Tansley Review

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Review

Tansley review
Blackwell Publishing Ltd

Photosynthetic flexibility and


ecophysiological plasticity: questions
and lessons from Clusia, the only CAM
tree, in the neotropics

Author for correspondence: Ulrich Lüttge


Ulrich Lüttge
Institute of Botany, Technical University of Darmstadt, Schnittspahnstrasse 3–5, D-64 287 Darmstadt,
Tel: +49 6151163700
Fax: +49 6151164630 Germany
Email: luettge@bio.tu-darmstadt.de
Received: 12 January 2006
Accepted: 14 March 2006

Contents

Summary 7 V. The environmental control of photosynthetic flexibility 13

I. The discovery of crassulacean acid metabolism 8 VI. Phenotypic plasticity: physiotypes and morphotypes 16
(CAM) in the trees of Clusia: arrival in the
limelight of international research VII. Ecological amplitude and habitat impact 16

II. Phylogeny 8 VIII. Conclusions and outlook 21

III. Photosynthetic physiotypes 10 Acknowledgements 22

IV. Metabolic flexibility: organic acid variations 12 References 22

Summary

Key words: Clusia, crassulacean acid It is the aim of this review to present a monographic survey of the neotropical genus
metabolism (CAM), ecophysiology, Clusia on scaling levels from molecular phylogeny, metabolism, photosynthesis and
neotropics, plasticity, photosynthesis, autecological environmental responses to ecological amplitude and synecological
phylogeny. habitat impact. Clusia is the only dicotyledonous genus with real trees performing
crassulacean acid metabolism (CAM). By way of introduction, a brief historical reminis-
cence describes the discovery of CAM in Clusia and the consequent increase in interest
in studying this particular genus of tropical shrubs and trees. The molecular phylogeny
of CAM in the genus is compared with that in Kalanchoë and the Bromeliaceae. At
the level of metabolism and photosynthesis, the great plasticity of expression of
photosynthetic physiotypes, i.e. (i) C3 photosynthesis, (ii) CAM including CAM idling,
(iii) CAM cycling and (iv) C3/CAM-intermediate behaviour, as well as metabolic flexibil-
ity in Clusia is illustrated. At the level of autecology, the factors water, irradiance and
temperature, which control photosynthetic flexibility, are assessed. The phenotypic plas-
ticity of physiotypes and morphotypes is described. At the level of synecology, the
ecological amplitude of Clusia in the tropics and the relations to habitat are surveyed.

New Phytologist (2006) 171: 7–25

© The Authors (2006). Journal compilation © New Phytologist (2006)


doi: 10.1111/j.1469-8137.2006.01755.x

www.newphytologist.org 7
8 Review Tansley review

atmosphere in spite of stomatal closure. It remained for the


I. The discovery of crassulacean acid metabolism Mexicans Tinoco Ojanguren & Vazquez-Yanez (1983) to
(CAM) in the trees of Clusia: arrival in the demonstrate and clearly explain the performance of CAM in
limelight of international research Clusia for the first time, and with an additional seminal study
As we shall see, species of Clusia are handsome woody plants. by Ting et al. (1985) this became widely known and created
Some have small flowers only, but others develop rather wide interest, moving Clusia into the limelight of interna-
large beautiful flowers; for example, the flowers of Clusia tional research activities. This now allows me to present Clusia
grandiflora Engl. have a diameter of 150 mm. Nevertheless, in this review, and we are also preparing a monographic book
Clusia most likely would have remained just one genus among on Clusia covering all aspects of its biology, i.e. its taxonomy
a vast number of others in the large biodiversity of trees and and phylogeny, diversity and phytogeography, ecology and
shrubs in the tropics if it had not turned out that species of physiology, and circadian clock (Lüttge, 2006c).
Clusia are performing crassulacean acid metabolism (CAM).
This mode of photosynthesis is basically characterized by
nocturnal uptake of CO2 and dark fixation via phosphoenol-
II. Phylogeny
pyruvate carboxylase (PEPC), where the resulting fixation The question is why there are not more trees with CAM.
product, mainly malic acid, is stored in the central cell sap Similarly, although there are woody plants, shrubs and small
vacuole. Behind closed stomata and in the absence of any trees with C4 photosynthesis, there are no real large trees with
overt gas exchange, the organic acid is remobilized during this mode of photosynthesis. The two modes, C4 and CAM,
the light period and decarboxylated, and the CO2 regained share many biochemical features, particularly the primary
is refixed via ribulose-bis-phosphate carboxylase/oxygenase fixation of CO2 by PEPC and synthesis of organic acids
(Rubisco) and assimilated in the Calvin cycle. (mainly malate and citrate in CAM and malate and aspartate
Clusia is the only genus of trees with this mode of photosyn- in C4 photosynthesis) and the subsequent decarboxylation
thesis. True, there are a number of sizeable plants with CAM of the organic acids and refixation and assimilation of the
in the families of Cactaceae, Euphorbiaceae and Didieraceae CO2 by Rubisco in the Calvin cycle. There are, of course,
as well as the monocotyledonous Yucca, all of which have important differences between C4 photosynthesis and
been considered ‘fantastic trees’ (Menninger, 1967; Ellenberg, CAM which are well known and need not be reiterated here.
1981). Moreover, giant cactus communities in Venezuela have However, the observation that trees do not use phosphoenol-
been called ‘cactus forests’ (Vareschi, 1980). However, Clusia pyruvate (PEP) carboxylation as widely as CAM and C4
is the only genus of trees sensu stricto with typical dicotyledo- photosynthesis are otherwise distributed among angiosperms
nous secondary growth that performs CAM. shows an intriguing parallel between both modes of photo-
Clusia early raised the curiosity of researchers because of the synthesis. Phylogenetic studies address the question of why
peculiarities of its photosynthetic physiology. During field this is so, but we must note here, at the outset, that in both
work in February 1800 in Venezuela, Alexander von Hum- cases the question has not been answered to date.
boldt discovered that Clusia rosea Jacq. had no obvious overt CAM is widespread among the taxa of vascular plants
gas exchange in the light period but built up a high internal including ferns and fern allies (Isoëtes). There is no doubt that
gas pressure with an oxygen concentration near 40% (Krätz, CAM arose polyphyletically many times during the evolution
2001; Lüttge, 2002). In 1937, Willy Hartenburg reported of vascular plants (Lüttge, 1987). This must have been facilitated
observations from glasshouse experiments showing that on by the fact that CAM does not require any particular metabolic
bright days Clusia mexicana Vesque did not take up CO2 in trait as compared with non-CAM plants. A reorganization
the light period but even released some CO2, which was not and specific management of metabolic modules common in
explicable by respiration. However, on overcast days the CO2 the housekeeping functions of all photosynthesizing plants
exchange pattern was as expected of photosynthesizing plants are sufficient for CAM performance (Lüttge, 2005). The
(Hartenburg, 1937; Lüttge, 1995b). Neither Alexander von evolution of CAM has been specifically studied in the genus
Humboldt nor Willy Hartenburg was able to really explain Kalanchoë and the family Bromeliaceae and more recently also
their observations. Now, as we know that Clusia species are in the genus Clusia. There are conspicuous differences among
performing CAM, we realize that they saw particular features the three taxa. A comparison (Lüttge, 2004, 2005) is useful as
of CAM, where stomata are closed in the light period when it reveals interesting peculiarities of Clusia.
CO2 remobilized from nocturnally stored organic acids is The radiation centre of Kalanchoë is in the moist forests of
photosynthetically assimilated, which is associated with the eastern Madagascar (Kluge et al., 1991; Gehrig et al., 2001;
build up of high internal partial pressures of oxygen (Spalding Kluge, 2005). The genus comprises species with obligate C3
et al., 1979). Decarboxylation of nocturnally stored organic photosynthesis and obligate CAM and there are a small
acids behind closed stomata also leads to high internal CO2 number of species that are C3/CAM intermediate and can
concentrations, i.e. 2–60 times atmospheric (Lüttge, 2002), switch between the two modes of photosynthesis (Brulfert
which may even cause some diffusive loss of CO2 to the et al., 1973, 1975). The majority of the species in the genus

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 9

are CAM species. Morphological (Boiteau & Allorge-Boiteau, significantly less negative than in C3 plants. However, it has
1995) as well as molecular (Gehrig et al., 1997, 2000, 2001) been noted that C3/CAM-intermediate species often make
evidence suggests that CAM must have evolved monophylet- little use of their CAM option in the field, so that δ13C values
ically in the genus from a C3-photosynthesis ancestor during are C3-like. Although Winter & Holtum (2002) and Winter
radiation from the moist east to the drier areas up to the arid et al. (2005) have presented a calibration for deducing from
areas in the west and south-west of Madagascar, where CAM δ13C values the extent of primary dark fixation (PEPC) and
functions as an important adaptation to the limited availability light fixation (Rubisco) of CO2, respectively, for a definitive
of water. identification of a CAM potential in a species additional
Conversely, in the neotropical family of the Bromeliaceae physiological evidence is required, for example from gas
with its three subfamilies Pitcairnioideae, Bromelioideae and exchange measurements and analyses of diurnal organic acid
Tillandsioideae and very many epiphytic species, again as fluctuations. It is much more tedious to obtain such informa-
supported by morphological/anatomical, biochemical and tion for many species than to just analyse δ13C values, but it
molecular genetic evidence (Smith, 1989; Crayn et al., 2004), appears to be very important to do this for as many species of
CAM and the epiphytic habit evolved a minimum of three Clusia as possible. This would not only provide insights into
times. Epiphytism and CAM have evolved independent of the evolution of CAM within the genus. In addition, because
each other, although evidently CAM is an adaptation to the of the extraordinary inherent evolutionary plasticity in the genus
water shortage stress of the epiphytic habitat (Zotz & Hietz, and its high speciation rate (Gustafsson et al., 2006), the
2001). The last common ancestor of the Bromeliaceae was a investigation of Clusia may also advance understanding of
terrestrial C3 mesophyte. It is important to note that there has CAM evolution in general.
also been a reversion from CAM back to C3 photosynthesis In summary, the comparison of the three taxa Kalanchoë,
in some taxa of the bromelioid line following subsequent Bromeliaceae and Clusia, in which CAM phylogeny has been
radiation into less xeric habitats (Crayn et al., 2004). Diversity studied in the greatest depth, shows that the evolution of
is very large in the family in many respects but photosynthetic CAM not only occurred independently in the three taxa but
plasticity is low. There are obligate C3 and CAM species but also went rather different ways, i.e. monophyletically within
in the entire large family there appears to be only one clearly the genus Kalanchoë, polyphyletically but without genera-
C3/CAM-intermediate species, namely Guzmania monos- tion of much plasticity in the Bromeliaceae and polyphylet-
tachia (L.) Rusby ex Mez (Maxwell et al., 1994, 1995, 1999; ically with production of enormous plasticity in the genus
Maxwell, 2002). Clusia.
The subfamily Clusioideae with the genus Clusia in the The key enzyme of CAM as well as C4 photosynthesis is
family Clusiaceae has a minimum phylogenetic age of PEPC. Recently, the molecular characterization of PEPC
90 × 106 years (Gustafsson et al., 2002). Clusia is a large isoforms has been investigated in detail in C4 photosynthesis
genus of shrubs and trees with about 300–400 species (Pipoly (Engelmann et al., 2003; Svensson et al., 2003; Westhoff &
et al., 1998). The genus is considered monophyletic. However, Gowik, 2004) and also in CAM plants, with a strong phylo-
internal transcribed spacer (ITS) sequencing of nuclear genetic bias. It has been found that Kalanchoë pinnata (Lam.)
ribosomal DNA has shown that the evolution of CAM in the Pers. has seven PEPC isoforms, only one of which is CAM
genus was polyphyletic (Vaasen et al., 2002; Gehrig et al., relevant (Gehrig et al., 2005). Taybi et al. (2004) studied the
2003; Holtum et al., 2004). ITS analyses of some 80 species expression of PEPC and PEPC-kinase (PEPCK ) genes in four
of Clusia and a derived phylogenetic tree suggest that CAM different species of Clusia. PEPCKs regulate the activity of
evolved at least twice in the genus. A reversion from CAM to PEPC which is more active and less sensitive to feedback
C3 was noted in eight individual species and in one group of inhibition by malate in the phosphorylated state. The species
three species (Gustafsson et al., 2006). Thus the situation in compared were C. rosea, an obligate CAM species, Cluisa
Clusia resembles that in the Bromeliaceae and not that in minor L., a C3/CAM-intermediate species, Clusia aripoensis
Kalanchoë. However, in contrast to the Bromeliaceae, Clusia Britt., a weakly CAM-inducible species, and Clusia multiflora
is characterized by an enormous plasticity. There are very H.B.K., an obligate C3 species. They found that transcriptional
many C3/CAM-intermediate species. It appears that this control of PEPC abundance is a key factor in determining the
even may be the rule in the genus and obligate C3 and especi- genotypic capacity of CAM, while the PEPCK gene is under
ally obligate CAM species are much fewer (Holtum et al., the control of metabolic feedback, and there are no C3- or
2004). CAM-specific isoforms of the PEPCK genes. The obligate
This expectation is, however, based on a much smaller CAM species Clusia hilariana Schlecht. produces three
sample size. To some extent, the occurrence of CAM can be isoforms of PEPC, one of which is root-specific and is
screened by analysing carbon isotope ratios (δ13C) because, as probably a housekeeping isoform in this organ, while the C3
a result of the much lower 13C discrimination of PEPC (−7‰ species C. multiflora and the C3/CAM-intermediate species
relative to CO2) than of Rubisco (+27‰ relative to CO2) C. minor only have one isoform (Vaasen, 2005; Vaasen et al.,
(Ziegler, 1994), in CAM-performing plants δ13C values are 2006). For C. minor the observation that there is only one

