You are on page 1of 11

The Plant Cell, Vol. 16, S61–S71, Supplement 2004, www.plantcell.

org ª 2004 American Society of Plant Biologists

Sex-Determining Mechanisms in Land Plants

Milos Tanurdzic and Jo Ann Banks1


Purdue Genetics Program and Department of Botany and Plant Pathology, Purdue University, West Lafayette, Indiana 47906

INTRODUCTION THE BRYOPHYTES

Sex determination is a process that leads to the physical The bryophytes are a group of plants that includes the modern
separation of male and female gamete-producing structures to liverworts, hornworts, and mosses. In this group, the haploid
different individuals of a species. Even though sexually repro- gametophyte is the dominant phase of the life cycle, which is
ducing species have only three possible options—to relegate the illustrated in Figure 1. Their diploid spore-producing sporophytes
two sexes to separate individuals, to keep them together on are very small, short-lived, and parasitic upon the independent
the same individual, or to have a combination of both—plants gamete-producing gametophyte generation. All bryophytes are
in particular display a great variety of sexual phenotypes. In homosporous, producing only one type of spore, yet all three
angiosperms, a sex-determining process is manifest in species groups of extant bryophytes are represented by species that are
that are monoecious, in which at least some flowers are homothallic, with one individual gametophyte producing both
unisexual but the individual is not, or dioecious, in which male and female sex organs (gametangia), and species that are
unisexual plants produce flowers of one sex type. In plants that heterothallic, with one individual producing only male or female
produce no flowers and are homosporous, sex determination is gametangia (Smith, 1955). Sexual dimorphism in heterothallic
manifest in the gametophyte generation with the production of species can be extreme, as exemplified by members of the
egg- and sperm-forming gametangia on separate individual genus Micromitrium, in which the dwarf male gametophyte
gametophytes. The determinants of sexual phenotype in plants grows on the leaves of the markedly larger female plant (Smith,
are diverse, ranging from sex chromosomes in Marchantia 1955).
polymorpha and Silene latifolia to hormonal regulation in Zea In many species of bryophytes, heterothallism (unisexuality)
mays and Cucumis sativa to pheromonal cross-talk between has been correlated with the presence of sex chromosomes
individuals in Ceratopteris richardii. Here, we highlight recent (Smith, 1955). Although the extent of heterothallism and sex
efforts aimed at understanding the genetic and molecular chromosomes in the bryophytes has not been assessed
mechanisms responsible for sex determination in several plant systematically, this is the only known group of homosporous
species that separate their sexes into two individuals or flowers. plants that uses sex chromosomes in sex determination. To
Representatives of all major land plant lineages are included to date, studies of bryophyte sex determination have focused on
give an evolutionary perspective, which is important in under- the heterothallic liverwort Marchantia polymorpha. In this
standing how different sex-determining mechanisms evolved species, the male and female thalli (vegetative gametophytes)
and their consequences in plant development and evolution. look alike, although males and females can be distinguished
Although great progress has been made in genetically identifying easily by differences in the morphology of the sexual structure
the genes that regulate sex expression in these species, few of each produces. A gametophyte bears gametangia on stalked
them have been cloned. Because a ‘‘one-size-fits-all’’ mecha- branches called antheridiophores (if male) or archegonio-
nism of sex determination will not account for the variety of phores (if female) that arise from the upper surface of the
sexual systems in plants, future efforts at cloning these genes thallus (Figure 1). Antheridiophores produce sperm-forming
in several well-chosen model systems will be necessary to antheridia, and archegoniophores produce egg-forming arche-
understand these processes at the molecular level. There are gonia. The sex of each haploid gametophyte is determined by
several excellent recent reviews of sex determination that cytologically distinct sex chromosomes, with males having one
describe species that have not been included to which the very small Y chromosome and no X chromosome and females
reader is directed (Ainsworth, 1999, 2000; Geber et al., 1999; having one X chromosome and no Y chromosome (Lorbeer,
Matsunaga and Kawano, 2001; Negrutiu et al., 2001; Barrett, 1934). In addition to its rapid growth (it is often an invasive
2002; Charlesworth, 2002). We begin with the most basal lineage weed in greenhouses), its ability to be propagated vegetatively
of the land plants. by gemma cups (Figure 1), and its ability to be transformed
(Takenaka et al., 2000), Marchantia has a relatively small
genome size of 280 Mbp distributed among eight autosomes
plus one sex chromosome (Okada et al., 2000), making it
1 To
a worthy model organism amenable to genomics-style inves-
whom correspondence should be addressed. E-mail banksj@
purdue.edu; fax 765-494-5896.
tigations.
Article, publication date, and citation information can be found at Working on the assumption that sex-determining factors exist
www.plantcell.org/cgi/doi/10.1105/tpc.016667. on the Marchantia sex chromosomes, Okama and colleagues
S62 The Plant Cell

Figure 1. The Life Cycle of the Liverwort Marchantia polymorpha.


Haploid gametophytes develop gametangia in antheridiophores and archegoniophores that produce the sperm and egg, respectively. Upon
fertilization, which is facilitated by raindrops, the diploid sporophyte remains attached to the archegoniophore and produces yellow sporangia, in which
haploid spores are formed after meiosis. The spores are liberated and germinate to form a new gametophyte thallus. The sex of the thallus depends on
which sex chromosome it inherits. Photographs courtesy of Katsuyuki Yamato, Kyoto University.

(2000) set out to identify these factors by constructing separate M2D3.4 encodes a putative protein similar to a Lilium longiflorum
male and female P1-derived artificial chromosome (PAC) gene that is expressed exclusively in the male gametic cells. The
libraries and identifying clones specific to either the male or the remaining three genes are not sex specific in their expression.
female genome. Their screen resulted in 70 male-specific PAC The functions of the Y chromosome–encoded genes are as yet
clones that hybridized only the Y chromosome by fluorescence in unknown.
situ hybridization. No female-specific clones were found, in- Although only a small portion of the Marchantia Y chromosome
dicating that the X chromosome does not harbor long stretches has been sequenced, it is sufficient to make meaningful
of unique sequences, as does the Y chromosome. To date, two comparisons with the euchromatic male-specific region (MSY)
male-specific PAC clones with insert sizes totaling 126 kb have of the human Y chromosome, the sequence of which was
been sequenced (Okada et al., 2001; Ishizaki et al., 2002). This published recently (Skaletsky et al., 2003; see also Hawley,
and other analyses have revealed that approximately one-fourth 2003). The sequence of the human MSY region and limited
to one-third of the 10-Mb Y chromosome of Marchantia consists comparative sequencing of the MSY regions in great apes
of an estimated 600 to 15,000 copies of an element of variable (Rozen et al., 2003) have added new insights to our understand-
length (0.7 to 5.2 kb) that contains other smaller repetitive ing of how the mammalian testis gene families on the Y
elements (Okada et al., 2001; Ishizaki et al., 2002). Of the six chromosome have been maintained over the course of evolution.
putative protein-encoding genes found embedded within the As will be shown, the similarities between the human and
repeats, all are present in multiple copies on the Y chromosome liverwort Y chromosomes are striking and may reflect a common
based on DNA gel blot hybridization. Two of these genes, named mechanism underlying the evolution of the Y chromosome in
ORF162 (Okada et al., 2001) and M2D3.5 (Ishizaki et al., 2002), these two disparate organisms. The MSY region of the human Y
are unique to the Y chromosome; the remaining four genes are chromosome is made up of three classes of sequences: X
present in low copy number on the X chromosome or the transposed, X degenerate, and ampliconic, the latter represent-
autosomes. ORF162 encodes a putative protein with a RING ing 30% of the MSY euchromatin. Because the Marchantia X
finger domain; M2D3.5 is a member of the same gene family. chromosome has not been sequenced, it is only possible to
ORF162 transcripts are detectable only in the male sexual make comparisons between the human ampliconic sequences,
organs, indicating that the gene family represented by ORF162 which are Y specific, and the Marchantia Y chromosome
and M2D3.5 may be important in the development of the sequences. Like the Marchantia Y chromosome sequences,
antheridiophore. Of the four genes also present on the X the ampliconic regions of the MSY consist of highly repetitive
chromosome or the autosomes, only one (M2D3.4) is restricted sequences unique to the Y chromosome, although the sizes,
in its expression to the male gametophyte, indicating that the sequences, and stoichiometries of the repetitive elements very
M2D3.4 X or automosmal homolog might be a pseudogene. considerably between the two species. Protein-encoding genes
Sex Determination S63

