SciELO - Scientific Electronic Library Online

 
vol.95 número2A review of Bulbinella (Asphodelaceae): distribution, conservation status, and economic importanceEl papel de las características foliares y la depredación por aves en los patrones de herbivoría en un ambiente semiárido del centro de México índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Botanical Sciences

versión On-line ISSN 2007-4476versión impresa ISSN 2007-4298

Bot. sci vol.95 no.2 México abr./jun. 2017

https://doi.org/10.17129/botsci.668 

Ecology

Wild Vanilla planifolia and its relatives in the Mexican Yucatan Peninsula: Systematic analyses with ISSR and ITS

Vanilla planifolia silvestre y sus parientes en la Península de Yucatán, México: análisis sistemáticos con ISSR e ITS

Sara Villanueva-Viramontes1 

Mariana Hernández-Apolinar2 

Germán Carnevali Fernández-Concha1  3  4 

Alfredo Dorantes-Euán1 

Gabriel R. Dzib1 

Jaime Martínez-Castillo1  * 

1 Centro de Investigación Científica de Yucatán, A.C. Mérida, Yucatán, México.

2 Departamento de Ecología y Recursos Naturales, Facultad de Ciencias, UNAM. Cd. México, México.

3 Herbario CICY, Centro de Investigación Científica de Yucatán, A.C. Mérida, Yucatán, México.

4 Orchid Herbarium of Oakes Ames, Harvard University Herbaria, Cambridge, Massachusetts, U.S.A. Corresponding author’s e-mail address: saravilvir@cicy.mx; mariana.hernandez.a@ciencias.unam.mx; carneval@cicy.mx; badip@cicy.mx; gabrid@cicy.mx.


Abstract:

Background:

Vanilla planifolia, a threatened species, is distributed naturally in semi-deciduous and evergreen rain forests of southeastern Mexico and parts of Central America. In the wild, it is difficult to diagnose from another sympatric Vanilla species, and individuals with reproductive structures are usually required.

Questions:

Can ISSR discriminate wild individuals of Vanilla planifolia from other sympatric species of the genus? Can phylogenetic analyses of ITS recover the monophyly of V. planifolia and help identify Vanilla species?

Studied species:

The vanilla (V. planifolia), the second-most important economically spice in the world.

Study site and years of study:

Plant material was collected during 2014 in the Mexican Yucatan Peninsula.

Methods:

We analyzed 88 wild individuals of several Vanilla species. Individuals of five species of Vanilla reported for the region were used as reference. 84 ISSR loci were analyzed using three clustering methods. A phylogenetic inference using ITS was performed.

Results:

ISSR markers clearly discriminate wild Vanilla planifolia, finding definite genetic structure within the species. The three clustering methods identified genetic relationships with congruent patterns. Five groups were found and they corresponded with the species studied. Phylogenetic inference of ITS sequences supported the monophyly of the Vanilla and the resulting cladograms were coherent with the clustering pattern observed in the ISSR studies.

Conclusions:

Both ISSR and ITS analyses are able to identify V. planifolia. Molecular data suggest the presence of V. pompona and a new species of Vanilla in the Yucatan Peninsula.

Keywords: Genetic identification; molecular markers; phylogenetic analysis; vanilla; wild populations

Resumen:

Antecedentes:

Vanilla planifolia es una especie amenazada que se distribuye naturalmente en las selvas tropicales perennifolias y subperennifolias del sureste de México y parte de Centroamérica. En estado silvestre, es difícil de diferenciar de otras especies simpátricas de Vanilla, y usualmente se requieren individuos con estructuras reproductivas.

Preguntas:

¿Pueden los marcadores ISSR discriminar individuos silvestres de Vanilla planifolia de otras especies simpátricas de vainilla?, ¿Puede el análisis de ITS recobrar la monofilia de V. planifolia y contribuir a identificar las especies de Vanilla?

Especie en estudio:

Económicamente, la vainilla (Vanilla planifolia) es la segunda especia más importante en el mundo.

Sitio de estudio y fecha:

La colecta se realizó en 2014 en la Península de Yucatán.

Métodos:

Se analizaron 88 de individuos silvestres de especies de Vanilla. Se utilizaron como testigos individuos de cinco especies de vainilla reportadas para la región. 84 loci de ISSR fueron analizados usando tres métodos de agrupamiento. Se realizó una inferencia filogenética con ITS.

Resultados:

Los ISSR discriminaron claramente a los individuos silvestres de Vanilla planifolia, detectando una estructura genética al interior de la especie. Los tres métodos de agrupamiento mostraron patrones congruentes, encontrándose cinco grupos que correspondieron con las especies estudiadas. El análisis filogenético de las secuencias de ITS mostró la monofilia del género Vanilla, y los cladogramas obtenidos fueron coherentes con los resultados obtenidos con los ISSR.

Conclusiones:

Tanto ISSR como ITS identificaron a Vanilla planifolia. Los análisis moleculares sugieren la presencia de V. pompona y de una nueva especie de Vanilla en la Península de Yucatán, México.

Palabras clave: Análisis filogenético; identificación genética; marcadores moleculares; poblaciones silvestres; vainilla

The genus Vanilla Mill. is composed of possibly more than 150 species. In 1954 Portères reported 110 distinct taxa but many more have been described since (e.g. Soto-Arenas & Dressler 2010), and several more still await formal taxonomic recognition. The genus features a Pantropical distribution and it probably originated in America (Cameron 2000, Ramírez et al. 2007, Bouetard et al. 2010). Vanilla taxonomy has been hampered by several factors, including the following:

  • The difficulty of finding and collecting individuals with fertile structures in the field as a result of the ephemeral, gregarious flowering that usually happens high in the canopy of the forests (Schlüter 2002, Soto-Arenas & Dressler 2010).

  • Flowers, which are fundamental for identification of species, are poorly preserved in herbaria. Usually, herbarium species are sterile or include a single flower, often poorly preserved and irretrievably flattened (or altogether glued to the herbarium sheet).

  • The rarity or inaccessibility of many of the species.

  • The fact that many species are vegetatively indistinguishable.

  • The enormous vegetative variation and phenotypic plasticity associated with the hemiepiphyte growth habit including leaves of different morphologies and sizes in the same individual as a response to age, vegetative development, and light exposure (Putz & Holbrook 1986, Ray 1990).

Vanilla planifolia Andrews has great economic importance, being the source of approximately 95 % of the world’s vanilla extract production (Lubinsky et al. 2008a). The vanilla is the second-most important spice in gastronomy and the cosmetic industry and is one of the most popular flavorings (Cid-Pérez & López-Malo 2011, Álvarez et al. 2013). Vanilla cultivation typically features vegetative propagation and hand-pollination, which ensures fruit production, albeit seeds are never harvested, and plants are never raised from seeds. Vegetative reproduction, plus the lack of new genotypes incorporated via seed, have in combination drastically reduced genetic variation in vanilla plantations (Smith et al. 1992, Soto Arenas 1999, Minoo et al. 2008). Because of this, FAO (1995) lists cultivated V. planifolia as species with a high degree of genetic erosion. Furthermore, wild populations of V. planifolia have been severely damaged by natural habitat destruction and illegal extraction for replenishment of commercial plantations, nearly driving the species to the brink of extinction, at least locally (Soto-Arenas 2006, Soto-Arenas & Solano-Gómez 2007, Menchaca 2010). Due to this, V. planifolia has a “special protection” status (SEMARNAT 2010) in Mexico. As a member of the Orchidaceae, it is included in Appendix 2 of CITES (CITES 2016).

Vanilla planifolia is reported as native from southeastern Mexico, Guatemala, Belize, and Costa Rica. However, a clear picture of its distribution is still uncertain (Soto-Arenas 1999) because relatively few wild individuals have apparently been collected (fewer than 30 individuals, according to Soto-Arenas 2006 and Schlüter et al. 2007). There are also several vegetatively similar species. To make matters worse, flowers in the herbaria are difficult to reconstruct; thus, a significant portion of specimens are probably misidentified as V. planifolia in many herbaria (Soto-Arenas 2009, Soto-Arenas & Dressler 2010). These errors of identification have also been reported in vanilla plantations in Mexico, as we have been able to corroborate in the Mexican states of Quintana Roo (Villanueva-Viramontes, pers. obs.), and Oaxaca (M. Hernández-Apolinar pers. Obs.). The literature also reports this from other countries, such as Ecuador and Guatemala (Soto-Arenas & Dressler 2010). Some of these plantations include, along with true V. planifolia, individuals of species such as Vanilla insignis Ames, Vanilla cribbiana Soto Arenas, Vanilla odorata C. Presl., Vanilla pompona Schiede, and Vanilla sp. nov. aff. V. phaeantha (Soto-Arenas & Dressler 2010).

In the Mexican Yucatan Peninsula (MYP), wild individuals of Vanilla planifolia have been collected, but only a few of them have been included in previous studies of this species (Cibrián 1999, Schlüter et al. 2007). Furthermore, at least four additional species of the genus have been reported as growing actually or potentially sympatrically with V. planifolia in this region: Vanilla insignis, Vanilla odorata, Vanilla inodora Schiede, and Vanilla sp. nov. aff. V. phaeantha (Carnevali et al. 2001, Soto-Arenas 2009). This sympatry complicates sampling of wild individuals of V. planifolia whenever fertile structures are wanting.

