigremontiana enriched formalate transport activity has been used to identify

    32- and 33-kD common ice plant polypeptides thatare induced or enhanced in the CAM state (Steiger etal., 1997; Luttge et al., 2000). These low abundancepolypeptides could be candidates for the vacuolarmalate transporter. Amino acid sequence informa-tion from these polypeptides may facilitate the isola-tion of the corresponding genes.

    CIRCADIAN CONTROL OF CAM

    The circadian rhythm of CO2 fixation, primarilystudied in K. fedtschenkoi, is one of the earliest and

    best documented examples of circadian rhythms in

    higher eukaryotes (Wilkins, 1992). Diel oscillations inthe activity of PEPC, controlled in part by circadianchanges in its phosphorylation state, play a key rolein directing carbon flux through the CAM pathway

    by changing the enzymes sensitivity to allostericinhibitors such as malate (Nimmo et al., 1987;Nimmo, 1998). PEPC phosphorylation state is con-trolled largely by changes in the activity of PEPCkinase (PPcK; Carter et al., 1991). In common iceplant, PPcK activity is induced concomitantly with aCAM-specific isoform of PEPC (Li and Chollet, 1994).Recent cloning of the gene for PPcK first in K.

    fedtschenkoi (Hartwell et al., 1999) and then in com-

    mon ice plant (Taybi et al., 2000) demonstrated di-rectly that this kinase is itself regulated at the level oftranscript abundance by a circadian oscillator. A dis-sociable protein inhibitor of PPcK activity has also

    been described from K. fedtschenkoi that may functionto suppress basal kinase activity during the lightperiod and early stages of the dark period whencarbon flux through PEPC is not needed (Nimmo etal., 2001a). In contrast to C

    4plants, elevations in

    cytosolic pH appear to have little (Bakrim et al., 2001)or no influence (Paterson and Nimmo, 2000) on PPcKactivity in common ice plant or K. fedtschenkoi, re-spectively. However, circadian control of PPcK tran-script abundance may be merely a secondary re-sponse to other factors such as the cytosolic malateconcentration, which has been hypothesized to reg-ulate the transcript abundance and activity of PPcK(Borland et al., 1999; Nimmo, 2000). Cytosolic malateconcentrations are likely to be controlled by transportof malate across the tonoplast, a view that is wellsupported by temperature effects on tonoplast func-tion and modeling studies (Rascher et al., 1998;Luttge, 2000). Thus, response to environmental fac-tors that alter organic acid content or malate parti-tioning between the vacuole and cytosol may be ableto override circadian rhythms of PPcK activity, pro-viding a possible mechanism for the rapid alterations

    in PEPC activity observed in some CAM species (Bor-

    land et al., 1999; Nimmo, 2000). In addition, feeding ofdetached K. fedtschenkoi and common ice plant leaveswith various pharmacological reagents implicates theinvolvement of a phosphoinositide-dependent phos-pholipase C, inositol 1,4,5 P-gated tonoplast calcium

    channels, a putative Ca

    2

    -dependent/calmodulinprotein kinase, and RNA and protein synthesis aspossible components in the signaling cascade that reg-ulates PPcK activity on a circadian basis (Hartwell etal., 1999; Bakrim et al., 2001; Nimmo et al., 2001b).However, these studies fail to address the influence ofsuch inhibitors on the functioning of the underlyingcircadian oscillator, and so, observed changes in PEPCactivity may not reflect alterations in the PEPC kinase-signaling cascade per se. One great challenge to un-derstanding circadian regulation of CAM will be todissect the mechanisms responsible for controlling thecircadian oscillations in malate uptake and releaseacross the tonoplast membrane. In particular, it will be

    important to understand how tonoplast malate trans-port is controlled by an underlying nuclear-controlledcircadian clock. Rapid molecular identification ofmalate transport components in the tonoplast and cir-cadian clock components from CAM species will beessential for this effort.

    A GENETIC MODEL FOR CAM?

    To date, ecophysiological investigations have sur-veyed a wide variety of CAM species to determinewhich ones actually perform CAM. Alternatively,

    studies have focused on comparative analysis of spe-cific aspects of CAM such as the degree of CAMinduction by water limitation (Cushman and Bor-land, 2001), intercellular localization of carboxylationand decarboxylation processes (Borland et al., 1998),or the patterns of carbohydrate partitioning within aparticular family (Christopher and Holtum, 1996,1998). However, unlike C3 and C4 plants, which havethe well-developed genetic models Arabidopsis andmaize (Zea mays), respectively, there has been, untilrecently, no investment in the development of a ge-netic model for CAM. This deficiency has hinderedour understanding of many of the molecular mecha-nisms that regulate CAM. In the past, CAM modelswere selected for their physiological characteristics.For example, certain obligate CAM species such as K.daigremontiana are often favored for gas exchangeand biochemical studies due to their reproducible

    behavior. Other CAM models such as common iceplant can show hyperplastic stress responsiveness toslight changes in growth conditions, which can be aproblem for reproducible physiological studies.Kalanchoe species, however, lack potential for devel-opment as a genetic system as well as any significantmolecular genetic resources.

