AMERICAN FERN JOURNAL QWi January-March 2013 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Microsite Factors and Spore Dispersal Limit Obligate Mycorrhizal Fern Distribution: Habitat Islands of Botrychium pumicola (Ophioglossaceae) Susan M. Roe-Andersen and Darlene Southworth Revealing a CrypUc Fei E. L. Peredo. M. Mendez-Couz, M. A. Revilla, and H. Femdndi rgh DNA Sequencing: Pityrogramma trif achys zeylanica (L.) Hook. in Burachapori Wildlife Sanctuary, J Foms nf Sniithem Africa: A Comorehensive Guide George Yatskievych The American Fern Society Council for 2013 Microsite Factors and Spore Dispersal Limit Obligate Mycorrhizal Fern Distribution: Habitat Islands of Botrychium pumicola (Ophioglossaceae) '^MISSOURI BOTANICAL JUL 2 9 2013 GARDEN LIBRARY ROE-ANDERSEN & SOUTHWORTH: MICROSITE FACTORS AND SPORE DISPERSAL IN BOTRYCHIUM PUMICOLA 3 without B. pumicola differ with regard to density of associated plant species, and B. pumicola spore bank; (3) B. pumicola dispersion is clumped; and (4) B. pumicola spore dispersal is predominantly local. Materials and Methods Plant natural history. — Leaves of Botrychium pumicola emerge after spring snowmelt, with peak emergence mid-July to mid-August (Hopkins et al, 2001; Roe-Andersen, 2010). Leaves reach heights of 1-6 cm above the soil surface, with 7-10 cm of the frond below ground. The stem with leaf primordia at the apex remains underground. The single leaf that matures each year is divided into a sterile trophophore and fertile sporophore. Reproduction occurs through spores, gemmae and intragametophytic selfing (Camacho, 1996; Camacho and Liston, 2001). Spores require a period of darkness for germination (Whittier, 1973; Johnson-Groh, 2002). Photoinhibition of spore germination ensures below-ground germination where the likelihood of sufficient moisture and proximity to mycorrhizal symbionts improve chances for colonization (Whittier, 2006). Botrychium species form obligate s3mibioses with arbuscular mycorrhizal (AM) fungi, predominately Glomus spp. (Kovacs et al., 2007; Winther and Friedman, 2007). Sporophytes of other Botrychium species remain under- ground for several years, relying on mycorrhizal partners to obtain fixed carbon from neighboring plants through an AM fungal network (Johnson-Groh, 1998; Winther and Friedman, 2007). Study sites. — Populations were studied at eight subalpine sites and one alpine site: Broken Top (BR), Mt. Bachelor (BA), and the Dome (DO) in the Deschutes National Forest (DNF); Liao Rock with two populations, Liao East (LE) and Liao West (LW), Grotto Cove (GR), Skell Head (SK), Cloud Cap with two populations. Cloud Cap North (CN) and Cloud Cap South (CS), and Dutton Ridge (DU) in Crater Lake National Park (CLNP), Oregon; and Shastina Cone (SH) in the Shasta-Trinity National Forest (STNF), California (Fig. 1, Table 1; Roe-Andersen, 2010). Montane populations at lower elevations were not included. All sites receive heavy snow accumulation (up to 3 m) with average minimum temperatures below freezing for eight months of the year (Hopkins et al., 2001; WRCDC, 2012). Average summer precipitation is low (3.1 cm total) and daily temperatures fluctuate widely (1.1-20.9°C). Plants are exposed to intense heat, cold, ultraviolet radiation, and strong winds. Sites in CLNP were located in exposed, wind-swept pumice fields surrounding the caldera (Fig. IB). Substrates consisted of volcanic “popcorn” pumice and mixed volcanic gravel from the Mt. Mazama pumice outfall c. 7700 years ago (Fig. 2A; Klimasauskas et al., 2002). In the DNF, the subalpine site (DO) in Newberry National Volcanic Monument, is a volcanic cinder cone approximately 11,000 years old, covered with 0.5-1 m of Newberry “popcorn” pumice from the Big Obsidian Flow eruption c. 1300 years ago (Sherrod et al., 1997). On the southeast flank of Broken Top Mountain, the BR site consisted of eroded andesite, dacite and rhyodacite from 100,000-year-old Broken Top AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) eruptions, mixed with pumice and volcanic gravel from the South Sister eruption c. 2000 years ago (Scott et al, 2003; Fig. 2B). The alpine site (BA) on the summit of Mt. Bachelor was in open, windswept volcanic rubble from Mt. Bachelor summit eruptions c. 12,000 years ago, and small amounts of pumice from the Mt. Mazama tephra plume c. 7700 years ago (Scott, 1990). The SH site was located on an exposed, windy, southwest flank near the upper end of Diller Canyon and consisted of andesitic boulder fields creating rock shelters for B. pumicola (Fig. 2C). This population was rediscovered 13 Jul 2009 (Eric Table 1. Botrychium pumicola study sites arranged by elevation in the Cascade Range of central and southern Oregon and northern California. CLNP, Crater Lake National Park; DNF, Deschutes National Forest; STNF, Shasta-Trinity National Forest. Size Elev (m) Study Site Location (ha) asl Aspect Grotto Cove CLNP Skell Head CLNP The Dome DNF Broken Top DNF Cloud Cap CLNP Liao Rock CLNP Dutton Ridge CLNP Shastina Cone STNF Mt. Bachelor DNF 2100 12° 2180 250° 2190 18° 2330 82°-140° 2400 185°-250° 2440 110°, 250-290° 2460 190° 2750 220° 2760 129°-172° ROE-ANDERSEN & SOUTHWORTH: MICROSITE FACTORS AND SPORE DISPERSAL IN BOTRYCHIUM PUMICOLA Fig. 2. Botrychium pumicola in the Cascade Range of central and southern Oregon and northern California. A-C, B. pumicola habitat. (A) Classic subalpine “popcorn” pumice, Liao Rock, Crater Lake National Park (CLNP). (B) Clay soil and few neighboring plants. Broken Top, Deschutes National Forest. (C) Andesitic rock shelter, Shastina Cone, Shasta-Trinity National Forest. (D) B. pumicola spore tetrad from soil sieving. Grotto Cove, CLNP. White, U.S. Forest Service, pers. comm.). Shastina Cone formed 9700- 9400 years ago as a subsidiary andesite cone on the west flank of Mt. Shasta (Christiansen, 1990). Sampling design— At each site, plots with B. pumicola were paired with adjacent plots without B. pumicola, but similar in slope, aspect, elevation, and vegetation. Because environmental variables at fine spatial scales influence the distribution and abundance of ferns (Karst et al, 2005), paired sites were located within 5-10 m of plots where B. pumicola occurred. Paired sites were chosen in areas where no B. pumicola had been observed for the past three consecutive years. At each site, 10 1-m^ plots (5 with B. pumicola) were chosen at random and sampled for environmental variables (Elzinga et al., 1998). At LE and LW with larger populations, 10 plots with B. pumicola and 10 without were sampled. Populations were mapped using a Garmin GPSmap 60CSx (Garmin International, Olathe, Kansas). Maps were created in ArcGIS v. 9.3 (ESRI, Redlands, California). Elevation was determined with the GPS unit and field checked with USGS 7-minute topographic maps (Crater Lake East, Crater Lake West, Broken Top, Mt. Bachelor, East Lake, and Mt. Shasta topographic ROE-ANDERSEN & SOUTHWORTH: MICROSITE FACTORS AND SPORE DISPERSAL IN BOTRYCHIUM PUMICOLA properties in paired plots Matched Pairs test for alpir 1 B. pumicola present (P) or absent (A) ad subalpine populations in the Cascade Population Density (g/L) OM % HgO % Gravel % Sand % Silt % Clay % pumicola P APAPAP A PAPAPA LW 371.3 505.1 2.8 1.5 4.3 1.0 23.7 17.9 91.3 93.8 7.5 3.8 1.3 2.5 LE 371.7 559.6 2.5 0.9 11.9 8.1 20.0 35.4 91.3 97.5 5.0 1.3 3.8 1.3 BR 405.6 508.4 2.8 2.7 2.5 1.9 34.7 49.1 83.8 82.5 12.5 13.8 3.8 3.8 GR 532.4 460.4 1.0 0.9 1.0 1.1 25.1 25.3 92.5 97.5 3.8 1.3 3.8 1.3 DU 499.7 542.1 1.8 1.5 4.3 7.5 13.7 31.3 92.5 92.5 5.0 6.3 2.5 1.3 SK 357 !o 483’6 l!o 1.0 5.1 9.4 21.7 23.0 96.3 90.0 2.5 6.3 1.3 3.8 BA 494.7 493.7 1.9 1.9 9.0 11.3 20.4 20.3 86.3 82.5 10.0 10.0 3.8 3.8 P 0.03 “ 0.07 0^68 005 089 0.48 0.80 ^ Significant at P < .05 percent sand, silt, or clay did not differ between paired plots with and without B. pumicola; soil density was lower in plots with B. pumicola in seven out of nine pairs (Wilcoxon Matched Pairs, P = 0.032; Table 3). Paired plots did not differ in pH, phosphate, magnesium, sodium, calcium, nitrate, ammonium (Table 4), or in percent organic matter (Table 3). Potassium was higher in plots with B. pumicola (Wilcoxon Matched Pairs, P = 0.024; Table 4), although similar values of K were found in some plots without B. pumicola. NMS ordination of all abiotic factors showed a high degree of overlap between plots with and without B. pumicola. No differences in abiotic factors were observed using Multi-Response Permutation Procedure (MRPP) for all populations (P = 0.180). When outliers (values >2 standard deviations from the mean) were removed, differences in abiotic factors were significant (MRPP , P = 0.049; Fig. 3). Biotic /actors.— Thirty-five plant species occurred in plots with B. pumicola and 32 in plots without B. pumicola (Table 5). The most abundant species was Raillardella argentea, and the most frequent were Carex brewed and Lupinus lepidus. No plant species, except B. pumicola, was present at all sites. Species- area curves indicated that plots with B. pumicola were similar in species richness of associated plants (first order jackknife estimate, 44.9) to plots without B. pumicola (first order jackknife estimate, 42.5). For an accumulated eight sites, species-area curves of plots with and without B. pumicola had 32 associated species. The density of associated plants showed a high range of variability in plots with and without B. pumicola (Table 5). NMS ordination of all plant associates showed differences between plots with and without B. pumicola (MRPP, P = 0.015). When abundance values were relativized to minimize the effect of plots with disproportionately large 11 ^ AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) Table 4. Soil pH and soil nutrients in paired plots with B. pumicola present (P) or absent (A) compared by Wilcoxon Matched Pairs test for alpine and subalpine populations in the Cascade Range of central and southern Oregon. P K Mg Na Ca NO3 NH4 Population pH (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) Botrychium pumicola PAPAPAP A PA P A PAPA P 0.87 0.95 0,02 “ 0,59 0,09 0,59 0,40 0.12 ® Significant at P < 0.05 ROE-ANDERSEN & SOUTHWORTH: MICROSITE FACTORS AND SPORE DISPERSAL IN BOTRYCHIUM PUMICOLA numbers, plots with and without B. pumicola still differed (MRPP, P = 0.007). When associated plant species occurring in fewer than 5% of plots were removed, plots with and without B. pumicola differed in community composition (MRPP, P = 0.018; Fig. 4). NMS ordination of all plant associates per m^ with singletons removed showed no differences between sites with and without B. pumicola (MRPP, P = 0.875). NMS ordination of abiotic factors plus plant associates with singletons removed showed no difference between plots with and without B. pumicola (MRPP, P = 0.875). Nearest-neighhor distances between B. pumicola conspecifics ranged from 9.1 to 71.6 cm with clumped and random dispersions (Table 6). In four populations, NNIs were below 1.00 indicating clumping; in five populations, NNIs of 1.10-1.33 indicated a random distribution. Nearest neighbor distances between B. pumicola individuals and the nearest associated plant species ranged from 1.6 to 15.4 cm (Table 6). Nearest neighbor indices between B. pumicola individuals and the nearest other plant species showed a clumped dispersion (NNI < 1.00) in all populations except at BR which had an NNI of 2.04 indicating a uniform dispersion (Table 6). Spore tetrads of Botrychium pumicola were found in soil samples from plots with B. pumicola (Fig. 2D) at LE (20); LW (1); SK (4); CN (1); CS (2); BR and BA (1 each) and in plots without B. pumicola at GR (3); BA (1). No B. pumicola spores were found at DU. One single spore (not in tetrad) was found during sampling of BR [B. pumicola present) soils. Spore tetrads were found on spore traps at distances up to 10 m. At SK, spores lodged on slides 10 m from the source plant with most spores lodging within 30 cm. At DU spores lodged on slides within 1 m from the source plant. At LNE spores lodged on slides 10 m from the source plant with most spores lodging within 5 m (Fig. 5). Discussion Microsite factors— In paired plots with and without B. pumicola, K levels were higher in the plots with B. pumicola. Potassium is essential for plant growth, development and regulation of stomata, with the role of K in photosynthesis becoming more important at higher CO2 levels and light intensity (Larcher, 2002; Cakmak, 2005). A higher K level may help B. pumicola regulate water loss in habitats where plants are subject to low summer precipitation, intense UV radiation, temperature extremes, and desiccating winds. Potassium is essential for spore germination in ferns (Miller and Wagner, 1987). Lower K levels may decrease gametophyte success in habitats that appear suitable for B. pumicola. Although several Botrychium species are associated with calcareous habitats (Cayouette and Farrar, 2009; Zika et al, 2002), B. pumicola is not a calcicole, and calcium did not differ between sites with and without B. pumicola. Soil density was lower in plots with B. pumicola. Greater porosity may enable B. pumicola spores to move down the soil column. This may assist spores in finding the darkness necessary for germination and the AM fungal partners for continued growth (Johnson-Groh, 2002; Whittier, 2006). ROE-ANDERSEN & SOUTHWORTH: MICROSITE FACTORS AND SPORE DISPERSAL IN BOTRYCHIUM PUMICOLA f JOURNAL: VOLUME 103 NUMBER 1 (2013) sisisiiiii ):21-26 Pellaea flavescens, a Brazilian Endemic, is a Synonym of Old World Pellaea viridis Jefferson Prado Institute de Botanica, C.P. 68041, 04045-972, Sao Paulo, SP, Brazil, e-mail: jprado.01@uol.com.b Eric Schuettpelz Department of Biology and Marine Biology, University of North Carolina Wilmington, 601 SoutI College Road, Wilmington, NC, 28403-5915, USA, e-mail; schuettpelze@uncw.edu Regina Y. Hirai Institute de Botanica, C.P. 68041, 04045-972, Sao Paulo, SP, Brazil, e-mail: regina.hirai@gmail.cor Alan R. Smith University Herbarium, 1001 Valley Life Sciences Building #2465, University of California, Berkeley, CA, 94720-2465, USA, e-mail: arsmith@berkeley.edu eilanthoid, rbcL, phylogeny, ferns . The morphology of these two lym of the older P. viridis. This in Brazil or, alternatively, the Pellaea flavescens Fee, described from Rio de Janeiro and recently found in Sao Paulo, Brazil (Prado and Hirai, 2011), is morphologically similar to Pellaea viridis (Forssk.) Prantl from Africa, Madagascar, the Comoros, and the Mascarenes (Moran and Smith, 2001). Nonetheless, Tryon and Tryon (1982) placed P. flavescens in Pellaea section Ormopteris, a small group of species mostly endemic to central Brazil, while referring P. viridis to Cheilanthes. Several recent phylogenetic analyses (Gastony and Rollo, 1995; Kirkpatrick, 2007; Prado et al, 2007; Schuettpelz et al, 2007; Eiserhardt et al, 2011) have demonstrated that both Pellaea and Cheilanthes, as defined by Tryon and Tryon (1982), are polyphyletic. These studies also showed that P. viridis and Pellaea section Ormopteris are not especially close relatives. Previously sampled members of Pellaea section Ormopteris nested, albeit with poor support, within a large Doryopteris clade, sister to Doryopteris section Lytoneuron (Prado et al.. 2007). Pellaea viridis was resolved (support values were not reported) in a separate clade of primarily African species variously assigned to Cheilanthes and Pellaea (Eiserhardt et al., 2011). The morphological similarity of Pellaea flavescens and P. viridis is at odds with their presumed placement in separate, distantly related clades. However, P. flavescens itself has not been sampled in any previously published molecular phylogenetic study, leaving the question of its evolutionary position unanswered. Here, we examine the relationships of this species, and the taxonomic implications thereof, through an analysis of plastid rbcL sequences. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) A plastid rbcL sequence was obtained from a recent collection of Pellaea flavescens following established protocols (Schuettpelz and Pryer, 2007). A dataset was then assembled, consisting of this sequence, 22 exemplars from clade D of Eiserhardt et al. (2011) obtained from GenBank, and an additional published sequence of P. viridis (Table 1). These sequences were manually aligned using Mesquite version 2.75 (Maddison and Maddison, 2011) and subsequently analyzed using MRBAYES version 3.2.1 (Huelsenbeck and Ronquist, 2001; Ronquist and Huelsenbeck, 2003). Regions at the ends of the alignment containing copious amounts of missing data were excluded, leaving a dataset of 1143 characters to which a GTR+G model of sequence evolution was assigned. Our Bayesian analysis comprised two independent runs, each utilizing four chains of two million generations. Trees were sampled every 1000 generations. To identify when the runs had reached stationarity, the standard deviation of the split frequencies between the two runs was examined, and the output parameter estimates were plotted using Tracer 1.5 (Rambaut and Drummond, 2009). Based on these convergence diagnostics, the first 500 trees were (very conservatively) excluded from each analysis before obtaining a majority rule consensus phylogeny with clade posterior probabil- ities. The resulting tree was rooted with Trachypteris pinnata (Hook, f.) C. Chr. Results and Discussion The rbcL sequence obtained from our Brazilian Pellaea flavescens sample was identical to two Old World exemplars of P. viridis (EF452147 and GU935494; Table 1). These sequences compose a well-supported (posterior probability, PP = 1.00) clade in our recovered phylogeny (Fig. 1) that is, in turn, robustly supported (PP = 1.00) as sister to a third exemplar of P. viridis. Pellaea section Ormopteris, here represented by P. gleichenioides (Gardn.) Christ and P. pinnata (Kaulf.) Prantl, is not closely related to this clade. Instead, it is well-supported as sister to Doryopteris section Lytoneuron (PP = 1.00), consistent with the findings of Prado et al. (2007). Thus, the placement of P. flavescens in Pellaea section Ormopteris by Tryon and Tryon (1982) is not supported by our molecular analysis. The phylogenetic position of Pellaea flavescens as closely related to P. viridis, is supported by morphology. In fact, these taxa were recognized as a “species pair” by Moran and Smith (2001), who noted only slight structural differences. Here, however, we find P. flavescens to be nested within P. viridis, having an rbcL sequence identical to those of two P. viridis individuals (Fig. 1). Although rbcL is among the more slowly evolving plastid genes, a recent study of candidate barcodes (Li et al., 2011) found it to be comparable to a much more rapidly evolving gene [matlO in its discriminatory power. Upwards of 40% of the species in the Cheilanthes marginata group (now recognized as the genus Gaga; Li et al., 2012) could be separated based on rbcL alone (Li et al., 2011). Unfortunately, but not surprisingly, discriminatory power was weakest fijiiiljiiiillllliiiiliil 1 1 f I ! •3 I I PRADO ET AL.: PELLAEA FLAVESCENS IN BRAZIL therefore superfluous and cannot be used unless conserved or sanctioned, according to the International Code of Nomenclature for algae, fungi, and plants (McNeill et al., 2012). In any case, the figure cited for P. bongardiana in the Flora Brasiliensis (tab. 55, fig. II) is wrong, with the illustrated plant representing Pellaea riedelii Baker. Pellaea viridis is commonly cultivated (Hoshizaki and Moran, 2001), and escape from cultivation, followed by naturalization, has been documented (e.g., in Hawaii, by Palmer, 2003; in eastern Australia, by Bostock, 1998; and in Florida, see Atlas of Florida vascular plants: http://florida.plantatlas.usf.edu/). Thus, it may be that the existence of this species in Brazil represents yet another introduction from cultivation into the wild. Despite this apparently reasonable possibility, P. viiidis is not currently cultivated in Brazil, so far as we are aware, and the Brazilian populations generally occur in undisturbed environments. Furthermore, while the first Brazilian collection {Glaziou 2473) we have seen dates from 1867, the current distribution of the species in Brazil, nearly 150 years later, remains highly restricted (Prado and Hirai, 2011). This stands in stark contrast to what has been encountered for most other introduced species. For example, despite a much more recent introduction into the New World of the Old World fern Thelypteris dentata (Forssk.) E. P. St. John (the earliest New World collection dates from 1908; Strother and Smith, 1970), this species is today one of the most common ferns in many areas of the Neotropics (Mickel and Smith, 2004). Although the distribution of the apparently introduced Marsilea hirsuta in the Azores is even more restricted than that of P. flavescens in Brazil (Schaefer et al., 2011), the absence of Marsilea in historical collections from the Azores would point to a very recent introduction of the genus there. It is possible, then, that the populations of P. viridis in Brazil are native, originating via long distance spore dispersal from the Old World. Although uncommon, such a disjunction is not unheard of; Moran and Smith (2001) documented 27 pteridophyte species that naturally occur in both the Neotropics and Africa. Moving forward, it is clear that the Pellaea viridis complex is in need of further examination (Moran and Smith, 2001; Eiserhardt et al., 2011), not only to clarify species boundaries but also to identify the affinities of the Brazilian and other New World collections. Only through a more thorough analysis will it be possible to uncover the true origin of this lineage in Brazil. This research was funded in part by the National Science Foundation (awards DEB-0717398, DEB-0717430, DEB-1145614, and DEB-1145925), as well as by Conselho Nacional de Desenvolvi- mento Cientifico e Tecnoldgico (301157/2010-3) and Fundagao de Amparo h Pesquisa do Estado de Sao Paulo (2011/07164-3). Laboratory assistance was provided by Alex Davila. Two anonymous reviewers and the associate editor provided helpful comments on the manuscript. Literature Cited Atlas OF Florida Vascular Plants, http://florida.plantatlas.usf.ee Bostock, P. D. 1998. Pellaea. Pp. 266-269 in Flora of Australia, iu/ Accessed 25 July 2012. V. 48. ABRS/CSIRO, Melbourne. I (2013) Mating System in Blechnum spicant and Dryopteris affinis ssp. affinis Correlates with Genetic Variability E. L. Peredo Department of Ecology and Evolutionary Biology, BioPharm 319, University of Connecticut, U-3043 75 North Eagleville Road, Storrs, CT 06269-3043, USA M. MEndez-Couz Laboratory of Neuroscience, Faculty of Psychology, Dpt. Psychology, PI Feijoo s/n 33800 Oviedo, M. A. Revilla^ and H. FernAndez Dpt. Biologia de Organismos y Sistemas, Universidad de Oviedo, C/ Catedrdtico R. Uria s/n 33071 Oviedo, Spain Abstract.— The monilophytes Blechnum spicant (L.) Sm. and Dryopteris affinis ssp. affinis (Lowe) Fraser-Jenkins show different reproductive strategies under in vitro conditions. While B. spicant exhibits asexual and sexual reproduction, with an antheridiogen system promoting outcrossing, D. affinis ssp. affinis reproduces only asexually through apogamy. Individuals of several populations of these species, collected in Principado de Asturias (Spain), were analyzed to test the influence of their mating system in the genetic variability displayed by each species. This study shows that the genetic diversity assessed in populations of each species collected in situ is in concordance with differences among localities. This result indicates high fixation of the detected alleles within each locality, as expected for a clonal reproductive system. In the sexual species B. spicant the genetic diversity was higher. Our results confirm the importance of reproduction system in the genetic diversity present in populations of these fern species making essential to consider the definition Key Words.— AFLP, Blechnun spicant, clonal growth, Dryopteris affinis, sexual reproduction The formation of a new fern sporophyte can be achieved through asexual or sexual pathways. During asexual reproduction, cells other than gametes develop a sporophyte through an asexual process without meiosis or fertilization called apogamy. The fern complex Dryopteris affinis (Lowe) Fraser-Jenkins includes diploid and triploid subspecies, all of them apogamic (Fraser-Jenkins, 1980). Apogamy has been reported to be a way of escaping hybrid sterility by alloploids (Fraser-Jenkins, 1980; Chao et ah, 2012). On the other hand, the mating system of the fern species Blechnum spicant (L.) Sm. implies the development of gametophytes that produce male and female gametes in the sexual organs, archegonia and antheridia, which form a sexual embryo through fertilization. * Corresponding author: arevilla@uniovi.es AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) Q 15/1/2009 12/3/2009 S 1 1 Ifl fill |S| Jill Is 1 2 2 2 - - - 3° 25' 58.327" N 3° 23' 29.706" N 38' 51.171" W “ 23' 50.639" N 41' 35.684" W ° 26' 0.615" N 56' 30.064" W “ 27' 52.154" N ° 23' 32.083" N 1 UTM coordini (30T) 4810718 0366566 4805620 0362878 4806340 0342852 4810789 0319177 4814827 0362944 4805766 Sampling site Road to Purdn Pur6n river, road to 1 ! B 1 B 2 B 3 D 8 D 9 D 10 1 1 1, 1 II 1 PEREDO ET AL.: BLECHNUM AND DRYOPTERIS MATING SYSTEM AND GENETIC VARIABILITY Hg. 1. Lett: Bar plot ot the proportion ot an individual s ot Blechnum spicant genome belonging to one or other cluster inferred by the STRUCTURE analysis. Cluster 1 is represented in green and includes individuals from sampled sites B2 (Novales) and B3 (Puron). Right: Plot of AK for each K value calculated as described in Evanno et al. (2005). calculated for Blechnum and 0.26 (0.07-0.33) for Dryopteris. HE and I calculated for each population were: B1 0.20, 030; B2 0.15, 0.22; B3 0.18, 0.28; D8, 0.13, 0.18; D9 0.05, 0.07, and DlO 0.02, 0.04 (consult Table 3 for a detailed list). To test the existence of significant differences between the HE, I, and percentage of polymorphism (%P) values calculated for each species, the Mann-Whitney U test and Krustal-Wallis test were applied. Firstly, the suitability of the data to these tests was corroborated by checking whether the data fit to a normal distribution. To eliminate the possibility of variation due to other causes than the species as a source for the differences in the genetic variation (e.g. primer combinations chosen or population related variation) several tests were performed prior the comparison species to species. No significant differences were detected in any of the comparison within D. affinis ssp. affinis (data not shown), whereas for B. spicant statistically significant differences (%P p = 0.03) were found only in the percentage of polymorphism calculated for each primer combination. No significant differences were found using the same approach (Krustal-Wallis test) to the values calculated within each species. The existence of differences between the values of heterozygosity (HE), Shannon index (I), and percentage of polymorphism (%P) in D. affinis ssp. affinis and B. spicant was tested using two-tailed Mann-Whitney U test. Significant differences were detected in each comparison: HE p<0.0001 (Monte Carlo p<0.0001); I p<0.0001 (Monte Carlo p<0.000l); and %P p<0.0001 (Monte Carlo p<0.000l). Genetic structure in Blechnum spicant and Dryopteris affinis.