ArticlePDF AvailableLiterature Review

Deconstructing Kranz anatomy to understand C4 evolution

Authors:

Abstract and Figures

C4 photosynthesis is a complex physiological adaptation that confers greater productivity than the ancestral C3 photosynthetic type in environments where photorespiration is high. It evolved in multiple lineages through the coordination of anatomical and biochemical components, which concentrate CO2 at the active site of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). In most C4 plants, the CO2-concentrating mechanism is achieved via the confinement of Rubisco to bundle-sheath cells, into which CO2 is biochemically pumped from surrounding mesophyll cells. The C4 biochemical pathway relies on a specific suite of leaf functional properties, often referred to as Kranz anatomy. These include the existence of discrete compartments differentially connected to the atmosphere, a close contact between these compartments, and a relatively large compartment to host the Calvin cycle. In this review, we use a quantitative dataset for grasses (Poaceae) and examples from other groups to isolate the changes in anatomical characteristics that generate these functional properties, including changes in the size, number, and distribution of different cell types. These underlying anatomical characteristics vary among C4 origins, as similar functions emerged via different modifications of anatomical characteristics. In addition, the quantitative characteristics of leaves all vary continuously across C3 and C4 taxa, resulting in C4-like values in some C3 taxa. These observations suggest that the evolution of C4-suitable anatomy might require relatively few changes in plant lineages with anatomical predispositions. Furthermore, the distribution of anatomical traits across C3 and C4 taxa has important implications for the functional diversity observed among C4 lineages and for the approaches used to identify genetic determinants of C4 anatomy.
Content may be subject to copyright.
Journal of Experimental Botany, Vol. 65, No. 13, pp. 3357–3369, 2014
doi:10.1093/jxb/eru186 Advance Access publication 5 May, 2014
Review papeR
Deconstructing Kranz anatomy to understand C4 evolution
Marjorie R.Lundgren, Colin P.Osborne and Pascal-AntoineChristin*
Department of Animal and Plant Sciences, University of Sheffield, Sheffield S10 2TN, UK
* To whom correspondence should be addressed. E-mail: p.christin@sheffield.ac.uk
Received 13 January 2014; Revised 15 March 2014; Accepted 25 March 2014
Abstract
C4 photosynthesis is a complex physiological adaptation that confers greater productivity than the ancestral C3 photo-
synthetic type in environments where photorespiration is high. It evolved in multiple lineages through the coordination
of anatomical and biochemical components, which concentrate CO2 at the active site of ribulose-1,5-bisphosphate car-
boxylase/oxygenase (Rubisco). In most C4 plants, the CO2-concentrating mechanism is achieved via the confinement
of Rubisco to bundle-sheath cells, into which CO2 is biochemically pumped from surrounding mesophyll cells. The C4
biochemical pathway relies on a specific suite of leaf functional properties, often referred to as Kranz anatomy. These
include the existence of discrete compartments differentially connected to the atmosphere, a close contact between
these compartments, and a relatively large compartment to host the Calvin cycle. In this review, we use a quantitative
dataset for grasses (Poaceae) and examples from other groups to isolate the changes in anatomical characteristics
that generate these functional properties, including changes in the size, number, and distribution of different cell types.
These underlying anatomical characteristics vary among C4 origins, as similar functions emerged via different modifi-
cations of anatomical characteristics. In addition, the quantitative characteristics of leaves all vary continuously across
C3 and C4 taxa, resulting in C4-like values in some C3 taxa. These observations suggest that the evolution of C4-suitable
anatomy might require relatively few changes in plant lineages with anatomical predispositions. Furthermore, the dis-
tribution of anatomical traits across C3 and C4 taxa has important implications for the functional diversity observed
among C4 lineages and for the approaches used to identify genetic determinants of C4 anatomy.
Key words: C4 photosynthesis, complex trait, convergent evolution, co-option, Kranz anatomy, leaf.
Introduction
During the diversication of owering plants, C4 photosyn-
thesis evolved from C3 ancestors more than 62 times indepen-
dently in several distantly related groups (Sage etal., 2011). C4
photosynthesis is characterized by a biochemical CO2 pump
formed by the coordination of several evolutionary novelties,
which increase the relative concentration of CO2 around rib-
ulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) to
nearly eliminate photorespiration (Ludwig and Canvin, 1971;
Hatch, 1987; von Caemmerer and Furbank, 2003; Skillman,
2008; Sage et al., 2012). The CO2-concentrating mechanism
relies on the primary xation of atmospheric carbon by phos-
phoenolpyurvate carboxylase (PEPC) coupled with carbonic
anhydrase. These reactions are spatially separated from the
secondary rexation of CO2 by Rubisco (Hatch, 1987; von
Caemmerer and Furbank, 2003). An efcient segregation of
these C4 biochemical reactions requires specic leaf functions
(Hattersley, 1984; Dengler etal., 1994; Muhaidat etal., 2007).
As a result of its multiple origins, C4 photosynthesis does
not present a consistent and discrete phenotype, so is better
considered a functional trait involving a suite of coordinated
leaf anatomical and biochemical characteristics (Brown and
Smith, 1972; Laetsch, 1974). These components can assem-
ble differently during each origin of C4 photosynthesis, and
these divergent evolutionary histories result in high anatomi-
cal and biochemical diversity among, and sometimes within,
C4 lineages (Hattersley and Watson,1992; Sinha and Kellogg,
© The Author 2014. Published by Oxford University Press on behalf of the Society for Experimental Biology. All rights reserved.
For permissions, please email: journals.permissions@oup.com
Abbreviations: PEPC, phosphoenolpyurvate carboxylase; Rubisco, ribulose-1,5-bisphosphate carboxylase/oxygenase.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3358 | Lundgren etal.
1996; Kadereit et al., 2003; Muhaidat et al., 2007; Edwards
and Voznesenskaya, 2011; Freitag and Kadereit, 2014). An
understanding of the evolutionary transitions leading to the
recurrent assembly of C4 photosynthesis requires investiga-
tion of the individual characteristics that together generate
C4 function, not only in C4 species but also in C3 species vari-
ously related to C4 taxa (Christin and Osborne, 2013). It is
particularly important to differentiate the present function of
each component from its identity and developmental origin.
In this work, we focus on the variation observed in both C3
and C4 plants in each of the anatomical traits that together
generate leaf functions compatible with C4 photosynthe-
sis. We combine a review of the literature with analyses of
a quantitative leaf anatomy dataset compiled from 155 C3
and C4 grass species (Christin etal., 2013). The C4 grasses in
this dataset encompass eight of the nine structural C4 forms
described for this family (Edwards and Voznesenskaya, 2011).
What is C4 leaf anatomy?
Differential arrangements of cells and organelles within the
leaves of taxa that we now recognize as C3 and C4 were rst
observed and published more than 80years before the C4 path-
way itself was discovered (Duval-Jouve, 1875; Haberlandt,
1884). The association between specic cell and organelle
arrangements and the C4 pathway was then identied soon
after the discovery of C4 photosynthesis (El-Sharkawy and
Hesketh, 1965; Downton and Tregunna, 1968; Berry etal.,
1970; Welkie and Caldwell, 1970). Since then, C4 photosyn-
thesis has usually been afliated closely with a suite of leaf
properties referred to as ‘Kranz’ anatomy (after Haberlandt’s
description in German of a wreath-like arrangement of cells).
Kranz anatomy can be described as two distinct concentric
layers of chlorenchyma cells, formed by a bundle sheath con-
taining most of the chloroplasts, surrounded by an outer
layer consisting of a small number of mesophyll cells. The
visual identication of such arrangements in transverse sec-
tion has been used in numerous anatomical surveys of leaves
to identify the photosynthetic pathway for hundreds of spe-
cies (Welkie and Caldwell, 1970; Carolin et al., 1973, 1975,
1977; Brown, 1977; Hattersley etal., 1982; Renvoize, 1987a).
Surveys of numerous C3 and C4 species over the past ve
decades have shown that leaf anatomies cannot be easily and
consistently grouped into discrete categories corresponding
to the two photosynthetic types but come in many avours
(Brown, 1975; Edwards and Voznesenskaya, 2011). It is true
that the leaf anatomy of a randomly selected C3 plant is
highly likely to deviate signicantly from that of a randomly
selected C4 plant. For example, Viburnum punctatum, like
most C3 eudicots, has distinct horizontal layers of mesophyll
cells in its leaves (Fig.1A), arranged such that it does not con-
form to the general anatomical pattern generally present in C4
plants, whereby the bundle-sheath and mesophyll cells form
concentric circles around the vasculature (Fig.1C). This con-
centric arrangement of cells can be found in many C3 grasses
though (Figs 1B and 2) (Hattersley etal., 1982; Dengler etal.,
1994; Besnard etal., 2013) and, as detailed below, individual
leaf characteristics that are usually associated with a C4 func-
tion can be found in at least some C3 plants. Furthermore,
some plants achieve C4 photosynthesis without the segre-
gation of photosynthetic reactions into different types of
cells (Bowes and Salvucci, 1984; Bowes and Salvucci, 1989;
Freitag and Stichler, 2000; Edwards et al., 2004). Despite
this variation, C4 physiology is still associated with a suite
Fig.1. Examples of C3 and C4 leaf cross-sections. The C3/C4 pair on the left (A, C) are unrelated, belonging to different major groups of flowering
plants. By contrast, the C3/C4 pair on the right (B, D) is composed of closely related species, belonging to the same subfamily of grasses. (A) Viburnum
punctatum (C3, Adoxaceae), (B) Sartidia angolensis (C3, Poaceae), (C) Centropodia mossamedensis (C4, Poaceae), and (D) Aristida mollissima (C4,
Poaceae). Black arrows indicate the mesophyll, red arrows the outer bundle sheath, and blue arrows the inner sheath of grasses (=mestome sheath). The
four cross-sections are shown at the same scale. Bars, 100μm. Picture (A) was kindly provided by Dr David Chatelet from Brown University and pictures
(B), (C) and (D) come from the collections of Professor J.Travis Columbus from Rancho Santa Ana Botanic Garden, CA, USA, with permission.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3359
of functional properties (Brown and Smith, 1972; Edwards
and Voznesenskaya, 2011), which must rst be considered
before analysing diversity in the identity and developmental
origins of the characteristics that generate them. Based on
the literature, the following functional properties of leaves
are considered essential requirements for C4 photosynthesis
(Hattersley etal., 1977; Leegood, 2002; von Caemmerer and
Furbank, 2003; Edwards and Voznesenskaya, 2011; Nelson,
2011). Note that these apply equally to all C4 plants, whether
or not they use distinct types ofcells.
1. There must be two distinct compartments arranged so
that atmospheric gases reach the rst compartment more
easily than the second. The rst compartment houses the
PEPC reactions, while the second, with characteristics that
restrict CO2 efux, houses the Calvin cycle.
2. The two compartments must be in close contact to allow
the rapid exchange of metabolites.
3. The compartment where the Calvin cycle occurs must
occupy a large enough fraction of the leaf to accommo-
date a signicant number of chloroplasts.
4. Chloroplasts must be abundant in the Calvin cycle
compartment.
These functional properties are extremely important for C4
physiology and biochemistry. However, to understand the
gradual evolutionary changes leading to the recurrent assem-
bly of C4 photosynthesis, it is important to account for exact
changes in cellular characteristics and the genetic determi-
nants of these characteristics. In the following sections, we
therefore discuss how each of the four functional properties
listed above is generated from underlying characteristics. We
look at how these characteristics vary qualitatively and quan-
titatively among C3 and C4 lineages, and show how there is
an overlap between the values observed in C3 and C4 species.
Two compartments differentially connected to the
atmosphere
In C3 plants, the Calvin cycle occurs in most of the leaf,
while it is restricted to specic locations in C4 plants. It is well
known that the identity of the compartments co-opted for the
segregation of the atmospheric CO2 xation by PEPC and its
rexation by the Calvin cycle differs among C4 origins (e.g.