    A comparison of the attributes of well-studied orcommercially important CAM models from diverse

    families indicates that common ice plant has many

    Cushman

    1442 Plant Physiol. Vol. 127, 2001

  • 8/8/2019 Cushman 2001

    5/10

    desirable features that make it an attractive geneticmodel (Table II). This fast-growing annual produceslarge quantities of small seeds (typically 10,00015,000plant1) under standard greenhouse or growth cham-

    ber conditions in 1-L pots. The plant is self-fertile, yet

    outcrossing is possible. In contrast, the perennial orsemi-perennial pineapple, Kalanchoe, and Clusia spe-cies grow more slowly and are poor seed producers.Although the common ice plant grows more slowlythan models such as Arabidopsis, compared withother CAM models, the common ice plant life cycle isquite rapid. Furthermore, it is possible to acceleratethe normal life cycle of common ice plant from 4 to 5mo under natural conditions (Winter et al., 1978) toapproximately 7 weeks under growth chamber condi-tions under continuous light or extended photoperi-ods and limited rooting volumes (Cheng and Ed-wards, 1991). Acceleration of the life cycle isconveniently accompanied by a miniaturization of the

    plant. This is an important consideration when con-ducting genetic screening because growth chamber orgreenhouse space is often a limiting factor. Alterna-tively, genetic screens could be conducted in a recentlyidentified dwarf mutant background that displaysCAM (see below). Finally, mutant collections have

    been established in common ice plant from fastneutron- or gamma-irradiated (Cushman et al., 2000b)or ethylmethane sulfonate-treated seeds (Adams et al.,1998). Expansion of existing fast neutron collectionswould create a useful resource for a fast neutronmutagenesis-based reverse genetic screening systemin the common ice plant, similar to related resources

    recently developed in Arabidopsis and rice (Oryza

    sativa; Li et al., 2001). Facile screening procedures have been developed for the isolation of CAM-defectivemutants (Cushman et al., 2000b). Identification ofCAM-defective mutants is based on a simple pH assaythat detects a failure in nocturnal C4 acid accumula-

    tion. Mutant collections are not currently available inother CAM models. A useful by-product of such mu-tant screens is the identification of mutants with mor-phological (e.g. dwarfism and absence of epidermal

    bladder cells) or physiological defects (e.g. salt sensi-tivity; J.C. Cushman, unpublished data).

    Another desirable feature of the ideal CAM modelis the availability of an efficient transformation sys-tem, preferably one that employs a non-tissueculture-based methodology such as vacuum infiltra-tion or floral dipping in Agrobacterium tumefacienssuspensions (Bechtold et al., 1993; Clough and Bent,1998). A transformation system with adequate effi-

    ciency would allow systematic functional genomicinvestigations to be performed involving reverse ge-netic screens for T-DNA insertion/activation-taggedgene knockouts, suppression or overexpression stud-ies, and ultimately targeted gene replacement of reg-ulatory or structural genes of interest with key rolesin CAM. Of the possible candidate model CAM spe-cies, several are amenable to genetic manipulationusing an A. tumefaciens-mediated transformation sys-tem (Truesdale et al., 1999). However, given the iceplants susceptibility to A. tumefaciens transformationin tissue culture (Andolfatto et al., 1994; Ishimaru,1999) and the availability of a high efficiency regen-

    eration system (Cushman et al., 2000b), an experi-

    Table II. Comparison of desirable attributes of well-studied CAM models

    AttributeCommon Ice Plant(M. crystallinum)

    Mother-of-Thousands(K. daigremontiana, K.

    blossfeldiana, K.fedtschenchoi)

    Balsam Apple (Clusia minor,C. major, C. rosea)

    Pineapple(A. comosus)

    Family Aizoaceae (dicot) Crassulaceae (dicot) Clusiaceae (dicot) Bromeliaceae (monocot)Growth habit Annual Perennial Perennial SemiperennialCommercial/horticultural

    importance (relative value)Ornamental, ground

    cover, and fire break ($)Ornamental ($$) Ornamental () Edible crop ($$$$)

    Mode of CAM C3-CAM and stress inducible Obligate anddevelopmentally

    regulated

    C3-CAM, rapid andreversible

    Obligate anddevelopmentally

    regulatedPropagation mode Seed Clonal Clonal ClonalSeed production 10,000 15,000 plant1a Poor to none None Poor to noneGrowth rate (to adulthood) Rapid (6 weeks) Intermediate

    (23 mo)Intermediate

    (23 mo)Slow (18 24 mo)

    Mutant collections? Ethylmethane sulfonate,fast neutron, and gamma

    irradiated

    No No No

    Transgenic plants? No (callus, yes) Yes No YesTransformable? Yes Yes ?b YesGenome size (Mb) 390 790 1,500 ? 526No. of expressed sequence

    tags (ESTs) available15,000 0 0 0

    Microarray availability? Yes No No Noa

    Soil-grown plants under standard laboratory conditions in 1-L pots.b

    Question marks indicate that no information is available.