— The Bayesian-clustering was applied to the datasets as implied in the STRUCTURE program assuming an admixture model, with the assumption of the possibility of mixed ancestry of the individuals. In B. spicant, Evanno’s AK indicate the possibility of a K=2 grouping as the value for AK =2 was 25.4 and low values were calculated for AK >2 (<0.5) (Fig. 1). In each of the 10 replicates calculated for K=2 individuals from the locality B1 (Nueva) grouped in cluster 1 with a high probability (>0.904) while individuals from B2 (Novales) and B3 (Road to Puron) belonged to cluster 2 (probability over 0.928 and 0.958, respectively). This high degree of similarity among the individuals from populations B2 and B3 was also confirmed by the detected grouping in the PEREDO ET AL.: BLECHNUM AND DRYOPTERIS MATING SYSTEM AND GENETIC VARIABILITY populations. The high Fst values also corroborate the high local fixation of variability (D8, 0.538; D9, 0.628; DlO, 0.685). This high degree of similarity among the individuals from the same sampling site was confirmed by the high bootstrap support (>90%) of the different branches in the dendrogram calculated with PAST (Fig. 3) or the grouping in the PCoA (data not shown). Discussion The genetic diversity calculated for the ferns Blechnum spicant and Dryopteris affinis ssp. affinis is in agreement with the reproductive mechanisms observed previously when cultured in vitro (Fernandez and Revilla, 2003). Also, our results indicate the importance of small-scale sampling. A relatively high percentage of polymorphism was estimated with AFLP for two species of ferns (81.38% for 5. spicant and 54,73% for D. ajfinisssp. affinis) in a small geographical area in the Principado de Asturias; no populations were found more than 45 km apart. However, when the distribution of this variation was analyzed in relation to the sampling sites, the polymorphism was drastically reduced to at least half and even one tenth in the case of D. affinis (28.66% (D8), 11.43% (D9) and 6.10% (DlO). In the case of B. spicant a less notable effect was observed, with recalculated polymorphism ranging between 45.75% (B2) and 60.03% (Bl). These changes in the percentage of polymor- phism are explained by high genetic fixation, making the variation in each I Journal 103{l):40-48 I Revealing a Cryptic Fern Distribution Through DNA Sequencing: Pityrogramma trifoliata in the Western Andes of Peru AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) Materials and Methods Plants were obtained in the field, from a site in the Department of Lima, Province Canete, in the Mala river basin, above the town of Calango (12°31'17.54"S, 76°30'1.11"W), at near 400 m elevation (Fig. lA-B; Fig. 2), where climate conditions are hyper-arid (Rundel et al., 2007). They were present as small, undeveloped gametophytes and very young sporophytes. Plants and their surrounding soil were collected and maintained in a plastic- covered container for nearly 5 months, from August 2011 until February 2012. Two samples were taken in November 2011 to prepare herbarium vouchers and to extract DNA. Plants did not survive past February 2012. DNA was isolated from dry tissue using the MP FastDNA® SPIN Kit and FastPrep® instrument (MP Biomedicals LLC, Solon, OH, USA). Portions of two plastid loci — rbcL and the trnG-trnR intergenic spacer (henceforth, trnG-R ) — in 21 jiL reactions were amplified following established protocols (Rothfels et al. 2013). PGR was performed with an initial four-minute denaturation step (95“C), followed by 35 cycles of 30 seconds denaturation (95°C), 30 seconds elongation (40°C), and one minute elongation (71°C). The reaction was concluded with a final elongation step at 71°C, for 10 minutes. The rbcL amplifications used the primers ESRBCLlF and ESRBCL654R (Schuettpelz and Pryer, 2007), and the trnG-R reactions used TRNGlF and TRNG63R (Nagalingum et al., 2007). PGR products were purified using Shrimp AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) LE6n ET AL.: cryptic distribution of PITYROGRAMMA in western PERU - ^ BO, ^ ^ Bommeria hispida (EU268671) Cheilanthes alabamensis (EU268672) Pellaea truncata (EU268714) Notholaena trichomanoides (EU268710) a triangularis subsp. maxonii (EU268716) -Cryptogramma crispa (EU268687) of Cystopteris supported the latter, as did the presence of toothed margins, although Cystopteris fragilis has veins ending in a sinus rather than in a tooth. A third possibility was a member of the Pteridaceae, either Anogramma leptophylla (L.) Link or Pityrogramma chaerophylla (Desv.) Domin, but in both these species the lamina is gradually reduced, and in the former, the rhizome does not bear scales, but hairs; additionally, the latter has not yet been JOURNAL: VOLUME 103 NUMBER recorded in Peru. Another possibility among Pteridaceae, was either Pityrogramma calomelanos (L.) Link or P. trifoliata (L.) R. M. Tryon; the former has a leaf apex that is gradually reduced and it bears glandular hairs on the laminae, whereas in the latter the terminal segment has a similar shape to the lateral pinnae, and also, for some individuals, laminar glandular hairs can be absent. BLAST searches supported the identity of the mystery fern as a species of Pteridaceae, apparently closest to Pityrogramma. In both datasets [rbcL and trnG- R], the mystery fern matches closely with accessions of Pityrogramma (Fig. 4 A- B). In the more densely sampled rbcL dataset, the mystery sequence is particularly closely related to Pityrogramma trifoliata (Fig. 4B); these two sequences differ by a single substitution across the 607 aligned sites. Pityrogmmma trifoliata is sometimes treated as Trismeria trifoliata (L.) Diels (e.g. Zuloaga et al., 2008), the only commonly recognized member of that genus, and is morphologically highly distinctive within Pityrogramma (at least as mature sporophytes!). The mystery plant then, while not exactly identical in sequence to the published rbcL sequence from P. trifoliata, is almost certainly that species. Discussion The discovery of Pityrogramma trifoliata in river beds in the xeric belt of the Andean foothills is a novelty due to the underexplored microhabitat it inhabited and the stark bareness of the surrounding hills (Fig. lA). Watkins et al. (2007) suggested that for terrestrial ferns, soil disturbance is related to gametophyte establishment success and growth. This also appears to be the case for our findings, as microsites among the riparian cobble provide shade and constantly humid soil essential for gametophyte colonization and development. In turn, these sites are likely highly unstable due to seasonal fluctuation in river discharge. Pityrogramma trifoliata is known from an altitudinal range of 50 to 2300 m in Peru, but no previous collections are known from the xeric elevational belt itself. Populations of P. trifoliata are found in sparse clusters where humidity is constant, such as margins of waterfalls and irrigation channels, and from which a few collections of this species in western Peru have been previously reported. This species and other western Andes ferns are likely characterized by higher dispersability, resilience, and based on this study, the capacity to survive as a gametoph 3 ^e. If we had been restricted to using morphological data for identification, we would have required more time and cultivation efforts for the plants to express those morphological characters associated with P. trifoliata. An important note as to the state of our knowledge of the fern flora is the scarcity of collections for ferns that are considered to be common. A more complete set of data on the natural history of ferns is also needed, especially those verifying and updating information on gametophytes, such as recent work within a phylogenetic framework in Pteridaceae (e.g. Gabriel y Galan, 2011; Johnson et al., 2012). AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) 0 leptosporangiate specie; SCHUETTPELZ, E. and K. M. Fryer. 2007. Fem phylogeny infe and three plastid genes. Taxon 56:1037-1050. SCHUETTPELZ, E., H. ScHNEiDER, L. HuiET, M. D. WiNDHAM and K. M. Fryer. 2007. A molecular phylogeny of the fem family Fteridaceae: Assessing overall relationships and the affinities of previously unsampled taxa. Molec. Fhylogen. Evol. 44:1172-1185. SwoFFORD, D. L. 2002. FAUF*: Fhylogenetic analysis using parsimony (*and other methods). Tryon, R. M. 1960. The ecology of Feruvian ferns. Amer. Fem J. 50:46-55. Vasco, A. 2011. Taxonomic revision of Elapboglossum subsection Muscosa (Dryopteridaceae). Blumea 56:165-202. Watkins Jr. J. E., M. K Mack and S. S. Mulkey. 2007. Gametophyte ecology and demography of epiph)dic and terrestrial tropical ferns. Amer. J. Bot. 94:701-708. Yansura, D. G. and B. J. Hoshizaki. 2012. The tree fem Highland Lace is a cultivar of Sphaeropteris cooperi. Amer. Fem J. 102:69-77. ZuLOAGA, F. O., O. Morrone, M. j. Belgrano, C. Marticorena and E. Marchesi (eds.) 2008. Catalogo de las Plantas Vasculares del Cono Sur (Argentina, Sur de Brasil, Chile, Paraguay y Uruguay). Monogr. Missouri Bot. Card. 107(l):l-983. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER I f I ! i 1 I I I I ii ill ssss r rii si ill 2sii i" iii ill si ill ills ill is ill ill iPi immim IpSiiiiiii I immUht liPiipilH Jsiiiiiil's liiiiiiiiiil Jiiiisiiiil ipiiiipj 3i ill soli +1 +i +1 +1 +1 +1 +1 +1 +1 " r ill liliii !i ? ilisilillll I Ipiiiipij I AMERICAN FERN JOURNAL: VOLUME 103 NUMBER Fig. 1. A. Type specimen of Polypodium gamerianum Vareschi [Vareschi &- Magdefrau 6839, VEN); B. Abaxial surface of largest fronds; C. Ceradenia intonsa, mixed with C. gameriana-, D. Fragments of C. intonsa which have been excluded, and contained in the packet placed in the lower left comer of type. Neotropics, the anhydathodous genera are Ceradenia L. E. Bishop (Bishop, Amer. Fern J. 78:1-5. 1988), Enterosora Baker (Bishop and Smith, Syst. Bot. 17:345-362. 1992), and Zygophlebia L. E. Bishop (Bishop, Amer. Fern J. 79:103-118. 1989b), which form a monophyletic clade according to the most recent phylogenetic hypotheses (Ranker et al. 2004; Sundue et al. 2010). Nearly all other remaining Neotropical grammitid genera have adaxial hydathodes, except possibly a few species of the Cochlidium (A. R. Smith, com. pers.). It is clear that P. gamerianum does not belong in Enterosora (which has laminar setae and lacks glandular paraphyses among the developing sporangia; Bishop and Smith, 1992). Zygophlebia, usually has areolate venation (regular anastamoses), and is also characterized by having glandular paraphyses, but these trichomes are brown, and thickly viscid, with sori adhering in a single sticky mass (Bishop, 1989b). In contrast, the type of P. gamerianum has larger fronds and a coriaceous texture, free veins, laminae covered with white to yellowish, sessile, spherical wax-like glandular hairs on both surfaces, and sori that have wax-like gland-bearing receptacular paraphyses. These last charac- ters are typical in species of Ceradenia subg. Ceradenia (Bishop, 1988). Additionally, these gland-bearing paraphyses show no tendency to adhere to SHORTER NOTES each other or to the sporangia, as observed by Bishop (1989b) in Ceradenia, and by Rakotondrainibe and Deroin in Zygophlebia (Taxon 55:145-152. 2006). Examination of the type of P. gamerianum shows that it should be placed in the genus Ceradenia. However, another problem arose in a more detailed examination of the type specimen. The morphology of the largest fronds on the sheet match the original diagnosis, but not the rhizome previously attached to the type (Fig. lA; see the dark spot on the lower left, and below the eight largest fronds and retained in a packet). It does not belong with the eight fronds, and thus the type of P. gamerianum is a mixed collection with another species of the genus Ceradenia (Fig. IC). The rhizome, which exhibits dorsiventral symmetry, is attached to a few fragmentary fronds, which have a herbaceous texture, and are provided on both sides of the laminae, rachis, costae, and margins with cylindrical, castaneous setae, 1.0-1. 5 mm long. Furthermore, setae are numerous abaxially, and at the apex of petioles, they are intermixed with some branched hairs. Fronds lack spherical wax-like glandular hairs on both surfaces, but they do have white, wax-like gland- bearing paraphyses in the sori. Other features of this packet-fragment are: scales of rhizome dark brown, shiny, linear-lanceolate, to 6 mm long, ca. 0.6 mm wide, bearing on the margins abundant hyaline to yellowish setulae; petioles black to castaneous at apices, to 6.5 cm long; rachis sclerenchyma not exposed abaxially; pinnae 2 cm long, ca. 6 mm wide, basally sursumcurrent, margins entire to slightly crenate (when fertile), slightly auriculate acrosco- pically, little dilated basiscopically, and with apices rounded. The traits of the rhizome and the laminae suggest this other species is a member of Ceradenia subg. Filicipecten L. E. Bishop (Bishop, 1988), and was collected with and included in the original description (in part), accidentally mixed by Vareschi in the type specimen of P. gamerianum. The features described are consistent with the species currently known as Ceradenia intonsa L. E. Bishop ex A. Leon & Mostacero in Leon (Flora de Colombia 29:38. 2012). Another specimen of C. intonsa, Vareschi & Magdefrau 6836 [YEN] was collected from the same locality, suggesting that both species grow together, which possibly resulted in the mixed collection. Because the rhizome does not belong to P. gamerianum, whose description is based entirely on the fronds that are in the sheet and not in the packet, it is necessary to designate a lectotype on the same specimen (McNeill et al. International Code of Nomenclative for algae, fungi, and plants:Art. 9.14. 2012), excluding the rhizome and frond fragments contained in the packet (Fig. ID). This is done simultaneously with the following new combination also required, and adding a more detailed description of the largest fronds, constituting the type: Ceradenia gameriana (Vareschi) Mostacero, comb. nov. Polypodium gamerianum Vareschi, Acta Bot. Venez. 1(2):117, f.l6. 1966. Type: Venezuela. Merida: Laguna Los Anteojos, selva del Pico Bolivar, [approx. 8°32'N, 71°03'W], 4100 m, 3 April 1958, V. Vareschi & K. Magdefrau 6839 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 1 (2013) (lectotype VEN!, here designated, excluding the rhizome, leaf portions still attached to it, and other fragments of similar texture in the packet located in the lower left comer, and the fig. 16C on the original description; ever 3 dhing excluded corresponding to another species, C. intonsa). Grammitis gameriana (Vareschi) Duek & Lellinger, Amer. Fern J. 68:120. 1978. Plants epipetric; rhizomes unknown; fronds 28-52 cm long, determinate, monomorphic, apparently erect, lacking setae throughout, covered with minute wax-like glandular hairs on both surfaces, these more dense abaxially, to 0.2 mm long, sessile, adpressed, branched, comprising 3-4 spherical glands less than 0.1 mm at the apex, appearing white to pale yellow; petioles 8-20 cm X ca. 0.5 mm, brown-stramineous, dull, glabrous; laminae coriaceous, 1- pinnatifid to -pinnatisect, narrowly linear, gradually reduced proximally, 21- 50 X 1.2— 3.5 cm, with 50—80 pairs of pinnae; rachis sclerenchyma slightly exposed and rounded abaxially, darkened, ca. 1 mm wide; pinnae 8-17 X 0.15—0.3 mm, set 45—90° to the rachis, each spaced 1—2 times its width, linear- lanceolate, oblong to linear, the bases decurrent basiscopically, straight or nearly so acroscopically, pinna margins entire (sterile) to undulate or slightly crenate (fertile), the apices rounded (sterile) to subacute (fertile); veins free, pinnate, branched once, not visible without transmitted light; hydathodes lacking adaxially; sori brown, round, 8-12(-18) per pinna, medial to submarginal, 0.8— 1.2 mm wide, with wax-like gland-bearing receptacular paraphyses, these whitish, like those elsewhere on the fronds; sporangia not setose; spores tetrahedral, with trilete scars. Ceradenia gameriana is known only from the type. It grows in paramos, at 4100 m, on mossy rocks. It appears to be more closely related to a group of mostly of terrestrial species growing in paramos, e.g., C. herrerae (Copel.) L. E. Bishop, C. terrestris L. E. Bishop, and C. maxoniana L. E. Bishop. The type of C. maxoniana (Lehmann 2400, from Dept. Tolima, Colombia, holotype US!, isotype B!, both online photo) with erect fronds is very similar, and it may prove to be the same species or conspecific (a possible later synonym), judging by the description (Bishop, Amer. Fern J. 79:24, Fig. 2D. 1989a). At first glance C. gameriana differs from all known species of Ceradenia by having a greater stature and longer petioles, characters that can be phenotypically plastic. The rhizome of C. gameriana is unknown, but presumably radially symmetrical (typical of subg. Ceradenia). More complete collections from the type locality are necessary in order to reach a better understanding on the subject of the relationships. I thank A. R. Smith for doing a critical reading of the manuscript and helping me with the English grammar, as well as anon 5 rmous reviewers for their useful comments regarding this work.— JuuAn Mostacero Giannangeu, Universidad Central de Venezuela, Fundacion Institute Botanico de Venezuela, Herbario Nacional de Venezuela (VEN), Apartado 2156, CaracaslOlO-A, Venezuela, e-mail: julian.mostacero@ucv.ve. I Fern Journal 103(l):57-58 (2013) Review Ferns of Southern Africa: A Comprehensive Guide, by Neil R. Crouch, Ronell R. Klopper, John E. Burrows, and Sandra M. Burrows. Struk Nature, Capetown, South Africa. 2011 (actually published on 12 January 2012). Price: $50.00 (available on Amazon.com) ISBN 978-1-77007-910-6 (for the softcover edition). Enthusiasts of South Africa’s diverse fern flora have been blessed with a number of exceptional publications over more than a century. In recent decades, in addition to the Flora of South Africa volume on Pteridophyta (Schelpe and Anthony, 1986), Werner B.G. Jacobsen (1983) set a very high standard with The Ferns and Fern Allies of Southern Africa, a large volume that was monographic in its level of detail and was illustrated with black-and- white photographs. John and Sandra Burrows (1990) extended this tradition with another large-format volume. Southern African Ferns and Fern Allies, equally detailed in its presentation and with the added advantage of color photographs and beautiful line drawings. Both of these have gone out of print in recent years, and there also has been a great deal of floristic and taxonomic research on ferns of the region during the past two decades, creating the need for an updated pteridoflora for South Africa. The new book is smaller in its format than its predecessors, but is even more impressive in its content. Because the publisher is an imprint of the South African arm of publishing giant Random House, the book is both readily available internationally and very reasonably priced. The softcover edition is sturdily bound and has a weatherproof cover. Originally, limited numbers of copies were printed in hardcover and leatherbound editions, but these both were sold out before the book ever appeared. The authors have a wealth of experience with South African ferns and lycophytes, and their experience shows in the precise descriptions and useful discussions, especially those based on their firsthand knowledge of ecological relationships of the species. The number of accepted taxa is 321 for the region, down from 343 in the earlier volume by Burrows because of extensive taxonomic revisions. Despite this, during the research six new taxa were discovered in the genera Cheilanthes, Isoetes, Ophioglossum, Pilularia, and Selaginella. These are formally described in a section on Taxonomic Notes toward the conclusion of the volume, in addition to their treatment in the main body of the book. The familial classification more or less follows that in the recent Synopsis of the Lycopodiophyta and Pteridophyta of Africa, Madagas- car, and Neighbouring Islands (Roux, 2009), as do most of the generic and species circumscriptions. Genus and species determinations are accomplished using dichotomous keys that are nicely set apart from the text with colored backgrounds, as well as tables differentiating closely related taxa. Throughout the book, from tables to maps to keys, there has been excellent use made of color in the layout. The key to families is an unusual construct consisting of a lengthy illustrated table that PTERIDOLOGIA ISSUES IN PRINT The following issues of Pteridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 plus postage and handling. 2 A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choco (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 plus postage and handling. 3. Lellinger, David B. 2002. A Modem Multilingual Glossary for Taxonomic Pteri- dology. 263 pp. $28.(X) plus postage and handling. 14. Hirai, Regina Y., and Jefferson Prado. 2012. Monograph of Moranopteris (Poly- podiaceae). 1 1 3 pp. $28.00 plus postage and handling. For orders and more information, please contact our authorized agent for sales at: Missouri Botanical Garden Press, P.O. Box 299, St. Louis, MO 63166-0299, tel. 314-577- 9534 or 877-271-1930 (toll free). For online orders, visit: http://www.mbgpress.org. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns. SPORE EXCHANGE Mr. Brian S. Aikin, 3523 Federal Ave, Everett, WA 98201-4647 (spores.afs@comcast. net), is Director. Spores exchanged and lists of available spores sent on request, http:// amerfemsoc.org/sporexy.html GIFTS AND BEQUESTS Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Back issues of the Journal and cash or other gifts are always welcomed and are tax-deductible. Inquiries should be addressed to the Membership Secretary. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://amerfemsoc.org/ AMERICAN FERN JOURNAL Q V( ^ Volume 103 Number 2 April-June 2013 The American Fern Society Council for 2013 MISSOURI BOTANICAL „erica„ Fem Journal 103(2).59-m (2013) NO V 1 4 2013 Current Status of the Ferns and Lyco|^y{ey?iHfiil& Hawaiian Islands University of Haw Abstract.— The Hawaiian Islands are well known for ha percentages of endemic plants in the world. Hawaiian femi large percentage of the endemic flora with approximately species considered endemic. In addition, at least 40 taxa are naturalized synopsis of the Hawaiian fem and lycophyte flora that includes new recent taxonomic updates and problems, and a summary of known and h origins of fem and lycophyte lineages. We also provide a checl to the native and naturalized ferns and lycophytes of the Hawai hawaii.edu/lucid/fems/introduction.html). Key Words. — Hawaii, fem, lycophyte, checklist, interactive ke one of the highest documented lycophytes represent a relatively of the native fem and lycophyte access to an interactive key There are 159 native species of ferns and lycophytes in the Hawaiian Islands. Although these represent only about 15% of the native vascular plant flora of the islands (159 fem and lycophyte species: 1032 flowering plant species) (W. L. Wagner, pers. comm, and W. L. Wagner et ah, 2005-), fern species are often common components of the vegetation of many mesic to wet plant communities and are sometimes the dominant species both in numbers of individuals and in total biomass (e.g., see Gagne and Cuddihy, 1999). In addition to the native fern and lycophyte taxa, there are now 40 introduced and naturalized taxa, some of which are becoming invasive. Thus, an accurate knowledge of fern and lycophyte diversity in the Hawaiian Islands is important for both basic ecological and evolutionary studies and for conservation management purposes. During the course of preparing an interactive, online key to the ferns of the Hawaiian Islands (see: http://www. herbarium.hawaii.edu/lucid/fems/introduction.html), we conducted a thor- ough review and analysis of the taxonomic history of Hawaiian ferns and lycophytes. We present here a summary of those studies, including the history of collecting, an overview of fern and lycophyte diversity in the islands, estimates of the geographical origins of native Hawaiian fem and lycoph3de lineages, additions to the known flora since the publication of the last floristic inventory (Palmer, 2003), and recent taxonomic changes. Synopsis of Hawaiian Fern and Lycophyte Floras and Enumerations It was upon Captain James Cook’s voyage of the HMS Resolution in 1778 that the first collections of Hawaiian plants were made by the naturalist William Anderson (W. L. Wagner et oL, 1999). Anderson made a single collection. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Argemone glauca Pope, and noted 22 plant taxa occurring on the island of Kaua’i. The following year, David Nelson, an apprentice botanist aboard Cook’s Resolution sister ship the Discovery, collected approximately 136 vascular plant specimens (St. John, 1978; W. L. Wagner et ah, 1999). Although the ship made excursions to Ni’ihau, Kaua’i, Maui, and Hawai’i, Nelson was able to explore only the island of Hawai’i. The specimens from these voyages are housed at the Natural History Museum in London. Archibald Menzies and Albert Chamisso also were early collectors of Hawaiian plants. Menzies explored five of the Hawaiian Islands from 1788 to 1794 (W. L. Wagner et al, 1999). Chamisso collected in the 1810s on O’ahu and later published species descriptions in the journal Linnaea (Robinson, 1912a). Charles Gaudichaud was the botanist aboard Louis de Freycinet’s voyage of the Uranie that explored O’ahu, Maui, and Hawai’i for twenty days in 1819, although only seven days were spent on land (St. John and Titcomb, 1983). Gaudichaud collected 4,175 specimens throughout the journey (1817-1820), but unfortunately 2,500 of those were lost in a shipwreck. In 1826, he published his account of Hawaiian ferns and lycoph5rtes in Botanique du Voyage autourdu monde (Gaudichaud, 1826). He included descriptions for 52 genera and 268 species of ferns and lycophytes. Of these, 17 genera and 37 species were from the Hawaiian Islands. Between 1825 and 1834, James Macrae, George T. Lay, and David Douglas made collections of Hawaiian ferns and lycophytes (Robinson, 1912a; W. L. Wagner et al, 1999). The botanist William Brackenridge visited the Hawaiian Islands from 1840—1841, as part of the United States Exploring Expedition under the command of Charles Wilkes (W. L. Wagner et al, 1999). Throughout the expedition, he recorded numerous ferns and lycophytes and published brief descriptions for all species, and he also included 46 plates with detailed illustrations (Brackenridge, 1971). Unfortunately the majority of the copies of this publication were destroyed in a fire and only 12 copies were salvaged (Robinson, 1912a). Images of Brackenridge’s plates may be viewed at the Smithsonian Institution Libraries website: http://www.sil.si.edu/imagegalaxy/ imageGalaxy_MoreImages.cfm?book_id=19-24a. Several enumerations and diagnostic keys of Hawaiian ferns were made between 1864 and 1887. William T. Brigham and Horace Mann visited Kaua’i, O’ahu, Maui, Moloka’i, and Hawai’i and made a checklist of 113 fern species (Robinson, 1912a). John M. Lydgate (1873) published a synopsis of Hawaiian ferns that included descriptions of 24 genera and keys to 111 taxa. Derby (1875) published a checklist that included 124 fern and 15 lycophyte taxa. Several years later, Edward Bailey (1883) wrote a summary of the Hawaiian ferns that included 149 species, each with descriptions and several with elevational and habitat data, although no identification keys were included. Lorenzo G. Yates (1887) published an enumeration of Hawaiian ferns (129 species) that lacked species descriptions, but which incorporated some habitat and elevational information for most species. William Hillebrand (1888) spent 20 years in the Hawaiian Islands working as a physician and developing his skills as a botanist. At the time of his death in AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) S. Wagner. In 1981, he presented a checklist of approximately 180 fern and lycophyte taxa (W. H. Wagner, 1981). Seven years later he reported 198 taxa, 172 native (118 endemic) and 26 naturalized, although he acknowledged uncertainties about the exact total number of taxa (W. H. Wagner, 1988). The two most recent Hawaiian fern and lycophyte floras are from Kathy Valier (1995) and Daniel D. Palmer (2003). Valier did not treat all the Hawaiian ferns and lycophytes, but rather provided descriptions for 61 of the most commonly encountered taxa. She reported, in accordance with W. H. Wagner (1988), about 200 fern and lycophyte taxa, 172 being native. The most up-to- date and comprehensive Hawaiian fern and lycophyte flora is that of Palmer (2003), which included descriptions of fern and lycophyte families, genera, and species along with keys to genera and species and numerous illustrations. Palmer recognized 73 genera and 221 taxa (including hybrids) of Hawaiian ferns and lycophytes. The species descriptions highlight characters that separate each species from similar species and include information on synonyms, current distribution, habitat, elevation, ethnobotanical knowledge, and vernacular names. Palmer was able to record the variability found throughout Hawaiian ferns and lycophytes by observing numerous specimens in the field and in herbaria and including first-hand observations of many type specimens. Current Treatment The Hawaiian Islands have representatives from all major lineages of monilophytes except the Equisetopsida and from all three lineages of lycophytes. The endemic fern genera of the Hawaiian Islands have long been thought to consist of Adenophorus (Polypodiaceae), Diellia (Aspleniaceae), and Sadleria (Blechnaceae). [N.B. Authorities for all Hawaiian taxa are given in Appendix 1.] However, molecular work by Schneider et al. (2004a, 2005) supported Diellia as a monophyletic clade nested within Asplenium (Aspleniaceae), therefore decreasing the number of endemic fern genera to two. This classification is recognized in the majority of recent treatments (Christenhusz et al, 2011; Smith et al, 2006, 2008; Snow et al, 2011; Viane and Reichstein, 1991). There are presently 144 native fern species (167 taxa) and 15 native lycophyte species (16 taxa), not including hybrids in the Hawaiian Islands. The genera with the most Hawaiian fern taxa are Asplenium (29 taxa), Dryopteris (20 taxa), and Adenophorus (12 taxa). A synopsis of the statistics of the Hawaiian fern and lycophyte flora is given in Table 1. New Fern Records in the Hawaiian Islands Several new state and island records of native and naturalized ferns have been published since Palmer’s flora (2003), including seven new state records of naturalized ferns (Table 2). In addition, 20 new island records have been reported, most of which are also naturalized ferns. One new island record of a native species is of the endemic Asplenium haleakalense, which was VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS federally listed or proposed for listing. As discussed later, the presumed extinct fern, Asplenium dielmannii, was also rediscovered on Kaua’i (Aguraiuja and Wood, 2003}. Another rediscovery on Kaua’i is that of Ctenitis squamigera, a species previously known from all main Hawaiian Islands with the exception of Kaho’olawe and Ni’ihau. It was last collected on Kaua’i in the late 1800’s and thus presumed extinct on the island until its rediscovery in two separate locations in 2011 (K. Wood, 2012). Pteris lidgatei was rediscovered on the island of Moloka’i (Oppenheimer, 2007). Its last known collection on Moloka’i was in 1912, and since then had been considered extinct on the island (Oppenheimer, 2007; Palmer, 2003; W. H. Wagner et ah, 1999). On Moloka’i, eight individuals are known from a single population (USFWS, 2009b). An interesting island record is that of Psilotum nudum (Psilotaceae). Although present on all main Hawaiian Islands including Ni’ihau and Kaho’olawe, P. nudum was never reported from any Northwestern Hawaiian Island (Palmer, 2003). A collection of P. nudum made in 1923 from Midway Atoll was discovered in 2008 in the collections of the Herbarium Pacificum (Kennedy et al., 2010). Two other possible island records, based on herbarium specimens, are of Doryopteris angelica (Pteridaceae) and D. takeuchii. Yesilyurt (2005) proposed that the endangered Kaua’i endemic, Doryopteris angelica, has a more extensive distribution than had been previously reported (Palmer, 2003) occurring on O’ahu, Moloka’i, and Lana’i. She also determined a range extension for the endangered (USFWS, 2012a) Doryopteris takeuchii, to the island of Lana’i, previously thought to be endemic to Diamond Head Crater on O’ahu. It is imperative to note that Yesilyurt’s (2005) conclusions were based on relatively few older herbarium specimens (5 specimens of D. takeuchii and 11 specimens of D. angelica), and she did not observe the type specimens for either species. For these reasons and because both of these ferns are currently federally listed endangered species it is important that new field observations be made to assess the current distribution of these two species before we accept the range extensions proposed by Yesil)mrt (2005). Newly Described Hawaiian Fern Taxa Only one native species, Cyclosorus pendens (syn. Pneumatopteris pendens-, Thelypteridaceae), has been described since Palmer’s (2003) manual was published (Palmer, 2005). This species is found on Kaua’i, O’ahu, Moloka’i, Maui, and Hawai’i at elevations of 368-1220 m and is most similar morphologically to C. sandwicensis (Palmer, 2005). Characters that define C. pendens are a habitat of damp, mossy rocks often near streams and pendent, lanceolate leaves with obtuse pinnae and rachises and costae that are densely covered with hairs (Palmer, 2005). Another species not included in Palmer (2003) is Deparia cataracticola (Athyriaceae), of which the author was unaware when his book was in press (D. Palmer, pers. comm.). This species is endemic to Kaua’i and is found on wet, mossy cliffs in waterfalls, a habitat similar to where the extinct D. VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Table 3. Current distribution of naturalized fern and lycophyte taxa in the Hawaiian Islands. O Mo L Ka Adiantum hispidulum Marsilea minuta * Nephrolepis brownii Nephrolepis falcata ‘Furcans’ Nephrolepis hirsutula ‘Super Pblebodium aureum Phymatosorus grossus Phymatosorus scolopendiia * Pityrogramma austroamericai Pityrogramma calomelanos Platycerium bifurcatum Platycerium superbum Pteris tremula * Pyrrosia piloselloides * Salvinia molesta Selaginella kraussiana Selaginella stellata Selaginella umbrosa Sphaeropteris cooperi Tectaria incisa * indicates new state record subsequ Hawaiian Island abbreviations: N Ma=Maui; Ka=Kaho’olawe; H=Haw Blechnum orientale (Blechnaceae) was discovered naturalized in two locations on the island of O’ahu in 2010 (Lau and Frohlich, 2012). It is native to tropical Asia, Australia, and several Pacific Islands (Chambers and Farrant, 2001; Chiou et al, 1994). Diagnostic characters include rhizomes that are short. VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS preserved in the “Godet herbarium,” although Robinson (1912a) reported that no such herbarium was known. Further investigation into the location of this specimen has been unsuccessful, although other collections from the Godet herbarium have been found at the Botanic Garden and Botanical Museum Berlin-Dahlem herbarium (B). The most recent record of Marsilea minuta was made in 1994 when it was found growing in taro patches at the Hawai’i Nature Center in Makiki Valley on O’ahu (Wilson, 2002, 2003). Marsilea minuta is native to Africa, Australia, India, and Southeast Asia (Jacona and Johnson, 2006) . It is morphologically similar to the endemic M. villosa, which also has an aquatic habit, pinnae that are clover-like (divided into four equal pinnae), and sori enclosed in hardened sporocarps located at the bases of long stipes. This naturalized species is readily distinguished from the endemic species by its usually glabrous blades (sometimes with a few scattered hairs) with crenate pinnae margins (entire margins in aquatic forms) (Jacona and Johnson, 2006), in contrast with the glabrous blades and slightly denticulate pinnae margins of M. villosa (Johnson, 1986; Palmer, 2003). Furthermore, the naturalized species has roots located at nodes and internodes (Jacona and Johnson, 2006), whereas M. villosa has roots restricted to the nodes (Johnson, 1986). Two populations of Phymatosorus scolopendria (Polypodiaceae) were reported on Maui (Oppenheimer, 2006). A collection of this species was made [Oppenheimer Gr Hansen H80308, BISH), but cannot be located for verification. Phymatosorus scolopendria is native to tropical Africa and Asia, Australia, and the Pacific (Brownlie, 1977). This species closely resembles and can easily be mistaken for a common naturalized fern, P. grossus, but is distinguished by its epiph3^ic habit (vs. terrestrial), smaller size, and fewer laminar lobes (Brownlie, 1977). Pyrrosia piloselloides was found naturalized in a 2-3 acre area in the Woodlawn section of Manoa Valley on the island of O’ahu in 2010 (Lau and Frohlich, 2012). This species is native to China, India, and Malesia (Ravensberg and Hennipman, 1986). Pyrrosia piloselloides is a small, epiphytic, colony-forming fern diagnosable by its sterile succulent leaves reaching 1-7 cm long, fertile linear leaves from 4-16 cm long, and linear, submarginal sori (Hovenkamp, 1986). Pteris tremula was first reported as naturalized on Maui (Oppenheimer, 2007) , when it was noted as becoming widespread throughout the central Kula areas at elevations from 914 to 1220 m. No further report has been found concerning its distributional status in the Hawaiian Islands. This species is native to Australia, New Zealand, and the Fiji Islands; it is identified by its erect leaves that attain one meter in length and blades that are 2-4 times pinnately divided with linear, toothed pinnules (Kramer and McCarthy, 1998). Taxonomic Updates Smith et al. (2006) revised the familial classification for all extant ferns based on the results of numerous molecular and morphological studies. They recognized 37 families with approximately 265 to 312 genera. A more recent AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) classification for extant ferns and lycophytes by Christenhusz et al. (2011) built upon the classification of Smith et al. (2006) with recent phylogenetic studies. Christenhusz et al. (2011) recognized 3 families with 5 lycophyte genera and 45 families with approximately 280 fern genera. Snow et al. (2011) addressed the need to update the taxonomy and nomenclature of the Hawaiian ferns and lycophytes. Their treatment primarily followed the classification by Smith et al. (2006, 2008). Rothfels et al. (2012b) reexamined the classification of the Eupolypods II and recognized 10 families with 36 genera. A recent classification of the Lycopodiaceae by 011gaard (2012) accepts 9 Neotropical genera. Although these publications clarify the classification of extant ferns and lycophytes, it is important to note that many fern and lycophyte genera lack molecular studies and are in need of phylogenetic analyses to assess proposed generic delimitations (Christenhusz et al., 2011; Smith et al., 2006). Below, we summarize recent taxonomic and nomenclatural changes and problems that still require further study for Hawaiian fern taxa. Ophioglossaceae. — One example of fern genera in need of further phylogenetic analysis is within the family Ophioglossaceae. The majority of treatments include Sceptridium within the genus Botrychium, and Ophioderma within Ophioglossum (Christenhusz et al, 2011; Smith et al., 2006, 2008). However, a molecular study by Hauk et al. (2003), using plastid DNA sequences from rbcL and trnL-F, showed strong support for the segregation of Sceptridium and Ophioderma as distinct genera. Snow et al. (2011) included Sceptridium within Botrychium and did not address the segregation of Ophioderma. Further phylogenetic analyses are needed to resolve these issues. Blechnaceae.— Another example of uncertain generic boundaries is evident with the circumscription of Blechnum, in which Shepherd et al. (2007) determined Blechnum to be paraphyletic and including Doodia and Sadleria. In accordance with earlier work by Cranfill (2001), Sadleria, an endemic Hawaiian genus, was found nested within Blechnum. However, S. cyatheoides was the only species of the genus included in the study and its placement had only weak support. Christenhusz et al. (2011) accept the genus Sadleria, but state that the status is not clear due to Blechnum being likely paraphyletic. Rothfels et al. (2012b) accept the genus Sadleria based on the inferred phylogeny of Rothfels ef al. (2012a). Strong support was found for Doodia as a monophyletic group within Blechnum with B. brasiliense Desv. as the sister species (Shepherd et al., 2007). Further analyses of Doodia are needed before making taxonomic changes because only six of the 15-18 species of Doodia were sampled, including only one of the two Hawaiian endemic species [D. kunthiana). In addition, only one small region of plastid DNA was sequenced [tmL-F) and results differed depending on how the data were analyzed. Furthermore, according to the maximum likelihood tree of Shepherd et al. (2007), Blechnum brasiliense could be placed within Doodia to produce a monophyletic genus. Rothfels et al. (2012b) conserve the genus Doodia, whereas Christenhusz et al. (2011) recognize Doodia within the genus Blechnum. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) the monophyly of Oreogrammitis. Work is in progress to further resolve the generic circumscriptions within Grammitis s. 1. (Sundue et al., unpublished). Thus, the placement of the Hawaiian members of Grammitis s. 1. is currently unclear, although the Hawaiian species clearly do not belong to Grammitis s. s. due to the lack of black, sclerified leaf margins. We treat the Hawaiian species here as within Oreogrammitis pending further study. Ranker et al. (2004) found Lellingeria to be divided into two non-sister clades; one including the L. apiculata (Kunze ex Klotzsch) A. R. Sm. & R. C. Moran and L. myosuroides (Sw.) A. R. Sm. & R. C. Moran groups (sister to Melpomene A. R. Sm. & R. C. Moran) and the other including the L. mitchelliae (Baker) A. R. Sm. & R. C. Moran group, now called Leucotrichum Labiak, (sister to Alansmia M. Kessler, Moguel, Sundue and Labiak) (Kessler et al, 2011; Rouhan et al, 2012). Lellingeria saffordii, the sole Hawaiian representative, is an endemic species that occurs on all major islands except Ni’ihau and Kaho’olawe. It is nested phylogenetically within the L. myosuroides group, which has recently been described as the new genus Stenogrammitis (Labiak, 2011), and thus the Hawaiian species is treated here as S. saffordii. Pteridaceae. — Ceratopteris is represented in the Hawaiian Islands only by the naturalized species, C. thalictroides, which is found on Kaua’i, O’ahu, and Hawai’i (Imada, 2007; Palmer, 2003). Species in the genus are generally polymorphic and this is especially so with C. thalictroides (Masuyama and Watano, 2010). The genus has been variously recognized as including 11 species (e.g,, Christensen, 1906), four species (Lloyd, 1974), with Lloyd lumping four previously described species into C. thalictroides, and 3 species (Tryon and Tryon, 1982). Masuyama and Watano (2010) proposed that five taxa be segregated from C. thalictroides based on differences in the nucleotide sequences of plastid DNA, cytological evidence, and morphological differenc- es (Masuyama et al., 2002; Masuyama and Watano, 2005; Masuyama, 2008). The Ceratopteris found in the Hawaiian Islands included a newly described taxon, C. gaudichaudii var. vulgaris, also known from China, Guam, Japan, Korea, Nepal, and Taiwan (Masuyama and Watano, 2010). Masuyama and Watano (2010) examined two specimens from O’ahu and determined Hawaiian members to be this new variety. Although C. gaudichaudii var, vulgaris occurs on O’ahu, it is important to note that additional Ceratopteris taxa may be present in the Hawaiian Islands; further study is needed. Distribution The majority of fern and lycophyte taxa are found on six of the eight main Hawaiian Islands (i.e., Kaua’i, O’ahu, Moloka’i, Lana’i, Maui, and Hawai’i), with few found on Ni’ihau and Kaho’olawe (Figure 1). Two species, the endemic Doryopteris decipiens and the indigenous Psilotum nudum are distributed on all eight main Hawaiian Islands. [NB: Indigenous is defined here as taxa that are native to a particular area, but are also native elsewhere.] In total there are 140 endemic fern and lycophyte taxa, representing VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Fig. 2. Endemic fem taxa on each main Haw approximately 77% of the 183 native taxa. This level of endemism is high compared to other oceanic islands, which usually have endemic rates of less than 50% (Moran, 2008). The greatest numbers of endemic fem and lycophyte taxa are found on Kaua’i and Maui (Figure 2). Despite the high level of island- wide endemism, Hawaiian ferns and lycophytes have relatively low numbers of single-island (28 taxa) and two-island (19 taxa) endemics (Figure 3). Only Kaua’i, O’ahu, and Maui have single-island endemics, with Kaua’i harboring the greatest number (17 taxa). The majority of taxa that are restricted to two islands are found on Maui (15 taxa) and Hawai’i (12 taxa). in the eight main Hawaiian Islands. The AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) 1 ferns and lycophytes of dry habitats. Adiantum hispidulum Asplenium adiantum-nigrum Asplenium aethiopicum * Asplenium dielerectum Asplenium caudatum Asplenium nidus Asplenium normale Asplenium trichomanes subsp. dt Blechnum appendiculatum Cheilanthes viridis Cyrtomium caryotideum Cyrtomium falcatum Doodia kunthiana Doryopteris decipiens Doryopteris decora Doryopteris takeuchii Dryopteris fuscoatra var. fuscoatr Lepisorus thunbergianus Lycopodium venustulum var. vert Microlepia speluncae Microlepia strigosa var. strigosa Nephrolepis brownii Pellaea temifolia 500-1700 200-1000 40-610 375-1680 1200-2700 30-560 10-700 300-2100 0-1525 90-1220 30-915 120-700 100-120 500-2100 10-2100 1220-2440 480-1280 0-1770 425-1830 600-3500 0-700 45-1525 0-1000 340-2450 450-1920 Elevational Ranges and Habitat The majority of Hawaiian ferns and lycophytes occur in high elevation mesic-wet to wet forests (MacCaughey, 1918; Palmer 2003; W. H. Wagner, 1995). Approximately 31 fern and lycophyte taxa occur in dry habitats, defined as areas receiving 1200 mm or less of annual rainfall (Gagne and Cuddihy, 1999) (Table 4). Coastal elevations, defined by Gagne and Cuddihy (1999) as 0-^ 300 m, harbor approximately 48 fern and lycophyte taxa, although many of these taxa are also distributed in higher elevations (Palmer, 2003). The native species with the lowest elevational range is Ophioglossum polyphyllum (Ophioglossaceae), occurring at 2-160 m. This species is seasonally common, appearing only after heavy rains in sand dunes, grasslands, and lava cobble (Palmer, 2003). Native species that occur at some of the highest elevations (above 3000 m) include: Asplenium adiantum-nigrum (Aspleniaceae), Cystopteris VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Summary of fern and lycophyte ti Terrestrial Lithophytic Epiphytic* Epigeous 39 ( 20 . 75 ) 9 ( 25 . 00 ) 15 ( 22 . 73 ) douglasii (Cystopteridaceae), Pellaea ternifolia [Pteridaceae), and Polystichum haleakalense (Dryopteridaceae). The majority of Hawaiian ferns are terrestrial, although many are epiphytic and occur mostly in high elevation, wet forests (Table 5). Several species are restricted to aquatic habitats, including the endangered endemic water clover fern, Marsilea villosa and the rare lycophyte, Isoetes hawaiiensis (Isoetaceae). Other interesting habitats include those that are epigeous (growing above, but near the surface of the ground in moss mats at the base of trees) and lithoph 5 ^ic/epipetric (growing on rocks). Only three taxa have an epigeous habit, and all are within the endemic genus Adenophorus: A. epigaeus, A. tamariscinus var. tamariscinus, and A. tripinnatifidus. Several examples of lithophytic species include Asplenium dielerectum, Cyclosorus boydiae, C. pendens, and Doryopteris takeuchii. Origins of the Hawaiian fern and lycoph 3 ^e flora The original number of colonizing species of angiosperms to the Hawaiian Islands is estimated to have been a minimum of 291 species (Sakai et ah, 1995), which gave rise to approximately 1032 native species (W. L. Wagner, pers. comm, and see updates at W. L. Wagner et ah, 2005-) in the flowering plant flora (i.e., an average of —3.6 derived species per colonizing ancestor). We used data from molecular phylogenetic studies and/or information on the geographical distribution of species and genera to estimate the origins of 78 taxa or lineages (33 endemic and 45 indigenous) of Hawaiian ferns and lycoph 3 des (Appendix 2). Molecular phylogenetic data were available for 22 taxa and/or lineages (see references cited in Appendix 2). We inferred the origin of the remaining 56 taxa and/or lineages based on the current distribution of the species (i.e., for indigenous species of limited distribution outside of the Hawaiian Islands) or of the genus (i.e., for Hawaiian endemics with congeneric species of limited distribution outside of the Hawaiian Islands). The 78 taxa or lineages comprise 103 taxa in total, thus there are an average of —1.3 extant taxa per inferred colonizing species (Table 6 and Appendix 2). We estimate that the majority of Hawaiian fern and lycophyte VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Table 8. wild pop Reported individuals of federally ] 1 endangered Hawaiian ferns and lycophytes in Adenophorus penen. Asplenium dielmannii Asplenium dielpallidum Asplenium peruvianum var. insulare : squamig Ctenit Diplai Doryopteris angelica 15-123 12 278 1148 234-242 ~65 29-54 50-100 32-17 thousands 51 USFWS, 2010a USFWS, 2009c USFWS, 2011a K. Wood, pers. < USFWS, 2008 USFWS, 2009d USFWS, 2010b USFWS, 201 Id USFWS, 201 le USFWS, 2009e USFWS, 201 If USFWS, 2009b In 2005, DOF AW and the Division of Aquatic Resources (DAR) within the Department of Land and Natural Resources (DLNR) developed a Comprehen- sive Wildlife Conservation Strategy (CWCS) to review the status of the native terrestrial and aquatic species in the Hawaiian Islands (Mitchell et al, 2005). The CWCS identifies the Hawaiian species of greatest conservation need (SGCN) and provides strategic plans for their conservation. The criteria CWCS used to determine a SGCN need may he found at: http://www.state.hi.us/ dlnr/dofaw/cwcs/process_strategy.htm. The CWCS includes 22 ferns and 3 lycoph3^es as species of greatest conservation need (Table 7). Eight of these species are listed as SGCN because they play an important role in their native habitats as either a dominant member of the native plant community or by hosting or providing habitat for other native species (Mitchell et al., 2005). Along with these species, three genera {Cibotium, Dryopteris, and Elaphoglos- sum] also have this designation (Mitchell et al., 2005). The Plant Extinction Prevention (PEP) program of Hawai’i reported nine fern taxa and one lycophyte with less than 50 individuals remaining in wild populations: Adenophorus periens, Asplenium dielerectum, A. dielmannii, A. dielpallidum, Asplenium peruvianum var. insulare, Diplazium molokaiense, Doryopteris angelica, Dryopteris crinalis var. podosora, Huperzia stemmer- manniae, and Pteris lidgatei (PEP, 2011). Approximate numbers of currently known individuals of endangered Hawaiian ferns have been compiled from reports of the USFWS and additional resources (Table 8). Extinct Hawaiian Ferns W. L. Wagner et al. (2005-) reported three ferns and one lycophyte from the Hawaiian Islands as extinct (Table 7). A single plant of one previously VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS SUNDUE, M. A., M. B. Islam and T. A. Ranker. 