Leaf Anatomical Traits Wich Accommodate The Facultative Engagement of Crassulacean Acid Metabolism in Tropical Trees of The Genus Clusia

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Experimental Botany, Vol. 65, No. 13, pp.

3513–3523, 2014
doi:10.1093/jxb/eru022  Advance Access publication 7 February, 2014

Research paper

Leaf anatomical traits which accommodate the facultative


engagement of crassulacean acid metabolism in tropical
trees of the genus Clusia
V. Andrea Barrera Zambrano1, Tracy Lawson2, Enrique Olmos3, Nieves Fernández-García3 and

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Anne M. Borland1,4,*
1 
School of Biology, Newcastle University, Newcastle upon Tyne NE17RU, UK
2 
School of Biological Sciences, University of Essex, Colchester CO4 3SQ, UK
3 
CEBAS–CSIC Campus Universitario de Espinardo, Department of Abiotic Stress and Plant Pathology, 30100 Murcia, Spain
4 
Biosciences Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831-6407, USA

* To whom correspondence should be addressed. E-mail: Anne.Borland@ncl.ac.uk

Received 1 November 2013; Revised 17 December 2013; Accepted 19 December 2013

Abstract
Succulence and leaf thickness are important anatomical traits in CAM plants, resulting from the presence of large
vacuoles to store organic acids accumulated overnight. A higher degree of succulence can result in a reduction in
intercellular air space which constrains internal conductance to CO2. Thus, succulence presents a trade-off between
the optimal anatomy for CAM and the internal structure ideal for direct C3 photosynthesis. This study examined how
plasticity for the reversible engagement of CAM in the genus Clusia could be accommodated by leaf anatomical
traits that could facilitate high nocturnal PEPC activity without compromising the direct day-time uptake of CO2 via
Rubisco. Nine species of Clusia ranging from constitutive C3 through C3/CAM intermediates to constitutive CAM were
compared in terms of leaf gas exchange, succulence, specific leaf area, and a range of leaf anatomical traits (% inter-
cellular air space (IAS), length of mesophyll surface exposed to IAS per unit area, cell size, stomatal density/size).
Relative abundances of PEPC and Rubisco proteins in different leaf tissues of a C3 and a CAM-performing species
of Clusia were determined using immunogold labelling. The results indicate that the relatively well-aerated spongy
mesophyll of Clusia helps to optimize direct C3-mediated CO2 fixation, whilst enlarged palisade cells accommodate
the potential for C4 carboxylation and nocturnal storage of organic acids. The findings provide insight on the optimal
leaf anatomy that could accommodate the bioengineering of inducible CAM into C3 crops as a means of improving
water use efficiency without incurring detrimental consequences for direct C3-mediated photosynthesis.

Key words: CAM, Clusia, leaf anatomy, PEPC, photosynthesis, stomata.

Introduction
The photosynthetic organs of plants with crassulacean acid from the vacuole and decarboxylated during the day to satu-
metabolism (CAM) share a number of anatomical traits rate Rubisco with CO2 behind closed stomata, thereby con-
which reflect functional requirements of this metabolic spe- serving water. Positive relationships between succulence and
cialization. The generally thick and succulent leaves and the magnitude of CAM have been demonstrated for diverse
stems of CAM plants accommodate large central vacuoles phylogenetic lineages (Winter et  al., 1983; Nelson et  al.,
which serve as storage reservoirs for malic acid produced at 2005). Leaf and stem succulence can improve water storage
night from dark CO2 uptake mediated via phosphoenolpyru- and capacitance (von Willert et al., 1990; Griffiths, 2013) and
vate carboxylase (PEPC). Malic acid is subsequently released it has been suggested that possession of this anatomical trait

Published by Oxford University Press on behalf of the Society for Experimental Biology.
3514 | Zambrano et al.

might have predisposed ancestral CAM taxa towards the evo- and water storage parenchyma. To determine which leaf
lution of this photosynthetic specialization in water-limited anatomical characteristics might accomodate both noctur-
habitats (Sage, 2002). nal net CO2 uptake and direct day-time uptake of atmos-
The degree of succulence also has important consequences pheric CO2 in Clusia, measurements were compared of leaf
for net CO2 exchange. Within the undifferentiated (i.e. no succulence, specific leaf area (SLA), % IAS, Lmes/area, and
distinct layers of palisade or spongy mesophyll) leaves and/ the cell sizes of spongy mesophyll, palisade mesophyll, and
or stems that typify most CAM species, increased succulence water storage parenchyma across nine species. The species
tends to reduce internal air space (IAS) and the length of of Clusia that were studied ranged from constitutive C3
the mesophyll surface that is directly exposed to intercellular through C3/CAM intermediates to constitutive CAM The
air spaces per unit area (Lmes/area). Whilst these anatomical relative abundance of PEPC and Rubisco proteins in dif-
traits improve the carbon economy of CAM during day-time ferent leaf tissues of a C3-performing and a CAM species
decarboxylation because net CO2 efflux from the leaf is mini- of Clusia was assessed using immunogold labelling. The
mized, the reduced leaf internal conductance to CO2 curtails density and size of stomata were also measured in the nine
direct uptake of atmospheric CO2, particularly during late species to determine if the potentially lower internal con-