Brown, 1975; Dengler etal., 1985). For instance, some single-
celled C4 species have evolved separate compartments for the
PEPC and Calvin cycle reactions through the rearrangements
of organelles or vacuoles within individual photosynthetic
cells (Edwards etal., 2004). In the majority of C4 plants, how-
ever, the PEPC and Calvin cycle reactions are segregated in
different types of cells. In C3 species, the mesophyll and bun-
dle sheath represent two physiologically distinct types of cells,
and the central position of bundle-sheath cells within the leaf
gives the opportunity for minimal contact with the atmos-
phere (Figs 1A, B and 3, and Supplementary Fig. S1 available
at JXB online). The bundle sheaths have consequently been
co-opted for Calvin cycle reactions across most C4 origins,
while the mesophyll cells, which are better connected to the
atmosphere, are used for the PEPC reactions. Despite this
convergence in function, the bundle-sheath cells recruited for
C4 photosynthesis are not homologous among all C4 origins.
In some C4 species within the grass genera Arundinella,
Garnotia, Arthropogon, Achlaena, Dissochondrus, Anrthraxon,
Fig.2. Examples of C3 grasses with leaf anatomy close to the C4 requirements. (A) Panicum pygmaeum (C3), (B) Panicum malacotrichum (C3). The
mesophyll (M) and vascular tissue (V) are indicated on the sections. Red arrows indicate the outer bundle sheath while blue arrows indicate the inner
sheath (=mestome sheath). Bars, 500μm.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3360 | Lundgren etal.
and Microstegium, the Calvin cycle also occurs in distinctive
cells, which are atypical bundle-sheath-like cells, differenti-
ated within the mesophyll but not associated with vascular
bundles (Fig. 3A) (Tateoka, 1958; Hattersley and Watson,
1992; Ueno, 1995; Dengler et al., 1996; Wakayama et al.,
2003). In addition, grasses and sedges possess multiple lay-
ers of sheath cells, with inner layers derived from procam-
bium (often referred to as the ‘mestome sheath’) and outer
layers from ground meristem (Dengler etal., 1985; Soros and
Dengler, 2001; Martins and Scatena, 2011). In studies of C4
photosynthesis, consideration of the different cells is often
based on their function. However, for evolutionary studies,
the ontogenic origin of each type of cell needs to be estab-
lished independently of its function. The C4 lineages within
grasses and sedges have alternatively co-opted one or both
of these cell types, while the second cell layer is often lost, for
example in the numerous C4 grasses with a single sheath layer
(Fig.3AE) (Brown, 1975; Dengler et al., 1996; Soros and
Dengler, 2001; Martins and Scatena, 2011). This diversity in
the identity of the two compartments co-opted for the segre-
gation of C4 reactions, together with phylogenetic analyses,
has been used previously to argue for multiple independent
C4 origins, rather than fewer origins followed by reversals in
closely related C3 species (Kellogg, 1999; Christin etal., 2010).
The limited connection of the Calvin cycle compartment to
the atmosphere is also achieved via different mechanisms in
the different C4 lineages. First, tightly packing mesophyll cells
around the bundle sheath reduces the fraction of cells from
the latter that are in contact with the atmosphere (Dengler
etal., 1994; Muhaidat etal., 2007), although similar packing
also occurs in some C3 grasses (Fig.1B) (Dengler etal., 1994)
and some C3 eudicots (Muhaidat et al., 2007). In addition,
the bundle-sheath cell walls can also be covered with a layer
of suberin, which limits gas diffusion. This is the case in C4
monocots that have co-opted the inner sheath layer for a C4
function (Hattersley and Browning, 1981; Ueno etal., 1988b).
However, the presence of suberin layers on the inner sheath
cell walls can also be found in most C3 grasses (Hattersley
and Browning, 1981). Neither of the characteristics reducing
contact of the Calvin cycle with the atmosphere is therefore
found exclusively in C4 plants.
Distance between the two compartments
Close contact between the PEPC and Calvin cycle com-
partments is guaranteed in plants with a single-celled C4
system. In plants with a dual-celled C4 system, the pres-
ence of mesophyll cells not directly adjacent to the bun-
dle sheaths will increase the average distance between the
compartments containing PEPC and Rubisco. This prob-
lem is usually solved in C4 plants by limiting the number
of cells separating consecutive Calvin cycle compartments,
and by organizing mesophyll cells into one or two layers
around the bundle sheath (Fig.1C, D), which produces the
classical pattern of Kranz anatomy. In some species, this
conguration is achieved through the development of a
Fig.3. Leaf anatomy for selected cross-sections of grasses. (A) Arundinella nepalensis (C4), (B) Anthaenantia lanata (C4), (C) Axonopus compressus
(C4), (D) Ischaemum afrum (C4), (E) Chrysopogon pallidus (C4), (F) Alloteropsis cimicina (C4), (G) Panicum pygmaemum (C3), (H) Bouteloua stolonifera
(C4) and (I) Panicum malacotrichum (C3). The diagrams highlight the mesophyll cells (yellow), outer bundle sheaths (red), inner bundle sheaths (blue), and
distinctive cells (purple). Uncoloured central areas are composed of mesophyll cells and intercellular airspace. Vein (light grey), epidermis (dark grey), and
sclerenchymatous girders (solid black) are also shown. Where only one bundle sheath is present, it is assumed that the outer bundle sheath has been
lost and the inner bundle sheath remains. All cross-sections are drawn at the same scale, indicated at the bottom. The corresponding pictures can be
found in Supplementary Fig. S1 available at JXB online.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3361
single bundle-sheath layer that encompasses all the vascu-
lature within the leaf and often water-storage cells as well,
and a single layer of mesophyll that surrounds the bundle
sheath. Variations on this anatomical theme are common
among C4 eudicots and have been found in the Asteraceae,
Amaranthaceae, and Cleomaceae families (Carolin et al.,
1975; Das and Raghavendra, 1976; Kadereit et al., 2003;
Peter and Katinas, 2003; Edwards and Voznesenskaya,
2011; Koteyeva et al., 2011). Some C4 grasses have similar
bundle sheaths that extend horizontally from the vascular
tissue and join together, such that the mesophyll becomes
isolated in small patches (Renvoize, 1983).
For C4 lineages with multiple photosynthetic units formed
by concentric cell layers of mesophyll, bundle sheath,
and vascular tissue, the presence of fewer mesophyll cells
between consecutive veins can be achieved via two different
developmental mechanisms. First, the number of cells that
develop between consecutive bundle sheaths can be directly
reduced during ontogeny. Second, extra Calvin cycle compart-
ments, such as distinctive cells or minor veins, can be added to
decrease the average distance between compartments, as has
been documented in both monocots (e.g. Poaceae; Fig.3A
E; Renvoize, 1987a; Dengler etal., 1994; Ueno etal., 2006;
Christin etal., 2013) and eudicots (e.g. Asteraceae; McKown
and Dengler, 2007; McKown and Dengler, 2009; Cleomaceae;
Marshall etal.,2007).
Interveinal distance (or vein density) is often considered
a proxy for the number of mesophyll cells between consecu-
tive bundles, and largely overlaps between C3 and C4 grasses
(Christin etal., 2013) and eudicots (Muhaidat etal., 2007).
However, the relationship between interveinal distance and
the number of mesophyll cells is only partial. First, because
Fig.4. Multidimensionality of C4 anatomy in grasses. Scatter plots for anatomical variables associated with the C4 syndrome are shown, along with
frequency distributions for each trait, arranged along the axes. For each pair of variables, dot size is proportional to a third variable. C3 grass species are
shown in yellow, C4 grass species using the outer sheath for the Calvin cycle in red, and C4 grass species using the inner sheath for the Calvin cycle in
blue. Relationships are shown between means of: (A) distance between consecutive bundle sheaths (μm) and inner bundle-sheath cell width (μm), with
dot size proportional to the percentage of inner bundle-sheath area; (B) distance between consecutive bundle sheaths (μm) and outer bundle-sheath
cell width (μm), with dot size proportional to the percentage of outer bundle-sheath area; (C) number of mesophyll cells between consecutive bundles
and mesophyll cell length (μm), with dot size proportional to the distance between consecutive bundle sheaths (μm); and (D) outer bundle sheath cell
width (μm) and area of vasculature (μm2), with dot size proportional to the outer bundle sheath area (μm2) per vein number. The data for 170 grasses
(representing 155 species) come from Christin etal. (2013).
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3362 | Lundgren etal.
interveinal distance is inuenced both by the diameter of
the veins and the size of the bundle sheaths, measuring
the actual distance between bundle sheaths is more rel-
evant. This distance is inuenced by the size of individual
mesophyll cells, their orientation, and nally their number
(Fig.4C). Some C4 species, such as Alloteropsis cimicina,
have relatively large interveinal distances but with only a few
large mesophyll cells between consecutive bundles (Fig.3F,
4C). In addition, the number of mesophyll cells contain-
ing PEPC below and above veins can inuence the aver-
age distance between the PEPC and Calvin cycle reactions
independently of the distance between consecutive bun-
dles. Some thick C4 leaves, such as those of Anthaenathia
lanata (Fig.3B) or some Portulaca (Ocampo etal., 2013),
consequently require a three-dimensional venation system.
Finally, leaf thickness is often reduced between veins so
that there are few mesophyll cells in positions most dis-
tant from the bundle sheaths, and interveinal distance can
greatly exceed the average distance between photosyntheti-
cally active mesophyll cells and bundle-sheath cells (Figs 1
and 3). For instance, in leaves of the C3 grass Panicum pyg-
maemum, the average number of mesophyll cells between
bundles greatly exceeds four. However, because its leaf
thickness decreases between veins, the number of meso-
phyll cells separated from the bundle sheath by more than
one cell is smaller than the number of mesophyll cells sepa-
rated from the bundle sheath by zero or one cell (38 versus
73 cells between the three veins in Fig.2). Finally, the dis-
tance between consecutive bundles can be increased by the
presence of achlorophyllous cells that do not inuence the
average path length from PEPC to Calvin cycle cells (e.g.
Fig.1D).
The number of mesophyll cells between consecutive bun-
dles will distinguish C3 from C4 taxa with a high success rate
and has consequently been proposed as a criterion to rec-
ognize C4 plants (Hattersley and Watson, 1975; Renvoize,
1987a; Sinha and Kellogg, 1996). However, the C3 and C4 dis-
tributions for this trait also overlap (Fig.4C). For instance,
Panicum malacotrichum is a C3 grass with less than four meso-
phyll cells between veins (Fig. 2). The variation observed in
both C3 and C4 taxa is probably due to the importance of vas-
cular architecture for both photosynthetic types. While the
distance between consecutive bundles affects the efciency of
C4 photosynthesis (Ogle, 2003), vein density also inuences
the transport of metabolites, leaf hydraulics and other physi-
ological characteristics in C3 plants (Sack and Scoffoni, 2013;
Sack etal., 2013). In summary, both interveinal distance and
the number of mesophyll cells between consecutive bundles
overlap in C3 and C4 taxa, so that C4 values represent only
a subset of those observed among all photosynthetic types
(Fig.4AC) (Muhaidat etal., 2007; Christin etal., 2013).
The transport of metabolites between the PEPC and
Calvin cycle compartments in C4 plants is also facilitated by
a number of plasmodesmata connecting mesophyll and bun-
dle-sheath cells that exceeds the number found in C3 plants
(Olesen, 1975; Weiner et al., 1988; Botha, 1992). However,
plasmodesmata frequency is known in only a few C3 species,
so the overall variation in this trait cannot be established with
condence.
