Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Rhizosphere 7 (2018) 49–56

Contents lists available at ScienceDirect

Rhizosphere
journal homepage: www.elsevier.com/locate/rhisph

Fungal diversity in the roots of four epiphytic orchids endemic to Southwest T


Mexico is related to the breadth of plant distribution
María de los Angeles Beltrán-Namboa, Miguel Martínez-Trujilloa, Juan Carlos Montero-Castrob,

Rafael Salgado-Garcigliac, Joel Tupac Otero-Ospinad, Yazmín Carreón-Abuda,
a
Genetic and Microbiology Laboratory, Biology Faculty, Universidad Michoacana de San Nicolás de Hidalgo, Francisco J. Múgica s/n, Col. Felicitas del Río, C.P. 58060
Morelia, Michoacán, Mexico
b
Plants´ Molecular Systematics Laboratory, Biology Faculty, Universidad Michoacana de San Nicolás de Hidalgo, Francisco J. Múgica s/n, Col. Felicitas del Río, C.P.
58060, Mexico
c
Vegetal Biotechnology Laboratory, Chemical-Biological Researchs Institute, Universidad Michoacana de San Nicolás de Hidalgo, Francisco J. Múgica s/n, Col. Felicitas del
Río, C.P. 58060, Mexico
d
Biological Sciences Department, Agricultural Sciences Faculty, Universidad Nacional de Colombia, Palmira sede, Colombia

A R T I C LE I N FO A B S T R A C T

Keywords: Fungi associated with plant roots are important for different plant processes, including germination and de-
Nonmycorrhizal fungi velopment. In Mexico, the knowledge about fungal species that interact with orchid roots and their possible roles
Mycorrhizal fungi in plant success and dispersion in different geographical locations is still very limited. For this reason, this work
Shannon-Wiener index aimed to determine the community composition and diversity of fungi associated with orchid roots at seven sites
Beta diversity
in the Transversal Volcanic Belt region in Michoacán, México. The roots of four endemic orchid species were
Fungal diversity
Orchid
analyzed: Laelia autumnalis, L. speciosa, Euchile citrina and P. squalida. In total, 71 isolates were obtained and
Mycorrhiza classified into 20 genera, including one mycorrhizal genus (Tulasnella) and 19 genera classified as basidiomy-
Endemic cetes and ascomycetes, such as Coprinus, Trichoderma and Xylaria. The diversity among different orchid species
and sampling sites was compared using the Chao 1, Shannon-Wiener and Whittaker indexes. In this work, species
such as L. autumnalis, L. speciosa and P. squalida were found in different sites and seemed to be more generalists
with regard to endophytes; they were typically associated with and found to be in symbiosis with species from
the orchid mycorrhizal genus Tulasnella. Other orchids seem to be more specific, such as E. citrina, which showed
the lowest diversity index and poor local distribution. This situation has been reported for species with a high
degree of endemism.

1. Introduction antagonistic relationships with other organisms to provide advantages


to plants and their potential as a source of bioactive compounds
A large number of fungi associated with the Orchidaceae family are (Bayman and Otero-Ospina, 2006; Khamchatra et al., 2016;
described in some reviews (Dearnaley et al., 2012; Liu, et al., 2010; Nontachaiyapoom et al., 2011).
Rasmussen and Rasmussen, 2007; Xiaoya et al., 2015). However, re- In other plant families, fungi from both groups (mycorrhizal and
search about fungal diversity in orchid roots is mainly focused on the nonmycorrhizal) present in the roots support water and nutrient ab-
fungi forming orchid mycorrhiza (OM), due to their primordial role in sorption, increase plant resistance to various stresses by releasing me-
seed germination and their contribution to plant development and tabolites, and stimulate germination and seedling development. For this
nutrition (Freudenstein and Chase, 2015; Rasmussen et al., 2015; reason, these organisms have a profound impact on the evolution,
Wright et al., 2011). Recently, it was reported that beneficial associa- ecology, health, structure and diversity of plant communities (Jiménez
tions with other endophytic fungi might also coexist (Herrera et al., et al., 2011; Mapperson et al., 2014; Otero-Ospina and Bayman, 2009;
2017), so interest in the nonmycorrhizal fungi of orchids has increased, Otero et al., 2011; Ye et al., 2014). Therefore, knowledge about the
due to their possible physiological functions, the synergistic or different fungi associated with orchid roots can provide a better


Corresponding author.
E-mail addresses: angelesb2008@gmail.com (M.d.l.A. Beltrán-Nambo), codigogenetico@gmail.com (M. Martínez-Trujillo),
cestrum2003@yahoo.com.mx (J.C. Montero-Castro), rafael.salgadogarciglia@gmail.com (R. Salgado-Garciglia), jtoteroo@unal.edu.com (J.T. Otero-Ospina),
ycabud@gmail.com, ycabud@umich.mx (Y. Carreón-Abud).

https://doi.org/10.1016/j.rhisph.2018.07.001
Received 31 May 2018; Received in revised form 22 July 2018; Accepted 23 July 2018
Available online 24 July 2018
2452-2198/ © 2018 Elsevier B.V. All rights reserved.
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