or gene families (six genes in Marchantia and nine gene families great potential as a useful comparative system for the study of
in human) occur within repetitive elements, and all are present sex determination in plants, in part because many species have
in multiple copies on the Y chromosome. In both organisms, very small genome sizes (J.A. Banks, unpublished observations).
homologs also may be present on the X or autosomal chromo- Selaginella produces microsporangia and megasporangia that
somes in low copy number. Although all protein-encoding are born on the same strobilus or in different strobili of the same
genes found within the human ampliconic sequences are ex- plant, depending on the species; strictly unisexual species have
pressed only in the testis, at least some of the protein-encoding not been reported. In Selaginella moellendorfii, for example, each
genes identified in Marchantia are male organ specific in their strobilis bears both kinds of sporangia, with microsporangia at
expression. the bottom and macrosporangia toward the top of each strobilis.
According to the prevailing theory of mammalian sex chromo- In such plants, sex determination would be viewed as the
some evolution (Graves and Schmidt, 1992; Jegalian and Page, process that regulates the sexual identity of the sporangia in the
1998; Lahn and Page, 1999), the X and Y chromosomes are strobilis, a mechanism that has clear parallels with floral organ
derived from an ancient autosomal pair of chromosomes. The Y identity in angiosperms that produce perfect flowers. Although
chromosome acquired genes, especially those that enhance we know virtually nothing about this group of plants beyond
male fertility, by a series of autosomal transpositions that were descriptive biology, this lineage holds important clues for
then amplified on the Y chromosome, whereas the X chromo- understanding the evolution of heterospory from homospory,
some maintained its ancestral genes. Other genes present on the a switch that occurred many times during land plant evolution
Y chromosome (i.e., X homologs) were lost, probably aided by (Stewart and Rothwell, 1993) and that has had a major impact on
a lack of X-Y recombination, leading to a mostly degenerate Y the timing of sex determination from the gametophyte to
chromosome. The recent sequencing results suggest that the sporophyte generation in plants (Sussex, 1966). The question
uniformity of the ampliconic repetitive sequences of the human Y of how heterospory evolved from homospory is difficult to study
chromosome is maintained from generation to generation by in the heterosporous angiosperm lineage because their homo-
intrachromosomal Y-Y gene conversion. This occurs at relatively sporous progenitors are probably extinct.
high rates: 600 nucleotides per newborn human male are
estimated to have undergone Y-Y gene conversion (Rozen et al., THE HOMOSPOROUS FERNS
2003). Although it is not known if there are higher order
palindromic repeated sequences in Marchantia as there are in Recent phylogenetic analyses of vascular seed–free plants
humans (these palindromic sequences may be necessary for group the leptosporangiate and eusporangiate ferns and
gene conversion), the homogeneity of repetitive sequences in the members of Equisetum and Psilotum into a monophyletic clade
Marchantia Y chromosome and the relatively high frequency of that is sister to the seed plants (Pryer et al., 2001). With few
male-specific genes within these repetitive elements suggest exceptions, they are homosporous plants. The one plant for
gene conversion playing a role in maintaining the repetitive which a sex-determining pathway has been genetically well
elements of the Marchantia Y chromosome. defined is the leptosporangiate fern Ceratopteris richardii. Like
One important distinction between Marchantia and humans is Marchantia, Ceratopteris is homosporous and produces only
that X-X recombination cannot occur in Marchantia because all one type of haploid spore. Although the sex of the Marchantia
diploid sporophytes are X-Y. This suggests that forces other than gametophyte is determined genetically by sex chromosomes,
homologous interchromosomal recombination are responsible the sex of the Ceratopteris gametophyte (male or hermaphro-
for preventing the degeneration of the X chromosome, although ditic) is determined epigenetically by the pheromone antheridi-
size alone may not be an indicator of chromosome degeneracy. ogen. Since their discovery by Dopp (1950) in the fern Pteridium
Future efforts to sequence the entire Y and X chromosomes in aquilinum, antheridiogens have been identified and character-
Marchantia will be very important in understanding how sex ized from many species of leptosporangiate ferns (reviewed by
chromosomes specify sexual phenotype in this species and how Naf, 1979; Yamane, 1998), suggesting that it is a common mode
both the X and Y chromosomes evolved in this ancient lineage of of regulating sexual phenotypes in this group of plants. Although
plants. the structure of the Ceratopteris antheridiogen is unknown, all
other fern antheridiogens characterized to date are mostly novel
THE LYCOPHYTES gibberellins (Yamane, 1998).
The Ceratopteris male and hermaphroditic gametophytes are
Another group of land plants deserving attention from an easy to distinguish not only by the type of gamete produced but
evolutionary perspective is the lycophyte lineage, which includes also by the presence or absence of a multicellular meristem. As
the modern Lycopodiales genera Selaginella and Isoetes. This illustrated in Figure 2, the hermaphrodite forms a single meri-
lineage is most closely related to the earliest vascular plants that stem and meristem notch that gives the hermaphrodite its
first appeared on land 250 to 400 million years ago (Kenrick heart-shaped appearance. Cells of this meristem differentiate as
and Crane, 1997). Although the earliest lycophytes and extant egg-forming archegonia, sperm-forming antheridia, or simply
members of the Lycopodiales are homosporous and produce enlarge, adding to the growing sheet of photosynthetic paren-
only one type of spore, Selaginella and Isoetes are heterospo- chyma cells that make up most of the hermaphrodite prothallus.
rous, with their sporophytes producing free-living megaspores A Ceratopteris spore grown in isolation always develops as
and microspores that give rise to the female and male a hermaphrodite. A male gametophyte develops from a spore
gametophytes, respectively. Of the lycophytes, Selaginella has only if the spore is placed in medium that had previously
S64 The Plant Cell

Figure 2. The Sex-Determining Mutants of Ceratopteris richardii.