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


photoperiod when stomata may re-open and direct Rubisco- ductance within thicker-leaved CAM Clusia species could
mediated CO2 uptake occurs (Griffiths, 1992; Nelson and be countered by the possession of more and/or larger sto-
Sage, 2008). A perceived incompatibility between the optimal mata, thereby creating greater potential maximum stomatal
anatomy for high nocturnal PEPC activity and the internal conductance compared with the C3 species.
structure ideal for C3 photosynthesis may account for the
bimodal distribution of weak and strong CAM plants that
is indicated by δ13C across various families known to contain Materials and methods
both C3 and CAM species (Winter and Holtum, 2002; Crayn
Plant material
et al., 2004; Silvera et al., 2005).
The concept of a functional anatomical threshold for the The species of Clusia chosen for this study were based on the degree
of leaf thickness, previous reports of CAM activity, and phylogeny
dominance of C3 or C4 carboxylation in all CAM species (species were sampled from five different sections within the genus).
is confounded by tropical trees of the genus Clusia, many Species included were: C. multiflora H. B. K., an obligate C3 species
of which exhibit a dynamic range of C3 and C4 carboxyla- (Grams et al., 1998) and C. tocuchensis Britton 1921, an obligate C3
tion patterns. It has been proposed that Clusia species are species (Lüttge, 2007). Both of these species are found in section
adapted to drought and high irradiance through the use of Anandrogyne, the most derived section according to the phylogeny
of Gustafsson et  al. (2007). From section Retinostemon, C.  minor
CAM and that this is associated with the degree of leaf suc- L. a C3/CAM intermediate (Borland et al., 1992) and C. alata Pl et
culence (Borland et al., 1998). Clusia can exist in a variety of Tr. (1860), a CAM species (Lüttge, 1999), were studied. From sec-
habits (i.e. lianas, epiphytes, shrubs, trees) and habitats (i.e. tion Phloianthera, C. lanceolata Camb., a C3/CAM intermediate spe-
semi-arid coastal restingas to montane cloud forest) which cies (Roberts et  al., 1996; Lüttge, 1999) and C.  hilariana Schlecht.,
might reflect the physiological plasticity in photosynthetic a CAM species, were included. From the Omphalanthera complex,
C. aripoensis Britton 1923, a C3/CAM intermediate species (Borland
characteristics that exist across the genus (Luttge, 2006, et  al., 1998), was studied. From section Chlamydoclusia, C.  rosea
2008). The genus is monophyletic, but, within it, CAM has Jacq 1760 a CAM species (Lüttge, 2007) and C.  grandiflora Splitg
appeared and possibly been lost several times (Gehrig et al., 1842, a C3 species that has reportedly lost CAM (Gehrig et al., 2003)
2003; Gustafsson et al., 2007). Some Clusia species are fully were included.
C3 (e.g. C. multiflora), others (e.g. C. minor, C. pratensis) are Cuttings of each species, originally taken from mature plants
growing at Moorbank Botanic Garden, Newcastle University,
capable of reversibly switching between C3 photosynthesis were maintained in heated glasshouses (day-time minimum 25  °C,
and strong CAM (Winter et  al., 2008) whilst other species night-time minimum 15  °C) with supplementary lighting provided
often considered as strong CAM (e.g. C. rosea, C. quadran- in the winter months (Oct–Feb). Rooted cuttings that were between
gula) may exhibit an almost continuous range of CO2 uptake 2–3-years-old and about 80–100 cm in height were grown in 20 cm
strategies as indicated by highly variable δ13C within indi- diameter pots in a mix of commercial loam-based compost (John
Innes No. 2, Sinclair Horticulture Ltd, Lincoln, UK) and sand
vidual species (Holtum et al., 2004; Vargas-Soto et al., 2009). (3:1 v/v) and watered with a commercial nutrient mix (Phostrogen,
Thus, anatomical traits that can accommodate potentially Bayer UK, Newbury, UK) every 3 weeks. Experimental plants were
strong CAM activity without compromising direct day-time transferred to a controlled environment chamber with a 12 h pho-
uptake of CO2 by Rubisco must exist in Clusia species that toperiod, 65–75% relative humidity, 300 μmol m–2 s–1 PFD at plant
show a propensity for facultative and reversible engagement height (provided by fluorescent lamps, F100W/840, Polylux cool
white, Light Emission Technology, New York, USA) and day/night
of CAM. temperatures of 27/18 °C. The plants were maintained under these
In contrast to the majority of other CAM species, the conditions for 4–6 weeks before gas exchange measurements and
leaves of Clusia possess clearly differentiated palisade and anatomical studies commenced.
mesophyll tissues along with an adaxial layer of water stor-
age parenchyma (Borland et  al., 1998). Tissue differentia- Leaf gas exchange
tion presents the potential for variation both in cell size
CAM expression was assessed by monitoring diel gas exchange
and intercellular air space distribution through the leaf and patterns of net CO2 uptake in all nine species under compara-
for accommodating differential distribution of PEPC and ble controlled conditions. Gas exchange profiles were obtained
Rubisco proteins in palisade mesophyll, spongy mesophyll, for plants grown under well-watered conditions (water supplied
Functional leaf anatomy of Clusia | 3515

every day) and under conditions of drought (9 d without water). Microscopy Sciences) diluted 1:50 in PBS supplemented with 1%
Each plant (including soil and pot) was weighed every day follow- BSA plus 1% goat serum. The grids were washed in PBS and dis-
ing the withdrawal of watering to establish the time taken for soil tilled water. Samples were viewed with a Philips Tecnai 12 elec-
drying to occur. In all cases, there was no further weight loss from tron microscope (Philips, Eindhoven, The Netherlands) equipped
the plant/pot after 4–6 d without water. Extending the period of with a CCD SIS MegaView III camera (1280 × 1280 pixel, 12 bit).
drought beyond 9 d resulted in a gradual decline in nocturnal Morphometrical data were obtained as described by Fernández-
and day-time net CO2 uptake in all species (data not shown). Net García et  al. (2009). Gold particle densities were counted in the
CO2 uptake was measured using a compact mini cuvette system, cytoplasm for PEPC and in the chloroplast for Rubisco, as well in
Central Unit CMS-400 with a BINOS-100 infra-red gas ana- the mitochondria and vacuole as a background.
lyser, working in an open mode (Heinz Walz GmbH, Effeltrich,
Germany). A  fully expanded leaf was clamped in the cuvette,
ensuring it received full light (i.e. 300  µmol m–2 s–1) within the Stomatal measurements
growth chamber. The temperature of the cuvette was set to track The leaves of Clusia are hypostomatous (A Barrera, unpublished
environmental conditions within the growth room. Data for net observations) so in order to measure stomatal characteristics, impres-
CO2 uptake were collected every 15 min. The gas flow through the sions of the abaxial surface of the leaf, were taken using Xantropen
cuvette was maintained between 400 and 500 ml min–1 to avoid VL plus silicone impression material and hardener (Beyer Dental