perspective on the interactions that occur in orchids’ natural habitats at risk category at this country (SEMARNAT, Secretaría del Medio
(Gamboa-Gaitán, 2006) and how these fungi can contribute to orchid Ambiente y Recursos Naturales, 2010). Because of this, it is important
survival and adaptation (Swarts and Dixon, 2009). to develop new strategies for orchid conservation. One of these con-
Some works have pointed out that the diversity of mycorrhizal and servation strategies is the use of fungal symbionts that could increase
nonmycorrhizal fungi in roots changes according to the orchid species germination rates and stimulate the orchids’ development and adaptive
and life cycle and that different patterns can be present. Some orchids capacity, because it was reported that some of the orchid distribution is
maintain an association with a single mycobiont, while others can ex- influenced by the disturbance degree of sites where they develop, the
pand their associations or can change symbionts during the transition species adaptability and their association with widely dispersed fungi,
from the juvenile to adult stage or during some environmental dis- among other factors (Beltrán-Nambo et al., 2012; Oliveira et al., 2013).
turbance (Rasmussen et al., 2015). For this reason, it is considered that In this country, more than 40% of orchid species are endemic, and
the study of endophytes that can stimulate the development of the their mycorrhizal and nonmycorrhizal fungi associations have been
embryo and promote plant growth in combination with mycorrhizal poorly studied, although some works have emphasized that the abun-
fungi or on their own is a topic that remains little explored, although it dance, distribution and heterogeneity of symbiont fungi are important
is very promising (Teixeira et al., 2015). for orchid adaptation and distribution on a local scale (McCormick
One of the central challenges facing ecology and conservation et al., 2018). Therefore, this research aimed to compare the diversity of
biology is the identification of factors influencing the spatial distribu- fungi associated with the roots of four epiphytic orchid species endemic
tion and abundance of rare or endangered plant species (Waud et al., to Mexico with those of populations threatened due to their ornamental
2014). Additionally, little is known about the physiological role and and artisanal use: Laelia autumnalis (Lex.) Lindl., L. speciosa (Kunth)
degree of association or specificity of fungi in the roots of epiphytic Schltr, Euchile citrina (Lex.) Withner and Prosthechea squalida (La Llave
orchids compared with those of terrestrial species. However, most y Lex.) Soto Arenas and Salazar. Here, we also analyze whether the
orchids are epiphytic and tropical or subtropical (Ávila-Díaz et al., presence of a high fungal diversity (generalist association) is related to
2013). At a global scale, it is mentioned that mycorrhizal fungi are a wider orchid distribution or if a more efficient association with fewer
found less consistently, but they seem to be strongly associated with fungi (specific association) is a factor that allows greater adaptive
epiphytic orchids compared to their terrestrial counterparts (Bailarote success for orchids.
et al., 2012; McCormick et al., 2018). Furthermore, many of the fungi in
these plants are not mycorrhizal endophytes, and these make up a
largely overlooked component of fungal biodiversity within roots (Yuan 2. Materials and methods
et al., 2009).
In Mexico, the depletion of large areas of pine, oak and mesophilic 2.1. Plants and sample sites
mountain forests have caused the loss of entire orchid populations,
whose major diversity is found in this kind of ecosystem (Gual-Díaz and Sampling was performed in seven sites located in different muni-
Rendón-Correa, 2014). A significant proportion of orchid species are cipalities of the physiographic province called the Transversal Volcanic
mainly found in the Transversal Neovolcanic Belt mountain chain, and Belt in the state of Michoacán (Fig. 1). Sites were selected based on
they are threatened due to the conversion of native forests to orchards distribution records of the four selected species in the state (Instituto de
for the cultivation of avocado and other plants (CONABIO and SUMA, Ecología AC-INECOL) and field trips to select sites where plants popu-
2005), causing approximately 15% of the orchid species to be reported lations were located. The reference data for the sampling sites and the
orchid species found in each one are shown in Table 1.

Fig. 1. Sampling site locations. The locations of the seven collection sites and the limits of the regions (municipalities) in which they are located are shown. (Dark
green = physiographical province; pale green = municipalities; black circles = sampled sites) (Map elaborate using QGIS program V. 9) (QGIS Development Team,
2009).

50
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

Table 1
Collection sites. Reference data for the different collection sites and orchid species analyzed in each of the sampling points. (Types of vegetation observed and
corroborated according to INEGI, 1985).
Collection sites Reference data (coordinates) Vegetation and altitude Orchid

# Municipality Nearby locality

1 001 Acuitzio Acuitzio N 19°29´36.35´´ Forest of pine- oak, cedar and oyamel. 2145 msnm. L. autumnalis
W 101°20´49.73´´

2 032 Erongaricuaro Oponguio N 19°38´42´´ Forest of oak 2114 msnm. L. autumnalis


W 101°39´43.9´´ L. speciosa

3 053 Morelia Cuanajillo N 19°38´34.18´´ Forest of oak 2200 msnm. L. speciosa


W 101°20´ 45.09´´

4 073 Quiroga Sta. Fe de la Laguna N 19°41´37.37´´ Forest of pine-oak and oak. 2281 msnm L. speciosa
W 101°33´45´´

5 100 Tzintzuntzan Cucuchucho N 19°36´18.4´´ Mixed forest of pine, oak and cedar, some nopal and copal. 2049 msnm L. autumnalis
W 101°37´18.62´´

6 016 Coeneo Carretera N 19°43´02.16´´ Forest of pine-oak. 2239 msnm L. autumnalis


W 101°37´45.53´´ P. squalida

7 111 Ziracuaretiro San Andrés Coru N 19°27´41.5´´ Mixed pine-oak forest, deciduous tropical forest ceiba, cedar, parota, tepeguaje. L. autumnalis
W 101°56´55.2´´ 1700 msnm. E citrina
P. squalida

The orchids Laelia speciosa and Euchile citrina are listed as species at mycorrhizal fungi, fungal coils and hyphae not forming pelotons were
risk by the Official Mexican Norm 059 (SEMARNAT, 2010) and as extracted by dissection, and the velamen and exodermis of selected root
priority species for conservation by the National Commission for fragments were removed under a stereoscopic microscope and laminar
Knowledge and Use of Biodiversity (CONABIO, 2012). In the case of L. flow hood (Rasmussen and Whigham, 2002). Sterilized distilled water
autumnalis and P. squalida, their current conservation status is un- drops containing coils or hyphae were dispersed in petri dishes con-
known. Orchids were located in different phorophytes or substrata, taining fungal isolation culture medium (FM) (Clements, 1988;
depending on the species, but mainly on oak (Quercus desertícola Trel., Mitchell, 1989). Then, samples were incubated at 26 °C under dark
Quercus sp. L.). In the case of L. autumnalis, in some areas, it was found conditions. For morphologic characterization and molecular determi-
growing on prickly pear (Opuntia ficus-indica Linnaeus Miller), cedar nation of the fungal isolates, hyphae segments were cut and transferred
(Callitropsis lusitanica (Mill.) D.P. Little) and ash (Fraxinus sp. Tourn. Ex to Petri dishes with potato dextrose agar (PDA) culture medium and
L.), together with P. squalida, and in other sites growing on rocks. incubated at 26 °C for two or three days until hyphal growth was ob-
Samples were collected during the flowering season in the years served.
2014–2016. Their identity was confirmed by INECOL specialists from
photographs and/or flower collection (Fig. A1a–f, Supplementary ma-
terial). The taxonomic names given for the orchid species described in 2.3. Morphological characterization
this work are based on the accepted nomenclature of the Kew Royal
Garden, reported in "The Plant List" (2013), which was consulted The color of the surface colony in culture media was determined
electronically. after 15 days using Munsell (2000) soil color charts. Furthermore, other
A 100 m2 quadrant was established at each collection site and di- macroscopic characteristics of fungal culture were analyzed: brightness,
vided into 20 × 33 m areas. A single tree with plants of any of the texture, odor and growth mode (Pereira et al., 2005). Additionally,
chosen species was selected (Fig. A1g, Supplementary material), for a growth rates were determined using the Currah et al., (1987) and were
total of 15 trees sampled per site (when this was possible; not all orchid represented in average values based on three replicates per strain. After
species were present in all the sites). Subsequently, three to five roots of 30 days, the presence of sclerotia, ascospores, basidiospores or mon-
each orchid species were randomly chosen from each tree for a total of ilioid cells were analyzed, including their dimensions (Shan et al.,
63 roots analyzed from L. autumnalis, 75 from L. speciosa, 30 from P. 2002). Fixed preparations on slides were stained with trypan blue and/
squalida and 10 from E. citrina. The roots were properly labeled and or acid fuchsin to measure the length and width of monilioid cells and
transported to the laboratory under refrigeration. spores. A Leica microscope with an integrated camera (Z1000) and the
AMScope V. 3.7 software was used for both measurements. Enzymatic
2.2. Isolation capacity tests were performed by reaction with polyphenyl oxidases in
culture medium with tannic acid as proposed by Davidson et al. (1938)
Roots were cut transversely using a scalpel to obtain 1 cm length and Zelmer (1994). Three petri dishes were inoculated with each strain
fragments. Then, a 1 mm transversal section from each fragment was obtained from the PDA isolation by placing a mycelium fragment of
mounted in polyvinyl-lactoglycerol alcohol and observed under an 1 mm3 in the dish and incubating it at 25 °C for 5–15 days. Subse-
optical microscope to detect the presence of hyphae or coils. For each quently, the reaction was observed; those that changed the growth
root, three to five colonized fragments were selected for fungal isola- medium coloration were considered positive.
tion. The isolates were grouped using a WARD hierarchical clustering
The selected fragments were superficially disinfected following the analysis based on Euclidean distances that were estimated from the
methodology of Ortega-Larrocea (2008) using commercial sodium hy- similarity of both qualitative and quantitative characteristics, including
pochlorite diluted at 10%, followed by an antibiotic solution (2% ery- their coefficient of cophenetic correlation. This procedure was per-
thromycin and 1% gentamicin), and rinsed three times with sterile formed using R version 3.3 software (R-Development Core Team,
distilled water. To ensure that isolates represented endophytic or 2008).

