• No se han encontrado resultados

Genetic relationships within and among Iberian fescues (Festuca L.) based on PCR-amplified markers

N/A
N/A
Protected

Academic year: 2020

Share "Genetic relationships within and among Iberian fescues (Festuca L.) based on PCR-amplified markers"

Copied!
14
0
0

Texto completo

(1)1170. Genetic relationships within and among Iberian fescues (Festuca L.) based on PCR-amplified markers Pedro J.G. de Nova, Marcelino de la Cruz, Juan V. Monte, and Consuelo Soler. Abstract: The genus Festuca comprises approximately 450 species and is widely distributed around the world. The Iberian Penninsula, with more than 100 taxa colonizing very diverse habitats, is one of its main centers of diversification. This study was conducted to assess molecular genetic variation and genetic relatedness among 91 populations of 31 taxa of Iberian fescues, based on several molecular markers (random amplified polymorphic DNA, amplified fragment length polymorphisms, and trnL sequences). The analyses showed the paraphyletic origin of the broad-leaved (subgenus Festuca, sections Scariosae and Subbulbosae, and subgenus Schedonorus) and the fine-leaved fescues (subgenus Festuca, sections Aulaxyper, Eskia, and Festuca). Schedonorus showed a weak relationship with Lolium rigidum and appeared to be the most recent of the broad-leaved clade. Section Eskia was the most ancient and Festuca the most recent of the fine-leaved clade. Festuca and Aulaxyper were the most related sections, in concordance with their taxonomic affinities. All taxa grouped into their sections, except F. ampla and F. capillifolia (section Festuca), which appeared to be more closely related to Aulaxyper and to a new independent section, respectively. Most populations clustered at the species level, but some subspecies and varieties mixed their populations. This study demonstrated the value in combining different molecular markers to uncover hidden genetic relationships between populations of Festuca. Key words: Festuca, AFLPs, RAPDs, trnL. Résumé : Le genre Festuca comprend environ 450 espèces et il est largement répandu dans le monde. L’Ibérie, avec plus de 100 taxons occupant des habitats très différents, constitue un des principaux centres de diversification. Cette étude a été réalisée pour évaluer la variation génétique moléculaire et le degré de parenté génétique entre 91 populations appartenant à 31 taxons de fétuques ibériennes à l’aide de plusieurs types de marqueurs moléculaires (RAPD, AFLP et des séquences de trnL). Ces études ont révélé une origine paraphylétique des fétuques « à feuilles larges » (sous-genre Festuca, sections Scariosae et Subbulbosae ainsi que le sous-genre Schedonorus) des fétuques « à feuilles étroites » (sous-genre Festuca, sections Aulaxyper, Eskia et Festuca). Le sous-genre Schedonorus montre une relation étroite avec le Lolium rigidum et semble le plus récent au sein du clade « à feuilles larges ». La section Eskia est la plus ancienne et la section Festuca la plus récente au sein du clade « à feuilles étroites ». Les sections Festuca et Aulaxyper sont les plus apparentées conformément à leurs affinités taxonomiques. Tous les taxons ont été classés au sein de leurs sections à l’exception du F. ampla et du F. capillifolia (section Festuca) qui ont semblé plus proches, respectivement, de la section Aulaxyper et d’une nouvelle section indépendante. La plupart des populations étaient groupées ensemble au niveau de l’espèce, mais certaines sous-espèces et variétés étaient entremêlées. Cette étude démontre l’utilité de combiner différents types de marqueurs pour élucider les relations génétiques cachées entre les populations de Festuca. Mots clés : Festuca, AFLP, RAPD, trnL. [Traduit par la Rédaction]. Introduction Genus Festuca L. (family Poaceae, subfamily Pooideae, Received 2 March 2006. Accepted 14 June 2006. Published on the NRC Research Press Web site at http://genome.nrc.ca on 8 November 2006. Corresponding Editor: J.P. Gustafson. P.J. de Nova, J.V. Monte, and C. Soler.1 Departamento de Biotecnologı́a, INIA Apdo, 1045-28800, Alcalá de Henares, Spain. M. de la Cruz. Departamento de Biologı́a Vegetal, E.T.S.I. Agrónomos, Universidad Politécnica de Madrid, 28040 Madrid, Spain. 1Corresponding. author (e-mail: llinares@inia.es).. Genome 49: 1170–1183 (2006). tribe Poeae) is comprised mainly of rhizomatous and cespitose perennial (sometime annual) allogamous grasses, with a genome size ranging from diploid (2n = 14) to dodecaploid (2n = 84) (Kerguélen and Plonka 1989). The number of species varies, according to different authors, between approximately 300 and 450 (Tzvelev 1976; Clayton and Renvoize 1986; Watson and Dallwitz 1992), and increases periodically as new species are described (Kerguélen et al. 1993; de la Fuente and Ortúñez 1994a, 1994b; Chusovljanov 1998; González-Ledesma et al. 1998; Kulikov 1998, 2000; de la Fuente et al. 1999a, 1999b; Fletcher and Stace 2000). The genus has a broad geographic distribution in both hemispheres, and the area of distribution of its taxa varies from narrow endemic to subcosmopolitan. Some species of Festuca are valued as fodder, as ornamental plants, or as grass. doi:10.1139/G06-077. #. 2006 NRC Canada.