During the last few decades, DNA markers have proven to be a powerful tool for accurate taxonomic determination (Nagy et al. 2012). The ISSR (Inter Simple Sequence Repeats; Zietkiewicz et al. 1994) is one of these markers. They are especially robust because allow discriminating between individuals of genetically close species or between varieties of the same species (Pharmawati et al. 2005, Elmeer & Almaki, 2011, Reyes-Alemán et al. 2013). Also, the ISSR, compared with other molecular tools, generate a high number of molecular markers with little effort; and is a relatively simple, cost-effective technique (Godwin et al. 1997, Bornet et al. 2002). Verma et al. (2009) used 10 ISSR primers to assess the genetic relationships of seven species and two hybrids of Vanilla: V. albida Blume, V. aphylla Eggers, V. andamanica Rolfe, V. planifolia, V. wightii Lindl. ex Wight, V. parishii Rchb. f., V. walkeriae Wight, V. x tahitensis JW. Moore and the hybrid V. planifolia ♀ ×V. wightii ♂. They found low levels of variation among individuals of V. planifolia, V. tahitensis, and V. aphylla, suggesting close relationships whereas there was also high affinity among V. planifolia and V. tahitensis, and between V. albida y V. aphylla. Vanilla andamanica was only distantly related to the other species, possibly reflecting a more distant phylogenetic relationship. Ramos-Castella et al. (2016), using five of the same ISSR primers of Verma in three species of Vanilla from México, found a clear differentiation between V. planifolia and the other three cultivated species from the wild (V. pompona, V. insignis, and V. aff. V. odorata).

On the other hand, the sequences of the ITS (Internal Transcribed Spacer) 1 and 2 of the 18S - 26S region of ribosomal DNA, usually display a high variability that allows for the broader molecular identification of organisms (Coleman 2003). ITS have been used in several molecular phylogenies of Vanilloideae (Cameron 2009, Poeaim et al. 2011, Soto-Arenas & Dressler 2010), providing a rich source of variable characters at several taxonomic levels, including species.

Wild populations of Vanilla planifolia are of utmost importance for the species’ conservation and its genetic improvement. As discussed earlier, the preservation of this species (and of related taxa) is currently hindered by the difficulties associated with identifying the species of Vanilla in the wild. Taking into account that ISSR molecular markers are relatively easy to use and powerful in genetic discrimination, they may provide accurate determinations of non-fertile, wild-collected individuals of Vanilla taxa, as well as of V. planifolia populations and varieties. Thus, the objectives of this study were the following:

  1. To assess the power of ISSR markers to discriminate between wild individuals of V. planifolia and those belonging to other species growing naturally in the Mexican Yucatan Peninsula.

  2. Contrast the results with those of a phylogenetic analysis of Vanilla ITS sequences.

Methods

Plant material. Leaves of 88 wild specimens of Vanilla spp. from the MYP were considered in this study (Table 1). These samples were deemed wild because they were collected in natural settings without evidence of human management. Hereafter, these samples will be referred to as MYP, followed by an identifying consecutive numbering plus a key to the region where they were collected (Table 1). Of all specimens studied, only two of them were found fertile and positively confirmed as V. planifolia (“MYP1-CY” and “MYP2-CY”, both from the same population) (Figure 1-A). The other specimens were infertile and only leaves were collected from them. However, we could observe morphological differences in leaf and stem in some of these samples. The detailed information about the collecting site of the samples is not provided for protection reasons. Sterile vouchers were collected by location and deposited at herbarium CICY.

Table 1 Geographical origin and description of the control samples and wild specimens of Vanilla collected in the MYP. 

Species Code Number of samples Origin Taxonomic characteristics
V. planifolia Soto-Arenas 2009
var. Mansa M 1 Puebla; MH
var. Variegata V 1 Puebla; MH
var. Oreja de Burro OB 1 Veracruz; MH
V. pompona 1 Puebla; MH
V. odorata 1 Veracruz; C-AR
V. insignis 2 Puebla; MH /JBR*; GC
V. sp. nov. aff. V. phaeantha 4 CY- JBR; GC No taxonomic
V. spp. MYP 88 CY, CQ, SQ, SC y SET; SV determination

Note: JBR: Jardín Botánico Regional Roger Orellana. MPY: Mexican Yucatan Peninsula. Regions: CY: Center of Yucatan; CQ: Center of Quintana Roo; SQ: Southern Quintana Roo; SC: South Campeche; and SET: Southeastern Tabasco. *: Unknown origin. MH: Mariana Hernández Collection. SV: Sara Villanueva Collection. GC: Germán Carnevali Collection. AR: Alma Ramos Collection.

Figure 1 Flowers and vegetative structures of the Vanilla species reported for the Yucatan Peninsula and used in the analysis with ISSR molecular markers. A) V. planifolia wild (both photographs by Sara Villanueva [SV]). B) V. insignis (flower photography by Leon Ibarra González, and leaves photography by SV). C) V. odorata (flower photography by Manfred Speckmaier, and leaves photography by SV). D) V. sp. nov. aff. V. phaeantha (both photographs by SV). E) V. pompona (both photos by Verónica Borbolla). 

In the ISSR analysis, we included 11 cultivated individuals as controls (standard samples). These individuals belong to the four species of Vanilla reported for the MYP and distributed as follows:

  1. Three individuals of Vanilla planifolia, originating from Puebla and Veracruz (“varieties” “Mansa” [M], “Variegata” [V], and “Oreja de Burro” [OB]).

  2. Two individuals of Vanilla insignis (one from Puebla [P], another from CICY’s Jardín Botánico Regional Roger Orellana [JBR-RO] – Figure 1-B).

  3. A single individual of Vanilla odorata from Veracruz (Figure 1-C).

  4. Four individuals of Vanilla sp. nov. aff. V. phaeantha grown at the JBR-RO (Figure 1-D). These have locality data and have been documented in bloom. This taxon still awaits formal description.

Also, a single individual of Vanilla pompona originally from the state of Puebla (Figure 1-E), a species that could potentially occur in the MYP.

DNA extraction and ISSR technique. Total genomic DNA was extracted from fresh leaves following a modification of the Dellaporta et al. (1983) procedure, which involved a single protein washout using phenol:chloroform:isoamylic. DNA was quantified with an ND-1000 spectrophotometer (Nanodrop, USA) and its quality assessed by 0.8 % agarose gel electrophoresis (Sambrook et al. 1989) and by PCR amplification of the 18S constitutive gene.

Three of the 10 ISSR primers specifically designed for Vanilla by Verma et al. (2009) were employed in this study: C07, C09 and T06. Amplifications were performed in a total volume of 20 μl, including 20 ng/μl of ADN, 200 mM of dNTPs (Invitrogen), 1.5 mM MgCl2, 60 pg. of primer, 2.5 μl buffer 10x of Taq ADN polymerase, and 1.5 U Taq DNA polymerase (Invitrogen). Amplification conditions were as follows: an initial step of 5 minutes at 94°C, followed by 35 cycles, each consisting of 30 seconds at 94°C for denaturing, 90 seconds at 50-54°C (depending upon primer used) for aligning and 90 seconds at 72°C for elongation, and 5 minutes at 72°C for final extension. Amplifications were performed at least twice and only reproducible products (bands) were considered for data analyses. Amplified PCR products were screened by agarose gel electrophoresis with 1.5% buffer TBE 0.5X (pH 8.3) and finally stained with ethidium bromide (Sambrook et al. 1989). Resulting gels were photographed and visually read with a 1 kb molecular marker as a reference for fragment length assessment.

Genetic relationships based on cluster analysis of the ISSR. Both monomorphic and polymorphic bands were considered for analyses and only well-defined bands (color intensity and reproducibility) included. The basic assumption was that bands of identical molecular weight regardless of whether belonging to individuals of the same or different population/species, represented the same allele. We compiled a database for each primer, recording the presence/absence of particular bands as 1or 0, respectively.

We performed a clustering analysis of both the standard samples as well as the MPY samples. Three different approaches followed for analyses:

  1. An assignment test of individuals using a Bayesian approach implemented in Structure 2.3.1 (Pritchard et al. 2000). Because individuals included in the analyses were referable to several species, the program was run under the no-admixture model. This model assumes that all of the genetic material of a particular individual comes from a single genetic pool (i.e. that no individual is hybrid). Based on this, we used the sampling location as LOCPRIOR in all the analyses performed (Tucker et al. 2012). The program ran under the following specifications: independent allelic frequencies, burn-in a period of 100,000, and 200,000 iterations after this period to allow the Markov chain to reach stationarity. Because there are at least four different Vanilla species reported for the MYP (see above), the analysis tested several values of K (K = 4 through K = 10). The most likely K value (K optimal) was estimated according to Evanno et al. (2005) implemented in the Structure-Harvester program (Earl & von Holdt 2012). Graphics output from Structure was edited in Microsoft PowerPoint 2013.

  2. A Neighbor-Joining analysis (Saitou & Nei 1987) to generate similarity clusters with Jaccard’s Similarity Index (Jaccard 1908). A frequency tree (Hampl et al. 2001) was constructed with the program FreeTree 0.9.1.50 (Pavlícek et al. 1999), with a bootstrap resampling of 1000 iterations. The output trees visualized in FigTree version 1.4.2 (Rambaut 2006-2014).

