ArticlePDF Available

Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus (Orchidaceae) in a fragmented cloud forest

Authors:

Abstract and Figures

Genetic differentiation in space can be detected at various scales. First, habitat fragmentation can produce a mosaic genetic structure. Second, life history aspects of a species such as dispersion, mating system, and pollination can generate a genetic structure at a finer level. The interplay of these levels has rarely been studied together. In order to assess the effects of forest fragmentation we analyzed the genetic structure at two spatial scales of the terrestrial orchid Cyclopogon luteoalbus, which lives in patches inside forest fragments in a cloud forest of eastern Mexico. We hypothesized high differentiation between forest fragments and strong spatial genetic structure within fragments under this scenario of strong fragmentation and restricted dispersal patterns. Using 11 allozymic loci we found high genetic diversity at fragment level with moderate differentiation among fragments, and at patch level, strong and variable spatial genetic structure among life cycle stages with high inbreeding coefficients. We also found bottlenecks indicating recent population size reductions. While both inbreeding and restricted seed dispersal may explain the strong spatial genetic structure at patch level, reduction in population size may explain the genetic structure at fragment level. However, the levels of genetic diversity indicate that some between-fragment gene flow has occurred. Bottlenecks and high inbreeding at patch level may result in local extinctions, but as long as an important number of fragments remain, patch recolonization through immigration is possible in C.luteoalbus. KeywordsAllozymes–Autocorrelation–Heterozygosity–Inbreeding–Outcrossing rate–Spatial genetic structure
Content may be subject to copyright.
1 23
Plant Systematics and Evolution
ISSN 0378-2697
Volume 297
Combined 3-4
Plant Syst Evol (2011) 297:237-251
DOI 10.1007/s00606-011-0511-6
Genetic structure at patch level of the
terrestrial orchid Cyclopogon luteoalbus
(Orchidaceae) in a fragmented cloud forest
Lilian Juárez, Carlos Montaña & Miriam
M.Ferrer
1 23
Your article is protected by copyright and
all rights are held exclusively by Springer-
Verlag. This e-offprint is for personal use only
and shall not be self-archived in electronic
repositories. If you wish to self-archive your
work, please use the accepted author’s
version for posting to your own website or
your institution’s repository. You may further
deposit the accepted author’s version on a
funder’s repository at a funder’s request,
provided it is not made publicly available until
12 months after publication.
ORIGINAL ARTICLE
Genetic structure at patch level of the terrestrial orchid
Cyclopogon luteoalbus (Orchidaceae) in a fragmented cloud forest
Lilian Jua
´rez Carlos Montan
˜aMiriam M. Ferrer
Received: 7 March 2011 / Accepted: 13 July 2011 / Published online: 9 August 2011
ÓSpringer-Verlag 2011
Abstract Genetic differentiation in space can be detected
at various scales. First, habitat fragmentation can produce a
mosaic genetic structure. Second, life history aspects of a
species such as dispersion, mating system, and pollination
can generate a genetic structure at a finer level. The
interplay of these levels has rarely been studied together. In
order to assess the effects of forest fragmentation we
analyzed the genetic structure at two spatial scales of the
terrestrial orchid Cyclopogon luteoalbus, which lives in
patches inside forest fragments in a cloud forest of eastern
Mexico. We hypothesized high differentiation between
forest fragments and strong spatial genetic structure within
fragments under this scenario of strong fragmentation and
restricted dispersal patterns. Using 11 allozymic loci we
found high genetic diversity at fragment level with mod-
erate differentiation among fragments, and at patch level,
strong and variable spatial genetic structure among life
cycle stages with high inbreeding coefficients. We also
found bottlenecks indicating recent population size reduc-
tions. While both inbreeding and restricted seed dispersal
may explain the strong spatial genetic structure at patch
level, reduction in population size may explain the genetic
structure at fragment level. However, the levels of genetic
diversity indicate that some between-fragment gene flow
has occurred. Bottlenecks and high inbreeding at patch
level may result in local extinctions, but as long as an
important number of fragments remain, patch recoloniza-
tion through immigration is possible in C. luteoalbus.
Keywords Allozymes Autocorrelation
Heterozygosity Inbreeding Outcrossing rate
Spatial genetic structure
Introduction
Genetic structure is defined by differences in allele fre-
quency within and among populations (Loveless and
Hamrick 1984). It can be detected at landscape, population,
or patch level (Epperson 1993; Fenster et al. 2003; Manel
et al. 2003) and is influenced by ecological factors (Kalisz
et al. 2001) such as pollinator availability (Van Rossum
and Triest 2006), pollination mechanism (Hamrick and
Godt 1996), habitat availability (Sun and Wong 2001), and
breeding system and dispersal mechanisms (Hamrick et al.
1993; Kalisz et al. 2001; Wright 1943). At landscape and
population scales, isolation by distance (due to genetic drift
associated with limited gene flow) may also influence
genetic structure (Wright 1943). Another factor that is
becoming important for genetic structure is habitat frag-
mentation, as isolation contributes to genetic differentia-
tion between fragmented populations and diminishes
genetic variation (Aguilar et al. 2008).
At fine scale or patch level, spatial genetic structure
(SGS) is promoted when relatedness decreases with dis-
tance and genotypes are not randomly distributed within a
population (Epperson 1993). SGS can be the result of
limited gene flow, genetic drift, microenvironmental
selection, breeding system, dispersal mechanisms, and
inherently ecological proximal causes acting at that level
(Hamrick et al. 1993; Kalisz et al. 2001; Wright 1943).
L. Jua
´rez C. Montan
˜a(&)
Red de Biologı
´a Evolutiva, Instituto de Ecologı
´a A.C.,
Apartado Postal 63, 91000 Xalapa, Veracruz, Mexico
e-mail: carlos.montana@inecol.edu.mx
M. M. Ferrer
Departamento de Ecologı
´a Tropical, Universidad Auto
´noma
de Yucata
´n, Km. 15.5 Carretera Me
´rida Xmatkuil,
97000 Me
´rida, Yucata
´n, Mexico
123
Plant Syst Evol (2011) 297:237–251
DOI 10.1007/s00606-011-0511-6
Author's personal copy
SGS in terrestrial orchids has been attributed to: (1) the
presence of pollinaria, because this ensures that seeds
within a capsule are full sibs, (2) seedling establishment
aggregated around mother plants where the necessary
mycorrhiza are found (this mechanism was inferred by
Chung et al. 2004a for Cephalanthera longibracteata, and
by Jacquemyn et al. 2006 for Orchis purpurea), (3) polli-
nation among near neighbors or flowers of the same plant
promoting inbreeding (e.g., Caladenia tentaculata, Peakall
and Beattie 1996 and Cymbidium goeringii, Chung et al.
1998), (4) clonal propagation (e.g., Cremastra appendicu-
lata, Chung et al. 2004b), and (5) recent founding of the
population (e.g., Spiranthes spiralis, Machon et al. 2003
and Cymbidium goeringii, Chung et al. 1998). Also, SGS
may vary with life cycle stage or cohort age (A
´lvarez-
Buylla et al. 1996; Epperson 1993; Kalisz et al. 2001); for
instance, SGS in seedlings may be lower than in adults
when after-establishment thinning preferentially eliminates
full sibs or highly related seedlings (Kalisz et al. 2001;
Tonsor et al. 1993). In these cases, seeds produced by
adults are highly inbred because seeds produced in polli-
naria are full siblings and seedlings mostly establish near
established plants.
Distributed mostly in the mountains of eastern Mexico,
cloud forest is home to half of the orchids known in Mexico
(IUCN/SSC 1996), but this vegetation type has been sub-
jected to a long and strong deforestation process to make
grazing lands, coffee plantations, and urban settlements,
particularly since the 19th century until now. In cloud forest,
as in most tropical forests, fragmentation may disrupt
genetic processes of some species, diminishing the genetic
diversity of remnant populations as local drift and/or
inbreeding increases and gene flow decreases (Aguilar et al.
2008; Honnay and Jacquemyn 2007). Orchidaceae is one of
the largest angiosperm families (17,000–35,000 species),
with around one-third of the species displaying terrestrial
growth habit (Dressler 1993). Cyclopogon luteoalbus
(A. Rich. & Galleoti) Schltr. is a locally abundant terrestrial
orchid patchily distributed in remnant cloud forest fragments
of very variable sizes but larger than 2 ha that are immersed
in a matrix of coffee plantations, pasturelands, and urban
settlements in the mountains of eastern Mexico. As its
generational time is 25 years, as shown by a demographic
study from 2005 to 2009 (Jua
´rez and Montan
˜a, unpublished
data), it is expected that its current genetic structure has been
affected by the long and strong fragmentation process.
This genetic study of C. luteoalbus was conducted, at
both fragment and patch (within-fragment) levels, in the
cloud forest of central Veracruz. In this highly fragmented
landscape we hypothesized high genetic differentiation
between fragments as a consequence of limited gene flow,
and strong spatial genetic structure at patch level because
seedling establishment occurs nearby already established
plants. We also expected to find differences among the
SGS of the different life stages due to nonrandom thinning
of individuals in transitions from each life stage to the next
one.
Materials and methods
Species and study sites
Cyclopogon luteoalbus (Orchidoideae: Cranichideae) is a
Neotropical species distributed from Mexico to El Salva-
dor. It is composed of a rosette (fewer than eight leaves)
and a simple system of tuberoid roots. Clonal propagation
is rare (5 out of 500 revised individuals in a 5-year
demographic study had connecting roots; Jua
´rez and
Montan
˜a, unpublished data). Flowering occurs synchro-
nously in winter (February), while fruits are set in March.
Nectar-producing flowers last 10 days. Fruits contain
2,510 ±1,593 SD seeds (N=8 fruits), which are wind-
dispersed during the dry season. The species has a mixed
mating system: autonomous self-pollination is possible, as
corroborated experimentally (0.83 ±0.23 SD fruit set in
bagged inflorescences, N=14 and 0.85 ±0.13 SD fruit
set in nonbagged inflorescences, N=17), and outcrossing
is promoted by sequential opening of flower within inflo-
rescences. Two halictid bees (Caenaugochlora cupriventris
and Augochlora sp.) are potential pollinators. Five-year
(2005–2009) demographic studies conducted in two
populations (where the fate of a total of 950 individuals
was monitored) showed that the generational time of
C. luteoalbus is around 25 years (Jua
´rez and Montan
˜a,
unpublished data).
The species lives in unevenly distributed patches inside
forests fragments of eastern Mexico. In a surface area of
approximately 15 km 915 km (which included the SBN
and Martinica study sites),
´az-Toribio (2009) reported
the presence of C. luteoalbus in 14 cloud forest fragments
of more than 2 ha surface area. The mean surface area of
and the mean distance between the 14 fragments were
13.3 ±10.7 SD ha and 4.8 ±2.7 SD km. The values of
some soil parameters were: litter depth 5.7 ±1.8 SD cm,
pH 4.6 ±0.8 SD, herbaceous cover 52.6 ±22.2 SD%,
total N 1.2 ±0.6 SD%, total C 18.7 ±12.7 SD%, C/N
14.2 ±2.5 SD, available P 3.9 ±3.4 SD ppm, and K
0.8 ±0.4 SD cmol/kg (N=14 in all cases). Most proba-
bly the patchy distribution inside fragments is due to the
very special soil conditions needed for the development of
the mycorrhizae required for seed germination (Rasmussen
1995). These soil conditions under forest canopies may be
locally absent for natural reasons or may have been elim-
inated by management in forested areas used for shadow
coffee plantations, where under a canopy of native and
238 L. Jua
´rez et al.
123
Author's personal copy
introduced trees, the understory is permanently eliminated
with consequent loss of the upper soil layer. The eventual
use of fertilizers (mainly phosphorus) and/or pesticides also
affects soil conditions. Even if coffee plantation is aban-
doned, reconstruction of these upper soil layers may take
several decades (or even centuries).