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
10 Review Tansley review

isoform each of PEPC and PEPCK is most noteworthy to have been a starting point for CAM evolution (Guralnick
because this implies that just one isoform can support both et al., 1986; Guralnick & Jackson, 2001), but this now takes
the housekeeping functions of the C3 state and the particular us to the next section discussing photosynthetic plasticity in
requirements of the CAM state. This may facilitate C3–CAM Clusia.
switches. PEPC transcript abundance is increased in the
CAM state (Taybi et al., 2004; Vaasen, 2005; A. Vaasen, D.
Begerov, R. Hampp, Botanisches Institut, Universität Tübingen,
III. Photosynthetic physiotypes
Tübingen, Baden-Württemberg, Germany). The observation In Fig. 1, I have reduced metabolic complexity to the extent
that a single isoform each of PEPC and PEPCK can support of only using two modules, namely PEPC and organic acids
performance of both C3 photosynthesis and CAM also and Rubisco and carbohydrate ([CH2O]n), respectively, and
strengthens the speculation that this has facilitated CAM external and internal CO2 for a schematic illustration of four
evolution and that C3/CAM flexibility, which is so wide- different photosynthetic physiotypes, all of which are found
spread in the genus Clusia, has played a role in this. However, among Clusia species.
whether the absence of more isoforms in C. minor is a result 1 C3 photosynthesis, where stomata are open during the light
of a lack of evolution of more isoforms or the loss of redun- period and CO2 taken up is fixed directly via Rubisco in the light.
dant isoforms during evolution remains unresolved at 2 CAM, where stomata are open in the dark period and CO2
present. More comparative work clearly is highly desirable. taken up is fixed via PEPC leading to the production of
However, in the earlier literature it was not C3/CAM- organic acid (both malic and citric in Clusia species), organic
intermediate behaviour but ‘CAM cycling’ that was suggested acid is decarboxylated during the light period and the CO2

Fig. 1 Minimal scheme of metabolism


using two metabolic modules, namely
phosphoenolpyruvate carboxylase (PEPC)
and nocturnal organic acid accumulation
(org.ac.) and ribulosebisphosphate
carboxylase/oxygenase (Rubisco) and
daytime carbohydrate synthesis ([CH2O]n)
and two pools of inorganic carbon, namely
external and internal CO2, to illustrate the
basic features of four photosynthetic
physiotypes and their connections: from left
to right, (i) crassulacean acid metabolism
(CAM) cycling (CAM cycl), (ii) full CAM with
CAM idling (CAM idl), (iii) C3/CAM-
intermediate behaviour, and (iv) C3
photosynthesis. Dark arrows, carbon flow in
the dark; white arrows, carbon flow in the
light.

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 11

regained is refixed via Rubisco. Some direct C3-type fixation


is also possible, i.e. in the later light period, when ecological
conditions, particularly water relations, allow stomatal opening
after nocturnally accumulated organic acid is consumed.
CAM idling, as also shown in Fig. 1, is not considered a
separate physiotype but a variant of normal full CAM, where
stomata remain closed night and day when water relations are
adverse and respiratory CO2 is refixed in the dark period and
recycled to carbohydrate during the light period.
3 C3/CAM-intermediate behaviour with reversible switches
between the two modes of photosynthesis in perennial Clusia,
where leaves are used for at least two growth periods (Olivares,
1997).
4 CAM cycling, where CO2 uptake and fixation occur
mainly during the light period as in C3 photosynthesis, but
respiratory CO2 is recuperated via PEPC behind closed stomata
during the dark period and organic acid is produced which is
decarboxylated during the light period, supplementing the CO2
taken up from the atmosphere via the open stomata as substrate
for Rubisco. This use of PEPC to prevent loss of respiratory
CO2 in the dark period was thought to have been an early step
towards the evolution of CAM, as alluded to in the previous
section (Guralnick et al., 1986; Guralnick & Jackson, 2001).
The scheme illustrated in Fig. 1 symbolizes the modular
nature of photosynthetic physiotypes and can be used to
elucidate the links that might be important in long-term and
medium-term evolutionary and ontogenetic development,
respectively, as well as in short-term environmental responses
in ecophysiology. For example, (i) CAM can also involve a
strong C3 element; (ii) without light-period CO2 uptake
CAM cycling would become CAM idling.
Some representative curves of day:night CO2 exchange in
Clusia species are shown in Fig. 2. A bona fide obligate C3
species is C. multiflora and obligate CAM species are Clusia
alata Pl. et Tr., Clusia major L., C. rosea and C. hilariana. As
noted in the previous section, there are very many C3/CAM-
intermediate species in the genus Clusia. The most widely Fig. 2 CO2 gas exchange curves ( JCO ) of Clusia minor L. under
2
studied species in this respect is C. minor. Under different condi- three different conditions (A–C) and Clusia venosa Jacq. (D), Clusia
alata Pl. et Tr. (E) and Clusia major L. (F), showing performance of C3
tions with respect to a combination of different environmental photosynthesis (A, D), crassulacean acid metabolism (CAM) (B, E, F),
factors, such as irradiance during growth and a variety of with the four phases of (I) nocturnal CO2 uptake, (II) daytime
day:night temperature regimes, it produced a vast number stomatal closure and organic acid remobilization, (III) a transition with
of different responses including full CAM with CAM idling CO2 uptake in the early light period, and (IV) CO2 uptake in the later
(Haag-Kerwer et al., 1992). In another experiment, irradiance light period (Osmond, 1978), and CO2 uptake almost around the
clock and strong expression of all four CAM phases (C). In (A–C) for
during growth and nitrogen nutrition were varied. Plants C. minor, the irradiance and water vapour pressure deficit (VPD) of
grown at high irradiance without supplemental nitrogen the atmosphere in µmol m−2 s−1/mbar bar−1, respectively, were as
showed no CO2 uptake and a minimal loss of CO2 but accu- follows: (A) 1700/6.6, (B) 400/13.5, (C) 400/3.5, i.e. high irradiance
mulated appreciable amounts of citrate plus a small amount of and low VPD favoured C3 photosynthesis, medium irradiance and
malate in the dark period, while light-period CO2 uptake was high VPD favoured CAM, and medium irradiance and very low VPD
favoured CO2 uptake around the clock. Dark bars, night-time; white
substantial, i.e. there were clear indications of CAM cycling bars, daytime. (After data of Lee et al., 1989, where C. minor is
(Franco et al., 1991). Most of the experiments cited here wrongly named C. rosea, and Franco et al., 1990.)
were performed with one clone of vegetatively propagated
C. minor. The enormous phenotypic plasticity with respect to
expression of photosynthetic physiotypes of the genome is
evident.