2010. Systematics of grammitid ferns (Polypodiaceae): using morphology and plastid sequence data to resolve the circumscription of Melpomene and the polyphyletic genera Lellingeria and Terpsichore. Syst. Bot. 35:701-715. Tryon, R. M. and A. F. Tryon. 1982. Ferns and Allied Plants: With Special Reference to Tropical America. Springer, New York. [USFWS] U. S. Fish and Wildufe Service. 1998. Recovery Plan for Four Species of Hawaiian Ferns. U.S. Fish and Wildlife Service, Portland, OR. [USFWS] U. S. Fish and Wildufe Service. 2004. Endangered and Threatened Wildlife and Plants; Review of Species That Are Candidates or Proposed for Listing as Endangered or Threatened; Annual Notice of Findings on Resubmitted Petitions: Annual Description of Progress on Listing Actions. Federal Register 69. [USFWS] U. S. Fish and Wildufe Service. 2008. 5-Year Review Summary and Evaluation: Diellia pallida, http:/ / ecos.fws.gov/ speciesProfile/ profile/speciesProfile.action?spcode = S022. [ac- cessed April, 2011]. [USFWS] U. S. Fish and Wildlife Service. 2009a. Endangered and Threatened Wildlife and Plants; Initiation of 5-Year Reviews of 103 Species in Hawaii. Federal Register 74. [USFWS] U. S. Fish and Wildufe Service. 2009b. 5-Year Review, short form summary: Pteris lidgatei. http ://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode= SOOW. [accessed April, 2011]. [USFWS] U. S. Fish and Wildlife Service. 2009c. 5-Year Review, short form summary: Diellia erecta. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOOP. [accessed April, 2011]. [USFWS] U. S. Fish and Wildlife Service. 2009d. 5-Year Review, short form summary: Ctenitis squamigera http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=S01I. [accessed April, 2011]. [USFWS] U. S. Fish AND Wildufe Service. 2009e. 5-Year Review, short form summary: Lycopodium nutans. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOlL [ac- cessed June, 2012]. [USFWS] U. S. Fish and Wildufe Service. 2010a. 5-Year Review, short form summary: Adenophorus periens. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOOM. [ac- cessed April, 2011]. [USFWS] U. S. Fish and Wildlife Service. 2010b. 5-Year Review, short form summary: Diplazium molokaiense. http://ecos.fws.gov/speciesProfile/profile/speciesProflle.action?spcode=SOOR. [accessed June, 2012]. [USFWS] U. S. Fish and Wildlife Service. 2011a. 5-Year Review, short form summary: Diellia falcata. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOOQ [ac- cessed June, 2012]. [USFWS] U. S. Fish and Wildlife Service. 2011b. 5-Year Review, short form summary: Diellia unisora http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=S01T [ac- cessed June, 2012]. [USFWS] U. S. Fish and Wildlife Service. 2011c. Species profile: Doryopteris angelica. http://ecos. fws.gov/speciesProfile/profile/speciesProfile.action?spcode=S02H. [accessed April, 2011]. [USFWS] U. S. Fish AND Wildufe Service. 201 id. Species profile: Dryopteris crinalis var. podosorus. http://ecos. fws.gov/speciesProfile/profile/speciesProfile. action?spcode=S02l. [accessed April, 2011]. [USFWS] U. S. Fish and Wildufe Service. 2011e. 5-Year Review, short form summary: Huperzia mannii. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOlK [ac- cessed June, 2012]. [USFWS] U. S. Fish and Wildufe Service. 201 If. 5-Year Review, short form summary: Marsilea villosa. http://ecos.fws.gov/speciesProfile/profile/speciesProfile.action?spcode=SOOU [ac- cessed June, 2012]. [USFWS] U. S. Fish AND Wildlife Service. 2012a. Endangered and Threatened Wildlife and Plants; Endangered Status for 23 Species on Oahu and Designation of Critical Habitat for 124 Species. Federal Register 77(181):57647-57862. VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Appendix 1. Checklist for the native and naturalized ferns and lycophytes of the Hawaiian Islands. The checklist follows the linear sequence for extant ferns and lycophytes by Christenhusz et al (2011) with several modifications (Rothfels et al, 2012b). ® endemic; ^ indigenous: naturalized Hawaiian Island abbreviations: N=Ni’ihau; K=Kaua’i; 0=0’ahu; Mo= Moloka’i; L=Lana’i; Ma=Maui; Ka=Kaho’olawe; H=Hawai’i (ex) = extinct synonyms listed are those that represent updates subsequent to Palmer (2003) Lycophytes 1. Lycopodiaceae P. Beauv. ex Mirb. Huperzia Bemh. ® Huperzia erosa Beitel & W. H. Wagner Distribution: K/O/Mo/L/Ma/H ^Huperzia erubescens (Brack.) Holub Distribution: K/O/Mo/Ma/H ^Huperzia filiformis (Sw.) Holub Distribution: K/O/Mo/L/Ma/H ^Huperzia haleakalae (Brack.) Holub Huperzia mannii (Hillebr.) Kartesz & Gandhi Distribution: K (ex)/Ma/H ^ Huperzia nutans (Brack.) Rothm. Distribution: K (ex)/0 Distribution: K/O/Mo/L/Ma/H zia serrata (Thunb. ex Murray) Trevis. Distribution: K/O/Mo/L/H H. Wagner & Hobdy) Kartesz Distribution: Ma/H ^ Huperzia subintegra (Hillebr.) Beitel & W. H. Wagner Distribution: K/O/Mo/Ma ^Lycopodiella cemua (L.) Pic. Serm. Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 I Lycopodium L. ^Lycopodium venustulum Gaudich. var. venustulum Distribution: K/O/Mo/L/Ma/H ^ Lycopodium venustulum Gaudich. var. verticole W. H. Wagnei Distribution: H 2. Isoetaceae Rei ® Isoetes hawaiiensis W. C. Taylor & W. H. Wagner Distribution: Ma/H 3. Selaginellaceae Willk. ^Selaginella arbuscula (Kaulf.J Spring Distribution: K/O/Mo/L/Ma/H note; Selaginella arbuscula was previously considered an occurring in the Society Islands, Ualan, Santa Cruz Island (Vanikoro), and the Marquesas Islands (Hassler and Swale, 2003). ® Selaginella deflexa Brack. Distrihution; K/O/Mo/Ma/H Selaginella kraussiana (Kunze) A. Braun Distribution: O/Ma/H Selaginella stellata Spring Distribution: H Distribution: H Sceptridium Lyon ^ Sceptridium subbifoliatum (Brack.) Lyon Distribution: K (ex)/0 (ex)/Mo (ex)/L (ex)/Ma Ophioderma (Blume) Endl. Ophioderma pendulum (L.) C. Presl ' subsp. falcatum (C. Presl) R. T. Clausen (ex)/H (ex) Distribution; K/O/Mo/L/Ma/H VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Ophioglossum L. tion: 0/Mo/Ma/H : K/O/H ilatum Hook. Distribution: K/O/Mo/L/Ma/H ^Ophioglossum polyphyllum A. Braun Distribution: K/O/Mo/L (ex)/Ka/Ma/H Psilotaceae J.W.Griff. & Henfr. ^Psilotum complanatum Sw. Distribution: K/O/Mo/L/Ma/H ^Psilotum nudum (L.) P. Beauv. Distribution: N/K/O/Mo/L/Ka/Ma/H 3. Marattiaceae Kaulf. Angiopteris Hoffin. Angiopteris evecta (G. Forst.) Hoffin. Distribution: K/O/Mo/L/Ma/H 1 douglasii (C. Presl) Baker istribution: K/O/Mo/L/Ma/H 4. Hymenophyllaceae Mart. Callistopteris Copel. ^ Callistopteris baldwinii (D. C. Eaton) Copel. Distribution: K/O/Mo/L/Ma/H S3m. Vandenboschia draytoniana (Brack.) Copel. Distribution: K/O/Mo/L/Ma/H ^Crepidomanes minutum (Blume) K. Iwats. Syn. Gonocormus Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) ^Crepidomanes proliferum (Blume) Bostock Syn. Gonocormus prolifer (Blume) Prantl Distribution: Ma/H Hymenophyllum Sm. ^ Hymenophyllum lanceolatum Hook. & . Syn. Sphaerocionium lanceolatui Distribution: K/O/Mo/L/Ma/H . & Am.) Copel. Syn. Sphaerocionium obtusum (Hook. & Am.) Copel. Distribution: 0/Mo/L/Ma/H ^ Hymenophyllum recurvum Gaudich. Syn. Mecodium recurvum (Gaudich.) Copel. Distribution: K/O/Mo/L/Ma/H Vandenhoschia Copel. ^ Vandenhoschia cyrtotheca (Hillebr.) Copel. Distribution: K/O/Mo/L/Ma/H ^ Vandenhoschia davallioides (Gaudich.) Copel. Distribution: K/O/Mo/L/Ma/H Distribution: K icranopteris Bemh. Distribution: K/O/Mo/L/Ma/H Diplopterygium (Diels) Nakai ® Diplopterygium pinnatum (Kunze) Nakai Distribution: K/O/Mo/L/Ma/H ® Sticherus owhyhensis (Hook.) Chi ng Distribution: K/O/Mo/L/Ma/H Lygodiaceae M. Roem. Lygodium Sw. VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAHAN ISLANDS Schizaea Sm. ^ Schizaea robusta Baker Distribution: K/O/Mo/L/Ma/H tiarsileaceae Mirb. Marsilea L. Marsilea minuta L. Syn. Marsilea crenata C. Presl Distribution: O ^ Marsilea villosa Kaulf. Distribution: N/O/Mo lalviniaceae Martinov Salvinia molesta D. S. Mitch. Distribution: 0/H Azalia Lam. Azalia filiculaides Lam. Distribution: K/O/Mo/L/Ma/H ibotiaceae Korall Cibatium chamissai Kaulf. Distribution: 0/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H ^ Cibatium menziesii Hook. Distribution: K/O/Mo/L/Ma/H ® Cibatium nealiae O. Deg. Distribution: K . Cyatheaceae Kaulf. ^ Sphaerapteris caaperi (Hook, ex F. Muell.) R. M. Tryon Distribution: K/O/L/Ma AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 12. Dicksoniaceae M.R. Schomb. Dicksonia L’Her. Dicksonia fibrosa Colen Distribution: H 13. Lindsaeaceae C. Presl ex M.R. Scbomb. Lindsaea ensifolia Sw. Distribution: K/O/Mo/Ma/H Lindsaea repens (Bory) Tbwaites ® var. macraeana (Hook. & Am.) C. Cbr Distribution: K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H Dennstaedtiaceae Lotsy Hypolepis Bemb. Hypolepis hawaiiensis Brownsey ^ var. hawaiiensis Distribution: K/O/Mo/L/Ma/H Distribution: Ma Microlepia C. Presl ^ Microlepia speluncae (L.) T. Moore Distribution: K/O/Mo/Ma/H Microlepia strigosa (Tbunb.) C. Presl ® var. mauiensis (W. H. Wagner) D. D. Palmer Distribution: Ma/H Distribution: K/O/Mo/L/Ma/H Pteridium Gled. ex Scop. Pteridium aquilinum (L.) Kubn Distribution: K/O/Mo/L/Ma/H VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS 15. Pteridaceae E. D. M. Kirchn. Adiantum L. Adiantum ‘Edwinii’ Distribution: O/L/Ma * Adiantum capillus-veneris L. Distribution: N/K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ka/Ma/H antum raddianum C. Presl Distribution: K/O/Mo/L/Ma/H Distribution: O/Mo/Ma Distribution: K/O/H Distribution: K/O/L/Ma/H ^ Coniogramme pilosa (Brack.) Hieron. Distribution: K/O/Mo/L/Ma/H Doryopteris J. Sm. ® Doryopteris angelica K. R. Wood & W. H. Wagner ® Doryopteris decipiens (Hook.) J. Sm. Distribution: N/K/O/Mo/L/Ka/Ma/H ^ Doryopteris decora Brack. Distribution: K/O/Mo/L/Ka/Ma/H ^ Doryopteris takeuchii (W. H. Wagner) W. H. Wagner Distribution: O Haplopteris C. Presl 'Haplopteris elongate (Sw.) E. H. Crane Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) ^Pellaea temifolia (Cav.) Link Distribution: K/O/Mo/L/Ma/H gramma Link Pityrogramma austroamericana Domin Distribution: N/K/O/Mo/L/Ka/Ma/H Pityrogramma calomelanos (L.) Link Distribution: N/K/O/Mo/L/Ka/Ma/H Distribution: K/O/Mo/L/Ma/H ‘Pferis excelsa Gaudich. Distribution: K/O/Mo/L/Ma/H Pteris hillebrandii Copel. Distribution: K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H ® Pteris lidgatei (Hillebr.) Christ Distribution: O/Mo/Ma Pteris tremula R. Br. : K/O/Mo/L/Ma/H Cystopteridaceae (Payer) Shmakov ^ Cystopteris douglasii Hook. Distribution: Ma/H ^ Cystopteris sandwicensis Brae Distribution: K/O/L/Ma Asplenium L. ® Asplenium acuminatum Hook. & Am. Distribution: K/O/Mo/L/Ma/H VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Distribution: K/O/Mo/L/Ma/H ^Asplenium aethiopicum (Bunn, f.) Bech. Distribution: K/O/Mo/L/Ma/H Syn. Asplemum horndum KauU. v^. glabratum (W. J. Rob.) D. D. Distribution: K/O/Mo/L/Ma/H Asplenium contiguum Kaulf. Distribution: K/O/Mo/L/Ma/H ® var. hirtulum C. Chr. Distribution: K/Ma ^ Asplenium dielerectum Viane Syn. Diellia erecta Brack. Distribution: O/Mo/L (ex)/Ma/H ^ Asplenium dielfalcatum Viane S3m. Diellia falcata Brack. Distribution: O ® Asplenium diellaciniatum Viane Distribution: K Syn. Diellia mannii (D. C. Eaton) W. J. Rob. Distribution: K Syn. Diellia pallida W. H. Wagner ^Asplenium excisum C. Presl Distribution: K/O/Mo/L/Ma/H ® Asplenium haleakalense W. H. Wagner ® Asplenium hobdyi W. H. Wagner Distribution: K/Mo/Ma/H 'Asplenium insiticium Brack. Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME ; NUMBER : Asplenium kaulfussii Schltdl. Distribution: K/O/Mo/L/Ma/H ® Asplenium leucostegioides Baker Syn. Diellia leucostegioides (Baker) W. H. Wagner Distribution: K/O/Mo/L/Ma/H ^ Asplenium macraei Hook. & Grev. Distribution: K/O/Mo/L/Ma/H ^Asplenium monanthes L. Distribution: Ma/H Asplenium nidus L. Distribution: K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H ium peruvianum Desv. ^ var. insulare (C. V. Morton) E Distribution: K/O/Mo/L/Ma/H iC. Chr. Asplenium sphenotomum Hillebr. Distribution: K/O/Mo/L/Ma/H Asplenium trichomimes L. ® subsp. densum (Brack.) W. H. Wagner Distribution: Ma/H Distribution: K/O/Mo/L/Ma/H ® Asplenium unisorum (W. H. Wagner) Viane Syn. Diellia unisora W. H Distribution: O I. Wagner VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS 18. Thelypteridaceae Ching ex Pic. Serm. ^ Cyclosorus boydiae (D. C. Eaton) W. H. Wagner Syn. Christella boydiae (D.C. Eaton) Holttum Distribution: O/Ma Syn. Christella cyatheoides (Kaulf.) Holttum Distribution: K/O/Mo/L/Ma/H Cyclosorus dentatus (Forssk.) Ching Syn. Christella dentata (Forssk.) Brownsey & Jermy Distribution: K/O/Mo/L/Ma/H ® Cyclosorus hudsonianus (Brack.) Ching Syn. Pneumatopteris hudsoniana (Brack.) Holttum Distribution: K/O/Mo/L/Ma/H ^Cyclosorus interruptus (Willd.) H. Ito Distribution: K/O/Mo/L/Ma/H Cyclosorus parasiticus (L.) Farw. Syn. Christella parasitica (L.) H. L6v. Distribution: K/O/Mo/L/Ma/H Syn. Pneumatopteris pendens D. D. Palmer Distribution: K/O/Mo/Ma/H ^ Cyclosorus sandwicensis (Brack.) Copel. Syn. Pneumatopteris sandwicensis (Brack.) Holttum Distribution: K/O/Mo/L/Ma/H ® Cyclosorus wailele (Flynn) W. H. Wagner Syn. Christella wailele (Flynn) D. D. Palmer Distribution: K Macrothelypteris (H. It6) Ching Macrothelypteris torresiana (Gaudich.) Ching Distribution: K/O/Ma/H ® Pseudophegopteris keraudreniana (Gaudich.) Holttum Distribution: K/O/Mo/L/Ma/H 100 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Thelypteris Schmidel ^ Thelypteris globulifera (Brack.) C. F. Reed Syn. Amauropelta globulifera (Brack.) Holttum Distribution: K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H ^ Blechnum orientale L. Distribution: O ^ Doodia kunthiana Gaudich. Distribution: K/O/Mo/L/Ma/H ^ Doodia lyonii O. Deg. Distribution: K/O/Ma/H (ex) Sadleria Kaulf. ® Sadleria cyatheoides Kaulf. Distribution: K/O/Mo/L/Ma/H Distribution: K/O/Mo/L/Ma/H ^ Sadleria souleyetiana (Gaudich.) T. Moon Distribution: K/O/Mo/L/Ma/H " Sadleria squarrosa (Gaudich.) T. Moore Distribution: K/O/Mo/L/Ma/H ^ Sadleria unisora (Baker) W. J. Rob. 20. Ath 3 uiaceae Alston Athyrium Roth ^ Athyrim Distribution: K/O/Mo/L/Ma/H VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Deparia Hook. & Grev. ^ Deparia fenziiana (Luerss.) M. Kato Distribution: K/O/Mo/L/Ma/H ^ Deparia kaalaana (Copel.) M. Kato Distribution: K (ex)/Ma (ex)/H (e> Distribution: K/Mo/L/Ma/H Deparia petersenii (Kunze) M. Kato Distribution: K/O/Mo/L/Ma/H ^ Deparia prolifera (Kaulf.) Hook. & Grev. Distribution: K/O/Mo/L/Ma/H Diplazium Sw. ® Diplazium amottii Brack. Distribution: K/O/Mo/L/Ma/H Diplazium esculezttum (Retz.) Sw. Distribution: K/O/Mo/L/Ma/H Distribution: K (ex)/0 (ex}/Mo (ex)/L (ex)/Ma Distribution: K/O/Mo/L/Ma/H 21. Dryopteridaceae Herter ® Arachniodes insularis W. H. Wagner Distribution: K/O/Mo/Ma/H Ctenitis (C. Chr.) C. Chr. ® Ctenitis latifrons (Brack.) Copel. Distribution: K/O/Mo/L/Ma/H ® Ctenitis squamigera (Hook. & Am.) Copel. Distribution: K/O/Mo/L/Ma Cyrtomium C. Presl ' Cyrtomium caryotideum (Wall. : Hook. & Grev.) C. Presl Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Cyrtomium falcatum (L. f.) C. Presl Distribution: K/O/Mo/L/Ma/H Dryopteris Adans. Dryopteris crinalis (Hook. & Am.) C. Chr. Distribution: K/O/Mo/L/Ma/H r. podosora (W. H. Wagner) D. D. Pal Dryopteris fuscoatra (Hillebr.) W. J. Rob. Distribution: K/O/Mo/L/Ma/H Distribution: Ma Dryopteris glabra (Brack.) Kuntze ® var. alboviridis (W. H. Wagner) D. D. Palmer Distribution: K Distribution: K Distribution: K/O/Mo/L/Ma/H ® var. hobdyana (W. H. Wagner) D. D. Palmer ® var. nuda (Underw.) Fraser-Jenk. Distribution: K/O/Mo/Ma ^ var. pusilla (Hillebr.) Fraser-Jenk. var. soripes (Hillebr.) Herat ex Fra Distribution: K/O/Mo/Ma/H ^ Dryopteris hawaiiensis (Hillebr.) W. J. Rol Distribution: K/O/Mo/Ma/H Distribution: K/O/Mo/L/Ma/H VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS ® Dryopteris rubiginosa (Brack.) Kuntze Syn. Nothoperanema rubiginosum (Brack.) A.R. Sm. & D. D. Distribution: K/O/Mo/L/Ma/H ^ Dryopteris sandwicensis (Hook. & Am.) C. Chr. Distribution: K/O/Mo/L/Ma/H ^ Dryopteris subbipinnata W. H. Wagner & Hobdy Dryopteris tetrapinnata W. H. Wagner & Hoi Distribution: Ma Dryopteris unidentata (Hook. & Am.) C. Chr. ^ var. paleacea (Hillebr.) Herat ex Fra Distribution: K/O/Mo/Ma/H Distribution: K/O/Mo/L/Ma/H ^Dryopteris wallichiana (Spreng.) Hyl. Distribution: K/O/Mo/Ma/H Schott ex J. Sm. Elaphoglossum aemulum (Kaulf.) Brack. Distribution: K/O/Mo/L/Ma/H (ex) Gaudich. Distribution: O ^ Elaphoglossum crassifolium (Gaudich.) W. R. And Distribution: K/O/Mo/L/Ma/H ^ Elaphoglossum fauriei Copel. Distribution: O/Mo ^Elaphoglossum paleaceum (Hook. & Grev.) Sledge Distribution: K/O/Mo/L/Ma/H Distribution: Mo/L/Ma/H ^ Elaphoglossum pellucidum Gaudich. Distribution: K/O/Mo/Ma/H AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) ^ Elaphoglossum wawrae (Luerss.) C. Chr. Distribution: K/O/Mo/Ma/H Polystichum Roth Distribution: Ma/H ^ Polystichum haleakalense Brack. Distribution: Ma/H ^ Polystichum hillebrandii Carruth. Distribution: Ma/H Nephrolepidaceae Pic. Serm. Nephrolepis Schott Nephrolepis brownii (Desv.) Hovenkamp & Miyam. Syn. Nephrolepis multiflora (Roxb.) F. M. Jarrett ex C. V. Morton Distribution: N/K/O/Mo/L/Ka/Ma/H Distribution: K/O/Mo/L/Ma/H fa (L.) Schott subsp. hawaiiensis V Distribution: N/K/O/Mo/L/Ma/H ^ Nephrolepis falcata (Cav.) C. Chr. ‘Furcans’ Distribution: K/O/Mo/L/Ma/H fuia (G. Forst.) C. Presl ‘Superha’ Tectariaceae Panigrahi Tectaria Cav. ^ Tectaria gaudichaudii (Mett.) Maxon Distribution: K/O/Mo/L/Ma/H Tectaria incisa Cav. Distribution: K/O/Ma/H Davalliaceae M. R. Schomb. ex A. B. Frank ^ Davallia fejeensis Hook. Distribution: O/Ma VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS 25. Polypodiaceae Bercht. & J. Presl ^ Adenophorus abietinus (D. C. Eaton) K. A. Wilson Distribution: K/O/L/Ma ^ Adenophorus haalilioanus (Brack.) K. A. Wilson Distribution: K/O ® Adenophorus hymenophylloides (Kaulf.) Hook. & Grev. Distribution: K/O/Mo/L/Ma/H ^ Adenophorus oahuensis (Copel.) L. E. Bishop ns L. E. Bishop Distribution: K/O (ex)/Mo/L (ex)/Ma (ex)/H Adenophorus pinnatifidus Gaudich. Distribution: K/O/Mo/L/Ma/H ^ var. rockii (Copel.) D. D. Palmer Distribution: K/O/Mo/Ma Adenophorus tamariscinus (Kaulf.) Hook. & Grev. ® var. montanus (Hillebr.) L. E. Bishop Distribution: Mo/Ma/H Distribution: K/O/Mo/L/Ma/H Grammitis tenella Kaulf. Distribution: K/O/Mo/L/Ma/H ® Adenophorus tripinnatifidus Gaudich. Distribution: K/O/Mo/L/Ma/H Lepisorus (J. Sm.) Ching * Lepisorus thunbergianus (Kaulf.) Ching Distribution: K/O/Mo/L/Ma/H AMERICAN FERN JOURNAL: VOLUME 1 ^ var. pentadactylum (Hillebr.) D. D. Pali i NUMBER 2 (2013) Distributio] : K/O/Mo/L/Ma/H Oreogrammitis Copel. Syn. Grammitis baldwinii (Baker) Copel. ® Oreogrammitis forbesiana (W. H. Wagner) Parris Syn. Grammitis forbesiana W. H. Wagner Distribution: K/Mo/Ma ^ Oreogrammitis hookeri (Brack.) Parris Syn. Grammitis hookeri (Brack.) Copel. Distribution: K/O/Mo/L/Ma/H note: Oreogrammitis hookeri was previously t indigenous species, but is now considered as dis found in Fiji and American Samoa (B. Parris, pers. Phlebodium (R. Br.) J. Sm. Phlebodium aureum (L.) J. Sm. Distribution: K/O/Mo/L/Ma/H Phymatosorus grossus (Langsd. & Fisch.) Brownlie Distribution: K/O/Mo/L/Ma/H m. f.) Pic. Serm. Platycerium Desv. Platycerium bifurcatum (Cav.) C. Chr. Platycerium superbum de Jonch. & Hennipman Polypodium L. Polypodium pellucidum Kaulf. ^ var. acuminatum D. D. Palmer Distribution: K/Ma VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS ^ var. pellucidum Distribution: K/O/Mo/L/Ma/H Distribution: Mo/Ma/H Pyrrosia Mirb. Pyrrosia piloselloides (L.) M. G. Price Distribution: O Syn. Lellingeria saffordii (Maxon) A. R. Sm. & R. C. Distribution: K/O/Mo/L/Ma/H I 1111 VERNON & RANKER: FERNS AND LYCOPHYTES OF HAWAIIAN ISLANDS Appendix 2. Continued. Huperzia filiformis Huperzia haleakalae Huperzia pbyllantha Huperzia serrata Lycopodiella cemua Tropical America Boreal Asia/Indopacific American Fern Journal 103(2):112-117 (2013) Myriopteris windhamii sp. nov., a New Name For Cheilanthes villosa (Pteridaceae) Amanda L. Grusz Department of Biology, Duke University, Durham, North Carolina 27708, USA, e-mail; alg3@duke.edu Cheilanthes (Pteridaceae) reveal that the genus is highly polyphyletic. To achieve a monophyletic generic classification of the 400-1- taxa of cheilanthoid ferns, it is necessary to transfer species that are only distantly related to the type species [Cheilanthes micropteris) to other genera. One of these species is Cheilanthes villosa Davenp. ex Maxon, which needs to he reassigned to the genus Myriopteris. Because the epithet villosa is preoccupied in Myriopteris and there are no synonyms for this distinctive taxon, a new name is required. The species is herein renamed Myriopteris windhamii. Key Words.— apomixis, cheilanthoid, myriopterid, triploid Recent molecular systematic studies of cheilanthoid ferns (Pteridaceae) have confirmed that the globally distributed genus Cheilanthes Sw. is highly polyphyletic (Eiserhardt et al. 2011; Gastony and Rollo 1998; Li et al 2012; Link-Perez et al. 2011; Prado et al. 2007; Rothfels ef al. 2008; Schuettpelz et al. 2007; Windham et al. 2009; Zhang et al. 2007). Species currently assigned to Cheilanthes occur in five of the six major cheilanthoid clades identified by Windham et al. (2009) and Eiserhardt et al. (2011). To achieve a monophyletic generic classification of the 400-t taxa of cheilanthoid ferns that includes more than a single genus, it is necessary to transfer to other genera those species that are only distantly related to the South American generitype, Cheilanthes micropteris Sw. A series of recent papers (e.g., Li et al. 2012; Perez-Link et al. 2011; Yatskievych and Arbelaez 2008) have initiated this process of narrowing and refining the circumscription of Cheilanthes. Continuing this effort, we are preparing to resurrect the genus Myriopteris Fee (Grusz et al. in prep), which, in its emended form, will include most of the North American taxa currently residing in Cheilanthes. The clade that will become Myriopteris Fee emend Grusz & Windham is a well supported monophyletic group that comprises roughly 10% of chei- lanthoid species diversity. It is among the earliest diverging lineages of cheilanthoid ferns and only distantly related to the type species of Cheilanthes. Most of the 40-f species in this group are easily transferred from Cheilanthes to Myriopteris, the most obvious exception being Cheilanthes villosa Davenp. ex Maxon. The name Myriopteris villosa was previously published by Fee (1852) based on a collection now associated with Cheilanthes lendigera (Cav.) Sw. (= Myriopteris lendigera (Cav.) Fee). Because the epithet villosa is thus preoccupied in Myriopteris and there are no synonyms for the GRUSZ: MYRIOPTERIS WINDHAMII SP. NOV. 113 distinctive taxon known as Cheilanthes villosa, a new name is required. This apomictic triploid species is here renamed Myriopteris windhamii. Myriopteris windhamii Grusz, sp. nov. Type. — USA. Arizona: Cochise Co., SSW of Sierra Vista in the Huachuca Mts. along Copper Canyon, ca. 0.25 trail miles NE of its intersection with W Montezuma Canyon Road. Lat./ Long/: 31.36385N 110.29794W (WCS84 Datum). Elevation 6175 feet. 16 August 2013, Windham 4165 (holotype: DUKE; isotypes: ARIZ, ASC, ASU, GH, MO, NMC, NY, TEX/LL, UNM, US, UT). Figs. 1-2. Cheilanthes villosa Davenp. ex Maxon, Proc. Biol. Soc. Wash. 31: 142. 1918; non Myriopteris villosa Fee Rhizomes compact, scales linear-lanceolate, entire to slightly erose, 0.2- 0.6 mm wide, often with a well defined, lustrous, dark central stripe and light brown margins, with tufted concolorous scales at rhizome apices and stipe bases, straight to slightly contorted, loosely appressed, persistent; leaves clustered, 7-30 cm long, with non-circinate vernation; blades narrowly oblong-lanceolate, 3^-pinnate at the base; petioles dark reddish brown, lustrous, terete, with scattered, ascending, lanceolate scales, largely paleac- eous, darker at point of attachment, mixed with abundant, hair-like scales; rachises with mostly hair-like scales adaxially and broader lanceolate to ovate scales abaxially, otherwise similar to petioles; pinnae 8-20 pairs, not articulate, with dark color of rachises continuing into pinna bases, basal pair usually slightly smaller than adjacent pair, equilateral, appearing villous above and scaly below; costae adaxial surface greenish except at base, with scattered hairs and hair-like scales, abaxial surface of costae dark and lustrous except at apex, with conspicuous ovate-lanceolate, truncate to shallowly cordate scales, these 0.7-1. 2 mm wide, strongly imbricate, spreading laterally, with erose- dentate margins, not ciliate, brown proximally, paleaceous distally, attenuate to hair-like apices; ultimate segments round to elliptical, the largest 1-2 mm wide, with coarse, contorted, whitish or translucent hairs adaxially and occasional hairs and narrow scales abaxially; leaf margins recurved, forming a weakly differentiated marginal false indusium to 0.2 mm wide; sporangia attached near vein endings, becoming confluent at maturity, partially covered by false indusium; spores 32 per sporangium, 60.8 ± 2.47 pm in diameter; n = 2n = 9011 (apomictic; Windham and Yatskievych 2003). Parafypes.