According to Rzedowski (1978), cloud forest in the
mountains of eastern Mexico has been occupied and
exploited by man during centuries. Forest clearing for
maize and bean cultivation was done during pre-Colum-
bian times, and for cane sugar cultivation since colonial
times (16th century). Challenger (1998) mentions that
deforestation for introduction of coffee plantations was
very pronounced between the last quarter of the 19th
century and 1960. Using satellite images of a watershed
which included two of our three study sites, Mun
˜oz-Villers
and Lo
´pez-Blanco (2008) documented a process of forest
conversion to grazing and agriculture lands that occurred in
central Veracruz between 1990 and 2003 (this process
began by the mid 20th century and is still going on). In this
1,325 km
2
watershed (1,316 of which has the potential to
sustain several types of forest, but mainly cloud forest)
these authors report that there remained only 561 km
2
of
forests in 1990 and that this figure had decreased to
487 km
2
in 2003. This variation was due to a small
increase in oak–pine (25 km
2
) and conifer (48 km
2
) forests
due to reforestation programmes and to a decrease from
427 to 279 km
2
of cloud forest cleared to be used as pas-
ture and agriculture lands. These anthropogenic changes in
natural habits had led to fragmentation (a complex process
that involves habitat loss, an increase in number of patches,
a decrease in patch size, and patch isolation; Fahrig 2003)
of the original continuous habitat for cloud forest orchids,
affecting the connection between them.
The study was conducted in three cloud forest fragments
in the central part of the State of Veracruz, Mexico [San-
tuario del Bosque de Niebla (SBN), Martinica, and Zon-
golica sites]. SBN is a 30-ha protected area (since 1975)
located 2.5 km southwest from the City of Xalapa
(19°3100500N, 96°5600300W, 1,350 m a.s.l., 1,517 mm rain-
fall, 18°C mean annual temperature). The dominant species
at the forest canopy are: Liquidambar styraciflua,Quercus
xalapensis, and Carpinus caroliniana, and at the under-
story are Palicourea padifolia and Piper auritum with
some remnants of farming such as coffee, orange, loquat,
and guava. Martinica is located 8 km northwest from SBN
at the outskirts of the City of Banderilla (19°3404900N,
96°5602800W, 1,550 m a.s.l., 1,470 mm rainfall, 16°C mean
annual temperature), comprising ca. 10 ha of privately
owned land that has not been farmed for the last 20 years.
The vegetation is similar to SBN.
Zongolica is a ca. 50 ha nonprotected area located ca.
200 km northwest from Martinica and SBN, in the Sierra
Zongolica, in western Veracruz (18°3900000N, 96°4904900W,
1,330 m a.s.l., 2,270 mm rainfall, 17°C mean annual
temperature). Cloud forest in this zone is mixed with
tropical semi-evergreen forest and is subjected to illegal
wood extraction, clearing for farming purposes, and
browsing by goats at the understory. Canopy dominant
species include Alchornea latifolia,Cupania dentata, and
Guettarda elliptica.
The size of the studied populations is variable (Marti-
nica ca. 1,500, SBN ca. 1,000, and Zongolica ca. 200
plants/ha).
SBN was partially used (away from the sites occupied
by patches of C. luteoalbus) for coffee and citrus planta-
tions from the 19th century until the 1970s. In contrast, at
Martinica and particularly Zongolica, deforestation was
triggered by conversion of cloud forest into pasturelands
that began in the second half of the 20th century.
Sample collection
At fragment level, leaf tissue from 30 reproductive adults
from each of the three fragments (SBN, Martinica, and
Zongolica) was collected.
At patch level, SGS was analyzed in the largest patch
within SBN (SBN-1 henceforth) containing 258 individuals
in a plot of 30 m 940 m (Fig. 1). At SBN-1 all individ-
uals were sampled and mapped using Cartesian coordi-
nates. Four life cycle stages were used to categorize
individuals as follows: 31 recruits or seedlings (nonrepro-
ductive with only one tuberoid root), 60 juveniles (nonre-
productive individuals with foliar area \25.0 cm
2
), 106
nonreproductive adults (nonreproductive individuals with
foliar area [25.1 cm
2
), and 61 reproductive adults (dis-
playing reproductive structures at least once during
2005–2007 with foliar area [25.1 cm
2
). No additional
C. luteoalbus plants were observed in the immediate
neighborhood of the patch. Maximum distance between
individuals was 8.3 m, and spatial analysis (not shown)
using the Kstatistic (Ripley 1976) after 19 Monte Carlo
simulations indicated that individuals were aggregated.
Enzymatic extraction and electrophoresis
Three hundred milligrams of foliar tissue from sampled
individuals (transported in ice from field and stored at
-70°C) were ground in a mortar and pestle, adding 250 ll
extraction buffer consisting of 0.05 mM Tris–HCl pH 7.5,
2% PVP-40, 5% sucrose, and 0.1% b-mercaptoethanol,
modified from Soltis et al. (1983). Homogenized tissue
embedded in wicks (2 920 mm) was placed in starch gels
(11.5% w/v) and run in an electrophoresis chamber with
constant current of 35 and 50 mA for the R (Chao-Luan
et al. 1999) and PK buffer systems (Soltis et al. 1983)at
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 239
123
Author's personal copy
4°C during 9 and 7 h, respectively. Six legible loci were
found in the R system: anodic peroxidase (Apx-1,Apx-2,
and Apx-3), phospoglucoisomerase (Pgi), esterase (Est),
and acid phosphatase (Acph) and five in the PK system:
malate dehydrogenase (Mdh), malic enzyme (Me), diaph-
orase (Dia), and glutamate oxalacetate transaminase (Got-1
and Got–2). Loci and alleles were designated with
ascending ordered numbers and letters, respectively, from
the most anodally to the most cathodally migrating. No
genotype was identical, thus clonal propagation was ruled
out in the C. luteoalbus sampled individuals.
Genetic diversity measurements
For all fragments and SBN-1 we measured: (1) percentage
of polymorphic loci (P), (2) mean number of alleles per
locus (A), (3) observed mean heterozygosity (H
O
), and (4)
expected mean heterozygosity (H
E
) (Hartl and Clark 2007).
Hardy–Weinberg equilibrium for genotype frequency was
tested using the Monte Carlo method (Weir 1990). Com-
parisons of H
O
and H
E
among life cycle stages and popu-
lations were done with Kruskal–Wallis nonparametric
analysis of variance (ANOVA) (Zar 1999). Additionally,
inbreeding coefficients (F
IS
) were estimated for each of the
life cycle stages and populations, and significant differ-
ences from 0 for each estimator were determined by chi-
squared tests following Weir (1990).
Genetic structure at fragment level
Genetic variation among and within fragments was
assessed with the Weir and Cockerham (1984) modified
estimators of Wright (1965)Fstatistics. Departures from
0 for each locus were determined by chi-squared tests
following Weir (1990) for F
IS
and F
IT
and Workman
and Niswander (1970) for F
ST
. Confidence intervals
(95%) of Fstatistics were obtained from bootstrapping
for all loci, and from jackknifing for single-locus esti-
mates. All these analyses were done using TFPGA 1.3
(Miller 1997). To test for isolation by distance, the
coefficient of regression between pairwise F
ST
values
and pairwise geographical distances was calculated and
compared with those obtained from 1,000 permutations
of individual genotypes among fragments (Hardy and
Vekemans 2002).
Evidence of recent bottlenecks
After a bottleneck, the heterozygosity expected under
Hardy–Weinberg equilibrium (H
E
) should be greater than
that expected under mutation–drift equilibrium (H
eq
)in
more than 50% of loci (Cornuet and Luikart 1996). This
hypothesis was assessed at each fragment and for each life
cycle stage at SBN-1 using the BOTTLENECK program
1.2 (Cornuet and Luikart 1996).
Fig. 1 Spatial distribution of
Cyclopogon luteoalbus plants
within patch SBN-1 with inset
showing SBN fragment and the
four patches existing inside it
(full squares, the upper of them
being the studied SBN-1 patch).
Reproductive adults (white
triangles), nonreproductive
adults (black triangles),
juveniles (circles), and
seedlings (stars). Axes scale in
meters
240 L. Jua
´rez et al.
123
Author's personal copy
Spatial genetic structure at patch level
To analyze SGS by life cycle stage, the number of dis-
tance classes to be used for kinship analyses was
obtained using Sturge’s rule (Legendre and Legendre
1998). These distance intervals were chosen to maximize
the number of pairs of plants (at each life cycle stage)
contained in the distance interval. The chosen lags were
2 m for the first 11 classes and 25, 30, and 40 m for the
last three distance classes. Pairs of seedlings were found
only in 8 distance classes, while pairs of individuals of
the remaining life cycle stages were found in all 14
distance classes.
The mean kinship coefficients (F
ij
) proposed by Loiselle
et al. (1995) were estimated for pairs of individuals within
each of the 14 distance intervals using the program SPA-
GeDI 1.3 (Hardy and Vekemans 2002). The coefficients of
the regressions (b
F
) between the kinship coefficients and
the logarithm of the distances between paired individuals
were estimated within each life cycle stage and also for all
individuals in the pooled dataset. SGS is significant when
the slopes of these regressions are negative and different
from 0. The hypothesis that SGS is significant was tested
by comparing the values of the slopes with those obtained
from 1,000 permutations of individual genotypes. To allow
nonstatistical assessment of differences between different
studies, the S
p
statistics were estimated using the formula
S
p
=-b
F
/(F
1
-1), where F
1
is the average kinship
coefficient for the shortest distance class (Vekemans and
Hardy 2004).
SGS differences between life cycle stages were made
using autocorrelation coefficients (which are another
measure of SGS) between geographical distance and
genetic similarity using the heterogeneity test proposed by
Smouse et al. (2008, GENEALEX 6.4). Significance of the
between life cycle stages differences in autocorrelation
coefficients within each distance class were tested with the
t
2
statistic using sequential Bonferroni correction, while the
significance of between life cycle stages differences in
autocorrelation coefficients in the pooled dataset was tested
using the multiclass test statistic x(Peakall and Smouse
2006).
Results
Genetic diversity
At both fragment and patch level all loci were polymorphic
(P=100%), with mean number of alleles per locus of
2.12 ±0.13 and 2.15 ±0.10 (Table 1), respectively.
Genetic diversity was high with mean H
O
of 0.46 ±0.04
and 0.40 ±0.02 at fragment and patch level, respectively
(Table 1). The observed heterozygosity at fragment level
was higher than expected, but at patch level the reverse was
true.
Table 1 Genetic diversity statistics for patch SBN-1 and for three fragments in Veracruz (SBN, Martinica, and Zongolica) of Cyclopogon
luteoalbus
NPAH
O
H
E
F
IS
?
Life cycle stage
Seedlings 28.36 100 2.00 0.365 0.443 0.176** 3 0
Juveniles 51.18 100 2.20 0.404 0.463 0.127*** 5 0
Nonreproductive
adults
92.18 100 2.20 0.425 0.473 0.101*** 3 2
Reproductive adults 53.18 100 2.20 0.410 0.453 0.095*** 3 2
Mean 56.22 100 2.15 0.401 0.458 0.125
SD 26.48 0.10 0.026 0.013 0.037
Fragments
SBN 27.70 100 2.27 0.415 0.442 0.061 ns 2 1
Martinica 26.30 100 2.00 0.509 0.461 -0.104* 1 2
Zongolica 25.30 100 2.10 0.479 0.433 -0.106* 1 2
Mean 26.43 100 2.12 0.468 0.445 -0.05
SD 1.20 0.13 0.048 0.014 0.09
If all loci are found in every plant, Nwill equal the number of plants, but it will decrease if some loci are missing in some individuals. Number of loci
(total 11 loci) displaying significant deviations from Hardy–Weinberg equilibrium: heterozygosity deficiency (-) and heterozygosity excess (?)