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


water condensation inside the cuvette. The leaf was maintained Leverkusen, Germany) with a mixture of approximate ratio 10:1
inside the cuvette for at least 48 h to get a complete 24 h gas (Weyers and Johansen, 1985). Stomatal impressions were made for
exchange profile. Data were analysed using DIAGAS software 2–3 leaves of two plants per species. Sections were photographed
based on the area of leaf inside the cuvette. Each gas exchange under a light microscope (Leica DM RB). At least 25 stomata per
curve presented is representative of that obtained from three bio- leaf were measured in terms of dimensions and over 50 areas of
logical replicates per species. 1.65 mm2 each were used to estimate stomatal density and stomatal
index. Stomatal dimensions (pore length and width) were measured
under 40× magnification and stomatal density was measured under
Leaf anatomy 10× magnification. All the images were analysed using Image J soft-
In addition to leaf succulence (kg m–2), measurements were made of ware. The theoretical maximum stomatal conductance (GH2O) was
specific leaf area (SLA: leaf area (cm2)/dry mass (g), a trait which has estimated using the equation described previously by Lawson et al.
been reported as a good indicator of leaf tissue density and resource (1998) where:
use strategies (Garnier and Laurent, 1994). To study internal leaf
anatomy, transverse sections of 2–3 leaves of each species were [mean formula mass of air] × [effective diffusion coefficient] × [stomatal density] × [pore area]
made, following the methodology of Morison et  al. (2005). Small GH 2 O =
pieces of leaf (approximately 2–3 × 8–10 mm) were fixed in 5% glu- (pore depth) + (end correction)
taraldehyde via vacuum infiltration and left overnight at 5 °C. The
leaf pieces were then dehydrated using an ethanol series and embed-
ded in resin (LR White acrylic resin, Sigma-Aldrich, Gillingham, The mean formula of air was taken as 40.9 mol m–3 at 25 °C and the
UK). Sections of 1 μm thickness were cut using an ultra-microtome effective diffusion coefficient for water vapour as 2.49 × 10–5 m2 s–1
and stained with toluidine blue. Images of each section were pho- at 25 °C. Pore area was calculated assuming the pore as an ellipse,
tographed under 5×, 10×, and 40× magnifications in a light micro- with a constant minor axis (pore length) at different apertures and
scope (Leica DM RB, Leica Microsystems CMS GmbH, Wetzlar, maximal aperture as a major axis. The end correction was (pore
Germany) in order to measure mesophyll thickness, [i.e. depth of area/π)0.5.
palisade and spongy mesophylls and of water storage parenchyma All statistical analyses were performed using Minitab 16 (Minitab
(WSP)], cell size (i.e. cell area), % intercellular air space (IAS), and Ltd. Coventry, UK).
length of mesophyll surface exposed to IAS per unit area (Lmes/area,
μm–1). All images were analysed using Image J software (National
Institutes of Health, USA). Results
Leaf gas exchange profiles
Immunolabelling for PEPC and Rubisco in leaf sections
To localize and determine the relative abundance of PEPC and Patterns of net CO2 uptake were measured for mature leaves
Rubisco in WSP and in palisade and spongy mesophyll cells, trans- of all nine Clusia species over an entire 24 h day/night cycle
verse sections of leaves of C. aripoensis (weak CAM-inducible but under well-watered and droughted conditions (Fig. 1). Three
operating in the C3 mode when sampled) and C. rosea (CAM) were species, C. minor, C. lanceolata, and C. aripoensis showed an
taken. Previous trials had shown that these two species presented increase in net dark CO2 uptake after drought and so were
better tissue preservation after fixation than the other C3 and CAM
species investigated in this study. categorized as C3/CAM intermediates (Fig.  1d–f; Table  1).
Leaf pieces (1 × 1 mm) were fixed in 2.5% paraformaldehyde and There was no increase in net dark CO2 uptake in C. hilariana,
0.5% glutaraldehyde in 0.06 M phosphate buffer (pH 7.2) for 2 h at C. alata or C. rosea after 9 d of drought so these species were
4 °C, rinsed in the same buffer and dehydrated in an ethanol series. categorized as CAM (Fig. 1a–c). Of the three CAM species,
Finally, samples were embedded in LR White resin (Fernández- C.  rosea engaged in the most day-time net CO2 take during
García et al., 2009).
Ultra-thin sections (60–70 nm) were obtained with a Leica EM phases 2 and 4 (Fig. 1c). Drought did not elicit any net dark
UC6 ultra-microtome (Leica Mikrosysteme, Vienna, Austria) and CO2 uptake or accumulation of acidity (data not shown) in
collected on formvar-coated nickel grids. The grids were placed in C. multiflora, C. tocuchensis or C. grandiflora so these species
phosphate-buffered saline (PBS) with 5% bovine serum albumin were categorized as C3 (Table 1). The proportion of net 24 h
(BSA) for 30 min at room temperature and then incubated for 2 h CO2 uptake that occurred at night was calculated by integrat-
with the primary antibody, anti-PEPC or Rubisco rabbit polyclonal
immunoglobulin G, diluted 1:250 in PBS containing 5% BSA. The ing the relevant areas under each of the gas exchange curves
sections were washed three times in PBS and incubated for 2 h with and expressed as a percentage of total 24 h net CO2 uptake
the secondary antibody (goat anti- rabbit, IgG, 10 nm, Electron (Table 1).
3516 | Zambrano et al.