— U.S.A. Arizona: Cochise Co., west wall of Copper Canyon in the Huachuca Mts. ca. 2.42 km NNW of the summit of Coronado Peak, Coronado National Forest, Montezuma Pass Quad (7 Vz min.), UTM - 3469960 m. N by 566775 m. E (Zone 12), elevation 6250 ft., 5 August 1983, Windham 458. Pima Co., Waterman Mts., N-facing slope on NE end of range, T12S R8E S24, pitted limestone with Larrea, Cereus giganteus, Cercidium, 3200 ft., Bruce Parfitt 2786, 24 Mar 1979 (NY); Cochise Co., Coronado National Forest, Huachuca Mts., east side of Copper Canyon at a point 3.04 km. NNW of the summit of Coronado Peak, Montezuma Pass Quadrangle (7 ‘A min.), Universal Transverse AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) 395986 GRUSZ: MYRIOPTERIS WINDHAMII SP. NOV. Fig. 2. Myriopteris windhamii [Pringle 459, NY). A. Rhizome and partial frond. B. Stipe detail. C. Rhizome scale. D. Pinnule, adaxial surface. E. Pinnule, abaxial surface. F. Pinnule, abaxial surface, with scales removed. G. Blade scale. Scale bars = 1 mm. Reproduced from The Pteridophytes of Mexico (Mickel and Smith, 2004) with permission of The New York Botanical Garden Press. Mercator- 3470625m.N by 566650m.E (Zone 12), growing from crack under overhang on cliff face ca. 6 meters E. of streambed, about 25° facing west, locally common on limestone outcrops at lower elevations., associated genera Quercus, Bhckellia, Pinus, Garrya, Cercocarpus, and Junipems, Windham 0236, 17 March 1981 (UT, ASC). New Mexico: Eddy Co., Lincoln National Forest, along rd. 409 to Sitting Bull Falls, open NW facing (309°) limestone cliff with Quercus, Dasylirion, Opuntia, Rhus. Gutierrezia, N 32.2630, W 104.6807, 1356 m. Beck 1050, 02 May 2008 (NY, DUKE). Texas: El Paso Co., Tom Mays County Park, Franklin Mountain, northwest edge of El Paso, common under rocks in ravine on limestone hillside in Chihuahuan Desert, with Notholaena, Cheilanthes, Pellaea, Agave. Yucca, Dasylirion, Quercus, and grasses, elevation 4400 feet, Wollenweber & Yatskievych 81-510, 15 December 1981 (ARIZ). Etymology.— This name honors Dr. Michael D. Windham, acknowledging his lifelong dedication to the study of cheilanthoid ferns and, in particular, his devotion to understanding the evolution and cytogenetics of apomictic polyploids (including his new namesake in Myriopteris). American Fern Journal 103(2j:118-130 | Polystichum montevidense Demystified: Molecular and Morphological Data Reveal a Cohesive, Widespread South American Species AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) insights into the species biology and taxonomy of this long-misunderstood taxon. Methods Taxon sampling— Out sample consists of 17 Polystichum accessions from across the northern and central Andean region, Argentina, Uruguay, and Brazil. We sampled all taxa shown to have close evolutionary relationships to P. montevidense in the phylogeny of McHenry (2012). Also included are three accessions from the senior author’s dense sampling of the Serra do Mar region, and one accession of P. montevidense from the tropical Andes, one from Argentina, and one from Uruguay, to adequately represent the morphological diversity of the group across its range. We excluded accessions that showed signs of hybridity (intermediate morphology and misshapen spores). Voucher data for all of our collections are listed in the Appendix. DNA extraction, amplification and sequencing— Total genomic DNA was extracted from fresh (O.lg) or silica-dried (0.02g) material. Leaf material collected in the field was preserved fresh at 4°C or in silica desiccant gel and stored at -80°C until extraction. Total genomic DNA was extracted from pinnules following a modified CTAB protocol (Doyle and Doyle, 1987). Four plastid DNA regions were amplified using by PGR: two genes [rps4 and rbcL] and two intergenic spacers (the region between trnL and trnF [trnLF] and the region between trnS and rps4 {trnS-rps4]). Primers for amplification and sequencing were taken from the literature: rbcL (Little and Barrington, 2003), rps4 (Shaw et ah, 2005), trnLF (Taberlet et al, 1991), and trnS-rps4 (Souza- Chies et al., 1997). PGR amplification was performed in a TG-312 or TG-3000 thermal cycler (Techne, Burlington, New Jersey, USA) following protocols in McKeown et al. (2012). PGR products were cleaned using ExoSAP-IT (USB Gorporation, Gleveland, OH, USA). Sequencing of the cleaned PGR products employed a cycle-sequence reaction using the BigDye Terminator Gycle Sequence Ready Reaction Kit v. 3.1 (Perkin-Elmer/ Applied Biosystems, Foster Gity, GA, USA). Sequences were resolved on an ABI Prism 3100-Avant Genetic Analyzer (Vermont Gancer Genter DNA Analysis Facility, Burlington, VT, USA). Gonsensus sequences from the raw chromatographs (using both the forward and reverse reads) were assembled for each gene using Sequencher 4.5 (Genes Gode Gorporation, Ann Arbor, MI, USA) or Geneious Pro v.5.0.3 (Drummond et al., 2007). Sequence alignment and phylogenetic analysis. — Gonsensus sequences were aligned with MUSGLE (Edgar, 2004) as implemented in Geneious Pro v.5.0.3 (Drummond et al., 2007). All phylogenetically informative indels were coded following the simple gap coding of Simmons and Ochoterena (2000) and added as additional binary characters at the end of the NEXUS file. The concatenated sequences were analyzed by Bayesian inference (BI) using MrBayes v. 3.2 (Ronquist et al., 2012). The data were partitioned by plastid region, and optimal models (Table 1) were applied to each of the molecular partitions. The model selection was done using jModelTest v. 2.1 under the Akaike CONDACK ET AL.: POLYSTICHUM MONTEVIDENSE Table 1. Characteristics of the i ! used in the phylogenetic analysis. K80+G TPM2uf+G TPM3uf+G TrN+G Length (hp) Variable sites Sampled taxa 1165 33 (2.8%) 17 331 29 (8.8%) 17 450 25 (5.6%) 17 460 50 (10.9%) 17 Information Criterion (AIC; Darriba et al. 2012). We employed a mixed-model Bayesian analysis with different models allowed for each partition. The parameters for each data partition were able to vary freely. Indels were treated as single, independent, and binary characters. BI was preformed in Mr. Bayes by running two independent analyses for 5 million generations with trees sampled every 1,000 generations. Stationarity was calculated by plotting the log-likelihood scores for each run against generation in the program Tracer vl.5 (Rambault and Drummond, 2007). All trees prior to this point (the first 500,000) were discarded as the burn-in phase and a 50% majority rule consensus tree was calculated for the remaining trees. Chromosome analysis. — Sporophytes collected from wild populations from Cordoba province, Argentina, were cultivated at the Universidad de Cordoba; voucher specimens were deposited at CORD. For the analysis of mitotic chromosomes, croziers were excised into small fragments approximately 2 mm wide. The fragments were treated in a solution of 2mM 8-hydroxiquinoline for 1 h at room temperature, followed by 8 h at 14°C, then fixed in ethanol-acetic acid (3:1) and stored at -20°C. For chromosome analysis, the crazier fragments were washed in distilled water 4 to 5 times, then hydrolyzed in 3 ml of cellulase 2%-pectinase 20%, for 2 h at 37°C. The hydrolyzed cells were stained in alcoholic hydrochloric acid-carmine and squashed following established protocols (Manton, 1950). Results Molecular analysis.— The analysis of South American accessions relevant to the disposition of the name P. montevidense yielded a highly resolved phylogeny (Fig. 1). In this phylogeny, our accession from the area of the type collection near Montevideo, Uruguay, resolves in a clade with Argentine and Bolivian accessions of P. montevidense (Clade 2; posterior probability 1PP] = 1.0). This trio of accessions is sister to a clade of Andean plants including P. solomonii and P. alhomarginatum. Together they form a well- supported clade (PP=0.96) with P. opacum, the Brazilian representative of the lineage that includes the widespread P. platyphyllum. In contrast, our accessions of specimens matching plants commonly determined as P. montevidense from the northern Serra do Mar, Brazil, resolve in a well- supported clade (Clade 1, PP=1.0) with the tropical Southeast Brazilian endemic species P. platylepis and P. pallidum (Condack, 2012). Morphological analysis.— \^e found that the Sello collections of Polypodium montevidense from near Montevideo, Uruguay (Brazil in Sello’s time), which 122 (2013) CONDACK ET AL.: POLYSTICHUM MONTEV1DENSE 123 124 LECTOTYPE (here designated).-Uruguay, n. ■•Brasilia, prope Monte Video’’!, Sello d 654 (B 20 0156391, B 20 0156394, B 20 0156395, B Syst. Veget. ed. 16. 4(1): : AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Santibanez, F. 2004. Preliminary Map of Arid Zones of South America Demo Version. Facultad de Ciencias Agronomicas Universidad de Chile. Available at http://www.cazalac.org/eng/ mapa_alc_eng.php. Sehnem, a. 1979. Aspidiaceas, In: R. Reitz, ed. Flora Ilustrada Catarinense. Itajai, Herbario Barbosa Rodrigues. Shaw, J., E. B. Lickey, J. T. Beck, S. B. Farmer, W. Liu, J. Miller, K. C. Siripun, C. T. Winder, E. E. Schilling and R. L. Small. 2005. The tortoise and the hare II: relative utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. Amer. J. Bot. 92:142-166. Simmons, M. P. and H. Ochoterena. 2000. Gaps as characters in sequence-based phylogenetic analyses. Syst. Biol. 49:369-381. Souza-Chies, T. T., G. Bittar, S. Nadot, L. Carter, E. Besin and B. Lejeune. 1997. Phylogenetic analysis of Iridaceae with parsimony and distance methods using the plastid gene rps4. Pi. Syst. Evol. 204:109-123. Sprengel, K. P. j. 1827. Carol! Linnaei Systema Vegetabilium, Editio Decima Sexta. 4(1). Dieterich, Gottingen. Taberlet, P., L. Gielly, G. Pautou and J. Bouvet. 1991. Universal primers for amplification of three non-coding regions of chloroplast DNA. PI. Molec. Biol. 17:1105-1109. Thiers, B. 2012. Index Herbariorum, a global directory of public herbaria and associated staff. New York: Botanical Garden’s Virtual Herbarium. Available at http://sweetgum.nybg.org/ih/ (accessed 1 August 2012). Tryon, R. M. and R. G. Stolze. 1991. Pteridophyta of Peru. Part IV. 17. Dryopteridaceae. Fieldiana, Bot. N. S. 27:1-176. Wagner, D. H. 1979. Systematics of Polystichum in western North America and north of Mexico. Pteridologia 1:1-64. Zhang, L. B. and D. S. Barrington, in press. Polystichum. In Wu, Z. Y., P. H. Raven and D. Y. Hong, eds. Flora of China. Vol. 2-3 (Lycopodiaceae through Polypodiaceae). Science Press, Beijing, and Missouri Botanical Garden Press, St. Louis. Selected specimens examined including all those used in the molecular analysis, which have GenBank accession numbers (marker sequence is rbcL, rps4 gene, tmLF, tmS-rps4). Herbarium abbreviations throughout follow Index Herbariorum (Theirs, 2012). GAN = GenBank accession number, Polystichum albomarginatum M, Kessler & A.R. Smith. ECUADOR. Zamora- Chinchipe: Bombuscaro, Podocarpus National Park, M. Lehnert 1257 (VT) (GANs: KC819964; KC819994; KC819949; KC819979). Polystichum maximum M. Kessler & A.R. Smith. BOLIVIA. Cochabamba: Prov, Ayopaya, San Cristobal, 16°39'S, 66°43'W, 2550, 1. Jimenez 1292 (LPB) (GANs: KC819965; KC819995; KC819950; KC819980). Polystichum montevidense (Spreng.) Rosenst. BOLIVIA. Santa Cruz: Caballe- ro, M. Sundue 844 (VT, NY). Cochabamba: Prov, Carrasco, 3.3 km NW of Kayarani, M. Sundue 621 (VT, NY). La Paz: Calle Sorata a Laripata, M. McHenry 10-95 (VT); Sorata, M. McHenry 10-100 (VT) (GANs: KC819967; KC819997; KC819952; KC819982). Tarija: Arce, Padcaya, H. Huaylla 1514 (NY). ARGENTINA. Cordoba: Dpto. Punilla, Los Gigantes, R. Morero 342 CONDACK ET AL.: POLYSTICHUM MONTEVIDENSE (CORD), 344 (CORD) (GANs: KC819968; KC819983; KC819953; KC819998). Tucuman: Dpto. Chicligasta. Cuesta del Clavillo, R. Morero 352 (CORD). URUGUAY. Montevideo: without precise locality, F. Sello d 654 (B, P); Maldonado: Cerro Pan de Aziicar, Gibert 857 (K); Idem, A. Lombardo (BM ex MVM10907); Idem, A. Lombardo 2568 (MVJB); Idem, A. Lombardo 4732 (MVJB); Idem, A. Lombardo 5436 (MVJB); Idem, /. Arechavaleta 784 (MVFA); Idem, O. del Puerto et ah 9702 (MVFA); Idem, E. Marches! 1379 (MVFA); Idem, M. Berro 1261 (MVFA); Idem, /. Condack & F Munoz 687 (R) (GANs: KC819966; KC819996; KC819951; KC819981); Idem, /. Condack 694 (R). Tacuarembo: Montes de Tacuarembo, /. Arechavaleta 4187 (MVM). URUGUAY/BRASIL. Artigas: Cuarein (border of Brazil and Uruguay), M. Berro 5617 (MVFA). Polystichum nudicaule Rosenst. COLOMBIA. Boyaca: Mun. Chisacd-San Pedro de Iquaque, M. Sundue 1285 (VT) (GANs: KC819969; KC819999; KC819954; KC819984). Polystichum opacum Rosenst. BRASIL. Rio Grande do Sul: Santa Cruz do Sul, /. Condack S^M. Brito 651 (R, VT) (GANs: KC819970; KC820000; KC819955; KC819985). Polystichum pallidum Gardn. BRASIL. Rio de Janeiro: Parque Nacional da Tijuca, /. Condack 662 (R, VT) (GANs: KC819971; KC820001; KC819956; KC819986). Polystichum platylepis Fee BRASIL. Bahia: Amargosa, /. Paixao S' M. Nascimento 1377 (UEFS). Espirito Santo: Castelo, Parque Estadual do Forno Grande, P. Labiak et al. 4820 (MBML, RB, UPCB). Mato Grosso: Vila Bela da Santissima Trindade, Serra Ricardo Franco, P. Windisch 1350 (GH, HB, ICN); Minas Gerais: Bocaina de Minas, Parque Nacional do Itatiaia, Alto dos Brejos, /. Condack et al. 328 (RB, NY); Parana: Clevelandia, G. Hatschbach 22695 (MBM, NY, PACA); Rio de Janeiro: Itatiaia, /. Condack & A. Vasco 636 (R, RBR, VT) (GANs: KC819974; KC820004; KC819959; KC819989). Rio Grande do Sul: Sao Jose dos Ausentes, /. Condack S' P. Schwartsburd 588 (R, RBR, VT) (GANs: KG819973; KC820003; KC819958; KC819988). Santa Catarina: Urubici, Morro da Igreja, /. Condack S' P. Schwartsburd 579 (R, RBR, VT) (GANs: KC819972; KC820002; KC819957; KC819987). Sao Paulo: Bananal, Serra da Bocaina, A. Brade 15203 (RB). Polystichum platyphyllum (Willd.) C. Presl. BRASIL. Goias: Caiaponia, Serra do Caiapo, H. Irwin et al. 17873 (GH, NY, RB, UB, US). Polystichum pycnolepis (Kunze ex Klotzsch) T. Moore. COLOMBIA. Boyaca: Mun. Villa de Leyva, M. Sundue 1275 (VT) (GANs: KC819975; KC820005; KC819960; KC819990). 130 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Polystichum rochaleanum Fee. BRASIL. Rio de Janeiro: Itatiaia, Pico das Agulhas Negras, /. Condack d‘ C. Ramos 516 (RB, VT) (GANs: KC819976; KC820006; KC819961; KC819991). Polystichum rufum M. Kessler & A. R. Smith. BOLIVIA. F. Tamaya: PN-ANMI Madidi, I. Jimenez 1078 (UC) (GANs: KG819977; KG820007; KC819962; KG819992). Polystichum solomonii M. Kessler & A. R. Smith. BOLIVIA. Caballero: 1.5 km down from Empalme, M. Sundue 775 (VT) (GANs: KG819978; KG820008; KG819963; KG819993). 132 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) Fig. 1. The best ML tree from the analysis of combined afpA, atpB, rbcL and rps4 data. Bootstrap support is shown for each node. Lomariopsidaceae is highlighted and the position of Dracoglossum within it is indicated. Classification of ferns has been contentious since the beginnings of fern taxonomy, and even though general morphology remains extremely useful for establishing natural groupings in ferns (e.g. Schneider, 2013; Schneider et al, 2009; Smith, 1995), some genera remained unplaced even after general phylogenetic studies of ferns that included most fern genera, have been carried out (e.g. Hasebe et al, 1995; Manhart, 1995; Pryer et al, 2004; Schuettpelz et al, 2006; Schuettpelz and Pryer, 2007; Lehtonen, 2011). Subsequent placement of enigmatic genera has varied. Some genera have been embedded within larger genera, such as Hymenophyllopsis K. 1. Goebel that is now loiown to belong to Cyathea Sm. (Korall et al, 2007; Christenhusz, 2009a); CHRISTENHUSZ ET AL.: PLACEMENT OF DRACOGLOSSUM 133 in Other cases clades have been divided into separate genera (e.g. Lehtonen et ah, 2010; Moran et al., 2010). A number of genera were found to be isolated, resulting in new families needing to be (re-)erected to accommodate them, such as Hypodematiaceae (e.g. Liu et al., 2007; Tsutsumi and Kato, 2006; Schuettpelz and Pryer, 2007), Lonchitidaceae (e.g. Christenhusz, 2009b; Lehtonen et al., 2010; Lehtonen et al., 2012), Diplaziopsidaceae and Rhachidosoraceae (Li et al., 2011; Christenhusz et al., 2011). In a few cases the ultimate placement was found to be far from where they were originally placed. Psilotaceae were originally placed in the lycopodio- phytes (e.g. Reimers, 1954), and were even associated with the earliest fossil land plants, but it has now been shown (Hasebe et al., 1995; Manhart, 1995) that this family forms a clade with Ophioglossaceae. Another good example of the use of molecular phylogeny in the placement of a genus is Cystodium J. Sm. in Hook. Originally thought to be a tree fern belonging to Dicksoniaceae (Christensen, 1938; Sen and Mittra, 1966), numerous morphological characters had shown it did not belong there (e.g. Croft, 1986), but nevertheless Cystodium remained within Dicksoniaceae (e.g. Kramer & Green, 1990) until a molecular phylogeny showed that it rather belonged among early branching polypods (Korall et al., 2006; Lehtonen et al., 2012). Only a few of these enigmatic genera now remain among ferns. Most remaining unsampled genera are either impossible to obtain material for, simply because they have not been recently collected, such as Hypoderris R.Br. in Wallich, Oenotrichia Copel., Psomiocarpa C. Presl, Thysanosoria Gepp and Xyropteris K. U. Kramer, or have simply slipped the attention of phylogenetic treatments. Dracoglossum is such an unsampled genus and is one of the few remaining genera that are of unknown affinity. Material from the original collection of Christenhusz (2005) was not available for analysis so a sample of Dracoglossum plantagineum (M. Jones 1018, TUB] from recent fieldwork on fern communities in central Panama (Jones et al., 2013) was included. A small population of the species was found growing on stones in a shallow stream in the San Lorenzo protected area, ca. 5 km from the Caribbean coast. We analysed this sample to establish the placement of the genus in the general phylogeny of ferns. Materials and Methods Taxon and character sampling.— We sampled four chloroplast genes {atpA, atpB, rbcL, rps4) from Dracoglossum in order to reveal its phylogenetic position. Total genomic DNA was extracted with an E.Z.N.A. SP Plant DNA Kit (Omega Bio-tek, Doraville, GA). PCR reactions were performed using PureTaq RTG PCR beads (Amersham Biosciences, Piscataway, NJ, USA) and primers ATPF412F and TRNR46F for atpA (Schuettpelz et al., 2006), ATPF412F and TRNR46F for atpB (Schuettpelz et al., 2006), ESRBCLlF and ESRBCL1361R for rbcL (Korall et al, 2006), and tmS^^ and rps4.5' for rps4 (Shaw et al., 2005). Purification of PCR reactions and DNA sequencing were carried out under BigDye™ terminator cycling conditions by Macrogen Inc., Seoul, South Korea/ CHRISTENHUSZ ET AL.: PLACEMENT OF DRACOGLOSSUM phylogenetic analysis on tl K, K. M., E. SCHUETTPELZ, P. G. iBiAK and M. Sundue. 2010. S5mopsis of Mickelia, a newly recognized genus of ms (Dryopteridaceae). Brittonia 62:337-356. S. Aluru and A. Stamatakis. 2007. Large-scale maximum likelihood-based IBM BlueGene/L. Proceedings of ACM/IEEE Supercomputing Reimers, H. 1954. XV. Abteilung: Pteridophyta. Farapflanzen, Pp. 269-311, in H. Melchior and E. Werdermann, eds. A. Engler’s Syllabus der Pflanzenfamilien. Bomtraeger, Berlin-Nikolassee. Schneider, H. 2013. Evolutionary morphology of ferns (monilophytes). Ann. Plant Rev. 45:115-140. Schneider, H., A. R. Smith and K. M. Pryer. 2009. Is morphology really at odds with molecules in estimating fern phylogeny? Syst. Bot. 34:455-475. Schuettpelz, E., P. Korall and K. M. Pryer. 2006. Plastid atpA data provide improved support for deep relationships among ferns. Taxon 55:897-906. Schuettpelz, E. and K. M. Pryer. 2007. Fem phylogeny inferred from 400 leptosporangiate species and three plastid genes. Taxon 56:1037-1050. Sen, U. and D. Mittra. 1966. The anatomy of Cystodium. Amer. Fem J. 56:97-101. Shaw, J., E. B. Lickey, J. T. Beck, S. B. Farmer, W. Liu, J. Miller, K. C. Siripun, C. T. Winder, E. E. Schilling and R. L. Small. 2005. The tortoise and the hare II: relatively utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. Amer. }. Bot. 92:142-166. Smith, A. R. 1995. Non-molecular phylogenetic hypotheses for ferns. Amer. Fem J. 85:104-122. Smith, A. R., K. M. Pryer, E. Schuettpelz, P. Korall, H. Schneider and P. G. Wolf. 2006. A classification for extant ferns. Taxon 55:705-731. Stamatakis, A. 2006. RAxML-Vl-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22:2688-2690. Stamatakis, A. 2008. The RAxML 7.0.4 Manual, The Exelixis Lab. LMU Munich. Tryon, R. M. and A. F. Tryon. 1982. Ferns and allied plants, with special reference to Tropical America. Springer Verlag, New York. Tsutsumi, C. and M. Kato. 2006. Evolution of epiphytes in Davalliaceae and related ferns. Bot. J. Linn. Soc. 151:495-510. Vaidya, G., D. j. Lohman and R. Meier. 2011. SequenceMatrix: concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics 27:171-180. Walker, T. G. 1985. Cytotaxonomic studies of the ferns of Trinidad 2. The cytology and taxonomic implications. Bull. Brit. Mus. (Nat. Hist.), Bot. 13:149-249. Appendix 1. GenBank accessions used in this study (taxa, atpA, atpB, rbcL, rps4). Arachniodes denticulata, EF463664, EF463380, AF537223, Arthropteris palisotii, JF304014, AB575230, Arthropteris parallela, EF463862, EF463522, EF463266, Blechnum gracile, EF463615, EF463351, EF463158, AF313606; Ctenitis eatonii, JF304011, EF450515, U05614, EF540713; Cyclo- peltis crenata, JF304016, EF450534, DQ054517, EF540718; Cyclopeltis semi- cordata, JF832107, EF463480, EF463234, Cyrtomium falcatum, EF463671, EF463387, AF537226, Davallia griffithiana, EF463649, EF463371, EF463165, Davallia mariesii, AB212706, U05617, JX103759; Davallia tyermannii, JX103677, JX103719, JX103761; Davallia yunnanensis, JX103676, JX103718, JX103760; Davallodes borneense, EF463650, AB212694, AB212694, Davallodes hymenophylloides, EF463648, AB212689, AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) AB212689, Didymochlaena truncatula, JF832112, EF452030, JF303975, AF425161; Dracoglossum plantagineum, KC914562, KC914563, KC914564, KC914565; Dryopteris filix-mas, JF832119, JF832164, JN189507, HQ680978; Elaphoglossum peltatum, EF463701, EF463417, EF463201, Lomariopsis pollicina, EF463776, EF463481, EF463235, HM748162; Lomariopsis sorbifolia, EF463777, EF463482, EF463236, Lomariopsis spectabilis, JF304015, EF450533, AB575226, Loxogramme abyssinica, EF463826, EF463498, EF463252, DQ164474; Nephrolepis biserrata, HQ157268, DQ646105, HQ157305, HQ157329; Nephrolepis cordifolia, EF452103, EF452041, U05637, HM748173; Nephrolepis davallioides, JF832131, JF832172, JF832075, HM748175; Nephrolepis hirsutula, EF463778, EF463483, U05638, HM748179; Oleandra articulata, EF463792, EF463487, EF463242, Onoclea struthiopteris, JF832130, JF832171, AB232415, AF425158; Pleopeltis macro- carpa, EF463838, EF463507, U21152, AY96233; Pleopeltis polypodioides, EF463839, EF463508, JN189568, AY362665; Polybotrya alfredii, EF463717, EF463433, JN189564, Polypodium vulgare, JF832137, JF832178, AB044899, EF551081; Psammiosorus paucivenius, EF463864, EF463524, EF463268, Pteridrys lofouensis, EF450527, EF460687, Serpocaulon triseriale, EF463850, EF463516, EF463263, AY362681; Synammia intermedia, EF463851, EF463517, EF463264, DQ168816; Tectaria (Ctenitopsis) fuscipes, EF450521, DQ508764, Tectaria (Fadyenia) prolifera, EF463869, EF463529, EF463273, Tectaria (Hemigramma) decurrens, EF450529, AB575231, Tectaria (Heterogonium) pinnatum, EF463863, EF463523, EF463267, Tectaria (Quercifilix) zeilanica, JF832143, EF463531, EF463275, Tectaria antioquoiana, EF463865, EF463525, EF463269, Tectaria apiifolia, EF463866, EF463526, EF463270, Tectaria fimbriata, EF463867, EF463527, EF463271, Tectaria incisa, EF463868, EF463528, EF463272, HQ157325; Thelypteris oligocarpa, EF463890, EF463550, EF463294, AF425162; Tricholepidium maculosum, JX103668, JX103710, JX103752; Triplophyllum funestum, EF463872, EF463532, EF463276, 140 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 2 (2013) The moths were identified to genus with the kind assistence of Mr. Alfred Moser (Sao Leopoldo, Brazil), and confirmed hy Dr. Victor O. Becker (Federal University of Brasilia, Brazil, voucher specimen) as Lepidoptera, Crambidae, genus Erupa Walker, Erupa cf. bilineatella (Walker, 1866). In one of the boxes the emergence a parasitoid wasp was also observed. According to Dr. Angelica Pentado-Dias (Federal University of Sao Carlos - UFSCAR, Brazil) this vespid Hymenoptera probably is a parasite of the Erupa larvae (specimen at UFSCAR). Between July 2004 to June 2005, three additional bracken stands in the northeastern region of the State of Rio Grande do Sul (Sapiranga, ca. 40 m alt, ca. 20 km from Sao Leopoldo; Canela, ca. 900 m alt, ca 55 km; Sao Francisco de Paula, 950 m alt., ca. 64 km, forming a poligonal area with ca 650 km^) were visited. In a total sample of 78 leaves, 32.05% were infested, all presenting similar morphological characters as to the damage on the plant material, suggesting a single infesting species and documenting that the infestation was not restricted to the first population. Aside from the damage on the petiole bases, no differences were observed on the infested leaves when compared to those without larvae. Images of the moth and larvae can be provided by the second author. The authors thank the cited entomologists for the identifications and the reviewers for helpful suggestions. — Gabriela Fausti Landoni, (Via Sant’Antonio 66, 21040 Cislago - VA, Italy), and Paulo G. Windisch, (Brazilian Research Council - CNPq grant), PPG-Botanica, Univ. Federal do Rio Grande do Sul, 91540-970 Porto Alegre-RS, Brazil, e-mail: pteridos@gmail.com.br. PTERIDOLOGIA ISSUES IN PRINT The following issues of Pteridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 plus postage and handling. 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choco (Part 1 : Psilotaceae through Dicksoniaceae). 364 pp. $32.00 plus postage and handling. 3. Lellinger, David B. 2002. A Modem Multilingual Glossary for Taxonomic Pteri- dology. 263 pp. $28.00 plus postage and handling. 4. Hirai, Regina Y., and Jefferson Prado. 2012. Monograph of Moranopteris (Polypo- diaceae). 1 13 pp. $28.00 plus postage and handling. For orders and more information, please contact our authorized agent for sales at: Missouri Botanical Garden Press, P.O. Box 299, St. Louis, MO 63166-0299, tel. 314-577- 9534 or 877-271-1930 (toll free). For online orders, visit: http://www.mbgpress.org. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fem Society welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns. SPORE EXCHANGE Mr. Brian S. Aikin, 3523 Federal Ave, Everett, WA 98201-4647 (spores.afs® Comcast, net), is Director. Spores exchanged and lists of available spores sent on request, http:// amerferasoc.org/sporexy.html GIFTS AND BEQUESTS Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Back issues of the Journal and cash or other gifts are always welcomed and are tax-deductible. Inquiries should be addressed to the Membership Secretary. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://amerfernsoc.org/ AMERICAN FERN JOURNAL Volume 103 Number 3 JulySeptember 2013 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY i-You Guo and Hong-Mei Liu 153 Conservation Status of Three Ka David H. Lorence, Kenneth R. 1 f, and Ruth Aguraiuja 166 First Record of S from Mesoamerica Sven Peter Ratke and Nicholas Hill First Report of the Aphid, Amphorophora ampullata (Homoptera: Aphididae) on the Fern, Hypolepis polypodioides (Hypolepidaceae) from Western Himalyas (India) S. G. E. Reddy, A. Kumari, and B. Lai Lygodium japonicum (Thunherg) Swartz in the Piedmont of Georgia Robert Wyatt and Graham E. Wyatt 182 185 188 The American Fern Society Council for 2013 Fem Journal 103(3):141-152 (2013) Molecular Phylogenetic Relationships of Cibotium and Origin of the Hawaiian Endemics Jennifer M. O. Geiger Department of Natural Sciences, Carroll College, Helena, MT 59625 USA, e-mail: jgeiger@carroll.edu Petra Korall Systematic Biology, Evolutionary Biology Centre, Uppsala Uni\ Uppsala, Sweden Tom a. Ranker University of Hawai’i at Manoa, Department of Botany, 3190 M Annabelle C. Kleist Department of Plant Sciences, MS4, University of California, 1 r, Norbyvagen 18D, SE-752 36 Nay, Honolulu, HI 96822 USA OneMISeOWR^WTANICAL MAR 2 1 2014 garden library Abstract.— The tree fem genus Cibotium comprises nine species distributed in tropical regions of Asia, Mesoamerica, and the Hawaiian Islands. The four Hawaiian species are endemic to the Hawaiian Islands. The goals of this paper were to determine the relationships among the Cibotium species, determine whether the Hawaiian species are monophyletic, and infer the dispersal pathway likely responsible for delivering an ancestral Cibotium species to the Hawaiian Islands. Molecular phylogenetic analyses based on four coding and five non-coding plastid DNA sequences supported Hawaiian Cibotium as monophyletic, suggesting a single colonization of the Hawaiian Islands. Hawaiian Cibotium are most closely related to species in Mesoamerica. If the ancestor of Hawaiian Cibotium dispersed to the Hawaiian Islands via wind dispersed spores, our analyses suggest the trade winds or storms delivered spores from Mesoamerica or the Hawaiian Islands were colonized first by a species from Asia, followed by subsequent dispersal to Mesoamerica fi-om Hawai’i. Our analyses do not allow us to favor one hypothesis over the other. Key Words. — Hawai’i, historical biogeography, Cibotium, spore dispersal, molecular phylogenetics. The current high islands of the Hawaiian Islands are located approximately 4000 km from the nearest continent, North America, and 1600 km from the nearest archipelago, the Marquesas Islands of Polynesia. The Hawaiian Island chain is about 80 million years old and was produced in a conveyer-belt-like manner by the Hawaiian hotspot beneath the Pacific Ocean. Islands are removed from the hotspot as the Pacific tectonic plate moves to the northwest (e.g., Clague and Dalrymple, 1987). Once removed from the hotspot, the islands eventually erode to or below sea level. Geological evidence suggests that the islands’ separation from the mainland has been relatively constant (Garson and Clague, 1995) and their extreme isolation has been a barrier to colonization by terrestrial organisms. Because the islands are volcanic in origin and are so isolated, colonizing species must have dispersed to the islands via wind or water. 142 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER : The Hawaiian Islands have the highest levels of species endemism of any regional flora in the world (Wagner et al., 1999), which can be partially explained by the continual opening of new habitats accompanied by dispersal of organisms throughout the developing islands. Of the 136 species of ferns and lycophytes native to the Hawaiian Islands, 67% are endemic to the archipelago (Vernon, 2011). Ferns and lycoph)des comprise about 12% of the native vascular-plant species of the Hawaiian Islands (Warren L. Wagner, 2011 pers. comm, and http://botany.si.edu/pacificislandbiodiversity/hawaiianflora/), although they constitute only about 3% of the vascular-plant flora of the world. During the past 80 Ma the size of islands produced by the activity of the Pacific hotspot has varied (Price and Clague, 2002). Price and Clague (2002) indicate that there was a lull in island formation from 33 Ma to 23 Ma ago. During this time, the lack of available habitats would have resulted in extinction of most or all Hawaiian terrestrial biota such that there was essentially a renewal of terrestrial life beginning 23 Ma ago. Prior to 18 Ma ago, there were few islands with peaks over 1000 m, but between approximately 18 Ma and 11 Ma ago there were numerous larger peaks over 1000 m, including some greater than 2000 m. However, due to subsequent erosion of those larger peaks, the models produced by Price and Clague (2002) indicate there were few islands higher than 1000 m during the 5—6 Myr preceding the formation of Kauai 5.2 Ma ago. Because of this. Price and Clague (2002) suggested that montane taxa, including most of the ferns, probably arrived from outside the Hawaiian archipelago or evolved after the formation of Kauai approximately 5.2 Ma ago because appropriate mid- and high-elevation montane habitats did not exist on islands older than Kauai by the time Kauai was high enough to support these habitats. Geiger et al. (2007; and references therein) identified four potential weather- or climate-based pathways for spore dispersal that could explain the biogeographic origins of native Hawaiian fern lineages. They suggested the northern subtropical jetstream could carry spores to the Hawaiian Islands from areas of the Indo-Pacific; spores could be moved from Central and North America via the trade winds; spores could be carried by storms from southern Mesoamerica; or spores could be transported from the Southern Hemisphere (S. America or the South Pacific) northward across the equator by a combination of mechanisms involving a seasonal southern shift of the Intertropical Convergence Zone (ITCZ), movement of air via Hadley Cells, and the trade winds (see Wright et al, 2001). Reviews of studies by Geiger et al. (2007) that included Hawaiian endemic species {Dryopteris (Geiger and Ranker, 2005), Hymenophyllum (Ebihara et al., 2004; Hennequin et al., 2006), Polystichum (Driscoll and Barrington, 2007), Adenophorus, Grammitis, and Stenogrammitis (Ranker et al., 2003, 2004; Stenogrammitis treated as Lellingeria in the works cited)) found evidence for each of the four described weather/climate-based dispersal mechanisms. In the current study, using molecular phylogenetics, we evaluated the biogeographical history of the AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 (2013) Taxa examined in study Cibotiaceae Cibotiaceae JX485654 ^ PC485655 ^ AM176429^ AM176589^ Cibotium menziesii Hook. Cibotium nealiae Degen. Ranker 2070 (COLO; TAIF) Wolf 266 (Sttybing Herbarimn) Ranker 1996 (COLO; BISH) Fagerlind & Skottsberg 6494 (S) Oxford Cibotiaceae Mortei AM176431^ AM176591^ NA JX485657 ® AM176432^ AM176592^ 2485 AM176433^ AM176593 ^ Turner & R. A. White Dicksonia antarctica Labill. Lophosoiia 0. F. Gmel.) C. Chr. (Type sp.) A dash ( — ) indicates th XXX = new to this stuc ^ from Korall et al., 200 ^ from Korall et al., 200 ^ sequenced by JMOG sequenced by PK Bold - from other studii (UTC) Grantham 006-92 (uq 2254 AM176428' AM176588^ 134 AM176442^ U93829 424 AM176450^ AM176609^ Li and Lu (2006). PCR products were purified with ExoSAP-IT (USD Corp.) and sequenced by Macrogen Inc. in South Korea. Sequences produced by Korall were amplified and sequenced using the primers and methods described in Korall et al. (2006, atpA, atpB, rbcL, rps4, and rps4-tmS IGS GEIGER ET AL.: PHYLOGENETICS OF CIBOTIUM Table 1. Extended. rbcLr-accD rbcL- atpB JX485690 ^ AMI 77328^ JX485692 ^ AM177330" JX485693 ® AM177331^ AMI 76485' JX485694 AMI 76486' JX485695 “ AMI 76487' PC485696 “ JX485659 JX485660 “ JX485674 ^ JX485682 '' JX485663'' JX485677 “ JX485685 ^ JX485666“ AM176488' JX485697 JX485661 ' NA NA JX485678 * JX485686 ' JX485679 '' JX485687 JX485667^ JX485668^ JX485680 ^ JX485688 JX485669' AMI 76489' PC485698 ^ JX485662 PC485681 * JX485689 ^ AM177327' AM176484' AM410497^ AM410290" AM410426=‘ AM410354^ U05919 AF101303 AF313596 AM410498^ AM176502' AM410505'' AM410291^ AM410298^ AM410427^ AM410355^ AM410434^ AM410361^ Pt485670'‘ JX485672'' JX485671'' JX485673^ (amplified and sequenced within rps4)) and Korall et al. (2007, rbcL-accD, rbcL-atpB, trnG-R, and trnL-F). Sequence editing and alignment —Sequence fragments were assembled and edited using Sequencher version 4.2.2 (Gene Codes, Ann Arbor, Michigan, USA). The edited consensus sequences were aligned manually using MacClade version 4.07bl3 (Maddison and Maddison, 2005). Ambiguously aligned regions (due to insertions or deletions) in the noncoding regions {rbcL-accD, rbcL-atpB, tmG-R, tmh-F, and rps4-tmS IGS) were excluded from the analyses. No “gap coding” was performed. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 (2013) I in Bayesian (B/ atpA atpB HKY+I HKY HKY+G GTR+I HKY+G HKY HKY+G HKY+G HKY+I ExcL^ded Analyzed Phylogenetic analyses . — We tested for incongruence among the nine single- region data sets by analyzing the data sets separately using both a Bayesian Markov chain Monte Carlo approach (B/MCMC) and equally weighted Maximum Parsimony (MP) with settings as described below for the combined data set. We compared the resultant topologies and incongruence supported by a Bayesian posterior probability of > 70% was considered a conflict. For each analytical method we compared all nine resulting topologies and no conflicts were found in the ingroup. The nine single-region data sets were combined into a single data set and analyzed using a Bayesian Markov chain Monte Carlo approach (B/MCMCj and equally weighted Maximum Parsimony (MP). B/ MCMC analyses were performed with MrBayes 3.1.1 (Huelsenbeck and Ronquist, 2001; Ronquist and Huelsenbeck, 2003), using a single partition for each region (i.e., with nine partitions). The most appropriate nucleotide substitution models for each of the regions were determined using the Perl script MrAIC version 1.4.4 (Ny lander, 2004) in combination with PHYML version 2.4.4 (Guindon and Gascuel, 2003). The choice of model was based on the corrected Akaike information criterion (AICc) (see Table 2 for a summary of models used). The analysis was run for five million generations, on four parallel chains, with the temperature parameter set to 0.2. Four independent analyses of each region were run simultaneously. To determine whether parameters had converged, the values sampled for different parameters were examined using the program Tracer v. 1.2.1 (Rambaut and Drummond, 2005). We also examined the standard deviation of the split frequencies among the independent runs as calculated by MrBayes. For each analysis, every 1000‘^ tree was sampled, and burn-in was conservatively set to one million generations. A majority-rule consensus tree was calculated based on the pool of trees resulting from the four independent analyses (except those trees discarded as bum-in). GEIGER ET AL.: PHYLOGENETICS OF CIBOTIUM 147 The MP analyses were performed with PAUP* version 4.0bl0 (Swofford, 2002) and included a heuristic search for the most parsimonious trees with 1000 random-sequence-addition replicates with MulTrees activated and tree- bisection-reconnection (TBR) branch swapping. Support for nodes was calculated by bootstrap analysis with 5000 bootstrap replicates with 10 random-sequence-addition replicates each. All trees were rooted with the three outgroup species. Results The number of characters included in the analyses and the nucleotide substitution model used in B/MCMC analyses are summarized in Table 2. For the combined MP analysis 407 (4.3%) characters were parsimony informative. The heuristic MP search resulted in three most parsimonious trees, each with a length of 977 steps in one island. Our results indicated that, based on plastid data alone, Cibotium was strongly supported as monophyletic and most relationships among the Cibotium species were strongly supported (posterior probability, PP = 1.00) according to the B/MCMC analysis, and they were moderately (bootstrap percentage, BP = 70-89%) to strongly supported by the MP analysis (BP > 90%) (Fig. 1). Few indels were found, and these were either autapomorphies or synapomorphies for clades already well supported. Most indels supported the monophyly of the ingroup. Only two indels were synapomorphies of subclades within the ingroup: a five-base indel joining Cibotium regale and C. schiedei and a seven-base indel shared by C. chamissoi and C. menziesii. Within Cibotium, there were three strongly supported clades that corre- spond to the geographic distributions of the included species: the Asian clade, with three species, was sister to a clade with the four Hawaiian species in one subclade, and the two Mesoamerican species in the other. Within the Asian group, C. arachnoideum was well supported by B/MCMC analysis and moderately supported by MP analysis as sister to the pair C. barometz and C. cumingii. The sister relationship between the Hawaiian species and those from Mesoamerica was strongly supported by both analyses. Within the Hawaiian clade, C. chamissoi and C. menziesii were strongly and moderately supported as sister species by B/MCMC and MP analyses, respectively. However, the relationships among those two taxa and C. nealiae and C. glaucum were unresolved. Discussion The goals of this study were to determine if the Hawaiian species of Cibotium are monophyletic and to determine the relationships among species of Cibotium to infer the biogeographical history of the Hawaiian species. Using five coding and four non-coding plastid DNA loci, this study has resulted in a relatively well-resolved molecular phylogeny for Cibotium. Within Cibotium, three clades were resolved, each of which corresponds to the geographical Species of Cyrtogon and Development of Three n Ching (Dryopteridaceae) 154 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER : Vietnam (Liu et al, 2010). The genus was found to be nested within the polystichoid fern clade as putative sister to Polystichum sect. Sphaenopolys- tichum (Liu et al, 2010). The phylogenetic results were consistent with established arguments concerning the close relationships to Polystichum (Copeland, 1947; Ching, 1978; Tryon and Tryon, 1982; Kramer et al, 1990; Xie, 1990; Tryon and Lugardon, 1991). Limited understanding of the phylogenetic relationships of polystichoid ferns especially Cyitogonellum motivated several studies targeting the discovery of additional morphological, cytological, and AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 prothallus cell. The initial prothallus cell contained many plastids, whereas chloroplasts were rare or absent in the rhizoids (Fig. 2D). Filamentous and laminar phase— About nine days after sowing, a uniseriate filament was formed. The first prothallus cell was divided by a transverse wall creating an apical cell. Transverse divisions of the apical cell resulted into 2-6 cells long uniseriate filaments (Fig. 2C-D). The cells were pyriform and had a length to width ratio ranging from 1:1 to 3:1. They contained abundant chloroplasts. Occasionally there were two rhizoids produced per filament. Both were attached to the first prothallus cell. The differentiation of the laminar phase was observed at about 17 days after sowing. The terminal cell of the filament divided longitudinally several times and formed a slightly lopsided spatulate prothallus. This prothallus was 6-16 cells wide. Prothallus development followed the Aspidium-type. Unicellular glandular hairs were formed along the margin while the meristem developed a notch and two wings. These young prothalli were elongate-spathulate in all three species (Fig. 2E-H, 3A-C, 4A-B). Prothallus. — Prothalli emerged ca. 60 days after sowing. Mature prothalli of the three species differed in shape (Fig. 2E, 2K, 3D, 4C). Prothalli of Cyrtogonellum fraxinellum had elongate-spatulate shape with irregularly branched wings, a dense cover of rhizoids, and usually one (Fig. 2E) or sometimes several meristems. The prothalli of C. caducum were spatulate with fewer rhizoids, a slightly irregular margin, and a single meristem (Fig. 3D), while those of C. inaequale were symmetrically cordiform with two fully developed wings, a dense cover of rhizoids, and a single well-developed meristem (Fig. 4C). Gametophyte of all three species had papillate-glandular hairs along the margins and on their lower surface (Fig. 2E-G, 2L, 2N-P, 3A-C, 3H-J, 4A- B, 4F, 4H-I, 4K). The globose apices of the papilli were 29-30 pm in diameter. Sometimes the tip was forked resulting in two apical papillae (Fig. 4K). Chloroplasts were located mainly at the center and basal region of the trichomes, which degenerated as the trichomes matured. Uni-seriate to multi-seriate, lobed extensions were found on the margin of prothalli of Cyrtogonellum fraxinellum (Fig. 20-P), C. caducum (Fig. 3J-K), and C. inaequale (Fig. 4H-I). These extensions terminated with a glandular hair that was unforked, filamentous and uniseriate or forked with two or more basal cells. The three species differed in the form of the extensions. Filamentous outgrowths were usually found in C. caducum (Fig. 3K), whereas lobed, wing-like outgrowths were found mainly in C. fraxinellum (Fig. 2H). Rhizoids were found on the prothalli of all three species. The primary rhizoid arose from germination of the spore while secondary rhizoids were produced by cells of the filament and the prothallus. Rhizoid formation was usually restricted to the ventral surface and adjacent regions of the adult prothallus. Occasionally swollen and forked rhizoids were detected in Cyrtogonellum fraxinellum (Fig. 2Q-S) and C. caducum (3L-M). Reproductive organs and young sporophytes.— Prothalli of all three species produced antheridia (Fig. 21, 3E. 4D-E) but no archegonia were observed for annular or ring cell, and an opercular cell. Spermatozoids were liberated by detachment of the operculum. After release, the spermatozoids were observed actively swimming. Mature prothalli of Cyrtogonellum fraxinellum produced both antheridia and archegonia at the same prothallium (Fig. 2J). Up to a maximum of five archegonia were located near the midrib on the cordate-thalloid prothallium. The terminal neck of the archegonia became brown with age. Mature prothalli of all three species produced sporophytes. These embryos were formed by outgrowth of somatic cells without the involvement of archegonia (Fig. 3F, 162 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 (2013) which rules out effects of differences in cultivation conditions. The prothalli of C. inaequale appear to he close to the regular type of prothalli reported in Dryopteridaceae, whereas those of C. fraxinellum differ by the occurrence of irregularly shaped wings. The prothalli of C. caducum deviate from symmetrical prothalli found in the majority of Dryopteridaceae but similar gametoph 3 de shapes are found in young prothalli and/or male gametophytes in response to antheridogenes (see Schneller, 2008). It is possible that culture conditions may have influenced the observed gametophyte shapes (Nayar and Kaur, 1971; Zhang and Shi, 2005; Deng et ah, 2009). The morphology of the sex organs of the studied Cyrtogonellum taxa is clearly of types known to occur in derived leptosporangiates (e.g., Schneider et ah, 2009). Only antheridia are formed in C. caducum and C. inaequale, whereas C. fraxinellum possess both antheridia and archegonia. All three species are able to form embryos without the involvement of archegonia. These observations support the notion that some species of Cyrtogonellum reproduce via apomixis (= apogamy) and not sexually. Apomixis has been suggested as the common reproductive mode of this genus based on cytological evidence, spore number (Lu et ah, 2006) and direct observations of gametophyte growth and development (Bao et al, 1999; Liu et al, 2012). Only a few prothalli of C. fraxinellum are observed with archegonia but our observations do not provide evidence for functional archegonia. Future work focusing on the detection of functional archegonia will be conducted because non-functional archegonia have also been reported for apomictic ferns such as Pteris cretica (Laird and Sheffield, 1986). In our study, all embryos are found to originate from somatic cells of the prothalli. The loss of functional archegonia in sexually reproducing ferns is considered as the putative trigger for the establishment of obligate apomictic ferns (Moggie, 1990). All three studied species are reported to be triploids (2n = 3x = 123) with 32 spores per sporangium as shown by Lu et al. (2006) and verified by our spore counts. Bao et al. (1999) has confirmed apomictic reproduction for Cyrtogo- nellum inaequale, whereas C. caducum and C. fraxinellum are inferred previously as apomicts based on their chromosome number and the number of spores per sporangium. In general, 32 spores per sporanginm have been considered as an indicator for apomictic reproduction in derived ferns. However, it has been argued that observations on the occurrence of sexual organs and the development of young sporophytes are required to confirm the hypothesis of apomictic reproduction (Huang et al., 2011; Liu et al., 2012). In C. caducum and C. inaequale, the lack of archegonia may be considered as evidence confirming apomictic reproduction. This is consistent with the fact that all observed sporoph)des are formed from somatic cells. Many apomictic ferns possess antheridia but archegonia are usually absent (Manton, 1950; Walker, 1966; Liu et al., 2012). However, non-functional archegonia have been reported for some apomictic fern prothalli (Laird and Sheffield, 1986; Chiou et al, 2006). Thus, the presence of archegonia is insufficient evidence to claim sexual reproduction in Cyrtogonellum frax- inellum. However, we need to consider the possibility of karyological variation American Fern Journal ):166-174 Taxonomic Reassessment and Conservation Status of Three Kaua’i Species of Asplenium in the Diellia Alliance David H. Lorence^ and Kenneth R. Wood N ational Tropical Botanical Garden, 3530 Papalina Road, Kalaheo, HI 96741, USA Ruth Aguraiuja Tallinn Botanic Garden, Kloostrimetsa Rd. 52, Tallinn 11913, Estonia of three Kaua’i species of Asplenium in the Dielli I observations and herbarium specimens. Their t iscussed, and a lectotype is selected for Undsaya i s.— Aspleniaceae, lectotypification, morphology, Kaua’i, Hawai’i, The Hawaiian endemic fern genus Diellia Brack, was previously considered to include six species (Palmer 2003; Wagner 1952, 1993). The origin of this group of six species is thought to date to ca. 2 Myr ago and coincide with the renewal of Hawaiian terrestrial life in the Miocene (Schneider et al. 2005al. However, recent molecular phylogenetic studies have shown Diellia to be deeply nested within Asplenium L. (Schneider et al. 2005b), and all names in Diellia have been transferred to or given new names in Asplenium (Viane & Reichstein 1991; Snow 2011; Snow et al, 2011). However, the new names or combinations were made by these authors essentially to provide valid names for these taxa in Asplenium based on Palmer’s and Wagner’s species concepts, without any further taxonomic evaluation. Consequently, several of the species complexes in this group have not yet been satisfactorily resolved taxonomically, including one with three species on Kaua’i. Recent intensive field work on Kaua’i by the second and third authors has revealed the presence of new populations of taxa in this rare Asplenium alliance. Observations and specimens deposited in the National Tropical Botanical Garden Herbarium (PTBG) have helped reassess the considerable morphological variation in this complex. We here discuss the taxonomy, nomenclature, and synonymy of the three Kaua’i species in this alliance and select a lectotype for Undsaya knudsenii. As currently circumscribed, three species in this alliance occur on Kaua’i, separable by characters in the key below. They are: Asplenium dielmannii ce@ntbg.org LORENCE ET AL.: ASPLENIUM IN DIELUA ALLIANCE Viane, A. dielpallidum Viane, and A. diellaciniatum Viane. The first two species are relatively uniform morphologically and easily recognized. However, population studies of the third species reveal an amazing range of variation in leaf morphology and degree of lohing. Hillebrand (1888) named four different species and six varieties (not all were validly published) in the genus Lindsaya Kaulf. [= Lindsaea Sm.l in order to encompass this astonishing array of variation: Lindsaya alexandri (Hillebr.) Hillebr. including one variety from Kaua’i, var. bipinnata Hillebr. (fronds bipinnate); Lindsaea centifolia Hillebr. from Kaua’i ; Lindsaea knudsenii Hillebr. from Kaua’i; and Lindsaea laciniata Hillebr. from Kaua’i, with var. subpinnata (fronds subpinnate). Except for the two varieties of Lindsaya alexandri from Maui, all the specimens Hillebrand used to typify these names were collected from essentially the same vicinity on Kaua’i, called “Halemanu” by island resident Valdemar Knudsen who sent plant specimens to Hillebrand. The specimens were subsequently deposited in the Berlin herbarium and images are now available online (Ropert 2000-present, http://www.bgbm.org/). Halemanu was the site of the Knudsen family’s mountain home in western Kaua’i. Searches in the vicinity of the old Knudsen homestead have failed to reveal any extant populations of Asplenium diellaciniatum around Halemanu (K. Wood, pers. obs.) However, two new populations numbering ca. 90 plants were recently discovered at Kawai’iki, located about 7.25 km from Halemanu in a similar habitat type. Acacia koa A. Gray mesic forest. Study of these two subpopulations at Kawai’iki and additional herbarium collections reveals that these morphotypes are neither consistent nor stable and intergrade with each other. We conclude that a single species, Asplenium diellaciniatum, is represented with extremely variable frond morphology and dissection, depending on age and stage of development of the plant and possibly also microhabitat (K. Wood and R. Aguraiuja, pers. obs.). A single plant may even possess fronds representing several morphotypes corresponding to taxa described by Hillebrand (e.g.. Wood S' Query 1174, 1175, 1177, all PTBG; Wood S' Perlman 9015, 9060, 9062, 9062B, 9062C, all PTBG), (Figures 1, 2). A similar situation exists in A. daucifolium Lam., a species of Madagascar and the Mascarene Islands which displays a series intergrading frond forms ranging from 1- to 4-pinnate (Tardieu-Blot 2008). In his seminal study of the morphology and taxonomy of the genus Diellia, Wagner (1952) recognized this variation for Kaua’i plants but ascribed it to Diellia erecta Brack., as forma alexandri (Hillebr.) W. H. Wagner. He recognized two additional species from this region of Kaua’i, D. mannii (D.C. Eaton) W. J. Rob. and D. laciniata (Hillebr.) Diels, and later described a third species, D. pallida W. H. Wagner (1993). Wagner’s species concepts were generally followed by Palmer (2003) in his treatment of Hawaiian ferns. Palmer recognized Diellia pallida and D. mannii [= Asplenium dielmannii Viane] as distinct species, although he synonymized Hillebrand’s Lindsaya knudsenii |3 var. under the latter. He also subsumed the names Diellia centifolia and AMERICAN FERN JOURNAL: VOLUME 103 NUMBER : m LORENCE ET AL.: ASPLENIUM IN DIELUA ALLIANCE 173 they may hybridize, A. leucostegioides Baker from East Maui (presumed extinct), and Asplenium dielerectum Viane (all the major islands except Kaua’i, Kaho’olawe and Ni’ihau). Further study is needed to clarify the status of the forms of the Maui Nui plants within Asplenium dielerectum, including Diellia erecta f. alexandri for which the nomenclature and typification are summarized below. Diellia erecta Brack, f. alexandri (Hillebr.) W. H. Wagner, Univ. Calif. Publ. Bot. 26: 155. 