    Crassulacean Acid Metabolism

    Plant Physiol. Vol. 127, 2001 1443

  • 8/8/2019 Cushman 2001

    6/10

    mental platform for future transgenic analysis incommon ice plant appears highly feasible.

    The other major limitation for CAM research hasbeen the lack a genetic model with a wealth of avail-able molecular genetic information, such as the com-

    plete nucleotide sequence of the genome or at thevery least, sizeable collections of ESTs. The commonice plant genome is approximately 390 Mb, as esti-mated by flow cytometry (DeRocher et al., 1990) innine chromosomes (2n 18; Adams et al., 1998) orapproximately 2.5 times larger than the Arabidopsisgenome (approximately 145 Mb) and slightly smallerthan the rice genome (approximately 420 Mb). Thecommon ice plant genome is also smaller than allother CAM models for which such data are currentlyavailable. For example, the pineapple genome (2n 25) is somewhat larger, with a DNA content approx-imately 3.7 times the size of the Arabidopsis genome

    (Arumuganathan and Earle, 1991; Williams andFleisch, 1993), whereas K. fedtschenkoi and K. blossfel-diana are two (approximately 790 Mb) and four times(approximately 1,500 Mb) the size, respectively, ofthe common ice plant genome (DeRocher et al., 1990).Thus, the small size of the common ice plant genomemakes it a most attractive target for genomesequencing.

    In lieu of genomic sequence information, the avail-ability of information-rich sequence data from ESTcollections would add strong incentives for investi-gators to invest in a particular CAM model. AlthoughcDNA libraries are available for K. daigremontiana

    (Bartholomew et al., 1996) and K. fedtschenkoi (Hart-well et al., 1999), the most comprehensive collectionof cDNA libraries for any CAM plant is available forthe common ice plant. More than 30 cDNA librariesexist from tissues that span the entire life cycle, fromseedling to adult and flowering stages, as well asdifferent tissues such as meristems, roots, shoots,leaves, epidermal bladder cells, flowers and seedcapsules, and different stress treatments (Bohnertand Cushman, 2001). Furthermore, more than 15,000ESTs are now available (http://www.ncbi.nlm.nih.gov/dbEST/dbEST_summary.html; Bohnert andCushman, 2001). In addition, a gene index has been

    recently created that allows easy access to the ESTsequence information in the form of nonredundantgenes (singletons) and tentative consensus sequencesderived from redundant cDNAs (http://www.tigr.org/tdb/mcgi/). However, similar investmentsin other intensively studied models such as K. daigre-montiana and Clusia spp. in which cDNA libraries arealso under development (T. Taybi and A.M. Borland,personal communication) will be needed for compar-ative analyses of the functional significance of genesencoding signaling and regulatory components, en-zymes, and transporters and to extend cross-speciescomparison beyond current physiological or bio-

    chemical investigations.

    Abundant molecular genetic resources will facili-tate integrative approaches to phenomena rangingfrom gene expression to gas exchange characteristics.Such integration is required to identify and distin-guish the functional contribution and regulation of

    specific gene products, especially among circadianlyregulated genes. Large EST collections and associ-ated databases provide the foundation of nucleotidesequence information on which to build anticipatedgenome sequencing efforts (see below), as well asmaterials with which to print cDNA-based microar-rays or to synthesize oligonucleotide-based Gene-Chips for large-scale gene expression-profiling ex-periments. Extensive or comprehensive expressiondata can often provide important clues about thefunction of specific isogenes in CAM or implicateroles in CAM for previously uncharacterized genes.Analysis of the existing common ice plant EST data-

    base compiled from salinity-stressed, CAM-induced

    plants indicates the presence of large numbers ofgenes, perhaps up to several thousand, that are notrepresented in other plant databases (Bohnert andCushman, 2001). Such unknown or novel ESTs in thecommon ice plant database may arise, in part, fromthe evolutionary distance between common ice plantand the other plant models. We also expect that genefamily expansion has occurred in the common iceplant, a native of the Namib Desert, to meet theadditional requirements of CAM for long-term sur-vival and reproductive success in arid environments.Evidence for this can be seen in, for example, thePEPC gene family. In Arabidopsis, this gene family is

    comprised of four members. In the common ice plant,however, at least six members make up this genefamily, with only one specifically recruited to func-tion in CAM (Cushman and Borland, 2001).