(2) de Nova et al.. coverage for eroded soils. However, both the potential genetic resources and the natural genetic variability of most species remain unexplored. The Iberian Peninsula is considered one of the main centers of diversification of the genus (Saint-Yves 1930), where up to 109 taxa have been described (78 species and 31 subspecies and varieties), and it is recognised that the list is still incomplete (Cebolla and Rivas 2003). Current intrageneric taxonomy distinguishes 9 subgenera within Festuca (Clayton and Renvoize 1986), 3 of which appear in the Iberian Peninsula: Festuca, Drymanthele V. Krecz and Bobrov, and Schedonorus (P. Beauv.) Peterm (Kerguélen and Plonka 1989). Most Iberian species are classified within subgenus Festuca (in sections Aulaxyper, Eskia, Festuca, Scariosae and Subbulbosae); only 6 taxa belong to subgenus Schedonorus (sections Plantynia and Schedonorus) and 2 to subgenus Drymanthele (Clayton and Renvoize 1986, Cebolla and Rivas 2003). Within the Iberian Peninsula, genus Festuca has a broad ecological range, and its species appear from sea level to the highest summits of Sierra Nevada and the Pyrenees, growing in a variety of habitats, from damp to dry soil or rock crevices and from meadows to scrub and forest. Festuca frequently plays a role in the structure of plant communities. The great morphologic similarity between different species, the existence of numerous ecotypes, and the range of intrapopulational differentiation of taxa make the study and classification of the genus a difficult task. To overcome problems such as the interference of the environment over the morphoanatomical characteristics usually used in classification, several molecular tools have been recently employed: restriction fragment length polymorphism (Xu and Sleper 1994), random amplified polymorphic DNA (RAPD) (Stammers et al. 1995; Siffelova et al. 1997), internal transcribed spacer (Charmet et al. 1997; Gaut et al. 2000; Torrecilla and Catalán 2002), and chloroplast DNA variability analysis (Lehväslaiho et al. 1987; Darbyshire and Warwick 1992; Catalán et al. 2004; Torrecilla et al. 2004). In this paper, we try to shed light on the genetic variability and the relations and differences among populations of Iberian wild fescues, using the molecular random PCR amplification of RAPD, amplified fragment length polymorphisms (AFLPs) (Williams et al. 1990; Vos et al. 1995), and the sequencing of the intron of chloroplast gene trnL (Taberlet et al. 1991).. Materials and methods Plant material A total of 91 populations, belonging to 2 subgenera and 6 sections of the genus Festuca L. and growing in the Iberian Peninsula were selected. Samples of Lolium rigidum (GenBank accession No. DQ376043), Agrostis castellana, and Puccinellia fasciculata (acc. No. DQ376076) were added to the study as controls. To conserve the original genetic structure of the natural populations, specimens were collected using the method of Hawkes (1980). This plant material belongs to the collection of wild Poaceae obtained from natural environments and maintained at the INIA plant breed-. 1171. ing unit (La Canaleja, Madrid, Spain). The taxa, distribution of the populations in Spain, and ploidy level are reflected in Table 1 and Fig. 1. A population named calcicolous-F. indigesta was included in the section Festuca, according to its morphoanatomical characteristics, but it could not be assigned to any known taxa (T.E. Dı́az-González, personal communication, 2002). DNA extraction and RAPD amplification Total DNA was extracted from the young green leaves of 30 individual specimens of each population, in accordance with the method of Dellaporta et al. (1983). RAPDs were performed in a reaction volume of 25 mL, containing 12.5 ng total DNA, 2.5 mL 10 reaction buffer, 7.5 ng primer (Kits M, R, S, and T, Operon Technologies, Alameda, Calif.), 100 mmol dNTP/L, and 1.25 U Taq DNA polymerase (Sigma, St. Louis, Mo.). Amplifications were carried out in a PTC-100 thermocycler (MJ Research Inc.). PCR followed the method of Williams et al. (1990), with minor modifications: 40 cycles of 1 min at 94 8C, 1 min at 36 8C, and 2 min at 72 8C. Amplification products were then resolved by electrophoresis in 1.8% agarose gel, using TAE buffer, and visualized with ethidium bromide staining. Each reaction and corresponding electrophoresis was repeated at least twice. AFLP procedures AFLP analysis was performed in accordance with the methods described by Vos et al. (1995). Genomic DNA (350 ng) was double-digested with 2 U each of EcoRI and MseI (Roche Diagnostics GmbH, Mannheim, Germany), in a final volume of 35 mL 1 buffer A, and incubated for 3 h at 37 8C. Adaptors EcoRI and MseI (5 and 50 pmols, respectively) were ligated to the resulting fragments (30 mL of the digestion mix), using 1 U T4 DNA ligase (Roche Diagnostics GmbH) in a final volume of 50 mL 1 buffer ligase, and incubated for 2 h at 25 8C. The ligation mix was diluted 1/10, and 5 mL were added to the preamplification, containing 1 reaction buffer, 30 ng each of primers EcoRI+1 and MseI+1, 0.25 mmol each dNTP/L, and 1 U of Taq DNA polymerase (Sigma) in a final volume of 20 mL. Preamplification was performed in a PTC-100 thermocycler as follows: 2 min at 95 8C; 20 cycles of 30 s at 94 8C, 1 min at 56 8C, 1 min at 72 8C; and 5 min at 72 8C. The preamplification mix was diluted 1/50, and 5 mL were added to the selective amplification, containing 1 reaction buffer, 30 ng each of primers EcoRI+3 and MseI+3, 0.25 mmol each dNTP/L, and 1 U of Taq DNA polymerase in a final volume of 20 mL. Selective amplification was performed as follows: 13 touchdown cycles (–0.7 8C per cycle) of 30 s at 94 8C, 30 s at 65 8C, 1 min at 72 8C; 23 cycles of 30 s at 94 8C, 30 s at 56 8C, 1 min at 72 8C; and 10 min at 72 8C. Five combinations of primer+3 were used with either the EcoRI+3 or MseI+3 primer end-labelled with 6-FAM (Table 2). The PCR products were separated by electrophoresis in a 6% polyacrylamide gel and analysed in an ABI PRISM 310 Genetic Analyser (Applied Biosystems). The reproducibility of the AFLP results was tested by replicating the complete procedure in 5 DNA samples for all the primer+3 combinations. #. 2006 NRC Canada.

(3) 1172. Genome Vol. 49, 2006 Table 1. Taxa and populations included in the study. Taxa and population ID Subgen. Festuca Sect. Festuca F. ampla Hackel AMP1 AMP2 AMP3 AMP4 F. arvernensis Auquier, Kerguelén & Markgr.Dannenb. subsp. costei (St.-Yves) Auquier & Kerguelén ARVc F. capillifolia Dufour in Roemer & Schultes CAP1 CAP2 CAP3 CAP4 F. clementei Boiss CLE F. curvifolia Lag. ex Lange CUR1 CUR2 CUR3 CUR4 F. gracilior (Hackel) Markgr.-Dannenb. GRA1 GRA2 GRA3 F. hystrix Boiss HYS1 HYS2 HYS3 HYS4 F. indigesta Boiss. IND1 IND2 Calcicolous F. indigesta INDc F. lemanii Bastard, Ess. LEM F longiauriculata Fuente, Ortúñez & Ferrero LON1 LON2 F. marginata (Hackel) K. Richter subsp. alopecuroides (Hackel) K. Richter MARa1 MARa2 MARa3 MARa4. Origin. Latitude. Longitude. Cádiz Ciudad Real Jaén Jaén. 36845’49@N 38830’56@N 38803’45@N 38817’12@N. 5823’00@W 3837’14@W 4800’28@W 2835’20@W. Chromosome No.. GenBank acc. No.. 2n = 28 DQ376048. 2n = 28. Barcelona. 41828’27@N. ´ 1848’14́E. DQ376070 2n = 14. Almerı́a Almerı́a Almerı́a Cuenca. 37817’22@N 34842’26@N 37815’48@N 39857’23@N. 2830’55@W 2813’17@W 2828’15@W 1815’12@W. Granada. 37803’49@N. 3822’12@W. DQ376072. 2n = 14 DQ376073 2n = 42 Madrid Madrid Madrid Madrid. 40850’13@N 40849’21@N 40847’31@N 40842’26@N. 3849’50@W 3856’54@W 4800’15@W 4808’55@W. DQ376058 DQ376059 2n = 14. Cuenca Teruel Teruel. 39857’23@N 40810’10@N 40812’14@N. 1815’12@W 1818’25@W 1821’08@W. Jaén Albacete Guadalajara Guadalajara. 38807’05@N 38835’26@N 40850’40@N 40841’55@N. 2834’10@W 2825’38@W 1853’02@W 2813’15@W. Granada Granada. 37803’49@N 37806’48@N. 3822’12@W 3825’03@W. Oviedo. 42859’55@N. 5855’52@W. DQ376064 2n = 14 DQ376068. 2n = 42 DQ376071 2n = 42 DQ376065 2n = 28 Gerona. 42815’53@N. 3817’14@E. DQ376060 2n = 14. Granada Almerı́a. 37807’34@N 37813’25@N. 3802’07@W 2832’01@W. DQ376067 2n = 14. Cuenca Teruel Teruel Teruel. 40827’35@N 40829’16@N 40830’54@N 40830’02@N. 2802’08@W 1841’15@W 1839’44@W 1840’31@W. DQ376063. #. 2006 NRC Canada.