The her1 (hermaphroditic) mutant and the wild-type hermaphrodite are indistinguishable, as are the tra1 (transformer) mutant and the wild-type males,
except that the her1 and tra1 mutants are insensitive to the absence or presence of ACE. The ACE-insensitive fem1 (feminization) gametophyte produces
no antheridia. The man1 (many antheridia) mutant produces 10 times more antheridia than hermaphrodites, whereas the not1 (notchless) mutant
rarely produces antheridia. The meristem notch normally present on the hermaphrodite often is missing in the not1 mutant, giving it a cup-shaped
appearance. The novel phenotypes of the fem1 tra1 and fem1 not1 tra1 mutants are shown. an, antheridia; ar, archegonia; mn, meristem notch.

supported the growth of a hermaphrodite. Lacking a multicellular tra1 double mutant, also illustrated in Figure 2. Unlike other
meristem, almost all cells of the male gametophyte terminally double mutant combinations, this one has a novel phenotype
differentiate as antheridia. The male-inducing pheromone that unlike that of either parent. This finding suggests that these two
is secreted by the Ceratopteris hermaphrodite is called ACE for genes (FEM1 and TRA1) define two separate pathways, one
antheridiogen Ceratopteris. Based on physiological studies in specifying male development and the other female development.
Ceratopteris (Banks et al., 1993), ACE is not secreted by the The sex-determining mutants have been ordered into a genetic
hermaphrodite until after it loses the competence to respond to sex-determining pathway, illustrated in Figure 3, that is most
its male-inducing effects, which corresponds to the initiation of consistent with the genetic data. In this pathway, the sex of the
the meristem. A gametophyte will develop as a male only if it is gametophyte ultimately depends on the activities of two genes,
exposed continuously to ACE from a very young age, between 2 one specifying the development of male traits (FEM1) and the
to 4 days after spore inoculation. Thus, in a population of spores, other specifying the development of female traits (TRA). These
those that germinate first become ACE-secreting meristic genes also repress each other, so that when TRA is active, FEM1
hermaphrodites, whereas those that germinate later become is not and visa versa. What determines which of these two genes
ameristic males under the influence of ACE secreted by its is expressed in the gametophyte (and thus its sex) is ACE, which
neighboring hermaphrodites. ultimately represses the TRA genes, as described in the legend
To understand how ACE represses the development of fe- to Figure 3.
male traits (i.e., archegonia and meristem) and promotes the In comparing mechanisms of gametophytic sex determination
development of male traits (i.e., antheridia) in Ceratopteris, in homosporous bryophytes and ferns, one obvious question
a genetics approach has been used to identify the genes that arises is what drove Marchantia to an X-Y chromosomal
involved in this response (Banks, 1998; Strain et al., 2001). To mechanism of sex determination and Ceratopteris to an epi-
date, five phenotypic classes of mutants have been identified; genetically regulated mechanism dependent on pheromonal
representatives of each class are illustrated in Figure 2. In cross-talk between individuals? The answer to this question
addition to those that are always hermaphroditic (the hermaph- probably lies in the different ratios of males and females or
roditic mutants), always male (the transformer [tra] mutants), or hermaphrodites that occur in the populations of each species. In
always female (the feminization [fem] mutants) regardless of the Marchantia, the segregation of X and Y sex chromosomes during
absence or presence of ACE, there are mutants that produce meiosis in the sporophyte ensures that each gametophyte
excessive antheridia (the many antheridia mutants) as well as progeny has an equal probability of being either male or female,
the feminizing mutants that often lack a meristem notch (the barring selection. In Ceratopteris, the ACE response allows
notchless mutants). By comparing the phenotypes of double the ratio of males to hermaphrodites to vary depending on
mutant gametophytes to each single mutant gametophyte the density of the population, such that as the population den-
parent, the epistatic interactions among these genes have been sity increases, the proportion of males also increases. Although
assessed. One particularly informative phenotype is the fem1 the underlying sex-determining mechanism is inflexible in
Sex Determination S65

ratio of X to autosomal chromosomes. This ratio is read and


either activates or represses the activities of downstream genes
in each pathway. In both animals, the sex ultimately depends on
the state of the terminal gene in each linear pathway, TRA1 in the
case of C. elegans. The Ceratopteris sex-determining pathway is
distinctly different from those of D. melanogaster and C. elegans
in that it is not linear and there are two sex-determining genes,
one for male and one for female development. Their ability to
repress each other endows each Ceratopteris spore with the
flexibility to determine its sex upon germination based on
environmental cues.
So why would a flexible mechanism of sex determination that
allows sex ratios to vary be adaptive in ferns but not in
bryophytes? The answer to this question may lie in the ephemeral
nature of the fern gametophyte. Although the gametophytes of
bryophytes are persistent, the gametophytes of ferns are not. In
Figure 3. The Genetic Sex-Determining Pathways in the Fern Ceratop- Ceratopteris, for example, gametophytes reach sexual maturity
teris, the Fly Drosophila melanogaster, and the Nematode Caenorhabdi- only 14 days after spore inoculation and die once they are
tis elegans. fertilized. The limited time that a fern gametophyte is able to be
The genetic model of sex determination in Ceratopteris (Strain et al., crossed by another might favor a sex-determining system that
2001) is dependent on two genes, FEM1 and TRA (there are at least two would promote outcrossing by increasing the proportion of
TRA genes), which promote the differentiation of male (antheridia) and males when population densities are high and ensuring a high
female (meristem and archegonia) traits, respectively. FEM1 and TRA proportion of hermaphrodites capable of self-fertilization when
also antagonize each other such that if FEM1 is active, TRA is not, and population densities are low. Because there are a variety of sex-
vice versa. What determines which of these two genes prevails in the
determining mutants available in Ceratopteris, the hypothesis
gametophyte and thus its sex is the pheromone ACE, which activates the
related to the consequences of variable versus fixed sex ratios
HER genes, of which there are at least five, and sets into motion a series
of switches that ultimately result in male development (i.e., FEM1 on and
can be tested easily, at least under defined laboratory conditions.
TRA off). These switches are thrown in the opposite direction when Future studies to clone the sex-determining genes in Cera-
spores germinate in the absence of ACE. Although FEM1 represses TRA topteris will be necessary to understand their biochemical
and TRA represses FEM1, they do not do so directly. TRA activates functions and to test their interactions predicted by the genetic
MAN1, which represses FEM1, and FEM1 activates NOT1, which model. Although the size of the Ceratopteris genome is probably
represses TRA. Because TRA and FEM1 are the primary regulators of very large (n ¼ 37), the ability to inactivate genes in the
sex, NOT1 and MAN1 are considered regulators of the regulators. The Ceratopteris gametophyte by RNA interference (Stout et al.,
sex determination pathway in C. elegans (Hodgkin, 1987; Villeneuve and 2003; G. Rutherford, M. Tanurdzic, and J.A. Banks, unpublished
Meyer, 1990) is linear and consists of a series of negative control
observations) provides an important means to study the effects
switches. The state of the initial switch gene (xol-1) is dependent on the
of inactivating potential sex-determining genes in the Ceratop-
ratio of X to autosomal (A) chromosomes. If the ultimate downstream
gene in this pathway (TRA-1) is high, the nematode develops as
teris gametophyte.
a hermaphrodite. If TRA-1 is low, it develops as a male. The linear sex
determination pathway shown for Drosophila is from the mid-1980s
(Baker and Ridge, 1980; Cline, 1983). There are actually other sex- THE FLOWERING PLANTS
determining pathways that account for most aspects of sexual
phenotype (reviewed by Oliver, 2002); the pathway shown is the somatic Although unisexuality is very common in animals, hermaphro-
pathway. In the soma, Sxl is the key regulator of sex, and its state of ditism is the rule in angiosperms. Approximately 90% of all
activity is determined by the X:A ratio. The dsx gene is the downstream angiosperm species have perfect flowers with specialized
regulatory gene that ultimately determines whether male or female genes
organs producing microspores or megaspores from which the
are expressed in the soma.
male or female gametophytes develop. Of the remaining species,
approximately half are monoecious, producing unisexual flowers
Marchantia, it is flexible enough in Ceratopteris to allow each of both sexes on the same individual, and the other half are
individual to determine its sex according to the size of the dioecious, with unisexual male and female flowers arising on
population in which it resides and the speed at which it germi- separate individuals (Yampolsky and Yampolsky, 1922). The
nates relative to its neighbors. The flexibility of the Ceratopteris distribution of dioecy and monoecy within the angiosperm
sex-determining mechanism is reflected in its sex-determining phylogenetic tree strongly favors the evolutionary scenario in
pathway, and this becomes especially apparent compared with which unisexual flowers evolved from perfect flowers multiple
the sex-determining pathways known from other organisms, times in the angiosperm lineage (Lebel-Hardenack and Grant,
including Drosophila melanogaster and Caenorhabditis elegans, 1997; Charlesworth, 2002). Not surprisingly, there are a variety of
which are illustrated in Figure 3. In both of these animals, an sex determination mechanisms in the angiosperms. For organi-
individual’s sex (male or female in D. melanogaster and male or zational purposes only, sex determination in monoecious and
hermaphrodite in C. elegans) is determined genetically by the dioecious species are treated separately in this review.
S66 The Plant Cell