  3. A Principal Coordinates Analysis (PCoA), which allows for the recognition of spatial clustering of genotypes, without altering the data and only inputting the genetic similarity matrix. This analysis was conducted in GenAlEx 6.2 (Genetic Analysis in Excel) (Peakall & Smouse 2006). The two-dimensional PCoA graphic was constructed with the Gnuplot 3.7 freeware <http://www.gnuplot.info/>.

Phylogenetic inference based on ITS. With the goal of obtaining an independent corroboration of the clustering patterns emerging from the ISSR analyses, we performed a phylogenetic analysis of the nuclear ITS (ITS 1 and 2) region. This study included all the individuals considered in the ISSR analysis and relevant outgroups (see below), representing both tribes (Pogonieae and Vanilleae) of Subfamily Vanilloideae (Chase et al. 2015). Taxon sampling, consisting of 121 terminals, putatively representing 25 species, was devised to test the following three hypotheses:

  1. That Vanilla is monophyletic.

  2. That the Vanilla’s Afro-Caribbean Clade (Cameron 1999) is monophyletic and sister to the rest of the genus.

  3. That all individuals tentatively referred to V. planifolia (on morphological grounds or ISSR data) fall in a clade.

To polarize phylogenetically informative character states, a species of the distantly related genus Cleistes Rich. ex Lindl. (tribe Pogonieae), Cleistes paranaensis Schltr. was employed as a functional outgroup. This taxon was identified as basal in a recent phylogenetic analysis of the genus (Pansarin et al. 2008). The outgroup of the phylogenetic analysis consisted of the following taxa (Table 2):

Table 2 GenBank specimens used as outgroups in the analysis of ITS. 

Accession Species Reference
EU498152.1 C. paranaensis Pansarin et al. 2008
EU498153.1 C. pusilla
EU498158.1 C. uliginosa
FJ425836.1 E. lucidum Cameron 2009
FJ425837.1 E. subrepens
FJ425828.1 E. parviflorum
FJ425838.1 C. smilacifolium
FJ425835.1 V. barbellata
FJ425830.1 V. imperialis
FJ425834.1 V. africana
FJ425840.1 V. roscheri
GQ867246.1 V, planifolia - clone 19 Belanger & Havkin-Frenkel 2011
GQ867237.1 V. pompona – clone 8
AF151006.1 V. aphylla Cameron & Chase1999
JF825978.1 V. siamensis Li et al. 2011
JF796930.1 V. shenzhenica
AF391785.1 V. hirsuta Clements et al. 2002
C-GC V. grandiflora Panamá; Germán Carnevali Collection, 2016
C-GC V. dressleri
C-GC V. cribbiana

Note: C-GC: Germán Carnevali Collection.

  1. Two species of Cleistes (Tribe Pogonieae): C. uliginosa Pabst and C. pusilla Pansarin.

  2. Three species of Epistephium Kunth: E. lucidum Cogn., E. subrepens Hoehne, and E. parviflorum Lindl.

  3. The single species of Clematepistephium N. Hallé: C. smilacifolium (Rchb. f.) N. Hallé.

The ingroup of the analysis consisted of several different Vanilla species, representing both morphological variation and geographical distribution of the genus, although Neotropical taxa thought to be related to V. planifolia constitute the bulk of the sample. Species included are:

  1. Five members of the Afro-Caribbean clade of Vanilla: V. roscheri Rchb. f., V. africana Lindl., V. siamensis Rolfe ex Downie, V. imperialis Kraenzl. and V. barbellata Rchb. f.

  2. Two taxa occurring in Asia: V. shenzhenica Z.J. Liu & X. Qi Chen and V. aphylla.

  3. Nine Neotropical taxa occurring in the continental mainland: V. sp. (Venezuela), V. grandiflora Lindl. (Panama), V. dressleri Soto Arenas (Panama), V. cribbiana Soto Arenas (Panama), V. planifolia (Puebla-Veracruz; and Genbank’s clone 19), V. pompona (Puebla; Genbank’s clone 8), V. odorata, V. insignis, V. sp. nov. aff. V. phaeantha.

  4. The hybrid Vanilla × hirsuta M.A Clem. & D.L. Jones (V. odorata × V. planifolia).

  5. All the wild-collected Vanilla individuals also analyzed in the ISSR study.

The ITS region was amplified following Poeaim et al. (2011), and using primers ITS1 and ITS4 (White et al. 1990). Amplification reactions performed in 30 μl containing 50 ng/μl of ADN, 1.25 nm of dNTPs, 3 mM of MgCl2, 20 pmol of each primer, 1.5 U Taq polymerase, and the buffer provided by the manufacturer (Qiagen). The PCR protocol was as follows: an initial denaturalizing at 95°C for 5 minutes, followed by a sequence of 30 cycles consisting each of 1.5 minutes of denaturalizing at 94 °C, 2 minutes of annealing at 55 °C, and 1 minute of extension at 72 °C, ending in a 10 minute final extension period at 72 °C. PCR products amplified were screened and evaluated in an agarose gel stained with ethidium bromide, using a molecular weight marker of 100 pb. PCR products were later sent for sequencing at Macrogen (http://www.macrogen.com).

DNA sequences were edited with Sequencher 5.3 <http://www.genecodes.com>. Alignments were performed with Muscle (Edgar 2004) with the default parameters as available at the CIPRES Science Gateway (Miller et al. 2010). Afterward, the resulting alignment was visually refined in BioEdit v. 5.0.6 (Hall 2001). Average sequence length was 746 bp. To select the model of nucleotidic substitution that best fit the data, jModelTest2 (Darriba et al. 2012, Guindon & Gascuel 2003) was employed. The model selected was GTR + G.

Nucleotide sequence data of the ITS region used for phylogenetic inferences were analyzed under three different paradigms: Bayesian Inference, Maximum Likelihood, and Maximum Parsimony. The first two were implemented through the CIPRES Science Gateway (Miller et al. 2010). Conditions for each analysis were as described below:

  1. Bayesian inference (Yang & Rannala 1997): This analysis was conducted in MrBayes 3.2.6 (Ronquist et al. 2012). Two independent threads of four chains each were run with a MCMC chain length of 50,000,000 generations. Parameters and convergence of trees were assessed with Tracer v.1.6 (Rambaut et al. 2014).

  2. Maximum likelihood: Implemented in RAxML 8 (Stamatakis 2014). The support for nodes was assessed with 1,000 bootstrap iterations.

  3. Maximum Parsimony: The shortest trees were estimated with NONA (Goloboff, 1999) through the Winclada 1.00.08 shell (Nixon 1999-2000). 1,000 iterations of the Parsimony Ratchet (Nixon 1999, Vos 2003) algorithm were run. Characters were coded as unordered and of uniform weight (Fitch parsimony); gaps coded as missing characters. Support for clades was assessed with 1,000 iterations of a Jackknife analysis (Felsenstein 2004).

The primary objective of the three analyses was the qualitative assessment of the degree of congruence between the topologies yielded by them. Consensus trees were visualized in Figure 3 and edited with Paint of Microsoft Windows version 1,511 to include clade support values 80 % or higher of posterior probabilities, parametric bootstrap, and Jackknife.

Figure 2 Assignment test of 88 wild individuals of Vanilla spp. collected in the Mexican Yucatan Peninsula and 11 standard samples of Vanilla, using ISSR markers and the Structure program. A) K=5 based in the Evanno’s method. B) K = 7 (second K optimal). CY: Center of Yucatan; CQ: Center of Quintana Roo; SQ: Southern Quintana Roo; SC: South Campeche; and SET: Southeastern Tabasco. 

Figure 3 Neighbor-Joining dendrogram based on Jaccard similarity index (Jaccard 1908) of 88 wild individuals of the Mexican Yucatan Peninsula and 11 standard samples of Vanilla, using ISSR markers. The color in the branches is according to with the Figure 2A

Results

This section will present the results of the ISSR analyses first and then the phylogenetic ITS analysis, along with a description of their congruence or lack thereof.

Genetic relationships based on cluster analysis of the ISSR. 84 loci amplified with the three ISSR primers. Evanno’s method estimated a K = 5, which correlate with the following color-coded aggregations (Figure 2A):

  1. “YELLOW”: This group aggregated some of the standard samples of Vanilla planifolia (the “Mansa” and “Variegata” cultivars) along with 31 wild-collected specimens from the MYP (including the MYP1-CY y MYP2-CY specimens identified as V. planifolia using floral structures). These individuals share characteristics in leaf morphology (elliptic-obtuse in shape) and smooth stem-like V. planifolia. Within this group, the individuals MYP24-SQ and MYP80-CQ presented some degree of “green” ancestry (25 and 38 % respectively).

  2. “GREEN”: This group aggregated the standard sample of Vanilla planifolia var. “Oreja de Burro” along with 23 additional wild specimens from the MYP. This group has the same leaf morphological characteristics of the “yellow” group.

  3. “RED”: This color signaled the standard samples of Vanilla odorata, V. insignis, and V. pompona, along with 14 wild-collected individuals. Within this group differences in leaf (falcate-lanceolate, linear, and elliptic) and stem (rough, thick, smooth, thin) morphology were found.

  4. “BLUE”: This color category characterized the JBR-RO cultivated specimens of Vanilla sp. nov. aff. V. phaeantha (originally from central Yucatan) and an additional seven wild individuals. All member of this group share characteristics in leaf shape (linear) and stem surface (rough-thick) morphology.