51
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

2.4. Molecular determination


Peroxidase reaction

Representative isolates of each group obtained from clustering


analysis were selected. From the isolates grown in PDA, a 1 mm3
Negative

Negative
Negative
fragment was cut and placed in potato-dextrose broth medium (PDB)
Positive

and incubated for 15–20 days at 25 °C with a 50 rpm agitation speed


and exposure to light. Fungi grown in PDB medium were liquefied in
Morphological characterization. Principal qualitative and quantitative characteristics of fungal strains obtained from epiphytic orchid roots and classified according to WARD clustering analysis.

150 ml of distilled water for 1 min, and aliquots of 1 ml were placed


No. of nuclei

into a dialysis membrane mounted in vacuum pump (manifold). The


2 or more
2 or more

species identification was made based on the internal transcribed


spacers (ITS) of ribosomal DNA. Total DNA was extracted using the
2

DNeasy Plant Mini Kit (Qiagen) from fresh tissue previously grown in
PDB medium and washed with a vacuum pump. For amplification, the
Growth average rates (mm) n = 268 (P≤ 0.05)

universal primers ITS1 and ITS4 (White et al., 1990) were used, fol-
lowing the protocol suggested by Swarts et al. (2010) but using a DNA
polymerase from the Qiagen company. Sequencing in both directions
(5´-3´ and 3´-5´) was performed at Macrogen (Seul, Korea).
All sequences were edited in Sequencher 5.2.4 (GeneCodes, Ann
Arbor, Michigan, USA), and the identification of similar sequences was
performed using the NCBI database (https://www.ncbi.nlm.nih.gov/)
with MegaBLAST option (Morgulis et al., 2008).
0.9 to 2.5 ± 1.4

1.2 to 2.5 ± 0.7

2.5. Diversity analysis


2 to 3 ± 1.2
3 or more

Fungal diversity was compared among orchid species and among


sampling sites. Alpha richness and diversity were determined using the
nonparametric method Chao 1 (Chao, 1984), Shannon-Wiener index
Spores dimensions (µm)

(Baev and Penev, 1995; Shannon and Weaver, 1949) and Pielou (1981)
species evenness index (J´), which refers to how similar in number each
species in an environment is. Additionally, the sampling effort was
determined by cumulative curves using EstimateS V.9.0 software
8.4–13.9

8.3–13.1

(Colwell, 2013). General beta diversity, which indicates the magnitude


8–12.2

10–15

of change in fungal species composition among all studied orchid spe-


cies and among the seven sites was calculated using the Whittaker
Basidiospore

(1972), with PAST V. 3.17 software (Hammer et al., 2001). Further-


Spore type

Ascospore
Ascospore
Monilioid

more, beta diversity was determined for each pair of orchid species and
each pair of sampling sites by using the same index.

3. Results
Waxy flat to plush, irregular growth
Plush to cottony, radial growth

3.1. Isolation and characterization of fungal isolates

A total of 268 fungal samples were isolated from 83 of the 168


processed roots (49.4%) of four orchid species analyzed. Based on
Plush to cottony
Plush to cottony

morphological characterization, the constructed distance tree generated


Colony aspect

four main groups of isolates (A–D) (Fig. A2 Supplementary material).


Group A included isolates with characteristics similar to those of
orchid mycorrhizal fungi, such as 90° branching angles, a slight con-
striction at the branching point, the absence of sporulation, and mon-
ilioid cell formation, among other features (Table 2). Groups B and C
Gray to olive or sepia (3/1 2.5Y, 6/4 10YR)

included isolates with ascospore formation, and group D included those


Colony color (Munsell catalogue, 2000)

isolates with basidiospore and fruiting body formation in some cases as


White to gray (8/1 2.5Y, 3/1 2.5Y)

its main characteristics (Table 2).


Pale yellow (8/2 to 8/3 5Y)

3.2. Molecular determination


Pale yellow (8/2 5Y)

From 268 isolates obtained, 100 were selected for molecular iden-
tification. All isolates from groups A and D (6 and 12, respectively)
were included; 40 isolates from group C were selected according to
colony similarity, and the remaining samples were taken from group B.
Identities were determined for 71 morphotypes (6 from group A, 31
from group B, 30 from group C and 4 from group D) (Table A1,
Supplementary material); 32 morphospecies were classified, with 18
Subgroups

showing identity at the species level in the GenBank database, ten at


Table 2

genus level and four at the order level. The strains were grouped
D
A

C
B

into three genera of Basidiomycota, including one recognized as an

52
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

Table 3
Fungal genera by site and orchid species. Distribution of fungal genera identified in axenic cultures from the species of orchids studied by sampling site and number of
fungal strains obtained from each.
Endophyte Host