    SEQUENCING A CAM PLANT GENOME?

    Recent technological improvements in high-throughput, automated DNA sequencing systemsand access to large capacity sequencing facilitiesmake it reasonable to call for the sequencing of thecomplete genome of a CAM plant in the near future.The common ice plant is a logical choice for such anundertaking because it has the smallest genomeamong well-studied CAM models and the largestEST collection for gene identification (Table II). Thiseffort will also provide important genomic informa-tion for comparative genomic studies of a specieswithin the Caryophyllales. Most genome sequencingefforts target the major crop species in the Cruciferae,Poaceae, and Solanaceae. In contrast, very fewCaryophyllales, which includes such plant familiesas the Aizoaceae, Amaranthaceae, Cactaceae, Che-nopodiaceae, Caryophyllaceae, Phytolaccaceae, andPortulacaceae, are targets for genomic sequencingefforts because most are crop or ornamental species

    of relatively minor economic value. Yet, many spe-

    Cushman

    1444 Plant Physiol. Vol. 127, 2001

  • 8/8/2019 Cushman 2001

    7/10

    cies in the order Caryophyllales have evolved tocolonize environments characterized by water defi-cit, salinity, or extreme temperatures. As such, thesespecies can be expected to be useful sources of novelgenes involved in extending unusual biochemical

    pathways for plant secondary metabolites or abioticstress tolerance. For example, many species of theCaryophyllales accumulate chromogenic betacyaninsinstead of anthocyanins and other complex substi-tuted flavonoids. Thus, access to complete sequenceinformation for the common ice plant would facili-tate discovery of genes with CAM-specific functionsor regulation (e.g. circadianly regulated genes), aswell as of new gene products for abiotic stress adap-tation and natural product biosynthesis and chemis-try (Vogt et al., 1999a, 1999b).

    PERSPECTIVES

    The C3 and C4 photosynthetic pathways have beenextensively investigated at the molecular geneticlevel. Much of this research has been greatly facili-tated by the availability of excellent and well-studiedgenetic models and an abundance of cDNA andgenomic sequence information. In contrast, our un-derstanding of the complex regulation of the CAMphotosynthetic pathway has lagged behind theseother models. However, recent advances toward thecreation of one or more viable genetic models forCAM, coupled with increasing availability of genesequence and expression information, forecast a

    bright and productive future for CAM researchers.

    Future development and application of genomic,proteomic, and metabolic profiling technologies inselected CAM models such as the common ice plantis expected to rapidly improve our understanding ofCAM induction by environmental and developmen-tal influences and the circadian rhythms that dictatethe diel patterns of CO2 fixation characteristic ofCAM plants. Thus, the greatest challenge facingCAM researchers in the future will be to developteams of interdisciplinary researchers using genomic,

    biochemical, and physiological research approachesin selected CAM models. This approach will providean integrated view of the complex regulatory dynam-ics that allow such remarkably plastic responses tothe environment that has become one of the greathallmarks of CAM plants.

    ACKNOWLEDGMENTS

    I would like to thank Mary Ann Cushman and JamesHartwell for their critical reading of the manuscript.

    Received September 7, 2001; returned for revision Septem- ber 10, 2001; accepted September 16, 2001.

    LITERATURE CITED

    Adams P, Nelson DE, Yamada S, Chmara W, Jensen RG,

    Bohnert HJ, Griffiths H (1998) Tansley Review No. 97:

    growth and development of Mesembryanthemum crystalli-num (Aizoaceae). New Phytol 138: 171190

    Andolfatto R, Bornhouser A, Bohnert HJ, Thomas JC(1994) Transformed hairy roots of Mesembryanthemumcrystallinum: gene expression patterns upon salt stress.Physiol Plant 90: 708714

    Arumuganathan K, Earle ED (1991) Nuclear DNA contentof some important plant species. Plant Mol Biol Rep 9:208218

    Bakrim N, Brulfert J, Vidal J, Chollet R (2001) Phos-phoenolpyruvate carboxylase kinase is controlled by asimilar signaling cascade in CAM and C4 plants. Bio-chem Biophys Res Commun 286: 11581162

    Barkla BJ, Vera-Estrella R, Maldonaldo-Gama M, PantojaO (1999) Abscisic acid induction of vacuolar H-ATPaseactivity in Mesembryanthemum crystallinum is develop-mentally regulated. Plant Physiol 120: 811819