The Monoecious Angiosperms

There are numerous terms to describe the variety of sexual


phenotypes observed in monoecious plant species (defined and
compiled by Sakai and Weller, 1999). Although this rich no-
menclature is appropriate, it tends to confound the problem
of sex determination in this group of plants. For simplicity,
monoecious species here are grouped into two categories: those
that produce only unisexual male and female flowers on the same
plant, and those that produce both unisexual and perfect flowers
on the same plant.
Zea mays (maize) is an example of a monoecious species that
produces only unisexual flowers in separate male and female
inflorescences, referred to as the tassel and ear, respectively.
Unisexuality in maize occurs through the selective elimination of
stamens in ear florets (flowers) and by the elimination of pistils in
tassel florets (reviewed by Irish, 1999). Two general classes of Figure 4. The ts4 Mutant of Maize.
sex-determining mutants have been identified in maize, including
The wild-type male tassel (left) produces only male florets, whereas
those that masculinize ears and those that feminize tassels. The
plants homozygous for the ts4 allele (right) form hermaphroditic flowers
anther ear (an1) and dwarf (d1, d2, d3, and d5) mutants of maize with functional pistils, allowing seeds to form in the tassel. Photographs
are recessive and masculinize ears by preventing stamen courtesy of Erin Irish, University of Iowa.
abortion in the female florets (Wu and Cheung, 2000). The
dominant dwarf mutation, d8, has a similar phenotype. The D1,
D3, and AN1 genes encode enzymes involved in the biosynthesis female-promoting phenotypic effects of GA. How the synthesis
of the plant hormone gibberellin (GA) (Bensen et al., 1995; or transport of these molecules is regulated, and the identity
Winkler and Helentjaris, 1995; Spray et al., 1996). The D8 gene of the downstream targets of these hormones, remain to be
encodes the maize homolog of GIBBERELLIN INSENSITIVE/ discovered.
REPRESSOR OF gal-3 (Peng et al., 1999), a family of Another monoecious plant, Cucumis sativus (cucumber), has
transcription factors that negatively regulate GA responses in served as a model system for sex determination studies since the
plants (Richards et al., 2001; reviewed by Olszewski et al., 2002). 1950s and early 1960s, driven by breeding programs for hybrid
GA signaling is thought to be derepressed in the d8 mutant, seed production. Cucumber plants are mostly monoecious but
resulting in a dominant phenotype. The molecular identity of can be dioecious or hermaphroditic, depending on the genotype.
these genes provides direct evidence that endogenous GAs Regardless of their sex, all floral buds are initially hermaphroditic,
have a feminizing role in sex determination in maize. and it is the arrest of stamen or pistil development that leads to
The tasselseed1 (ts1) and ts2 mutants of maize feminize male unisexual flowers. In monoecious cucumbers, male flowers form
florets in the tassel, as illustrated in Figure 4. In addition to the at the bottom and female flowers form at the top of each shoot.
tassel phenotype, the second floret of the ear spikelet, which There are three major genes that affect the arrangement of
normally fails to develop, develops normally in the ts mutants, unisexual flowers or their sex; they are designated the F, A, and M
leading to a double-kerneled spikelet in the ear (Dellaporta and genes (following the nomenclature of Pierce and Werner, 1990).
Calderon-Urrea, 1994; Irish, 1999). The TS2 gene has been The F gene is semidominant and affects the expression of
cloned and shown to encode a putative short-chain alcohol femaleness along the plant, causing it to extend the gradient
dehydrogenase with signature motifs of steroid dehydrogenases of femaleness toward the bottom of the plant. The A gene
(DeLong et al., 1993). TS2 mRNA is expressed in the sub- is epistatic to F, and it too is required for the expression of
epidermal layer of the gynoecium before its abortion, which femaleness. The M gene is required for maleness in that it is not
correlates well with the timing and location of cell death required for the establishment of the gender gradient along the
responsible for the pistil abortion in wild-type male florets shoot but rather for the selective abortion of pistils and stamens
(Calderon-Urrea and Dellaporta, 1999). The expression of ts2 in female and male flowers, respectively (Perl-Treves, 1999).
requires the wild-type TS1 gene. Although the pistils of ear florets With respect to the F and M genes, M-ff plants are monoecious,
also express TS1 and TS2, they are protected from a deathly fate M-F- plants are female, mmF- plants are hermaphroditic, and
by the action of another gene, SILKLESS1, that interacts directly mmff plants are andromonoecious (with male and hermaphro-
or indirectly with TS2 (Calderon-Urrea and Dellaporta, 1999). ditic flowers).
Although the cloning of the sex-determining genes in maize In addition to the sex-determining genes, plant hormones have
demonstrates that GAs and possibly other steroid-like hormones long been implicated in the sex-determining process in cucum-
play a pivotal role in stamen abortion and feminization of flowers, ber. GA and ethylene application and the use of GA and ethylene
the spatial distribution of these molecules could have an effect on inhibitors can subvert the genotypic constitution of the plant, with
the sex determination process, as exemplified by a steep GA acting mainly as a masculinizing agent and ethylene acting
gradient in GA abundance along the maize shoot (Rood et al., as a feminizing agent (Perl-Treves, 1999). By treating monoe-
1980), which correlates well with the male-suppressing and cious and andromonoecious cucumber plants with various
Sex Determination S67