Naverage number of genotypes analyzed, Ppercentage of polymorphic loci, Anumber of alleles per locus, H
O
and H
E
mean observed and
expected heterozygosity, respectively, F
IS
inbreeding coefficient
Asterisks indicate the probability of v
2
under the null hypothesis that F
IS
is equal to zero (Weir 1990), * P\0.05, ** P\0.01, *** P\0.001
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 241
123
Author's personal copy
Hardy–Weinberg equilibrium was found in only 2 out of
11 loci in the patch- and fragment-level analysis. At frag-
ment level, 4 and 5 loci had deficiency and excess of
heterozygotes, respectively, while at patch level these
figures were 14 and 4, respectively. There were no sig-
nificant differences in genetic diversity among fragments
[Kruskal–Wallis tests: H
(2,31)
=1.19, P=0.55 and H
(2,31)
=
0.21, P=0.89 for H
O
and H
E
, respectively]. Similarly, no
significant differences were found for genetic diversity
among different life cycle stages [Kruskal–Wallis tests:
H
(3,44)
=0.64, P=0.88 and H
(3,44)
=1.06, P=0.78 for
H
O
and H
E
, respectively].
At fragment level H
O
was 0.415, 0.509, and 0.479 (SBN,
Martinica, and Zongolica fragments, respectively). The
average of H
O
over life cycle stages at patch level was
0.401, decreasing from 0.41 in reproductive adults to 0.365
in seedlings.
At fragment level, F
IS
was negative and different from 0
(except for at SBN, Table 1). At patch level, all F
IS
values
for life cycle stages were positive and significantly differ-
ent from 0 (Table 1). The value of F
IS
for seedlings was
1.86 times higher than that of reproductive adults, while the
F
IS
values of juveniles and nonreproductive adults were
1.34 and 1.07 times higher than that of reproductive adults
(Table 1). Allele frequencies are given in Appendix 1.
Genetic structure at fragment level
CIs (95%) for the mean values of F
IT
and F
IS
included
zero, while mean values of F
ST
were positive and signifi-
cantly different from zero, indicating that ca. 12% of
genetic diversity was distributed among fragments
(Table 2). The F
ST
value between the SBN and Zongolica
fragments was the highest (F
ST
=0.208), followed by the
F
ST
value between the SBN and Martinica fragments,
which was two times higher than the F
ST
value between
Martinica and Zongolica (F
ST
=0.128 and 0.066, respec-
tively). The distance between the SBN and Martinica
fragments is ca. 8 km, while Martinica and Zongolica are
more than 100 km apart. No isolation by distance was
found: the slope (b=0.000013) of the regression between
paired F
ST
values and geographic distance did not differ
from that obtained from permutation test (P=0.67).
Evidence of recent bottlenecks
Both within fragments and within all life cycle stages at
SBN-1, most of the 11 loci analyzed had higher expected
heterozygosity under Hardy–Weinberg (HW) equilibrium
than under drift–mutation equilibrium (Table 3). Similar
results were obtained using the infinite allele model and the
stepwise mutation model.
Spatial genetic structure at patch level
Spatial genetic structure analyses indicated significant
positive autocorrelation for all life cycle stages (and for the
pooled data set) at the shortest distance class (Fig. 2a–e).
Different life cycle stages showed different levels of
relatedness at the shortest distance class, which includes
the following significant kinship coefficients (F
ij
):
0.142 ±0.023 SE for all life cycles pooled (Fig. 2a),
0.03 ±0.025 SE (marginally significant, P=0.052,
Fig. 2b) for seedlings, 0.192 ±0.041 SE (Fig. 2c) for
juveniles, 0.22 ±0.044 SE (Fig. 2d) for nonreproductive
adults, and 0.154 ±0.027 SE (Fig. 2e) for reproductive
adults. Significant slopes of the regression between the
logarithm of distances and the kinship coefficients for the
pooled data set (b
F
=-0.045, P\0.001), for seedlings
(b
F
=-0.023, P\0.05), and for juveniles, nonreproduc-
tive adults, and reproductive adults (b
F
=-0.067, -0.062,
and -0.059, P\0.001, respectively; Table 4) confirmed
that genetic relatedness decreases with distance. The S
p
value for the pooled data set was 0.053, increasing from
seedlings to juveniles to reproductive adults and then
decreasing slightly in nonreproductive adults (S
p
=0.024,
0.084, 0.08, and 0.069, respectively; Table 4), suggesting
that spatial genetic structure increases with age.
The heterogeneity test was not significant after Bon-
ferroni sequential correction for within distance–class
comparisons between the life cycle stages (Appendix 2).
However, the global test (x) was significant for all
Table 2 Wright’s Fstatistics for three Cyclopogon luteoalbus pop-
ulations in Veracruz cloud forest
Loci F
IT
F
ST
F
IS
Got-1 0.363*** 0.167* 0.235***
Got-2 -0.432*** -0.006 ns -0.423***
Apx-1 0.331* 0.374*** -0.067 ns
Apx-2 0.123* 0.068 ns 0.059 ns
Apx-3 0.102*** -0.001 ns 0.103*
Me 0.388*** 0.092* 0.326***
Pgi 0.250*** 0.266*** -0.022 ns
Acph -0.272*** 0.054 ns -0.344***
Dia 0.419*** 0.249*** 0.226***
Mdh -0.214*** 0.029 ns -0.251***
Est -0.493*** 0.006 ns -0.502***
Mean 0.068 ns 0.128* -0.069 ns
SD 0.10 0.04 0.09
Upper limit 95% CI 0.240 0.029 0.087
Lower limit 95% CI -0.121 0.053 -0.223
ns not significant, CIs confidence intervals
Probability of the v
2
goodness-of-fit test with respect to Hardy–
Weinberg equilibrium: * P\0.05; ** P\0.01; *** P\0.001
242 L. Jua
´rez et al.
123
Author's personal copy
comparisons between life cycle stages except seedlings
versus reproductive adults, suggesting differences in
strength for SGS (Appendix 2).
Discussion
Genetic diversity
The high genetic diversity found in C. luteoalbus is par-
tially explained by the fact that only polymorphic loci were
considered here. Comparing the average values of H
O
(0.46
and 0.40 in the fragments and at patch SBN-1) with those
obtained in other studies in which only polymorphic
loci were analyzed and \5 populations were studied,
C. luteoalbus genetic diversity was higher than that observed
for a narrowly distributed and obligated outcrossing orchid
(H
O
=0.33 ±0.02, N=2, Pleurothallis fabiobarossii,
Borba et al. 2001), similar to a widespread and preferen-
tially outcrossing orchid (H
O
=0.46 ±0.15, N=3,
Govenia superba, Garcı
´a-Cruz et al. 2009) but lower than
another widespread and preferentially outcrossing orchid
(H
O
=0.52 ±0.18, N=2, Govenia mutica, Garcı
´a-Cruz
et al. 2009), lower than that for orchids with mechanical
barriers to self-pollination (H
O
=0.61, N=1, Bulbo-
phyllum bidentata, and H
O
=0.49, N=1, B. rupicolum,
Azevedo et al. 2007), but higher than that for orchids
where pollination by deceit prevents inbreeding (H
O
=
0.21 ±0.06, N=5, Sophronitis sincorana, Borba et al.
2007). However, some caution about these comparisons
should be applied, because only polymorphic loci were
analyzed and only between 36% and 55% of those found in
the cited species were shared with those used in C. lute-
oalbus. High genetic diversity seems to be related with
cross-pollination, and it has also been related to high out-
crossing rates and floral synchrony as a result of increased
gene flow via pollen (Domı
´nguez et al. 2005; Ferrer et al.
2004), and this could be true in C. luteoalbus, a species
with synchronous blooming and whose negative F
IS
values
indicate high outcrossing rates.
Between-fragment gene flow through pollen exchange
is expected in species with a mechanism of pollination
by deceit which promotes long-distance pollen flow. In
food-rewarding species such as C. luteoalbus, pollinator
foraging pattern is spatially limited as pollinators spend
more time foraging within the same patch (Cozzolino
and Widmer 2005). Although this foraging pattern pro-
motes local mating and positive inbreeding coefficients
at patch level (as was found in C. luteoalbus), the high
genetic diversity and negative F
IS
values (indicating high
outcrossing rates) found at fragment level suggest that
some interpatch and interfragment pollen exchange is
Table 3 Expected (Exp.) and observed (Obs.) values for the number
of loci showing higher expected heterozygosity under HW than under
drift–mutation equilibrium for the infinite allele model (IAM) and the
stepwise mutation model (SMM), and Pvalues for the Wilcoxon test
comparing the observed and expected values
IAM SMM
Number of loci with heterozygosity
excess
Pvalue Number of loci with heterozygosity
excess
Pvalue
Exp. Obs. Wilcoxon test Exp. Obs. Wilcoxon test
Life cycle stage
Seedlings 4.67 11
0.0002
5.24 11 0.0002
Juveniles 4.60 11
0.0002
5.37 10 0.0005
Nonreproductive adults 4.39 11
0.0002
5.14 11 0.0002
Reproductive adults 4.64 11
0.0002
5.25 10 0.0007
Fragments
SBN 4.99 11
0.0002
5.66 9 0.004
Martinica 4.27 10
0.0005
4.77 10 0.0005
Zongolica 4.38 9
0.0009
4.94 8 0.0068
Results from bottleneck analyses in SBN-1 patch, and in three fragments of Cyclopogon luteoalbus from cloud forest of Veracruz. Only the
Wilcoxon test is reported, as all other tests used by BOTTLENECK program (sign, standardized departure, and Wilcoxon tests) gave similar
results
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 243
123
Author's personal copy
occurring as in other terrestrial orchids (Caladenia ten-
taculata, Peakall and Beattie 1996). Also, even if seed
dispersal is leptokurtic, some between-fragment gene
flow through wind-dispersed seeds (that may travel long
distances) has been found (Jersakova and Malinova
2007). This is also possible in C. luteoalbus,asa
between-fragment distance of 4.8 ±2.7 SD km has been
reported by
´az-Toribio (2009) for 14 fragments in a
15 km 915 km area which included the SBN and
Martinica fragments.
Fig. 2 Correlograms showing
kinship coefficients F
ij
and their
95% confidence intervals
(vertical bars), calculated by
1,000 permutations within each
distance class. Kinship
coefficients for all life cycle
stages pooled (a), for seedlings
(b), for juveniles (c), for
nonreproductive adults (d), and
for reproductive adults (e). Data
from Cyclopogon luteoalbus
plants found at patch SBN-1
244 L. Jua
´rez et al.
123
Author's personal copy
Bottleneck analyses showed evidence of recent reduc-
tions in effective population size of C. luteoalbus at both
fragment and patch level. A decrease in genetic diversity is
expected after a reduction of effective population size
(Aguilar et al. 2008; Ghazoul 2005; Honnay and Jacque-
myn 2007) due to random fixation and loss of alleles and
consequent reduction of heterozygosity, but this would be
evident in C. luteoalbus if monomorphic loci were also
included in this analysis. Genetic diversity is also affected
if the efficiency of pollinator services decreases as a con-
sequence of fragmentation (Aguilar et al. 2008; Ghazoul
2005; Honnay and Jacquemyn 2007). In C. luteoalbus the
effect of reduction of population size and genetic drift in
the remnant fragments are counterbalanced by gene flow
among them. Also, the high genetic diversity in C. lute-
oalbus could be partially due to selection favoring het-
erozygotes, as discussed in the next subsection.