L. autumnalis (sites) L. speciosa (sites) P. squalida (sites) E. citrine (sites)

Division Order Genera 1 2 5 6 7 2 3 4 6 7 7 Total

Basidiomycota (3 genera) Agaricales Clitopilus giovanellae 2 2


Coprinus sp. 10 10
Tulasnellales Tulasnella calospora 1 1 1 3

Ascomycota (17 genera) Eurotiales Paecilomyces inflatus 1 1


Glomerellales Colletotrichum sp. 1 1
Helotiales Lachnum sp. 1 3 4
Neofabraea actinidiae 1 1
Neofabraea sp. 1 4 5
Ascomycete sp. 4 3 8 1 1 2 19
Hypocreales Trichoderma atroviride 10 10
Trichoderma rossicum 39 39
Trichoderma viride 2 8 13 4 27
Trichoderma viridialbum 2 1 3
N/D 3 6 3 12
Pleosporales Alternaria alternata 6 6
Preussia minima 5 3 6 14
Preussia cymatomera 2 2
Sordariales Chaetomium nigricolor 6 1 7
Chaetomium sp. 3 2 5
Fusarium tricinctum 1 2 3
Fusarium sp. 4 4
Trichocladium opacum 1 1
Trichocladium sp. 2 2
Ascomycete (uncultured) 2 3 1 2 2 2 12
Xylariales Xylaria sp. 4 2 12 7 3 3 31
Ascovirgaria occulta 16 6 22
Muscodor albus 7 1 8
Nemania sp. 3 4 7
Virgaria nigra 3 3
Xylaria enteroleuca 2 2
Not defined Spegazzinia sp. 1 1

Uncultured N/D 1 1
Morphospecies/site (Total) 5 12 14 69 30 8 71 31 10 13 5

Total of strains 130 110 23 5 268


Total genera 15 11 4 1
Total morphospecies 22 18 7 2

orchid mycorrhizal genus (Tulasnella) and 17 genera of Ascomycota highest richness and diversity of fungi in their roots were those asso-
(Table 3). ciated with the two species of Laelia. However, the evenness index (J´)
indicated the existence of dominant fungal species in some orchid
3.3. Diversity of fungal isolates species (Fig. 2a).
With the analysis of fungal alpha diversity by sampling site, it was
The results for alpha diversity showed that orchid species with the observed that the highest richness and diversity of fungal species was

Fig. 2. Alpha diversity and shared species. A) Graph of richness and fungal diversity among orchid species, b) Graph of richness and fungal diversity among sampling
sites, c) Venn diagram that shows the number of fungal species shared among orchids. (U = Shannon-W Evenness index).

53
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

present in site 7 (San Andrés Coru), where three of the orchids analyzed comprised of 18 accepted species (Teik-Khiang and Hyde, 1999), most
were found (L. autumnalis, P. squalida and E. citrina) (Fig. 2b). The of them saprotrophs of wood, bark or other plant tissues. Although this
lowest species uniformity according to the evenness index occurred in genus is not considered mycorrhizal, many species in this genus un-
site 3 (Cuanajillo). dergo asexual reproduction, and it appears to be beneficial in at least
The general beta diversity for the sampled orchids was 2, according two Trichocladium species (T. lignicola I. Schmidt and T. opacum (Corda)
to the calculated Whittaker index (this parameter does not have a S. Hughes). It has been observed to have monilioid-like cell formation
maximum value). When the beta diversity between pairs of orchid (Teik-Khiang and Hyde, 1999) and for this reason, the cladistic analysis
species was analyzed, it was observed that orchids of the same genus or included it in group A. In America, this genus has been found in plants
those that shared a substrate exhibited the lowest index values (Table such as Fraxinus, Quercus, Pseudotsuga and Castanea (Seidl, 2009),
A2, Supplementary material) and share fungal species, as L. autumnalis which are like the phorophytes in this work. Due to the limited number
with L. speciosa, as is demonstrated in Fig. 2c. of studies of endophytes associated with orchids, the role that many of
When analysis by sampling site was performed, the global beta di- them play is unknown (Rodríguez et al., 2009).
versity was 3.4. The pair of sites analyzed with the lowest index values In the present work, most of the fungal isolates were nonmycor-
of fungal species turnover was site 2 and sites 1 and 6 (Table A3, rhizal. In the Ascomycota phylum, diverse groups were found, in-
Supplementary material). L. autumnalis was present at all three sites. cluding the pathogenic species Fusarium, Alternaria, Colletotrichum and
The species accumulation curves indicated that in the case of L. Neofabraea; the beneficial or synergistic endophytes Trichoderma,
speciosa, a better sampling effort is required since the number of fungal Preussia, Chaetomium, and Paecilomyces; and the saprobes Lachnum and
species would increase if collection sites were expanded (Fig. A3a, Xylaria. Among the endophyte genera, the most widely reported species
Supplementary material). Finally, the curve for sampling sites indicated in orchids are Xylaria, Trichoderma and Colletotrichum (Almanza-Álvarez
that the sampling effort was satisfactory since the generated curve et al., 2017; Ávila-Díaz et al., 2013; Gamboa-Gaitán and Otero-Ospina,
is asymptotic, which indicates that the fungal species number 2016). In this study, the dominant genera in both the number of species
will remain constant even with a greater number of samples (Fig. A3b, and number of isolated strains were Trichoderma and Xylaria, mostly
Supplementary material). isolated from Laelia speciosa and L. autumnalis. These genera are con-
sidered to stimulate plant germination and development by releasing
4. Discussion growth factors or metabolites that stimulate auxin production or act as
antagonists for other organisms that could damage these plant species
4.1. Fungal isolation and orchid rarity (Rivera-Orduña et al., 2011). It has been suggested that replacement of
the Rhizoctonia complex for ectomycorrhizal fungi or other endophytes
Because the identification was performed in axenic cultures of fungi could be an orchid strategy because these fungi are more stable and
from orchid roots, it is possible that fungal species were underestimated high-throughput carbon/nutrient resources (Stark et al., 2009).
compared to a metagenomic approach that includes noncultivable Group D included 2 genera of the Basidiomycota phylum, Coprinus
species. The highest number of fungal isolates obtained in this work and Clitopillus. Some studies have pointed out that some orchids can be
were extracted from Laelia autumnalis and L. speciosa. The first orchid mycorrhized by fungi not included in the Rhizoctonia complex, such as
species was present in five of seven sample sites with large populations, Coprinus and Russula (Kikuchi et al., 2008; Osgura-Tsujita et al., 2009);
contrasting with the E citrina species that was only located in site 7 (San Coprinus have been characterized as a mycorrhizal species of orchids
Andrés Coru). Although the number of samples of each orchid was an such as Epipogium roseum (Yamato et al., 2005). Previous reports in-
important factor in the number of fungi obtained (Kassem and dicate that endophytic fungi act as a communication bridge between
Nanniperi, 1995), other factors have been reported to influence fungal several plants and fulfil different functions, such as promoting orchid
isolation, such as special requirements preventing their development in seed germination (Dearnaley et al., 2012; Marmeisse and Girlanda,
standard culture media (Garden and Whitbeck, 2007) and sensitivity to 2016; Rasmussen et al. 2015).
environmental changes (Gadd et al., 2007). Plant species that are cri-
tically dependent on these symbionts are more selective and sensitive 4.3. Fungal diversity and orchid distribution
than others during symbiosis establishment, so symbiont sensitivity can
be expected to have a major impact on successful colonization and The Shannon-Wiener diversity indexes between orchids and sites
establishment, and thus ultimately on species rarity (Bonnardeaux obtained in this study were similar or slightly lower than those reported
et al., 2007; Otero et al., 2013; Roche et al., 2010; Swarts et al., 2010; in other studies for terrestrial and epiphytic orchids, where values from
Waud et al., 2017). E. citrina is an at-risk species that has experienced 1.3 to 3.5 are reported (Stark et al., 2009; Tao et al., 2008). According
habitat deterioration, and only two different fungi morphospecies were to beta diversity results between orchids and sites, it was observed that
isolated from its roots. the fungal species diversity in sampling sites was related to the orchid
richness and the presence of certain orchid species. Species of Laelia are
4.2. Orchid fungal identity and their reported ecological role evidently capable of maintaining diverse fungi in their roots, and this
ability could be a significant asset as it increases the likelihood of
Only one genus of mycorrhizal fungi was obtained, identified as forming a partnership with whichever fungus is present or obtaining
Tulasnella calospora, which was isolated from three of the orchid species some benefit at a given site and thus the likelihood of seeds starting to
analyzed. This genus is widely reported, together with Ceratobasidium germinate; it may also enhance dispersion processes (Waud et al.,
and Serendipitia (Jacquemyn et al., 2017), as one of the principal my- 2017). L. autumnalis not only showed the highest fungal community
corrhizal fungi of different orchids in the subtribe Laeliinae (which values of richness and diversity but also was the species with the
includes the species in this work), and its role in germinative and de- highest territorial coverage. This confirms what has been established
velopment processes in these plants is well known (Ding et al., 2014; for some orchid species, which is that their adaptive talent to disperse
Linde et al., 2014; Pereira et al., 2014; Phillips et al., 2014). In Mexico and colonize new environments is related to their ability to interact
and Central and South America, Tulasnella is reported to be mainly with a large variety of fungi to gain advantages in the substrate where
associated with some rupicolous and epiphytic orchids such as Epiden- they develop and probably with seasonal variation (Ercole et al., 2015;
drum, Coppensia, Vanilla, Acineta and Oncidium spp. (Jacinto-Hernández Waud et al., 2016). Additionally, the plasticity in fungal partners en-
and Ortega-Larrocea, 2013; Moreno-Martínez, 2011; Nogueira et al., sures their survival during different stages of their life cycle or under
2014; Rendón-Lara et al., 2013; Segundo, 2016). Group A included different environmental conditions (Rasmussen et al., 2015). The lower
Trichocladium Harz, in addition to mycorrhizal fungi. This genus is values showed by E. citrina could be explained by the fact that this