(4) de Nova et al.. 1173. Table 1 (continued). Taxa and population ID F. rivas-martinezii Fuente & Ortúñez subsp. rivas-martinezii RIVr1 RIVr2 RIVr3 RIVr4 RIVr5 F. rivas-martinezii Fuente & Ortúñez subsp. rectifolia Fuente & Ortúñez RIVre F. summilusitana Franco & Rocha Afonso SUM1 SUM2 SUM3 SUM4 SUM5 F. valentina (St-Yves) Markgr.-Dannenb. VAL1 VAL2 VAL3 VAL4 VAL5 Sect. Aulaxyper F. heterophylla Lam. subsp. heterophylla HETh F. iberica (Hack.) K. Richter IBE1 IBE2 IBE3 IBE4 F. nevadensis (Hackel) K. Richter var. gaetula (Maire ex St. Yves) Al-Bermani & Stace NEVg1 NEVg2 NEVg3 F. nevadensis (Hackel) K. Richter var. nevadensis NEVn1 NEVn2 NEVn3 F. nigrescens Lam. subsp. nigrescens NIGn1 NIGn2 NIGn3 F. nigrescens Lam. subsp. nigrescens NIGn4. Origin. Latitude. Longitude. Madrid Madrid Madrid Madrid Ávila. 41803’32@N 41805’22@N 40858’35@N 40852’51@N 40840’55@N. 3828’11@W 3829’34@W 3849’18@W 3846’19@W 4841’03@W. Chromosome No. 2n = 14. GenBank acc. No.. DQ376062. 2n = 28. Teruel (Tormón). 40812’14@N. 1821’08@W. DQ376061 2n = 42. Palencia Palencia León León León. 43802’04@N 43800’47@N 42854’50@N 42853’31@N 42852’43@N. 4844’32@W 4843’41@W 5859’15@W 5858’32@W 5857’48@W. DQ376069. 2n = 14 Jaén Jaén Jaén Jaén Jaén. 38806’51@N 38806’30@N 38807’05@N 38807’28@N 38806’44@N. 2833’43@W 2834’46@W 2834’10@W 2833’03@W 2832’50@W. DQ376066. 2n = 28 Barcelona. 41828’27@N. 1848’14@E. DQ376050 2n = 28. Granada Madrid Madrid Guadalajara. 37806’48@N 41805’22@N 40852’51@N 41811’42@N. 3825’03@W 3829’34@W 3846’19@W 3825’38@W. DQ376046 2n = 14. Jaén Jaén Jaén. 37852’47@N 37853’02@N 37854’24@N. 2852’39@W 2853’00@W 2854’28@W. DQ376049. 2n = 28. Granada Jaén Jaén. 37858’16@N 37849’12@N 38807’11@N. 2827’06@W 2858’33@W 2840’21@W. DQ376051. 2n = 42 Palencia León Oviedo. 43800’47@N 42853’31@N 43802’04@N. 4843’41@W 5858’32@W 5842’56@W. DQ376055 2n = 14. León. 42842’16@N. 6855’15@W. #. 2006 NRC Canada.

(5) 1174. Genome Vol. 49, 2006 Table 1 (continued). Taxa and population ID NIGn5 NIGn6 NIGn7 NIGn8 NIGn9 NIGn10 F. rothmaleri (Litard.) Markgr.-Dannenb. ROT F. rubra L. subsp. pruinosa (Hackel) Piper RUBp F. trichophylla (Ducros ex Gaudin) K. Richter subsp. asperifolia (St.-Yves) Al-Bermani TRIa F. trichophylla (Ducros ex Gaudin) K. Richter subsp. scabrescens (Hackel ex Trabut) Catalán & Stace TRIs1 TRIs2 TRIs3 TRIs4 TRIs5 TRIs6 TRIs7 TRIs8 F. trichophylla (Ducros ex Gaudin) K. Richter subsp. trichophylla TRIt1 TRIt2 TRIt3 TRIt4 Sect. Eskia F. burnatii St-Yves BUR1 BUR2 F. elegans subsp. merinoi (Pau) Fuente & Ortúñez ELEm Sect. Scariosae F. scariosa (Lag.) Ascherson & Graebner SCA1 SCA2 SCA3 Sect. Subbulbosae F. paniculata (L.) Schinz & Tell. subsp. paui Cebolla & Rivas Ponce PANpa. Origin Lugo León Lugo Lugo Lugo Lugo. Latitude 42853’04@N 42853’32@N 42855’19@N 42845’30@N 42844’07@N 42843’35@N. Longitude 7804’42@W 6848’25@W 6851’56@W 7801’00@W 6859’55@W 6858’47@W. Chromosome No.. GenBank acc. No.. DQ376056. 2n = 28 Madrid. 41804’53@N. 3827’09@W. DQ376047 2n = 42. Oviedo. 43839’20@N. 5850’56@W. DQ376057 2n = 28. Guadalajara. 40840’02@N. 2817’13@W. DQ376052 2n = 14. Guadalajara Guadalajara Cuenca Teruel Cuenca Teruel Teruel Teruel. 40856’41@N 40855’34@N 40836’56@N 40828’05@N 40825’52@N 40821’09@N 40830’54@N 40830’05@N. 2851’52@W 2855’21@W 2816’40@W 1836’33@W 1853’31@W 1844’32@W 1839’44@W 1839’01@W. DQ376053 2n = 28. Palencia Soria Teruel Guadalajara. 42848’46@N 41808’00@N 40821’09@N 40841’46@N. 4843’35@W 2826’48@W 1844’32@W 1854’14@W. León Oviedo. 42853’31@N 42859’08@N. 5858’32@W 5853’27@W. DQ376054. 2n = 14 DQ376075 2n = 28. León. 42840’19@N. 6847’29@W. DQ376074 2n = 14. Almerı́a Almerı́a Almerı́a. 37802’04@N 37818’53@N 37816’07@N. 2857’03@W 2830’26@W 2827’27@W. DQ376045. 2n = 14. Cuenca. 40808’49@N. 2841’28@W. DQ376044. #. 2006 NRC Canada.