It is interesting to note that the inbreeding coefficient F
IS
is negative at the Martinica and Zongolica fragments (but
not significant at the SBN fragment), while it is positive
and significant at the SBN-1 patch (Table 1). This suggests
that within-fragment heterozygotes are favored by natural
selection (Lewontin 1974; Mitton and Grant 1984) while at
patch level local mating inside family groups favors a
Wahlund effect.
Genetic structure at fragment level
Genetic differences between fragments were found to be
moderate (F
ST
=0.12 ±0.04 SD). However, no evi-
dence of isolation by distance was found. First, the
F
ST
=value for C. luteoalbus falls within the range of
F
ST
values for data reported for allozyme studies made in
70 orchid species (G
ST
or F
ST
) of 0.012–0.75 (Forrest
et al. 2004). This considerable variation has been attrib-
uted to a high variability in life histories characteristics
such as reproductive strategies and generation times, and
to the degree of isolation (Forrest et al. 2004). Bottle-
necks and isolation have been found to increase genetic
structure among populations, resulting in among-popula-
tion differentiation in allelic frequencies (Honnay and
Jacquemyn 2007).
Fig. 2 continued
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 245
123
Author's personal copy
SBN fragment, which is immersed in an urban matrix
and surrounded by coffee plantations, is the more differ-
entiated fragment, as shown by the larger F
ST
value
between this fragment and Martinica and Zongolica
(F
ST
=0.128 and F
ST
=0.208, respectively). Moreover,
larger portions of SBN (away from the sites where
C. luteoalbus is present) were used for coffee and citrus
plantations from the 19th century until the 1970s. In con-
trast, at Martinica and particularly Zongolica, deforestation
began more recently (the second half of the 20th century)
by the conversion of cloud forest into pasturelands.
This moderate but significant genetic differentiation
suggests that, even if C. luteoalbus is a locally abundant
species, colonization of new fragments depends on (1)
specific habitat requirements (see Species and study
sites’), and (2) even if abandoned, coffee plantation
physicochemical soil properties may remain unsuitable for
orchid colonization as the continuous removal of under-
story during coffee cultivation eliminates soil surface lay-
ers. The lack of suitable sites for recruitment of
C. luteoalbus individuals in grazing lands, urban environ-
ments, and present or past coffee plantations confined this
species to settle on relatively undisturbed cloud forest
fragments, which are becoming less frequent.
Despite the fragmentation process, there is still some
between-fragment gene flow that hinders the development
of stronger differentiation between them. This is suggested
by the small F
ST
values and by the lack of evidence of
isolation by distance (i.e., there was no progressive
increase of genetic differentiation with geographic dis-
tance). In this sense, it is important to note that the frag-
mentation process has developed unevenly in space and
time and so its probable effect on F
ST
values is by now
spatially variable. Taking into account that the generational
time of C. luteoalbus is 25 years, it can be expected that, in
places where fragmentation started in the 18th or early 19th
century for coffee cultivation, its current effect should be
stronger than in places that were cleared during the last
decades of the 20th century to be used as pasturelands.
Spatial genetic structure at patch level
As we expected, strong spatial genetic structure was found at
patch SBN-1. The S
p
statistic for all life cycle stages pooled
(S
p
=0.05 ±0.02 SD) was similar to that found for mixed
mating species (S
p
=0.03 ±0.03 SD, N=7) and herba-
ceous plants (S
p
=0.04 ±0.06 SD, N=24) but higher
than those expected for species with gravity seed dispersal
and animal-dispersed pollen (S
p
=0.02 ±0.01 SD,
N=17 and S
p
=0.01 ±0.01 SD, N=6, respectively), as
reported by Vekemans and Hardy (2004).
The short distance at which SGS was detected (\6m)
falls within the range recorded for five other terrestrial
orchids (range of minimal distance at which SGS equals
zero in the correlograms: 5–16 m; Peakall and Beattie
1996; Chung et al. 1998,2005; Chung et al. 2004a,b).
This distance reflects the spatial structure of the family
group derived from restricted seed and pollen dispersal
(Sokal and Wartenberg 1983).
Inbreeding in SBN-1 (F
IS
=0.12) could be a conse-
quence of the foraging behavior of the pollinators that
promotes local mating. In C. congestus, the halictid bee
Pseudoaugochloropsis graminea visits 1–4 flowers per
inflorescence and 1–2 inflorescences per visit on a patch
(Singer and Sazima 1999). As a similar foraging pattern
was observed in bee visitors to C. luteoalbus (Caenaug-
ochlora cupriventis and Augochlora sp.; Jua
´rez and
Montan
˜a, personal observation), capsules set within an
inflorescence could be derived from both geitonogamous
pollination and within-patch pollination. On the other hand,
experimental and observational studies show that seed
dispersal is highly leptokurtic in orchids, as most of the
seeds are dispersed by gravity near to the mother plant
(Chung et al. 2005; Jersakova and Malinova 2007). Both
the pollination mechanism and the leptokurtic seed dis-
persal may lead to outcrossed, biparental-inbred, and self-
fertilized seeds (and also to a Wahlund effect derived from
within-patch breeding). All these factors may serve to
explain the strong spatial genetic structure.
Four scenarios involving proximal causes such as pollen
and seed dispersal have been envisaged in the literature for
SGS development: (1) strong SGS and inbreeding, when
dispersal of both pollen and seeds is limited (Caujape
´-
Castells and Pedrola-Monfort 1997; Chung et al. 2004a),
(2) weak SGS without inbreeding, when dispersal of both
pollen and seeds is widespread (Epperson and Allard
1984), (3) SGS without inbreeding, when pollen dispersal
is random and seed dispersal limited (Hamrick and Nason
1996), and (4) SGS absent, when pollen dispersal is limited
but seed dispersal is not (Chung et al. 2003a). The case of
Table 4 Statistics of spatial genetic structure in Cyclopogon
luteoalbus within SBN-1
Life cycle stage F
ij
at 2 m b
F
(slope) S
p
Seedlings 0.039 -0.023* 0.024
Juveniles 0.192 -0.067* 0.084
Nonreproductive adults 0.220 -0.062* 0.080
Reproductive adults 0.154 -0.059* 0.069
All samples 0.142 -0.045* 0.053
The kinship coefficient (F
ij
) between individuals in the shortest dis-
tance class, and the slope (b
F
) of the regression between the kinship
coefficients and the logarithm of the distances between paired
individuals
*P\0.05
246 L. Jua
´rez et al.
123
Author's personal copy
C. luteoalbus seems to points to the first scenario where
SGS is promoted by restricted pollen and seed dispersal.
SGS varies across life cycle stages because demographic
processes affect each life cycle stage differently (Jacque-
myn et al. 2007; Kalisz et al. 2001). If reproductive adults
have strong SGS, within-patch mating will result in marked
SGS in the other stages (Kalisz et al. 2001). According to
Kalisz et al. (2001) the lower SGS of seedlings as com-
pared with other life cycle stages can be due to historical
contingencies, local selection, or random processes. The
current reproductive cohort could have been the product of
a single population founding event (that triggered a genetic
bottleneck; Nei et al. 1975; Cornuet and Luikart 1996),
creating a pattern of genetic structure in adults that is
decaying in offspring cohorts. Microlocal environmental
conditions (such as fine-scale genetic interactions of
mycorrhizal associations; Taylor and Bruns 1999) may
favor survival of spatial aggregates of relatives. Kalisz
et al. (2001) mention that also balancing selection favoring
increasing levels of heterozygosity with age could partially
explain differences between offspring and adults.
Also the differences in SGS among life cycle stages
suggest a differential reproductive contribution over time
which is due to the fact that a different set of adults
reproduce each year. In other words, each year seedlings
descend from the same set of adults while juveniles and
adults (recruited from different cohorts of seedlings) des-
cend from a different set of adults, and are established
close to their mother plant. Five-year demographic data of
C. luteoalbus show that only 0.4% of adults set fruits all
years while 50% set fruits only 1 out of the 5 years (7, 18,
and 25% set fruit in 4, 3, and 2 out of 5 years, respectively;
Jua
´rez and Montan
˜a, unpublished data).
Because inbreeding and the consanguinity level affect the
kinship coefficient (F
ij
) (coancestry coefficient sensu Crow
and Kimura 1970), the expected values of 0.250 for full sibs
and 0.125 for half-sibs under random mating (Loiselle et al.
1995) must be corrected (Weir et al. 2006). Seedlings and
juveniles at SBN-1 are not originated by random mating
(F
IS
=0.176 and 0.127, respectively; Table 1), and their
expected F
ij
values are 0.213 and 0.222 assuming mating
between full sibs, 0.106 and 0.111 assuming mating between
half-sibs, 0.053 and 0.055 assuming mating between first
cousins, and 0.027 and 0.028 assuming mating between
second cousins (estimated from Crow and Kimura 1970,
eqn 3.3.1, p. 69). The observed values of the mean kinship
coefficient for seedlings and juveniles at 0–2 m distance
class were F
ij
=0.039 and 0.192, respectively, suggesting
that recruited seedlings are most likely progeny from first or
second cousins while juveniles are most likely progeny from
full siblings. The kinship coefficient of seedlings is lower
than that of juveniles, which suggests differential survival
favoring closely related individuals.
Differences in the length of residence in each life cycle
stage can contribute to between life cycle stages SGS
changes: seedlings came from only one cohort and remain
at this stage only 1 year, while juveniles came from at least
five cohorts (minimum age of first reproduction is 5 years)
and may remain more than 5 years at this stage. The
accumulation of juveniles during this time interval arising
from within-patch mating may reinforce the juvenile SGS,
and a similar reasoning applies for adult stages.
The heterogeneity test shows that SGS of seedlings and
reproductive adults do not differ (Appendix 2), most
probably due to the genetic similarity between reproductive
adults and their progeny and to the fact that seedlings of
different mother plants are more of less evenly distributed
in space because their seed shadows largely overlap inside
the patch. Afterwards, the mechanisms described above
may be responsible for creating the difference between
SGS of seedlings and the other life cycle stages (Tonsor
et al. 1993; Kalisz et al. 2001).
After fragmentation, plants from a formerly continuous
population remain isolated in patches and inbreeding
becomes the predominant mechanism of mating. As plants
of these patches senesce and die, genetic variability is
eroded and inbreeding depression may appear. Seedlings
produced in these populations have a high inbreeding
coefficient which decreases as life cycle develops, probably
due to the expression of deleterious allele combinations at
different life cycle stages or due to the negative effects of
fragmentation, as Aguilar et al. (2008) suggests.
Nevertheless, the genetic viability of populations with a
history of bottlenecks and isolation may not be severely
affected by inbreeding because deleterious alleles in the
homozygous condition are purged by natural selection
(Wallace 2003).
In conclusion, this study shows the existence of spatial
genetic structure inside patches, and this coincides with a
scenario where gene flow via pollen and seed dispersal is
restricted. Also, there are high levels of genetic diversity
and moderate between-fragment genetic differentiation that
points to a scenario of moderate gene flow via between-
fragment pollen and seed dispersal. At patch level, local
mating may result in high levels of autogamy, geitonogamy
or biparental inbreeding. By contrast, all the few between-
fragment matings are between unrelated plants. The neg-
ative effects of fragmentation revealed by recent bottle-
necks and the high inbreeding coefficients (that decrease as
the life cycle develops) may result in local extinctions at
patch or even fragment level (when there are few and/or
small patches in the fragment).
However, as long as pollinator services remain in an
important number of fragments bearing suitable habitats,
long-distance pollen and seed dispersal (even if not
very frequent) will counterbalance negative effects of
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 247
123
Author's personal copy
fragmentation by promoting gene flow and fragment recol-
onization. Our information was collected in one element of
the current landscape mosaic (forest fragments), but man-
agement of other elements of the mosaic (shadow coffee
plantations) may put in motion another process of coloni-
zation and local extinctions. In these plantations the canopy
is never removed but the understory is regularly cleared to
improve coffee production. However, the vagaries of coffee
prices trigger temporary abandonment of these clearing
practices in periods of low prices, when coffee grains are not
collected. During these periods (whose timing and duration
are unpredictable, but can last for a couple of decades)
C. luteoalbus (and many other herbaceous species) can
colonize the understory and contribute to regional gene flow
until the clearing practices are reestablished.