Large Calvin cycle compartment
The amount of CO2 that can be re-xed by Rubisco in the
Calvin cycle will depend on the number of chloroplasts
within the compartment co-opted for this function. The size
of this compartment, not including the volume occupied by
the vacuole, will inuence the number of chloroplasts that
can be accommodated. Thus, C4 plants tend to have enlarged
bundle-sheath cells able to accommodate numerous chloro-
plasts. More than the size of individual bundle sheath cells,
the cumulative volume of bundle sheath relative to the PEPC
compartment (mesophyll) is relevant, and seems to be con-
strained within a given range in C4 plants (Hattersley, 1984;
Dengler etal., 1994; Muhaidat etal., 2007). This might repre-
sent a trade-off between having sufcient chloroplasts in the
Calvin cycle compartment and still conserving enough meso-
phyll volume forPEPC.
Similar bundle sheath:mesophyll ratios can be achieved
through different combinations of the numerator (volume
of bundle sheath) and denominator (volume of mesophyll).
For instance, similar proportions of bundle sheath can be
achieved through alternative developmental mechanisms,
involving the production of either larger or more numerous
bundle-sheath cells (the latter is generally achieved through
a proliferation of veins; Fig.3) (Hattersley, 1984; McKown
and Dengler, 2009). The cross-sectional area of mesophyll per
vein is mainly a function of the distance between veins, the
thickness of the leaf (including the thickness between veins
in comparison to that at the veins) and the presence of achlo-
rophyllous cells (Christin et al., 2013). On the other hand,
when viewed in transverse section, the total area of a given
type of bundle sheath per vein is a function of the size of the
bundle-sheath cells, the diameter of the veins, and, in some
cases, the completeness of the bundle sheath (Fig.4) (Christin
etal., 2013). For instance, the external bundle sheath of many
grasses is not developed on the abaxial side of the leaf, which
reduces the total volume of this tissue (Fig.5) (e.g. Renvoize,
1985, 1987b). Thus, the relative amount of bundle-sheath tis-
sue is a function of at least ve distinct traits, which may all
vary independently. Functionally similar characteristics can
consequently arise through different developmental modi-
cations, as highlighted by the diversity of C4 leaf anatomy
(Fig.4).
Fig.5. Detail of a cross-section for Dactylis glomerata. The mesophyll (M)
and vascular tissue (V) are indicated on the section of this C3 species. The
red arrow Indicates the outer bundle sheath, while the blue arrow indicates
the inner sheath (=mestome sheath). Bar, 100μm. Note the incomplete
outer sheath.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3363
The ve components that dictate the relative amount of
bundle-sheath tissue are important determinants of the gross
leaf anatomy associated with C4 photosynthesis. However,
each component shows an essentially continuous distribu-
tion across C3 and C4 values, such that C4-compatible ranges
merely represent a subset of the distribution found in C3 taxa
(Fig.4; Marshall etal., 2007; McKown and Dengler, 2007).
The C4-suitability of one parameter depends on the values of
the other parameters. For instance, large volumes of bundle-
sheath tissue can arise in the presence of signicant distances
between consecutive bundles if the bundle-sheath cells are
enlarged (Fig.4A, B). This is highlighted by a comparison
of Alloteropsis cimicina and Axonopus compressus (Fig.3F
and C, respectively), which achieved similar ratios of bun-
dle sheath per mesophyll area [BS/(BS+M) of 0.26 and 0.21,
respectively] through different means. Alloteropsis cimicina
has very large outer bundle sheaths that are separated by
long distances of mesophyll, while Axonopus compressus has
small inner sheaths that are separated by very short meso-
phyll distances in particularly thin leaves (Fig. 3F and 3C,
respectively).
During the course of evolution, numerous alterations in
the characteristics that generate each leaf function occur
either stochastically or in response to selective pressures.
For instance, leaf thickness often represents an adaptation
to the amount of light received by plants (Boardman, 1977;
Terashima et al., 2001). The number and size of veins alters
the hydraulics of a plant, which, in turn, affects the sorting of
plants across environments (McKown etal., 2010; Sack etal.,
2012). Finally, the bundle sheath controls water ux between
the mesophyll and vascular tissue such that an increase in
bundle-sheath size might provide better protection against
cavitation in arid environments (Sage, 2001; Leegood, 2008;
Grifths et al., 2013). Recurrent and independent changes
in different leaf properties repeatedly led to the emergence
of tissues suitable for C4 photosynthesis, which characterize
numerous extant C3 plants (Muhaidat et al., 2007; Edwards
and Voznesenskaya, 2011; Muhaidat et al., 2011; Kadereit
etal., 2012; Christin etal., 2013; Grifths etal., 2013).
Distribution of organelles
One of the most important requirements for C4 photosynthe-
sis probably lies in the distribution of chloroplasts. Although
they are present in all photosynthetic cells of C3 plants, chlo-
roplasts are especially abundant in mesophyll cells and can
vary from equally abundant to completely absent in bundle-
sheath cells (Figs 1, 2 and 5) (Crookston and Moss, 1970).
In C4 plants, the light-dependent and light-independent func-
tions of chloroplasts are often decoupled, and chloroplasts of
the PEPC and Calvin cycle compartments can become mor-
phologically and functionally differentiated (Woo etal., 1970;
Laetsch, 1974; Hattersley etal., 1977; Bowman etal., 2013).
Although the characteristics and distribution of organelles
vary among C4 lineages (Ueno et al., 1988b; Voznesenskaya
etal., 2006; Edwards and Voznesenskaya, 2011), the Calvin
cycle compartment of C4 plants consistently has a high con-
centration of chloroplasts, where the enzymes of the Calvin
cycle are preferentially expressed.
No quantitative census of chloroplast distribution is avail-
able for randomly selected plants; however, the organelle dis-
tribution has been investigated in species closely related to
C4 lineages, which shows that some plants maintain signi-
cant numbers of chloroplasts in bundle-sheath cells, despite
lacking a functional C4 pathway (Hattersly etal., 1982; Ueno
and Sentoku, 2006; Christin etal., 2013). This is particularly
common in plants using C2 photosynthesis, a weak CO2-
concentrating mechanism based on a glycine shuttle from
mesophyll to bundle-sheath cells (Edwards and Ku, 1987;
Sage et al., 2012). When chloroplast abundance in bundle-
sheath cells is compared among taxa, there is a gradient
from closely related C3 to C2, and then from C2 to C4 spe-
cies (Muhaidat et al., 2011; Sage et al., 2013). The C2 trait
is consequently often considered an evolutionary intermedi-
ate between C3 and C4 types (Hylton etal., 1988; Sage etal.,
2012; Williams etal., 2013). Therefore, as for other anatomi-
cal traits, the number of chloroplasts in bundle-sheath cells
varies and may form a continuum between C3 and C4 species.
Despite this, a high concentration of chloroplasts in bundle-
sheath cells might be the only trait that occurs systematically
within dual-celled C4 photosynthesis that is never present
in non-C4 plants. The tight association between C4 physiol-
ogy and chloroplast distribution is explained by the fact that
C4 physiology results from a differential distribution of the
Calvin cycle (among other biochemical reactions), which is
usually linked to the distribution of chloroplasts.
Other ultrastructural properties associated with some C4
plants include the distribution of mitochondria and peroxi-
somes among compartments, the distribution of organelles
within compartments and the ultrastructure and photochem-
ical properties of the chloroplasts (Bruhl and Perry, 1995;
Edwards and Voznesenskaya, 2011). Some of these properties
are also observed in non-C4 species closely related to C2 and
C4 taxa (Sage etal., 2012)
Plasticity for C4-suitable anatomy
Phenotypic plasticity to environmental cues creates an addi-
tional layer of variation and further blurs the dichotomy
between C4 and non-C4 anatomy. Specically, plasticity for
the anatomical traits relevant to photosynthesis (e.g. compart-
mentalization, interveinal distance, mesophyll cell size and
number, bundle-sheath cell size, and organelle distribution)
could partially explain the variation found in these anatomical
characteristics or, more importantly, the shift of C3 plants into
the C4-suitable space. Plasticity for these traits has been docu-
mented in the literature. For example, the C3 grass Phragmites
australis acquires C4-like traits when it grows at low soil water
potentials (Gong etal., 2011). Specically, interveinal distance
decreases, chlorophyll content within bundle-sheath cells
increases, and the activity of C4-related enzymes increases as
soil water potential becomes more negative across a natural
precipitation gradient (Gong etal., 2011). The C4-like Flaveria
brownii lacks the complete suite of anatomical characteris-
tics required for a fully functioning C4 system (Araus et al.,
1990). However, this species can plastically increase its degree
of C4 photosynthesis by nearly doubling its investment in
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3364 | Lundgren etal.
bundle-sheath tissue relative to mesophyll in response to high
irradiance compared with when it is grown at low irradiance
(1200 vs 80μmol m–2 s–1 photosynthetic photon ux density,
respectively; Araus etal., 1991). Furthermore, interveinal dis-
tances decreased in the C3 grasses, Festuca arundinaceae (43%
decrease), and the C3/C4 intermediate grass Panicum miliodies
(34% decrease), when grown in high versus low nitrogen levels
(Bolton and Brown, 1980).
In addition to the plasticity of individual anatomi-
cal components, two different modes of environmentally
induced C4 photosynthesis exist. First, several aquatic spe-
cies of Hydrocharitaceae, and possibly some Alismataceae
and Cyperaceae, are able to switch from C3 to single-cell C4
photosynthesis (Bowes et al., 2002). The environmental cue
for this plasticity may be exposure to low-CO2 conditions as
they become submerged under water, or seasonal variation in
temperature (Bowes et al., 1979; Bowes, 2011). In contrast,
some aquatic Eleocharis species use C3 or C3/C4 intermedi-
ate photosynthesis when submerged but induce C3/C4 or C4
photosynthesis by developing C4-compatible leaf anatomy
and expressing C4 enzymes in the emergent leaves (Ueno
etal., 1988a; Ueno, 2001; Murphy etal., 2007). Finally, some
amphibious C4 grasses seem to switch from a C4 system that
functions without C4-associated leaf anatomy in aquatic
leaves to a classical dual-cell C4 cycle in aerial leaves (Keeley,
1998; Boykin etal., 2008).
Phenotypic plasticity for C4-associated traits might have
important implications for the evolution of C4 photosyn-
thesis (Sultan, 1987; West-Eberhard et al., 2011). First, the
direction and degree of phenotypic change in response to an
environmental gradient is heritable (Schlichting and Levin,
1986; Schlichting and Pigliucci, 1993), and the reaction norm
for a trait is genetically distinct from the trait itself. Selection
can therefore act independently on both a trait and on the
plasticity for that trait. Plasticity may thus deter the evolu-
tionary transition from C3 to C4 photosynthesis by diluting
the effects of natural selection. However, adaptive phenotypic
plasticity may promote C4 evolution if the plastic expression
of C4-suitable anatomical traits in C3 plants allows the colo-
nization of new niches, leading to selective pressures for the
gradual acquisition of C4 biochemistry (Heckmann et al.,
2013). Indeed, Sage and McKown (2006) reviewed the litera-
ture to nd that C3 plants seem to be inherently more plas-
tic than C4 plants overall. Thus, this capacity for phenotypic
plasticity might affect the probability of evolving C4 photo-
synthesis. For instance, differential capacity in the phenotypic
plasticity for important C4 anatomical traits among plant lin-
eages may explain the differential propensity for C4 evolution.
However, the plasticity of anatomical traits associated with
C4 photosynthesis remains mostly unknown in C3 species,
and more comparative work is required.