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Fig. 1.  Gas exchange curves showing 24 h patterns of net CO2 uptake for nine species of Clusia under well-watered (closed symbols) and droughted
(9 d without water, open symbols) conditions. The shaded areas in each graph represent the dark period and each of the gas exchange curves presented
is representative of three replicate runs for each species.

Table 1.  The percentage of 24 h net CO2 uptake that occurred at night in nine species of Clusia under well-watered conditions and after
9 d without water

The data were calculated from the gas exchange profiles illustrated in Fig. 1, from which species were categorized as CAM, C3/CAM or C3.
Values for leaf succulence and specific leaf area are also shown as the mean of six biological replicates ±standard error of the mean.

Species Net 24 h CO2 uptake Photosynthetic Leaf succulence Specific leaf area
over night (%) category (kg m–2) (cm g–1 dry mass)
C. alata 74 ± 5 CAM 1.641 ± 0.018 91 ± 9
–H2O 89 ± 8
C. hilariana 88 ± 6 CAM 1.018 ± 0.019 112 ± 14
–H2O 92 ± 5
C. rosea 41 ± 10 CAM 0.903 ± 0.017 118 ± 20
–H2O 62 ± 11
C. minor 21 ± 7 C3/CAM 0.693 ± 0.032 132 ± 18
–H2O 77 ± 10
C. lanceolata 14 ± 4 C3/CAM 0.597 ± 0.020 160 ± 15
–H2O 72 ± 10
C. aripoensis 0 C3/CAM 0.569 ± 0.007 190 ± 17
–H2O 31 ± 4
C. grandiflora –8 ± 2 C3 0.512 ± 0.013 160 ± 15
–H2O –10 ± 1
C. tocuchensis –2 ± 1 C3 0.531 ± 0.011 196 ± 20
–H2O –3 ± 1
C. multiflora –5 ± 1 C3 0.554 ± 0.021 212 ± 22
–H2O –7 ± 1
Functional leaf anatomy of Clusia | 3517

Leaf anatomy
The CAM species of Clusia had higher values for leaf succu-
lence (kg m–2) than the C3/CAM intermediates which, in turn,
were higher than those of C3 species (Table 1). Specific leaf
area showed an inverse relationship with succulence and with
the proportion of CO2 taken up at night under well-watered
conditions across the nine Clusia species (P<0.01; Fig.  2a).
Thus, leaves of CAM Clusias were thicker than those of the
C3/CAM species which were thicker than those of the C3
species.
There was a weak (but not significant, P>0.1) negative rela-
tionship between the percentage of internal air space (% IAS)
in the leaf mesophyll and the propensity for dark CO2 uptake

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


(Fig. 2b). The length of the mesophyll surface exposed to IAS
per unit area (Lmes/area) was found to show a strong negative
relationship (P<0.01) with the magnitude of CAM in the nine
species of Clusia (Fig.  2c). Those species which engaged in
more dark CO2 uptake had lower Lmes/area compared with the
species which engaged in more day-time net CO2 uptake. The
values of Lmes/area calculated for the nine Clusia species were
compared with the mean Lmes/areas calculated by Nelson et al.
(2005) for a phylogentically diverse assemblage of 18 CAM
species and six C3 species (Fig. 2c). The Lmes/areas of all the
Clusia species were higher than the reported CAM average
(Fig.  2c). Values for Lmes/area in most of the Clusia species
examined in the present study (including all the C3/CAM spe-
cies as well as the CAM-performing C. rosea) were in the range
of Lmes/areas reported for C3 species (Nelson et al., 2005).
The leaves of Clusia possess clearly differentiated layers of
water storage parenchyma, palisade mesophyll, and spongy
mesophyll and the proportions of these tissues vary between
photosynthetic types as illustrated in Fig. 3. The relative con-
tributions from the different tissue types to leaf thickness was
quantified by measuring the depth of the layers of water stor-
age parenchyma (WSP), palisade mesophyll, and spongy mes-
ophyll and expressing this as a percentage of leaf thickness
(Fig. 4a, c, e). The size of the cells in each of the three tissue
types was also measured (Fig.  4b, d, f). Each of these ana-
tomical characteristics was plotted against the propensity for
CAM (measured as the proportion of CO2 taken up at night).
In those species that engaged in a greater proportion of dark
CO2 uptake, the palisade mesophyll represented a greater
percentage of leaf thickness (40–45%) compared with the C3
species of Clusia where palisade mesophyll constituted ~25–
Fig. 2.  Relationships between the proportion of net CO2 uptake
30% of leaf thickness (Fig. 4a, P=0.02). Cell size of palisade conducted at night in nine species of Clusia and (a) specific leaf area,
mesophyll also showed a positive relationship with the pro- (b) intercellular air space, and (c) Lmes/area (mean±SE, n=3). Average Lmes/
pensity for CAM (Fig. 4b, P<0.01). The cells of the palisade areas for C3 and CAM species were taken from Nelson et al. (2005) and
mesophyll in the four species that showed the highest rates of are shown in relation to Lmes/areas of the nine Clusia species (c).
net dark CO2 uptake (i.e. C. alata, C. hilariana, C. rosea, and
droughted C. minor; Fig. 1) were >4× larger than those of the
C3 species and the weak C3/CAM intermediate, C. aripoensis CAM species as reported by Nelson et al. (2005). The spongy
(Fig. 4b). The size of palisade cells in the C3 Clusias and the mesophyll constituted ~40–60% of leaf thickness across all
weak C3/CAM intermediate C. aripoensis fell within the range the different photosynthetic types and there was a trend (not
of average cell size for C3 species as reported by Nelson et al. statistically significant) for increased cell size of the spongy
(2005). The size of palisade cells of the CAM Clusias and mesophyll in those species which engaged in more dark CO2
the C3/CAM intermediate C.  minor cells fell just below the uptake (Fig. 4c, d). A weak inverse correlation (P=0.15) was
range of average cell size reported for a diverse range of other found between the magnitude of CAM and the percentage
3518 | Zambrano et al.