    Bartholomew DM, Rees DJG, Rambaut A, Smith JAC(1996) Isolation and sequence analysis of a cDNA encod-ing the c subunit of a vacuolar-type H-ATPase from theCAM plant Kalanchoe daigremontiana. Plant Mol Biol 31:435442

    Bechtold N, Ellis J, Pelletier G (1993) In plantaAgrobacterium-mediated gene transfer by infiltration ofadult Arabidopsis thaliana plants. Mol Biol Genet 316:11941199

    Bohnert HJ, Ayoubi P, Borchert C, Bressan RA, BurnapRL, Cushman JC, Cushman MA, Deyholos M, FischerR, Galbraith DW et al. (2001) A genomics approachtowards salt stress tolerance. Plant Physiol Biochem 39:295311

    Bohnert HJ, Cushman JC (2001) The ice plant cometh:lessons in abiotic stress tolerance. J Plant Growth Regul19: 334346

    Borland AM (1996) A model for the partitioning of photo-synthetically fixed carbon during the C-3-CAM transi-tion in Sedum telephium. New Phytol 134: 433444

    Borland AM, Hartwell J, Jenkins GI, Wilkins MB,Nimmo HG (1999) Metabolite control overrides circa-dian regulation of phosphoenolpyruvate carboxylase ki-nase and CO

    2fixation in crassulacean acid metabolism.

    Plant Physiol 121: 889896Borland AM, Tecsi LI, Leegood RC, Walker RP (1998)

    Inducibility of crassulacean acid metabolism (CAM) inClusia species: physiological/biochemical characteriza-

    tion and intercellular localization of carboxylation anddecarboxylation processes in three species which exhibitdifferent degrees of CAM. Planta 205: 342351

    Carter PJ, Nimmo HG, Fewson CA, Wilkins MB (1991)Circadian rhythms in the activity of a plant protein ki-nase. EMBO J 10: 20632068

    Cheng S-H, Edwards GE (1991) Influence of long photo-periods on plant development and expression of crassu-lacean acid metabolism in Mesembryanthemum crystalli-num. Plant Cell Environ 14: 271278

    Christopher JT, Holtum JAM (1996) Patterns of carbohy-drate partitioning in the leaves of crassulacean acid me-tabolism species during deacidification. Plant Physiol

    112: 393399

    Crassulacean Acid Metabolism

    Plant Physiol. Vol. 127, 2001 1445

  • 8/8/2019 Cushman 2001

    8/10

    Christopher JT, Holtum JAM (1998) Carbohydrate parti-

    tioning in the leaves of Bromeliaceae performing C3

    pho-tosynthesis or crassulacean acid metabolism. Aust J Plant

    Physiol 25: 371376

    Clough SJ, Bent AF (1998) Floral dip: a simplified method

    for Agrobacterium-mediated transformation of Arabidopsisthaliana. Plant J 16: 735743

    Cockburn W (1998) Rapid-cycling CAM: an hypotheticalvariant of photosynthetic metabolism. Plant Cell Environ

    21: 845851

    Cockburn W, Whitelam GC, Broad A, Smith J (1996) The

    participation of phytochrome in the signal transductionpathway of salt stress responses in Mesembryanthemumcrystallinum L. J Exp Bot 47: 647653

    Cushman JC, Bohnert HJ (1999) Crassulacean acid metab-

    olism: molecular genetics. Annu Rev Plant Physiol PlantMol Biol 50: 305332

    Cushman JC, Bohnert HJ (2001) Induction of crassulacean

    acid metabolism by salinity molecular aspects. In ALauchli, U Luttge, eds, Salinity: Environment, Plants,

    Molecules. Kluwer Academic Publishers, Dordrecht, TheNetherlands (in press)

    Cushman JC, Borland AM (2001) Induction of crassu-lacean acid metabolism by water limitation. Plant Cell

    Environ (in press)

    Cushman JC, Meyer G, Michalowski CB, Schmitt JM,

    Bohnert HJ (1989) Salt stress leads to the differentialexpression of two isogenes of phosphoenolpyruvate car-

    boxylase during crassulacean acid metabolism inductionin the common ice plant. Plant Cell 1: 715725

    Cushman JC, Michalowski CB, Bohnert HJ (1990) Devel-

    opmental control of crassulacean acid metabolism induc-ibility by salt stress in the common ice plant. PlantPhysiol 94: 11371142

    Cushman JC, Taybi T, Bohnert HJ (2000a) Induction ofcrassulacean acid metabolism: molecular aspects. In RC

    Leegood, TD Sharkey, S von Caemmerer, eds, Photosyn-thesis: Physiology and Metabolism. Kluwer Academic

    Publishers, Dordrecht, The Netherlands, pp 551582

    Cushman JC, Wulan T, Kuscuoglu N, Spatz MD (2000b)