1952. Basionym: Diellia alexandri (Hillebr.) Diels, Nat. Pflanzenfam. [Engler & Prantl} 1(4]: 212 f. 114 G-K. 1899. Undsaya alexandri Hillebr., FI. Hawaiian Isl. 622. 1888. Lectotype, designated by Wagner (1952): USA: Hawaiian Islands. Maui: Northern slope of Haleakala, 3000- 4000 ft, Alexander S' Lydgate (Lectotype B, image seen). For their assistance in the field and valuable logistical support we thank Michelle Clark, Kaua’i Partnerships Biologist, Pacific Islands Fish and Wildlife Office; Wendy Kishida, Kaua’i Coordinator for the Plant Extinction Prevention Program; and Katie Cassel, Koke’e Resource Conservation Program. We are grateful to the State of Hawaii, Division of Forestry and Wildlife (DOF AW), Kaua’i branch, for permission to visit sites and collect specimens. Dominik Ropert (Herbarium Berolinense) kindly provided a digital image of the holotype specimen of Undsaya laciniata. We thank reviewers of the manuscript for providing valuable comments. Literature Cited Aguraiuia, R. and K. R. Wood. 2002. The Critically Endangered Endemic Fern Genus Diellia Brack. (Aspleniaceae) from Hawai’i: Its population structure and distribution. Special paper presented at the Pteridophyte S 3 miposium, Fern Flora Worldwide; Threats and Responses. Fern Gaz. 16:330-334. Aguraiuia, R. and K. R. Wood. 2003. Diellia mannii (D. C. Eaton) Robins. (Aspleniaceae) rediscovered in Hawaii. Am. Fern J. 93:154-156. Hillebrand, W. 1888. Flora of the Hawaiian Islands. London. McNeill, J. C., F. R. Barrie, W. R. Buck, V. Demoulin, W. Greuter, D. L. Hawksworth, P. S. Herendeen, S. Knapp, K. Marhold, J. Prado, W. F. Prud’homme van Reine, G. F. Smith, J. H. Wiersema and N. J. Turland. 2012. International Code of Botanical Nomenclature for algae, fungi, and plants (Melbourne Code). 2012. Regnum Vegetabile Volume 154. Koeltz, Germany. Lidgate, j. M. 1873. Short Synopsis of Hawaiian Ferns. Honolulu. [Note: John M. Lydgate (1854- 1922) collected in Hawai’i. His surname is customarily spelled “Lydgate”.] Palmer, D. D. 2003. Hawai’i’s Ferns and Fern Allies. University of Hawai’i Press, Honolulu. 324 p. Plant Extinction Prevention Program of Hawaii. 2011. Hawaii State PEP list, http://pepphi.org [accessed June, 2012). ROpert, D. (ed.) 2000- (continuously updated): Digital specimen images at the Herbarium Berolinense. Published on the Internet http://ww2.bgbm.oig/herbarium/ (Barcode: B 20 0046501 / Imageld: 268001) [accessed 29-May-13j. Schneider, H., T. A. Ranker, S. J. Russell, R. Cranfill, J. M. O. Geiger. R. Aguraiuia, K. R. Wood, M. Grandmann, K. Kloberdanz and J. C. Vogel. 2005a. Origin of the endemic fern genus Diellia coincides with the renewal of Hawaiian terrestrial life in the Miocene. Proc. Roy. Soc. London, ser. B, Biol. Sci. 272:455^60. A New Hybrid of Serpocaulon (Polypodiaceae) 176 AMERICAN FERN JOURNAL; VOLUME 103 NUMBER I Smith et al. (2006) divided the genus into four informal groups: 1) S. loriceum, 2) S. fraxini folium, 3) S. subandinum and 4) S. lasiopus. The new hybrid described here, between S. fraxinifolium (group 2) and S. ptilorhizon (group 3), provides evidence that the limits between these groups are weak. Serpocaulon X sessilipinnum A. Rojas & J.M. Chaves, hyb. nov. TYPE. — Costa Rica. Puntarenas: Coto Brus, San Vito, Cerro Paraguas, orillas de la laguna Los Gamboa, 8°47'16.3"N, 82°59'20"W, 1447 m, 3 Nov 2012, J.M. Chaves et al. 299 (Holotype: CR; Isotypes: HLDG, MO) (Figs. IB and E, 2B and E). Diagnosis: The new hybrid differs from S. fraxinifolium (Jacq.) A.R. Sm. by having rhizome scales that are relatively smaller, moderately dense, with dark brown central portions and narrower pale margins; blades that are 1-pinnate basally to pinnatisect distally and relative narrower; pinnae that are sessile and debate to deltate-lanceolate; fewer series of areoles and sori, spores that are whitish, ellipsoidal, most well developed and a few collapsed. Similar characters differentiate the new hybrid form S. ptilorhizon (Christ) A.R. Sm. Additionally, S. ptilorhizon has rhizome scales that are smaller and less dense than the hybrid, and scales have a blackish central portion and a narrower pale margin; blades that are relatively narrower and pinnatisect; pinnae that are lanceolate to narrowly debate; fewer series of areoles and sori; and spores that are yellow and regularly ellipsoidal. Description. — Rhizome long creeping, 2.5-4 mm in diameter, non-pruinose, moderately scaly; rhizome scales 1-2 X 0.9-1. 5 mm, orbicular to ovate, clathrate, dark brown centrally, with yellowish to light brown margin 0.1- 0.3 mm, appressed, marginally entire to irregularly-lobulate, apically obtuse to rounded; fronds 44-53.5 cm long, separated by 0.5-8. 5 cm; stipe 21.6-25.2 X 0.1-0. 2 cm, ribbed, stramineous to light brown, glabrous except for trichomes 0.2-0.3 mm, sparse, brown; blade 22.5-28.2 X 13.6-21 cm, narrowly debate, 1- pinnate to pinnatisect distally, basally truncate, apically subconform; pinnae 1.9-10.3 X 0.5-1. 7 cm, 8-13 pairs, linear-lanceolate, straight to falcate, basal pinnae slightly deflexed, marginally entire; apical pinnae 3.8-11 X 0.8-1. 3 cm, similar to lateral pinnae, with 1-2 basal lobules; rachis and costae stramineous to light brown, sparsely scaly, scales similar to rhizome scales; laminar tissue glabrous; veins reticulate, forming 2-3 series of areoles between costa and margin; sori round, located in (1-) 2 lines between costa and margin; sporangia glabrous; spores 47-51 X 28-38 pm, bilateral, ellipsoidal, convex to slightly concave-convex, exospore prominently verrucate, vermcae 3-6.2 pm, translu- cent, some with yellow patches, others irregular. Additional Specimen Examined. — Costa Rica. Guanacaste: Liberia, Parque Nacional Guanacaste, sector Santa Maria, sendero que va al volcdn Santa Maria, zona de campamento, 10°48'03"N, 85°19'02"W, 1600-1700 m, 3 Sep. 2012, A. Rojas et al. 10249 (CR, MO, USJ). Etymology. — The specific epithet refers to the sessile pinnae. Serpocaulon X sessilipinnum has characters that are intermediate between S. fraxinifolium (Jacq.) A.R. Sm. and S. ptilorhizon (Christ) A.R. Sm. including ROJAS-ALVARADO & CHAVES-FALLAS: SERPOCAULON X SESSILIPINNUM HYB. NOV. AMERICAN FERN JOURNAL: VOLUME 103 NUMBER : and D. Serpocaulon fraxinifolium [/.M. Chaves et al. 297, CR); A. Rhizc scale. B and E. S. X sessilipinnum [J.M. Chaves et al. 299, CR); B. Rhiz( scale. C and F. S. ptilorhizon (J.M. Chaves et al. 298, CR); C. Rhizo ROJAS-ALVARADO & CHAVES-F ALLAS: SERPOCAULON X SESSILIPINNUM HYB. NOV. 179 Table 1. Morphological comparison among Serpocaulon fraxinifolium, Serpocaulon X sessilipinnum, and Serpocaulon ptilorhizon. Width of the Blade division S. fraxinifolium (1.5-) 2-3 X (1-) 1.5-2 S. X sessilipinnum 1-2 X 0.9-1.5 1-pinnate basally to pinnatisect distally debate to deltate-lanceolate Pinnae shape Pinnae width (cm) axis of spore (pm) Mean of equatorial axis of spore (pm) Spore shape 50 34.5 0.5-1.1 72.3 size, color of central portion, width of pale margin, density of rhizome scales, shape and division of blade, shape and width of pinnae, number of areolae and sori series, and shape and color of spores (Table 1). The new hybrid differs from S. fraxinifolium by having smaller (1-2 X 0.9-1. 5 mm vs. (1.5-) 2-3 X (1-) 1.5-2 mm) rhizome scales with dark brown (vs. light brown) central portions and narrower (0.1-0.3 mm vs. 0.2-0.5 (-0.7) mm) pale margins with moderate (vs. high) density, 1-pinnate basally to pinnatisect distally (vs. 1-pinnate throughout) blade divisions, relatively narrower (0.5-1. 7 cm vs. (1-) 1.8-3. 3 cm), debate to deltate-lanceolate (vs. ovate to lanceolate) and sessile (vs. free) pinnae, fewer series (2-3 vs. 4-5) of areoles and fewer series ((!-) 2 vs. 3-5) of sori, and whitish (vs. yellow) and ellipsoidal to collapsed (vs. ellipsoidal) spores (Table 1). From Serpocaulon ptilorhizon, S. X sessilipinnum differs by its longer (1-2 X 0.9-1. 5 mm vs. 0.5-1 mm in diameter) rhizome scales with dark brown (vs. blackish) central portion and broader (0.1-0.3 mm vs. :£0.1 mm) pale margin, moderate (vs. sparse) scale density, 1-pinnate basally to pinnatisect distally . j. Bot. ; Shorter Notes SHORTER NOTES Shorter Notes AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 (2013) Fig. 1. Amphorophora ampullata congregating and sucking the sap from Hypolepis polypodioides. during last week of September and October (14.5 and 17.3 aphids/pinna respectively) and peaked during the last week of November of 2012 (24.8 aphids/pinna) which may due to high temperature (28-30 °C) and humidity (65—75%), later the infestation was decreased in the month of December (5.9 aphids/pinna) and January 2013 (0.10 aphids/pinna) due to low temperature (4-10 °C). The fern aphid, A. ampullata is greenish to brown, bigger in size as compared to other aphids. The alate male body is about 2.71 mm length, 0.99 mm width and antennae (n=l) are longer than the body (3.42 mm) (Raychaudhuri et al, Ins. Matsum. n.s., 20:1-42. 1980). In the present study, the alate male body is about 2.91 mm length, 1.03 mm width and antennae are longer than the body (3.45 mm); whereas, the alate female body is about 4.15 mm length, 2.04 mm width and antennae are 5.38 mm long (n=10). Both nymphs and adults stages of aphid are responsible for causing damage to H. polypodioides and were found congregating the under (lower) surface of fronds (Fig. 1). Apparently the aphids injure the plant by piercing and sucking the sap, which results in yellowing, loss of vigour, drying, and dropping of fronds. In severe infestations, entire plants start drying and wilting. In India, A. ampullata is distributed in Arunachal Pradesh, Himachal Pradesh, Meghalaya, Sikkim and West Bengal on different host plants other than H. polypodioides (Raychaudhuri et al., Ins. Matsum. n.s., 20:1-42. 1980), of A. Notes Review AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 3 (2013) book from many general ones. All delineations are accompanied by the scientific binomial, the Chinese name, and the pinyin transliteration. In the key, the stressed syllables in the Latin names are underlined, which is reminiscent of Werner Rothmaler’s Exkursionsflora von Deutschland and is much appreciated. Every taxon is illustrated with at least one photograph, thus facilitating the accurate identification of every currently known species, subspecies, and hybrid in Taiwan. Even very rarely and poorly documented species are discussed with several photos, some of which are a full page in size, bringing the total number of color images to over 4,700, some of which are of types of species or infraspecific taxa making this book particularly valuable for not only fern lovers but also professional taxonomists! A thorough index includes numerous synonyms, misapplied names, and alternative taxonomic combinations. As with every book published, Knapp’s (2011) book could have been out of date right after it was submitted to the press. Readers, however, will be glad to see that a supplemental volume to his 2011 book - Ferns and Fern Allies of Taiwan - Supplement (Taipei: KBCC Press. 212 pp. Paperback. Price: $35.00. ISBN 978-986-87098-1-2) was published in March 2013 with 560 new color images. The Supplement contains some additions and corrections, bringing the total number of species of ferns and lycophytes in Taiwan to 747. Many field images used in the book are available for free at https://picasaweb. google. com/116136418529949606360. In China’s current 32 provinces, Taiwan has 4,383 species of vascular plants and is the seventh most species-rich province for vascular plants (the first six are Yunnan, Sichuan, Guangxi, Xizang, Guizhou, and Guangdong; Wu et al., 1994-2013). When only species of ferns and lycophytes are counted, Taiwan is the fourth richest province in Ghina (the first three are Yunnan, Sichuan, and Guizhou; Lin et al., 2013). Interestingly, in terms of species numbers of ferns and lycophytes per square mile, Taiwan has not only the highest fern diversity of any province in China but also is one of the most fem-diverse regions in the world, in comparison with, e.g., 509 species and infraspecific taxa of ferns and lycoph 3 hes in 70 genera in the whole of North America, north of Mexico [Flora of North America Editorial Committee, 1993). With Knapp’s (2011, 2013) publications this rich flora of ferns and lycophytes in Taiwan is now well documented and is definitely one more big step closer to the final version. With the new version of Flora of China vol. 2-3 (pteridophytes) published (Lin et al., 2013) and more and more molecular data accumulated, further revision of the fern and lycophyte flora of Taiwan is expected, with new familial and generic delimitations incorporated. We look forward to seeing updates of Knapp’s work on the flora of ferns and lycophytes of Taiwan in the near future.— Li-Bing Zhang, Missouri Botanical Garden, P.O. Box 299, St. Louis, Missouri 63166-0299, U.S.A. and Chengdu Institute of Biology, Chinese Academy of Sciences, P.O. Box 416, Chengdu, Sichuan 610041, P. R. China; Hai He, College of Life Sciences, Chongqing Normal University, Shapingba, Chongqing 400047, P. R. China. fii Ititif Itliifl ili INFORMATION FOR A PTERIDOLOGIA ISSUES IN PRINT The following issues of Pteridologia, the memoir series of the American Fern Society, are available for purchase: 1. Wagner, David H. 1979. Systematics of Polystichum in Western North America North of Mexico. 64 pp. $10.00 plus postage and handling. 2 A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choco (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 plus postage and 3. Lellinger, David B. 2002. A Modem Multilingual Glossary for Taxonomic Pteri- dology. 263 pp. $28.(X) plus postage and handling. 4. Hirai, Regina Y., and Jefferson Prado. 2012. Monograph of Moranopteris (Polypo- diaceae). 1 13 pp. $28.00 plus postage and handling. For orders and more information, please contact our authorized agent for sales at: Missouri Botanical Garden Press, P.O. Box 299, St. Louis, MO 63166-0299, tel. 314-577- 9534 or 877-271-1930 (toll free). For online orders, visit: http://www.mbgpress.org. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns. SPORE EXCHANGE Mr. Brian S. Aikin, 3523 Federal Ave, Everett, WA 98201-4647 (spores.afs@comcast. net), is Director. Spores exchanged and lists of available spores sent on request, http:// amerfemsoc.org/sporexy.html GIFTS AND BEQUESTS Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Back issues of the Jourual and cash or other gifts are always welcomed and are tax.deductihle. Inquiries should be addressed to the Membership Secretary. VISIT THE AMERICAN FERN SOCIETY’S WORLD WroE WEB HOMEPAGE: http://amerfemsoc.org/ AMERICAN FERN JOURNAL Volume 103 Number 4 October-December 2013 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY and Lycophyte Richness at Three Taxonomic Levels Marc Bogonovich, Scott Robeson, and Maxine Watson 193 Chemical Composition of Essential Oils from Two Fern Species of Anemia Marcelo Guerra Santos. Caio Pinho Fernandes. Luis Armando Candido Tietbohl. Rafael Garrett, Jonathas Felipe Revoredo Lobo, Alphonse I^Iecom, and Leandro Rocha 215 Helical CeU Wall Thickenings in Root Cortical CeUs of Polypodiaceae Species from North- western Argentina Marcela A. Herndndez, Lucrecia Thrdn, Marisol Mata, Olga G. Martinez, Shortek Notes Hydrochemical Characterization of A Stand of the Threatened Endemic /so«os malitiver’ niana T Abeli. S. Orsenigo, N, M. G, Ardenghl, F.C.HM.T Lucassen, and AJ,P. Smolders 241 Flora of China, vole. 2-3, Arthur V Gilman and Miahael A. Sundue 245 Errata Referees for 2013 Table of Contents for Volume 103 m 255 The American Fern Society Councii for 2013 Fern and Lycophyte MISSOURI BOTANICAL APR 2 3 2014 garden library 194 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) which we refer to as the productivity-diversity hypothesis. In this study we examine the relationship between climate and monilophyte (fern) richness at three taxonomic levels, in order to test the productivity-diversity hypothesis. We use the Smith et al, (2006) fern taxonomic classification system. First, we test a number of regression models that employ linear combinations of simple climate variables to determine whether these variables predict fern richness. We also test whether one particular richness-climate model, the interim general model I (IGM I), accurately predicts fern richness. We employ the IGM I because it successfully predicts richness gradients in other plant groups and it has a plausible theoretical basis (O’Brien, 2006). A theoretical basis for linking the generation of plant diversity with climate factors is provided in water-energy dynamics theory (WED; O’Brien, 1998; O’Brien, 2006), which can be included among productivity-diversity hypoth- eses. WED states that plant biological activity is related to liquid water and energy availability, which foster higher productivity, and finally lead to greater biological richness. Climate factors associated with greater plant productivity also lead to greater rates of molecular evolution (see Rohde, 1992; Wright et al., 2003; Wright et al., 2006) or greater amounts of biological activity in general (generation time, faster growth, more individuals). Higher rates of molecular evolution lead to more population divergence (Martin and Mckay, 2004), which eventually leads to greater numbers of species. The WED hypothesis is related to the evolutionary rates hypothesis (ERH) which predicts that diversity will be associated with conditions that cause higher rates of molecular evolution (Evans and Gaston, 2005). For plants, regions with higher (or optimal) temperatures and more liquid water are likely to be associated with higher rates of molecular evolution and higher rates of productivity. The interim general model (IGM I) is a climate model designed to predict plant richness based on the principles of WED theory (Field et al., 2005) taking the following form: Richness = p^RAN + P^PETmin - p^PETmin^ + p^ Where PETmin is the minimum monthly potential evapotranspiration and RAN is annual rainfall. Before each term are empirically fitted coefficients. Rainfall is included to reflect that plants require the liquid form of water. PET (potential evapotranspiration) is an index of energy; its relationship with richness is expected to be nonlinear. WED predicts greatest richness where water and energy climate conditions are most favorable for plant productivity and biological activity, while the IGM provides a hypothesis for, and quantification of, what these conditions would be. For many organisms, species richness correlates with water and energy variables in a similar way as taxonomic richness at higher levels. We know of no reason a priori why ferns would be unusual in this regard. Family richness, additionally, may serve as a proxy of species richness patterns (Balmford et al., 1996; Francis and Currie, 2003; Gaston and Blackburn, 1995; Qian and tirHBJHjHHniliiKHHfliinn BOGONOVICH ET AL.: NORTH AMERICAN FERN . I LYCOPHYTE RICHNESS 197 be considered separate species by some taxonomists (Tryon, 1969). It is common for fern taxonomists to use subspecific designations to distinguish similar but distinct non-intergrading taxa (Hickey et al, 1989; Yatskievych and Moran, 1989). Regardless, many of these types are geographical subspecies and so the species would not be counted twice in any one location. Richness patterns are largely unchanged when hybrids are excluded and subspecies or varieties are considered as single species. Range maps.—^Ne used published non-GIS fern range maps from the Flora of North America North of Mexico Vol. 2, Pteridophytes and Gymnosperms (FNA editorial committee, 1993) to produce geo-referenced range map polygons for GIS analysis. Image files were obtained online at http://www.fna.org/. The image files were used as a template on which to produce 479 polygon shape files in the Arc View 9.1 GIS software (Esri, 2005). The non-GIS range maps were produced by the author of each taxon treatment in the Flora by hand drawing shaded regions from herbarium records (see Flora of North America volume 2). Richness maps.— All analyses were performed on a North American Lambert conformal conic projection with standard parallels at 20°and 60 N latitude and the central meridian at 96°W. A polygon shapefile layer with a grid of 50 km X 50 km (2,500 km^) was created over North America using Hawth’s Analysis Tools for ArcGIS (Beyer, 2004) and used as a base-map. All maps were generated with this grid with a total of 8,760 squares. Regression analyses were applied to a sub-sample of this grid, 88 squares, in order to account for the effect of spatial autocorrelation on the regression results (described below). The Lambert conformal conic projection has low areal distortion in our region of analysis (Bolstad, 2005) and is not expected to affect the results. Richness in each grid square was determined by summing the number of species range map polygons that overlapped each grid square. This was accomplished using Hawth’s tools polygon in polygon analysis (Beyer, 2004). The number of families and genera occurring in each grid square was tabulated similarly. _ j j Climate variables and regression models.— The climate variables tested and their sources are listed in Table 1. Worldclim variables represent climate normals for the period 1950-2000 (Hijmans et al., 2005), while actual evapotranspiration (AET) and potential evapotranspiration (PET) ^e the normals from 1931-1960 (Leemans and Cramer, 1991). Annual Rainfall (RAN) is calculated as the sum of precipitation of months with a mean temperature greater than zero degrees Celsius. Worldclim climate maps were 2.5 minute resolution. Maps of climate variables were converted to vector point maps when necessary and summarized in each 2,500 km^ grid square using a spatial join in Arcview 9.1 (Esri, 2005); the average value of the points in the climate layer in each grid square was then determined. Ordinary least-squares (OLS) regressions were performed between richness variables and climate variables with 2,500 km^ grid squares forming the sample points. Regressions also were performed on an 88 square sub-sample of the total set of 8,760 squares to assess AMERICAN FERN JOURNAL: VOLUME 103 NUMBER spatial autocorrelation issues. Regression analyses were performed with the R statistical package (R core development team, 2009). We present the univariate regressions between each richness level and each climate variable (Table 1) with the highest R-squared values (Table 2). We also perform and present several multiple regressions (Table 2). We performed two, two-variable multiple regressions with the first regression using mean annual temperature (MAT) and AET and the second using MAT and RAN. Finally, we test the interim general model I (IGM I) on our fern richness data (equation above). The models presented in table 2 correspond closely with the best performing models based on Mallow’s CP (Mallow, 1973) and a best subsets regression, performed using variables from Table 1. AIC values for each regression model are provided. We report adjusted R-squared values throughout. Biogeographical patterns are frequently explored with summary statistics without considering geographical pattern (Ruggiero and Hawkins, 2006). Here we present and discuss maps and explicitly evaluate model perfor- mances with respect to geography. Mapping residuals of richness-climate regression models allows us to identify specific regions and taxa responsible for model shortcomings, potentially identifying causal factors through spatial association. Spatial autocorrelation . — In a spatial regression, if residuals are spatially autocorrelated this can potentially mean that the effective sample size is lower than the actual sample size (Fortin and Dale, 2005). As a result, we assess the impacts of spatial autocorrelation by calculating the effective sample size of our data set. Then we use a subsample of grid squares from our full data set (that is even smaller than the effective sample size) to evaluate the robustness of our results. Equation 5.17 from Fortin and Dale, (2005) allows one to find the effective sample size of a spatial data set for a given spatial autocorrelation p between adjacent grid squares: n' « = n0 1+p Where n is the sample size, n’ is the effective sample size, and 0 is the approximate correction factor © = (l-p)/(l-Hp). To find the effective sample size of our data set, we first calculated the autocorrelation coefficient p between the regression residuals of adjacent squares in a regression using all 8,760 squares. We used residuals from the regression model with fern family richness as a response and MAT, RAN, and Biol5 (precipitation seasonality) as predictors, mapped in Fig. ID. For this model p = 0.8946 between adjacent squares. We obtain similar results for the residuals of other regressions presented in the study. Plugging this value into equation 5.17 from Fortin and Dale (2005) yields an effective sample size n’ = 487. We conservatively selected only 88 grid squares from our map using a hexagonal sampling grid with points spaced ~500km apart (which is equivalent to a p = 0.98). Using this conservative subsample, we perform standard OLS regressions in the same fashion as the 8,760 grid values. Equation coefficients and R-squared values were very similar whether we used the subsample, or the full dataset. !! Hi ill !ii 111 ill iil iil 11 I ^11 ill III 111 Hi III •1 1 ip 1 1 1 PI !! 