  5. “FUCHSIA”: This group consisted of 13 wild specimens. These specimens have a morphology similar to individuals assigned to V. planifolia.

According to Meirmans (2015), considering different values of K may reflect different genetic and demographic processes and thus ensure a better biological interpretation of the data. In this study, we found that the second-highest K value was 7 (Figure 2B):

  1. “BLUE”: In this group, the yellow and fuchsia previous color categories (K = 5; Figure 2A) were aggregated, including the standard samples of the “Mansa” and “Variegata” cultivars of V. planifolia.

  2. “OCHRE”: This group is practically equal that former green group (K = 5; Figure 2A) including standard sample of variety “Oreja de Burro” of V. planifolia. Within this group appear the wild individuals MYP24-SQ and MYP80-CQ, which have a higher degree “ochre” ancestry (45% and 30% respectively).

  3. “AQUA”: This color signaled the standard sample of V. odorata with five wild-collected individuals. Three of these (MYP82-SC, MYP83-SC, and MYP84-SC) individuals present different degree of “green” and “yellow” ancestry (i.e. admixed individuals). All these individuals were present in the former red group (K = 5; Figure 2A).

  4. “YELLOW”: This color category is composed by V. insignis specimens and two wild-collected individuals. These individuals also were grouped in the former red category (K = 5; Figure 2A).

  5. “RED”: This group consisted of the standard sample of V. pompona and seven wild specimens. In the same way, these individuals are grouped into the former red category (K = 5; Figure 2A).

  6. “FUCHSIA”: This group is maintained as in the K = 5 (the old blue group).

  7. “GREEN”: An unknown wild genotype that composes part of the ancestry of three individuals, namely MYP82-SC, MYP83-SC, and MYP84-SC (i.e. admixed individuals). These individuals were assigned to the “aqua” group according to their high ancestry coefficient derived from this group.

The Neighbor-Joining analysis aggregated the genotypes analyzed in two (A and B) highly supported (100 % bootstrap) major clusters (Figure 3), within which several subgroups were retrieved; these are color coded according to what was observed in the Structure analysis (Figure 2A).

  1. The A cluster aggregated all the standard samples of Vanilla planifolia and 67 wild MYP specimens. Three subclusters were contained: A1 (13 individuals), A2 (13 individuals, including MYP2-CY, identified as V. planifolia), and A3 (41 wild individuals along with the standard samples of the three varieties of V. planifolia, and the MYP1-CY, also identified as V. planifolia using floral material).

  2. The B cluster consisted of the remaining standard samples and further wild MYP specimens. Two highly supported subclusters were included: B1 (86 % support by the bootstrap) this subcluster contained the standard samples of V. odorata, V. pompona, and V. insignis, along with 14 wild specimens. B2 (99 % bootstrap) included the JBR individuals of Vanilla sp. nov. aff. V. phaeantha and seven wild MYP samples.

The PCoA analysis identified four principal clusters (Figure 4). Also, color coded according to with the Figure 2A. These clusters were mostly “pure-colored”; except an assorted-color (yellow/fuchsia) cluster. The YELLOW/FUCHSIA group included the “Mansa” and “Variegata” cultivars of Vanilla planifolia and 44 wild-collected MYP specimens, the MYP1-CY and MYP2-CY among them. The GREEN group comprised the standard samples of V. planifolia var. “Oreja de Burro” and 23 additional wild MYP individuals. The BLUE group contained the standard samples of Vanilla sp. nov. aff. V. phaeantha and seven wild MYP individuals. Lastly, the RED group included V. pompona, V. odorata, and V. insignis with the remaining (14) wild individuals. These groups feature morphological coherence, except for the red group. The principal coordinates 1 and 2 explained 38.69 % and 12.77 % of total variation, respectively.

Figure 4 Principal Coordinates Analysis (PCoA) of 88 wild individuals of Vanilla of the Mexican Yucatan Peninsula and 11 standard samples of Vanilla, using ISSR markers. The color points are according to with the Figure 2A

In summary, the three clustering methods identify genetic relationships with relatively congruent patterns where there are at least four genetic groups and Vanilla planifolia isolated from the other species that occur wild in the MYP. It also identifies a genetic group that appears to correlate with morphologically distinctive plants and that will eventually be proposed as a new taxon (Vanilla sp. nov. aff. V. phaeantha).

Phylogenetic inference of Vanilla based on ITS. Including gaps, the matrix analyzed consisted of 1,018 sites with an average of 698 pb per taxon. We show here the Bayesian inference majority (80 %) consensus tree (Figure 5), including the bootstrap values from the ML analysis and jackknife values exceeding 80 from the parsimony analysis (589 informative sites [57.7 %], L: 3775, CI: 35, RI: 54). Basal nodes were fully resolved and strongly supported. However, there was not resolution within the American Vanilla taxa, due to collapse/lack of support for several clades.

Figure 5 Vanilliodeae phylogeny. Tree with the highest posterior probability resulting from the Bayesian phylogenetic inference. ML bootstrap values to the left of the diagonal; Parsimony jackknife values to the right of the diagonal; and posterior probability values above the branches. The phylogram in the lower left corner shows the length of the branches. Trees were obtained from multiple sequence alignments of nucleotides of Vanilla species using ITS. 

In our analyses, the most highly supported topology retrieves Vanilla as monophyletic (1.0:87/88) and sister to Clematepistephium (1.0), being both members of tribe Vanilleae, in congruence with previous analyses (e.g. Cameron 2009, Bouetard et al. 2010). The Afro-Caribbean clade, represented in this analysis by Vanilla aphylla, V. shenzhenica, V. imperialis, V. siamensis, V. africana, V. roscheri and V. barbellata, is strongly supported (1.0:97/100). This clade is sister to the remainder of the genus, composed in this study only of American species, hereafter referred to as the American Clade (AC). The Afro-Caribbean clade includes West Indian taxa as well as Paleotropical, mainly from Africa and tropical Asia, some of which are aphyllous.

The AC resolved as a trichotomy in our analyses, including Vanilla bahiana in an isolated position, along with a highly supported (p.p. 1.0) metaphyletic complex (i.e., monophyletic but lacking internal resolution) composed of V. grandiflora, V. pompona (Puebla and clone 8) and several MYP wild, sterile genotypes most likely attributable to V. pompona (1.0: 93/95). This clade is vegetatively characterized by succulent plants with large, massive leaves. It roughly corresponds with the “RED” group identified by the ISSR analysis. It is noteworthy that V. pompona has not been reported from the MYP, and if eventually confirmed by flowering material, it would constitute a first report for the area.

The third clade contains all the remaining species, represented by one or several samples, in the analysis. Within the third clade, two sister groups are readily identifiable. One of these clades is formed only by V. dressleri + V. cribbiana (DC-clade), composed of tropical rainforest, shade-loving species with thinly textured leaves and narrow stems. The DC-clade is sister to a poorly supported, metaphyletic group, consisting of four clades, one lacking support, whereas the other three have posterior probabilities equal or smaller than 0.85. Each of these clades roughly corresponds to additional four species of Vanilla known from the MYP. Similarly, these clades correspond with the color-coded groups recovered by the ISSR analyses. They are described below:

  • The first, or BLUE-clade, includes all the samples positively identified as Vanilla sp. nov. aff. V. phaeantha, along with six wild MYP individuals, which most likely represent the same species. This clade includes a sample of a Venezuelan species (Carnevali s.n, CICY), also closely related to V. phaeantha but differing in floral details.

  • RED-clade A: This highly supported (1.0:98/97) clade aggregates the standard sample of V. odorata, along with three wild MYP individuals (MYP76-CQ, MYP75-CQ, and MYP74-CQ). The analyses retrieve V. hirsuta as sister to the core of this clade. Vanilla hirsuta is thought to be a hybrid of V. odorata and V. planifolia. This relationship, however, is poorly supported.

  • RED-clade B. This clade contains the two standard samples of Vanilla insignis and receives an intermediate to moderately high level of support (0.85: --/95).

  • GREEN-YELLOW/FUCHSIA clade: This moderately supported (0.82) clade includes individuals positively identified or presumed to be Vanilla planifolia. Two subclades are retrieved, the first of which (green) includes the standard samples of V. planifolia var. “Oreja de Burro” along with 24 wild MYP individuals. Support is high at 0.96. The second subclade (yellow/fuchsia) contains the standard samples of the “Mansa” and “Variegata” cultivars of V. planifolia. It also includes 47 wild individuals, two of which are noteworthy (MYP1-CY, and MYP2-CY) because these have been seen in bloom and positively identified as V. planifolia.

Discussion

There is currently a void in our knowledge of the genus Vanilla. This vacuum ranges from basic biology including aspects of its systematics, phylogenetics, ecology, and physiology through more applied aspects such as the conservation status of its species (Gigant et al. 2011, Herrera-Cabrera et al. 2012, Azofeifa-Bolaños et al. 2014). At the most basic level, the identity, distribution, and circumscription of Vanilla taxa constitute the cornerstone to addressing problems related to the successful conservation and sustainable exploitation of the species of the genus.