Consequences for the evolution of
C4-associated anatomy
When comparing the anatomy of a randomly selected
C3 taxon with that of a highly efcient C4 species, the
evolutionary transition from C3 to C4 anatomy can seem
extraordinary (Fig.1A, C). However, it is important to note
that C4 photosynthesis did not emerge from the average C3
taxon but from C3 ancestors with leaf anatomical properties
much closer to the C4 requirements (Figs 1B and 5) (Muhaidat
et al., 2011; Christin et al., 2013; Sage et al., 2013). In the
Poaceae, some species apparently using the C3 photosynthetic
type have gross leaf anatomies that closely resemble those of
C4 plants. For instance, Panicum malacotrichum and Panicum
pygmaeum (Fig.2) are two C3 grasses (δ13C values of –27.4
and –29.7, respectively), which are closely related to several
C4 lineages (namely Alloteropsis and Echinochloa; Grass
Phylogeny Working Group II, 2012). These species possess
large proportions of bundle-sheath tissue that are rmly
in the C4 range [BS/(BS+M) of 0.26 and 0.23, respectively;
Christin et al., 2013], and most mesophyll cells are directly
adjacent to the bundle sheath or separated by only one mes-
ophyll cell (Fig.2). Chloroplasts are still almost completely
restricted to the mesophyll in these species. However, because
the gross leaf anatomy is in place, fewer anatomical changes
are necessary for the evolution of C2 or C4 photosynthesis. In
other cases, such as the grass tribe Neurachninae, C3 species
that are closely related to C4 species have both C4-like gross
anatomy [BS/(BS+M) of 0.14–0.16; Christin etal., 2013] and
the presence of conspicuous chloroplasts in the inner sheath,
which was co-opted for C4 photosynthesis in this group
(Hattersley etal., 1982). These examples show that the evolu-
tion of C4-suitable anatomy might not always require drastic
modications, as C3 lineages may possess C4-like values for
individual traits that can generate C4 leaf functions.
Each component of C4-compatible leaf anatomy may vary
independently within C3 ancestors, such that any combination
of mesophyll cell size, bundle-sheath cell size, leaf thickness
and interveinal distance could theoretically occur. However,
the observed range is obviously more limited (Fig.4), for a
number of reasons. First, multiple traits might be inuenced
by the same gene (pleiotropy). For instance, genome size the-
oretically affects the size of all cells (Grime and Mowforth,
1982; Masterson, 1994; Beaulieu et al., 2008; Šímová and
Table1. Degrees of co-variation among anatomical variables
Co-variation in grasses between the mean distance between
consecutive bundle sheaths (μm), outer bundle-sheath (OBS) cell
width (μm), inner bundle-sheath (IBS) cell width (μm), number of
mesophyll (M) cells between consecutive bundles, and leaf thickness
(μm). R2 values are provided for pairs of variables with significant
correlations. Regressions with P values less than 0.05 are considered
significant, while those with P values greater than 0.05 are indicated
by NS. Phylogenetically controlled analyses were performed with the
pgls function of the caper R package (Orme etal., 2012), using the
data for 155 grass species from Christin etal. (2013).
BS distance
0.02 OBS cell width
NS 0.23 IBS cell width
0.58 NS 0.05 No. M cells
NS 0.28 0.43 0.12 Leaf thickness
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3365
Herben, 2012), so that an increase in bundle-sheath cell size
might co-occur with increases in the sizes of mesophyll cells.
Plants often escape this constraint via cell-specic endoredu-
plication, which allows an increase of one type of cell relative
to others (Sugimoto-Shirasu and Roberts, 2003), and com-
parative analyses show that variation in the cell sizes of dif-
ferent components of C4 anatomy is only partially correlated
(Table1). However, endoreduplication is not involved in the
increase of bundle-sheath cell size, at least in the C4 Cleome
gynandra (Aubry etal., 2013). It is also likely that some com-
binations of traits are not viable, as the whole-leaf structure
inuences plant tness (Noblin etal., 2008), not its individual
components.
The multidimensionality of leaf characteristics associated
with C4 photosynthesis, as highlighted for the grass family,
means that different combinations of underlying traits will
generate C4-compatible leaf anatomies (Fig.4). For instance,
both a proliferation of veins with small bundle-sheath cells
and an increase of bundle-sheath cell size without additional
veins would increase the relative amount of bundle-sheath
cells (Fig.6). This potential for alternative anatomical com-
binations to achieve the same functional outcome means that
C3 ancestors will repeatedly reach C4-compatible areas of
the multidimensional trait space (Fig.6), and increases the
likelihood of C4 anatomy evolving (Williams et al., 2013).
Asample of evolutionary trajectories in the Poaceae shows
lineages for which repeated and independent alterations of
the distance between bundle sheaths and bundle-sheath size
led into different C4-compatible regions of the anatomical
space (Fig.6). Obviously, not all C3 lineages that acquired C4-
suitable leaf anatomical characteristics have evolved C4 bio-
chemistry. For example, Panicum malacotrichum and Oryza
coarctata have C4-suitable mesophyll distances between con-
secutive bundle sheaths and proportions of bundle-sheath tis-
sue but have not developed the C4 syndrome (Figs 2 and 6)
(Christin etal., 2013). Furthermore, Cleome violacea, C.afri-
cana, and C.paradoxa have small interveinal distances, and
C. africana and C. paradoxa also display enlarged bundle-
sheath cells similar to their C4 congener C.gynandra, yet these
three species do not employ the C4 photosynthetic system
(Marshall etal., 2007). However, the presence of these char-
acteristics probably enables C4 evolution (pre-adaptation or
exaptation sensu Gould and Vrba, 1982; Christin etal., 2013;
Grifths etal., 2013; Sage etal., 2013). Once a C4-compatible
anatomy is in place, the C4 biochemical pathway can evolve
from a C3 background in a stepwise sequence, where each
step incrementally increases the efciency of photosynthesis
(Heckmann et al., 2013). However, the multiple anatomical
requirements for C4 photosynthesis do not usually co-occur
in C3 plants. Interesting exceptions include plants with a C2
physiology, which were probably co-opted for the evolution
of C4 photosynthesis (Christin et al., 2011; Muhaidat et al.,
2011; Sage etal., 2012).
Functional C4 diversity as a consequence
of evolutionary diversity
Because C4-compatible leaf anatomy engages multiple compo-
nents, each C4 origin may involve different modications and
co-opt different compartments for the Calvin cycle (Brown,
1975; Dengler et al., 1994; Edwards and Voznesenskaya,
2011; Christin et al., 2013). The anatomy present in the C3
ancestor might affect which C4 phenotypes are possible. For
instance, C3 ancestors with enhanced water storage tissue are
likely to give rise to C4 leaves that maintain the same capac-
ity to store water, with the PEPC and Calvin cycle compart-
ments occupying other parts of the leaves (Voznesenskaya
et al., 1999; Kadereit et al., 2003; Freitag and Kadereit,
2014). Similarly, C4 species that use the inner bundle sheath
for the Calvin cycle must evolve from C3 ancestors that pos-
sessed two differentiated sheaths, as is the case with grasses
and sedges (Dengler etal., 1994; Soros and Dengler, 2001).
Furthermore, C4 phenotypes that are functionally similar can
be achieved through different modications, even when start-
ing with similar C3 ancestors.
Different modications to full the same C4 requirements
might have functional consequences. Indeed, the adaptation
of C4 photosynthesis through the evolution of thick leaves
with large bundle-sheath cells (Fig.3F) is likely to have dif-
ferent consequences from the evolution of thin leaves with
small cells but very short interveinal distance (Fig.3C). An
increase in vein density will affect not only the hydraulics but
Fig.6. Evolutionary trajectories toward C4-compatible anatomical traits.
Phylogenetic relationships are plotted in anatomical space for grass
species selected to represent a diversity of anatomical traits. Values are
the distance between consecutive bundle sheaths, and the width of outer
bundle-sheath cells, which are observed for the tips and inferred for the
internal nodes. The black point represents the root of the tree (see Christin
etal., 2013, for details). Yellow branches indicate a C3 state, red branches
a C4 state using the outer sheath for the Calvin cycle, and blue branches
a C4 state using the inner sheath for the Calvin cycle. Numbers refer
to the extant species: 1, Dichanthelium acuminatum (C3); 2, Danthonia
spicata (C3); 3, Heteropogon contortus (C4); 4, Aristida congesta (C4); 5,
Stipagrostis obtusa (C4); 6, Eleusine indica (C4); 7, Panicum malacotrichum
(C3); 8, Oryza coarctata (C3); 9, Panicum miliaceum (C4), 10, Arundo donax
(C3). Data from Christin etal. (2013).
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3366 | Lundgren etal.
also the distribution of stomata, which tend to be located in
between veins (Taylor etal., 2012). Leaf thickness will have
consequences for light-capture efciency as well as ecologi-
cally meaningful traits such as specic leaf area (Wilson etal.,
2002). Similarly, light capture will also be affected by the dif-
ferent distribution of chloroplasts in mesophyll and bundle-
sheath cells, and the relative abundance of each cell type,
together with the orientation of mesophyll cells (Vogelmann
etal., 1996). The path length from stomata to the photosyn-
thetically active cells will also be inuenced by leaf thickness,
interveinal distance, and amount of intercellular airspace
(Noblin etal., 2008). Finally, co-opting some areas of the leaf
for C4 photosynthesis while maintaining water storage cells
will probably allow the C4 descendants to thrive in more arid
conditions (Voznesenskaya etal., 1999; Kadereit etal., 2012).
All of these characteristics, which can be directly affected by
the evolutionary path a species took to achieve C4 function,
will determine the physiology of a plant and thus its ecologi-
cal preferences. Therefore, the diversity of evolutionary tra-
jectories toward C4-compatible leaf anatomy might partially
explain the ecological diversity associated with distinct C4 lin-
eages (e.g. Taub, 2000; Kadereit etal., 2012; Liu etal., 2012).
Consequences for putative genetic
determinism
A detailed discussion of genetic determinants is beyond the
scope of this paper. However, it is worth pointing out that,
despite recent important developments (e.g. Slewinski etal.,
2013; Wang et al., 2013; Lundquist etal., 2014), the genetic
mechanisms necessary to introduce C4-compatible anatomy
into C3 species remain largely unknown. This has particu-
lar implications for the bioengineering of C4 photosynthesis
into major C3 crops, such as rice and wheat, which has the
potential to greatly enhance yield (Covshoff and Hibberd,
2012; von Caemmerer etal., 2012). First, the multiplicity of
traits means that there are probably multiple genes involved.
For instance, a phylogenetic analysis shows that the distance
between consecutive bundle sheaths and the size of these bun-
dle sheaths vary independently in grasses (Table1), suggesting
different underlying genetic changes. Second, as the variation
in most traits presents a continuum from C3 to C4 plants, the
determinism is likely to involve multiple genes with small
effects and no master switch. Third, the diversity of strate-
gies used to achieve leaf functions that are compatible with
C4 photosynthesis means that genetic determinism is likely
to differ among C4 lineages. Finally, the genetic changes that
occur during the evolution of C4 photosynthesis are likely to
vary as a function of the condition in the C3 ancestor.
Interestingly, similar variation in some of the underlying
traits exists in C3 and C4 species, which suggests that useful
genetic variants may be identied from the analysis of C3
taxa that vary in only some of the traits, even if these C3 taxa
do not present C4-like anatomies. For instance, a C3 taxon
with variation in the number of mesophyll cells between
consecutive veins would be a good study system, even if
the bundle sheath and distribution of chloroplasts were not
C4-compatible. Considering variation within C3 taxa that are
unrelated to C4 lineages might therefore expose new ways to
identify the adaptive signicance of individual C4 compo-
nents, as well as their genetic determinism.
Conclusions
Overall, C4 leaves can be dened by a set of important func-
tions that characterize all C4 plants. However, the underlying
developmental characteristics that generate these functional
properties are extremely variable, as a consequence of the tax-
onomic diversity of C4 plants. The same functionally impor-
tant traits are not homologous among all C4 plants, and this
has important implications for the evolution and underlying
genetics of C4-specic leaf anatomy. In addition, the devel-
opmental modications that generate each of the essential
requirements of C4 leaf anatomy can happen independently.
Thus, distantly related C4 groups might arrive at the same
phenotype for one of these requirements (e.g. both groups
co-opt the same compartment for the Calvin cycle) but not
another (e.g. they achieve small distances between the two
compartments through either a reduction in the number of
cells between veins or the development of additional veins).