    Efficient plant regeneration of Mesembryanthemum crys-

    tallinum via somatic embryogenesis. Plant Cell Rep 19:459463

    DeRocher EJ, Bohnert HJ (1993) Developmental and envi-

    ronmental stress employ different mechanisms in theexpression of a plant gene family. Plant Cell 5: 16111625

    DeRocher EJ, Harkins KR, Galbraith DW, Bohnert HJ

    (1990) Developmentally regulated systemic en-

    dopolyploidy in succulents with small genomes. Science

    250: 99101

    de Saussure T (1804) Recherches chimiques sur la vegeta-tion. Chez la V.e Nyon, Paris

    Drennan PM, Nobel PS (2000) Responses of CAM species

    to increasing atmospheric CO2

    concentrations. Plant CellEnviron 23: 767781

    Ehleringer JR, Monson RK (1993) Evolutionary and eco-logical aspects of photosynthetic pathway variation.

    Annu Rev Ecol Syst 24: 411439

    Eller BM, Ferrari S (1997) Water use efficiency of twosucculents with contrasting CO

    2fixation pathways. Plant

    Cell Environ 20: 93100Gehrig H, Faist K, Kluge M (1998a) Identification of phos-

    phoenolpyruvate carboxylase isoforms in leaf, stem, androots of the obligate CAM plant Vanilla planifolia SALIB.(Orchidaceae): a physiological and molecular approach.Plant Mol Biol 38: 12151223

    Gehrig H, Heute V, Kluge M (1998b) Towards a betterknowledge of the molecular evolution of phosphoenol-pyruvate carboxylase by comparison of partial cDNAsequences. J Mol Evol 46: 107114

    Gehrig H, Heute V, Kluge M (2001) New partial sequencesof phosphoenolpyruvate carboxylase as molecular phy-logenetic markers. Mol Phylogenet Evol 20: 262274

    Gehrig H, Taybi T, Kluge M, Brulfert J (1995) Identifica-tion of multiple PEPC isogenes in leaves of the faculta-tive crassulacean acid metabolism (CAM) plant Kalanchoeblossfeldiana Poelln. cv. Tom Thumb. FEBS Lett 377:399402

    Golldack D, Dietz K-J (2001) Salt-induced expression ofthe vacuolar H-ATPase in the common ice plant isdevelopmentally controlled and tissue specific. PlantPhysiol 125: 16431654

    Griffiths H (1989) Carbon dioxide concentrating mecha-nisms and the evolution of CAM in vascular epiphytes.In U Luttge, ed, Vascular Plants as Epiphytes: Evolutionand Ecophysiology. Springer-Verlag, Berlin, pp 4286

    Hartwell J, Gill A, Nimmo GA, Wilkins MB, Jenkins GI,Nimmo HG (1999) Phosphoenolpyruvate carboxylase ki-nase is novel protein kinase regulated at the level ofexpression. Plant J 20: 333342

    Hausler RE, Baur B, Scharte J, Teichmann T, Eicks M,Fischer KL, Flugge U-I, Schuber S, Weber A, Fischer K(2000) Plastidic metabolite transporters and their physi-ological functions in the inducible crassulacean acid me-tabolism plant Mesembryanthemum crystallinum. Plant J24: 285296

    Herppich W, Herppich M, von Willert DJ (1992) Theirreversible C

    3to CAM shift in well-watered and salt-

    stressed plants of Mesembryanthemum crystallinum is un-der strict ontogenetic control. Bot Acta 105: 3440

    Heyne B (1815) On the deoxidation of the leaves of Coltyle-don calycina. Trans Linn Soc Lond 11: 213215

    Holtum JAM, Winter K (1982) Activities of enzymes of

    carbon metabolism during the induction of crassulaceanacid metabolism in Mesembryanthemum crystallinum.Planta 155: 816

    Ishimaru K (1999) Transformation of a CAM plant, thefacultative halophyte Mesembryanthemum crystallinum by

    Agrobacterium tumefaciens. Plant Cell Tiss Org Cult 57:6163

    Keeley JE (1996) Aquatic CAM photosynthesis. In K Win-ter, JAC Smith, eds, Crassulacean Acid Metabolism: Bio-chemistry, Ecophysiology and Evolution, Vol 114.Springer-Verlag, Berlin, pp 281295

    Keeley JE (1998) CAM photosynthesis in submergedaquatic plants. Bot Rev 64: 121175

    Kore-eda S, Kanai R (1997) Induction of glucose

    Cushman

    1446 Plant Physiol. Vol. 127, 2001

  • 8/8/2019 Cushman 2001

    9/10

    6-phosphate transport activity in chloroplasts of Mesem-bryanthemum crystallinum by the C

    3-CAM transition.