Recently, it has become apparent that DNA molecular markers are powerful to help in the characterization of Vanilla species. These markers provide a tool from the population to the species-level identification of individual plants (Nagy et al. 2012). This characteristic is particularly useful in this genus where plants are most frequently collected sterile, are intrinsically variable and phenotypically plastic according to growing conditions, are difficult to the get established and flowered in cultivation, and many species are vegetatively very similar or practically identical.

In Vanilla, only 47 of about 110 species recognized in the genus, have been included in any molecular analysis, and most of them only in phylogenetic studies (Soto-Arenas 2003, Minoo et al. 2008, Soto-Arenas & Cribb 2010; Soto-Arenas & Dressler 2010). A few studies have included purported Vanilla species (e.g. V. tahitensis; Besse et al. 2004) that were later more correctly identified as hybrid taxa (e.g. Lubinsky et al. 2008b). In other cases, these hybrids had proposed as new species only to be later more correctly recognized as nothotaxa (e.g. V. hirsuta; Soto-Arenas & Dressler 2010).

Molecular characterization for Vanilla species is particularly relevant for V. planifolia, the most commonly cultivated taxon of the genus and upon which most of the world’s commercial production of vanilla based (Bory et al. 2008a; Lubinsky et al. 2008a). Furthermore, this characterization is very important because natural populations of this species have become severely threatened due to illegal extraction of wild individuals for introduction into plantations, both historically and recently (Soto-Arenas 2006). Moreover, destruction of natural habitats has recently become an even larger threat to the continued wild survival of V. planifolia. This study is the first to employ molecular markers to characterize a relatively large number of wild V. planifolia individuals, along with its sympatric congeners in an area of its natural distribution, the MYP.

Clustering pattern of Vanilla species in the MYP. The several approximations used in this study with ISSR markers retrieved largely congruent clustering patterns. Vanilla planifolia and Vanilla sp. nov. aff. V. phaeantha were clearly identified with the ISSR markers. The standard samples of V. pompona, V. odorata, and V. insignis were closely related to each other, probably because only a few individuals per species were tested. Three individuals of V. odorata identified by vegetative morphological features (mainly the narrow, falciform leaves), upon analyses, demonstrated to have a close genetic relationship with the group constituted by V. pompona - V. insignis - V. odorata. These individuals were included in the clade of V. odorata in the ITS phylogeny. The small number of wild-collected individuals in the MYP referable to these species suggest they are extremely rare in the area.

It is interesting to notice that the second highest delta K value obtained revealed a K = 7. These seven genetically different groups of individuals are ascribable to (and contain standard samples of) species previously known from the MYP (V. planifolia, V. insignis, V. odorata, and Vanilla sp. nov. aff. V. phaeantha). In the fifth group, a few wild MYP individuals clustered together with the V. pompona standard sample, and if morphologically confirmed, it would constitute the first record for this species that is mainly known from the western drainage of Megamexico. Vanilla planifolia var. “Oreja de Burro” was retrieved in the sixth group. And finally the seventh ancestral genotypic contribution (20–35 %, “green”, as a yet unidentified wild genotype) is present in three individuals (MYP82, MYP83, and MYP84) from southern Campeche. These individuals include in their genetic makeup about 50 % V. odorata and a smaller genetic contribution from V. insignis. However, in the PCoA, these individuals are not genetically close to V. odorata even though featuring an ancestry coefficient exceeding 50 %. Similarly, this K = 7 brings together two wild individuals to V. insignis (MYP86-SC y MYP87-SC), which are placed by the PCoA closest to V. pompona. Interestingly, although there is no clear evidence that V. pompona occurs natively in the MYP (Carnevali et al., 2001, Soto-Arenas 2009), seven MYP wild individuals were identified as this species by the K = 7 (Figure 2B). We identified these sterile individuals as V. planifolia relying upon their vegetative structures. These two species are vegetatively similar, especially when juvenile, but eventually, V. pompona develops larger, heavier, much thicker leaves than those commonly found in V. planifolia (Figure 1-A and 1-E). Floral features of course, readily diagnose both species.

Regarding Vanuilla planifolia, the finding of three subgroups within this species is probably attributable to the inclusion of a larger number of samples with a similar vegetative morphology, and therefore a greater probability of identifying an increased number of genotypes referable to the species and thus of detecting genetic or phylogenetic structure within the species. It is known that both self-compatible and self-incompatible individuals occur in V. planifolia. The last type, includes the entity commonly referred to as “Oreja de Burro” (Soto-Arenas 2009). Although the literature describes significant morphological features for the “Mansa” and the “Oreja de Burro” entities (Castillo & Engleman 1993), we were unable to identify wild MYP morphotypes unambiguously referable to them. However, the standard sample “Oreja de Burro” morphotype was retrieved widely apart from the standard samples of the “Mansa” and “Variegata” morphotypes in the Structure and the PCoA analyses, thus suggesting phylogenetic structure within V. planifolia.

A noteworthy finding from both the ISSR and the phylogenetic analyses was the precise identification of a group of individuals morphologically referred to Vanilla sp. nov. aff. V. phaeantha. This entity, albeit closely related to the West Indian V. phaeantha and initially identified as such (Soto-Arenas & Dressler 2010) it is easily differentiated by the presence of several prominent, deep yellow rows of teeth emerging from the labellum throat and reaching the labellum margin, which is also slightly fimbriated (Figure 1-D). This new taxon is in the process of being formally proposed.

Although the ISSR markers displayed high efficiency for the discrimination of individuals as belonging to clades or genetic groups associated with the standard samples of species involved in the study, some individuals were difficult to assign. A preliminary classification based on stem and leaf macro-morphology indicated that some wild individuals possibly belong to V. insignis but the genetic analysis indicates they are best referred (or genetically closer) to Vanilla sp. nov. aff. V. phaeantha, and secondly a cluster of eight individuals similar to V. planifolia but genetically related to the group comprised by V. pompona-V. odorata-V. insignis. These examples reveal how the comparison of individuals at different stages of development can be misleading in this genus, where plants are phenotypically plastic according to age and exposition, as previously pointed out by Putz &Holbrook (1986) and Ray (1990).

Phylogenetic relationships of the Vanilla species in the MYP. The most probable cladograms obtained in our phylogenetic inferences retrieved clades whose composition was entirely coherent with patterns of aggregation identified in the ISSR analysis. An example of this is the recovery of clades consisting of Vanilla sp. nov. aff. V. phaeantha and V. planifolia (Figs. 4 and 5). This finding reinforces the notion that the ISSR markers are robust as species-level discriminators of taxa, in this particular case in Vanilla.

Only a fraction of the species of the genus have been included in phylogenetic inferences, and there has not been an exhaustive phylogenetic analysis of Vanilla to date (Poeaim et al. 2011), the closest being the one by Soto-Arenas & Dressler (2010), which included only 15 species-level taxa. Furthermore, novelties within the genus are still being reported (e.g. V. paludosa Pansarin JM, Aguiar & A. C. Ferreira (Pansarin et al. 2012), V. rivasii Molineros, R. T. González, Flanagan & J. T. Otero (Molineros-Hurtado et al. 2014), V. labellopapillata A.K. Koch, Fraga, J.U. Santos & Ilk.-Borg. (Koch et al. 2013).

The resolution of the clades with the different methods evaluated agree with that reported by Cameron & Molina (2006), Pansarin et al. (2008), Cameron (2009), and Soto-Arenas & Dressler (2010). It is important to note that Vanilla hirsuta is synonymous with V. tahitensis J.W. Moore, which has a recent hybrid origin (V. planifolia x V. odorata; Soto-Arenas & Dressler, 2010). Therefore, its proximity to the individual genetic relationship of the standard sample of V. odorata and wild individuals identified as belonging to this species, it is not surprising. Moreover, V. bahiana described in Brazil, it appears to be very similar both morphologically and genetically with V. phaeantha (Soto-Arenas 2009). However, in each of the methods evaluated in this study, this species showed a closer relationship with V. pompona, which strengthens the hypothesis that Vanilla sp. nov. aff. V. phaeantha is a different species to V. phaeantha.

Except for V. planifolia “Oreja de Burro” (in the parsimony analysis), all the specimens referred by us to V. planifolia on morphological grounds constituted a monophyletic assemblage. Flowers of this entity have yet to be analyzed by us to confidently assess its relationships. Other individuals remained unidentified at this stage and placed at intermediate positions. The phylogenetic analysis retrieved a clade consisting mainly of specimens of Vanilla planifolia. However, nested within it there were three individuals assigned to the V. pompona-V. odorata-V. insignis group (red group in K = 5) by the STRUCTURE analysis of the ISSR data. This same analysis (but with K = 7) grouped two of these individuals with V. odorata (red group). The remaining genotype is possibly a complex hybrid containing in its ancestry V. odorata (50% aqua of its genetic makeup), V. sp. (35 % green), and V. insignis (15 % yellow), where the unknown taxon is likely to be a wild genotype of V. planifolia. From this perspective, individuals that were retrieved in intermediate or conflicting positions could be interpreted either as belonging to additional Vanilla species. Yet to be formally identified (with floral material) in our area, or interspecific hybrids.