Most of the anatomical characteristics that can generate
functional properties of C4 leaves exist in at least some C3
plants. The only well-characterized exception is chloroplast
concentration in the compartment co-opted for the segre-
gation of the Calvin cycle, which seems to be specic to C4
plants, and to some extent C2 plants. Without considering the
distribution of chloroplasts and hence C4 physiology, leaves
of C3 and C4 plants cannot be placed into mutually exclusive
categories (see Fig.3, for example), and there is continuous
variation of the underlying traits among C4 and C3 species
(Fig. 4). Hard categorization is meaningful from a func-
tional perspective, but it wrongly suggests that the recurrent
emergence of C4 photosynthesis represents the same number
of drastic transitions between two distinct and homogene-
ous characteristic states. Acknowledging the diversity pre-
sent within both C3 and C4 taxa, and the continuum that
exists between these two physiological states, is paramount
to understanding the evolutionary processes that led to C4
plants, as well as the genetic mechanisms responsible for C4-
compatible leaf anatomy.
Supplementarydata
Supplementary data are available at JXB online.
Supplementary Fig. S1. Cross-sections corresponding to
the diagrams shown in Fig.3.
Acknowledgements
This work was funded by a University of Shefeld Prize Scholarship to
MRL, and a Royal Society University Research Fellowship UF120119 to
PAC. The authors thanks Dr David Chatelet from Brown University and
Professor Travis Columbus from Rancho Santa Ana Botanic Garden who
provided the leaf cross-sections reproduced in the gures.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3367
References
Araus JL, Brown HR, Byrd GT, Serret MD. 1991. Comparative effects
of growth irradiance on photosynthesis and leaf anatomy of Flaveria
brownii (C4-like), Flaveria linearis (C3–C4) and their F1 hybrid. Planta 183,
497–504.
Araus JL, Brown RH, Bouton JH, Serret MD. 1990. Leaf anatomical
characteristics in Flaveria trinervia (C4), Flaveria brownii (C4-like) and their
F1 hybrid. Photosynthesis Research 26, 49–57.
Aubry S, Kneřová J, Hibberd JM. 2013. Endoreduplication is not
involved in bundle-sheath formation in the C4 species Cleome gynandra.
Journal of Experimental Botany 65, 3557–3566.
Beaulieu JM, Leitch IJ, Patel S, Pendharker A, Knight CA. 2008.
Genome size is a strong predictor of cell size and stomatal density in
angiosperms. New Phytologist 179, 975–986.
Berry JA, Downton WJS, Tregunna EB. 1970. The photosynthetic
carbon metabolism of Zea mays and Gomphrena globosa: the location of
the CO2 fixation and the carboxyl transfer reactions. Canadian Journal of
Botany 48, 777–786.
Besnard G, Christin PA, Male PJ, Coissac E, Ralimanana,
Vorontsova MS. 2013. Phylogenomics and taxonomy of Lecomtelleae
(Poaceae), an isolated panicoid lineage from Madagascar. Annals of
Botany 112, 1057–1066.
Boardman NK. 1977. Comparative photosynthesis of sun and shade
plants. Annual Review of Plant Physiology 28, 355–377.
Bolton JK, Brown RH. 1980. Photosynthesis of grass species differing
in carbon dioxide fixation pathways V.Response of Panicum maximum,
Panicum milioides, and tall fescue (Festuca arundinacea) to nitrogen
nutrition. Plant Physiology 66, 97–100.
Botha CEJ. 1992. Plasmodesmatal distribution, structure and frequency
in relation to assimilation in C3 and C4 grasses in southern Africa. Planta
187, 348–358.
Bowes G, Haladay AS, Haller WT. 1979. Seasonal variation in the
biomass, tuber density, and photosynthetic metabolisms of Hydrilla
in three Florida lakes. Journal of Aquatic Plant Management 17,
61–65.
Bowes G, Rao SK, Estavillo GM, Reiskind JB. 2002. C4 mechanisms
in aquatic angiosperms: comparisons with terrestrial C4 systems.
Functional Plant Biology 29, 379–392.
Bowes G, Salvucci ME. 1984. Hydrilla: inducible C4-type photosynthesis
without Kranz anatomy. Advances in Photosynthesis Research 3,
829–832.
Bowes G, Salvucci ME. 1989. Plasticity in the photosynthetic carbon
metabolism of submersed aquatic macrophytes. Aquatic Botany 34,
233–266.
Bowes G. 2011. Single-cell C4 photosynthesis in aquatic plants. In:
Raghavendra AS, Sage RF, eds. C4 photosynthesis and related CO2
concentrating mechanisms. The Netherlands: Springer, 63–80.
Bowman SM, Patel M, Yerramsetty P, Mure CM, Zielinski AM,
Bruenn JA, Berry JO. 2013. A novel RNA binding protein affects rbcL
gene expression and is specific to bundle sheath chloroplasts in C4 plants.
BMC Plant Biology 13, 138.
Boykin LM, Pockman WT, Lowrey TK. 2008. Leaf anatomy of
Orcuttieae (Poaceae: Chloridoideae): more evidence of C4 photosynthesis
without Kranz anatomy. Madroño 55, 143–150.
Brown WV. 1975. Variations in anatomy, associations, and origins of
Kranz tissue. American Journal of Botany 62, 395–402.
Brown WV. 1977. The Kranz syndrome and its subtypes in grass
systematics. Memoirs of the Torrey Botanical Club 23, 1–97.
Brown WV, Smith BN. 1972. Grass evolution, the Kranz syndrome,
13C/12C ratios, and continental drift. Nature 239, 345–346.
Bruhl JJ, Perry S. 1995. Photosynthetic pathway-related ultrastructure
of C3, C4, and C3-like C3-C4 intermediate sedges (Cyperaceae), with
special reference to Eleocharis. Australian Journal of Plant Physiology 22,
521–530.
Carolin RC, Jacobs SWL, Vesk M. 1973. The structure of the cells of
the mesophyll and parenchymatous bundle sheath of the Gramineae.
Botanical Journal of the Linnean Society 66, 259–275.
Carolin RC, Jacobs SWL, Vesk M. 1975. Leaf structure
in Chenopodiaceae. Botanische Jahrbücher für Systematik,
Pflanzengeschichte und Pflanzengeographie 95, 226–255.
Carolin RC, Jacobs SWL, Vesk M. 1977. The ultrastructure of Kranz
cells in the family Cyperaceae. Botanical Gazette 138, 413–419.
Christin PA, Freckleton RP, Osborne CP. 2010. Can phylogenetics
identify C4 origins and reversals? Trends in Ecology and Evolution 25,
403–409.
Christin PA, Osborne CP. 2013. The recurrent assembly of C4
photosynthesis, an evolutionary tale. Photosynthesis Research 117,
163–175.
Christin PA, Osborne CP, Chatelet DS, Columbus JT, Besnard
G, Hodkinson TR, Garrison LM, Vorontsova MS, Edwards EJ.
2013. Anatomical enablers and the evolution of C4 photosynthesis in
grasses. Proceedings of the National Academy of Sciences, USA 110,
1381–1386.
Christin PA, Sage TL, Edwards EJ, Ogburn RM, Khoshravesh R,
Sage RF. 2011. Complex evolutionary transitions and the significance of
C3-C4 intermediate forms of photosynthesis in Molluginaceae. Evolution
65, 643–660.
Covshoff S, Hibberd, JM. 2012. Integrating C4 photosynthesis into C3
crops to increase yield potential. Current Opinion in Biotechnology 23,
209–214.
Crookston RK, Moss DN. 1970. The relation of carbon dioxide
compensation and chlorenchymatous vascular bundle sheaths in leaves of
dicots. Plant Physiology 46, 564–567.
Das VSR, Raghavendra AS. 1976. C4 photosynthesis and a unique type
of Kranz anatomy in Glassocordia boswallaea (Asteraceae). Proceedings
of the Indian Academy of Sciences 84B, 12–19.
Dengler NG, Dengler RE, Donnelly PM, Hattersley PW. 1994.
Quantitative leaf anatomy of C3 and C4 grasses (Poaceae): bundle sheath
and mesophyll surface area relationships. Annals of Botany 73, 241–255.
Dengler NG, Dengler RE, Hattersley PW. 1985. Differing ontogenetic
origins of PCR (“Kranz”) sheaths in leaf blades of C4 grasses (Poaceae).
American Journal of Botany 72, 284–302.
Dengler NG, Donnelly PM, Dengler RE. 1996. Differentiation of bundle
sheath, mesophyll, and distinctive cells in the C4 grass Arundinella hirta
(Poaceae). American Journal of Botany 83, 1391–1405.
Downton WJS, Tregunna EB. 1968. Carbon dioxide compensation-its
relation to photosynthetic carboxylation reactions, systematics of the
Gramineae, and leaf anatomy. Canadian Journal of Botany 46, 207–215.
Duval-Jouve J. 1875. Histotaxie des feuilles des Graminees. Annales des
Science Naturelles Botanical Series 61, 227–346.
Edwards GE, Franceschi VR, Voznesenskaya EV. 2004. Single-cell C4
photosynthesis versus the dual-cell (Kranz) paradigm. Annual Review of
Plant Biology 55, 173–196.
Edwards GE, Ku MSB. 1987. Biochemistry of C3-C4 intermediates.
In: Hatch MD, Boardman NK, eds. The biochemistry of plants:
a comprehensive treatise, Vol. 10. New York: Academic Press,
275–325.
Edwards GE, Voznesenskaya EV. 2011. C4 photosynthesis: Kranz
forms and single-cell C4 in terrestrial plants. In: Raghavendra AS, Sage RF,
eds. C4 photosynthesis and related CO2 concentrating mechanisms. The
Netherlands: Springer, 29–61.
El-Sharkawy M, Hesketh J. 1965. Photosynthesis among species in
relation to characteristics of leaf anatomy and CO2 diffusion resistances.
Crop Science 5, 517–521.
Freitag H, Kadereit G. 2014. C3 and C4 leaf anatomy types in
Camphorosmeae (Camphorosmoideae, Chenopodiaceae). Plant
Systematics and Evolution 300, 665–687.
Freitag H, Stichler W. 2000. A remarkable new leaf type with unusual
photosynthetic tissue in a central Asiatic genus of Chenopodiaceae. Plant
Biology 2, 154–160.
Gong CM, Bai J, Deng JM, Wang GX, Liu XP. 2011. Leaf anatomy and
photosynthetic carbon metabolic characteristics in Phragmites communis
in different soil water availability. Plant Ecology 211, 675–687.
Gould SJ, Vrba ES. 1982. Exaptation—a missing term in the science of
form. Paleontological Society 8, 4–15.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
3368 | Lundgren etal.
Grass Phylogeny Working Group II. 2012. New grass phylogeny
resolves deep evolutionary relationships and discovers C4 origins. New
Phytologist 193, 304–312.
Griffiths H, Weller G, Toy LFM, Dennis RJ. 2013. You’re so vein:
bundle sheath physiology, phylogeny and evolution in C3 and C4 plants.
Plant, Cell & Environment 36, 249–261.
Grime JP, Mowforth MA. 1982. Variation in genome size—an ecological
interpretation. Nature 299, 151–153.
Haberlandt G. 1884. Physiologische Pflanzenanatomie. Leipzig:
Engelmann.
Hatch MD. 1987. C4 photosynthesis: a unique blend of modified
biochemistry, anatomy and ultrastructure. Biochimica et Biophysica Acta
895, 81–106.
Hattersley PW, Browning AJ. 1981. Occurrence of the suberized lamella
in leaves of grasses of different photosynthetic types. I.In parenchymatous
bundle sheaths and PCR (“Kranz”) sheaths. Protoplasma 109, 371–401.
Hattersley PW, Watson L, Johnston FLS, Johnson CR. 1982.
Remarkable leaf anatomical variations in Neurachne and its allies
(Poaceae) in relation to C3 and C4 photosynthesis. Botanical Journal of the
Linnean Society 84, 265–272.