(6) de Nova et al.. 1175. Table 1 (concluded). Taxa and population ID Subgen. Schedonorus Sect. Schedonorus F. arundinacea Schreb. subsp. mediterranea (Hackel) K. Richter ARUm. Origin. Latitude. Longitude. Chromosome No.. GenBank acc. No.. 2n = 28. Málaga. 36856’34@N. 4827’55@W. DQ376042. Fig. 1. Map of the Iberian Penninsula with the localization of the studied sections: !, section Festuca; ~, section Aulaxyper; *, section Eskia; &, section Scariosae; ^, section Subbulbosae; +, section Schedonorus. Coordinates of individual taxa can be found in Table 1.. trnL intron amplification and sequencing The chloroplast noncoding region of the trnL (UAA) intron was used for phylogenetic analysis. Amplification was performed using 100 ng genomic DNA, 1 reaction buffer, 0.2 mmol each dNTP/L, 100 ng of each primer (5’-CGAAATCGGTAGACGCTACG-3’ and 5’-GGGGATAGAGGGACTTGAAC-3’) (Taberlet et al. 1991), and 2 U of Taq DNA polymerase in a final volume of 100 mL. PCR conditions were programmed as follows: 16 touchdown cycles (–0.5 8C per cycle) of 1 min at 94 8C, 1 min at 62 8C, 2 min at 72 8C; 29 cycles of 1 min at 94 8C, 1 min at 54 8C, 2 min at 72 8C; and 10 min at 72 8C. Amplification product was extracted with QIAquick Gel Extraction Kit (Qiagen GmbH, Hilden, Germany), after 1% agarose gel electrophoresis in TAE buffer. For each sample, direct and reverse sequences were obtained using the 2 amplification primers. RAPD and AFLP statistical analysis RAPD and AFLP data were scored on 2 matrices. Faint. amplified bands were discarded and fragments of identical size were considered to be the same. Presence (1) or absence (0) were used as character states. The data matrices were analysed with the NTSYS computer package, version 1.60 (Rolhlf 1989). The Jaccard similarity coefficient (Jaccard 1908) and the unweighted pair-group method, arithmetic averages (UPGMA) (Sneath and Sokal 1973) were combined to construct dendrograms. trnL alignment and sequence analysis First, sequences were aligned using the CLUSTAL W (1.82) program with default parameters (Thompson et al. 1994). After alignment, phylogenetic analysis was conducted with the PAUP 4.0 beta 8 program (Swofford 1998), using the parsimony criterion (Swofford and Berlocher 1987), a heuristic search procedure, and the MULPARS, TBR branch swapping, and CLOSEST addition options. Bootstrap support was estimated with 1000 replicates (Sanderson 1989). A sample of the species Puccinellia fasciculata was used in the study as the outgroup. The resulting tree was constructed using TreeView (Page 1996). #. 2006 NRC Canada.

(7) 1176. Genome Vol. 49, 2006. Table 2. Sequences of the adaptors and primers used in amplified fragment length polymorphisms (AFLPs). Type Adaptor EcoRI+ Adaptor EcoRI– Adaptor MseI+ Adaptor MseI– Primer EcoRI+1 Primer MseI+1 Primer EcoRI+3 Primer EcoRI+3 Primer EcoRI+3 Primer EcoRI+3 Primer MseI+3 Primer MseI+3 Primer MseI+3. Sequences 5’-CTCGTAGACTGCGTACC-3’ 3’-CATCTGACGCATGGTTAA-5’ 5’-GACGATGAGTCCTGAG-3’ 3’-TACTCAGGACTCAT-5’ 5’-GACTGCGTACCAATTCA-3’ 5’-GATGAGTCCTGAGTAAC–3’ 6-FAM-5’-GACTGCGTACCAATTCACC-3’ 6-FAM-5’-GACTGCGTACCAATTCACG-3’ 5’-GACTGCGTACCAATTCAGC-3’ 6-FAM-5’-GACTGCGTACCAATTCAGG-3’ 6-FAM-5’-GATGAGTCCTGAGTAACAG-3’ 6-FAM-5’-GATGAGTCCTGAGTAACTG-3’ 5’-GATGAGTCCTGAGTAACAA-3’. Results RAPDs PCR amplification with the 30 primers generated 1389 fragments. Of these, 724 (52.12%) fragments were present only within populations, 986 (70.99%) were present only within species, subspecies and varieties, and 1248 (89.85%) were present only within sections. None of the fragments was present in all populations analysed. Within populations, F. curvifolia had the maximum Jaccard similarity coefficient (average 0.780) and F. iberica had the lowest (average 0.293). F. gracilior and F. rivasmartinezii subsp. rectifolia were the most similar taxa (average 0.574), and the minimum similarity was between F. burnatii and both F. capillifolia and F. arundinacea subsp. mediterranea (0.000). The 3 minima of similarity between F. capillifolia and sections Aulaxyper, Eskia, and their own section Festuca was remarkable. Festuca ampla was much more similar to almost all populations of section Aulaxyper than to any population of its own section Festuca. The average similarity of genus Festuca to the controls L. rigidum, A. castellana, and P. fasciculata were 0.008, 0.016, and 0.00001, respectively. The highest similarity values within sections (average similarity of all populations) were in Scariosae and Eskia (0.316 and 0.331, respectively), and the lowest were in Aulaxyper and Festuca (0.235 and 0.231, respectively). The highest similarities between sections were between Aulaxyper and Festuca (0.081) and between Subbulbosae and Eskia (0.078); the least related sections were Schedonorus and Eskia (0.003). The dendrogram clearly discriminated all taxa, showing high levels of variability even inside individual species (Fig. 2). In general, populations of the same species clustered together, but some populations of F. trichophylla, F. iberica, F. marginata, and F. rivas-martinezii joined clusters of other taxa. In the case of F. rivas-martinezii, the populations that did not cluster were from different subspecies. Species from the same section appeared together in the same clusters, except in the case of F. capillifolia and F. ampla, both of section Festuca. Festuca ampla clearly joined the Aulaxyper cluster and F. capillifolia joined F. arundinacea subsp. mediterranea (section Schedonorus). Festuca panicu-. lata subsp. paui, the only representative taxon of section Subbulbosae, appeared close to section Eskia. The clusters of sections Festuca and Aulaxyper grouped first at the highest similarity, and the clusters of sections Subbulbosae, Eskia, Schedonorus, and Scariosae were added sequentially with progressive dissimilarity. The taxonomic markers Agrostis and Puccinellia appeared in the outermost positions of the dendrogram; Lolium joined section Scariosae. Within section Festuca, F. valentina, F. gracillior, and F. rivas-martinezii subsp. rectifolia grouped together in the innermost cluster, showing the greatest genetic relatedness (similarity over 0.488); F. summilusitana, F. indigesta, and calcicolous-F. indigesta grouped in another cluster (similarity over 0.462), and were joined by F. curvifolia (with similarity of 0.315). Another cluster was formed by F. arvenensis, F. marginata, and F. rivas-martinezii subsp. rivas-martinezii (similarity over 0.255), and a 4th one by F. hystrix and F. longiauriculata (similarity over 0.211). F. lemanii and F. clementei remained as the outermost species, joining the rest of the section with similarity indices of 0.156 and 0.102, respectively. Within section Aulaxyper, F. nigrescens populations appeared in 2 distinct groups related to ploidy level (3 populations with 2n = 42 and 7 with 2n = 14). They were joined by F. trichophylla populations (similarity of 0.263) and, in turn, were joined by the group of F. nevadensis (with similarity of 0.236). The group of F. rothmaleri and F. heterophylla joined with less similarity, and were subsequently joined by the populations of F. ampla (with similarity of 0.116). AFLPs Five primer combinations generated 1347 fragments between 70 and 488 bp. A high percentage of fragments (52.26%) were under 200 bp, and only 5.50% exceeded 400 bp. No fragment was shared by all populations; 419 fragments (31.11%) were present only within populations; 681 (50.56%) within species, subspecies, or varieties; and 1011 fragments (75.06%) were present only in sections. Similarity values were, in general, higher than in RAPD analysis. Festuca longiauriculata had the highest average similarity within species (0.829); F. trichophylla subsp. scabrescens and F. iberica had the lowest (both 0.388). Festuca gracilior and F. rivas-martinezii subsp. rectifolia showed the highest similarity between taxa (0.676), whereas the lowest was found between F. burnatii and F. scariosa (0.034). Average similarity between genus Festuca and controls were 0.083, 0.031, and 0.042 for Lolium, Agrostis, and Puccinellia, respectively. Within sections, again Scariosae and Festuca had the highest and lowest values, respectively (0.743 and 0.351). Between sections, Festuca and Aulaxyper had the highest average similarity (0.169), and Eskia and Scariosae had the lowest (0.047). The AFLP dendrogram also showed high variability within sections and species, although with higher similarities than in the RAPD case (Fig. 3). Again, Festuca and Aulaxyper appeared more related to themselves than to the remaining sections (0.177). Section Schedonorus was the most similar to both (0.098), followed by Scariosae (0.075) and EskiaSubbulbosae (0.071). The group of Agrostis and Puccinellia controls appeared most isolated, sharing a 0.037 similarity #. 2006 NRC Canada.