Natural hybridization has long been known to be frequent in several groups of the Orchidaceae (e.g. Adams & Anderson 1958) including Vanilla (Lubinsky et al. 2006). Examples include introgression of V. claviculata (W. Wright) Sw. with V. barbellata Rchb.f in Puerto Rico (Nielsen & Siegismund 1999, Nielsen 2000). Also, AFLP and SSR markers have suggested the possibility of interspecific hybrids between Neotropical species of Vanilla such as V. bahiana, V. planifolia, or V. pompona (Bory 2007, Bory et al. 2008b). So, interspecific hybridization may be common whenever closely related taxa occur in sympatry (Bory et al. 2008). Nevertheless, not all morphological variation (or polymorphisms) within Vanilla species is to be attributed to interspecific hybridization. As in other species, this variation could be due to genetic drift, disruptive selection, or even directional selection resulting from an evolutionary response to local conditions (Ackerman et al. 2011). Also, there is a component of this variation associated with phenotypic plasticity arising from differences in life stages and different growth conditions (van Kleunen & Fischer 2005).

Implications for the conservation and sustainable use of Vanilla. Wild relatives of crops constitute an important source of genes associated with the resistance to pests and diseases or with higher yields under a variety of conditions (Hajjar & Hodgkin 2007, Hunter & Heywood 2011, Fisher, 2012). They are also relevant as a source of nutrients, medicine and other traditional uses for communities and for the maintenance ecosystemic services (Hoyt & Brown 1992). However, before these wild relatives can be utilized in a sustainable manner, it is first necessary to determine which species belong.

In concordance with Verma et al. (2009), our study showed that the ISSR can be used to identify Vanilla species present in the MYP. We found higher levels of genetic variation in wild samples than commonly found in vanilla plantations and that might be a result of sexual reproduction and generation from seed, as indicated by the presence of pods in some of the wild populations studied. This point is an indicative of the relevance of these reservoirs of potentially important alleles for the improvement of V. planifolia as a crop, and as such, they should be protected.

Thus, the MYP, where the species is still plentiful (yet occurring as small, widely dispersed populations) should be considered of utmost importance for the conservation of V. planifolia as a genetically healthy wild taxon, and as a source of alleles for its improvement as a crop.

Acknowledgments

The authors thank Biol. Veronica Limones Briones for their collaboration in lab work, and Ana Maria Hernández Hernández and Carlos Fernando Regla Márquez in the edition of graphics, and Verónica Borbolla, Manfred Speckmaier (Universität Wien, Core Facility Botanischer Garten) and Leon Ibarra González (Research of the Brigada de Educación para el Desarrollo Rural no. 3 (DGETA) and the Fundación Produce, A.C., Quintana Roo for the pictures of the vanillas. The first author thanks the National Council for Science and Technology (CONACYT) for the scholarship for postgraduate studies.

Literature cited

Ackerman JD, Morales M, Tremblay R. 2011. Darwin’s orchids: their variation, plasticity, and natural selection. Lankesteriana 11: 184-211. [ Links ]

Adams H, Anderson E. 1958. A conspectus of hybridization in the Orchidaceae. Evolution 12: 512–518. [ Links ]

Álvarez CL, Osorio NW, Marín-Montoya M. 2013. Identificación molecular de microorganismos asociados a la rizosfera de plantas de vainilla en Colombia. Acta Biológica Colombiana 18: 293–305. [ Links ]

Azofeifa-Bolaños JB, Paniagua-Vásquez A, García-García JA. 2014. Importancia y desafíos de la conservación de Vanilla spp. (Orchidaceae) en Costa Rica. Agronomía Mesoamericana 25: 189–202. DOI: 10.15517/am.v25i1.14220 [ Links ]

Belanger FC, Havkin-Frenkel D. 2011. Molecular analysis of a Vanilla hybrid cultivated in Costa Rica. In: Havkin-Frenkel D, Belanger FC. Eds. Handbook of Vanilla Science and Technology. Oxford: Wiley-Blackwell. 256–265. DOI: 10.1002/9781444329353.ch15 [ Links ]

Besse P, Da Silva D, Bory S, Grisoni M, Le Bellec F, Duval MF. 2004. RAPD genetic diversity in cultivated Vanilla: Vanilla planifolia, and relationships with V. tahitensis and V. pompona. Plant Science 167: 379–385. DOI: 10.1016/j.plantsci.2004.04.007 [ Links ]

Bornet B, Muller C, Paulus F, Branchard M. 2002. High informative nature of Inter-Simple Sequence Repeat (ISSR) sequences amplified with tri- and tetra-nucleotide primers from cauliflower (Brassica oleracea var. botrytis L.) DNA. Genome 45: 890–896. [ Links ]

Bory S. 2007. Diversité de Vanilla planifolia G. Jackson dans I ocean Indien et de ses especies apparentées; aspects genétiques, cytogénetiques et epigénétiques. PhD Thesis. Faculty of Science and Technology. University of Reunion. [ Links ]

Bory S, Grisoni M, Duval MF, Besse P. 2008a. Biodiversity and preservation of Vanilla: present state of knowledge. Genetic Resources and Crop Evolution 55: 551–571. DOI: 10.1007/s10722-007-9260-3 [ Links ]

Bory S, Lubinsky P, Risterucci AM, Noyer JL, Grisoni M, Duval MF, Besse P. 2008b. Patterns of introduction and diversification of Vanilla planifolia (Orchidaceae) in Reunion Island (Indian Ocean). American Journal of Botany 95: 805–815. DOI: 10.3732/ajb.2007332 [ Links ]

Bouetard A, Lefeuvre P, Gigant R, Bory S, Pignal M, Besse P, Grisoni M. 2010. Evidence of transoceanic dispersion of the genus Vanilla based on plastid DNA phylogenetic analysis. Molecular Phylogenetics and Evolution 55: 621–630. DOI: 10.1016/j.ympev.2010.01.021 [ Links ]

Cameron KM. 2009. On the value of nuclear and mitochondrial gene sequences for reconstructing the phylogeny of vanilloid orchids (Vanilloideae, Orchidaceae). Annals of Botany 104: 377–385. DOI: 10.1093/aob/mcp024 [ Links ]

Cameron KM. 2000. Gondwanan biogeography of Vanilloideae (Orchidaceae). Program and Abstracts. New Zealand: Southern Connections Congress. [ Links ]

Cameron KM, Chase MW. 1999. Phylogenetic Relationships of Pogoniinae (Vanilloideae, Orchidaceae): A Herbaceous Example of the Eastern North America-Eastern Asia Phytogeographic Disjunction. Journal of Plant Research 112: 317–329. DOI: 10.1007/PL00013873 [ Links ]

Cameron KM. 1999. Biogeography of Vanilloideae (Orchidaceae). XVI International Botanical Congress, Abstracts, p. 749. St. Louis, Missouri. [ Links ]

Cameron KM, Molina MC. 2006. Photosystem II gene sequences of psbB and psbC clarify the phylogenetic position of Vanilla (Vanilloideae, Orchidaceae). Cladistics 22: 239–248. DOI: 10.1111/j.1096-0031.2006.00102.x [ Links ]

Carnevali G, Tapia-Muñoz JL, Jiménez-Machorro R, Sánchez-Saldaña L, Ibarra-González L, Ramírez IM, Gómez MP. 2001. Notes on the flora of the Yucatan Peninsula II: A synopsis of the orchid flora of the Mexican Yucatan Peninsula and a tentative checklist of the Orchidaceae of the Yucatán Peninsula Biotic Province. Harvard Papers in Botany 5: 383–466. [ Links ]

Castillo MR, Engleman EM. 1993. Caracterización de dos tipos de Vanilla planifolia. Acta Botánica Mexicana 25: 49–59. [ Links ]

Chase MW, Cameron KM, Freudenstein JV, Pridgeon AM, Salazar G, Van den Berg C, Schuiteman A. 2015. An updated classification of Orchidaceae. Botanical Journal of the Linnean Society 177: 151–174. DOI: 10.1111/boj.12234 [ Links ]

Cibrián JA. 1999. Variación genética de Vanilla planifolia en México. Undergraduate Thesis. Facultad de Ciencias, Universidad Nacional Autónoma de México. [ Links ]

Cid-Pérez TS, López-Malo A. 2011. Extractos de vainilla: una mezcla de componentes químicos de aroma y sabor. Temas Selectos de Ingeniería de Alimentos 5: 51–63. [ Links ]

CIPRES Science Gateway. <http://www.phylo.org/portal2 > (March, 2016). [ Links ]

CITES. 2016. Convention on International Trade in Endangered Species of Wild Fauna and Flora. <http://www.cites.org > (accessed: November 10, 2012). [ Links ]

Clements MA, Jones DL, Sharma IK, Nightingale ME, Garratt MJ, Fitzgerald KJ, Mackenzie AM, Molloy BPJ. 2002. Phylogenetics of Diurideae (Orchidaceae) based on the internal transcribed spacer (ITS) regions of nuclear ribosomal DNA. Lindleyana 17: 135–171. [ Links ]

Coleman AW. 2003. ITS2 is a double-edged tool for eukaryote evolutionary comparisons. Trends Genet.19:370–375. DOI: 10.1016/S0168-9525(03)00118-5 [ Links ]

Darriba D, Taboada GL, Doallo R, Posada D. 2012. jModelTest 2: more models, new heuristics, and parallel computing. Nature Methods 9 (8): 772. DOI:10.1038/nmeth.2109 [ Links ]

Guindon S, Gascuel O. 2003. A simple, fast and accurate method to estimate large phylogenies by maximum-likelihood. Systematic Biology 52: 696–704. DOI: 10.1080/10635150390235520 [ Links ]