Hattersley PW, Watson L, Osmond CB. 1977. In situ
immunofluorescent labelling of Ribulose-1,5-bisphosphate Carboxylase
in leaves of C3 and C4 plants. Australian Journal of Plant Physiology 4,
523–539.
Hattersley PW, Watson L. 1975. Anatomical parameters for predicting
photosynthetic pathways of grass leaves: the ‘maximum lateral cell
count’ and the ‘maximum cells distant count’. Phytomorphology 25,
325–333.
Hattersley PW, Watson L. 1992. Diversification of photosynthesis.
In: Chapman GP, ed. Grass evolution and domestication. Cambridge:
Cambridge University Press, 38–116.
Hattersley PW. 1984. Characterization of C4 type leaf anatomy in grasses
(Poaceae). Mesophyll: bundle sheath area ratios. Annals of Botany 53,
163–179.
Heckmann D, Schulze S, Denton A, Gowik U, Westhoff P, Weber
AP, Lercher MJ. 2013. Predicting C4 photosynthesis evolution: modular,
individually adaptive steps on a Mount Fuji fitness landscape. Cell 153,
1579–1588.
Hibberd JM. 2007. Cleome, a genus closely related to Arabidopsis,
contains species spanning a developmental progression from C3 to C4
photosynthesis. The Plant Journal 51, 886–896.
Hylton CM, Rawsthorne S, Smith AM, Jones DA, Woolhouse HW.
1988. Glycine decarboxylase is confined to the bundle-sheath cells of
leaves of C3-C4 intermediate species. Planta 175, 452–459.
Kadereit G, Ackerly D, Pirie MD. 2012. A broader model for C4
photosynthesis evolution in plants inferred from the goosefoot family
(Chenopodiaceae s.s.). Proceedings of the Royal Society B: Biological
Sciences 279, 3304–3311.
Kadereit G, Borsch T, Weising K. 2003. Phylogeny of Amaranthaceae
and Chenopodiaceae and the evolution of C4 photosynthesis. International
Journal of Plant Sciences 164, 959–986.
Keeley JE. 1998. C4 photosynthetic modifications in the evolutionary
transition from land to water in aquatic grasses. Oecologica 116,
85–97.
Kellogg EA. 1999. Phylogenetic aspects of the evolution of C4
photosynthesis. In: Sage RF, Monson RK, eds. C4 plant biology. San
Diego: Academic Press, 411–444.
Koteyeva NK, Voznesenskaya EV, Roalson EH, Edwards GE. 2011.
Diversity in forms of C4 in the genus Cleome (Cleomaceae). Annals of
Botany 107, 269–283.
Laetsch WM. 1974. The C4 syndrome: a structural analysis. Annual
Review of Plant Physiology 25, 27–52.
Leegood RC. 2002. C4 photosynthesis: principles of CO2 concentration
and prospects for its introductions into C3 plants. Journal of Experimental
Botany 53, 581–590.
Leegood RC. 2008. Roles of the bundle sheath cells in leaves of C3
plants. Journal of Experimental Botany 59, 1663–1673.
Liu H, Edwards EJ, Freckleton RP, Osborne CP. 2012. Phylogenetic
niche conservatism in C4 grasses. Oecologia 170, 835–845.
Ludwig LJ, Canvin DT. 1971. The rate of photorespiration during
photosynthesis and the relationship of the substrate of light respiration to
the products in sunflower leaves. Plant Physiology 48, 712–719.
Lundquist PK, Rosar C, Brautigam A, Weber APM. 2014. Plastid
signals and the bundle sheath:mesophyll development in reticulate
mutants. Molecular Plant 7, 14–29.
Marshall DM, Muhaidat R, Brown NJ, Liu Z, Stanley S, Griffiths H,
Sage RF, Hibberd JM. 2007. Cleome, a genus closely related to
Arabidopsis, contains species spanning a developmental progression from
C3 to C4 photosynthesis. The Plant Journal 51, 886–896.
Martins S, Scatena VL. 2011. Bundle sheath ontogeny in Kranz and
non-Kranz species of Cyperaceae (Poales). Australian Journal of Botany
59, 554–562.
Masterson J. 1994. Stomatal size in fossil plants: evidence for polyploidy
in majority of angiosperms. Science 264, 421–424.
McKown AD, Cochard H, Sack L. 2010. Decoding leaf hydraulics with a
spatially explicit model: Principles of venation architecture and implications
for its evolution. American Naturalist 175, 447–460.
McKown AD, Dengler NG. 2007. Key innovations in the evolution of
Kranz anatomy and C4 vein pattern in Flaveria (Asteraceae). American
Journal of Botany 94, 382–399.
McKown AD, Dengler NG. 2009. Shifts in leaf vein density through
accelerated vein formation in C4 Flaveria (Asteraceae). Annals of Botany
104, 1085–1098.
Muhaidat R, Sage RF, Dengler NG. 2007. Diversity of Kranz anatomy
and biochemistry in C4 eudicots. American Journal of Botany 94, 362–381.
Muhaidat R, Sage TL, Frohlich MW, Dengler NG, Sage RF. 2011.
Characterization of C3-C4 intermediate species in the genus Heliotropium
L.(Boraginaceae): anatomy, ultrastructure and enzyme activity. Plant, Cell
& Environment 10, 1723–1736.
Murphy LR, Barroca J, Franceschi VR, Lee R, Roalson EH, Edwards
GE, Ku MSB. 2007. Diversity and plasticity of C4 photosynthesis in
Eleocharis (Cyperaceae). Functional Plant Biology 34, 571–580.
Nelson T. 2011. Development of leaves in C4 plants: anatomical features
that support C4 metabolism. In: Raghavendra AS, Sage RF. eds. C4
photosynthesis and related CO2 concentrating mechanisms. The
Netherlands: Springer, 147–159.
Noblin X, Mahadevan L, Coomaraswamy IA, Weitz DA, Holbrook
NM, Zwieniecki MA. 2008. Optimal vein density in artificial and real
leaves. Proceedings of the National Academy of Sciences, USA 105,
9140–9144.
Ocampo G, Koteyeva NK, Voznesenskaya EV, Edwards GE, Sage
TL, Sage RF, Columbus JT. 2013. Evolution of leaf anatomy and
photosynthetic pathways in Portulacaceae. American Journal of Botany
100, 2388–2402.
Ogle K. 2003. Implications of interveinal distance for quantum yield in C4
grasses: a modeling and meta-analysis. Oecologia, 4, 532–542.
Olesen P. 1975. Plasmodesmata between mesophyll and bundle sheath
cells in relation to the exchange of C4-acids. Planta 123, 199–202.
Orme D, Freckleton R, Thomas G, Petzoldt T, Fritz S, Isaac N,
Pearse W. 2012. Caper: Comparative analyses of phylogenetics and
evolution in R Package version 0.5. Available at http://cran.r-project.org/
web/packages/caper/index.html.
Peter G, Katinas L. 2003. A new type of Kranz anatomy in Asteraceae.
Australian Journal of Botany 51, 217–226.
Renvoize SA. 1983. A survey of leaf-blade anatomy in grasses IV.
Eragrostideae. Kew Bulletin 38, 469–478.
Renvoize SA. 1985. A survey of leaf-blade anatomy in grasses V.The
bamboo allies. Kew Bulletin 40, 509–535.
Renvoize SA. 1987a. A survey of leaf-blade anatomy in grasses XI.
Paniceae. Kew Bulletin 42, 739–768.
Renvoize SA. 1987b. A survey of leaf-blade anatomy in grasses X:
Bambuseae. Kew Bulletin 42, 201–207.
Sack L, Scoffoni C, John GP, Poorter H, Mason CM, Mendez-Alonzo
R, Donovan LA. 2013. How do leaf veins influence the worldwide leaf
economic spectrum? Review and synthesis. Journal of Experimental
Botany 64, 4053–4080.
Sack L, Scoffoni C, McKown AD, Frole K, Rawls M, Harvan JC,
Tran H, Tran T. 2012. Developmentally based scaling of leaf venation
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
Deconstructing Kranz anatomy | 3369
architecture explains global ecological patterns. Nature Communications
3, 837.
Sack L, Scoffoni C. 2013. Leaf venation: structure, function,
development, evolution, ecology and applications in the past, present and
future. New Phytologist 198, 983–1000.
Sage RF. 2001. Environmental and evolutionary preconditions for the
origin and diversification of the C4 photosynthetic syndrome. Plant Biology
3, 202–213.
Sage RF, Christin PA, Edwards EJ. 2011. The C4 plant lineages of
planet Earth. Journal of Experimental Botany 62, 3155–3169.
Sage RF, McKown AD. 2006. Is C4 photosynthesis less phenotypically
plastic than C4 photosynthesis? Journal of Experimental Botany 57,
303–317.
Sage RF, Sage TL, Kocacinar F. 2012. Photorespiration and the
evolution of C4 photosynthesis. Annual Review of Plant Biology 63, 19–47.
Sage TL, Busch F, Johnson DC, Friesen PC, Stinson CR, Stata M,
Sultmanis S, Rahman BA, Rawsthorne S, Sage RF. 2013. Initial events
during the evolution of C4 photosynthesis in C3 species of Flaveria. Plant
Physiology 163, 1266–1276.
Schlichting CD, Levin DA. 1986. Phenotypic plasticity: an evolving plant
character. Biological Journal of the Linnean Society 29, 37–47.
Schlichting CD, Pigliucci M. 1993. Evolution of phenotypic plasticity via
regulatory genes. American Naturalist 142, 366–370.
Šímová I, Herben T. 2012. Geometrical constraints in the scaling
relationships between genome size, cell size and cell cycle length in
herbaceous plants. Proceedings of the Royal Society B: Biological
Sciences 279, 867–875.
Sinha NR, Kellogg EA. 1996. Parallelism and diversity in multiple origins
of C4 photosynthesis in the grass family. American Journal of Botany 83,
1458–1470.
Skillman JB. 2008. Quantum yield variation across the three pathways
of photosynthesis: not yet out of the dark. Journal of Experimental Botany
59, 1647–1661.
Slewinski TL, Anderson AA, Zhang C, Turgeon R. 2013. Scarecrow
plays a role in establishing Kranz anatomy in maize leaves. Plant and Cell
Physiology 53, 2030–2037.
Soros CL, Dengler NG. 2001. Ontogenetic derivation and cell
differentiation in photosynthetic tissues of C3 and C4 Cyperaceae.
American Journal of Botany 88, 992–1005.
Sugimoto-Shirasu K, Roberts K. 2003. “Big it up”: endoreduplication
and cell-size control in plants. Current Opinion in Plant Biology 6,
544–553.
Sultan SE. 1987. Evolutionary implications of phenotypic plasticity in
plants. Evolutionary Biology 21, 127–178.
Tateoka T. 1958. Notes on some grasses. VIII. On leaf structure of
Arundinella and Garnotia. Botanical Gazette 120, 101–109.
Taub DR. 2000. Climate and the U.S. Distribution of C4 grass subfamilies
and decarboxylation variants of C4 photosynthesis. American Journal of
Botany 87, 1211–1215.
Taylor SH, Franks PJ, Hulme SP, Spriggs E, Christin PA, Edwards
EJ, Woodward FI, Osborne CP. 2012. Photosynthetic pathway and
ecological adaptation explain stomatal trait diversity amongst grasses.
New Phytologist 193, 387–396.
Terashima I, Miyazawa SI, Hanba YT. 2001. Why are sun leaves thicker
than shade leaves? Consideration based on analyses of CO2 diffusion in
the leaf. Journal of Plant Research 114, 93–105.
Ueno O, Kawano Y, Wakayama M, Takeda T. 2006. Leaf vascular
systems in C3 and C4 grasses: a two-dimensional analysis. Annals of
Botany 97, 611–621.