    Plant Cell Physiol 38: 895901

    Kore-eda S, Yamashita T, Kanai R (1996) Induction oflight-dependent pyruvate transport into chloroplasts of

    Mesembryanthemum crystallinum by salt stress. Plant CellPhysiol 37: 257262

    Li B, Chollet R (1994) Salt induction and the partial puri-fication/characterization of phosphoenolpyruvate car-

    boxylase protein-serine kinase from an inducible crassu-lacean acid metabolism (CAM) plant, Mesembryanthemumcrystallinum L. Arch Biochem Biophys 314: 247254

    Li X, Song Y, Century K, Straight S, Ronald P, Dong X,

    Lassner M, Zhang Y (2001) A fast neutron deletionmutagenesis-based reverse genetics system for plants.

    Plant J 27: 235242

    Lu ttge U (1996) Clusia: plasticity and diversity in a genus

    of C3/CAM intermediate tropical trees. In K Winter, JAC

    Smith, eds, Crassulacean Acid Metabolism: Biochemis-try, Ecophysiology and Evolution, Vol 114. Springer-

    Verlag, Berlin, pp 296311

    Lu ttge U (2000) Light-stress and crassulacean acid metab-

    olism. Phyton (Horn) 40: 6582

    Lu ttge U, Pfeifer T, Fischer-Schliebs E, Ratajczak R (2000)

    The role of vacuolar malate-transport capacity in crassu-lacean acid metabolism and nitrate nutrition: higher

    malate-transport capacity in ice plant after crassulaceanacid metabolism induction and in tobacco under nitrate

    nutrition. Plant Physiol 124: 13351347

    Martin CE (1996) Putative causes and consequences of

    recycling CO2

    via crassulacean acid metabolism. In K

    Winter, JAC Smith, eds, Crassulacean Acid Metabolism:Biochemistry, Ecophysiology and Evolution, Vol 114.Springer-Verlag, Berlin, pp 192203

    McElwain EF, Bohnert HJ, Thomas JC (1992) Light mod-erates the induction of phosphoenolpyruvate carboxylase

    by NaCl and abscisic acid in Mesembryanthemum crystalli-

    num. Plant Physiol 99: 12611264

    Miszalski Z, Niewiadomska E, Slesak I, Lu ttge U, Kluge

    M, Ratajczak R (2001) The effect of irradiance on car-

    boxylating/decarboxylating enzymes and fumarase ac-tivities in Mesembryanthemum crystallinum L. exposed to

    salinity stress. Plant Biol 3: 1723

    Neuhaus E, Schulte N (1996) Starch degradation in chloro-

    plasts isolated from C3 or CAM (crassulacean acidmetabolism)-induced Mesembryanthemum crystallinum L.Biochem J 318: 945953

    Nimmo GA, Wilkins MB, Fewson CA, Nimmo HG (1987)Persistent circadian rhythms in the phosphorylation

    state of phosphoenolpyruvate carboxylase from Bryophyl-lum fedtschenkoi leaves and in its sensitivity to inhibition

    by malate. Planta 170: 408415

    Nimmo GA, Wilkins MB, Nimmo HG (2001a) Partial pu-

    rification and characterization of a protein inhibitor ofphosphoenolpyruvate carboxylase kinase. Planta 213:250257

    Nimmo HG (1998) Circadian regulation of a plant protein

    kinase. Cronobiol Int 15: 109118

    Nimmo HG (2000) The regulation of phosphoenolpyruvatecarboxylase in CAM plants. Trends Plant Sci 5: 7580

    Nimmo HG, Fontaine V, Hartwell J, Jenkins GI, NimmoGA, Wilkins MB (2001b) PEP carboxylase kinase is anovel protein kinase controlled at the level of expression.New Phytol 151: 9197

    Osmond CB (1978) Crassulacean acid metabolism: a curi-osity in context. Annu Rev Plant Physiol 29: 379414

    Paterson KM, Nimmo HG (2000) Effects of pH on theinduction of phosphoenolpyruvate carboxylase kinase inKalanchoe fedtschenkoi. Plant Sci 154: 135141

    Paul MJ, Loos K, Stitt M, Ziegler P (1993) Starch-degrading enzymes during the induction of CAM in

    Mesembryanthemum crystallinum. Plant Cell Environ 16:531538

    Pilon-Smits EAH, Hart Ht, van Brederode J (1996) Evo-lutionary aspects of crassulacean acid metabolism in theCrassulaceae. In K Winter, JAC Smith, eds, CrassulaceanAcid Metabolism: Biochemistry, Ecophysiology and Evo-lution, Vol 114. Springer-Verlag, Berlin, pp 349359