Dellaporta SL, Wood J, Hicks JB. 1983. A plant minipreparation: Version II. Plant Molecular Biology Reporter 1: 19–20. DOI:10.1007/BF02712670 [ Links ]

Edgar RC. 2004. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research 32: 1792–1797. DOI: 10.1093/nar/gkh340 [ Links ]

Earl DA, vonHoldt BM. 2012. Structure Harvester: a website and program for visualizing Structure output and implementing the Evanno method. Conservation Genetics Resources 4 (2): 359–361. DOI: 10.1007/s12686-011-9548-7 [ Links ]

Elmeer K, Almalki A. 2011. DNA fingerprinting of Prosopis cineraria and Prosopis juliflora using ISSR and RAPD techniques. American Journal of Plant Sciences 2: 527–534. DOI: 10.4236/ajps.2011.24062 [ Links ]

Evanno G, Regnaut S, Goudet J. 2005. Detecting the number of clusters of individuals using the software Structure: a simulation study. Molecular Ecology 14: 2611–2620. DOI: 10.1111/j.1365-294X.2005.02553.x [ Links ]

FAO, Food and Agriculture Organization of the United Nations. 1995. International Technical Conference on Plant Genetic Resources. Report of the Sub-Regional: Preparatory Meeting for Central America, Mexico, and the Caribbean. San José, Costa Rica. [ Links ]

Felsenstein J. 2004. Inferring phylogenies. Massachusetts. Sinauer Associates, Inc. [ Links ]

Fisher M. 2012. Crop wild relatives and their potential for crop improvement (Crops, Soils, Agronomy) CSA News. American Society of Agronomy. [ Links ]

Gigant R, Bory S, Grisoni M, Besse P. 2011. Biodiversity and evolution in the Vanilla genus. In: Grillo O, Venora G. eds. The dynamical processes of biodiversity: case studies of evolution and spatial distribution. Environmental Sciencies. DOI: 10.5772/24567 [ Links ]

Godwin ID, Aitken EA, Smith LW. 1997. Application of inter simple sequence repeat (ISSR) markers to plant genetics. Electrophoresis 18: 1524–1528. [ Links ]

Goloboff P. 1999. NONA ver. 2. Tucumán: Fundación e Instituto Miguel Lillo. [ Links ]

Hajjar R, Hodgkin T. 2007. The use of wild relatives in crop improvement: A survey of developments over the last 20 years. Euphytica 156: 1–13. DOI: 10.1007/s10681-007-9363-0 [ Links ]

Hall T. 2001. BioEdit: Biological sequence editor for Windows. Ibis Biosciences. California. <http://www.mbio.ncsu.edu/BioEdit/bioedit.html > (Accessed: February 10, 2016). [ Links ]

Hampl V, Pavlícek A, Flegr J. 2001. Construction and bootstrap analysis of DNA fingerprinting-based phylogenetic trees with the freeware program FreeTree: application to trichomonad parasites. International Journal of Systematic and Evolutionary Microbiology 51: 731–735. [ Links ]

Herrera-Cabrera BE, Salazar-Rojas VM, Delgado-Alvarado A, Campos-Contreras J, Cervantes-Vargas J. 2012. Use and conservation of Vanilla planifolia J. in the Totonacapan Region, México. European Journal of Environmental Sciences 2: 43–50. [ Links ]

Hoyt E, Brown S. 1992. Conservando los parientes silvestres de las plantas cultivadas. Delaware: Addison-Wesley Iberoamericana. [ Links ]

Hunter D, Heywood V. eds. 2011. Parientes silvestres de los cultivos: manual para la conservación in situ. Roma. Bioversity International. [ Links ]

Jaccard P. 1908. Nouvelles recherches sur la distribution florale. Bulletin de la Société vaudoise des sciences naturelles 44: 223–270. DOI: 10.5169/seals-268384 [ Links ]

Koch AK, Nicoletti de Fraga C, Ubiratan M, dos Santos J, Ilkiu-Borges AL. 2013. Taxonomic notes on Vanilla (Orchidaceae) in the brazilian amazon, and description of a new species. Systematic Botany 38 (4): 975-981. DOI: 10.1600/036364413X674706 [ Links ]

Li JH, Liu ZJ, Salazar GA, Bernhardt P, Perner H, Tomohisa Y, Jin XH, Chung SW, Luo YB. 2011. Molecular phylogeny of Cypripedium (Orchidaceae: Cypripedioideae) inferred from multiple nuclear and chloroplast regions. Molecular Phylogenetic and Evolution. 61: 308–320. DOI: 10.1016/j.ympev.2011.06.006 [ Links ]

Lubinsky P, Dam M van, Dam A van. 2006. Pollination of Vanilla and evolution in Orchidaceae. Lindleyana. 75: 926–929. [ Links ]

Lubinsky P, Bory S, Hernández-Hernández J, SeungChul K, Gómez-Pompa A. 2008a. Origins and dispersal of cultivated Vanilla (Vanilla planifolia Jacks. [Orchidaceae]). Economic Botany 62 :127–138. [ Links ]

Lubinsky P, Cameron MK, Molina MC, Wong M, Lepers-Andrzejewski S, Gómez-Pompa A, Kim SC. 2008b. Neotropical roots of a Polynesian spice: the hybrid origin of Tahitian vanilla, Vanilla tahitensis (Orchidaceae). American Journal of Botany 95: 1040–1047. DOI: 10.3732/ajb.0800067. [ Links ]

Macrogen. <http://www.macrogen.com > (February, 2016). [ Links ]

Meirmans PG. 2015. Seven common mistakes in population genetics and how to avoid them. Molecular Ecology 24: 3223–3231. DOI: 10.1111/mec.13243 [ Links ]

Menchaca-Garcia RA. 2010. Obtención y caracterización morfológica de híbridos de Vanilla planifolia G. Jackson in Andrews y V. pompona Schiede. PhD. Thesis. Center for Tropical Research, Universidad Veracruzana. [ Links ]

Miller MA, Pfeiffer W, Schwartz T. 2010. Creating the CIPRES Science Gateway for inference of large phylogenetic trees” in Proceedings 2011. TeraGrid Conference: Extreme Digital Discovery. Art. 41. Utah. DOI: 10.1145/2016741.2016785 [ Links ]

Minoo D, Jayakumar VN, Veena SS, Vimala J, Basha A, Saji KV, Nirmal Babu K, Peter KV. 2008. Genetic variations and interrelationships in Vanilla planifolia and few related species as expressed by RAPD polymorphism. Genetic Resources and Crop Evolution 55: 459–470. DOI: 10.1007/s10722-007-9252-3 [ Links ]

Molineros-Hurtado F, González Mina RT, Flanagan NS, Tupac-Otero J. 2014. Vanilla rivasii (Orchidaceae), a new species from the Colombian Pacific region. Lankesteriana 13 (3): 353—357. DOI: http://dx.doi.org/10.15517/lank.v13i3.14423 [ Links ]

Nagy S, Poczai P, Cernák I, Gorji AM, Hegedüs G, Taller J. 2012. PICcalc: an online program to calculate polymorphic information content for molecular genetic studies. Biochemical Genetics 50: 670–672. DOI: 10.1007/s10528-012-9509-1 [ Links ]

Nielsen RL. 2000. Natural hybridization between Vanilla claviculata (W.Wright) Sw. and V. barbellata Rchb.f. (Orchidaceae): genetic, morphological, and pollination experimental data. Botanical Journal of the Linnean Society. 133: 285–302. DOI: 10.1111/j.1095-8339.2000.tb01547.x [ Links ]

Nielsen RL, Siegismund HR. 1999. Interspecific differentiation and hybridization in Vanilla species (Orchidaceae). Heredity 83: 560–567. DOI: 10.1038/sj.hdy.6885880 [ Links ]

Nixon KC. 1999. The parsimony ratchet, a new method for rapid parsimony analysis. Cladistics 15: 407–414. DOI: 10.1111/j.1096-0031.1999.tb00277.x [ Links ]

Nixon KC. 1999-2002. WinClada ver. 1.0000 Ithaca: published by the author. [ Links ]

Pansarin ER, Salatino A, Salatino MLF. 2008. Phylogeny of South American Pogonieae (Orchidaceae, Vanilloideae) based on sequences of nuclear ribosomal (ITS) and chloroplast (psaB, rbcL, rps16, and trnL-F) DNA, with emphasis on Cleistes and discussion of biogeographic implications. Organisms Diversity and Evolution 8 (3): 171–181. DOI: 10.1016/j.ode.2007.09.003 [ Links ]

Pansarin ER, Aguilar JMRBV, Ferreira AWC. 2012. A new species of Vanilla (Orchidaceae: Vanilloideae) from São Paulo, Brazil. Brittonia 64: 157-161. DOI: 10.1007/s12228-011-9215-z [ Links ]

Pavlícek A, Hradá S, Flegr J. 1999. Free-Tree-freeware program for construction of phylogenetic trees on the basis of distance data and bootstrap/jackknife analysis of the tree robustness. Application in the RAPD analysis of genus Frenkelia. Folia Biológica 45: 97–99. [ Links ]

Peakall R, Smouse PE. 2006. GenAlEx 6. Genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Resources 6: 288–295. DOI: 10.1111/j.1471-8286.2005.01155.x [ Links ]