As the current structure of the fragmented landscape is
sufficient to maintain adequate genetic diversity, stopping
the fragmentation process seems necessary to improve the
viability of C. luteoalbus populations in the cloud forests
of central Veracruz.
Acknowledgments This study is part of L.J.’s PhD dissertation
funded by a CONACyT PhD scholarship and a CONACyT research
grant to C.M. Daniel Pin
˜ero, Luis Eguiarte, Oscar
´os, and Octavio
Rojas commented on early versions of the manuscript. Jorge Garcı
´a,
Francisco Reyes, Da
´nae Cabrera, and Juan Pablo Esparza helped with
laboratory and field work, Phil Brewster with pictures, and Ricardo
Ayala determined bee specimens. Julia Herna
´ndez helped in genetic
analysis performed at the Laboratorio de Gene
´tica de Poblaciones,
INECOL (Instituto de Ecologı
´a, A.C.).
Appendix 1
See Table 5.
Table 5 Allele frequency of 11 loci estimated for four life cycle stages (seedlings, juveniles, nonreproductive adults, and reproductive adults) of
Cyclopogon luteoalbus individuals found at patch SBN-1 and at three fragments in central Veracruz (SBN, Martinica, and Zongolica)
Locus Allele Life cycle stage Population
Seedlings Juveniles Nonreproductive adults Reproductive adults SBN Martinica Zongolica
Got-1 a 0.8462 0.8404 (-) 0.7118 0.8868 0.8519 (-) 0.5714 (-) 0.4231
b 0.15 0.1596 0.2882 0.1132 0.1481 0.4286 0.5769
Got-2 a 0.4750 0.5694 0.6250 0.6759 0.6304 0.6786 (?)
b 0.5250 0.4306b 0.3750 0.3241 0.3696 0.3214
Dia a 0.6667 0.4792 0.6118 (?) 0.6977 0.2778 0.7143 0.8833
b 0.3333 0.5208 0.3882 0.3023 0.4074 0.2857 0.1167
c 0.0000 0.0000 0.0000 0.0000 0.3148 0.0000 0.0000
Est a 0.7333 0.7203 0.6538 0.6639 (?) 0.5667 0.4833 (?)–
b 0.2667 0.2797 0.3462 0.3361 0.4333 0.5167
Mdh a 0.5323 0.5000 0.5051 0.4569 (?) 0.4655 (?) 0.6667 0.5862
b 0.4677 0.5000 0.4949 0.5431 0.5345 0.3333 0.4138
Apx-1 a 0.5769 (-) 0.5114 (-) 0.4247 (-) 0.4464 (-) 0.1667 0.7115 0.7353
b 0.4231 0.4886 0.5753 0.5536 0.8333 0.2885 0.2647
Apx-2 a 0.6935 (-) 0.5268 (-) 0.6857 (-) 0.5492 0.8000 0.6481 0.5200
b 0.3065 0.4732 0.3134 0.4508 0.2000 0.3519 0.4800
Apx-3 a 0.7097 0.5882 0.6404 0.5104 (-) 0.7321 0.6042 0.7045
b 0.2903 0.4118 0.3596 0.4896 0.2679 0.3958 0.2955
Acph a 0.6207 0.6724 (-) 0.5931 0.6356 0.6833 0.4333 (?) 0.5167 (?)
b 0.3793 0.3276 0.4069 0.3644 0.3167 0.5667 0.4833
Me a 0.7143 (-) 0.7143 (-) 0.6180 (-) 0.5700 (-) 0.5536 (-) 0.3333 0.2500 (-)
b 0.2857 0.2245 0.3539 0.3900 0.3929 0.6667 0.7500
c 0.0000 0.0612 0.0281 0.0400 0.0536 0.0000 0.0000
Pgi a 0.5500 0.4153 0.5316 (?) 0.6548 0.6538 0.7143 0.1500
b 0.4500 0.5508 0.4421 0.3333 0.3462 0.2813 0.7167
c 0.0000 0.0339 0.0263 0.0119 0.0000 0.0000 0.1333
Signs in parenthesis indicate significant excess (?) or deficiency (-) of heterozygotes
248 L. Jua
´rez et al.
123
Author's personal copy
Appendix 2
See Table 6.
References
Aguilar R, Quesada M, Ashworth L, Herrerias-Diego Y, Lobo J
(2008) Genetic consequences of habitat fragmentation in plant
populations: susceptible signals in plant traits and methodolog-
ical approaches. Mol Ecol 17(24):5177–5188
A
´lvarez-Buylla ER, Chaos A
´, Pin
˜ero D, Garay AA (1996) Demo-
graphic genetics of a pioneer tropical tree species: patch dynamics,
seed dispersal, and seed banks. Evolution 50:1155–1166
Azevedo MTA, Borba EL, Semir JO, Solferini VN (2007) High
genetic variability in neotropical myophilous orchids. Bot J Linn
Soc 153(1):33–40
Borba E, Felix J, Solferini V, Semir J (2001) Fly-pollinated
Pleurothallis (Orchidaceae) species have high genetic variabil-
ity: Evidence from isozyme markers. Am J Bot 88(3):419–428
Borba EL, Funch RR, Ribeiro PL, Smidt EC, Silva-Pereira V (2007)
Demography, and genetic and morphological variability of the
endangered Sophronitis sincorana (Orchidaceae) in the Chapada
Diamantina, Brazil. Plant Syst Evol 267(1):129–146
Caujape
´-Castells J, Pedrola-Monfort J (1997) Space–time patterns of
genetic structure within a stand of Androcymbium gramineum
(Cav.) McBride (Colchicaceae). Heredity 79(4):341–349
Challenger A (1998) Utilizacio
´n y conservacio
´n de los ecosistemas
terrestres de Me
´xico. CONABIO, Me
´xico
Chao-Luan L, Wang Q, Jiang S, Ge S, Wang K (1999) Genetic
diversity of allozymes in populations of Cycas panzhihuaensis l.
Zhou & sy. Yang. In: Paper presented at the biology and
conservation of cycads–proceedings of the 4th international
conference on cycad biology, Sichuan, Beijing
Chung MY, Chung GM, Chung MG, Epperson B (1998) Spatial
genetic structure in populations of Cymbidium goeringii
(Orchidaceae). Genes Genet Syst 73(5):281–285
Chung MY, Nason JD, Epperson BK, Chung MG (2003a) Temporal
aspects of the fine-scale genetic structure in a population of Cinna-
momum insularimontanum (Lauraceae). Heredity 90(1):98–106
Chung MY, Nason JD, Chung MG (2004a) Spatial genetic structure
in populations of the terrestrial orchid Cephalanthera longib-
racteata (Orchidaceae). Am J Bot 91(1):52–57
Chung MY, Nason JD, Chung MG (2004b) Implications of clonal
structure for effective population size and genetic drift in a rare
terrestrial orchid Cremastra appendiculata. Conserv Biol
18:1515–1524
Chung MY, Nason JD, Chung MG (2005) Spatial genetic structure in
populations of the terrestrial orchid Orchis cyclochila (Orchid-
aceae). Plant Syst Evol 254(3):209–219
Cornuet J, Luikart G (1996) Description and power analysis of two
tests for detecting recent population bottlenecks from allele
frequency data. Genetics 144(4):2001–2014
Table 6 Single-class (t
2
) and multiclass (x) test and associated Pvalues for the heterogeneity test of SGS between different life cycle stages of
Cyclopogon luteoalbus at patch SBN-1 in Veracruz, Mexico
Distance
class
123456 7 8 9 1011121314xP
Interval (m) 0–2 2–4 4–6 6–8 8–10 10–12 12–14 14–16 16–18 18–20 20–22 22–25 25–30 30–40
Life cycle stages pairs
Seedlings versus juveniles
t
2
11.14 1.71 2.35 9.64 2.23 2.60 1.72 1.31 2.48 1.52 3.75 1.67 11.14 1.71 70.51 0.01
P0.01 0.19 0.12 0.01 0.15 0.11 0.23 0.29 0.12 0.08 0.01 0.05 0.01 0.19
Seedlings versus reproductive adults
t
2
8.55 4.36 0.05 1.22 1.65 0.26 0.05 5.50 2.64 0.82 0.35 0.81 8.55 4.36 37.76 0.06
P0.01 0.02 0.85 0.24 0.22 0.67 0.78 0.04 0.12 0.77 0.98 0.87 0.01 0.02
Seedlings versus nonreproductive adults
t
2
4.08 0.17 2.91 1.16 4.29 0.15 1.81 6.55 1.69 0.84 2.07 0.27 4.08 0.17 41.76 0.02
P0.05 0.68 0.10 0.32 0.04 0.71 0.22 0.03 0.17 0.55 0.14 0.94 0.05 0.68
Juveniles versus reproductive adults
t
2
0.61 1.08 10.52 5.97 0.12 3.07 3.62 11.02 0.17 1.31 8.49 1.94 0.61 1.08 62.38 0.01
P0.40 0.35 0.01 0.01 0.78 0.09 0.06 0.01 0.68 0.31 0.01 0.15 0.40 0.35
Juveniles versus nonreproductive adults
t
2
4.05 2.02 25.82 20.47 20.29 6.25 12.33 14.35 0.41 0.90 0.60 3.08 4.05 2.02 73.57 0.01
P0.04 0.18 0.01 0.01 0.01 0.02 0.02 0.01 0.45 0.34 0.47 0.09 0.04 0.18
Reproductive adults versus nonreproductive adults
t
2
3.15 9.77 8.88 7.64 22.80 1.82 6.88 1.53 0.65 0.00 1.96 1.37 3.15 9.77 65.50 0.01
P0.07 0.01 0.02 0.01 0.01 0.20 0.02 0.22 0.37 0.90 0.16 0.27 0.07 0.01
xvalues in bold are significant at P\0.05. Pvalues of t
2
must be compared with 0.0035 =0.05/14, which is the Bonferroni-corrected
significance level (equivalent with 0.05 for only one comparison) for comparison within each of the 14 distance classes
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 249
123
Author's personal copy
Cozzolino S, Widmer A (2005) Orchid diversity: an evolutionary
consequence of deception? Trends Ecol Evol 20(9):487–494
Crow JF, Kimura M (1970) An introduction to population genetics
theory. Harper & Row, NY
´az-Toribio M (2009) Orquı
´deas terrestres como indicadoras de
calidad ambiental en fragmentos de bosque meso
´filo de montan
˜a.
Master Thesis. INECOL, Veracruz, Mexico
Domı
´nguez C, Abarca C, Eguiarte L, Molina-Freaner F (2005) Local
genetic differentiation among populations of the mass-flowering
tropical shrub Erythroxylum havanense (Erythroxylaceae). New
Phytol 166(2):663–672
Dressler RL (1993) Phylogeny and classification of the orchid family.
Dioscorides, Portland
Epperson BK (1993) Recent advances in correlation analysis of
spatial patterns of genetic variation. Evol Biol 27:95–155
Epperson BK, Allard RW (1984) Allozyme analysis of the mating
system in lodgepole pine populations. J Hered 75(3):212–214
Fahrig L (2003) Effects of habitat fragmentation on biodiversity.