combinations of GA and ethrel or GA and ethylene inhibitors, Yin sex-determining process. Even though this is a tempting
and Quinn (1995) demonstrated that ethylene is the main reg- speculation, it leaves the question of abortion timing unresolved.
ulator of sex determination, with GA functioning upstream of Several studies have addressed the role of the MADS box floral
ethylene, possibly as a negative regulator of endogenous homeotic genes in the sex determination process in many mono-
ethylene production. These findings led them to propose a model ecious and dioecious plants, including Asparagus officinalis
for how sex determination might occur (Yin and Quinn, 1995), (Park et al., 2003), Betula pendula (Elo et al., 2001), Gerbera
with ethylene serving both as a promoter of the female sex and an hybrida (Yu et al., 1999), Populus deltoides (Sheppard et al.,
inhibitor of the male sex. The basic tenets of the model are that 2000), Rumex acetosa (Ainsworth et al., 1995), Silene latifolia
the F gene should encode a molecule that would determine the (Hardenack et al., 1994), and maize (Heuer et al., 2001). In all
range and gradient of ethylene production along the shoot, cases, the expression of B- or C-function floral homeotic genes
thereby acting to promote femaleness, whereas the M gene in carpels or stamens was shown to decline in organs targeted
should encode a molecule that perceives the ethylene signal and for abortion in the unisexual flower. However, it is not clear if
inhibits stamen development above threshold ethylene levels. the changes in expression of these genes are a cause or a con-
This model is consistent with how unisexual flowers might arise sequence of organ abortion. Furthermore, evidence that sex-
very early and very late during shoot development; however, the determining mutants cosegregate with MADS box genes is
model also predicts an entire range of intermediate types rarely lacking. That plants may not use MADS box genes for sex
or never seen in cucumber. As suggested by Perl-Treves (1999), determination would not be surprising given that the unisexual
variations in the model of Yin and Quinn and the incorporation of flowers of most monoecious and dioecious plants are derived
additional factors in the sex-determining process in cucumber from bisexual flowers with all sex organ primordia present.
could account for the observed lack of intermediate types. Another plant within the monoecious group of angiosperms
Recent results from several laboratories have provided that has been studied is Carica papaya (papaya). Papaya is
molecular evidence in favor of the ethylene theory of sex a polygamous species with three sexes—females, males, and
determination in cucumber. Two 1-aminocyclopropane-1-car- hermaphrodites. In this species, sexual phenotype is controlled
boxylic acid synthase genes, CS-ACS1 and CS-ACS2, have by a single gene with three alleles: the dominant M (for male), the
been identified in cucumber, and one of them (CS-ACS1) maps dominant Mh (for hermaphrodite), and the recessive m (for
to the F locus (Trebitsh et al., 1997). The monoecious cucumber female) alleles (Storey, 1938). The progeny of self-fertilized
genome has only one copy (Cs-ACS1), whereas the gynoecious hermaphrodites (genotypically Mhm) segregate hermaphrodites
genome has both copies. The expression of both genes and females in a 2:1 ratio. The lack of male progeny indicate that
correlates with sexual phenotype, with gynoecious plants the MM genotype is lethal, perhaps because of a lethal gene
accumulating more transcript than monoecious or andromonoe- closely linked to the M locus. This and other crosses led early
cious plants (Kamachi et al., 1997; Yamasaki et al., 2001). workers to the conclusion that all males are genotypically MhM.
Although these studies are consistent with the female-promot- Thus, females are the homogametic and males the hetero-
ing effects of ethylene, they do not address the question of gametic sex. Papaya flowers of different sexes also display
how ethylene inhibits stamen abortion in gynoecious and not secondary sexual characteristics that cosegregate with the M
andromonoecious plants. Yamasaki et al. (2001) provided allele. The apparent lack of recombination between the loci
evidence suggesting that the product of the M locus mediates responsible for primary and secondary sex traits indicates that
the inhibition of stamen development by ethylene (i.e., M affects both loci are tightly linked and inherited en bloc, much like a sex
sensitivity to ethylene). This finding indicates that ethylene chromosome (Storey, 1953), although heteromorphic chromo-
concentration, which is likely to be dependent on the F locus, somes do not exist (Kumar et al., 1945).
and the differential sensitivity of males and females to ethylene, The economical importance of papaya has driven sex de-
which is likely to be dependent on the M locus, are both termination research in this species, because the pyriform fruits
important in regulating sexual phenotype in cucumber. Definitive from hermaphroditic trees are preferred by consumers more than
cloning of the M and F genes will allow these hypotheses to be the spherical fruits produced by the female trees. Because it is
tested directly. only upon flowering that the sex of the individual papaya tree can
Because sex determination in cucumber and most other be determined, molecular markers that cosegregate with the M
angiosperm species occurs via selective abortion of flower alleles have been sought intensively. To date, two groups have
organs, Kater et al. (2001) set out to establish whether this reported randomly amplified polymorphic DNA markers that are
abortion is based on organ identity or positional information highly specific for males and hermaphrodites but absent in
within the flower. The availability of cucumber homologs of the females (Deputy et al., 2002; Urasaki et al., 2002). One randomly
MADS box ABC homeotic genes and the ability to express them amplified polymorphic DNA marker was also shown by DNA gel
ectopically in cucumber allowed these authors to show that the blot hybridization to be completely absent from the female ge-
sex determination machinery in cucumber selectively aborts sex nome (Urasaki et al., 2002), providing the first molecular evidence
organs based on their position rather than their identity (i.e., in for genomic differentiation between the sexes. Although papaya
male flowers, carpels are aborted only in the fourth whorl, and in is a tropical tree and not considered a model plant system, it has
female flowers, stamens abort only in the third whorl). In addition, a small genome of 371 Mbp (Arumuganathan and Earle, 1991)
because nonreproductive organs that develop in the inner whorls and may be transformable (Fitch et al., 1992). These facts,
of a C-class homeotic mutant are not aborted, Kater et al. (2001) coupled with its economic importance in tropical and subtropical
speculated that C-class gene products might be targets of the regions of the world, make it is a species worthy of further study.
S68 The Plant Cell

The Dioecious Angiosperms genes have X or autosomal homologs. These data indicate
that the S. latifolia Y chromosome most closely resembles the
Silene latifolia is a dioecious species with individual plants X-degenerate class of sequences of the human Y chromo-
producing either all female or all male flowers. As it is by far the some. Genes within this class have an X homolog, the X and
best characterized dioecious species to date, our review of sex- Y homologs encode similar but nonidentical isoforms, many of
determining mechanisms in dioecious plants will focus on this them are ubiquitously expressed, and all are present in low copy
species. In male and female S. latifolia flowers, the gynoecium number in the genome (Skaletsky et al., 2003). As shown in Table
and androecium initiate but arrest development prematurely, 1, many of these features are shared with the Y-linked genes of
leading to functionally unisexual flowers (Grant et al., 1994). The S. latifolia.
sexual phenotype of individuals is determined by sex chromo- Although molecular approaches have not yet succeeded in
somes; males are heterogametic (XY) and females are homoga- identifying the major regulatory sex-determining genes in S.
metic (XX). Early cytogenetic studies of sex-determining mutants latifolia, this work has and will continue to test theories of how Y
in S. latifolia led Westergaard (1946, 1958) to conclude that the chromosomes evolved from an ancestral pair of autosomes in
Y chromosome is divided into three regions relevant to sex plants (Charlesworth, 1996, 2002; Negrutiu et al., 2001). These
expression: one required for the suppression of female de- theories state that once mutations that result in genetically
velopment and two required for the promotion of male de- determined males (where female genes are repressed) and
velopment. None of these regions would be necessary for the females (where male genes are repressed) occur, recombination
development of female reproductive organs, because these between the sex-determining genes must be suppressed to
functions would reside on the X or autosomal chromosomes. avoid recombinant asexual or hermaphroditic offspring. Another
Additional sex-determining mutants have been generated re- consequence of nonrecombination is the certainty that unisexual
cently by x-ray mutagenesis of pollen and selecting both male and female offspring will be produced in equal ratios.
hermaphrodites and asexual F1 progeny (Farbos et al., 1999; Recombination between homologous chromosomes often is
Lardon et al., 1999; Lebel-Hardenack et al., 2002). These suppressed by the accumulation of chromosomal inversions on
mutants verify the earlier work of Westergaard (1946, 1958; his one homolog (in this case, the Y). The lack of X-Y recombination
lines were apparently lost) and have resulted in the identification would eventually lead to degeneracy and loss of gene function on
of two additional classes of mutants, those that are not Y linked the Y chromosome, with the exception of genes required for male
and hermaphroditic and those that are Y linked and asexual. The fertility and those necessary to suppress female fertility. The S.
hermaphroditic deletion mutants are likely to contain a gene(s) latifolia Y chromosome appears not to fit this paradigm for
necessary for the female-suppressing function, whereas the several reasons. First, the Y chromosome is 1.4 times larger than
asexual deletion mutants likely contain the male-promoting the X chromosome and is largely euchromatic, indicating that it
gene(s). Genetic screens to identify mutant XX hermaphrodites may not be degenerate (Ciupercescu et al., 1990; Grabowska-
or asexuals have not been reported. Although these sex- Joachimiak and Joachimiak, 2002). Second, measurements of
determining genes have not been cloned, the construction of DNA polymorphism in sex-linked gene pairs have revealed that
an amplified fragment length polymorphism map of the Y although the DNA polymorphism of S4Y-1 is greater than that of
chromosome using lines deleted for overlapping regions of the S4X-1, the DNA polymorphism of SlY-1 is 20-fold lower than that
Y chromosome will be useful for genetic and physical mapping of of SlX-1 (Filatov et al., 2000, 2001; Filatov and Charlesworth
the sex-determining mutants (Lebel-Hardenack et al., 2002) and 2002). Given that the S. latifolia sex chromosomes diverged less
may ultimately lead to their cloning. than 20 million years ago (Desfeux et al., 1996; Charlesworth,
Another approach to identify sex-determining genes in S. 2002), which is much less than the 240- to 320-million-year
latifolia has been to clone genes that are expressed specifically in timeline for human sex chromosome evolution (Lahn and Page,
the male flowers and determine their linkage to the Y chromo- 1999), it is likely that the S. latifolia Y chromosome is at a relatively
some (reviewed by Charlesworth, 2002). Of the >50 isolated young stage of evolution (Charlesworth, 2002). Comparative
genes that have been correlated with sex expression in S. latifolia sequencing of the S. latifolia sex chromosomes will be important
to date, only the four listed in Table 1 have been shown to be in understanding the evolution of the Y chromosome in plants,
linked to the Y chromosome. Genes linked exclusively to the Y especially compared with the Y chromosome of Marchantia and
chromosome have not been found, because all of the Y-linked the sex-determining chromosome region of papaya.