Ueno O, Samejima M, Muto S, Miyachi S. 1988a. Photosynthetic
characteristics of an amphibious plant, Eleocharis vivipara: expression of
C4 and C3 modes in contrasting environments. Proceedings of the National
Academy of Sciences, USA 85, 6733–6737.
Ueno O, Takeda T, Maeda E. 1988b. Leaf ultrastructure of C4 species
possessing different Kranz anatomical types in the Cyperaceae. Botanical
Magazine 101, 141–152.
Ueno O. 1995. Occurrence of distinctive cells in leaves of C4 species in
Arthraxon and Microstegium (Andropogoneae-Poaceae) and the structural
and immunocytochemical characterization of these cells. International
Journal of Plant Sciences 156, 270–289.
Ueno O. 2001. Environmental regulation of C3 and C4 differentiation in the
amphibious sedge Eleocharis vivipara. Plant Physiology 127, 1524–1532.
Ueno O, Sentoku N. 2006. Comparison of leaf structure and
photosynthetic characteristics of C3 and C4 Alloteropsis semialata
subspecies. Plant, Cell & Environment 29, 257–268.
Vogelmann TC, Nishio JN, Smith WK. 1996. Leaves and light capture:
light propagation and gradients of carbon fixation within leaves. Trends in
Plant Science 1, 65–70.
von Caemmerer S, Furbank RT. 2003. The C4 pathway: an efficient CO2
pump. Photosynthesis Research 77, 191–207.
von Caemmerer S, Quick PW, Furbank RT. 2012. The development
of C4 rice: current progress and future challenges. Science 336,
1671–1672.
Voznesenskaya EV, Franceschi VR, Chuong SDX, Edwards GE.
2006. Functional characterization of phosphoenolpyruvate carboxykinase-
type C4 leaf anatomy: immuno-, chytochemical and ultrastructural analysis.
Annals of Botany 98, 77–91.
Voznesenskaya EV, Franceschi VR, Pyankov VI, Edwards GE. 1999.
Anatomy, chloroplast structure and compartmentation of enzymes relative
to photosynthetic mechanisms in leaves and cotyledons of species in the
tribe Salsoleae (Chenopodiaceae). Journal of Experimental Botany 50,
1779–1795.
Wakayama M, Ueno O, Ohnishi J. 2003. Photosynthetic enzyme
accumulation during leaf development of Arundinella hirta, a C4 grass
having Kranz cells not associated with veins. Plant and Cell Physiology 44,
1330–1340.
Wang P, Kelly S, Fouracre JP, Langdale JA. 2013. Genome-wide
transcript analysis of early maize leaf development reveals gene cohorts
associated with the differentiation of C4 Kranz anatomy. The Plant Journal
75, 656–670.
Weiner H, Burnell JN, Woodrow IE, Heldt HW, Hatch MD. 1988.
Metabolite diffusion into bundle sheath cells from C4 plants relation to
C4 photosynthesis and plasmodesmatal function. Plant Physiology 88,
815–822.
Welkie W, Caldwell M. 1970. Leaf anatomy of species in some
dicotyledon families as related to the C3 and C4 pathways of carbon
fixation. Canadian Journal of Botany 48, 2135–2146.
West-Eberhard MJ, Smith JAC, Winter K. 2011. Photosynthesis,
reorganized. Science 332, 311–312.
Williams BP, Johnston IG, Covshoff S, Hibberd JM. 2013.
Phenotypic landscape inference reveals multiple evolutionary paths to C4
photosynthesis. eLife 2, e00961.
Wilson PJ, Thompson K, Hodgson JG. 2002. Specific leaf area and
leaf dry matter content as alternative predictors of plant strategies. New
Phytologist 143, 155–162.
Woo KC, Anderson JM, Boardman N, Downton WJS, Osmond C,
Thorne SW. 1970. Deficient photosystem II in agranal bundle sheath
chloroplasts of C4 plants. Proceedings of the National Academy of
Sciences, USA 67, 18–25.
at University of Sheffield on April 3, 2015http://jxb.oxfordjournals.org/Downloaded from
... The stored carbon is then moved to a separate cell to be fixed into a four-carbon molecule [27]. The spatial separation of metabolic phases is facilitated by the evolution of Kranz anatomy, a unique cell organization characterized by a concentric layered organization, shown in Figures 2 and 3, where photosynthetic processes are divided into two distinct cell types [28]. CO 2 is first absorbed into the mesophyll (PM) cells, where fixation occurs entirely in C 3 plants, which have thin walls and large intercellular spaces [29]. ...
... This pausing of photosynthesis is detrimental in preserving an organism's essential resources during stress conditions. anatomy, a unique cell organization characterized by a concentric layered organization, shown in Figures 2 and 3, where photosynthetic processes are divided into two distinct cell types [28]. CO2 is first absorbed into the mesophyll (PM) cells, where fixation occurs entirely in C3 plants, which have thin walls and large intercellular spaces [29]. ...
... By drawing inspiration from the physiological aspects of plant biology, where cells serve as the fundamental building materials, it is possible to conceive of buildings as metabolic interfaces that engage in a dynamic exchange with the environment. In this project, the incorporation of the principles of compartmentalization and energetic phasing could enhance a building's environmental performance and adaptability to changing climatic anatomy, a unique cell organization characterized by a concentric layered organization, shown in Figures 2 and 3, where photosynthetic processes are divided into two distinct cell types [28]. CO2 is first absorbed into the mesophyll (PM) cells, where fixation occurs entirely in C3 plants, which have thin walls and large intercellular spaces [29]. ...
Article
Full-text available
In light of the escalating climate crisis, there is a pressing need for a significant shift in how we design the built environment to effectively confront global challenges. Natural systems have inspired scientists, architects, and engineers for centuries; however, conventional biomimetic approaches often focus on superficial aspects, disregarding the underlying complexities. While this approach may lead to a more efficient outcome, it operates under the assumption that the organism functions exclusively within the confines of human knowledge, which are inherently limited by established epistemological and technological systems. This study advocates for a departure from conventional biomimetic approaches and asks the mechanisms of the biological system to inform the process of translation, as opposed to simply defining the outcome. By relinquishing control to material properties and dynamic processes of the biological analog, this study explores the generation of novel, bio-inspired dynamic formworks through non-linear fabrication processes. Specifically, it investigates the thermal properties of accessible building materials, enabling them to respond to environmental conditions without sophisticated technology or human intervention. By embracing chance and unpredictability, translated behaviors are granted the same influence as human intervention. Drawing inspiration from adaptive plant physiology, this research seeks to inspire innovative, climate-responsive methodological practices within broader architectural systems.
... We reported these findings during conferences held in 1964 by the American Society of Agronomy. These ground-breaking discoveries facilitated the consequent research that established the biochemical details of the C 4 system (Hatch et al. 1967, Hatch and Slack 1970, Laetsch 1974, see also Berry 2012, Lundgren et al. 2014, Reeves et al. 2017. " The Biology Book: Big Ideas Simply Explained", DK publishers, London, 2021, The Biology Book: Big Ideas Simply Explained -DK -Google Libros, highlighted in the section on photosynthesis (page 52) the discovery of the link between a particular leaf mesophyll anatomy and the efficiency of photosynthesis among plant species, i.e. ...
... It is about time for younger scientists, scientific organizations, societies, journals, and publishers to take note and correct those mischievous acts! For correct citations, see for example, Berry (2012), Lundgren et al. (2014), Retta et al. (2016), Singh and Reeves (2020), Bano (2022), Liu et al. (2022). Berry et al. (1997), working with the grain amaranth, A. hypochondriacus, reported that Rubisco, C 4 PEPC, and PPDK mRNAs are abundant in meristems and leaf primordia, but are utilized only during specific developmental stages. ...
Preprint
Full-text available
Productivity of most improved major food crops showed stagnation in the past decades after their high-yielding capacity saved millions from famine due to the great achievements and contributions of the 1960s agricultural Green Revolution. As the human population is projected to reach 9-10 billion by 2050, and perhaps even greater by the end of the current century, agricultural productivity must be increased by as much as was achieved during the past 10,000 years to ensure these demands, especially considering escalating climate change challenges. Photosynthetic capacity is the basic process underlying primary biological productivity in green plants, and enhancing it might lead to increasing potential crop yield. Since yields are a function of plant genetics, crop management practices, and environmental conditions, there are several approaches to improve the photosynthetic capacity, including integrated agroecosystems management, to close the wide gaps between actual farmers and the optimum obtainable yield. Conventional and molecular genetic improvement to increase leaf P N is a viable approach, which has been recently shown in a few crops. Bioengineering the more efficient C 4 into the C 3 system is another ambitious approach and is currently being applied to the C 3 rice crop. Two under-researched, yet old important crops native to the tropic Americas (i.e., the C 4 amaranths and the C 3-C 4 intermediate cassava), have shown high potential P N , high productivity, high water use efficiency, and tolerance to heat and drought stresses. These physiological traits make them suitable for future agricultural systems, particularly in a globally warming climate. In the face of accelerating climate change and ever-increasing world population, there is an urgent need to further diversify food, feed, and energy resources by taping the potential of agriculturally unutilized plant species, particularly when genetic resources are at risk. Exciting contributions to the C 3 :C 4 photosynthetic phenomenon were made at Tucson AZ, and Davis CA, USA, and at Cali, Colombia, S.A., along with the work at the Hawaiian and Australian Sugar Planter Associations. Work on crop canopy photosynthesis including that on flowering genes, that control the formation and decline of the canopy photosynthetic activity, has contributed to the climate change research effort. Associated drought effects on crop canopy photosynthetic behavior were studied by scientists at Stoneville and Starkville MS, USA, where the first cotton models were developed. In other words, the so-called photosynthetic establishment became dispersed worldwide among agricultural scientists who are mainly breeders and agronomists. It is recommended therefore that the plant breeders need to select for higher P N 2 to enhance yield and crop tolerance to environmental stresses, as anticipated in globally warming climates whose adverse effect is greater in the tropic/subtropic compared to the temperate regions. Also, experts in plant diseases and pests, soil sciences, meteorology, and crop modeling must cooperate toward developing sound integrated crop management systems. The plant science instructors, and researchers, for various reasons, need to focus more on tropical species and to use the research, highlighted here, as an example of how to increase their yields.
... The NAD-ME subtype decarboxylate malate to pyruvate in mitochondria using a NAD-dependent malic enzyme (Rao and Dixon, 2016). In addition, NAD-ME present an outer bundle sheath (lacking of suberin lamella) and an inner non-photosynthetic mestome sheath covered by suberin (Lundgren et al., 2014;Fouracre et al., 2014;Lundgren et al., 2014;Mertz and Brutnell, 2014). Because of this anatomical arrangement, leakage (Φ) is probably lower in NADP-ME than in NAD-ME (Ghannoum et al., 2002). ...
... The NAD-ME subtype decarboxylate malate to pyruvate in mitochondria using a NAD-dependent malic enzyme (Rao and Dixon, 2016). In addition, NAD-ME present an outer bundle sheath (lacking of suberin lamella) and an inner non-photosynthetic mestome sheath covered by suberin (Lundgren et al., 2014;Fouracre et al., 2014;Lundgren et al., 2014;Mertz and Brutnell, 2014). Because of this anatomical arrangement, leakage (Φ) is probably lower in NADP-ME than in NAD-ME (Ghannoum et al., 2002). ...
... The mestome sheath cells are absent in the NADP-ME subtype. Differences in grana development and chloroplast arrangement can be seen within the subtypes (Gutierrez et al. 1974;Hattersley and Watson 1976;Lundgren et al. 2014). The cellular developmental origins of the same cells from different subtypes vary. ...