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Fig. 3.  Transverse leaf sections from three species of Clusia with different modes of photosynthetic metabolism and indicating the major tissue types
within the leaf that include water storage parenchyma, palisade mesophyll, and spongy mesophyll. (This figure is available in colour at JXB online.)

of total mesophyll occupied by WSP (Fig.  4c). The highest The maximum measured stomatal conductance (gH2O) of
percentage of water storage parenchyma was found in the C3 all nine Clusia species (measured at the start of the photo-
species C. multiflora, C. tocuchensis, and the weak C3/CAM period) was substantially lower than the maximum stomatal
species C. aripoensis. However, the C3 species C. grandiflora conductance that was predicted (GH2O) based on stomatal
(a species that is believed to have lost CAM) had a signifi- size and density (Fig. 7a, b). There was a negative relationship
cantly reduced percentage of WSP compared with the other between measured maximum stomatal conductance and the
C3 species. There was no relationship between the propensity propensity to engage in dark CO2 uptake (Fig. 7a). However,
for CAM and cell size of the WSP (Fig. 4f). there was no clear trend across the photosynthetic types in
terms of the potential for water loss based solely on stomatal
dimensions and density on the leaf (Fig. 7b).
Immunolocalization of PEPC and Rubisco
The abundance of Rubisco protein was 1.5 times higher in
cells of the palisade mesophyll compared with cells of the Discussion
spongy mesophyll in both C.  aripoensis operating in the C3 Succulence and leaf thickness are important anatomical
mode (t=4.436, P<0.0001) (Fig.  5a) and in the CAM spe- traits in CAM plants, resulting from the presence of large
cies C.  rosea (t= 3.619, P=0.001) (Fig.  5b). The abundance vacuoles to store organic acids accumulated during the night.
of PEPC was slightly higher in spongy mesophyll cells in A  higher degree of succulence means less intercellular air
C. aripoensis compared with palisade cells but was three times space (IAS) between mesophyll cells and a reduction in the
higher in palisade cells compared with spongy mesophyll cells length of mesophyll exposed to intercellular air spaces (Lmes/
in C. rosea (t=6.770, P<0.0001). As expected, the abundance area), traits that reduce internal conductance to CO2. Thus,
of PEPC protein was higher in C. rosea (Fig. 5a) compared succulence presents a ‘trade-off’ between the optimal leaf
with C.  aripoensis (Fig.  5b). Rubisco and PEPC proteins anatomy for CAM and the internal structure ideal for C3
were not detectable in cells of the WSP using immunogold photosynthesis. The aim of the present study was to investi-
labelling. gate how plasticity for the reversible engagement of CAM in
the genus Clusia could be accommodated by leaf anatomical
Stomatal characteristics traits that, in principle, could facilitate high nocturnal PEPC
activity without compromising the direct day-time uptake of
Stomatal density showed a significant (P<0.01) negative cor- CO2 via Rubisco.
relation with the propensity for dark CO2 uptake across the
nine species of Clusia (Fig. 6a) but there was no clear rela-
Specific leaf area and implications for internal
tionship between photosynthetic type and stomatal index
conductance to CO2 in Clusia
(Fig.  6b). There was a trend for increased stomatal pore
size in those species that engaged in more dark CO2 uptake As predicted, Clusia species with more succulent and
(Fig. 6c) but the C3 species C. grandiflora (reported to have thicker, denser leaves (lower specific leaf area, SLA) engaged
lost CAM) was an outlier on this trend (Fig. 6c). in more dark CO2 uptake than the thinner-leaved species.
Functional leaf anatomy of Clusia | 3519

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Fig. 4.  Relationships between the proportion of net CO2 uptake conducted at night in nine species of Clusia and the percentage of leaf thickness
attributed to (a) palisade mesophyll, (c) spongy mesophyll, and (e) water storage parenchyma. Also shown are cell sizes for (b) palisade mesophyll, (d)
spongy mesophyll, and (f) water storage parenchyma (mean ±SE, n=3). Average cell sizes for C3 and CAM species were taken from Nelson et al. (2005)
and are shown in relation to palisade cell size of the nine Clusia species (b). The dotted circle in (b) highlights the four Clusia species with the higest rates
of net dark CO2 uptake.

SLA serves as a useful indicator of the way in which plants The trend of thicker leaves in CAM Clusias was accompa-
invest carbon and nutrients (dry biomass) in a given area of nied by a reduced length of mesophyll surface exposed to IAS
light-intercepting foliage (Vendramini et al., 2002). Species per unit area (Lmes/area). This observation is consistent with
with lower SLA may be considered to incur a higher leaf- the hypothesis that, by constraining leaf internal conductance
level cost for light interception (Poorter, 2009), a strategy to CO2, a reduced Lmes/area will enhance photosynthetic effi-
that is common in species that inhabit environments where ciency during decarboxylation in phase III of CAM because
drought and/or nutrient limitation hamper growth. Thus, net CO2 efflux from the leaf is minimized (Maxwell et  al.,
possession of a low SLA may be a trait that predisposed 1997). Thus, the propensity for CAM in Clusia showed a clear
Clusia species towards the evolution of CAM in water and relationship with Lmes/area. Previous measurements gathered
nutrient-limited habitats. for a phylogentically diverse group of CAM species indicated
3520 | Zambrano et al.