54
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

species was only found in one site, as their natural populations have encounters, lasting relationships and alien’s invasions. Mycol. Res. 111, 51–61.
been diminished due to habitat deterioration, and that only two dif- Chao, A., 1984. Nonparametric estimation of the number of classes in a population.
Scand. J. Stat. 11, 265–270.
ferent fungi were obtained in culture from its roots. Some researchers Clements, M.A., 1988. Orchid mycorrhizal associations. Lindleyana 3, 73–86.
have found that orchids were rarely limited by the availability of ap- Colwell, R.K., 2013. Statistical estimation of species richness and shared species for
propriate fungi at a geographic scale. The analyzed orchids are asso- samples. Version 9. Persistent URL 〈purl.ocic.org/estimates〉, consulted January 12,
2018.
ciated locally with many different fungal partners along their species CONABIO, SUMA, 2005. La biodiversidad en Michoacán. In: Villaseñor-Gómez, L. (Ed.),
distribution range, which allows a local species to remain an ecological Estudio de Estado, 1ª edición. CONABIO, SUMA, UMSNH, Morelia, Mich, pp. 258.
generalist. In such pattern, the association with rare fungi could also CONABIO, 2012. Portal de Geoinformación. Sistema Nacional de Información sobre
Biodiversidad. Cartografía digital No. comercial 2.5. Acceso en línea: 〈http://www.
cause an orchid species to be more susceptible to decline, even at a conabio.gob.mx/informacion/gis/〉, consulted December 13, 2017.
small scale (Bailarote et al., 2012; McCormick and Jacquemyn, 2014). Currah, R.S., Sigler, L., Hambleton, S., 1987. New records and new taxa of fungi from the
Our results also suggest the existence of segregation processes, be- mycorrhizae of terrestrial orchids of Alberta. Can. J. Bot. 65, 2473–2482.
Dearnaley, J.D., Artos, F.M., Elosse, M.S., 2012. Orchid mycorrhizas: molecular ecology,
cause although different orchid species were found to be associated
physiology, evolution and conservation aspects. In: Hock, B. (Ed.), Fungal
with the same fungus within a genus, in sites where these plants co- Associations. The Mycota IX, 2ª edition. Springer-Verlag, Berlin Heidelberg,
incided, they were associated with different fungal species. One ex- Queensland, Toowoomba, pp. 207–224.
ample is the association of Laelia autumnalis and L. speciosa with Davidson, R.W., Cambell, A., Blaisedel, D.J., 1938. Differentiation of wood-decaying
fungi by their reactions on gallic or tannic acid medium. J. Agric. Res. 57, 683–695.
Trichoderma viride in different locations. It has been determined that Ding, R., Chen, X.H., Zhang, L.J., Yu, X.D., Qu, D., Duan, R., Xu, Y.F., 2014. Identity and
segregation processes exist between orchids and are mediated by their specificity of Rhizoctonia-like fungi from different populations of Liparis japonica
interactions with different groups of symbionts. This ensures that nu- (Orchidaceae) in northeast China. PlosONE 9 (8), e105573. https://doi.org/10.1371/
journal.pone.0105573.
trients are obtained from different fungi, avoiding competition between Ercole, E., Adamo, M., Rodda, M., Gebauer, G., Giralda, M., Perotto, S., 2015. Temporal
orchid species and allowing their coexistence and dispersion (Ercole variation in mycorrhizal diversity and carbon and nitrogen isotope abundance in the
et al., 2015; Jacquemyn et al., 2014; Tesitelova et al., 2013; Waud et al., wintergreen meadow orchid Anacamptis morio. New Phytol. 205, 1308–1319.
Freudenstein, J.V., Chase, M.W., 2015. Phylogenetic relationships in Epidendroideae
2016). (Orchidaceae), one of the great flowering plant radiations: progressive specialization
and diversification. Ann. Bot. 115 (4), 665–681. https://doi.org/10.1093/aob/
5. Conclusions mcu253.
Gadd, G., Watkinson, S., Dyler, P., 2007. Fungi in the Environment, 1a edition. Cambridge
University Press, Cambridge, pp. 379.
The studied orchid species, with local generalist associations and Gamboa-Gaitán, M., 2006. Hongos Endófitos Tropicales: conocimiento actual y per-
diverse fungi, managed to develop in different habitats successfully. spectivas. Acta Biol. Colomb. 11 (1), 3–20.
Gamboa-Gaitán, M.A., Otero-Ospina, J.T., 2016. Colombian vanilla and its microbiota. III.
Species such as L. autumnalis, L. speciosa and P. squalida seem to be more
Diversity and structure of the endophytic community. Acta Bot. Hung. 58 (3–4),
generalists with respect to endophytes; they are associated with and 241–256. https://doi.org/10.1556/ABot.58.2016.3-4.2.
were found in symbiosis with the orchid mycorrhizal genera Tulasnella. Garden, Z., Whitbeck, J., 2007. The Rhizosphere. An Ecological Perspective, 1a. edition.
Although they are endemic species of Mexico, they are found and Elsevier Press, New Orleans, U.S.A., pp. 201 (ISBN-13: 978-0-12-088775-0).
Gual-Díaz, M., Rendón-Correa A., 2014. Bosques Mesófilos de Montaña de México. In:
widely distributed in the central and southern zones of the country, CONABIO (Ed.), Diversidad, ecología y manejo. 1a Edición. Comisión Nacional para
while other orchids, such as E. citrina, seem to be more specific. This el Conocimiento y Uso de la Biodiversidad, México, D.F. Retrieved from 〈http://