Article
Full-text available
Plant physiologists set about comprehending the genesis of the C4 photosynthetic pathway after its discovery by Hatch and Slack. They discovered that a sophisticated combination of morphological and biochemical adaptations allowed the plant to concentrate CO 2 around RuBisCO to achieve maximum efficiency. We categorize the evolutionary events leading to C4 photosynthesis, beginning with anoxygenic photosynthesis and the evolution of RuBisCO to the cooling of Earth by the Great Oxygenation Event that led to the oxygenic photosynthesis. The evolutionary descent of the C4 plants is a phenomenon that occurred around 30 million years ago. Due to industrialization and population growth, improved photosynthetic efficiency and carbon fixation of C4 plants could contest the current global scenario of rising CO 2 concentration. C3 crops engineered with C4 traits, implemented on a large scale, could impact the climate globally. Here we discuss the various strategies used to introduce C4 traits in the C3 plants and the potential techniques to be considered for successful hybridization.
... Plants using C 4 photosynthesis have specialised leaf anatomy, typically characterised by high vein density and an enlarged layer of BS cells, rich in specialised chloroplasts, surrounding each vascular bundle (Lundgren et al., 2014;Muhaidat et al., 2007). Thus, chloroplast proliferation in BS cells is one of the key requirements to engineer C 4 photosynthesis in C 3 crops. ...
Article
Full-text available
Societal Impact Statement The human population is expected to reach 9.7 billion in the next 30 years, increasing the strain on our already precarious food system. Climate change is shifting weather patterns, leading to unpredictable and catastrophic events that further threaten the agronomic sector. Plant scientists are implementing biotechnological tools to sustainably increase both the production and nutritional content of our crops. Engineering GOLDEN2‐LIKE (GLK) transcription factors is a promising route to improve photosynthesis, as well as other important agronomical traits, to achieve food security for a growing population under an unpredictable climate. Summary Using agricultural biotechnology to increase the photosynthetic efficiency of crops has been a focus of plant science research over the last two decades. Transcription factors coordinate the expression of gene networks that are the basis of plant development and physiological responses and, as such, are good targets to help improve photosynthesis. Among the known plant transcriptional regulators, GOLDEN2‐LIKE transcription factors (GLKs) may be ideal candidates to improve photosynthesis in crops, as they are master regulators of genes associated with photosynthesis and chloroplast biogenesis across a broad diversity of plant lineages. Moreover, recent work has revealed their involvement in environmental response, pathogen defence and development regulation across the plant's whole life cycle. Thus, manipulating GLK expression and activity, alone or likely in combination with other modifications, has clear potential to improve plant development and growth. Here, we review the research into GLK function and discuss the potential of these key transcription factors as biotechnological tools to enhance photosynthetic efficiency and stress tolerance in crops. Additionally, we take advantage of the vast plant genome and transcriptome datasets available to explore the evolutionary history of GLKs across the plant kingdom and discuss the implications for their adoption into crop engineering projects.
Article
Full-text available
Kodo millet (Paspalum scrobiculatum L.) is an underutilized crop that encompasses nutritional benefits and climate resilience, making it a viable option for future crop development with nutraceutical properties. The cultivation of this crop has ancient roots, where it was revered for its ability to thrive in times of famine and was a vital companion crop to rice. Dishes made with Kodo millet are highly palatable and can be easily integrated into mainstream rice-based dishes. Among all cereals, Kodo millet is distinguished by its gluten-free composition, high phosphorus content, and significant antioxidant potential, which contributes to a diet that may reduce cardiovascular disease risk. Often grown in rainfed zones by marginal farmers, Kodo millet is valued for its grain and fodder. This less demanding crop can tolerate both biotic and abiotic stress, allowing it to thrive in soils with low organic matter and with minimal inputs, making it an ideal dual-purpose crop for rainfed areas. Despite its nutritional and agricultural benefits, Kodo millet's popularity is hindered by challenges such as low yield, market demand, lodging at harvest, and poor dehulling recovery, which necessitate the development of high-yielding varieties through the latest breeding advancements. Systematic investment and concerted breeding efforts are essential to harness the full potential of this nutrient-dense crop. The absence of whole genome sequence for Kodo millet poses a barrier to uncovering novel genetic traits. Consequently, there is an imperative to establish a millet-based value chain that elevates these underutilized crops, shaping smart cropping patterns and enhancing nutritional profiles for sustainable diets. Accordingly, this review highlights the significance of Kodo millet and the impact of breeding to establish it as a smart food choice for the future.
Article
Full-text available
Carbon concentrating mechanisms enhance the carboxylase efficiency of the central photosynthetic enzyme rubisco by providing supra-atmospheric concentrations of CO2 in its surrounding. In the C4 photosynthesis pathway, this feat is realised by combinatory changes to leaf biochemistry and anatomy. In contrast to the C4 pathway, carbon concentration can also be achieved by the photorespiratory glycine shuttle which requires fewer and less complex modifications. Plants displaying CO2 compensation points between 10 to 40 ppm are often considered to utilize such a photorespiratory shuttle and are termed 'C3-C4 intermediates'. In the present study, we perform a physiological, biochemical and anatomical survey of a large number of Brassicaceae species to better understand the C3-C4 intermediate phenotype, including its basic components and its plasticity. Our phylogenetic analysis suggested that C3-C4 metabolism evolved up to five times independently in the Brassicaceae. The efficiency of the pathway showed considerable variation between tested plant species. Centripetal accumulation of organelles in the bundle sheath was consistently observed in all C3-C4 classified taxa indicating a crucial role of anatomical features for CO2 concentrating pathways. Leaf metabolite patterns were strongly influenced by the individual species, but accumulation of photorespiratory shuttle metabolites glycine and serine was generally observed. Analysis of PEPC activities and metabolite composition suggests that C4-like shuttles have not evolved in the investigated Brassicaceae. Convergent evolution of the photorespiratory shuttle indicates that it represents a distinct and fit photosynthesis type.
Article
Although the unique tissue required for C4 photosynthesis in nonsucculent plants is often described as being modified leaf parenchyma sheath, which is positioned meaningfully between the mesophyll externally and the vascular tissues internally, the actual range of locations and known associations make that concept untenable. In origin the Kranz tissue develops from procambium as well as ground parenchyma. It is found in stems as well as leaves. In position Kranz tissue can lie in the parenchyma sheath, in the mestome sheath, isolated in the mesophyll, peripherally in some thick leaves, or within the veins. It can be associated with mesophyll only, mesophyll and colorless parenchyma, mesophyll and sclerenchyma, other Kranz tissue and vascular tissues, mesophyll and mestome sheath, mesophyll and phloem, mesophyll and xylem, epidermis, and, finally, mestome sheath and xylem and phloem. The use of the term Kranz is expounded.
Article
C4 photosynthesis is thought to be an adaptation to warm environments, involving complex changes in the expression of genes governing photosynthesis, intermediary metabolism, and leaf anatomy and histology. Such complexity should be difficult to evolve, yet the pathway has arisen multiple times in the history of the flowering plants and at least four times in the grass family alone. We have used immunolocalization techniques to compare photosynthetic gene expression across all four origins, to determine which genetic changes occur in parallel and which are unique to a particular lineage. The only gene expression patterns common to all origins of the pathway are up-regulation of PEP carboxylase and down-regulation of RuBisCO in mesophyll cells. Both NAD-malic enzyme and NADP-malic enzyme are expressed in bundle sheaths. Expression patterns of light-harvesting chlorophyll a/b binding proteins and pyruvate orthophosphate dikinase appear to be lineage specific, and may be localized to bundle sheaths or to mesophyll or expressed throughout the photosynthetic tissue of the leaf. We suggest that future studies of parallel origin of the C4 pathway concentrate on regulation of the two carboxylases, as well as the increased density of vascular tissue, which is the only histological characteristic common to all origins of the pathway.
Article
The C4 grass Arundinella hirta is characterized by unusual leaf blade anatomy: veins are widely spaced and files of bundle-sheath-like cells, the distinctive cells, form longitudinal strands that are not associated with vascular tissue. While distinctive cells (DCs) appear to function like bundle sheath cells (BSCs), they differ developmentally in two ways: they are derived from ground meristem rather than procambium and they are formed 1–2 plastochrons later. This study describes ultrastructural features of differentiating of BSCs, DCs, and associated mesophyll cells (MCs) during leaf development. BSCs and DCs differ from adjacent MCs by undergoing earlier cell enlargement, greater rates of chloroplast enlargement, reduction of chloroplast thylakoids at late stages of differentiation, more extensive starch formation, greater wall thickening, and deposition of a suberin lamella. The precocious delimitation of the bundle sheath layer is reflected in earlier BSC enlargement and vacuole growth. Derivation of DCs from ground meristem is correlated with late developmental changes in chloroplast size, wall thickness, and plasmodesmatal density. Despite these differences in timing of events, particularly at early stages, the development of the specialized structural features of BSCs and DCs is essentially similar. Thus, proximity to vascular tissue appears to be nonessential for the coordination and regulation of BSC- and MC-specific developmental events.
Article
The origin and early development of procambium and associated ground meristem of major and minor veins have been examined in the leaf blades of seven C4 grass species, representing different taxonomic groups and the three recognized biochemical C4 types (NAD-ME, PCK, and NADP-ME). Comparisons were made with the C3 species, Festuca arundinacea. In “double sheath” (XyMS+) species (Panicum effusum, Eleusine coracana, and Sporoboìus elongatus), the procambium of major veins gives rise to xylem, phloem, and a mestome sheath; associated ground meristem differentiates into PCA (“C4 mesophyll”) tissue and the PCR (“Kranz”) sheath. Development in the C3 species parallels this pattern, except that associated ground meristem differentiates into mesophyll and a parenchymatous bundle sheath. In contrast, major vein procambium of “single sheath” (XyMS–) species (Panicum bulbosum, Digitaria brownii, and Cymbopogon procerus) differentiates into xylem, phloem and a PCR sheath; associated ground meristem gives rise to PCA tissue. These observations of major vein development support W. V. Brown's hypothesis that the PCR sheaths of “double sheath” (XyMS+) C4 grasses are homologous with the parenchymatous bundle sheaths of C3 grasses, while in “single sheath” (XyMS–) C4 species they are homologous with the mestome sheath. Although there are some similarities in the development of the major and minor vascular bundle procambium in the C4 species examined, the ontogeny of the smaller minor veins is characterized by a precocious delineation of the PCR sheath layer that may even precede the appearance of the distinctive cytological features of ground meristem and procambium. This contracted development in minor veins appears to be related to their close spacing in mature leaves and to their comparatively late appearance during leaf ontogeny.
Article
The cross-sectional area of ‘primary carbon assimilation’ (PCA) (or mesophyll) tissue and of ‘photosynthetic carbon reduction’ (PCR) (or parenchymatous bundle sheath, PBS) tissue associated with each vein has been measured in transverse sections of leaf blades of 124 grass species (Poaceae). The species sample is representative of all major grass taxa, and of all photosynthetic types found in this family, viz. C3, C3/C4 intermediate, C4 NADP-malic enzyme type (NADP-ME), C4 NAD-malic enzyme type (NAD-ME) and PEP carboxykinase type (PCK). Mean PCA (or mesophyll) area per vein varies between photosynthetic types in the order C3 > NAD-ME > PCK = NADP-ME, mean PCR (or PBS) area per vein in the order NAD-ME > PCK = C3 > NADP-ME, and mean PCA/PCR (or mesophyll/PBS) area ratio in the order C3 > NADP-ME > NAD-ME > PCK. Since grass leaves have parallel venation, tissue areas and area ratios are directly proportional to tissue volumes and volume ratios. Regression analyses of plots of PCA (or mesophyll) area per vein against PCR (or PBS) area per vein yield characteristic slopes for photosynthetic types. Differences between types in all these parameters are nearly always statistically significant, even within high level taxonomic groups (Eupanicoids and Chloridoids). However, differences between major taxa (Eupanicoids, Andropogonoids, Chloridoids), within a photosynthetic type, are frequently not significant. This histometric characterization of photosynthetic types is discussed in relation to the co-operation of PCA and PCR tissues in C4 photosynthesis, to possible differences between C4 types in PCR spatial requirements and to the developmental origin of PCR tissue.