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
12 Review Tansley review

(Franco et al., 1991). At low irradiance there was no malate


IV. Metabolic flexibility: organic acid variations and only citrate accumulation with and without supplemental
Clusia also shows flexibility at the level of metabolites in the nitrogen nutrition. At high irradiance the ratio of malate to
CAM cycle. Normally the precursor for glycolytic formation citrate was 0.22 without and 3.07 with nitrogen.
of phosphoenolpyruvate (PEP) as a CO2 acceptor for dark 3 In C. minor the ratio of malate to citrate was also affected
fixation is glucan (starch) and the acid produced is malic acid by day:night (D:N) temperature regimes and irradiance
(malate). In Clusia, however, free sugars, i.e. glucose, fructose during growth. For example, low irradiance and D:N tem-
and sucrose, are also used in addition to starch as PEP peratures of 30 : 15°C produced only malate accumulation;
precursors (Popp et al., 1987; Ball et al., 1991; Berg et al., high irradiance and D:N temperatures of 25 : 15°C or 30 : 15°C
2004). Although noteworthy, this situation is not unique as it produced high malate and low citrate accumulation; high
also occurs occasionally in other taxa, for example in the irradiance and D:N temperatures of 25 : 20°C produced
bromeliad pineapple Ananas comosus (L.) merr. (Black et al., equal malate and citrate accumulation; low irradiance and D:N
1996). More remarkable is the fact that Clusia, in addition or temperatures of 30 : 30°C produced only citrate accumulation;
alternatively to malate, can also synthesize citrate. There are high irradiance and D:N temperatures of 20 : 20°C produced
other CAM plants that nocturnally form citrate in their CAM high citrate and low malate accumulation (Haag-Kerwer
cycle but the amounts are far lower than those found in Clusia; et al., 1992).
for example, day:night oscillations of citrate concentrations Naturally one would expect that a switch in metabolism as
in some species of Kalanchoë may be up to 26 mM, and this important as a change from malate to citrate accumulation
is considered relatively high (Lüttge, 1988), but C. minor should have adaptive benefits in response to the various
and C. rosea can show values of up to 200 mM (Franco et al., environmental cues. However, the picture created by the
1992). This really amounts to a qualitative rather than only a above three examples of environmental responses remains
quantitative difference. The relative amounts of malate and very fuzzy, and we can only try to evaluate comparatively a
citrate produced can vary within the same species, clone or number of roles of malate and citrate, respectively, which I
plant depending on environmental conditions, where summarize below in eight points (Table 1; Franco et al., 1992;
again C. minor is an excellent example (Haag-Kerwer et al., Lüttge, 1996, 2006a).
1992). 1 With respect to day:night carbon balance, the CAM cycle
I can list three cases where differential responses of with citrate is futile. Nocturnal citrate accumulation does not
nocturnal malate and citrate accumulation in the CAM of contribute to carbon gain because, starting with a C6-hexose
Clusia species to environmental factors have been observed. unit as precursor, the result is C6-citrate. Conversely, two
1 Under drought conditions in experiments by Franco et al. molecules of C4-malate are formed by nocturnal fixation of
(1992) and de Mattos et al. (1999) with three Clusia species 2 CO2 per C6-hexose unit used (Lüttge, 1988).
(Clusia lanceolata Camb., C. rosea and C. minor) the ratio of 2 As it does not contribute to carbon gain, citrate, of course,
malate to citrate accumulation declined by a factor of 1.8–3.4 also does not contribute to water saving during CO2 acquisition,
after 10 to 16 d without watering. This was, however, not unlike malate synthesis by nocturnal CO2 uptake when
observed in a study by Borland et al. (1998) with C. minor, evaporative demand is much lower than during the day.
C. rosea and C. aripoensis, but in this case the ratios of malate 3 Like malate, citrate can contribute to carbon recycling
to citrate were already very low in the nondroughted controls. during the dark period, for example in CAM idling. With
2 In C. minor the ratio of malate to citrate accumulated also malate accumulation respiratory CO2 is recycled; with citrate
depended on nitrogen nutrition and irradiance during growth accumulation entire carbon skeletons are recycled.

Table 1 Potential advantages and disadvantages, respectively, of night:day change in the concentrations of malate (∆mal ) and citrate (∆citr )
during the crassulacean acid metabolism (CAM) cycle (Franco et al., 1992; Lüttge, 1996, 2006a)

Function ∆mal ∆citr

1. CO2 acquisition Yes No


2. H2O saving during CO2 acquisition Yes No
3. Carbon recycling in the dark period CO2 Carbon skeleton
4. Carbon recycling in the light period and CO2 CO2
increase in internal CO2 concentration Yes Yes, more than for malate
5. Energy budget in the dark period ATP consumption ATP consumption and production of redox power
6. Energy budget in the light period ATP consumption Larger ATP consumption and consumption of redox power
7. Osmotic changes Yes Yes, less than for malate
8. Buffer capacity Low High

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 13

4 Like malate, citrate can contribute to CO2 recycling in the (point 6) in this respect may also be considered beneficial as
light period. Potentially more CO2 is generated from citrate it contributes to energy dissipation, reducing the danger of
breakdown in the light period so that citrate can make a larger photoinhibition and photodestruction at high irradiance. In
contribution to the build-up of high internal CO2 concentrations this respect, observations in the field in Trinidad are interesting,
behind closed stomata than malate. where citrate breakdown in C. minor during the day showed
5 Energetically, citrate accumulation is marginally cheaper in a much closer correlation with irradiance as compared with
the dark period. ATP is consumed in similar amounts in both malate (Borland et al., 1996). Thus, one might expect that
malate and citrate accumulation but reduction equivalents citrate accumulation might be more favourable than malate
are generated in citrate synthesis which can be used for ATP accumulation under conditions of drought, when stomata
production in the respiration chain. partially or fully close during the dark period and internal
6 Energetically, the return of carbon from citrate to carbohy- carbon recycling and CAM idling become increasingly
drate in the light period appears to be more costly than in the important, and also because drought is often associated with
case of malate. The decarboxylation of two molecules of high-irradiance stress. However, apart from the observation
malate gives rise to two molecules of pyruvate, which can be that malate:citrate ratios during nocturnal acid accumulation
used to regenerate two molecules of triose and then one under drought conditions may decrease, the other results on
molecule of hexose via gluconeogenesis. The decarboxylation differential effects of environmental factors on malate and
of one molecule of citrate may generate three molecules of citrate accumulation, respectively, presented above do not show
CO2 and only one molecule of pyruvate. The refixation of any obvious relations to the theoretically expected benefits
CO2 to generate triose is much more costly energetically than from citrate accumulation.
gluconeogenesis starting from pyruvate. It is astonishing that such a fascinating problem as the
7 The nocturnal accumulation of malate in the vacuoles of putative functional advantage of the metabolic plasticity of
CAM plants has osmotic consequences and can support performing CAM with malate or citrate is not understood.
nocturnal water storage (Eller & Ruess, 1986; Lüttge, 1986; Even information about the metabolic pathway involved in
Ruess et al., 1988; Eller et al., 1992; Murphy & Smith, 1998). nocturnal citrate synthesis is scarce (Olivares et al., 1993), and
The CAM cycle with citrate is less effective in this respect; for we have no information on light-period citrate metabolism in
example, for one hexose unit from glucan, two osmotically Clusia, so that we are left with known general biochemistry of
active malate molecules and only one citrate molecule are intermediate metabolism (see also Holtum et al., 2005), for
formed. If the precursor is free hexose, citrate formation is example in the assessment of energy budgets, as mentioned
osmotically neutral. above. Some classical biochemical studies would be very
8 Citrate is a more effective pH-buffer compound than interesting, and rigorously systematic comparative studies
malate. The buffer capacity of the vacuole is important for varying environmental parameters may provide important
nocturnal organic acid accumulation because it determines new insights. However, current fashions in plant biology have
the trans-tonoplast proton gradient against which the proton a very different focus from motivating funding of the study of
pumps energizing acid accumulation must work (Lüttge such an exciting problem in biochemical ecology.
et al., 1981). A higher buffer capacity may therefore allow an
increased capacity for organic acid accumulation (Franco
et al., 1992). Thus, citrate accumulation in Clusia species may
V. The environmental control of photosynthetic
be an explanation of the fact that in Clusia species we observe
flexibility
the highest nocturnal accumulation of acids ever found Environmental input is always received by the phenotypes,
among CAM plants, namely 1410 mM titratable protons in and in the case of the different modes of photosynthesis these
one example of C. minor in the field in Trindiad (Borland are the photosynthetic physiotypes distinguished above
et al., 1992). (Fig. 1). There is then a possible feedback to the genotype
Hence, there is no benefit from citrate accumulation for which may respond with changes of the phenotype expressed
carbon gain and improvement of water relations (points 1, 2 (Lüttge, 2005). More straightforwardly, we can ask the
and 7 above). Conversely, citrate is useful as it could facilitate questions of which environmental factors determine the
a larger total accumulation of acids in the dark period (point reversible switches between C3 photosynthesis and CAM in
8) which contributes to the build-up of high internal CO2 Clusia species and how this works. As an example, we have
pressures in the light period, in which citrate is more effective already seen above that quite specifically the expression of
than malate (point 4). This is useful because high internal PEPC may change during a C3 to CAM transition. The major
CO2 concentrations may bring Rubisco closer to substrate external factors are water, irradiance and temperature (Lüttge,
saturation, or actually achieve substrate saturation, minimizing 2004). In the field they never act individually but always
photorespiration, photoinhibition and photodestruction at interactively. Even in experiments they are often hard to
high irradiance during the light period (Lüttge, 2002). The separate. However, individual factors may become limiting
higher energy demand of citrate cycling in the light period and we may try to consider them separately.

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
14 Review Tansley review

CAM option in plastic Clusia species. However, this is not


1. Water
always so. In well-watered C. minor an increase of the
Although the earliest environmental pressure for the irradiance from 360 to 1200 µmol m−2 s−1 suppressed CO2
evolution of CAM was probably the problem of CO2 dark fixation (Schmitt et al., 1988). In studies of the return
acquisition with the advantage of the 60 times higher CO2 from CAM to C3 photosynthesis in C. minor, de Mattos &
affinity of PEPC than Rubisco and the related function of Lüttge (2001) found that, provided water is not limiting and
CAM as a CO2-concentrating mechanism (Lüttge, 2002), it the plant can afford increased transpiration, daily photon use
is evident that the driving force for the polyphyletic evolution increases when there is unrestricted C3-like CO2 uptake,
of CAM in so many taxa of the vascular plants must have been where the plant may overcome the limitations of the storage
limited availability of water. capacity of the vacuole for nocturnal organic acid accumulation,
In the C3/CAM-intermediate C. minor, CAM can be improving its daily carbon balance.
induced by drought. In an experiment where this was done it A comparative phytotron study of the obligate C3 species
was shown subsequently that even the two opposite leaves on C. multiflora and the C3/CAM-intermediate species C. minor
one node simultaneously could perform C3 photosynthesis (Herzog et al., 1999) is interesting because it also sheds light
and CAM, respectively. The two leaves were kept in gas on the behaviour of the species and their relative use of the
exchange cuvettes at a low (6.2 mbar bar−1) and high CAM option in the field. Plants of the two species were grown
(13.1 mbar bar−1) water vapour pressure deficit (VPD) of the at a low daily dose of irradiance (4 mol m−2 d−1) and then trans-
air, respectively, i.e. subjecting the two leaves to different ferred to two higher daily doses of 24.5 and 33.5 mol m−2 d−1,
evaporative demand. When the plant was re-watered after 4 d respectively. Daily irradiance was applied in a bell-shaped
of drought, the leaf in the dry air continued to perform CAM pattern as in nature, i.e. increasing from dawn towards noon
with dominating CO2 uptake and fixation in the dark, while and decreasing again towards dusk. Figure 3 shows some
the leaf in the moist air within a few hours switched to CO2 results obtained 5 d after the transfer. We can see that the C3
uptake in the light (Schmitt et al., 1988, where C. minor is species which had maximum rates of CO2 uptake ( J CO2)
wrongly called C. rosea). between 2 and 4 µmol m−2 s−1 at 4 mol m−2 d−1 could make
While this illustrates the plasticity of a single plant, K. use of the higher irradiance of 24.5 mol m−2 d−1 for increasing
Winter and colleagues have tried to carry out a more general J CO up to 6 µmol m−2 s−1 but was strongly inhibited by the
2
survey of Clusia species, assessing water use efficiency (WUE) still higher irradiance of 33.5 mol m−2 d−1 showing only a J CO2
by carbon isotope analysis (δ13C). WUE is defined as the ratio of < 1 µmol m−2 s−1. Conversely, the C3/CAM-intermediate
of moles of CO2 fixed and assimilated to moles of water lost species which had performed C3 photosynthesis with maximum
by transpiration. WUE is high in nocturnal CO2 fixation during rates of J CO2 of c. 2–3 µmol m−2 s−1 at 4 mol m−2 d−1
CAM because CO2 acquisition in the dark occurs at low had switched to CAM under the higher daily doses of irradiance
evaporative demand. WUE is much lower in C3 photosynthesis with high potential rates of J CO2 even in the transition phase
in the light and also in CAM plants when they make use of between nocturnal CO2 uptake and daytime stomatal closure
their potential for direct CO2 fixation via Rubisco in the later in the morning and during C3-like CO2 fixation in the later
light period (Fig. 1). As the 13C discrimination of PEPC is so part of the day, with no difference between the two high daily
much lower than that of Rubisco, analysis of δ13C values doses. At 4 mol m−2 d−1 neither species was photoinhibited.
allows screening (see section II) of the extent to which plants However, under the higher irradiances they were similarly
use primary CO2 fixation via PEPC and Rubisco, respectively photoinhibited and increasingly so in the second part of the
(Borland et al., 1993), and allows this to be related to WUE. light period after irradiance was increased from 24.5 to
Winter & Holtum (2002) and Winter et al. (2005) have 33.5 mol m−2 d−1 as given by the potential quantum yield of
shown in general that δ13C values in CAM species strongly photosystem II (Fv/Fm, where Fv is the variable and Fm is the
correlate with WUE. These authors also screened 38 species maximum fluorescence of dark-adapted leaves and a ratio
of Clusia from Panamá and found that under natural con- below 0.83–0.80 indicates photoinhibition; Björkman &
ditions in the field Clusia species only sparingly make use of Demmig, 1987). Nonphotochemical quenching (NPQ) of
their PEPC option of CO2 fixation (Holtum et al., 2004) and, chlorophyll (Chl) a fluorescence of photosystem II was also
thus, the related high WUE. This underlines the importance similar for the two species, as were the levels of the xanthophyll
of other factors in addition to water in the habitats of Clusia zeaxanthin, which is involved in harmless thermal energy dis-
species. sipation under irradiance stress (Hager 1980; Demmig-Adams,
1990; Demmig-Adams & Adams, 1992; Horton et al., 1994;
Pfündel & Bilger, 1994; Schindler & Lichtenthaler, 1996).
2. Irradiance
With the putative protective function of the high internal
High irradiance increases VPD and may cause over- CO2 concentrations built up during CAM in the light period
energization of the photosynthetic light-harvesting apparatus. (see section IV), one would have expected that Fv/Fm would
Thus, one would expect that it would support the use of the have been higher and NPQ and zeaxanthin levels lower in the