Annu Rev Ecol Evol Syst 34:487–515
Fenster CB, Vekemans X, Hardy OJ (2003) Quantifying gene flow
from spatial genetic structure data in a metapopulation of
Chamaecrista fasciculata (Leguminosae). Evolution 57(5):
995–1007
Ferrer MM, Eguiarte LE, Montan
˜a C (2004) Genetic structure and
outcrossing rates in Flourensia cernua (Asteraceae) growing at
different densities in the south-western Chihuahuan desert. Ann
Bot 94(3):419–426
Forrest AD, Hollingsworth ML, Hollingsworth PM, Sydes C,
Bateman RM (2004) Population genetic structure in European
populations of Spiranthes romanzoffiana set in the context of
other genetic studies on orchids. Heredity 92:218–227
Garcı
´a-Cruz J, Gonza
´lez-Astorga J, Sosa V, Herna
´ndez-Gonza
´lez O
(2009) Genetic diversity in six Govenia (Orchidaceae) species
with different pollinator attraction strategies. Int J Plant Sci
170(7):894–905
Ghazoul J (2005) Pollen and seed dispersal among dispersed plants.
Biol Rev 80(3):413–443
Hamrick JL, Godt MJW (1996) Effects of life history traits on genetic
diversity in plant species. Philos Trans R Soc Lond B
351(1345):1291–1298
Hamrick JL, Nason JD (1996) Consequences of dispersal in plants. In:
Rhodes OE, Chesser RK, Smith MH (eds) Population dynamics
in ecological space and time. University of Chicago Press,
Chicago, pp 203–236
Hamrick JL, Murawski DA, Nason JD (1993) The influence of seed
dispersal mechanisms on the genetic structure of tropical tree
populations. Plant Ecol 107(1):281–297
Hardy OJ, Vekemans X (2002) SPAGeDi: a versatile computer
program to analyse spatial genetic structure at the individual or
population levels. Mol Ecol 2:618–620
Hartl DL, Clark AG (2007) Principles of population genetics, 4th edn.
Sinauer Associates, Sunderland
Honnay O, Jacquemyn H (2007) Susceptibility of common and rare
plant species to the genetic consequences of habitat fragmenta-
tion. Conserv Biol 21(3):823–831
IUCN/SSC Orchid Specialist Group (1996) Orchid status survey and
conservation action plan. World Conservation Union, New York
Jacquemyn H, Brys R, Vandepitte K, Honnay O, Roldan-Ruiz I (2006)
Fine-scale genetic structure of life history stages in the food-
deceptive orchid Orchis purpurea. Mol Ecol 15(10):2801–2808
Jacquemyn H, Brys R, Vandepitte K, Honnay O, Rolda
´n-Ruiz I,
Wiegand T (2007) A spatially explicit analysis of seedling
recruitment in the terrestrial orchid Orchis purpurea. New Phytol
176(2):448–459
Jersakova J, Malinova T (2007) Spatial aspects of seed dispersal and
seedling recruitment in orchids. New Phytol 176(2):237–241
Kalisz S, Nason JD, Hanzawa FM, Tonsor SJ (2001) Spatial population
genetic structure in Trillium grandiflorum: The roles of dispersal,
mating, history, and selection. Evolution 55(8):1560–1568
Legendre P, Legendre L (1998) Numerical ecology, 2nd edn.
Elsevier, Amsterdam
Lewontin RC (1974) The genetic basis of evolutionary change.
Columbia University Press, NY
Loiselle BA, Sork VL, Nason J, Graham C (1995) Spatial genetic
structure of a tropical understory shrub Psychotria officinalis
(Rubiaceae). Am J Bot 82:1420–1425
Loveless MD, Hamrick JL (1984) Ecological determinants of genetic
structure in plant populations. Annu Rev Ecol Syst 15(1):65–95
Machon N, Bardin P, Mazer SJ, Moret J, Godelle B, Austerlitz F
(2003) Relationship between genetic structure and seed and
pollen dispersal in the endangered orchid Spiranthes spiralis.
New Phytol 157(3):677–687
Manel S, Schwartz MK, Luikart G, Taberlet P (2003) Landscape
genetics: combining landscape ecology and population genetics.
Trends Ecol Evol 18(4):189–197
Miller MP (1997) Tools for population genetic analysis (TFPGA) 1.3:
a windows program for the analysis of allozyme and molecular
population genetic data (computer software distributed by the
author)
Mitton JB, Grant MC (1984) Associations among protein heterozy-
gosity, growth rate, and developmental homeostasis. Ann Rev
Ecol Syst 15:479–499
Mun
˜oz-Villers LE, Lo
´pez-Blanco J (2008) Land use/cover changes
using landsat tm/etm images in a tropical and biodiverse
mountainous area of central-eastern Mexico. Int J Remote Sens
29(1):71–93
Nei M, Maruyama T, Chakraborty R (1975) The bottleneck effect and
genetic variability in populations. Evolution 29:1–10
Peakall ROD, Beattie AJ (1996) Ecological and genetic consequences
of pollination by sexual deception in the orchid Caladenia
tentaculata. Evolution 50(6):2207–2220
Peakall ROD, Smouse PE (2006) Genalex 6: genetic analysis in excel.
Population genetic software for teaching and research. Mol Ecol
Notes 6(1):288–295
Rasmussen HN (1995) Terrestrial orchids form seed to mycotrophic
plant. Cambridge University Press, UK
Ripley BD (1976) The second-order analysis of stationary point
processes. J Appl Probab 13:255–266
Rzedowski J (1978) La Vegetacio
´ndeMe
´xico. Limusa, Mexico City
Singer RB, Sazima M (1999) The pollination mechanism in the
Pelexia alliance’ (Orchidaceae: Spiranthinae). Bot J Linn Soc
131(3):249–262
Smouse PE, Peakall R, Gonzales E (2008) A heterogeneity test for
fine scale genetic structure. Mol Ecol 17(14):3389–3400
Sokal RR, Wartenberg DE (1983) A test of spatial autocorrelation
analysis using an isolation-by-distance model. Genetics
105(1):219–237
Soltis DE, Haufler CH, Darrow DC, Gastony GJ (1983) Starch gel
electrophoresis of ferns: a compilation of grinding buffers, gel
and electrode buffers, and staining schedules. Am Fern J 73:9–27
Sun M, Wong KC (2001) Genetic structure of three orchid species
with contrasting breeding systems using RAPD and allozyme
marker. Am J Bot 88(12):2180–2188
Taylor DL, Bruns TD (1999) Population, habitat and genetic correlates
of mycorrhizal specialization in the ‘cheating’ orchids Corallorh-
iza maculata and C. mertensiana. Mol Ecol 8:1719–1732
Tonsor SJ, Kalisz S, Fisher J (1993) A life-history based study of
population structure: seed bank to adults in Plantago lanceolata.
Evolution 47:833–843
Van Rossum F, Triest L (2006) Fine-scale genetic structure of the
common Primula elatior (Primulaceae) at an early stage of
population fragmentation. Am J Bot 93(9):1281–1288
250 L. Jua
´rez et al.
123
Author's personal copy
Vekemans X, Hardy OJ (2004) New insights of fine-scale spatial
genetic structure analyses in plant populations. 13:921–935
Wallace LE (2003) The cost of inbreeding in Platanthera leucophaea
(Orchidaceae). Am J Bot 9(2):235–242
Weir BS (1990) Genetic data analysis: methods for discrete popu-
lation analysis. Sinauer Associates, Sunderland
Weir BS, Cockerham CC (1984) Estimating F-statistics for the
analysis of population structure. Evolution 38(6):1358–1370
Weir BS, Anderson AD, Hepler AB (2006) Genetic relatedness
analysis: modern data and new challenges. Nat Rev Genet
7:771–780
Workman PL, Niswander JD (1970) Population studies on south-
western Indian tribes. II. Local genetic differentiation in the
Papago. Am J Hum Genet 22(1):24–29
Wright S (1943) Isolation by distance. Genetics 28(2):114–138
Wright S (1965) The interpretation of population structure by
F-statistics with special regard to systems of mating. Evolution
19:395–420
Zar J (1999) Biostatistical analysis. Prentice-Hall, Upper Saddle River
Genetic structure at patch level of the terrestrial orchid Cyclopogon luteoalbus 251
123
Author's personal copy
... At the population level, molecular tools clarify the action of different evolutionary factors involved in intraspecific genetic variability patterns. Molecular approaches have characterized the impacts of habitat fragmentation on gene flow and orchid population structure (Juárez et al. 2011;Trapnell et al. 2013;Minasiewicz et al. 2018), the evolutionary potential of populations (Kisel et al. 2012;Li et al. 2020), inbreeding (Blambert et al. 2016;Brys and Jacquemyn 2016;Picolo et al. 2016), genetic consequences of pollination syndromes (Gaskett 2012;Papadopulos et al. 2013;Phillips et al. 2020), apomixis (Campacci et al. 2017;Firetti 2018), genetic drift (Pinheiro et al. 2014;Jaros et al. 2016;Yun et al. 2020), and spatial genetic structure of populations on a fine-scale (Ilves et al. 2015;Cortés-Palomec et al. 2019;Hedrén and Lorenz 2019). Other studies have focused on elucidating evolutionary processes involved in the diversification of species and lineages, such as hybridization and polyploidization (Marques et al. 2014), and demographic factors originated by the historical dynamics of climatic fluctuations (Kolanowska et al. 2020). ...
Chapter
The varied patterns of genetic differentiation of vulnerable orchids populations (eg. Cattleya genus) urgently need to be known. Here, we present contextualized examples of genetic studies and discuss conservation challenges for orchids in the neotropical region, as the genetic diversity is essential for conservation and management actions of threatened species. In most of the sampled populations of Cattleya granulosa, an endangered orchid inhabiting the Atlantic Forest remnants, for example, there is a body of evidence of the occurrence of genetic bottlenecks. Our study showed that conservation genetics should consider the spatial distribution of genetic diversity in order to develop appropriate conservation strategies. Thus, further advances in the protection of genetic diversity are needed, since conventional databases and public policies may be insufficient in land use planning aiming at biodiversity conservation. In addition to the need to maintain genetic diversity, it is necessary to understand the reproductive and demographic characteristics caused by the interaction of a number of evolutionary and ecological mechanisms to maintain the evolutionary potential of an orchid species or population.KeywordsPlant conservation geneticsCattleya geneticsOrchidaceaeNeotropical orchidsThreatened orchids
... This reflects the findings of other studies that have reported restricted gene flow via orchid seed, despite their minute size and apparent potential for long-distance movement by wind (Jersaḱováand Malinova, 2007;McCormick and Jacquemyn, 2014;Pinheiro et al., 2014). Although prior research in Hainan revealed relatively low values for the Sp statistic (Zhang et al., 2019), Sp values derived here (0.0050-0.0498) indicate greater consistency with those reported for several other outcrossing orchids (Jacquemyn et al., 2006;Juaŕez et al., 2011;Gigant et al., 2016;Sujii et al., 2019), as well as with values published for a range of other outcrossing, wind-dispersed plant taxa, including trees (Vekemans and Hardy, 2004;Piotti et al., 2013;Soliani et al., 2016). At the scale of the populations examined here, this may reflect a rapid drop-off in detectable relatedness at distance classes above 5-10 m due to overlapping seed and pollen movement among multiple, highly heterozygotic individuals, thereby masking any fine-scale spatial genetic patterning, as Sujii et al. (2019) inferred for neotropical inselberg-dwelling Epidendrum orchids. ...