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Fig. 5.  The relative amounts of PEPC and Rubisco in terms of the density of gold particles in cells of the palisade mesophyll and spongy mesophyll for
C. aripoensis operating in the C3 mode (a) and the CAM species C. rosea (b). Gold particle densities were also counted in the mitochondria and vacuole
as background.

that photosynthetic divergence between weak and strong with the other C3/CAM and C3 Clusia species. Moreover,
CAM is mediated by % IAS and Lmes/area which presents a immunolocalization studies showed that the enhanced PEPC
functional threshold for predominantly Rubisco or predomi- abundance in C.  rosea compared with the weak C3/CAM
nantly PEPC-mediated net CO2 uptake (Nelson et al., 2005). intermediate C. aripoensis was predominantly associated with
However, the CAM Clusia species investigated in the present the palisade mesophyll tissue which contained 3× more PEPC
work showed substantially higher % IAS (mean of 30%) com- protein compared with the spongy mesophyll. Enhanced
pared with the diverse CAM lineages investigated by Nelson development of palisade mesophyll tissue is commonly found
et al. (2005) where average % IAS was ~15%. Similarly, the in thicker-leaved C3 species and is hypothesized to improve
measured values for Lmes/area in all of the nine Clusia species the harvesting of light, thereby helping to offset the increased
were higher than those of the CAM species studied by Nelson investment in biomass in thicker leaves (Smith and Hughes,
et  al. (2005). In C.  alata and C.  hilariana which conducted 2009). In addition, enhanced development of the palisade
70–90% of 24 h uptake at night, the average value of Lmes/area mesophyll could potentially accommodate the increased ener-
was 50% higher than the mean value for CAM species studied getic requirements of CAM. In CAM Clusias, the strategy of
by Nelson et al. (2005). The Lmes/area for the remaining seven investment in palisade mesophyll cells, alongside a relatively
Clusia species (including the CAM species C. rosea and the high Lmes/area could maintain a capacity for direct C3 car-
C3/CAM intermediate C.  minor which is capable of strong boxylation whilst the enlarged palisade mesophyll cells would
CAM activity under drought) fell within the range measured also offer potential for significant C4 carboxylation by serving
for C3 species (Nelson et  al., 2005). The ‘C3-like’ Lmes/area to accommodate more PEPC protein and large vacuoles to
in Clusia leaves indicates a relatively well-aerated mesophyll store organic acids overnight. In principle, such anatomical
which may be a critical anatomical component underpinning features could accommodate the dynamic range of photosyn-
the notable phenotypic plasticity of C3/CAM engagement thetic manifestations that distinguish Clusia from many other
within the genus. CAM lineages.

Development of the palisade mesophyll is linked to Water storage parenchyma and propensity for CAM in
propensity for CAM in Clusia Clusia
Whilst the relatively high (compared with other CAM species) In contrast to the close positive relationship that was estab-
Lmes/area of Clusia could help to optimize direct day-time lished between leaf thickness and the depth of palisade
C3-mediated CO2 uptake, adequate vacuolar storage capac- mesophyll tissues within Clusia, there was no such positive
ity is required to facilitate C4 carboxylation in those species relationship with water storage parenchyma (WSP). In fact,
that engage in constitutive or facultative CAM. The thicker the presence of a thick layer of WSP was particularly evi-
leaves (lower SLA) of the Clusia species that engaged in more dent in the C3 species C. tocuchensis and C. multiflora, both
nocturnal net CO2 uptake showed enhanced development of of which belong to the most derived section of the genus in
the palisade mesophyll, both in terms of cell size and num- which CAM does not appear to have evolved (Gustaffson
ber of cell layers. The cells of the palisade mesophyll were, et  al., 2007). Comparisons of WSP in the genus Peperomia
on average, ~4× larger in the four Clusia species that showed have shown that the thickness of this tissue shows a nega-
the highest rates of nocturnal net CO2 uptake compared tive relationship with CAM expression (Sipes and Ting,
Functional leaf anatomy of Clusia | 3521

in C3 species was, however, noted in C. grandiflora, a C3 spe-


cies believed to have lost CAM and belonging to section
Chlamydoclusia (Gustaffson et al., 2007). The depth of WSP
in C. grandiflora was comparable with that in the CAM spe-
cies C. rosea which also belongs to section Chlamydoclusia.
A  more detailed examination of the depth of WSP across
more species and sections of the Clusia genus will be neces-
sary to establish potential relationships between the thickness
of the WSP, phylogeny, and photosynthetic mode.

Stomatal traits and propensity for CAM in Clusia


It was hypothesized that the potentially lower internal con-
ductance to CO2 within the thicker leaves of CAM-performing

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


species of Clusia could be countered by the possession of more
and/or bigger stomata compared with the thinner-leaved C3
species. Previous studies have reported that, in general, more
succulent species show lower stomatal densities than less suc-
culent species (Sayed, 1998; Luttge, 2008). The present study
was in agreement with this finding but whilst stomata were
present in lower densities in the thicker-leaved CAM species
of Clusia, the stomatal pore areas tended to be larger in the
CAM Clusias compared with the C3 species. A negative rela-
tionship between stomatal density and stomatal size has been
reported for a diverse range of C3 species and is attributed
to spatial limits in the placing of stomata on the leaf surface
which ultimately constrains the maximum size and density of
stomata (Beaulieu et al., 2008; Franks et al., 2009). Since it has
been suggested elsewhere that smaller stomata are better at
improving WUE due to their more rapid response to changes
in environmental conditions such as humidity (Hetherington
and Woodward, 2003), smaller stomata might have been
predicted for CAM-performing Clusias compared with C3
Clusias. However, the production of more, smaller stomata
for a given leaf area may incur additional metabolic costs and
higher rates of guard cell respiration (Srivastava et al., 1995;
Franks et al., 2009). These extra costs could be compensated
for with high CO2 assimilation rates but only if environmen-
tal resources such as water and light are not limiting (Franks
et al., 2009). Given that the CAM-performing species of Clusia
commonly inhabit water-limited environments and, given the
extra energetic costs associated with CAM, the possession of
larger stomata in lower densities might be the most appropri-
ate strategy in terms of resource use.
The CAM species of Clusia studied here possessed both
larger stomata and palisade mesophyll cells which may be a
Fig. 6.  Relationships between the proportion of net CO2 uptake consequence of common genetic control of cell sizes, perhaps
conducted at night in nine species of Clusia (mean ±SE, n=3) and (a) via genome size (Beaulieu et al., 2008). It has been proposed
stomatal density, (b) stomatal index, and (c) stomatal pore area (mean for C3 plants that size correlations between different cell types
±SE, n=50).
(e.g. guard cells, epidermal cells, palisade, and xylem) provide
a highly efficient match between potential maximum water
1985). Thus, leaves of the C3-CAM intermediate P.  obtusi- loss (determined by stomatal conductance) and the leaf vas-
folia showed a thicker WSP (i.e. 63% of total mesophyll) cular system’s capacity to replace that water (determined by
compared with P. macrostachya, a constitutive CAM species vein density; Brodribb et al., 2013). Calculating the theoreti-
(i.e. WSP=19% of total mesophyll; Fondom et al., 2009). It cal maximum stomatal conductance to water in the different
is tempting to speculate that, by acting as a means of buffer- Clusia species showed no clear trend between cell size, pho-
ing against water shortage, a thick WSP might obviate the tosynthetic mode or the potential for water loss. However,
need for CAM. An exception to the trend for a thicker WSP actual measurements of maximal stomatal conductance
3522 | Zambrano et al.