Table 1. Sex Chromosome–Linked Genes in S. latifolia

Y-Linked Gene X-Linked Gene Male-Specific Expression Function Reference

SlY-1 SlX-1 SlY-1 yes, SlX-1 no WD-repeat protein Delichere et al., 1999
SlY-4 SlX-4 No Fructose-2,6-bisphosphatase Atanassov et al., 2001
MRO3-Ya MRO3-Xa Yes Unknown Matsunaga et al., 1996; Guttman and
Charlesworth 1998
DD44Y DD44X No Oligomycin sensitivity–conferring Moore et al., 2003
protein
a The Y homolog is inactive and the X homolog is active.
Sex Determination S69

FUTURE DIRECTIONS Arumuganathan, K., and Earle, E.D. (1991). Nuclear DNA content of
some important plant species. Plant Mol. Biol. Rep. 9, 208–218.
Sex determination in plants is a fundamental developmental Atanassov, I., Delichere, C., Filatov, D., Charlesworth, D., Negrutiu,
process that is particularly important for economic reasons, I., and Moneger, F. (2001). Analysis and evolution of two functional
because the sexual phenotypes of commercially important crops Y-linked loci in a plant sex chromosome system. Mol. Biol. Evol.
dictate how they are bred and cultivated. Although most crop 18, 2126–2168.
plants are not considered model systems—and sex determina- Baker, B.S., and Ridge, K.A. (1980). Sex and the single cell. I. On the
tion is not a problem that can be addressed in the model action of major loci affecting sex determination in Drosophila
angiosperm Arabidopsis—the economic value in manipulat- melanogaster. Genetics 94, 383–423.
Banks, J. (1998). Sex determination in the fern Ceratopteris. Trends
ing the sexual phenotypes of crop plants should continue to
Plant Sci. 2, 175–180.
drive interest in this area of research. Recent studies of sex-
Banks, J., Hickok, L., and Webb, M.A. (1993). The programming of
determining mechanisms have demonstrated clearly that an- sexual phenotype in the homosporous fern, Ceratopteris richardii. Int.
giosperms, including crop plants, have evolved a variety of J. Plant Sci. 154, 522–534.
sex-determining mechanisms that involve a number of different Barrett, S. (2002). The evolution of plant sexual diversity. Nat. Rev. 3,
genetic and epigenetic factors, from sex chromosomes to plant 274–284.
hormones. Although the determinants of sexual phenotype are Bensen, R., Johal, J., Crane, V.C., Tossberg, J.T., Schnable, P.S.,
diverse, determining whether the downstream master sex- Meeley, R.B., and Briggs, S.P. (1995). Cloning and characterization
regulatory genes that specify male or female development are of the maize An1 gene. Plant Cell 7, 75–84.
held in common or not will require cloning the sex-determining Calderon-Urrea, A., and Dellaporta, S.L. (1999). Cell death and cell
genes from a variety of plant species. protection genes determine the fate of pistils in maize. Development
126, 435–441.
Choosing to study sex determination in plants representing
Charlesworth, D. (1996). The evolution of chromosomal sex deter-
other major land plant lineages will allow several broader
mination and dosage compensation. Curr. Biol. 6, 149–162.
developmental and evolutionary questions to be addressed. Charlesworth, D. (2002). Plant sex determination and sex chromo-
One unresolved question is how heterospory evolved from somes. Heredity 88, 94–101.
homospory. By identifying the sex-determining genes in homo- Ciupercescu, D., Veuskens, J., Mouras, A., Ye, D., Briquet, M., and
sporous plants such as Ceratopteris and examining the Negrutiu, I. (1990). Karyotyping Melandrium album, a dioecious plant
expression of possible homologous genes in closely related with heteromorphic sex chromosomes. Genome 33, 556–562.
heterosporous species, one can test the hypothesis that the Cline, T.W. (1983). The interaction between daughterless and sex-lethal
switch from homospory to heterospory involved a heterochronic in triploids: A lethal sex-transforming maternal effect linking sex
shift in the timing of expression of these genes from the determination and dosage compensation in Drosophila melanogaster.
gametophyte to the sporophyte generation. Other questions to Dev. Biol. 95, 260–274.
Delichere, C., Veuskens, J., Hernould, M., Barbacar, N., Mouras, A.,
be resolved are how sex chromosomes evolved in plants and
Negrutiu, I., and Moneger, F. (1999). SlY1, the first active gene
whether similar processes led to distinct sex chromosomes in
cloned from a plant Y chromosome, encodes a WD-repeat protein.
plants and animals. Comparing the Y chromosome sequences of
EMBO J. 11, 4169–4179.
Marchantia and S. latifolia, for example, will be invaluable in Dellaporta, S.L., and Calderon-Urrea, A. (1994). The sex determina-
understanding how male-promoting genes and female-sup- tion process in maize. Science 266, 1501–1505.
pressing genes became localized to a Y chromosome and how DeLong, A., Calderon-Urrea, A., and Dellaporta, S.L. (1993). Sex
recombination between the Y and its homolog was and con- determination gene TASSELSEED2 of maize encodes a short-chain
tinues to be suppressed. alcohol dehydrogenase required for stage-specific floral organ
abortion. Cell 27, 757–768.
Deputy, J.C., Ming, R., Ma, H., Liu, Z., Fitch, M.M., Wang, M.,
ACKNOWLEDGMENTS
Manshardt, R., and Stiles, J.I. (2002). Molecular markers for sex
Support was provided by the National Science Foundation (MCB- determination in papaya (Carica papaya L.). Theor. Appl. Genet. 106,
9723154). This is journal paper 17271 of the Purdue University 107–111.
Agricultural Experiment Station. Desfeux, C., Maruice, S., Henry, J.P., Lejeune, B., and Gouyon, P.H.
(1996). Evolution of reproductive systems in the genus Silene. Proc.
R. Soc. Lond. Ser. B 263, 409–414.
Received August 25, 2003; accepted January 19, 2004. Dopp, W. (1950). Eine die Antheridienbildung bei farnen fordernde
Substanz in den Prothallien von Pteridium aquilinum L. Kuhn. Ber.
Dtsch. Bot. Ges. 63, 139–147.
REFERENCES Elo, A., Lemmetyinen, J., Turunen, M., Tikka, L., and Sopanen, T.
(2001). Three MADS-box genes similar to APETALA1 and FRUITFULL
Ainsworth, C., ed (1999). Sex Determination in Plants. (Oxford, UK: from silver birch (Betula pendula). Physiol. Plant. 112, 95–103.
BIOS Scientific Publishers). Farbos, I., Veuskens, J., Vyskot, B., Oliveira, M., Hinnisdaels, S.,
Ainsworth, C. (2000). Boys and girls come out to play: The molecular Aghmir, A., Mouras, A., and Negrutiu, I. (1999). Sexual dimorphism
biology of dioecious plants. Ann. Bot. 86, 211–221. in white campion: Deletion of the Y chromosome results in a floral
Ainsworth, C., Crossley, S., Buchanan-Wollaston, V., Thangavelu, asexual phenotype. Genetics 151, 1187–1196.
M., and Parker, J. (1995). Male and female flowers of the dioecious Filatov, D., and Charlesworth, D. (2002). Substitution rates in the
plant sorrel show different patterns of MADS box gene expression. X- and Y-linked genes of the plants Silene latifolia and S. dioica. Mol.
Plant Cell 7, 1583–1598. Biol. Evol. 19, 898–907.
S70 The Plant Cell