Article
Full-text available
Introduction Plants confined to island-like habitats are hypothesised to possess a suite of functional traits that promote on-spot persistence and recruitment, but this may come at the cost of broad-based colonising potential. Ecological functions that define this island syndrome are expected to generate a characteristic genetic signature. Here we examine genetic structuring in the orchid Phalaenopsis pulcherrima, a specialist lithophyte of tropical Asian inselbergs, both at the scale of individual outcrops and across much of its range in Indochina and on Hainan Island, to infer patterns of gene flow in the context of an exploration of island syndrome traits. Methods We sampled 323 individuals occurring in 20 populations on 15 widely scattered inselbergs, and quantified genetic diversity, isolation-by-distance and genetic structuring using 14 microsatellite markers. To incorporate a temporal dimension, we inferred historical demography and estimated direction of gene flow using Bayesian approaches. Results We uncovered high genotypic diversity, high heterozygosity and low rates of inbreeding, as well as strong evidence for the occurrence of two genetic clusters, one comprising the populations of Hainan Island and the other those of mainland Indochina. Connectivity was greater within, rather than between the two clusters, with the former unequivocally supported as ancestral. Discussion Despite a strong capacity for on-spot persistence conferred by clonality, incomplete self-sterility and an ability to utilize multiple magnet species for pollination, our data reveal that P. pulcherrima also possesses traits that promote landscape-scale gene flow, including deceptive pollination and wind-borne seed dispersal, generating an ecological profile that neither fully conforms to, nor fully contradicts, a putative island syndrome. A terrestrial matrix is shown to be significantly more permeable than open water, with the direction of historic gene flow indicating that island populations can serve as refugia for postglacial colonisation of continental landmasses by effective dispersers.
... Thus, we hypothesize that the negative inbreeding coefficient observed in our smaller populations is a result of a limited number of individuals surviving to flowering due to loss of individuals to inbreeding depression in early stages. This has been observed in other orchid species (Juárez et al., 2011) and will likely threaten the future viability of these populations (Spielman et al., 2004), emphasizing the need for conservation efforts within small P. leucophaea populations. ...
Article
Full-text available
Protecting biodiversity requires an understanding of how anthropogenic changes impact the genetic processes associated with extinction risk. Studies of the genetic changes due to anthropogenic fragmentation have revealed conflicting results. This is likely due to the difficulty in isolating habitat loss and fragmentation, which can have opposing impacts on genetic parameters. The well‐studied orchid, Platanthera leucophaea, provides a rich dataset to address this issue, allowing us to examine range‐wide genetic changes. Midwestern and Northeastern United States. We sampled 35 populations of P. leucophaea that spanned the species’ range and varied in patch composition, degree of patch isolation, and population size. From these populations we measured genetic parameters associated with increased extinction risk. Using this combined dataset, we modeled landscape variables and population metrics against genetic parameters to determine the best predictors of increased extinction risk. All genetic parameters were strongly associated with population size, while development and patch isolation showed an association with genetic diversity and genetic structure. Genetic diversity was lowest in populations with small census sizes, greater urbanization pressures (habitat loss), and small patch area. All populations showed moderate levels of inbreeding, regardless of size. Contrary to expectation, we found that critically small populations had negative inbreeding values, indicating non‐random mating not typically observed in wild populations, which we attribute to selection for less inbred individuals. The once widespread orchid, Platanthera leucophaea, has suffered drastic declines and extant populations show changes in the genetic parameters associated with increased extinction risk, especially smaller populations. Due to the important correlation with risk and habitat loss, we advocate continued monitoring of population sizes by resource managers, while the critically small populations may need additional management to reverse genetic declines. Protecting biodiversity requires an understanding of how anthropogenic changes impact the genetic processes associated with extinction risk. The well‐studied orchid, Platanthera leucophaea, provides a rich dataset to evaluate the potentially opposing impact of habitat loss and fragmentation on genetic parameters, allowing us to examine range‐wide genetic changes. The once widespread orchid has suffered drastic declines and extant populations show changes in the genetic parameters associated with increased extinction risk, especially in smaller populations.
... Although no indication of sexual reproduction was found in N. miniata at Sennes, the variation obtained by analysis of two SSR loci was still sufficient to describe SGS in the species at this site. Confidence intervals for calculation of F ij estimates were usually small (Table S2), and comparable to those obtained for sexual species using higher numbers of loci (Trapnell, Hamrick & Nason, 2004;Jacquemyn et al., 2006;Chung et al., 2011;Juárez, Montaña & Ferrer, 2011;Helsen et al., 2015). Confidence limits for N ij estimates were Table S2), but N ij estimates were still significant for several consecutive distance intervals, unlike in most studies of sexually reproducing orchids. ...
Article
Orchids have minute, air-filled seeds and are considered to be efficient dispersers and colonizers. However, empirical studies show that most seeds fall within a metre of the mother plant in orchids, and that individuals standing close to each other are often closely related. A poor contribution to gene dispersal by seeds may be compensated for by more efficient dispersal by pollen, but in autogamous or agamospermous orchids this component is not available. Here, we used two highly variable simple sequence repeat loci to analyse fine-scale genetic structure in the agamospermous orchid Nigritella miniata, which is widespread in alpine grasslands in the Dolomites. We studied a densely populated area of 30 × 50 m and an enlarged area of 100 × 500 m with more scattered occurences. Sp statistics, describing the decrease in relatedness with geographical distance, were 0.1073 and 0.0609, respectively, and were among the highest values recorded for the orchid family, revealing a strong genetic structuring. However, some genotypes in the studied area were distributed at distances up to 26.5 km in the Dolomites, which suggests that long-distance seed dispersal occurs occasionally and that populations separated by long distances can still be parts of the same meta-population system.
... Por último, bosque tropical subcaducifolio (BTSC), en el Ejido El Tigre, municipio de Ocampo (543-696 m snm, precipitación anual de 943 mm, temperatura media anual de 26.7 °C), aquí las mayores temperaturas se presentan antes de la temporada de lluvias y la precipitación tiene lugar entre agosto y septiembre (Figura 2 ), separadas entre sí por 15 m, estableciendo la primera a 5 m del inicio del transecto. Las subparcelas estuvieron dispuestas de manera alterna y a una separación de 1 m de la línea del transecto (Cruz-Fernández et al. 2010, Juárez et al. 2011, Pérez-Harguindeguy et al. 2013. En cada subparcela se registró el número de especies y su abundancia; y se colectó un ejemplar de respaldo para cada taxón, dejando cormos y raíces tuberosas para que los individuos se regeneraran. ...
Article
Full-text available
Antecedentes: Los rasgos funcionales de las plantas se relacionan con estrategias adaptativas y forman parte de la diversidad biológica. Por su diversidad de adaptaciones las orquídeas terrestres son un modelo para estudiar las relaciones entre su diversidad, rasgos funcionales y ambiente, pero esto ha sido poco estudiado. Preguntas: ¿Cómo cambia la diversidad de orquídeas y sus rasgos funcionales entre tipos de vegetación? ¿Sus rasgos funcionales permiten reconocer grupos de especies? ¿Su diversidad y rasgos funcionales se asocian con variables climáticas y de estructura de la vegetación? Sitio y años de estudio: Reserva de la Biosfera El Cielo, enero 2016-enero 2017. Métodos: Se registró la riqueza, abundancia y rasgos funcionales de orquídeas terrestres, variables climáticas y de estructura vegetal en tres tipos de vegetación. Se estimó diversidad verdadera, recambio de especies y diversidad de rasgos en cada tipo de vegetación, para reconocer grupos funcionales a nivel interespecífico. La diversidad y rasgos funcionales entre tipos de vegetación se asoció con variables ambientales medidas. Resultados: Los bosques de pino-encino y mesófilo de montaña presentaron mayor diversidad verdadera y tuvieron entre si mayor recambio de especies que con el bosque tropical subcaducifolio, éste tuvo la mayor diversidad de rasgos funcionales. Se reconocieron tres grupos funcionales; la diversidad verdadera y funcional se relacionó con temperatura, humedad y densidad del dosel. Conclusiones: La diversidad verdadera no determina la diversidad funcional; ambientes sometidos a mayor estrés ambiental presentan mayor diversidad funcional. La composición de grupos funcionales entre orquídeas estudiadas refleja eventos de convergencia en sus rasgos funcionales.
... In comparison to mean values for other species, the Sp statistic calculated here for P. pulcherrima (0.0030-0.0058) is significantly lower than for both self-pollinating (0.1431) and outcrossing (0.0126) species [5], even when compared with other orchids, e.g. Orchis purpurea (0.0144 to 0.0148) [56], Cyclopogon luteoalbus (0.053) [57] and Vanilla humblotii (0.020 to 0.045) [58]. There are three possible explanations for this [12]. ...
Article
Full-text available
Background: Gene flow in plants via pollen and seeds is asymmetrical at different geographic scales. Orchid seeds are adapted to long-distance wind dispersal but pollinium transfer is often influenced by pollinator behavior. We combined field studies with an analysis of genetic diversity among 155 physically mapped adults and 1105 F1 seedlings to evaluate the relative contribution of pollen and seed dispersal to overall gene flow among three sub-populations of the food-deceptive orchid Phalaenopsis pulcherrima on Hainan Island, China. Results: Phalaenopsis pulcherrima is self-sterile and predominantly outcrossing, resulting in high population-level genetic diversity, but plants are clumped and exhibit fine-scale genetic structuring. Even so, we detected low differentiation among sub-populations, with polynomial regression analysis suggesting gene flow via seed to be more restricted than that via pollen. Paternity analysis confirmed capsules of P. pulcherrima to each be sired by a single pollen donor, probably in part facilitated by post-pollination stigma obfuscation, with a mean pollen flow distance of 272.7 m. Despite limited sampling, we detected no loss of genetic diversity from one generation to the next. Conclusions: Outcrossing mediated by deceptive pollination and self-sterility promote high genetic diversity in P. pulcherrima. Long-range pollinia transfer ensures connectivity among sub-populations, offsetting the risk of genetic erosion at local scales.
... For example, a lack of genetic variation in epiphytic bromeliad (Guzmania monostachia) populations in Costa Rican forest patches was attributed primarily to anthropogenic barriers to gene flow but could also be influenced by life history traits such as its selective breeding system and limited seed dispersal ability (Cascante-Marin et al., 2014). Higher inbreeding and genetic bottlenecks within a small TMF fragment can also increase the extinction risk of a population (Juárez et al., 2011). ...
Article
Full-text available
Tropical montane forests (TMFs) are major centers of evolutionary change and harbor many endemic species with small geographic ranges. In this systematic map, we focus on the impacts of anthropogenic habitat degradation on TMFs globally. We first determine how TMF research is distributed across geographic regions, degradation type (i.e., deforestation, land-use conversion, habitat fragmentation, ecological level (i.e., ecosystem, community, population, genetic) and taxonomic group. Secondly, we summarize the impacts of habitat degradation on biodiversity and ecosystem services, and identify deficiencies in current knowledge. We show that habitat degradation in TMFs impacts biodiversity at all ecological levels and will be compounded by climate change. However, despite montane species being perceived as more extinction-prone due to their restricted geographic ranges, there are some indications of biotic resilience if the impacts to TMFs are less severe. Species richness and key species interactions can be maintained in mildly degraded sites, and gene flow can persist between TMF fragments. As such, minimally degraded areas such as secondary forests and restored areas could play a crucial role in maintaining the meta-community and ecosystem services of TMFs—either via resource provision or by linking patches of pristine forest. Research deficiencies highlighted include poor research representation in Asian and African TMFs, few assessments of population and genetic-level responses to fragmentation, and little assessment of the impacts of habitat fragmentation at all ecological levels. To address these concerns, we present a list of the top research priorities to urgently address the growing threat of habitat degradation in TMF.
... While heterozygosity requires longer periods of fragmentation before its effects become detectable, rare alleles are usually at a greater risk of loss during bottlenecks caused by fragmentation (Young et al. 1996;Kramer et al. 2008). This effect was shown in other long-lived orchid species (Juárez et al. 2011;Chung et al. 2014). Indeed we found that populations with the lowest value of private alleles and private alleles richness (OL1, OL2 and PR) recently passed through a bottleneck. ...