(8) de Nova et al.. 1177. Fig. 2. Unweighted pair-group method with arithmetic averages (UPGMA) dendrogram of Festuca populations, using the Jaccard similarity coefficient, based on 1389 random amplified polymorphic DNA (RAPD) markers.. #. 2006 NRC Canada.

(9) 1178. Genome Vol. 49, 2006. Fig. 3. UPGMA dendrogram of Festuca populations, using the Jaccard similarity coefficient based on 1347 amplified fragment length polymorphism (AFLP) markers.. #. 2006 NRC Canada.

(10) de Nova et al.. with the main group of fescues. F. paniculata subsp. paui (sect. Subbulbosae) joined section Eskia, as it did in the RAPD dendrogram. F. ampla appeared again inside the main group of section Aulaxyper, instead of within the Festuca group. F. capillifolia also quit section Festuca. Lolium rigidum appeared closer to F. arundinacea subsp. mediterranea (similarity of 0.113), instead of joining F. scariosa, as it did in the RAPD dendrogram. In general, and as in the RAPD dendrogram, populations of the same taxa grouped together, except again in the case of F. trichophylla, F. iberica, F. rivas-martinezii, F. marginata, and F. nigrescens. The genetic distance between the 2 ploidy groups of F. nigrescens was even greater than in RAPDs, and the separation of F. rivas-martinezii populations reflected its intraspecific taxonomy. As in the RAPD dendrogram, the innermost cluster within section Festuca was formed by F. gracilior, F. valentina, and F. rivas-martinezii subsp. rectifolia (similarity of 0.622); F. indigesta, calcicolous-F. indigesta, and F. summilusitana were grouped in another cluster, with similarity over 0.494, and F. arvernensis, F. marginata, and F. rivasmartinezii clustered with a similarity of 0.439. In contrast, F. hystrix and F. clementei were the outermost species and joined the main group with similarities under 0.300. Within section Aulaxyper, F. heterophylla and F. rothmaleri had the strongest interspecific relationship; these were joined by F. ampla (from section Festuca) at a similarity level of 0.371. As in the RAPD dendrogram, F. nigrescens subsp. nigrescens populations appeared in 2 groups but, in this case, with an even weaker relationship between themselves; the hexaploid populations grouped in the outermost cluster within Aulaxyper. Festuca valentina and F. scariosa populations had the strongest intraspecific relationship (they grouped at similarity levels of 0.751 and 0.703, respectively); F. capillifolia and F. nevadensis had the lowest (under 0.643 and 0.702, respectively). As in the RAPD case, relationships between intraspecific taxa, such as subspecies and varieties, were not the strongest in the dendrogram. trnL sequences The PCR amplification and sequencing of the homologue regions of the intron of the chloroplast trnL gene resulted in a product of unique sequences, ranging from 496 bp in F. summilusitana to 516 bp in F. burnatii (among the populations of fescues), and to 519 bp in L. rigidum. After the alignment and coding of gaps, the dataset comprised 530 characters. A remarkable gap was a 9 bp deletion shared by all populations of section Festuca and Aulaxyper except F. capillifolia and F. clementei. The relative frequencies of bases were A, 0.372; T, 0.303; C, 0.142; and G, 0.183. Transversions were more abundant than transitions (ratio, 0.635). Phylogenetic analysis detected 50 variable characters, of which 24 were parsimony informative. The most parsimonious tree (Fig. 4) had a total length of 66 (consistency index, 0.833; retention index, 0.903; homoplasy index, 0.167), and the bootstrap support of the nodes ranged between 51% and 98%. The tree showed 2 diverging branches; in one of them, sections Scariosae, Subbulbosae, and Schedonorus shared a common ancestor, and F. arundinacea. 1179. subsp. mediterranea (Schedonorus) appeared closely related to L. rigidum, both with a common and recent origin. In the other branch, the Eskia group showed their ancient origin, whereas sections Aulaxyper and Festuca diverged much more recently. All taxa of section Aulaxyper formed a paraphyletic grouping, except F. rubra subsp. pruinosa and F. nigrescens subsp. nigrescens, which shared a monophyletic origin. Section Festuca diverged even more recently, and all their taxa except F. ampla, F. capillifolia, and F. clementei formed a paraphyletic group. F. ampla joined the paraphyletic group of Aulaxyper taxa, whereas F. capillifolia and F. clementei diverged much earlier and formed a paraphyletic grouping with F. elegans subsp. merinoi of section Eskia. Both populations of F. curvifolia showed a monophyletic origin with a high bootstrap value. The 2 populations of F. nigrescens subsp. nigrescens also showed a monophyletic origin but, in this case, with little support of bootstrap, probably because of their different ploidy level.. Discussion The high level of polymorphism detected in this study reflects the high genetic diversity of genus Festuca. The degree of polymorphism was lower in AFLPs, probably because most fragments (52.26%) were under 200 bp; these smaller fragments are less likely to contain polymorphisms because they have fewer base pairs than the bigger RAPD fragments. However, with both RAPDs and AFLPs, all populations were clearly differentiated and, for almost every taxon, the geographic and the genetic distances between their populations were highly correlated (data not shown). In addition, the high values of consistence and retention indices of the phylogenetic tree, based on the trnL, together with the lower homoplasy index, showed the optimization of the cladistic analysis. The results of the analyses clarified the relationships between Festuca populations at different scales. At the section level, both RAPDs and AFLPs, together with the trnL sequencing, highlighted the close relationship between sections Aulaxyper and Festuca, consistent with the similar patterns of chromosome C-banding (Dawe 1989) and their morphoanatomical similarity (on this basis, both were included in section Ovinae by Hackel 1882). Although Darbyshire and Warwick (1992), based on chloroplast DNA restriction site variation, proposed a paraphyletic origin, the monophyletic origin of both sections was shown by Gaut et al. (2000), Torrecilla and Catalán (2002), and Catalán et al. (2004), using internal transcribed spacer sequences. Some authors (Tzvelev 1983; Holub 1984) proposed that the broad-leaved fescues (sections Scariosae, Schedonorus, and Subbulbosae) were ancestral to the ‘‘fine-leaved’’ fescues (sections Aulaxyper, Eskia, and Festuca). Our trnL phylogram showed the paraphyletic origin of both groups, and a similar divergence time, corroborating the results of Torrecilla and Catalán (2002). Section Eskia can be considered an intermediate group between broad- and fine-leaved fescues because of its anatomical similarities with section Scariosae (Torrecilla et al. 2004). In our study, section Subbulbosae joins section Eskia in both RAPDs and AFLPs, a fact that can be explained by #. 2006 NRC Canada.