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Fig. 7.  The relationship between the proportion of net CO2 uptake conducted at night in nine species of Clusia (mean ±SE, n=3) and (a) measured
maximal stomatal conductance (gH2O) taken at the start of the photoperiod and (b) theoretical maximal stomatal conductance (GH2O) based on the
density and dimensions of stomata and using the equation described by Lawson et al. (1998).

showed a strong negative relationship with the propensity for References


CAM. Since the difference between theoretical maximum sto- Beaulieu JM, Leitch IJ, Patel S, Pendharkar A, Knight CA. 2008.
matal conductance to water and measured stomatal conduct- Genome size is a strong predictor of cell size and stomatal density in
ance was greatest in those species which showed the greatest angiosperms. New Phytologist 179, 975–986.
level of CAM expression, endogenous control over stoma- Borland AM, Griffiths H, Maxwell C, Broadmeadow MSJ, Griffiths
NM, Barnes JD. 1992. On the ecophysiology of the Clusiaceae in
tal conductance appears to be more important than stoma- Trinidad: expression of CAM in Clusia minor L. during the transition from
tal patterning in determining the potential for water loss in wet to dry season and characterization of three endemic species. New
CAM as compared with C3 species. The lower Lmes/area in the Phytologist 122, 349–357.
CAM-performing species might also curtail the diffusion of Borland AM, Técsi LI, Leegood RC, Walker RP. 1998. Inducibility of
crassulacean acid metabolism (CAM) in Clusia species; physiological/
water vapour from leaves (Kaiser, 2009) biochemical characterisation and intercellular localization of carboxylation
and decarboxylation processes in three species which exhibit different
degrees of CAM. Planta 205, 342–351.
Conclusions Brodribb TJ, Jordan GJ, Carpenter RJ. 2013. Unified changes in cell
size permit coordinated leaf evolution. New Phytologist 199, 559–570.
A study of functional leaf anatomy within the genus Clusia, Crayn DM, Winter K, Smith JAC. 2004. Multiple origins of crassulacean
indicates that a relatively well-aerated spongy mesophyll acid metabolism and the epiphytic habit in the Neotropical family
helps to optimize direct C3-mediated CO2 fixation, whilst Bromeliaceae. Proceedings of the National Academy of Sciences, USA
101, 3703–3708.
enlarged palisade cells accommodate the potential for C4
De Rocher E, Harkins K, Galbraith D, Bohnert H. 1990.
carboxylation and nocturnal storage of organic acids. The Developmentally regulated systemic endopolyploid in succulents with small
findings provide insight into the optimal leaf anatomy that genomes. Science 250, 99–101.
could accommodate the bioengineering of inducible CAM De Veylder L, Larkin J, Schnittger A. 2011. Molecular control and
into C3 crops as a means of improving water use efficiency function of endoreduplication in development and physiology. Trends in
Plant Science 16, 624–634.
without incurring detrimental consequences for direct C3-
Fernández-García N, Martí MC, Jiménez A, Sevilla F, Olmos E. 2009.
mediated photosynthesis. Selecting genotypes of C3 target Sub-cellular distribution of glutathione in an Arabidopsis mutant (vtc1)
crops with increased ploidy levels, which is correlated expe- deficient in ascorbate. Journal of Plant Physiology 166, 2004–2012.
diently with increased cell size and biomass productivity Fondom NY, Castro-Nava S, Huerta AJ. 2009. Seasonal variation in
(De Veylder et  al., 2011) and with succulence (De Rocher photosynthesis and diel carbon balance under natural conditions in two
Peperomia species that differ with respect to leaf anatomy. Journal of the
et  al., 1990), along with a well-developed palisade meso- Torrey Botanical Society 136, 57–69.
phyll, could, in principle, enhance the efficacy of engineered Franks PJ, Drake PL, Beerling DJ. 2009. Plasticity in maximum
CAM. stomatal conductance constrained by negative correlation between
stomatal size and density: an analysis using Eucalyptus globulus. Plant,
Cell & Environment, 32, 1737–1748.
Garnier E, Laurent G. 1994. Leaf anatomy, specific mass and water
Acknowledgements content in congeneric annual and perennial grass species. New
VABZ was funded by Colfuturo, and Newcastle University. This material Phytologist 128, 725–736.
is based upon work supported by the Department of Energy, Office of Gehrig HH, Aranda J, Cushman MA, Virgo A, Cushman JC, Hammel
Science, Genomic Science Program under Award Number DE-SC0008834. BE, Winter K. 2003. Cladogram of Panamanian Clusia based on nuclear
Oak Ridge National Laboratory is managed by UT-Battelle, LLC for the DNA: implications for the origins of crassulacean acid metabolism. Plant
US Department of Energy under Contract Number DE–AC05–00OR22725. Biology 5, 59–70.
Functional leaf anatomy of Clusia | 3523
Grams TEE, Herzog B, Luttge U. 1998. Are there species in the genus Poorter HNU, Poorter L, Wright IJ, Villar R. 2009. Causes and
Clusia with obligate C3 photosynthesis? Journal of Plant Physiology 152, 1–9. consequences of variation in leaf mass per area (LMA): a meta-analysis.
Griffiths H. 1992. Carbon isotope discrimination and the integration New Phytologist 182, 565–588. Erratum Vol. 183, 1222.
of carbon assimilation pathways in terrestrial CAM plants. Plant, Cell & Roberts S, Griffiths H, Borland AM, Reinert F. 1996. Is crassulacean
Environment 15, 1051–1062. acid metabolism activity in sympatric species of hemi-epiphytic stranglers
Griffiths H. 2013. Plant venation: From succulence to succulents. Current such as Clusia related to carbon cycling as a photoprotective process?
Biology 23, R340–R431. Oecologia 106, 28–38.
Gustafsson MHG, Winter K, Bittrich V. 2007. Diversity, phylogeny Sage RF. 2002. Are crassulacean acid metabolism and C(4)
and classification of Clusia. In: Luttge U, ed. Clusia. A woody neotropical photosynthesis incompatible? Functional Plant Biology 29, 775–785.
genus of remarkable plasticity and diversity, Ecological Studies, Vol. 194. Sayed O. 1998. Succulent plants inhabiting desert depressions in
Heidelberg: Springer, 95–116. eastern Arabia were predominantly leaf succulents exhibiting a variety of
Hetherington A, Woodward F. 2003. The role of stomata in sensing and adaptations. Journal of Arid Environments 40, 177–189.
driving environmental change. Nature 424, 901–908. Silvera K, Santiago LS, Winter K. 2005. Distribution of crassulacean
Holtum JAM, Aranda J, Virgo A, Gehrig HH, Winter K. 2004. δ13C acid metabolism in orchids of Panama: evidence of selection for weak and
values and crassulacean acid metabolism in Clusia species from Panama. strong modes. Functional Plant Biology 32, 397–407.
Trees: Structure and Function 18, 658–668. Sipes DL, Ting IP. 1985. Crassulacean acid metabolism and