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 15

Fig. 3 Photosynthetic performance of Clusia


multiflora H.B.K. (solid lines) and Clusia
minor L. (broken lines) grown at a daily
irradiance of 4 mol m−2 d−1 5 d after transfer
to high irradiance of 24.5 and 33.5 mol m−2 d−1,
respectively, in a phytotron where irradiance
was increased up to midday and then
decreased again in a bell-shaped pattern
(not shown). JCO , CO2 exchange; Fv/Fm,
2
potential quantum yield of chlorophyll a of
photosystem II; NPQ, nonphotochemical
fluorescence quenching; Z, levels of
zeaxanthin, where data are only available for
24.5 mol m−2 d−1. Dark bars, night-time;
white bars, daytime. (After data of Herzog
et al., 1999.)

C3/CAM-intermediate species after its switch to CAM than cies, under higher daily irradiance leaves became necrotic and
in the C3 species. That this is not so shows that CAM species, died. However, the plant was not principally suffering from
like C3 species, are not exempt from photoinhibition and have the high irradiance. It only suffered from the sudden change
protective mechanisms, such as thermal energy dissipation, to from low irradiance during growth to higher irradiances,
deal with irradiance and oxidative stress, as was also observed probably because acclimation of the light-harvesting appara-
in the CAM species C. hilariana (Franco et al., 1999). Measure- tus takes a long time or is totally insufficient. However,
ments by Winter et al. (1990) showed that in the early light C. multiflora then produced new leaves from dormant buds
period in CAM-performing C. rosea the level of zeaxanthin which were adapted to the new situation. Conversely, the C3/
greatly increased from much lower levels maintained in the CAM-intermediate species C. minor was not damaged after
dark period, decreased in the middle of the day when stomata the transfer to higher irradiance. It was probably the much
closed and organic acid was remobilized, and subsequently more rapid switch of the biochemical machinery to CAM
increased again as stomata opened in the later light period. combined with protective mechanisms which saved it from
The levels of violaxanthin, the double epoxide of zeaxanthin, destruction of its leaves. In fact, separate experiments showed
showed the opposite pattern during the day. With respect that the capacity to produce zeaxanthin was even higher in
to the pair of Clusia species, C. multiflora and C. minor, com- C. minor (c. 130 µmol m−2 leaf area) than in C. multiflora
pared by Herzog et al. (1999), the long-term comportment (c. 60 µmol m−2) but it did not make use of this higher capacity,
after transfer to higher irradiance is important. In the C3 spe- at least up to 5 d after transfer (Fig. 3).

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
16 Review Tansley review

These phytotron experiments suggest that the C3/CAM- photosynthetic organs, the leaves, which, although they
intermediate species may have a larger niche width, performing may have greatly differing sizes in different species, are all
well and with high short-term flexibility both under full sun somewhat leathery and succulent. Diversity is revealed
exposure and in more shaded situations, while the C3 species when we consider the physiotypes with high plasticity of
performs well under high irradiance only when it has long-term modes of photosynthesis and metabolic pathways during
development under full exposure, for example starting with CAM performance, as described in sections III and IV. At the
seedling growth. Indeed, in the field it has been observed that morphotypic level diversity is very high in the generative
C. multiflora occupies exposed sites in a secondary savanna organs (Engler, 1925; Bittrich & Amaral, 1996; Gustafsson
while C. minor prefers semi-shaded sites in a deciduous dry & Bittrich, 2003; Gustafsson et al., 2006). Flower size may
forest but can also invade the exposed sites of C. multiflora range in diameter from 5 to 150 mm. There is substantial
(Franco et al., 1994; Grams et al., 1997). diversity in floral morphology with respect to size, degree of
fusion and number of floral parts, position of staminodes and
stamen and anther morphology. Pollination rewards are
3. Temperature
nectar, pollen and resin. Floral resin is a rare award. It has
It is widely assumed that lower night temperatures and evolved polyphyletically in the genus of Clusia (Gustafsson
higher day temperatures are favourable for the performance et al., 2006). Diversity at the morphological level is also
of CAM, based on the observation of effects of temperature revealed by the different life forms of Clusia species. They may
on the overall performance of the counteracting enzymes germinate terrestrially and directly develop into independent
of nocturnal carboxylation (PEPC) and daytime decarb- free-standing trees. They may also germinate epiphytically,
oxylation of CAM, where lower temperatures favour the predominantly in humus niches between branches of host
former and higher temperatures the latter (Brandon, 1967; trees or of other epiphytes such as epiphyte nests or even the
Kluge et al., 1973; Neales, 1973; Medina et al., 1977; Kluge tanks of bromeliads. Then they can establish soil contact via
& Ting, 1978; Buchanan-Bollig & Kluge, 1981; Buchanan- aerial roots, i.e. they are hemi-epiphytes. Other adventitious
Bollig et al., 1984; Nobel, 1988; Fetene & Lüttge, 1991; roots may strangle the bark of the host tree and thus murder
Carter et al., 1995). However, in the tropics many CAM the phorophyte, the stem of which may then rot away under a
species perform well under rather high and similar night and network of strangler roots so that the previous hemi-epiphyte
day temperatures. eventually becomes a free-standing tree (Lüttge, 1991). This
Studies on species of Clusia show that temperature effects diversity does not only occur among species, because even
are often closely related to effects of irradiance (Haag-Kerwer individual species can express many of these life forms, so that
et al., 1992). We have already seen in section IV that, depending we have here another example of plasticity in the genus
on irradiance during growth, day:night temperature regimes (Fig. 4).
may strongly determine the extent to which either malate or In contrast to other hemi-epiphytes, mainly in the genus
citrate or both are used in the CAM cycle. In C. minor it was Ficus, several morphological/anatomical traits, such as leaf
found that the amplitude of the day:night (D:N) temperature thickness, and stomatal density and size, did not differ much
difference is important in eliciting a C3 to CAM switch and between the epiphytic and rooted tree life form of C. minor
this is modulated by irradiance during growth. At low irradiance, (Holbrook & Putz, 1996). Often the hydraulic architecture
plants performed C3 photosynthesis when D:N temperatures must be adapted in hemi-epiphytes as compared with rooted
were 30 : 30, 30 : 25 and 30 : 20°C and switched to CAM at trees for sufficient water supply. In a comparative study of the
30 : 15°C. A smaller D:N temperature difference was required hemi-epiphytic C3/CAM-intermediate Clusia uvitana Pittier
for the switch in high-irradiance-grown plants which performed with other hemi-epiphytes, mainly of the genus Ficus (Zotz
C3 photosynthesis at 30 : 30°C but already switched to CAM & Winter, 1994a,b; Patiño et al., 1995; Zotz et al., 1997), it
at 30 : 25°C. At 15 : 15°C gas exchange was reduced under turned out that C. uvitana had less effective unit leaf area, a
both irradiances, but high-irradiance-grown plants performed rather large leaf area per unit of stem cross-sectional area and
CAM and low-irradiance-grown plants C3 photosynthesis. a lower specific stem conductance. This suggests that physio-
The changes in the mode of photosynthesis were always readily typic flexibility with the water-saving CAM option also relates
reversible when temperature regimes were changed again. to a morphotypic trait such as hydraulic architecture.

VI. Phenotypic plasticity: physiotypes and VII. Ecological amplitude and habitat impact
morphotypes Physiological ecology has always largely been autecology. We
Clusia species are woody shrubs and trees with a dichasial have noticed, however, that it is increasingly developing the
cyme. In an earlier review (Lüttge, 1999) I proposed potential to make contributions to synecology (Lüttge &
considering all species of Clusia to belong to a single Scarano, 2004). This is in the cast of mind of Sir Arthur
morphotype. Saying this, I referred to the vegetative Tansley, whose name is heading the series of reviews to which

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 17

Fig. 4 Life forms of Clusia. Top row


(left–right): terrestrial seedlings of Clusia
multiflora; terrestrial seedlings of Clusia
minor; seedling of Clusia sp. under the bark
of a fallen tree. Centre row (left–right): Clusia
rosea in an epiphyte nest; roots of C. rosea in
the tank of the bromeliad Aechmea lingulata,
where the tank was cut open for
photography; C. rosea growing epiphytically
in a tree branch. Bottom row (left–right): two
pictures of C. rosea as a strangler and C. rosea
as a free-standing tree.