situation has been reported for species with a high degree of endemism. www.biodiversidad.gob.mx/ecosistemas/pdf/BosquesMesofilos_montana_baja.pdf〉.
Hammer, Ø., Harper, D.A.T., Ryan, P.D., 2001. PAST: paleontological statistics software
package for education and data analysis. Palaeontol. Electron. 4 (1), 9.
Acknowledgments Herrera, H., Valadares, R., Contreras, D., Bashan, Y., 2017. Mycorrhizal compatibility and
symbiotic seed germination of orchids from the coastal range and Andes in South
central Chile. Mycorrhiza 27, 175–188.
This work was financed by the CECTI through Project no. 05. The INEGI, 1985. Instituto Nacional de Estadística Geografía e Informática: Síntesis
authors thank Dr. Pilar Ortega Larrocea for transmission of knowledge Geográfica del Estado de Michoacán, Iztacalco, México, D.F., pp. 315.
and training in techniques for the study of orchid mycorrhiza, which Jacinto-Hernández, A., Ortega-Larrocea M.P., 2013. Germinación simbiótica de Oncidium
affn. sphacelatum. In: CONABIO-Viccon-Ávila-Nava (Ed.), Memorias del Segundo
made the performance of this work possible; and to M.C. Aarón
encuentro mexicano de orquideología. Chilpancingo, Guerrero, p. 38.
Giovanni Munguía for helping to improve the translation of this work. Jacquemyn, H., Waud, M., Brys, R., Lallemand, F., Courty, P., Robionek, A., Selosse, M.,
2017. Mycorrhizal associations and trophic modes in coexisting orchids: an ecological
contiuum between auto and mixotrophy. Frontiers Plant Sci. 8, 1497. https://doi.
Appendix A. Supplementary material
org/10.3389/fpis.2017.01497.
Jacquemyn, H., Brys, R., Merckx, V.S., Waud, M., Lievens, B., Wiegand, T., 2014.
Supplementary data associated with this article can be found in the Coexisting orchid species have distinct mycorrhizal communities and display strong
online version at doi:10.1016/j.rhisph.2018.07.001. spatial segregation. New Phtytol. 202, 616–627.
Jiménez, C., Sanabria, N., Altuna, G., Alcano, M., 2011. Efecto de Trichoderma harzianum
(Rifai) sobre el crecimiento de plantas de tomate (Lycopersicon esculentum L.). Rev.
References Fac. Agron. ((LUZ)) 28, 1–10.
Kassem, A., Nanniperi, P., 1995. Methods in Applied Soil Microbiology and Biochemestry,
1a edition. Elsevier, London, pp. 575.
Almanza-Álvarez, J., Garibay-Origel, R., Salgado-Garciglia, R., Fernández-Pavía, S., Khamchatra, N., Dixon, K., Chayamarit, K., Apisitwanich, S., Tantiwiwat, S., 2016. Using
Lappe-Oliveras, P., et al., 2017. Identification and control of pathogenic fungi in in situ seed baiting technique to isolate and identify endophytic and mycorrhizal
neotropical valued orchids (Laelia spp.). Trop. Plant Pathol. 42 (5), 339–351. https:// fungi from seeds of a threatened epiphytic orchid, Dendrobium friedericksianum
doi.org/10.1007/s40858-017-0171-3. Rchb.f. (Orchidaceae). Agric. Nat. Resour. 50 (1), 8–13. https://doi.org/10.1016/j.
Ávila-Díaz, I., Orijel, R., Magaña-Lemus, R., Oyama, K., 2013. Molecular evidence reveals anres.2016.01.002.
fungi associated within the epiphytic orchid Laelia speciosa (HBK) Schltr. Bot. Sci. 91 Kikuchi, G., Higuci, M., Morota, T., Nagasawa, E., Suzuki, A., 2008. Fungal symbiont and
(4), 523–529. cultivation test of Gastrodia elata Blume (Orchidaceae). Jpn. J. Bot. 83, 88–95.
Baev, P.V., Penev, L.D., 1995. BIODIV: Program for Calculating Biological Diversity Linde, C.C., Phillips, R.D., Crisp, M.D., Peakall, R., 2014. Congruent species declination of
Parameters, Similarity, Niche Overlap and Cluster Analysis. Version 5.1. Pensoft, Tulasnella using multiple loci and methods. New Phytol. 201, 6–12.
Sofia, Moscow, pp. 57. Liu, H., Luo, Y., Liu, H., 2010. Studies of mycorrhizal fungi of Chinese orchids and their
Bailarote, B.C., Lievens, B., Jacquemyn, H., 2012. Does mycorrhizal specificity affect role in orchid conservation in China – a review. Bot. Rev. 76, 241–262. https://doi.
orchids decline and rarety? Am. J. Bot. 99 (10), 1655–1665. org/10.1007/s12229-010-9045-9.
Bayman, P., Otero-Ospina, J.T., 2006. Microbial endophytes of orchid roots. Soil Biol. 9, Mapperson, R., Kotiw, M., Davis, R., Dearnaley, J., 2014. The diversity and antimicrobial
153–177. https://doi.org/10.1007/3-540-33526-9. activity of Preussia sp. endophytes isolated from Australian dry rainforests. Curr.
Beltrán-Nambo, M.A., Ortega-Larrocea, M.P., Salgado-Garciglia, R., Otero-Ospina, J.T., Microbiol. 68 (1), 30–37.
Martinez-Trujillo, M., Carreón-Abud, Y., 2012. Distribution and abundance of ter- Marmeisse, R., Girlanda, M., 2016. Mycorrhizal fungi and the soil carbon nutrient cycling.
restrial orchids of the genus Bletia in sites with different degrees of disturbance, in In: Drozhinina, I.S., Kiubicek, C.P. (Eds.), Mycota IV. Environmental and Microbial
the Cupatitzio Natural Reserve, Mexico. Int. J. Biodivers. Conserv. 4 (8), 316–325. Relationship, 3a edition. Springer, New York, pp. 189–204. https://doi.org/10.1007/
Bonnardeaux, Y., M. Brundrett, Y., Batty, A., Dixon, K., Koch, J., Sivasithamparam, K., 978-3-319-29532-9.
2007. Diversity of mycorrhizal fungi of terrestrial orchids: compatibility webs, brief Mitchell, R., 1989. Growing hardy orchids from seed at Kew. Plants Man 3 (2), 152–169.