(11) 1180. Genome Vol. 49, 2006. Fig. 4. Strict consensus of the most parsimonious trees obtained from a phylogenetic analysis of trnL intron sequences. Numbers are bootstrap values of the nodes.. the existence of the homoplasy phenomena; however, both sections appear in separate branches of the phylogram and diverge at different times. Chandrasekharan et al. (1972) reported a high genetic affinity between section Scariosae and section Schedonorus but, in our study, this relationship appears only in the monophyletic origin of both within the broad-leaved group. A recent botanical study (de la Fuente and Ortúñez 2001) proposed the inclusion of F. scariosa in section Eskia, but all of our analyses suggest that it should remain in its own section. The species included as controls were only weakly associated with Festuca taxa, except in the case of L. rigidum, which was associated with section Schedonorus in AFLPs and in the phylogenetic tree and to section Scariosae in the. RAPD analysis. The relationship between Lolium and Schedonorus has been repeatedly described (Holub 1984; Lehväslaiho et al. 1987; Morgan et al. 1988; Darbyshire 1993; Stammers et al. 1995; Pasakinskiené et al. 1998). Cytological studies have noted the relationship between Lolium and ancestors of Festuca (Jenkin 1933; Malik and Thomas 1966), whereas other studies, based on comparative anatomy, have suggested a relationship between the reduced inflorescence of Lolium and some ancestor of F. pratensis (Essad 1962). On the basis of these relationships, some authors have concluded that Schedonorus should be treated as a subgenus of Lolium (Peto 1933; Crowder 1953; Stebbins 1956; Terrell 1966; Darbyshire 1993). In contrast, the morphology of the spikelet, the existence of only diploid genomes (proof of its more recent origin), and the differences #. 2006 NRC Canada.

(12) de Nova et al.. in the proteins of the endosperm have been stated as reasons to keep Lolium and Schedonorus as independent genera (Aiken et al. 1997, 1998; Holub 1998). Our study supports the idea of including Lolium as a subgroup of Schedonorus, as has been proposed by Terrell (1968). At the specific or intraspecific level, some interesting relationships between species have been revealed. For instance, Bulinska-Radomska and Lester (1986) presented both F. heterophylla and F. rubra as transitional between sections Schedonorus and Ovinae (as the summation of Aulaxyper and Festuca), but our study shows that they are fully Aulaxyper taxa. F. ampla has long been classified within section Festuca (Hackel, 1882) although it shows some of the anatomical features that Stace et al. (1992) consider characteristic of the ‘‘F. rubra aggregate’’: the absence of auricles and adaxial schlerenchyma strands in their leaf blades. Also, from an ecological point of view, the requirement of deep soil by F. ampla (de la Fuente and Ortúñez 1995) more closely resembles the ecological affinities of most Aulaxyper taxa than those of section Festuca, which are better adapted to shallow soils. All our analyses confirm these relationships and suggest that F. ampla should be placed in section Aulaxyper. F. clementei and F. capillifolia, both classified within section Festuca (and both with the adaxial schlerenchyma characteristic), appeared in an ancestral position on the branch of fine-leaved fescues in the phylogram (together with Eskia taxa), and quite separate from the rest of section Festuca in the RAPD and AFLP dendrograms. Saint-Yves (1922) questioned the taxonomic position of F. capillifolia and proposed the creation of a subsection (Exaratae) to distinguish it from the rest of Festuca (a proposal that did not succeed in the taxonomic literature); our results agree with that proposition but suggest that the rank of the taxonomic category should be greater than a subsection (see also Torrecilla and Catalán 2002; Catalán et al. 2004). Within section Festuca, the grouping of F. arvernensis subsp. costei, F. rivas-martinezii, and F. marginata subsp. alopecuroides could be related to their anatomical similarities (all of them bear discontinuous abaxial sclerenchyma strands in their leaf blades, a differential characteristic in Festuca taxonomy), although 1 population of F. marginata subsp. alopecuroides and 1 of F. rivas-martinezii subsp. rectifolia did not join the group. The separation of the subspecies of F. rivas-martinezii into different groups supports the idea of Cebolla and Rivas (2003) of 2 different species: F. longifolia subsp. rivas-martinezii (Fuente & Ortúñez) Cebolla et al. and F. rectifolia (Fuente et al.) Cebolla & Rivas Ponce (Cebolla and Rivas 2003). The genetic similarity of F. valentina and F. gracilior (both group in the AFLP and RAPD dendrogram) is consistent with their morphoanatomical similarity; some authors (de la Fuente and Ortúñez 1998) treat them as the same species, whereas others consider them different species (Cebolla and Rivas 2001, 2003). The absence of any special relationship in the trnL sequence prevents further conclusions about their status. The calcicolous F. indigesta population was genetically similar to genuine F. indigesta, both in RAPDs and AFLPs, which is consistent with their morphological similarities. F. trichophylla subsp. scabrescens populations showed a. 1181. high genetic polymorphism and have been grouped with populations of other subspecies of F. trichophylla, making their intraspecific taxonomic structure unrecognisable. The taxonomic relatedness of F. trichophylla subsp. scabrescens and F. iberica remains controversial. Al-Bermani et al. (1992) proposed that they were the same taxon, whereas de la Fuente et al. (1997) reaffirmed the independence of both taxa based, among other reasons, on their different genome size. The grouping patterns of both taxa did not confirm either hypotheses because, although some populations were grouped together, others remained entirely independent. The separation of F. nevadensis populations did not reflect their varietal status, although, in this case, the molecular techniques could not discriminate among varieties because their differential characteristics appear mixed in the populations (Kerguélen and Plonka 1989). In contrast, the 10 populations of F. nigrescens subsp. nigrescens formed 2 distinct groups, 1 for each ploidy level in the RAPD and AFLP dendrograms, suggesting that they could be considered different subspecies or even species. In addition, it is noteworthy that F. nigrescens associates with F. rubra subsp. pruinosa in the trnL phylogram, supporting the monophyletic origin of the F. rubra group, which AlBermani et al. (1992) discriminated from the F. trichophylla group on the basis of some morphoanatomical characteristics.. Acknowledgements The authors gratefully acknowledge Dr. T.E. Dı́az González (Universidad de Oviedo, Spain) for his advice on calcicolous F. indigesta, Dr. Pilar Catalán (Universidad de Zaragoza, Spain) for her helpful remarks about this study, and Mercedes Martı́nez for her linguistic assistance. We also acknowledge the financial support given by INIA (project SC97-065 and RTA02-038).. References Aiken, S.G., Dallwitz, M.J., McJannet, C.L., and Consaul, L.L. 1997. Biodiversity among Festuca (Poaceae) in North America: diagnostic evidence from DELTA and clustering programs, and an INTKEY package for interactive, illustrated identification and information retrieval. Can. J. Bot. 75: 1527–1555. Aiken, S.G., Gardiner, S.E., Bassett, H.C.M., Wilson, B.L., and Consaul, L.L. 1998. Implications from SDS-PAGE analyses of seed proteins in the classification of taxa of Festuca and Lolium (Poaceae). Biochem. Syst. Ecol. 26: 511–533. Al-Bermani, A.K.K.A., Catalán, P., and Stace, C.A. 1992. A new circumscription of Festuca trichophylla (Gaudin) K. Richter (Gramineae). An. Jard. Bot. Madr. 50: 209–220. Bulinska-Radomska, Z., and Lester, R.N. 1986. Phylogeny of chromosome races of Festuca arundinacea and F. mairei (Poaceae) as indicated by seed protein electrophoresis. Plant Syst. Evol. 152: 153–166. Catalán, P., Torrecilla, P., López Rodrı́guez, J.A., and Olmstead, R.G. 2004. Phylogeny of the festucoid grasses of subtribe Loliinae and allies (Poeae, Pooideae) inferred from ITS and trnL-F sequences. Mol. Phylogenet. Evol. 31: 517–541. PMID:15062792. Cebolla, C., and Rivas, M.A. 2001. Festuca michaelis (Poaceae), une nouvelle espèce pour la Péninsule Ibérique. Flora Mediterr. 11: 363–371. #. 2006 NRC Canada.