the current contribution belongs, because quite early ‘He morphotype, occupies a wide range of habitats in the neotropics,
made it clear that synecology and autecology are subsumed in as listed in Table 2 and illustrated in Fig. 5. Not only the
the study of the community by methods firstly descriptive genus as a whole but also some individual species, for example
and appreciative, and secondly, analytic and experimental’ the C3 species C. multiflora, the CAM species C. rosea and the
(Godwin, 1977). Thus, in this vein, in this section I shall C3/CAM-intermediate species C. minor, cover almost the
describe the community performance of Clusia. However, one entire range of different habitats listed in Table 2, and some
may bear in mind here that laboratory studies of autecology species occur in rather contrasting habitats, such as the coastal
mostly cover the dynamics of short-term acclimations, while restingas of Brazil and inselbergs (the C3 species Clusia parvi-
field observations represent momentary pictures of long-term flora Saldanha et Engl.) and the restingas and dry lowland
adaptations. Long-term laboratory observations are as rare as forest (the CAM species Clusia fluminensis Pl. et Tr.). In the
recordings of gradual responses under gradually varying, for following I shall list a number of sites occupied by Clusia
example seasonally, natural conditions. The continuous long- species and try to evaluate the environmental impact of the genus.
term studies in the field which I allude to below are quite rare. I shall do this by comparing the ecophysiological performance
With the great plasticity found in the genus Clusia we may of Clusia species with those of other co-occurring plants with
expect that it will span a large ecological amplitude. Indeed, similar life forms, i.e. shrubs and trees. Much field work is
it is observed that the genus, with its single simple vegetative in fact reported in the literature and an ecophysiological

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
18 Review Tansley review

Table 2 Ecological amplitude of Clusia with the range of habitats where Clusia is found, and occupation of a large number of these habitats
by three selected species with obligate C3 photosynthesis, obligate crassulacean acid metabolism (CAM) and C3/CAM-intermediate behaviour

Habitat C. muliflora (C3) C. rosea (CAM) C. minor (C3/CAM)

Restinga
Coastal rocks 
Savanna/cerrado  
Gallery forest – cerrado ecotone
Open shrubland 
Dry lowland forest  
Secondary shrub forest  
Dry montane karstic limestone forest   
Montane rainforest   
Upper montane rainforest 
Cloud forest/fog forest/elfin forest 
Inselberg

comparison can be based on traits such as carbon isotope C. hilariana (CAM) showed superior photosynthetic capacity
ratios, rates of CO2 exchange, Chl a fluorescence parameters (ETRmax) as compared with the other four species, and the C3
of photosytem II and day:night changes of organic acid levels species C. parviflora was similar to the other C3 species.
in CAM-performing species. Space does not allow me to go Chronic photoinhibition not reversible overnight was absent
into as much detail as elsewhere ( Lüttge, 1999, 2006b), but or low in all species, but an advantage of CAM in avoiding
the advantage of a briefer treatment is that it provides a more acute photoinhibition at midday was not seen. In the inter-
concise overview. mediate restinga of Barra de Maricá (22°53′ S, 42°52′ W)
C. lanceolata and C. fluminensis were studied (Roberts et al.,
1996; Reinert et al., 1997). Both are potential CAM species
1. The Brazilian restinga complex
but δ13C values indicated that C. lanceolata was mostly
The restinga complex consists of a mosaic of plant fixing CO2 by the C3 mode, while C. fluminensis was shifted
communities on the sandy coastal plains ranging from open more towards CAM. Chronic photoinhibition was absent and
formations to forest ecosystems ( Lacerda et al., 1993). The ETRmax and acute photoinhibition were similar in the two
open formation is characterized by shrub islands surrounded species so that the relative performance of the two modes of
by white sand. There are sand dune ridges with dry forest in photosynthesis did not make much difference. In the dry
between and on fixed dunes (Cirne & Scarano, 2001). Species restinga of Massambaba (22°56′ S, 42°13′ W), C. fluminensis
of Clusia are important floristic elements of the restingas, mainly performing CAM and for comparison A. legalis,
which also have been called ‘Clusia scrub’ (Ule, 1901). R. brasiliensis and M. parviflora were investigated (Duarte
According to annual rainfall, height of the ground water table et al., 2005; Geßler et al., 2005a; Scarano et al., 2005). All
and salinity of the ground water, we may distinguish wet, species showed similar photosynthetic capacity and acute
intermediate and dry restingas. The major Clusia species of photoinhibition and there was no obvious advantage of
the restingas are the C3 species C. parviflora, and the CAM CAM for C. fluminensis. In the dry forest of Buziós (22°49′ S,
species C. hilariana and C. fluminensis, the former of which is 41°59′ W), C. fluminensis was compared with three Fabaceae
much more dominant than the latter. trees, Caesalpinia echinata Lam., Caesalpinia ferrea Mart. ex
In the wet restinga of Jacarepiá (22°47′ to 22°57 ′ S, 42°20′ Tul. and Machaerium obovatum Kuhlm. et Hoehne, and the
to 42°43′ W), C. fluminensis has been compared with the C3 Euphorbiaceae Croton compressus Lam. (Geßler et al., 2005b;
Fabaceae species Andira legalis Vell. Toledo (Geßler et al., Scarano et al., 2005). C. fluminensis achieved somewhat
2005a; Scarano et al., 2005). C. fluminensis developed some- higher ETRmax than C. compressus and higher rates than the
what larger maximum apparent electron transport rates of other species. In terms of acute and chronic photoinhibition
photosynthesis (ETRmax) and with respect to photoinhibition it was not better off than the two Caesalpinia species or
the two species were very similarly affected at midday and C. compressus but was superior to M. obovatum. C. fluminensis
slightly before dawn. In the intermediate restinga of Jurubatiba in the very dry forest must be limited by the moisture regime
(22°00′ to 22°23′ S, 41°15′ to 41°45′ W), C. parviflora and (Scarano et al., 2005) and again there appeared to be no clear
C. hilariana were compared with the C3 species A. legalis, advantage of CAM at the dry site. ETRmax measurements
Rheedia brasiliensis (Mart.) Pl. et Tr. (another Clusiaceae) and (µmol m−2 s−1) indicated that C. fluminensis showed a strong
Myrsine parvifolia A. DC. (Myrsinaceae) (Franco et al., 1996; dependence on a moisture gradient from wet restinga (290) to
Duarte et al., 2005; Scarano et al., 2005). The dominant dry restinga (160) and dry forest (100).

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 19

Fig. 5 Ecological amplitude of Clusia. Top


row (left–right): coastal restinga in Brazil with
shrubs of Clusia hilariana in the centre;
free-standing tree of C. hilariana in a
semidecidous dry forest in Brazil. Second row:
Clusia criuva on the rock outcrops of an
inselberg in Brazil; C. criuva at the ecotone of
a gallery forest with the cerrado savanna
in Brazil. Third row: four sympatric Clusia
species on karstic limestone mountains in
northern Venezuela (Clusia multiflora, Clusia
rosea, Clusia alata and Clusia sp.) and roots
of Clusia in the crevasses of the limestone
rocks. Fourth row: Unidentified Clusia species
in a tropical cloud forest, Sierra Maigualida,
southern Venezuela; C. multiflora c. 60 cm
tall in an elfin forest, Cerro Santa Ana,
Paraguana peninsula, Venezuela.

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
20 Review Tansley review

Overall, the potential of Clusia species for CAM performance


may have some subtle benefits but does not appear to provide
a very obvious conspicuous advantage as compared with C3
species of similar life form. I have alluded to the fact that
Clusia may dominate the restingas in places. Moreover, several
studies have shown that Clusia species, especially C. hilariana,
can have important pioneer functions in the restingas, starting
vegetation islands on the bare sand plain and functioning as
nurse trees under the canopy of which other vegetation may
become established (Liebig et al., 2001; Dias et al., 2005). In
ecology as well as in evolution (Gould, 2002), subtle differences
can become decisive. However, functions other than photo-
synthetic capacity discussed here could be essential, for
example reproductive biology. However, in the restingas,
sexual reproduction of Clusia species is absent or very rare and
we must assume that all growth is clonal.

2. Northern coastal mountain ranges in Venezuela


In the northern coastal range of Venezuela near Caracas
(10°24′ N, 66°58′ W, 1500–1740 m asl) the same two Clusia
species, C. multiflora and C. minor, investigated in a phytotron
study in relation to the irradiance factor (section V.2) and
physiotypic plasticity described in section III were measured
in the field (Franco et al., 1994; Grams et al., 1997). In
comparison to the behaviour discussed in section V.2 the field
performance was similar (Fig. 6). The C3 species C. multiflora
potentially could achieve higher maximum rates of CO2
uptake, J CO2-max, in the light. Notwithstanding the higher
J CO -max in C. multiflora at higher irradiance, there was a
2
pronounced midday depression. Chronic photoinhibition at
dawn and acute photoinhibition at midday (Fv/Fm values Fig. 6 Photosynthetic performance of Clusia multiflora H.B.K
< 0.8) were similar in the two species. Nocturnal CO2 uptake and Clusia minor L. in the field with different daily irradiances as
in C. minor when performing CAM increased from low to indicated. JCO -max, maximum rates of CO2 uptake (positive values)
2
and CO2 release (negative values) observed in the light period (open
medium irradiance and then decreased again at the highest bars) and in the dark period (closed bars); JCO -integr, net CO2
irradiance, and this was reflected in the total integrated CO2 2
uptake integrated over 24 h; WUE, water use efficiency integrated
uptake for the dark periods and night:day oscillations of over 24 h; Fv/Fm, potential quantum yield of chlorophyll a of
organic acids (not shown). Integrated CO2 uptake over 24 h photosytem II, predawn (dark bars) and at midday (open bars). (After
was much higher under high than under low irradiance in data of Franco et al., 1994; Grams et al., 1997; Lüttge, 1999.)
C. multiflora, and in C. minor at the highest irradiance day-
time CO2 uptake was also somewhat higher and night-time
CO2 uptake was reduced as compared with the shaded plants concentrations built up during acid remobilization in the light
at the lowest irradiance (Franco et al., 1994; Grams et al., period. All four species showed slight chronic and pronounced
1997). C. minor when performing CAM always had much acute photoinhibition at dawn and midday, respectively.
higher WUE than C. multiflora.
In the Sierra de San Luis in the state Falcón (11°18′ N,
3. Semi-evergreen moist tropical forest, Barro Colorado
69°45′ W, 1200 m asl) on a karstic limestone ridge the
Island, Panama
performance of the C3 species C. multiflora was compared with
that of three CAM species, C. rosea, C. alata and an unidentified In this forest (9°10′ N, 79°51′ W) the hemi-epiphytic
species (Popp et al., 1987; Franco et al., 1994; Haag-Kerwer C. uvitana was compared with other hemi-epiphytes by Zotz
et al., 1996). The CAM species made considerable use of & Winter (1994a,b). These authors made an important
their CAM option at this site. Photosynthetic capacity was contribution to the understanding of a possible advantage
similar in all four species. However, ETRmax values were of the CAM option in C3/CAM-intermediate species. The
higher in the CAM species as a result of the high internal CO2 advantage of the CAM option in structuring hydraulic