55
M.d.l.A. Beltrán-Nambo et al. Rhizosphere 7 (2018) 49–56

McCormick, M.K., Jacquemyn, H., 2014. What constrains the distribution of orchid po- https://doi.org/10.3732./ajb 10000049.
pulations? New Phytol. 202, 392–400. Rodríguez, R.J., White, J.F., Arnold, A.E., Redman, R.S., 2009. Fungal endophytes: di-
McCormick, M.K., Whigham, D., Canchani-Viruet, A., 2018. Mycorrhizal fungi affect versity and functional roles. New Phytol. 182 (2), 314–330. https://doi.org/10.1111/
orchid distribution and population dynamics. New Phytol. Res. Rev. https://doi.org/ j.1469-8137.2009.02773.x.
10.1111/nph.15223. Segundo, R., 2016. Efecto de cultivos simbióticos y asimbióticos sobre la germinación,
Moreno-Martínez, D., 2011. Estudios de germinación in vitro e in situ de Epidendrum desarrollo de plántulas in vitro y supervivencia ex vitro de Laelia autumnalis, L.
parkinsonianum Hook. y Acineta barkeri (Bateman) Lindl. (Orchidaceae) (Tesis de Speciosa y Govenia superba (Orchidaceae). Tesis de Licenciatura. Facultad de
Maestría). Universidad Veracruzana, Xalapa, pp. 92. Biología, U.M.S.N.H., Michoacán, Méx, pp. 77.
Morgulis, A., Koulouris, G., Raytselis, Y., Madden, T.L., Agancuala, R., Schaffer, A., 2008. Seidl, M., 2009. Trichocladium. The Environmental Reporter 8. EMLab P. y K, pp. 1–3.
Database indexing for production MegaBLAST searches. Bioinformatica 24 (16), SEMARNAT, Secretaría del Medio Ambiente y Recursos Naturales, 2010. Norma Oficial
1757–1764. mexicana NOM-059. Protección Ambiental-Especies nativas de México de flora y
Munsell, 2000. Soil Color Charts. Gretag Macbeth, New Windsor, NY. fauna silvestres-categorias de riesgo y especificaciones para su inclusión, exclusión o
Nogueira, R.E., Van den Berg, C., Pereira, O.L., Kasuya, M.M., 2014. Isolation and mo- cambio-Lista de especies en riesgo. Diario Oficial de la Federación. Publicado el 30 de
lecular characterization of Rhizoctonia-life fungi associated with orchid roots in the diciembre de 2010.
Quadrilátero Ferrífero and Zona da Mata regions of the state of Minas Gerais, Brazil. Shan, X.C., Liew, E.C., Weatherhead, M.A., Hodgkiss, J.J., 2002. Characterization and
Acta Bot. Br. 28 (2), 298–300. taxonomic placement or Rhizoctonia-like endophytes from orchid roots. Mycologia
Nontachaiyapoom, S., Sasirat, S., Manoch, L., 2011. Symbiotic seed germination of 94 (2), 230–239.
Grammatophyllum speciosum Blume and Dendrobium draconis Rchb. f., native Shannon, C.E., Weaver, W., 1949. The Mathematical Theory of Communication. The
orchids of Thailand. Sci. Horticult. 130 (1), 303–308. https://doi.org/10.1016/j. University of Illinois Press, Urbana, pp. 117.
scienta.2011.06.04. Stark, C., Babik, W., Durka, W., 2009. Fungi from the root of the common terrestrial
Oliveira, S.F., Bocayuva, M.F., Veloso, T.G.R., Bazzoll, D.M.S., da Silva, C.C., Pereira, orchid Gymnadenia conopsea. Mycol. Res. 113, 952–959.
O.L., Kasuya, M.C.M., 2013. Endophytic and mycorrhizal fungi associated with roots Swarts, N.D., Dixon, K.W., 2009. Terrestrial orchid conservation in the age of extinction.
of endangered native orchids from the Atlantic Forest, Brazil. Mycorrhiza 24 (1), Ann. Bot. 104, 543–556. https://doi.org/10.1093/aob/mcp025.
55–64. https://doi.org/10.1007/s00572-013-0512-0. Swarts, N.D., Sinclair, E.A., Francis, A., Dixon, K.W., 2010. Ecological specialization in
Ortega-Larrocea, M.P., 2008. Propagación simbiótica de orquídeas terrestres con fines de mycorrhizal symbiosis leads to rarity in an endangered orchid. Mol. Ecol. 19,
restauración edafoecológica. In: Alvarez-Sánchez, J., Monroy-Ata, A. (Eds.), Técnicas 3226–3242. https://doi.org/10.1111/j.1365-294X.2010.04736.x.
de estudio de las asociaciones micorrízicas y sus implicaciones en la restauración. Tao, G., Liu, Z.Y., Hyde, K.D., Lui, X.Z., Yu, Z.N., 2008. Whole rDNA analysis reveals
Ciencias, UNAM, México, D.F., pp. 85–96. novel and endophytic fungi in Bletilla ochracea (Orchidaceae). Fungal Divers. 33,
Osgura-Tsujita, Y., Gebauer, G., Hashimoto, T., Umato, H., Yukawa, T., 2009. Evidence 101–122.
for novel and specialized mycorrhizal parisitism: the orchid Gastrodia confusa gains Teik-Khiang, G., Hyde, D., 1999. A synopsis of Trichocladium species, based on literature.
carbon from saprotrophic Mycena. Proc. Real Soc. Lond. B 276, 761–768. Fungal Divers. 2, 101–118.
Otero, J.T., Mosquera, A.T., Flanagan, N.S., 2013. Tropical orchid mycorrhizae: potential Teixeira, Da Silva J., Tsavkelova, E., Zeng, S., Bun, T.N., Parthibhan, S., Dobránszki, J.,
applications in orchid conservation, commercialization, and beyond. Lankesteriana Cardoso, J.C., Rao, M.V., 2015. Symbiotic in vitro seed propagation of Dendrobium:
13 (1–2), 57–63. fungal and bacterial partners and their influence on plant growth and development.
Otero, J.T., Thrall, P.H., Clements, M., Burdon, J.J., Miller, J.T., 2011. Codiversification Planta. https://doi.org/10.1007/s00425-015-2301-9.
of orchids (Pterostylidinae) and their associated mycorrhizal fungi. Aust. J. Bot. 59 Tesitelova, T., Jersakova, J., Roy, M., Kubatova, B., Tesitel, J., Urfus, T., Travnícek, P.,
(5), 480–497. https://doi.org/10.1071/BT11053. Suda, J., 2013. Ploidy-specific symbiotic interactions: divergence of mycorrhizal