Downloaded from https://academic.oup.com/jxb/article-abstract/65/13/3513/2877477 by guest on 23 March 2019


Kaiser H. 2009. The relation between stomatal aperture and gas crassulacean acid metabolism modifications in Peperomia camptotricha.
exchange under consideration of pore geometry and diffusional resistance Plant Physiology 77, 59–63.
in the mesophyll. Plant, Cell & Environment 32, 1091–1098. Smith W, Hughes N. 2009. Progress in coupling plant form and
Lawson T, James W, Weyers J. 1998. A surrogate measure of stomatal photosynthetic function. Castanea, 74, 1–26.
aperture. Journal of Experimental Botany 49, 1397–1403. Srivastava A, Lu Z, Zeiger E. 1995. Modification of guard cell properties
Luttge U. 2006. Photosynthetic flexibility and ecophysiological plasticity: in advanced lines of Pima cotton bred for higher yields and heat
questions and lessons from Clusia, the only CAM tree, in the neotropics. resistance. Plant Science 108, 125–131.
New Phytologist 171, 7–25. Vargas-Soto JG, Andrade JL, Winter K. 2009. Carbon isotope
Luttge U. 2008. Clusia: Holy Grail and enigma. Journal of Experimental composition and mode of photosynthesis in Clusia species from Mexico.
Botany 59, 1503–1514. Photosynthetica 47, 33–40.
Lüttge U. 1999. One morphotype, three physiotypes: sympatric species Vendramini F, Díaz S, Gurvich DE, Wilson PJ, Thompson K,
of Clusia with obligate C-3 photosynthesis, obligate CAM and C-3-CAM Hodgson JG. 2002. Leaf traits as indicators of resource-use strategy in
intermediate behaviour. Plant Biology 1, 138–148. floras with succulent species. New Phytologist 154, 147–157.
Lüttge U. 2007. Clusia: A woody neotropical genus of remarkable von Willert DJ, Eller BM, Werger MJA, Brinckmann E. 1990. Desert
plasticity and diversity, Ecological Studies, Vol. 194. Heidelberg: Springer. succulents and their life strategies. Vegetatio 90, 133–143.
Maxwell K, Von Caemmerer S, Evans J. 1997. Is a low internal Weyers JDB, Johansen LG. 1985. Accurate estimation of stomatal
conductance to CO2 diffusion a consequence of succulence in plants with aperture from silicone-rubber impressions. New Phytologist 101,
crassulacean acid metabolism? Australian Journal of Plant Physiology 24, 109–115.
777–786. Winter K, Garcia M, Holtum JAM. 2008. On the nature of facultative
Morison JIL, Gallouët E, Lawson T, Cornic G, Herbin R, Baker and constitutive CAM: environmental and developmental control of CAM
NR. 2005. Lateral diffusion of CO2 in leaves is not sufficient to support expression during early growth of Clusia, Kalanchoë, and Opuntia. Journal
photosynthesis. Plant Physiology 139, 254–266. of Experimental Botany 59, 1829–1840.
Nelson E, Sage R. 2008. Functional constraints of CAM leaf anatomy: Winter K, Holtum JAM. 2002. How closely do the δ13C values of
tight cell packing is associated with increased CAM function across crassulacean acid metabolism plants reflect the proportion of CO2 fixed
a gradient of CAM expression. Journal of Experimental Botany 59, during day and night? Plant Physiology 129, 1843–1851.
1841–1850. Winter K, Wallace BJ, Stocker GC, Roksandic Z. 1983. Crassulacean
Nelson E, Sage T, Sage RF. 2005. Functional leaf anatomy of plants with acid metabolism in Australian vascular epiphytes and some related
crassulacean acid metabolism. Functional Plant Biology 32, 409–419. species. Oecologia 57, 129–141.

You might also like