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 21

architecture in hemi-epiphytes has been mentioned in section that not so much CAM per se but the high flexibility in carbon
VI. However, the two authors also covered an entire 12- acquisition under variable environmental conditions in the
month period with numerous measurements, one of the rare tropics inherent in CAM (Fig. 1) is its real ecophysiological
studies meeting the need for continuous long-term studies in advantage. Plasticity with segregation and separation of
the field, and thus revealing an important seasonal effect. In ecotypes based on the high ecological amplitude might relate
the wet season, with respect to both CO2 uptake integrated to the high speciation rate in the genus (Lüttge, 1995a,b,
over 24 h and maximum rates of J CO2, the performance of 2005). The radiative power of the genus is also great. This can
C. uvitana was close to that of a C3 fern Polypodium crassifolium be seen in the Atlantic rain forest complex of Brazil, where the
L. and inferior to the that of the C3 orchid Catasetum Atlantic forest itself on tertiary ground is surrounded by a
viridiflavum Hook. However, in the dry season, when C. uvitana variety of plant communities (Fig. 7), some of which are also
switched to CAM it was superior to the other two species. older relicts, such as the dry forest, but some of which are
quite young, such as the restingas and swamp forests on
tertiary ground. The dioecious species C. hilariana in these
4. Tropical forest on the island of Trinidad
communities is a migrant from the Atlantic forest with
A similar advantage of the CAM option in response to missing or extremely limited generative propagation in the
fluctuations of habitat conditions in wet and dry seasons restingas because its specific pollinators from the Euglossini
was also shown for Clusia species in Trinidad: C. minor, tribe, for example the bee Trigona spinipes, did not follow it to
C. tocuchensis, C. aripoensis and C. intertexta. This is another the new habitat (Scarano et al., 2004).
example of rather continuous long-term studies, where My monographic review on the genus Clusia may have
especially for C. minor the advantage of C3/CAM switches for shown that there is a great potential for future study of this
performance over the seasons was demonstrated (Borland single diverse genus spanning all scaling levels from molecular
et al., 1992, 1993, 1994, 1996; Roberts et al., 1998). genetics to ecosystems.
1 While the molecular phylogeny of Clusia has advanced, its
scope and data base need to be extended and the molecular
5. Inselberg, Pão de Açúcar, Rio de Janeiro
approach must also be used in population genetics.
On this inselberg (22°57′ S, 43°59′ W) the C3 species 2 Studies of metabolism should evaluate the organization
C. parviflora and the two Euphorbiaceae C. compressus and and management of metabolic reaction modules in the flexi-
Styllingia dichotoma Muell. Arg. form bushes and shrubbery. bility of photosynthetic physiotypes. Metabolomics will be
The performances of all three species are quite similar (Duarte a useful approach, but some classical biochemistry is also
et al., 2005; Scarano et al., 2005). required, for instance regarding the plasticity in carbohydrate
metabolism and the citric acid enigma.
3 More rigorously comparative quantitative laboratory and
VIII. Conclusions and outlook field studies, such as the comparison of C. multiflora and
Clusia, the only dicotyledonous tree genus with CAM, C. minor in the phytotron and in the field discussed in this
presents all facets of CAM and with its extraordinary plasticity review, should be performed to evaluate ecophysiological
it allows the study of CAM in all its ramifications. It suggests capacities and potential environmental impact.

Fig. 7 The Atlantic rain forest (AF) of Brazil


with its peripheral ecosystems of dry forest
(DF), high-altitude rock outcrops (HA),
inselbergs (I), restinga (R) and swamp forest
(SF), all occurring within a distance of only
60–80 km from the sea (S). (Modified from
Scarano, 2002.)

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
22 Review Tansley review

4 In addition to assessment of ecophysiology, reproductive Buchanan-Bollig IC, Kluge M. 1981. Crassulacean acid metabolism
biology should be advanced to foster understanding of the (CAM) in Kalanchoë daigremontiana: temperature response of
phosphoenolpyruvate (PEP)-carboxylase in relation to allosteric effectors.
performance of Clusia at the community level. Planta 152: 181–188.
Buchanan-Bollig IC, Kluge M, Müller D. 1984. Kinetic changes with
temperature of phosphoenolpyruvate carboxylase from a CAM plant.
Acknowledgements Plant, Cell & Environment 7: 63–70.
I thank Dr Annie Borland, Newcastle-upon-Tyne, UK, for Carter PJ, Wilkins MB, Nimmo HG, Fewson CA. 1995. Effects of
temperature on the activity of phosphoenolpyruvate carboxylase and on
valuable comments. the control of CO2 fixation in Bryophyllum fedtschenkoi. Planta 196:
375–380.
Cirne P, Scarano FR. 2001. Resprouting and growth dynamics after fire of
References the clonal shrub Andira legalis (Leguminosae) in a sandy coastal plain in
Ball E, Hann J, Kluge M, Lee HSJ, Lüttge U, Orthen B, Popp M, southeastern Brazil. Journal of Ecology 89: 351–357.
Schmitt A, Ting IP. 1991. Ecophysiological comportment of the Crayn DM, Winter K, Smith JAC. 2004. Multiple origins of crassulacean
tropical CAM-tree Clusia in the field. II. Modes of photosynthesis acid metabolism and the epiphytic habit in the neotropical family
in trees and seedlings. New Phytologist 117: 483 – 491. Bromeliaceae. Proceedings of the National Academy of Sciences, USA 101:
Berg A, Orthen B, de Mattos EA, Duarte HM, Lüttge U. 2004. Expression 3703–3708.
of crassulacean acid metabolism in Clusia hilariana Schlechtendal in Demmig-Adams B. 1990. Carotenoids and photoprotection: a role for
different stages of development in the field. Trees 18: 553 –558. the xanthophyll zeaxanthin cycle. Biochimica et Biophysica Acta 1020:
Bittrich V, Amaral MCE. 1996. Flower morphology and pollination biology 1–24.
of some Clusia species from the Gran Sabana (Venezuela). Kew Bulletin 51: Demmig-Adams B, Adams WW. 1992. Photoprotection and other
681–694. responses of plants to high light stress. Annual Reviews of Plant Physiology
Björkman O, Demmig B. 1987. Photon yield of O2 evolution and and Plant Molecular Biology 43: 599–626.
chlorophyll fluorescence characteristics at 77 K among vascular plants of Dias ATC, Zaluar HLT, Ganade G, Scarano FR. 2005. Canopy
diverse origins. Planta 170: 489 –504. composition influencing plant patch dynamics in a Brazilian sandy coastal
Black CC, Chen J-Q, Doong RL, Angelov MN, Sung SJS. 1996. Alternative plain. Journal of Tropical Ecology 21: 343–347.
carbohydrate reserves used in the daily cycle of crassulacean acid Duarte HM, Geßler A, Scarano FR, Franco AC, de Mattos EA, Nahm M,
metabolism. In: Winter K, Smith JAC, eds. Crassulacean acid metabolism: Rennenberg H, Rodrigues PJFP, Zaluar HTL, Lüttge U. 2005.
biochemistry, ecophysiology and evolution. Ecological Studies, Vol. 114. Ecophysiology of six selected shrub species in different plant communities
Berlin, Heidelberg, New York: Springer, 31– 45. at the periphery of the Atlantic Forest of SE Brazil. Flora 200:
Boiteau P, Allorge-Boiteau L. 1995. Kalanchoë (Crassulacées) de 456–476.
Madagascar: systématique, écophysiologie et phytochimie. Paris, France: Ellenberg H. 1981. Ursachen des Vorkommens und Fehlens von
Editions Karthala. Sukkulenten in den Trockengebieten der Erde. Flora 171: 114–169.
Borland AM, Griffiths H, Broadmeadow MSJ, Fordham MC, Maxwell C. Eller BM, Ferrari S, Ruess BR. 1992. Spatial and diel variations of water
1993. Short-term changes in carbon-isotope discrimination in the relations in leaves of the CAM-plant Senecio medley-woodii. Botanica
C3-CAM intermediate Clusia minor L. growing in Trinidad. Oecologia 95: Helvetica 102: 193–200.
444–453. Eller BM, Ruess BR. 1986. Modulation of CAM and water balance of
Borland AM, Griffiths H, Broadmeadow MSJ, Fordham MC, Maxwell C. Senecio medley-woodii by environmental factors and age of leaf. Journal of
1994. Carbon-isotope composition of biochemical fractions and the Plant Physiology 125: 295–309.
regulation of carbon balance in leaves of the C3-crassulacean acid Engelmann S, Bläsing OE, Gowik U, Svensson P, Westhoff P. 2003.
metabolism intermediate Clusia minor L. growing in Trinidad. Molecular evolution of C4 phosphoenolpyruvate carboxylase in the genus
Plant Physiology 106: 493 –501. Flaveria – a gradual increase from C3 to C4 characteristics. Planta 217:
Borland AM, Griffiths H, Maxwell C, Broadmeadow MSJ, Griffiths MN, 717–725.
Barnes JD. 1992. On the ecophysiology of the Clusiaceae in Trinidad: Engler A. 1925. Guttiferae. In: Engler A, Prantl K, eds. Die natürlichen
expression of CAM in Clusia minor L. during the transition from wet to Pflanzenfamilien, 2nd edn, Vol. 21A. Leipzig, Germany: Wilhelm
dry season and characterization of three endemic species. New Phytologist Engelmann, 154–237.
122: 349–357. Fetene M, Lüttge U. 1991. Environmental influences on carbon recycling in
Borland AM, Griffiths H, Maxwell C, Fordham MC, Broadmeadow MSJ. a terrestrial CAM bromeliad, Bromelia humilis Jacq. Journal of
1996. CAM induction in Clusia minor L. during the transition from wet Experimental Botany 42: 25–31.
to dry season in Trinidad: the role of organic acid speciation and Franco AC, Ball E, Lüttge U. 1990. Patterns of gas exchange and organic
decarboxylation. Plant, Cell & Environment 19: 655 – 664. acid oscillations in tropical trees of the genus Clusia. Oecologia 85: 108 –
Borland AM, Laszlo IT, Leegood RC, Walker RP. 1998. Inducibility of 114.
crassulacean acid metabolism (CAM) in Clusia species; Franco AC, Ball E, Lüttge U. 1991. The influence of nitrogen, light and
physiological/biochemical characterization and intercellular localization of water stress on CO2 exchange and organic acid accumulation in the
carboxylation and decarboxylation processes in three species which exhibit tropical C3-CAM tree, Clusia minor. Journal of Experimental Botany 42:
different degrees of CAM. Planta 205: 342 –351. 597–603.
Brandon PC. 1967. Temperature features of enzymes affecting crassulacean Franco AC, Ball E, Lüttge U. 1992. Differential effects of drought and light
acid metabolism. Plant Physiology 42: 977–984. levels on accumulation of citric and malic acids during CAM in Clusia.
Brulfert J, Guerrier D, Queiroz O. 1973. Photoperiodism and enzyme Plant, Cell & Environment 15: 821–829.
activity: Balance between inhibition and induction of the crassulacean Franco AC, Haag-Kerwer A, Herzog B, Grams TEE, Ball E, de Mattos EA,
acid metabolism. Plant Physiology 51: 220 –222. Scarano FR, Barreto S, Garcia MA, Mantovani A, Lüttge U. 1996. The
Brulfert J, Guerrier D, Queiroz O. 1975. Photoperiodism and enzyme effect of light levels on daily patterns of chlorophyll fluorescence and
rhythms: kinetic characteristics of the photoperiodic induction of organic acid accumulation in the tropical tree Clusia hilariana. Trees 10:
crassulacean acid metabolism. Planta 125: 33 – 44. 359–365.