Otero-Ospina, J.T., Bayman, P., 2009. Germinación simbiótica y asimbiótica en semillas fungi between cytotypes of the Gymnadenia conopsea group (Orchidaceae). New
de orquídeas epifitas. Acta Agronóm. 58 (4), 270–276. Phytol. 199, 1022–1033.
Pereira, M.C., Da Silva-Coelho, I., Da Silva-Valadares, R.B., Oliveira, S.F., Bocayuva, M., The Plant List, 2013. Version 1.1. Published on the Internet; 〈http://www.theplantlist.
Pereira, O.L., et al., 2014. Morphological and molecular characterization of org/〉, (Accessed 1 January, 2017).
Tulasnella spp. fungi isolated from the roots of Epidendrum secundum, a widespread Waud, M., Busschaert, P., Ruyters, S., Jacquemyn, H., Lievens, B., 2014. Impact of primer
Brazilian orchid. Symbiosis 62 (2), 111–121. https://doi.org/10.1007/s13199-014- choice on characterization of orchid mycorrhizal communities using 454 pyr-
0276-0. osequencing. Mol. Ecol. Resour. 14, 679–699.
Pereira, O.L., Kasuya, M.C., Borges, A.C., Araújo, E.F., 2005. Morphological and mole- Waud, M., Busschaert, P., Lievens, B., Jacquemyn, H., 2016. Specificity and localized
cular characterization of mycorrhizal fungi isolated from neotropical orchids in distribution of mycorrhizal fungi in the soil may attribute to co-existence of orchid
Brazil. Can. J. Bot. 83 (1), 54–65. https://doi.org/10.1139/b04-151. species. Fungal Ecol. 20, 155–165.
Phillips, R.D., Peakall, R., Hutchinson, F., Linde, C.C., et al., 2014. Specialized ecological Waud, M., Brys R., Landuyt W., Lievens B., Jacquemyn H., 2017. Mycorrhizal specificity
interactions and plant species rarity: the role of pollinators and mycorrhizal fungi not limit the distribution of an endangered orchid species. 〈http://dx.doi.org/10.
across multiple spatial scales. Biol. Conserv. 220–223, 1691285–1691295. 1111/mec.14014〉.
Pielou, E.C., 1981. The usefulness of ecological models: a stock-taking. Q. Rev. Biol. 56, White, T.J., Burns, T., Lee, S., Taylor, J., 1990. Amplification and direct sequencing of
17–31. fungal ribosomal RNA genes for phylogenetics. In: Innis, M.A., Gelfand, D.H.,
QGIS Development Team, 2009. QGIS Geographic Information System. Open Source Sninsky, J.J., White, T.J. (Eds.), PCR Protocols: A Guide to Methods and Applications.
Geospatial Foundation. 〈http://qgis.osgeo.org〉, consulted on Juanary 12, 2018. Academic Press, New York, pp. 315–322.
Rasmussen, H.N., Dixon, K.W., Jersáková, J., Těšitelová, T., 2015. Germination and Whittaker, R.H., 1972. Evolution and measurement of species diversity. Taxon 21 (2/3),
seedling establishment in orchids: a complex of requirements. Ann. Bot. 116 (3), 213–251.
391–402. https://doi.org/10.1093/aob/mcv087. Wright, M., Cross, R., Cousens, R.D., May, T.W., McLean, C.B., 2011. The functional
Rasmussen, H. y, Rasmussen, F.N., 2007. Trophic relationships in orchid mycorrhiza – significance for the orchid Caladenia tentaculata of genetic and geographic variation
diversity and implications for conservation. Lankesteriana 7 (1–2), 334–341. https:// in the mycorrhizal fungus Sebacina vermifera s. lat. complex. Muelleria 29 (2),
doi.org/10.15517/lank.v7i1-2.19560. 130–140.
Rasmussen, H.N. y, Whigham, D.F., 2002. Phenology of roots and mycorrhiza in orchid Xiaoya, M., Kang, J., Nontachaiyapoom, S., Wen, T., Hyde, K.D., 2015. Non-mycorrhizal
species differing in phototrophic strategy. New Phytol. 154 (3), 797–807. endophytic fungi from orchids. Curr. Sci. 109 (1), 72–87.
R-Development Core Team, 2008. R: A Language and Environment for Statistical Yamato, M., Yagame, T., Suzuhi, A., Iwase, K., 2005. Isolation and identification of my-
Computing. R Foundation for Statistical Computing, Vienna, Austria(〈http://www.R- corrhizal fungi associating with an achlorophyllous plant, Epipogium roseum
project.org〉). (Orchidaceae). Mycoscience 43, 73–77. https://doi.org/10.1007/s10267-004-
Rendón-Lara, C.E., Ortega-Larrocea M.P., Menchaca-García R., Lozano M., 2013. 0218-4.
Colonización micorrízica de dos especies de orquídeas epífitas del Estado de Veracruz Ye, W., Shen, C.H., Lin, Y., Chen, P.J., Xu, X., Oelmüller, R., et al., 2014. Growth pro-
(Cycnoches ventricosum) Batemam y (Chysis bractescens) Lindley. In: CONABIO- motion-related miRNAs in oncidium orchid roots colonized by the endophytic fungus
Viccon-Ávila-Nava (Ed.), Memorias del Segundo encuentro mexicano de Piriformospora indica. PLoS One 9 (1). https://doi.org/10.1371/journal.pone.
orquideología. Chilpancingo, Guerrero, México, p. 45. 0084920.
Rivera-Orduña, F.N., Suárez-Sanchez, R., Flores-Bustamante, Z., Gracida-Rodríguez, J.N., Yuan, Z.L., Chen, Y.C., Yang, Y., 2009. Diverse non-mycorrhizal fungi endophytes in-
Flores-Cotera, L.B., 2011. Diversity of endophytic fungi of Taxus blobosa (Mexican habiting an epiphytic, medicinal orchid (Dendrobium nobile) estimation and char-
yew). Fungal divers. 47, 65–67. acterization. J. Microbiol. Biotechnol. 25, 295–303.
Roche, S., Carter, R.I., Peakall, R., et al., 2010. A narrow group of monophyletic Zelmer, C.D., 1994. Interactions Between Northern Terrestrial Orchids and Fungi in
Tulasnella (Tullasnelaceae) lineages are associated with multiple species of Nature (M.Sc. Thesis). University of Alberta, Canada.
Chiloglottis (Orchidaceae) implications on diversity. Am. J. Bot. 97, 200–209.

56

You might also like