    Ranson SL, Thomas M (1960) Crassulacean acid metabo-lism. Annu Rev Plant Physiol 11: 81110

    Rascher U, Blasius B, Beck F, Lu ttge U (1998) Temperatureprofiles for the expression of endogenous rhythmicityand arrhythmicity of CO

    2exchange in the CAM plant

    Kalanchoe daigremontiana can be shifted by slow temper-ature changes. Planta 207: 7682

    Rockel B, Jia C, Ratajczak R, Lu ttge U (1998a) Day-nightchanges of the amount of subunit-c transcript of theV-ATPase in suspension cells of Mesembryanthemum crys-tallinum. J Plant Physiol 152: 189193

    Rockel B, Lu ttge U, Ratajczak R (1998b) Changes in mes-sage amount of V-ATPase subunits during salt-stressinduced C

    3-CAM transition in Mesembryanthemum crys-

    tallinum. Plant Physiol Biochem 36: 567573Rowley G (1978) The Illustrated Encyclopedia of Succu-

    lents. Salamander, LondonSchuber M, Kluge M (1981) In situ studies on crassulacean

    acid metabolism in Sedum acre L. and Sedum mite Gil.Oecologia 50: 8287

    Schmitt AK, Lee HSJ, Lu ttge U (1988) The response of theC

    3-CAM trees, Clusia rosea, to light and water stress: I.

    Gas exchange characteristics. J Exp Bot 39: 15811590Smith JAC, Winter K (1996) Taxonomic distribution of

    crassulacean acid metabolism. In K Winter, JAC Smith,

    eds, Crassulacean Acid Metabolism: Biochemistry, Eco-physiology and Evolution, Vol 114. Springer-Verlag, Ber-lin, pp 427436

    Steiger S, Ratajczak R, Martinoia E, Lu ttge U (1997) Thevacuolar malate transporter of Kalanchoe diagremontiana:a 32-kDa polypeptide? J Plant Physiol 151: 137141

    Taybi T, Cushman JC (1999) Signaling events leading tocrassulacean acid metabolism induction in the commonice plant. Plant Physiol 121: 545555

    Taybi T, Patil S, Chollet R, Cushman JC (2000) A minimalserine/threonine protein kinase circadianly regulatesphosphoenolpyruvate carboxylase activity in crassu-lacean acid metabolism-induced leaves of the common

    ice plant. Plant Physiol 123: 14711481

    Crassulacean Acid Metabolism

    Plant Physiol. Vol. 127, 2001 1447

  • 8/8/2019 Cushman 2001

    10/10

    Ting IP (1985) Crassulacean acid metabolism. Annu RevPlant Physiol 36: 595622

    Truesdale MR, Toldi O, Scott P (1999) The effect of ele-vated concentrations of fructose 2,6-bisphosphate on car-

    bon metabolism during deacidification in the crassu-lacean acid metabolism plant Kalanchoe daigremontiana.Plant Physiol 121: 957964

    Vogt T, Grimm R, Strack D (1999a) Cloning and expressionof a cDNA encoding betanidin 5-O-glucosyltransferase, a

    betanidin- and flavonoid-specific enzyme with high ho-mology to inducible glucosyltransferases from the So-lanaceae. Plant J 19: 509516

    Vogt T, Ibdah M, Schmidt J, Wray V, Nimtz M, Strack D(1999b) Light-induced betacyanin and flavonol accumu-lation in bladder cells of Mesembryanthemum crystallinum.Phytochemistry 52: 8392

    Williams DDF, Fleisch H (1993) Historical review of pine-apple breeding in Hawaii. Acta Hortic 334: 6776

    Wilkins MB (1992) Circadian rhythms: their origin andcontrol. New Phytol 121: 347375

    Winter K, Lu ttge U, Winter E, Troughton JH (1978) Sea-sonal shift from C

    3photosynthesis to crassulacean acid

    metabolism in Mesembryanthemum crystallinum growingin its natural environment. Oecologia 34: 225237

    Winter K, Smith JAC (1996a) An introduction to crassu-lacean acid metabolism. In K Winter, JAC Smith, eds,Crassulacean Acid Metabolism: Biochemistry, Ecophysi-ology and Evolution, Vol 114. Springer-Verlag, Berlin, pp113

    Winter K, Ziegler H (1992) Induction of crassulacean acidmetabolism in Mesembryanthemum crystallinum increasesreproductive success under conditions of drought andsalinity stress. Oecologia 92: 475479

    Zotz G, Winter K (1993) Short-term regulation of crassu-lacean acid metabolism activity in a tropical hemiepi-phyte, Clusia uvitana. Plant Physiol 102: 835841

    Cushman

    1448 Plant Physiol. Vol. 127, 2001