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 23

Franco AC, Herzog B, Hübner C, de Mattos EA, Scarano FR, Ball E, Haag-Kerwer A, Grams TEE, Olivares E, Ball E, Arndt S, Popp M,
Lüttge U. 1999. Diurnal changes in chlorophyll a fluorescence, Medina E, Lüttge U. 1996. Comparative measurements of gas exchange,
CO2-exchange and organic acid decarboxylation in the tropical tree acid accumulation and chlorophyll a fluorescence of different species of
Clusia hilariana. Tree Physiology 19: 635 – 644. Clusia showing C3 photosynthesis, or crassulacean acid metabolism, at the
Franco AC, Olivares E, Ball E, Lüttge U, Haag-Kerwer A. 1994. same field site in Venezuela. New Phytologist 134: 215–226.
In situ studies of crassulacean acid metabolism in several sympatric Hager A. 1980. The reversible, light-induced conversions of xanthophylls in
species of tropical trees of the genus Clusia. New Phytologist 126: the chloroplast. In: Czygan FC, ed. Pigments in plants. Stuttgart, Germany:
203–211. G. Fischer, 57–79.
Gehrig HH, Aranda J, Cushman MA, Virgo A, Cushman JC, Hammel BE, Hartenburg W. 1937. Der Wasser- und Kohlensäurehaushalt tropischer
Winter K. 2003. Cladogram of Panamanian Clusia based on nuclear Regenwaldpflanzen in sommerlicher Gewächshauskultur. Jahrbücher
DNA: implications for the origins of crassulacean acid metabolism. Plant Wissenschaftlicher Botanik 85: 641–697.
Biology 5: 59–70. Herzog B, Hoffmann S, Hartung W, Lüttge U. 1999. Comparison of
Gehrig H, Gaussmann O, Marx H, Schwarzott D, Kluge M. 2001. photosynthetic responses of the sympatric tropical C3 species Clusia
Molecular phylogeny of the genus Kalanchoë (Crassulaceae) inferred from multiflora H.B.K. and the C3-CAM intermediate species Clusia minor L.
nucleotide sequences of the IST-1 and IST-2 regions. Plant Sciences 160: to irradiance and drought stress in a phytotron. Plant Biology 1:
827–835. 460–470.
Gehrig H, Brulfert J, Kluge M. 2000. In: Lourenco WR, Goodman SM, eds. Holbrook NM, Putz FE. 1996. From epiphyte to tree: differences in leaf
Diversity and endemism in Madagascar. Paris, France: Mémoires de la structure and leaf water relations associated with the transition in growth
Société de Biogéographie, 75 – 82. form in eight species of hemiepiphytes. Plant, Cell & Environment 19:
Gehrig HH, Rösicke H, Kluge M. 1997. Detection of DNA polymorphisms 631–642.
in the genus Kalanchoë by RAPD-PCR fingerprint and its relationships to Holtum JAM, Aranda J, Virgo A, Gehrig HH, Winter K. 2004. δ13C values
infrageneric taxonomic position and ecophysiological photosynthetic and crassulacean acid metabolism in Clusia species from Panama. Trees 18:
behaviour of the species. Plant Science 125: 41–51. 658–668.
Gehrig HH, Wood JA, Cushman MA, Virgo A, Cushman JC, Winter K. Holtum JAM, Smith JAC, Neuhaus HE. 2005. Intracellular transport and
2005. Large gene family of phosphoenolpyruvate carboxylase in the pathways of carbon flow in plants with crassulacean acid metabolism.
crassulacean acid metabolism plant Kalanchoë pinnata (Crassulaceae) Functional Plant Biology 32: 429–449.
characterised by partial cDNA sequence analysis. Functional Plant Biology Horton P, Ruban AV, Walters RG. 1994. Regulation of light harvesting in
32: 467–472. green plants. Indication by nonphotochemical quenching of chlorophyll
Geßler A, Duarte HM, Franco AC, Lüttge U, de Mattos EA, Nahm M, fluorescence. Plant Physiology 106: 415–420.
Rodrigues PJFP, Scarano FR, Rennenberg H. 2005b. Ecophysiology of Kluge M. 2005. Options of photosynthesis in the genus Kalanchoë : an
selected tree species in different plant communities at the periphery of the approach integrating the levels of biochemistry, ecophysiology, molecular
Atlantic Forest of SE Brazil. III. Three legume trees in a semi-deciduous taxonomy, and phytogeographic distribution of the species. Nova Acta
dry forest. Trees 19: 523 –530. Leopoldina Neue Folge 92, number 342: 195–205.
Geßler A, Duarte HM, Franco AC, Lüttge U, de Mattos EA, Nahm M, Kluge M, Brulfert J, Ravelomanana D, Ziegler H. 1991. Crassulacean acid
Scarano FR, Zaluar HLT, Rennenberg H. 2005a. Ecophysiology of metabolism in Kalanchoë species collected in various climatic zones of
selected tree species in different plant communities at the periphery of the Madagascar: a survey by δ13C analysis. Oecologia 88: 407–414.
Atlantic Forest of SE Brazil. II. Spatial and ontogenetic dynamics in Andira Kluge M, Lange OL, Eichmann M, Schmid R. 1973. Diurnaler
legalis, a deciduous legume tree. Trees 19: 510 –522. Säurerhythmus bei Tillandsia usneoides: Untersuchungen über den Weg
Godwin H. 1977. Sir Arthur Tansley: the man and the subject. Journal of des Kohlenstoffs sowie die Abhängigkeit des CO2-Gaswechsels von
Ecology 65: 1–26. Lichtintensität, Temperatur und Wassergehalt der Pflanze. Planta 112:
Gould SJ. 2002. The structure of evolutionary theory. Cambridge, MA, USA: 357–372.
Harvard University Press. Kluge M, Ting IP. 1978. Crassulacean acid metabolism. Analysis of an
Grams TEE, Haag-Kerwer A, Olivares E, Ball E, Arndt S, Popp M, ecological adaptation. Ecological Studies, Vol. 30. Berlin, Heidelberg, New
Medina E, Lüttge U. 1997. Comparative measurements of chlorophyll a York: Springer.
fluorescence, acid accumulation and gas exchange in exposed and shaded Krätz O. 2001. Alexander von Humboldt (1769–1859) auf Pflanzenjagd. In:
plants of Clusia minor L. and Clusia multiflora H.B.K. in the field. Trees 11: Palmengarten, Heft 35, Palmengarten, Frankfurt M, 33–46.
240–247. Lacerda LD, Araujo DSD, Maciel NC. 1993. Dry coastal ecosystems of the
Guralnick LJ, Jackson MD. 2001. The occurrence and phylogenetics of tropical Brazilian coast. In: Van der Maarel E, ed. Dry coastal ecosystems:
Crassulacean acid metabolism in the Portulacaceae. International Journal of Africa, America, Asia and Oceania. Amsterdam, the Netherlands: Elsevier,
Plant Science 162: 257–262. 477–493.
Guralnick LJ, Ting IP, Lord EM. 1986. Crassulacean acid metabolism in the Lee HSJ, Schmitt AK, Lüttge U. 1989. The response of the C3-CAM
Gesneriaceae. American Journal of Botany 53: 336 –345. tree Clusia rosea to light and water stress. II. Internal CO2 concentration
Gustafsson MHG, Bittrich V. 2003. Evolution of morphological diversity and water use efficiency. Journal of Experimental Botany 40:
and resin secretion in flowers of Clusia L. (Clusiaceae): insights from ITS 171–179.
sequence variation. Nordic Journal of Botany 22: 183 –203. Liebig M, Scarano FR, de Mattos EA, Zaluar HLT, Lüttge U. 2001.
Gustafsson MHG, Bittrich V, Stevens PF. 2002. Phylogeny of Clusiaceae Ecophysiological and floristic implications of sex expression in the
based on rbcL sequences. International Journal of Plant Science 163: dioecious neotropical CAM tree Clusia hilariana Schltdl. Trees 15:
1045–1054. 278–288.
Gustafsson MHG, Winter K, Bittrich V. 2006. Diversity, phylogeny and Lüttge U. 1986. Nocturnal water storage in plants having crassulacean acid
classification of Clusia. In: Lüttge U, ed. Clusia: a woody neotropical genus metabolism. Planta 168: 287–289.
of remarkable plasticity and diversity. Ecological Studies. Berlin, Heidelberg, Lüttge U. 1987. Carbon dioxide and water demand: crassulacean acid
New York: Springer. (In press.) metabolism (CAM), a versatile ecological adaptation exemplifying the
Haag-Kerwer A, Franco AC, Lüttge U. 1992. The effect of temperature and need for integration in ecophysiological work. New Phytologist 106:
light on gas exchange and acid accumulation in the C3-CAM plant Clusia 593–629.
minor L. Journal of Experimental Botany 43: 345 –352. Lüttge U. 1988. Day–night changes of citric-acid levels in crassulacean acid