Filatov, D., Laporte, V., Vitte, C., and Charlesworth, D. (2001). DNA Lebel-Hardenack, S., and Grant, S.R. (1997). Genetics of sex
diversity in sex-linked and autosomal genes of the plant species determination in flowering plants. Trends Plant Sci. 2, 130–136.
Silene latifolia and Silene dioica. Mol. Biol. Evol. 18, 1442–1454. Lebel-Hardenack, S., Hauser, E., Law, T., Schmidt, J., and Grant, S.
Filatov, D., Moneger, F., Negrutiu, I., and Charlesworth, D. (2000). (2002). Mapping of sex determination loci on the white campion
Low variability in a Y-linked plant gene and its implications for (Silene latifolia) Y chromosome using amplified fragment length
Y-chromosome evolution. Nature 404, 388–390. polymorphism. Genetics 160, 717–725.
Fitch, M.M., Manshardt, R., Gonsalves, D., and Slightom, J.L. (1992). Lorbeer, G. (1934). Die Zytologie der Lebermoose mit besonderer
Virus resistant papaya derived from tissue bombarded with the coat Berucksichtingung allgemeiner Chromosomenfragen. Jahrb. Wiss.
protein gene of papaya ringspot virus. Bio/Technology 10, 1466– Bot. 80, 567–817.
1472. Matsunaga, S., and Kawano, S. (2001). Sex determination by sex
Geber, M.A., Dawson, T.E., and Delph, L.F., eds (1999). Gender and chromosomes in dioecious plants. Plant Biol. 3, 481–488.
Sexual Dimorphism in Flowering Plants. (Berlin: Springer). Matsunaga, S., Kawano, S., Takano, H., Uchida, H., Sakai, A.,
Grabowska-Joachimiak, A., and Joachimiak, A. (2002). C-banded and Kuroiwa, T. (1996). Isolation and developmental expression of
karyotypes of two Silene species with heteromorphic sex chromo- male reproductive organ-specific genes in a dioecious campion,
somes. Genome 45, 243–252. Melandrium album (Silene latifolia). Plant J. 10, 679–698.
Grant, S., Hunkirchen, B., and Saedler, H. (1994). Developmental Moore, R.C., Kozyreva, O., Lebel-Hardenack, S., Siroky, J., Hobza,
differences between male and female flowers in the dioecious plant R., Vyskot, B., and Grant, S.R. (2003). Genetic and functional
Silene latifolia. Plant J. 6, 471–480. analysis of DD44, a sex-linked gene from the dioecious plant Silene
Graves, J., and Schmidt, M. (1992). Mammalian sex chromosomes: latifolia, provides clues to early events in sex chromosome evolution.
Design or accident? Curr. Opin. Genet. Dev. 2, 890–901. Genetics 163, 321–334.
Guttman, D., and Charlesworth, D. (1998). An X-linked gene with Naf, U. (1979). Antheridiogens and antheridial development. In The
a degenerate Y-linked homologue in a dioecious plant. Nature 393,
Experimental Biology of Ferns, A.F. Dyer, ed (New York: Academic
263–266.
Press), pp. 436–470.
Hardenack, S., Ye, D., Saedler, H., and Grant, S. (1994). Comparison
Negrutiu, I., Vyskot, B., Barbacar, N., Georgiev, S., and Moneger, F.
of MADS box gene expression in developing male and female flowers
(2001). Dioecious plants: A key to the early events of sex chromosome
of the dioecious plant white campion. Plant Cell 6, 1775–1787.
evolution. Plant Physiol. 127, 1418–1424.
Hawley, R.S. (2003). The human Y chromosome: Rumors of its death
Okada, S., et al. (2000). Construction of male and female PAC genomic
have been greatly exaggerated. Cell 113, 825–828.
libraries suitable for identification of Y-chromosome-specific clones
Heuer, S., Hansen, S., Bantin, J., Brettschneider, R., Kranz, E., Lorz,
from the liverwort, Marchantia polymorpha. Plant J. 24, 421–428.
H., and Dresselhaus, T. (2001). The maize MADS box gene
Okada, S., et al. (2001). The Y chromosome in the liverwort Marchantia
ZmMADS3 affects node number and spikelet development and is
polymorpha has accumulated unique repeat sequences harboring
co-expressed with ZmMADS1 during flower development, in egg
a male-specific gene. Proc. Natl. Acad. Sci. USA 98, 9454–9459.
cells, and early embryogenesis. Plant Physiol. 127, 33–45.
Oliver, B. (2002). Genetic control of germline sexual dimorphism in
Hodgkin, J. (1987). Sex determination and dosage compensation in
Drosophila. Int. Rev. Cytol. 219, 1–60.
Caenorhabditis elegans: Variations on a theme. Annu. Rev. Genet. 21,
Olszewski, N., Sun, T.-p., and Gubler, F. (2002). Gibberellin signaling:
133–154.
Biosynthesis, catabolism, and response pathways. Plant Cell 14
Irish, E.E. (1999). Maize sex determination. In Sex Determination in
(suppl.), S61–S80.
Plants, C. Ainsworth, ed (Oxford, UK: BIOS Scientific Publishers), pp.
Park, H.H., Ishikawa, Y., Yoshida, R., Kanno, A., and Kameya, T.
183–188.
(2003). Expression of AODEF, a B-functional MADS-box gene, in
Ishizaki, K., Shimizu-Ueda, Y., Okada, S., Yamamoto, M., Fujisawa,
M., Yamato, K.T., Fukuzawa, H., and Ohyama, K. (2002). Multicopy stamens and inner sepals of the dioecious species Asparagus
genes uniquely amplified in the Y chromosome-specific repeats of the officinalis L. Plant Mol. Biol. 51, 867–875.
liverwort Marchantia polymorpha. Nucleic Acids Res. 30, 4675–4681. Peng, J., et al. (1999). ‘‘Green revolution’’ genes encode mutant
Jegalian, K., and Page, D. (1998). A proposed path by which genes gibberellin response modulators. Nature 400, 256–261.
common to mammalian X and Y chromosomes evolve to become X Perl-Treves, R. (1999). Male to female conversion along the cucumber
inactivated. Nature 394, 776–780. shoot: Approaches to studying sex genes and floral development in
Kamachi, S., Sekimoto, H., Kondo, H., and Sakai, S. (1997). Cloning Cucumis sativus. In Sex Determination in Plants, C. Ainsworth, ed
of a cDNA for a 1-aminocyclopropane-1-carboxylate synthase that is (Oxford, UK: Bios Scientific Publishers), pp. 189–216.
expressed during development of female flowers at the apices of Pierce, L.K., and Werner, T.C. (1990). Review of genes and linkage
Cucumis sativus L. Plant Cell Physiol. 38, 1197–1206. groups in cucumber. HortScience 25, 605–615.
Kater, M.M., Franken, J., Carney, K., Colombo, L., and Angenent, Pryer, K.M., Schneider, H., Smith, A.R., Cranfill, R., Wolf, P.G.,
G.C. (2001). Sex determination in the monoecious species cucumber Hunt, J.S., and Sipes, S.D. (2001). Horsetails and ferns are a mono-
is confined to specific floral whorls. Plant Cell 13, 481–493. phyletic group and the closest living relatives to seed plants. Nature
Kenrick, P., and Crane, P.R. (1997). The origin and early evolution of 409, 618–622.
plants on land. Nature 389, 33–39. Richards, D.E., King, K.E., Ait-ali, T., and Harberd, N.P. (2001). How
Kumar, L., Abraham, A., and Srinivasan, V. (1945). The cytology of gibberellin regulates plant growth and development: A molecular
Carica papaya Linn. Indian J. Agric. Sci. 15, 242–253. genetic analysis of gibberellin signaling. Annu. Rev. Plant Physiol.
Lahn, B., and Page, D. (1999). Four evolutionary strata on the human X Plant Mol. Biol. 52, 67–88.
chromosome. Science 286, 964–967. Rood, S.B., Pharis, R.P., and Major, D.J. (1980). Changes of
Lardon, A., Seorgiev, S., Aghmir, A., Le Merrer, G., and Negruitiu, I. endogenous gibberellin-like substances with sex reversal of the
(1999). Sexual dimorphism in white campion: Complex control of apical inflorescence of corn. Plant Physiol. 66, 793–796.
carpel number is revealed by Y chromosome deletions. Genetics 151, Rozen, S., Skaletsky, H., Marszalek, J.D., Minx, P.J., Cordum, H.S.,
1173–1185. Waterston R.H., Wilson R.K., and Page D.C. (2003). Abundant gene
Sex Determination S71