Pharmawati M. Yan G, Finnegan PM. 2005. Molecular variation and fingerprinting of Leucadendron cultivars (Proteaceae) by ISSR markers. Annals Botany 95: 1163–1170. DOI: 10.1093/aob/mci127 [ Links ]

Poeaim S, Poeaim A, Soytong K. 2011. Determination of the genetic relationship of Vanilloideae (Orchidaceae) based on the sequence of the Internal Transcribed Spacer of ribosomal DNA. Research Magazine: 7-11 < http://ird.rmutto.ac.th > (Accessed October 2, 2015). [ Links ]

Portères R. 1954. Le genre Vanilla et ses espèces. In:. Bouriquet G. Le vanillier et la vanille dans le monde. Paris: Lechevalier P. 94–290. [ Links ]

Pritchard JK, Stephens M, Donnelly P. 2000. Inference of population structure using multilocus genotype data. Genetics 155: 945–959. [ Links ]

Putz FE, Holbrook NM. 1986. Notes on the natural history of hemiepiphytes. Selbyana 9: 61–69. [ Links ]

Rambaut A. 2006–2014. FigTree, Tree Figure Drawing Tool Version v1.4.2. <http://tree.bio.ed.ac.uk/software/figtree/ > (accessed: October 2, 2015). [ Links ]

Rambaut A, Suchard MA, Xie D, Drummond AJ. 2014. Tracer v1.6. <http://beast.bio.ed.ac.uk/Tracer > (Accessed February 11, 2016). [ Links ]

Ramírez SR, Gravendeel B, Singer RB, Marshall CR, Pierce NE. 2007. Dating the Origin of the Orchidaceae from a Fossil Orchid with Its Pollinator. Nature 448: 1042–1045. DOI:10.1038/nature06039 [ Links ]

Ramos-Castella AL, Iglesias-Andreu LG, Martínez-Castillo J, Ortiz-García M, Andueza-Noh RH, Octavio-Aguilar P, Luna-Rodríguez M. 2016. Evaluation of molecular variability in germplasm of vanilla (Vanilla planifolia G. Jackson in Andrews) in Southeast Mexico: implications for genetic improvement and conservation. Plant Genetic Resources. DOI: 10.1017/S1479262115000660 [ Links ]

Ray TS. 1990. Metamorphosis in the Araceae. American Journal of Botany 77: 1599–1609. [ Links ]

Reyes-Alemán JC, Valadez-Moctezuma E, Simuta-Velázco L, Barrientos-Priego AF, Gallegos-Vázquez C. 2013. Distinción de especies del género Persea mediante RAPD e ISSR de ADN. Revista Mexicana de Ciencias Agrícolas 4 (4): 517–529. [ Links ]

Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Hohna S, Larget B, Liu L, Suchard MA, Hueslsenbeck JP. 2012. MrBayes 3.2: Efficient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Systematic Biology 61: 539–42. DOI: 10.1093/sysbio/sys029 [ Links ]

Saitou N, Nei M. 1987. Neighbor-Joining method: A new method for reconstructing phylogenetic trees. Molecular Biology and Evolution 4: 406–425. [ Links ]

Sambrook J, Fritschi EF, Maniatis T. 1989. Molecular cloning: a laboratory manual. New York: Cold Spring Harbor Laboratory Press. DOI: 10.1016/0092-8674(90)90210-6 [ Links ]

Schlüter PM. 2002. RAPD variation in Vanilla planifolia Jackson (Orchidaceae) and assessment of the putative hybrid Vanilla tahitensis Moore. MSc Thesis. University of Oxford. [ Links ]

Schlüter PM, Soto-Arenas MA, Harris SA. 2007. Genetic Variation in Vanilla planifolia (Orchidaceae). Economic Botany 61: 328–336. [ Links ]

SEMARNAT. 2010. Norma Oficial Mexicana de Protección especial de especies nativas de México de Flora y Fauna silvestres. Diario Oficial de la Federación. <http://dof.gob.mx/nota_detalle.php?codigo=5420810&fecha=21/12/2015 > (accessed: February 9, 2016). [ Links ]

Smith NJH, Williams JT, Plucknett DL, Talbot JP. 1992. Tropical Forests and Their Crops. Ithaca: Cornell University Press. [ Links ]

Soto-Arenas MA. 1999. Filogeografía y recursos genéticos de las vainillas de México. Instituto Chinoin AC. Reprote final SNIB-CONABIO proyecto No. J101. [ Links ]

Soto-Arenas MA. 2003. Vanilla (generic treatment). In: Pridgeon A.M., Cribb P.J., Chase M.W. and Rasmussen F.N. (eds.), Genera Orchidacearum volume 3, pp. 321-334, Orchidoideae (Part two), Vanilloideae. Oxford University Press. Oxford. [ Links ]

Soto-Arenas MA, Cribb PJ. 2010. A new infrageneric classification and synopsis of the genus Vanilla Plum. ex Mill. (Orchidaceae, Vanillinae). Lankesteriana 9: 355-398. [ Links ]

Soto-Arenas MA. 2006. La vainilla: retos y perspectivas de su cultivo. CONABIO Biodiversitas 66: 1–9. [ Links ]

Soto-Arenas MA. 2009. Recopilación y análisis de la información existente sobre las especies mexicanas del género Vanilla. Instituto Chinoín, A.C. In: “Generación y recopilación de información de las especies de las que México es centro de origen y diversidad genética”. SEMARNAT, CONABIO. [ Links ]

Soto-Arenas MA, Dressler RL. 2010. A revision of the Mexican and Central American species of Vanilla plumier ex Miller with a characterization of their it region of the nuclear ribosomal DNA. Lankesteriana 9: 285–354. [ Links ]

Soto-Arenas MA, Solano-Gómez AR. 2007. Ficha técnica de Vanilla planifolia. In: Soto-Arenas MA. (comp.). Información actualizada sobre las especies de orquídeas del PROY-NOM-059-ECOL-2000. Instituto Chinoin A.C., Databases SNIB-CONABIO. Project No. W029. México. [ Links ]

Stamatakis A. 2014. RAxML Version 8: A Tool for Phylogenetic Analysis and Post-Analysis of Large Phylogenies. Bioinformatics 30: 1312-1313. DOI: 10.1093/bioinformatics/btu033 <http://bioinformatics.oxfordjournals.org/content/early/2014/01/21/bioinformatics.btu033.abstract > (accessed: February 9, 2016). [ Links ]

Tucker KP, Hunter ME, Bonde RK, Austin JD, Clark AM, Beck CA, McGuire PM, Oli MK. 2012. Low genetic diversity and minimal population substructure in the endangered Florida manatee: implications for conservation. Journal of Mammalogy 93:1504–1511. DOI: 10.1644/12-MAMM-A-048.1 [ Links ]

van Kleunen M, Fischer M. 2005. Constraints on the evolution of adaptive phenotypic plasticity in plants. New Phytologist 166: 49–6. DOI: 10.1111/j.1469-8137.2004.01296.x [ Links ]

Verma PC, Chakrabaarty D, Jena SN, Mishra DM, Singh PK, Sawant SV, Tuli R. 2009. The extent of genetic diversity among Vanilla species: Comparative results for RAPD and ISSR. Industrial Crops and Products 29: 581-589. DOI: 10.1016/j.indcrop.2008.11.006 [ Links ]

Vos RA. 2003. Accelerated likelihood surface exploration: the likelihood ratchet. Systematic Biology 52: 368-373. [ Links ]

White TJ, Bruns T, Lee S, Taylor JW. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ. eds. PCR Protocols: A Guide to Methods and Applications. New York. Academic Press, Inc. 315-322. DOI: 10.1016/B978-0-12-372180-8.50060-3 [ Links ]

Yang Z & Rannala B. 1997. Bayesian phylogenetic inference using DNA sequences: a Markov chain Monte Carlo method. Molecular Biology and Evolution 14:717-724. [ Links ]

Zietkiewicz E, Rafalski A, Labuda D. 1994. Genome fingerprint by sequence repeat (SSR)-anchored polymerase chain reaction amplification. Genomics 20: 176-183. [ Links ]

1This research was conducted at the Marcadores moleculares laboratory of the Unidad de Recursos Naturales, and Herbario CICY at the Centro de Investigación Científica de Yucatán, A. C.

Received: January 28, 2016; Accepted: November 29, 2016

* Corresponding author: Jaime Martínez-Castillo, e-mail: jmartinez@cicy.mx

Autor Contributions. Sara Villanueva-Viramontes: collected most of the botanical material, conceived, designed, and performed the experiments, analyzed the data, and drafted the first version of the paper as well as reviewed drafts of the manuscript. Mariana Hernández-Apolinar: provided botanical materials and reviewed drafts of the manuscript. Germán Carnevali Fernández-Concha: provided botanical materials, conceived and designed parts of the experiments and analyses. Translated and reviewed drafts of the paper. Funded the ITS analyses. Alfredo Dorantes-Euán: Participated in the botanical collections, and reviewed drafts of the first version. Gabriel Rolando Dzib: Participated in the botanical collections, and reviewed drafts of the first version. Jaime Martínez-Castillo: conceived and designed parts of the experiments and reviewed drafts of the manuscript. Funded the ISSR analysis and the collecting trips.

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License