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25
24 Review Tansley review

metabolism: phenomenon and ecophysiological significance. Plant, Cell & parameters using the pressure probe: extended theory and practice
Environment 11: 445– 451. of the pressure-clamp technique. Plant, Cell & Environment 21:
Lüttge U. 1991. Morphogenetische, physiologische und biochemische 637–657.
Strategien von Baumwürgern im tropischen Wald. Naturwissenschaften 78: Neales TF. 1973. Effect of temperature on the assimilation of carbon dioxide
49–58. by mature pineapple plants, Ananas comosus (L.) Merr. Australian Journal
Lüttge U. 1995a. Ecophysiological basis of the diversity of tropical plants: of Biological Science 26: 539–546.
the example of the genus Clusia. Scientia Guaianae 5: 23 –36. Nobel PS. 1988. Environmental biology of agaves and cacti. Cambridge, UK:
Lüttge U. 1995b. Clusia: ein Modellfall der ökophysiologischen Plastizität in Cambridge University Press.
einer tropischen Gattung. In: Rundgespräche der Kommission für Ökologie Olivares E. 1997. Prolonged leaf senescence in Clusia multiflora H.B.K. Trees
der Bayerischen Akademie der Wissenschaften, Vol. 10. Bayerische 11: 370–377.
Tropenforschung – Einst und jetzt. Munich, Germany: Dr. Pfeil, Olivares E, Faist K, Kluge M, Lüttge U. 1993. 14CO2 pulse-chase labelling
173–186. in Clusia minor L. Journal of Experimental Botany 44: 1527–1533.
Lüttge U. 1996. Clusia: plasticity and diversity in a genus of C3/CAM Osmond CB. 1978. Crassulacean acid metabolism: a curiosity in context.
intermediate tropical trees. In: Winter K, Smith JAC, eds. Crassulacean Annual Reviews of Plant Physiology 29: 379–414.
acid metabolism: biochemistry, ecophysiology and evolution. Ecological Patiño S, Tyree MT, Herre EA. 1995. Comparison of hydraulic architecture
Studies, Vol. 114. Berlin, Heidelberg, New York: Springer, 296 –311. of woody plants of differing phylogeny and growth form with special
Lüttge U. 1999. One morphotype, three physiotypes: Sympatric species of reference to free-standing and hemi-epiphytic Ficus species from Panama.
Clusia with obligate C3 photosynthesis, obligate CAM and C3-CAM New Phytologist 129: 125–134.
intermediate behaviour. Plant Biology 1: 138 –148. Pfündel E, Bilger W. 1994. Regulation and possible function of the
Lüttge U. 2002. CO2-concentrating: consequences in crassulacean acid violaxanthin cycle. Photosynthesis Research 42: 89–109.
metabolism. Journal of Experimental Botany 53: 2131–2142. Pipoly JJ, Kearns DM, Berry PE. 1998. Clusia. In: Berry PE, Holst BK,
Lüttge U. 2004. Ecophysiology of crassulacean acid metabolism (CAM). Yatskievych K, eds. Flora of the Venezuelan Guayana Caesalpiniaceae –
Annals of Botany 93: 629 – 652. Ericaceae, Vol. 4. St Louis, MO, USA: Missouri Botanical Garden Press,
Lüttge U. 2005. Genotypes – phenotypes – ecotypes: relations to 260–294
crassulacean acid metabolism. Nova Acta Leopoldina Neue Folge 92, Popp M, Kramer D, Lee H, Diaz M, Ziegler H, Lüttge U. 1987.
number 342: 177–193. Crassulacean acid metabolism in tropical dicotyledonous trees of the genus
Lüttge U. 2006a. Photosynthesis. In: Lüttge U, ed. Clusia: a woody Clusia. Trees 1: 238–247.
neotropical genus of remarkable plasticity and diversity. Ecological Studies. Reinert F, Roberts A, Wilson MJ, de Ribas L, Cardinot G, Griffiths H.
Berlin, Heidelberg, New York: Springer. (In press.) 1997. Graduation in nutrient composition and photosynthetic pathways
Lüttge U. 2006b. Physiological ecology. In: Lüttge U, ed. Clusia: a woody across the restinga vegetation of Brazil. Botanica Acta 110: 135–142.
neotropical genus of remarkable plasticity and diversity. Ecological Studies. Roberts A, Borland AM, Maxwell K, Griffiths H. 1998. Ecophysiology of
Berlin, Heidelberg, New York: Springer. (In press.) the C3-CAM intermediate Clusia minor L. in Trinidad: seasonal and
Lüttge U. 2006c. Clusia: a woody neotropical genus of remarkable plasticity short-term photosynthetic characteristics of sun and shade leaves. Journal
and diversity. Ecological Studies. Berlin, Heidelberg, New York: Springer. of Experimental Botany 49: 1563–1573.
(In press.) Roberts A, Griffiths H, Borland AM, Reinert F. 1996. Is crassulacean acid
Lüttge U, Scarano FR. 2004. Ecophysiology. Revista Brasileira de Bôtanica metabolism activity in sympatric species of hemi-epiphytic stranglers such
27: 1–10. as Clusia related to carbon cycling as a photoprotective process? Oecologia
Lüttge U, Smith JAC, Marigo G, Osmond CM. 1981. Energetics of malate 106: 28–38.
accumulation in the vacuoles of Kalanchoë tubiflora cells. FEBS Letters 126: Ruess BR, Ferrari S, Eller BM. 1988. Water economy and photosynthesis of
81–84. the CAM plant Senecio medley-woodii during increasing drought. Plant,
de Mattos EA, Herzog B, Lüttge U. 1999. Chlorophyll fluorescence Cell & Environment 11: 583–589.
during CAM-phases in Clusia minor L. under drought stress. Journal Scarano FR. 2002. Structure, function and floristic relationships of plant
of Experimental Botany 50: 253 –261. communities in stressful habitats marginal to the Brazilian Atlantic
de Mattos EA, Lüttge U. 2001. Chlorophyll fluorescence and organic acid Rainforest. Annals of Botany 90: 517–542.
oscillations during transition from CAM to C3-photosynthesis in Clusia Scarano FR, Cirne P, Nascimento MT, Sampaio MC, Villela D, Wendt T,
minor L. (Clusiaceae). Annals of Botany 88: 457–463. Zaluar HLT. 2004. Ecologia vegetal: integrando ecosistema,
Maxwell K. 2002. Resistance is useful: diurnal patterns of photosynthesis in comunidades, populações e organismos. In: Rocha CFD, Esteves FA,
C3 and crassulacean acid metabolism epiphytic bromeliads. Functional Scarano FR, eds. Pesquisas de longa duração na restinga de Jurubatiba:
Plant Biology 29: 679– 687. ecologia, história natural e conservação. São Carlos, Brazil: Editoria Rima,
Maxwell C, Griffiths H, Borland AM, Young AJ, Broadmeadow MSJ, 77–97.
Fordham MC. 1995. Short-term photosynthetic responses of the Scarano FR, Duarte HM, Franco AC, Geßler A, de Mattos EA, Nahm M,
C3-CAM epiphyte Guzmania monostachia var. monostachia to tropical Rennenberg H, Zaluar HLT, Lüttge U. 2005. Ecophysiology of selected
seasonal transitions under field conditions. Australian Journal of Plant tree species in different plant communities at the periphery of the Atlantic
Physiology 22: 771–778. Forest of SE Brazil. I. Performance of three different species of Clusia in an
Maxwell C, Griffiths H, Young AJ. 1994. Photosynthetic acclimation to array of plant communities. Trees 19: 497–509.
light regime and water stress by the C3-CAM epiphyte Guzmania Schindler C, Lichtenthaler HK. 1996. Photosynthetic CO2-assimilation,
monostachia: gas exchange characteristics, photochemical efficiency and chlorophyll fluorescence and zeaxanthin accumulation in field grown
the xanthophyll cycle. Functional Ecology 8: 746 –754. maple trees in the course of a sunny and a cloudy day. Journal of Plant
Maxwell K, Marrison JL, Leech RM, Griffiths H, Horton P. 1999. Physiology 148: 399–412.
Chloroplast acclimation in leaves of Guzmania monostachia in response to Schmitt A, Lee HSJ, Lüttge U. 1988. The response of the C3-CAM tree,
high light. Plant Physiology 121: 89 –95. Clusia rosea, to light and water stress. Journal of Experimental Botany 39:
Medina E, Delgado M, Troughton JH, Medina JD. 1977. Physiological 1581–1590.
ecology of CO2 fixation in Bromeliaceae. Flora 166: 137–152. Smith JAC. 1989. Epiphytic bromeliads. In: Lüttge U, ed. Vascular plants as
Menninger EA. 1967. Fantastic trees. New York, NY, USA: The Viking Press. epiphytes: evolution and ecophysiology. Ecological Studies, Vol. 76. Berlin,
Murphy R, Smith JAC. 1998. Determination of cell water-relation Heidelberg, New York: Springer, 109–138.

New Phytologist (2006) 171: 7–25 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Tansley review Review 25

Spalding MH, Stumpf DK, Ku MSB, Burris RH, Edwards GE. 1979. Molecular evidence for independent evolution of photosynthetic
Crassulacean acid metabolism and diurnal variations of internal CO2 and flexibitity. Plant Biology 4: 86–93.
O2-concentrations in Sedum praealtum DC. Australian Journal of Plant Vareschi V. 1980. Vegetationsökologie der Tropen. Stuttgart, Germany: Eugen
Physiology 6: 557–567. Ulmer.
Svensson P, Bläsing OE, Westhoff P. 2003. Evolution of C4 Westhoff P, Gowik U. 2004. Evolution of C4 phosphoenolpyruvate
phosphoenolpyruvate carboxylase. Archives of Biochemistry and Biophysics carboxylase. Genes and proteins: a case study with the genus Flaveria.
414: 180–188. Annals of Botany 93: 13–23.
Taybi T, Nimmo HG, Borland AM. 2004. Expression of Winter K, Aranda J, Holtum JAM. 2005. Carbon isotope composition and
phosphoenol-pyruvate carboxylase and phosphoenol-pyruvate carboxylase water-use efficiency in plants with crassulacean acid metabolism.
kinase genes. Implications for genotypic capacity and phenotypic plasticity Functional Plant Biology 32: 381–388.
in the expression of crassulacean acid metabolism. Plant Physiology 135: Winter K, Holtum JAM. 2002. How closely do δ13C values of crassulacean
587–598. acid metabolism plants reflect the proportion of CO2 fixed during day and
Ting IP, Lord EM, Sternberg L da SL, DeNiro MJ. 1985. Crassulacean night? Plant Physiology 129: 1843–1851.
acid metabolism in the strangler Clusia rosea Jacq. Science 229: Winter K, Lesch M, Diaz M. 1990. Changes in xanthophyll-cycle
969–971. components and in fluorescence yield in leaves of a crassulacean-
Tinoco Ojanguren C, Vazquez-Yanez C. 1983. Especies CAM in la selva acid-metabolism plant Clusia rosea Jacq., throughout a 12 hour
humeda tropical de Los Tuxtlas, Veracruz. Boletin Sociedad Botanica de photoperiod of constant irradiance. Planta 182: 181–185.
Mexico. 45: 150–153. Ziegler H. 1994. Stable isotopes in plant physiology and ecology. Progress in
Ule E. 1901. Die Vegetation von Cabo Frio an der Küste von Brasilien. Botany 56: 1–24.
Botanische Jahrbücher für Systematik. Pflanzengeschichte und Zotz G, Hietz P. 2001. The physiological ecology of vascular epiphytes:
Pflanzengeographie 28: 511–528. current knowledge, open questions. Journal of Experimental Botany 52:
Vaasen A, Begerow D, Hampp R. 2006. Phosphoenolpyruvate carboxylase 2067–2078.
(PEPC) genes in C3, CAM and C3/CAM intermediate species of the Zotz G, Patiño S, Tyree MT. 1997. Water relations and hydraulic
genus Clusia: rapid reversible C3/CAM switches are based on the C3 architecture of woody hemi-epiphytes. Journal of Experimental Botany 48:
housekeeping gene. Plant, Cell & Environment. (In press.) 1825–1833.
Vaasen A. 2005. Verwandtschaftsbeziehungen, Photosynthesemechanismen Zotz G, Winter K. 1994a. Annual carbon balance and nitrogen-use efficiency
und Phosphoenolpyruvat-Carboxylase bei der südamerikanischen in tropical C3 and CAM epiphytes. New Phytologist 126: 481–492.
Pflanzengattung Clusia. PhD thesis. Universität Tübingen, Tübingen, Zotz G, Winter K. 1994b. A one year study on carbon, water and nutrient
Germany. relationships in a tropical C3-CAM hemi-epiphyte, Clusia uvitana Pittier.
Vaasen A, Begerow D, Lüttge U, Hampp R. 2002. The genus Clusia L. New Phytologist 127: 45–60.

About New Phytologist

• New Phytologist is owned by a non-profit-making charitable trust dedicated to the promotion of plant science, facilitating projects
from symposia to open access for our Tansley reviews. Complete information is available at www.newphytologist.org.

• Regular papers, Letters, Research reviews, Rapid reports and both Modelling/Theory and Methods papers are encouraged.
We are committed to rapid processing, from online submission through to publication ‘as-ready’ via OnlineEarly – the 2004 average
submission to decision time was just 30 days. Online-only colour is free, and essential print colour costs will be met if necessary.
We also provide 25 offprints as well as a PDF for each article.

• For online summaries and ToC alerts, go to the website and click on ‘Journal online’. You can take out a personal subscription to
the journal for a fraction of the institutional price. Rates start at £109 in Europe/$202 in the USA & Canada for the online edition
(click on ‘Subscribe’ at the website).

• If you have any questions, do get in touch with Central Office (newphytol@lancaster.ac.uk; tel +44 1524 594691) or, for a local
contact in North America, the US Office (newphytol@ornl.gov; tel +1 865 576 5261).

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 7–25

You might also like