(13) 1182 Cebolla, C., and Rivas, M.A. 2003. Catálogo del género Festuca L. (Poaceae) en la Penı́nsula Ibérica. Candollea, 58: 189–213. Chandrasekharan, P., Lewis, E.J., and Borrill, M. 1972. Studies in Festuca. 2. Fertility, relationships between species of the sections Bovinae and Scariosae, and their affinities with Lolium. Genetica, 43: 375–386. doi:10.1007/BF00156133. Charmet, G., Ravel, C., and Balfourier, F. 1997. Phylogenetic analysis in the Festuca-Lolium complex using molecular markers and ITS rDNA. Theor. Appl. Genet. 94: 1038–1046. doi:10. 1007/s001220050512. Chusovljanov, D.V. 1998. A new species of the genus Festuca (Poaceae) from Altai. Bot. Z. 83: 113–115. Clayton, W.D., and Renvoize, S.A. 1986. Genera graminum. Grasses of the world. Kew Publishing, London, UK. Crowder, L.V. 1953. Interspecific and intergeneric hybrids of Festuca and Lolium. J. Hered. 44: 195–203. Darbyshire, S.J. 1993. Realignment of Festuca Subgenus Schedonorus with the Genus Lolium (Poaceae). Novon, 3: 239–243. doi:10.2307/3391460. Darbyshire, S.J., and Warwick, S.I. 1992. Phylogeny of North American Festuca (Poaceae) and related genera using chloroplast DNA restriction site variation. Can. J. Bot. 70: 2415–2429. Dawe, J.C. 1989. Sectional survey of Giemsa C-banded karyotypes in Festuca L. (Poaceae: Festuceae): systematic and evolutionary implications. Ph.D. thesis, Viena University. Dellaporta, S.L., Wood, J., and Hicks, J.B. 1983. A plant DNA minipreparation: version II. Plant Mol. Biol. Rep. 1: 19–21. Essad, S. 1962. Étude génétique et cytogénétique des espèces Lolium perenne L., Festuca pratensis Huds. et leurs hybrides. Ann. Inst. Nat. Rech. Agron. Ser. B. No. 8. Fletcher, R., and Stace, C.A. 2000. A new section and species of Festuca (Poaceae) naturalized in England. Watsonia, 23: 173–177. de la Fuente, V., and Ortúñez, E. 1994a. A new species of Festuca L. from Spain. Bot. J. Linn. Soc. 114: 23–30. doi:10.1006/bojl. 1994.1002. de la Fuente, V., and Ortúñez, E. 1994b. A new species of Festuca L. (Poaceae) from Pyrenees (Spain). Fontqueria, 40: 35–42. de la Fuente, V., and Ortúñez, E. 1995. Festuca henriquesii Hackel y F. ampla Hackel en la Penı́nsula Ibérica. Stud. Bot. 14: 129–141. de la Fuente, V., and Ortúñez, E. 1998. Biosistemática de la sección Festuca del género Festuca L. (Poaceae) en la Penı́nsula Ibérica. Ediciones de la Universidad Autónoma de Madrid, Cantoblanco, Madrid. de la Fuente, V., and Ortúñez, E. 2001. Festuca sect. Eskia (Poaceae) in the Iberian Peninsula. Folia Geobot. 36: 385–421. de la Fuente, V., Ortúñez, E., and Ferrero, L.M. 1997. Contribución al conocimiento del género Festuca L. (Poaceae) en el Paı́s Vasco y Sistema Ibérico septentrional (Penı́nsula Ibérica). Itinera Geobot. 10: 317–351. de la Fuente, V., Ortúñez, E., and Ferrero, L.M. 1999a. A new species of Festuca L. (Poaceae) from the Iberian Peninsula. Parlatorea, 3: 65–77. de la Fuente, V., Ortúñez, E., and Ferrero, L.M. 1999b. New species of Festuca L. section Festuca (Poaceae) in the Iberian Peninsula. Lazaroa, 20: 3–9. Gaut, B.S., Tredway, L.P., Kubik, C., Gaut, R.L., and Meyer, W. 2000. Phylogenetic relationships and genetic diversity among members of the Festuca-Lolium complex (Poaceae) based on ITS sequence data. Plant Syst. Evol. 224: 33–53. doi:10.1007/ BF00985265. González-Ledesma, M., Koch, S.D., and Gómez, M. 1998. Two new species of Festuca (Gramineae: Pooideae) from Mexico. Novon, 8: 147–151. doi:10.2307/3391986.. Genome Vol. 49, 2006 Hackel, E. 1882. Monographia festucarum Europaearum. Kassel und Berlin. p. 216. Hawkes, J.G. 1980. Crop genetic resources. Field collection manual for seed crops, root and tuber crops, tree fruit crops and related wild species. International Board for Plant Genetic Resources (I.B.P.G.R.) and the European Association for Research on Plant Breeding (EUCARPIA), Birmingham, UK. Holub, J. 1984. New genera in Phanerogamae. Folia Geobot. Phytotax. 19: 95–99. Holub, J. 1998. Reclassifications and new names in vascular plants 1. Preslia, 70: 97–122. Jaccard, P. 1908. Nouvelles recherches sur la distribution florale. Bull. Soc. Vaudoise Sci. Nat. 44: 223–270. Jenkin, T.J. 1933. Interspecific and intergeneric hybrids in herbage grasses. Initial crosses. J. Genet. 28: 205–264. Kerguélen, M., and Plonka, F. 1989. Les Festuca de la Flore de France (Corse comprise). Bull. Soc. Bot. Centre-Ouest, 10: 1–368. Kerguélen, M., Plonka, F., and Chas, É. 1993. Nouvelle contribution aux Festuca (Poaceae) de France. Lejeunia, 142: 1–42. Kulikov, P.V. 1998. A new species of the genus Festuca (Poaceae) from the southern Urals. Bot. Z. 83: 110–115. Kulikov, P.V. 2000. A new species of the genus Festuca (Poaceae) from north-western Caucasus. Bot. Z. 85: 124–127. Lehväslaiho, H., Saura, A., and Lokki, J. 1987. Chloroplast DNA variation in the grass tribe Festuceae. Theor. Appl. Genet. 74: 298–302. Malik, C.P., and Thomas, P.T. 1966. Karyotypic studies in some Lolium and Festuca species. Caryologia, 19: 167–196. Morgan, W.G., Thomas, H., and Lewis, E.J. 1988. Cytogenetic Studies of Hybrids between Festuca gigantea Vill., and Lolium multiflorum Lam. Plant Breed. 101: 335–343. doi:10.1111/j. 1439-0523.1988.tb00306.x. Page, R.D.M. 1996. TREEVIEW: An application to display phylogenetic trees on personal computers. Comput. Appl. Biosci. 12: 357–358. PMID:8902363. Pasakinskiené, I., Anamthawat-Jónsson, K., Humpreys, M.W., Paplauskiene, V., and Jones, R.N. 1998. New molecular evidence on genome relationships and chromosome identification in fescue (Festuca) and ryegrass (Lolium). Heredity, 81: 659–665. doi:10.1046/j.1365-2540.1998.00446.x. Peto, F.H. 1933. The cytology of certain intergeneric hybrids between Festuca and Lolium. J. Genet. 28: 113–156. Rolhlf, J.F. 1989. NTSYS.pc: numerical taxonomy and multivariate analysis system for the IBM pc microcomputer (and compatibles). Applied Biostatistics INC, Setavek, New York. Saint-Yves, A. 1922. Les Festuca (Eu-Festuca) de ĹAfrique du Nord et des Iles Atlantiques. Candollea, 1: 1–63. Saint-Yves, A. 1930. Aperçu sur la distribution géographique des Festuca (Subgen. Eu-Festuca). Candollea, 4: 146–165. Sanderson, M.J. 1989. Confidence limits on phylogenies: the bootstrap revisited. Cladistics, 5: 113–129. Siffelova, G., Pavelkova, M., Klabouchova, A., Wiesner, I., and Nasinec, V. 1997. Computer-aided RAPDs fingerprinting of accessions from ryegrass-fescue complex. J. Agric. Sci. 129: 257–265. Sneath, P.H.A., and Sokal, R.R. 1973. Numerical taxonomy. Freeman, San Francisco. Stace, C.A., Al-Bermani, A.K.K.A., and Wilkinson, M.J. 1992. The distinction between the Festuca ovina L., and Festuca rubra L. aggregates in the British Isles. Watsonia, 19: 107–112. Stammers, M., Harris, J., Evans, G.M., Hayward, M.D., and Forster, J.W. 1995. Use of random PCR (RAPD) technology to analyse phylogenetic reltionships in the Lolium/Festuca complex. Heredity, 74: 19–27. PMID:7852097. Stebbins, G.L. 1956. Taxonomy and evolution of genera with spe#. 2006 NRC Canada.