Article
Full-text available
Plant species that are capable of propagating clonally are expected to be less vulnerable to habitat fragmentation due to their long life span. Cypripedium calceolus L. is a rare, clonal, long-lived orchid species. It has suffered marked decline because of habitat loss and fragmentation and over-collection, yet an IUCN report on this species does not regard fragmentation as a major threat to the species. We applied 13 nuclear microsatellites and cpDNA sequences to identify the patterns of population structure, genetic diversity and connectivity of six remnant local populations of C. calceolus in highly fragmented Gdańsk Pomerania region (N Poland). Despite severe (80%) loss of localities in the studied area we found that the local populations retain high levels of clonal (R 0.86–1) and genetic diversity (He = 0.572). However, their differentiation is relatively high (FST = 0.132 for nuclear SSR and FST = 0.363 for cpDNA) despite close geographic proximity (0.6–57 km). Bayesian clustering classified populations according to their geographic origin with little admixture. Low genetic connectivity between the remnant populations shows that the current gene flow is too low to serve as a cohesive force in a fragmented habitat, which may impede a quick response to environmental change. The species’ ability to retain ancestral variation may help withstand fragmentation, but in the light of observed extirpation rate it should be rather considered as a factor that only delays local populations’ extinction. This leads to the conclusion that habitat loss and fragmentation should be regarded as a real threat to stability of C. calcelolus populations.
... Em nível populacional, as ferramentas moleculares esclarecem eventos de diversidade genética, estrutura genética, gargalos genéticos, fluxo gênico, endogamia e deriva genética. Pesquisas com essas abordagens têm resolvido os impactos da fragmentação do hábitat sobre fluxo gênico e estrutura da população de orquídeas (WALLACE, 2002;JACQUEMYNA et al., 2007;JUÁREZ et al., 2011), consequências genéticas de síndromes de polinização (TREMBLAY et al., 2005;VEREECKEN, 2010;GASKETT, 2011;2012), potencial evolutivo das populações (ACKERMAN; WARD, 1999;ACKERMAN et al., 2007;KISEL et al., 2012), endogamia (WALLACE, 2003;SMITHSON, 2006;BORBA et al., 2007), deriva genética (TREMBLAY; ACKERMAN, 2001;CHUNG;CHUNG, 2007;CHUNG;PARK, 2008;CHUNG, 2009;PHILLIPS et al., 2012) e estrutura genética espacial das populações em fina escala em orquídeas (CHUNG et al., 2004;2005;WALLACE, 2006;JACQUEMYN et al., 2012). Diversos estudos têm sido empregados no intuito de elucidar os processos demográficos e evolutivos envolvidos na diversificação de espécies e linhagens, como hibridação e poliploidia (HEDRÉN et al., 2008), e fatores demográficos originados pela dinâmica histórica de oscilações climáticas em populações do Mediterrâneo e Europa Ocidental de Anacamptis palustris (Jacq.) ...
Article
Full-text available
Esta revisão apresenta a importância da conservação genética para a elaboração de estratégias conservacionistas adequadas à distribuição espacial da diversidade genética, sendo apresentados e contextualizados exemplos de estudos de genética de populações e filogeografia na família Orchidaceae. O principal foco da biologia da conservação é a compreensão e a manutenção da diversidade genética, já que ela fornece o potencial adaptativo e evolutivo de uma espécie. Deste modo, o conhecimento da diversidade genética de uma espécie é primordial para as ações de conservação e manejo. Os níveis de variabilidade genética podem ser determinados pela estrutura populacional de uma espécie, sendo o resultado de características reprodutivas e demográficas, ocasionadas pela interação e ação uma série de mecanismos evolutivos e ecológicos. A avaliação da estrutura populacional das espécies permite descobrir as Unidades Evolutivas Significativas (UES), que devem ser explicitamente definidas em características que realcem o potencial para a sobrevivência da espécie. Por conseguinte, é necessário que haja um foco na preservação da diversidade funcional. Técnicas genéticas são essenciais, pois fornecem estimativas de fluxo gênico entre as populações e, assim, norteiam os esforços para sustentar os níveis genéticos e intercâmbio entre as populações. Os métodos filogenéticos e filogeográficos contribuem para responder diversas questões em biologia da conservação, como quais são os locais prioritários para a conservação, quais espécies conservar e quais são os esforços conservacionistas necessários. E, finalmente, delineiam as estratégias que devem ser tomadas para conservar a maior quantidade de diversidade genética, visando manter o potencial evolutivo de uma espécie ou população.
Article
Full-text available
Resumen Se presenta una revisión de las investigaciones en el campo de la Ecología Molecular en México. Entre 1990 y 2016 se identificaron 656 artículos científicos relacionados. Los temas mejor representados son la genética de poblaciones (35.3% de los estudios) y la filogeografía (30.3%), mientras que los campos emergentes de la Ecología Molecular, como la genómica del paisaje, la ecología trófica basada en secuencias de ADN y el análisis del parentesco y la conducta, estuvieron poco representados. Los sistemas más estudiados han sido los animales (58.5%) y las plantas (32.5%), mientras que otros organismos como hongos, protozoarios y bacterias han recibido mucho menos atención. En general, se observa un desarrollo considerable de la Ecología Molecular en nuestro país. Sin embargo, para continuar esta tendencia será necesario incorporar extensivamente los avances tecnológicos como la secuenciación de nueva generación y la bioinformática, así como incursionar en las áreas emergentes de esta disciplina.
Article
When a population experiences a reduction of its effective size, it generally develops a heterozygosity excess at selectively neutral loci, i.e., the heterozygosity computed from a sample of genes is larger than the heterozygosity expected from the number of alleles found in the sample if the population were at mutation drift equilibrium. The heterozygosity excess persists only a certain number of generations until a new equilibrium is established. Two statistical tests for detecting a heterozygosity excess are described. They require measurements of the number of alleles and heterozygosity at each of several loci from a population sample. The first test determines if the proportion of loci with heterozygosity excess is significantly larger than expected at equilibrium. The second test establishes if the average of standardized differences between observed and expected heterozygosities is significantly different from zero. Type I and II errors have been evaluated by computer simulations, varying sample size, number of loci, bottleneck size, time elapsed since the beginning of the bottleneck and level of variability of loci. These analyses show that the most useful markers for bottleneck detection are those evolving under the infinite allele model (IAM) and they provide guidelines for selecting sample sizes of individuals and loci. The usefulness of these tests for conservation biology is discussed.
Article
Analyses of fine-scale and macrogeographic genetic structure in plant populations provide an initial indication of how gene flow, natural selection, and genetic drift may collectively influence the distribution of genetic variation. The objective of our study is to evaluate the spatial dispersion of alleles within and among subpopulations of a tropical shrub, Psychotria officinalis (Rubiaceae), in a lowland wet forest in Costa Rica. This insect-pollinated, self-incompatible understory plant is dispersed primarily by birds, some species of which drop the seeds immediately while others transport seeds away from the parent plant. Thus, pollination should promote gene flow while at least one type of seed dispersal agent might restrict gene flow. Sampling from five subpopulations in undisturbed wet forest at Estación Biologíca La Selva, Costa Rica, we used electrophoretically detected isozyme markers to examine the spatial scale of genetic structure. Our goals are: 1) describe genetic diversity of each of the five subpopulations of Psychotria officinalis sampled within a contiguous wet tropical forest; 2) evaluate fine-scale genetic structure of adults of P. officinalis within a single 2.25-ha mapped plot; and 3) estimate genetic structure of P. officinalis using data from five subpopulations located up to 2 km apart. Using estimates of coancestry, statistical analyses reveal significant positive genetic correlations between individuals on a scale of 5 m but no significant genetic relatedness beyond that interplant distance within the studied subpopulation. Multilocus estimates of genetic differentiation among subpopulations were low, but significant (Fst = 0.095). Significant Fst estimates were largely attributable to a single locus (Lap-2). Thus, multilocus estimates of Fst may be influenced by microgeographic selection. If true, then the observed levels of IBD may be overestimates.
Article
We consider whether changes in population-genetic structure through the life cycle of Cecropia obtusifolia, a tropical pioneer tree, reflect its gap-dependent demography and the role of evolutionary processes that are important for this species. We asked whether the spatial scale at which population-genetic subdivision occurs corresponds to the scale of habitat patchiness created by gap dynamics; whether patterns of seed dispersal and storage in the soil affect spatial genetic patterns; and whether spatial genetic patterns change through the species life cycle. We estimated Wright's F-statistics for six successive life-history stages for individuals grouped into subpopulations according to occurrence in natural gaps, physical proximity, or occurrence within large quadrats. For each life stage, FST -statistics were significantly higher when individuals were grouped by gaps, although concordant patterns across life stages for the three grouping methods were obtained. This supports the hypothesis that patchy recruitment in gaps or among-gap heterogeneity influences the species' genetic structure. F-statistics of seeds collected from females before dispersal (tree seeds), seed-rain seeds, soil seeds, seedlings, juveniles, and adults grouped by gaps, were, respectively: FIT = 0.004, 0.160, 0.121, 0.091, -0.0002, -0.081; FIS = -0.032, 0.124, 0.118, 0.029, -0.016, -0.083; and FST = 0.035, 0.041, 0.003, 0.063, 0.015, 0.002. Spatial genetic differentiation in rain seeds was not significantly lower than that of tree seeds. The loss of genetic structure in the soil seed bank, relative to that found in the seed rain may be due to sampling artifacts, but alternative explanations, such as microsite selection or temporal Wahlund effect are also discussed. If structure among soil seeds is unbiased, the peak in seedling FST may be due to microsite selection. FIS of seeds in the rain and soil were significantly greater than zero. A Wahlund effect is the most likely cause of these positive FIS values. Such fine-scale substructuring could be caused by correlated seed deposition by frugivores. The decrease in FIS from seedlings to adults could result from loss of fine-scale genetic structure during stand thinning or from selection.
Article
Only orchids affect pollination by the deceptive sexual attraction of male insects, a syndrome particularly well developed in Australia. We examined the ecological and genetic consequences of exclusive pollination by sexually attracted male thynnine wasps in the orchid Caladenia tentaculata. Male wasps respond rapidly to flowers artificially presented in 1 × 1 m(2) experimental patches. Sixty of 287 wasps approached within centimeters of the flower, but did not land. Of the remaining 79% who made floral contact, only 7.5% attempted copulation, the step critical for pollination. Wasps only rarely moved among patches (19% of flights) and none attempted copulation a second time, resembling observations in natural populations. We confirmed outcrossing and long distance pollen flow by monitoring how colored pollen moved in natural populations. Pollen movements approximated a linear rather than a leptokurtic distribution (mean distance: 17 m; maximum: 58 m). Pollinator visits varied independently of flower density in three of four populations with most solitary flowers being visited. Allozyme analysis revealed within-population fixation indices (F) close to zero and low levels of differentiation (FST) among populations. Despite behavioral evidence for long distance pollen flow, significant local genetic structure exists, perhaps reflecting restricted seed dispersal. Long distance pollen flow in C. tentaculata may therefore promote outbreeding by minimizing pollen transfers among related neighbors. Although this species is self-compatible, outcrossed progeny develop significantly faster than selfed progeny. Effective pollination at low flower densities could accentuate this advantage. The data are consistent with the predictions that deceptive pollination will result in long distance pollen flow, which may be of selective advantage at low density. Comparative studies of how food reward, food deceptive, and sexual deceptive pollination systems vary within a phylogenetic framework could further illuminate the evolution of sexual deception.
Article
This paper provides a rigorous foundation for the second-order analysis of stationary point processes on general spaces. It illuminates the results of Bartlett on spatial point processes, and covers the point processes of stochastic geometry, including the line and hyperplane processes of Davidson and Krickeberg. The main tool is the decomposition of moment measures pioneered by Krickeberg and Vere-Jones. Finally some practical aspects of the analysis of point processes are discussed.