conversion between arms of palindromes in human and ape Y K.T., Fukuzawa, H., and Ohyama, K. (2000). Direct transformation
chromosomes. Nature 423, 873–876. and plant regeneration of the haploid liverwort Marchantia polymorpha
Sakai, A.K., and Weller, S.G. (1999). Gender and sexual dimorphism in L. Transgenic Res. 9, 179–185.
flowering plants: A review of terminology, biogeographic patterns, Trebitsh, T., Rudich, J., and Riov, J. (1997). Identification of a
ecological correlates and phylogenetic approaches. In Gender and 1-aminocyclopropane-1-carboxylic acid synthase gene linked to
Sexual Dimorphism in Flowering Plants, M.S. Geber, T.E. Dawson, the Female (F) locus that enhances female sex expression in cucum-
and L.F. Delph, eds (Berlin: Springer), pp. 1–31. ber. Plant Physiol. 113, 987–995.
Sheppard, L.A., Brunner, A., Krutovskii, K., Rottmann, W., Skinner, Urasaki, N., Tokumoto, M., Tarora, K., Ban, Y., Kayano, T., Tanaka,
J., Vollmer, S., and Strauss, S.H. (2000). A DEFICIENS homolog from H., Oku, H., Chinen, I., and Terauchi, R. (2002). A male and
the dioecious tree black cottonwood is expressed in female and male hermaphrodite specific RAPD marker for papaya (Carica papaya L.).
floral meristems of the two-whorled, unisexual flowers. Plant Physiol. Theor. Appl. Genet. 104, 281–285.
124, 627–640. Villeneuve, A.M., and Meyer, B.J. (1990). The regulatory hierarchy
Skaletsky, H., et al. (2003). The male-specific region of the human Y
controlling sex determination and dosage compensation in Caeno-
chromosome is a mosaic of discrete sequence clusters. Nature 423,
rhabditis elegans. Adv. Genet. 27, 117–188.
825–837.
Westergaard, M. (1946). Aberrant Y chromosomes and sex expression
Smith, G.M. (1955). Cryptogamic Botany. Vol. II. Bryophytes and
in Melandrium album. Hereditas 32, 419–443.
Pteridophytes. (New York: McGraw-Hill).
Westergaard, M. (1958). The mechanism of sex determination in
Spray, C.R., Kobayashi, M., Suzuki, Y., Phinney, B.O., Gaskin, P.,
dioecious flowering plants. Adv. Genet. 9, 217–281.
and MacMillan, J. (1996). The dwarf-1 (d1) mutant of Zea mays
Winkler, R.G., and Helentjaris, T. (1995). The maize Dwarf3 gene
blocks steps in the gibberellin-biosynthetic pathway. Proc. Natl. Acad.
encodes a cytochrome P450-mediated early step in gibberellin
Sci. USA 93, 10515–10518.
biosynthesis. Plant Cell 7, 1307–1317.
Stewart, W., and Rothwell, G. (1993). Paleobotany and the Evolution of
Wu, H.-M., and Cheung, A.Y. (2000). Programmed cell death in plant
Plants. (New York: Cambridge University Press).
Storey, W.B. (1938). Segregation of sex types in solo papaya reproduction. Plant Mol. Biol. 44, 267–281.
and their application to the selection of seed. Am. Soc. Hortic. Sci. 35, Yamane, H. (1998). Fern antheridiogens. Int. Rev. Cytol. 184, 1–32.
83–85. Yamasaki, S., Fujii, N., Matsuura, S., Mizusawa, H., and Takahashi,
Storey, W.B. (1953). Genetics of papaya. J. Hered. 44, 70–78. H. (2001). The M locus and ethylene-controlled sex determination in
Stout, S.C., Clark, G.B., Archer-Evans, S., and Roux, S.J. (2003). andromonoecious cucumber plants. Plant Cell Physiol. 42, 608–619.
Rapid and efficient suppression of gene expression in a single-cell Yampolsky, C., and Yampolsky, H. (1922). Distribution of the sex
model system, Ceratopteris richardii. Plant Physiol. 131, 1165–1168. forms in the phanerogamic flora. Bibl. Genet. 3, 1–62.
Strain, E., Hass, B., and Banks, J. (2001). Characterization of Yin, T., and Quinn, J.A. (1995). Tests of a mechanistic model of one
mutations that feminize gametophytes of the fern Ceratopteris. hormone regulating both sexes in Cucumis sativus (Cucurbitaceae).
Genetics 159, 1271–1281. Am. J. Bot. 82, 1537–1546.
Sussex, I. (1966). The origin and development of heterospory in Yu, D., Kotilainen, M., Pollanen, E., Mehto, M., Elomaa, P.,
vascular plants. In Trends in Plant Morphogenesis, E. Cutter, ed Helariutta, Y., Albert, V., and Teeri, T. (1999). Organ identity genes
(New York: John Wiley & Sons), pp. 140–152. and modified patterns of flower development in Gerbera hybrida
Takenaka, M., Yamaoka, S., Hanajiri, T., Shimizu-Ueda, Y., Yamato, (Asteraceae). Plant J. 17, 51–62.

You might also like