(14) de Nova et al. cial reference to family Gramineae. Evolution, 10: 235–245. doi:10.2307/2406009. Swofford, D.L. 1998. PAUP*. Phylogenetic analysis using parsimony (*and other methods). Version 4. Sinauer Associates, Sunderland, Massachusetts. Swofford, D.L., and Berlocher, S.H. 1987. Inferring evolutionary trees from gene frequency data under the principle of maximum parsimony. Syst. Zool. 36: 293–325. doi:10.2307/2413068. Taberlet, P., Gielly, L., Pauton, G., and Bouvet, J. 1991. Universal primers for amplification of three non-coding regions of chloroplast DNA. Plant Mol. Biol. 17: 1105–1109. doi:10.1007/ BF00037152. PMID:1932684. Terrell, E.E. 1966. Taxonomic implications of genetics in ryegrass (Lolium). Bot. Rev. 32: 138–164. Terrell, E.E. 1968. A taxonomic revision of the genus Lolium. USDA-ARS Tech. Bull. 1392: 1–65. Thompson, J.D., Higgins, D.G., and Gibson, T.J. 1994. CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22: 4673–4680. PMID:7984417. Torrecilla, P., and Catalán, P. 2002. Phylogeny of Broad-leaved and Fine-leaved Festuca Lineages (Poaceae) based on Nuclear ITS Sequences. Syst. Bot. 27: 241–251.. 1183 Torrecilla, P., López Rodrı́guez, J.A., and Catalán, P. 2004. Phylogenetic relationships of Vulpia and related genera (Poeae, Poaceae) based on analysis of ITS and trnL-F sequences. Ann. Mo. Bot. Gard. 91: 124–158. Tzvelev, N.N. 1976. Grasses of the Soviet Union. Zlaki SSSR. Translation for the Smithsonian Institution (1984), Amerind Publishing Co., New Delhi. Tzvelev, N.N. 1983. Grasses of the Soviet Union. Part II. Oxonian Press, New Delhi. Vos, P., Hogers, R., Bleeker, M., Rijans, M., van de Lee, T., Hornes, M., et al. 1995. AFLP: a new technique for DNA fingerprinting. Nucleic Acids Res. 23: 4407–4414. PMID:7501463. Watson, L., and Dallwitz, M.J. 1992. The grass genera of the world. C.A.B. International, Wallingford. pp. 1038. Williams, J.G.K., Kubelik, A.R., Livak, K.J., Rafalski, J.A., and Tingey, S.V. 1990. DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic Acids Res. 18: 6531–6535. PMID:1979162. Xu, W.W., and Sleper, D.A. 1994. Phylogeny of tall fescue and related species using RFLPs. Theor. Appl. Genet. 88: 685–690. doi:10.1007/BF01253971.. #. 2006 NRC Canada.

(15)

Referencias

Documento similar

Brain sections (frontal cortex) from ALS patient 11 and control 6 were observed with a confocal laser scanning microscope. Sections were obtained from fixed tissue

The seeds of both species are morphologically similar and generally slightly larger and wider in the Iberian populations ((0.9) 1.1 (1.25) mm long and Figure 3.

Márquez and Morrone (2003) distinguished a northern and a southern generalized track within the Mesoamerican dominion, based on an analysis of species of Homalolinus

In addition to creating a tool for DNA-based specimen identification, the Iberian dataset coupled with GMYC analysis permitted the first large-scale assessment of genetic patterns

examining the degree of effectiveness of inclusive education programs in private, semi-private and public centers for the neurodiverse L2 classroom; exploring students’

The within this study described upper Cretaceous fossil turtle and mosasaur remains originally belong to the private Mosbach collection and were found in 1915 within the Alsen

Log-ratio histograms comparing sheep/goat postcranial measurements from Late Roman Zornoztegi and other contemporary sites in the Iberian Peninsula. Only fused specimens are

Cuando estos bosques se degradan aparece junto al Quercus pyrenaicae, la Festuca elegans, dando lugar a la asociaci6n Festuco eleganti-Quercetum pyrenaicae, que