1 111 11 1 1 I s i!l ill i!i ill 11! iii ill I il2 iil ill ill Ml 111 t ill ill iil ill 111 ill Hi li ili 1 III IJI m III III ill lll^ II 1 i § 1 1 i ill AMERICAN FERN JOURNAL: VOLUME 103 NUMBER ■ BOGONOVICH ET AL.: NORTH AMERICAN ] [ AND LYCOPHYTE RICHNESS Species richness.— A coarse latitudinal pattern in fern species richness is observed with generally higher richness found at lower latitudes (Fig. lA). Species richness centers occur in the southern half of the continent whereas species poor areas are found in the northern third. Species richness does not reach its peak in a single latitudinal band, nor is the pattern unimodal. Instead there are four richness peaks each with between 60-82 species (per grid square). These include two mid-latitude centers, one surrounding Washington state (herein Northwest), which reaches 62 species per square and another surrounding the Appalachians and the Great Lakes region (herein Northeast), which reaches 75 species per square. Two richness centers are identifiable farther south, one centered around Arizona and New Mexico (herein Southwest) reaching 70 species per square and a second, the highest richness peak, in Florida with 82 per square. The Northeastern richness center is the largest in geographical area, stretching from Nova Scotia and New England to the Great Lakes and extending south along the Appalachians. Richness poor regions include the Great Plains, the Ganadian Arctic, and a richness trough between Florida and the northeast. Lycophytes, like ferns, have species richness peaks in the U.S. northeast and northwest, though lycophyte richness and lycophyte richness peaks are comparably more northern, and the Great Plains is a species poor region for both groups. Unlike ferns, lycophytes lack species richness centers in Florida and the Southwest (Fig. IF). The pattern of pteridophyte species richness is dominated by ferns, the larger constituent taxon (Fig. lE). Fern richness at higher taxonomic levels.— There are large differences in patterns of fern richness at the three taxonomic levels (Figs. lA-^). Richness patterns observed at genus and family levels more closely approach a unimodal pattern than does species richness. The southwestern richness pe^ observed for species does not exist at the family level, and is much reduced at the genus level. Similarly the rise in Northwestern species richness relative to surrounding areas is not as large at the family level. The Northeastern species richness center, observed at the species level, shifts south and is less pronounced at the family level and nearly merges with the Floridian richness peak. At the family level there is essentially a single richness center in the eastern United States, reaching its maximum in Florida, the subtropical region of the study area. Richness-climate relationships. -Regressions between climate variables and fern species, genus, and family richness reveal interesting patterns (Table 2l The strongest relationships found are with fern family richness (with AET, R = 0 678- MAT R^ = 0 676) The same climate variables explain substantially less variation in species richness (AET. = 0.449; MAT. = 0.539), Climate relationships are consistently stronger for fern family richness han genus richness. Most of our discussion below on richness-climate models concerns fern family richness. As expected, multiple regressions incorporating two variables explain more variation in richness than single variable regressions. MAT and AET explain 202 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER - 77.6% and MAT and RAN explain 78.1% of the variation in fern family richness (Table 2; Fig. 2], roughly 10% more than any single variable. Warmer and wetter regions generally contain more fern families (Fig. 2a-b). Warmer regions that are also wet have slightly more families than regions that are warm but not as wet. Additionally, dry regions that are very cold have fewer families than similarly dry regions that are warmer. These simple multiple regressions incorporating one temperature variable and one liquid water variable slightly out-performed the IGM I (R^ = 0.781 vs. 0.711 with family richness as the response variable). A best subsets regression on the variables included in Table 1, confirmed that the best two- variable model was simply MAT and RAN. The best three variable model included MAT, RAN, and precipitation seasonality (Biol 5, Table 1), and explained 80.7% of the variation in fern family richness (Table 2). Again, as found for univariate regressions, the multiple regressions consistently explained more variation in family richness than species richness. A map of the residuals of the recession of fern family richness as a function of MAT, RAN, and Biol 5 (precipitation seasonality) illustrates regions where this regression model over-predicts and under-predicts family richness (Fig. ID, presented are the residuals of the model using all 8,760 squares). There are more families along western North American coastal mountain ranges than predicted by this regression and fewer fern families in the mid-longitude Great Plains I mill AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) BOGONOVICH ET AL.: NORTH AMERICAN FERN AND LYCOPHYTE RICHNESS ■Ill AMERICAN FERN JOURNAL: VOLUME 103 NUMBER ■ Figs. 3-7. Contir mil BOGONOVICH ET AL.: NORTH AMERICAN FERN AND LYCOPHYTE RICHNESS 207 Figs. 3-7. Continued. 210 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER Broader scales may put these patterns in perspective. Mexico has over 1,000 fern species (Mickel and Smith, 2004), more than twice as many as in the much larger, more northern study region. On a global scale, fern species richness patterns may be less unusual and species richness-climate relationships stronger, as temperate richness centers are swamped out by much larger richness centers closer to the Equator. Fern richness-climate regression models should be tested in future studies that incorporate larger areas and lower latitudes. Further observations . — The proportion of the continental fern flora found in highly diverse regions differs among taxonomic levels. The highest fern species richness is found in regions around Florida followed by the northeast. They contain 82 and 75 species, respectively, or —20% of the continental pool of 387 species. The percentages increase with taxonomic level. The highest concentrations of genera and families occur in Florida, with 42 genera, or 62.7% of the continental pool of 67 genera, and 21 families, or 91.3% of the pool of 23 families. Together with the climate relationships, this suggests that most fern families are tropical or subtropical in origin with differential expansion (of families) into temperate regions. This can be thought of as a climatic filtering of families toward regions of lower rainfall and temperature. This pattern is consistent with the out of the tropics hypothesis (OTT) articulated by Jablonski (1993) and Jablonski et al, (2006) to explain the evolutionary construction of the latitudinal diversity gradient. However, this pattern is also consistent with the tropical conservatism hypothesis (TCH; Wiens and Donoghue, 2004). The species richness patterns of individual fern and lycophyte families may provide clues about overall fern and lycoph)^e richness patterns. The Southwest species richness center (Fig. lA) is largely composed of species from one family, the Pteridaceae (42 out of 70; Fig. 6B); many of which are triploids. This pattern has been recognized before (Tryon, 1969). On the other hand the Northeast richness center appears to be the result of high numbers of species from several families including the Dryopteridaceae, Osmundaceae, and Ophioglossaceae among others. Together these observations suggest that multiple explanations may be required to explain fern species richness patterns, but also suggests that the northeast and northwest richness peaks are a general phenomenon and not the result of one or two aberrant fern families. Interestingly, two lycophyte families, which are only distantly related to ferns, have richness peaks corresponding to the northeast and northwest fern richness peaks. Other than the Pteridaceae, only the lycophyte family Selaginellaceae (Fig. 6E) reaches its peak richness in the southwest. Other groups including mammals and other plants (trees) have lower richness in Florida than in surrounding areas (Currie and Paquin, 1987; Simpson, 1964). This pattern has been hypothesized to result from the so- called peninsular effect, related to island biogeography theory (Taylor and Regal, 1978; but see Jenkins and Rhine, 2008). Peninsulas share spatial properties with islands, namely separation and distance from continental species pools. Therefore, like islands, peninsulas are expected to have fewer BOGONOVICH ET AL.: NORTH AMERICAN FERN AND LYCOPHYTE RICHNESS 211 species than is typical for the environment due to limitations on dispersal of appropriate lineages into the peninsular regions. In contrast, ferns actually experience a high richness peak in Florida. Few Florida fern species are endemic and most also exist in the West Indies or South America, with fewer affinities to the Mexican fern flora. One explanation for the pronounced fern richness center in Florida may involve ferns’ long-range dispersal capacity (Spurr, 1941). Other geographical patterns of ferns are consistent with this hypothesis. The vascular plant floras of isolated oceanic islands tend to have higher percentages of pteridophytes than the mainland (Kreft et ah, 2010; Tryon 1970). The observed high species richness of ferns and low species richness of other groups in Florida is consistent with the interpretation that ferns are less limited by barriers to dispersal than other groups. Conclusion— North American ferns have unusual mid-latitude species richness peaks, while family richness patterns are more comparable to those reported for species in other organisms. The fern species richness peaks occur in diverse climates, with different composition and numbers of higher taxa, suggesting that there may be more than one explanation required to understand fern species richness patterns. In contrast, fern family richness is strongly correlated with a small number of climate variables, suggesting a parsimonious explanation is sufficient to explain family richness patterns. The best regression model included three variables — mean annual temperature, annual rainfall, and precipitation seasonality — and explained 80.7% of the variation in fern family richness. The result that fern richness, particularly at the family level, is strongly related to water and energy variables supports the most general prediction of the productivity-diversity hypothesis. More work is needed to explore and evaluate alternative fern richness-climate models. We are grateful to Kerry Woods, Elizabeth Middleton, and Leonie Moyle, James Bever, Eileen O’Brien, and anonymous reviewers for insightful comments on the manuscript. Ceorge Yatskievych, Michael Barker, and Cerald Castony shared their expertise on ferns and lycophytes. Literature Cited Aldasoro, J. J„ F. Cabezas and C. Aedo. 2004. Diversity and distribution of ft Africa, Madagascar and some islands of the South Atlantic. J. Biogeogr. Balmford, a., M. j. B. Creen and M. C. Murray. 1996. Using higher-taxon richa species richness. 1. Regional tests. Proc. Roy. Soc. London, Ser. B, Biol. 31:157»-1604. ess as a surrogate for Sci. 263:1267-1274. Barrington, D. S. 1993. Ecological and historical factors in fern biogeography. J. Biogeogr. 20:275-279. ,, . , , Beyer, H. L. 2004. Hawth’s Analysis Tools for ArcCIS. Available at http://www.spatialecology. Bhatt^, K°°r’, O. R. Vetaas and J. A. Crytnes. 2004. Fern species richness along a central Himalayan elevational gradient, Nepal. J. Biogeogr. 31:389-400. Bickford, S. A. and S. W. Laffan. 2006. Multi-extent analysis of the relationship between pteridophyte species richness and climate. Global Ecol. Biogeogr. 15:588-601. Hfii 216 AMERICAN FERN JOURNAL; VOLUME 103 NUMBER 4 (2014) Anemia is the only genus in the family Anemiaceae, a member of the order Schizaeales (schizaeoid ferns). The Schizaeales is characterized by fertile- sterile leaf blade differentiation, absence of well-defined sori, and sporangia possessing a transverse, subapical, continuous annulus (Smith et al, 2006; Rothfels et al, 2012). The genus Anemia contains about 100-120 species, mainly distributed within the tropical and subtropical regions (Smith et al., 2006; Skog et al., 2002). In Brazil, there are 70 species and seven varieties of Anemia (Barros et al., 2013) and many endemic species occur in the southeastern and central part of the country, which is known as the center of this genus diversity (Mickel, 1962; Tryon and Tryon, 1982; Moran and Mickel, 1995). Anemia is easily identified by the sporangia on a basal pair of skeletonized, highly modified, erect pinnae (Smith et al., 2006). Skog et al. (2002) concluded that species of the genus Anemia fall into two well- supported subgenera, Anemiorrhiza and Anemia (including subg. Coptophyl- lum). Many Anemia species have aromatic fronds (Page, 2002; Juliani et al., 2004; Santos and Sylvestre, 2006) that produce volatile substances by glandular hairs (Ribeiro et al., 2007). Essential oils are volatile substances responsible for the scent, odor or smell found in many plants, and are mainly comprised of terpenes. However, other classes of chemical substances may also be present, such as phenylpropanoids and hydrocarbons (Harborne 1984). Although several species of ferns possess a quite typical smell (Page, 2002), only a few works regarding chemical characterization of their volatile constituents have been reported in the literature (Briggs and Sutherland, 1947; Boeder, 1985; Cheng and Mao, 2005; Miyazawa et al., 2007), with special efforts carried out to study the species Anemia tomentosa (Sav.) Sw. var. anthriscifolia (Schrad.) Mickel (Juliani et al, 2004; Santos et al, 2006, 2008, 2010; Pinto et al, 2007; 2009a; 2009b; Joseph- Nathan et al, 2010). The aim of the present study was to analyze the chemical composition of essential oils from two fern species, Anemia hirsuta and Anemia raddiana. Anemia hirsuta (L.) Sw. and Anemia raddiana Link. (Fig. 1) occur in ravine banks at the forest edge, in open habitats and also on rocky outcrops. Both species belong to the subgenus Anemia. Anemia hirsuta has a once-pinnate foliar blade, fertile pinnae approximate to the sterile, free veins, deeply incised pinnae, hirsute, brown rhizome hairs and a Neotropical distribution (Mickel, 1982). Anemia raddiana has a bipinnate foliar blade, fertile pinnae that are remote from sterile pinnae, red rhizome hairs and an endemic geographical distribution in Southeastern and Southern Brazil (Mickel, 1962; Schwartsburd and Labiak, 2007). To our knowledge, this is the first phytochemical characterization of essential oils from these two species of ferns. Materials and Methods Plant material— Leaves of Anemia hirsuta were collected during the day (March, 2009) in ravine banks at “Serra das Emerencias” in the municipality of Armagao de Biizios, Rio de Janeiro state, Brazil (22°47'53.2"S,-^1°56' 0.5"W), SANTOS ET AL.; ESSENTIAL OILS OF ANEMIA SPECIES 217 Fig. 1. Anemia raddiana Link. (A] and Anemia hirsuta (L.) Sw. (B). and a voucher sample was deposited under the register number RFFP - 13.790. Leaves oi Anemia raddiana were collected (February, 2010) in ravine banks at “Institute Zoobotanico de Morro Azul”, municipality of Engenheiro Paulo de Frontin, Rio de Janeiro state, Brazil (22°29'42.28"S-43==34'2.37"W) and a voucher sample was deposited under the register number RFFP - 13.791. The voucher samples were deposited at the herbarium of Faculdade de Formagao de Professores, Universidade do Estado do Rio de Janeiro (RFFP). Essential oils extraction.— Fresh leaves of A. hirsuta (0.7 kg] and A. raddiana (0.9 kg) were individually turbolized with distilled water and then submitted to hydrodistillation for three hours using a Clevenger-type apparatus. The resulting essential oils extracts were dried over anhydrous sodium sulphate and stored at approximately 4°C until analysis. Gas chromatography/mass spectrometry analysis.— One microliter of each sample was dissolved in CH2CI2 (1:100 mg/pL) and analyzed by GC/MS (Shimadzu QP5000) using a DB-5MS column (5% phenylmediyl silicone, 30 m X 0.25 mm ID, 0.25 pm film thickness). Gas chromatographic (GC) conditions were as follows: injector temperature 260X; detector temperature 290=C; carrier gas (helium), flow rate 1 mL/min and split injection with a 1:40 split ratio. Oven temperature was initially held for 2 min at 60°C, and then raised to 290=C at a rate of 3°C/min. Mass spectra were recorded at 70eV with a mass range from m/z 35-450 and scan rate of 1 scan/sec. The Retention Indices (RI) were calculated by linear interpolation to the retention times of the eluting peaks with the retention times of a mixture of aliphatic hydrocarbons (C5-C24) AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) analyzed in the same conditions (Van Den Dool and Kartz, 1963). The compounds identification was performed by comparison of their retention indices and mass spectra with those reported in literature (Adams, 2007), The MS fragmentation pattern of compounds was also checked with NIST mass spectra libraries. Quantitative analysis of the chemical constituents was performed by flame ionization gas chromatography (GC/FID) under the same conditions of GC/MS analysis and percentages obtained by FID peak-area normalization method. Data analysis. — Sorensen’s similarity index was used to compare the essential oil composition between the Anemia species (Brower et al, 1997). Results Essential oils were obtained from fresh ferns and yielded 0.04% and 0.03% w/w for Anemia radianna and Anemia hirsuta, respectively. Anemia radianna essential oil was analyzed by GC/MS and nine compounds were identified, corresponding to 81.3% of the total oil composition (Table 1). The major compound was p-selinene (46.8%). Thirteen substances were identified by GC/MS analysis of A. hirsuta essential oil (Table 1). The substances corresponded to 92.3% of the total oil composition (Table 1). The most abundant constituent was the sesquiterpene p-caryophyllene (48.7%). The remaining substances found in both essential oils are listed in Table 1. No similarity between the essential oils constituents of A. hirsuta, A. radiana were detected (Tables 1 & 2). Discussion The present study found that the essential oils from Anemia hirsuta and Anemia radianna are mainly composed of monoterpene hydrocarbons, oxygenated monoterpenes, sesquiterpene hydrocarbons and oxygenated sesquiterpenes. The sesquiterpene p-caryophyllene, which is the most abundant constituent of the essential oil of A. hirsuta, is recognized by its antifeedant and insecticidal activity (Barakat, 2011; Glinwood et al., 2011); it also displays phytotoxic activity (Glinwood et ah, 2011) and can accumulate in soil (Asensio et al., 2008), Similarly, P-selinene, the major substance found in essential oil from A. radianna, is known to have allelopathic effects (Bewick et al., 1994). Many terpenoids are known to possess allelopathic properties (Reigosa et al., 2006), mediate direct and indirect plant defenses (Pichersky and Gershenzon, 2002) and improve thermotolerance in plants (Sharkey et al., 2001). Extracts obtained from A. villosa and Anemia tomentosa var. anthriscifolia were able to inhibit germination and development of lettuce seedlings (Moraes et al, 2003). Fern species, among them Pteridium aquilinum (L.) Kuhn and Dicranopteris linearis (Burm. f.) Underw. produce allelopathic substances in order to form dense populations (Walker and Sharpe, 2010). Several classes of secondary (or special) metabolites can be allelochemical SANTOS ET AL.: ESSENTIAL OILS OF ANEMIA SPECIES RI-Retention Indices on DB-5 5 (relative intensity): 41, 53, 67(100), 79, 93, s (relative intensity): 41, 43(100), 67, 81. 95, I (relative intensity): 41(100), 55, 67, 79, 93, z (relative intensity): 41(100), 55, 67, 79, 93^ z (relative intensity): 41, 55, 67, 81, 95(100), : (relative intensity): 55, 67, 81, 95(100). 107, 121 z (relative intensity): 43. 67. 71(100), 95, 107, 121 1 . 187, 205, 220. agents, including terpenoids and phenolics substances (Harborne, 1999; Ferreira and Aquila, 2000; Reigosa et al., 2002). More studies should be conducted to evaluate the action of the volatile mixture of Anemia species as allelopathic agents. Walker et al. (2010) emphasized the importance of allelopathy as an important area for research expansion in fern ecology. Volatile substances released by plants may directly protect them against insect herbivores or do so indirectly by attracting natural enemies of the herbivores (Pare and Tumlinson, 1999; Pichersky and Gershenzon, 2002). Additionally, plant volatiles may function as a form of communication, SANTOS ET AL.: ESSENTIAL OILS OF ANEMIA SPECIES growing in Argentina (Cordoba province), at 960m above sea level, were characterized by oxygenated sesquiterpenes with a-bisaholol (51.4%) as the major secondary compound (Juliani et al, 2004). In contrast, the major compound of the essential oil in A. raddiana and A. hirsuta was p-selinene (46.8%) and p-caryophyllene (48.7%), respectively (Table 1). In all studies above, the essential oils were obtained by hydrodistillation from fresh aerial parts in a Clevenger-type apparatus. Intraspecific variation in the quantitative and qualitative aspects of essential oil production can be influenced by abiotic factors like water and mineral nutrients and biotic factors such as insect damage (Gershenzon, 1983; Pare and Tumlinson, 1999). Hence, the existence of chemotypes is an important aspect of intraspecific essential oils variation (Thompson et al., 2003). The phylogenetic relationship between species of Anemia is an important factor to be considered when assessing interspecific variation in essential oils. Until recently, three subgenera were recognized within the genus Anemia, Coptophyllum (that included A. tomentosa var. anthriscifolia and A. raddiana), Anemia (that included A. hirsuta) and Anemiorrhiza (Tryon and Tryon, 1982). More recently, Skog et al. (2002) concluded that species of the genus Anemia fall into only two well-supported subgenera, Anemiorrhiza ^d Anemia (including subg. Coptophyllum). In accordance with the classification by Mickel (1962), the suhgenus Coptophyllum had six sections, and A. tomentosa var. anthriscifolia and A. raddiana were classified in the section Tomentosae. This same author noted that many phylogenetic problems exist within this section due to their ability to hybridize. To understand the biodiversity is important to evaluate macro and micromolecular characters, mapping and quantifying chemical and biological data (Gottlieb et at., 1996, 1998). In this context, the chemical data of the essential oils may help future ecological and evolutionary analyses in the genus Anemia. To our knowledge, these are the first contributions to the chemical characterization of the essential oils from Anemia raddiana and Anemia hirsuta. Thus, as part of our ongoing studies of Brazilian fern species, characterization of essential oils may be helpful in the study of the ecology and chemotaxonomy of Anemia and other ferns. The authors thank CNPq (Conselho Nacional de Dej CAPES (Coordenagao Carlos Chagas Filho de Amparo & Pesqui financial support, ani' species identification editors Gary K. Greer Superior), FAPERJ (Fundagao do Estado do Rio de Janeiro) and PROCIENCIA-UERJ for (The New York Botanical Garden) for the confirmation of d Valuable comments and suggestions of the American Fern Journal Warren D. Hauk. Literature Cited 1 Gomponents by Gas Chromatography/Mass 222 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) Asensio, D., S. M. Owen, J. Llusia and J. Penuelas. : the soil horizons around Pinus halepensis ti Barakat, D. a. 2011. Insecticidal and antifeedant a Edulis La Llave & Lex (Rutaceae) leaf extrac . Soil Biol. Biochem. 40:2937-2947. ities and chemical composition of Casimiroa id its fractions against Spodoptera littoralis larvae. Aust. J. Basic & Appl. Sci. 5:693-703. Barros, I. C. L., A. C. P. Santiago and A. F. N. Pereira. 2013. Anemiaceae in Lista de Especies da Flora do Brasil. Jardim Botanico do Rio de Janeiro. Available at: http://floradobrasil.jbrj.gov. br/2012//FB090589. Accessed on 10 January 2013. Bewick, T. A., D. G. Shilung, J. A. Dusky and D. Williams. 1994. Effects of celery [Apium graveolens] root residue on growth of various crops and weeds. Weed Technology 8:625-629. Briggs, L. H. and M. D. Sutherland. 1947. A terpene-type essential oil from a fern [Paesia scaberula]. Nature 160:333. Brower, J. E., J. H. Zar and C. N. von Ende. 1997. Field and Laboratory Methods for General Ecology. WCB/McGraw-Hill, Boston. Cheng, C. and J. Mao. 2005. Constitution of volatile oils from three kinds of pteridophyte plants. Chem. Industr. Forest Prod. 2:107-110. Ferreira, A. G. and M. E. A. Aquila. 2000. Alelopatia: uma drea emergente da ecofisiologia. Revista Brasil. Fisiol. Veg. 12:175-204. Gershenzon, j. 1983. Changes in the levels of plant secondary metabolites under water and nutrient stress. Recent Advances Phytochem. 18:273-320. Giluj, Y. G., R. M. Gleiser and J. A. Zygadlo. 2008. Mosquito repellent activity of essential oils of aromatic plants growing in Argentina. Bioresource Technol. 99:2507-2515. Gunwood, R., V. Ninkovic and J. Petterson. 2011. Chemical interaction between undamaged plants - Effects on herbivores and natural enemies. Phytochemistry 72:1683-1689. Gobbo-Neto, L. and N. P. Lopes. 2007. Plantas medicinais: fatores de influencia no conteiido de metabolitos secund4rios. Quim Nova 30:374-381. Gottlieb, O. R., M. A. C. Kaplan and M. R. M. B. Borin. 1996. Biodiversidade: Urn Enfoque Quimico- Biologico. Editora UFRJ, Rio de Janeiro. Gottlieb, O. R., M. R. M. B. Borin, C. L. A. C. Pagotto and D. H. T. Zocher. 1998. Biodiversidade: o enfoque interdisciplinar brasileiro. Ciencia & Saiide Coletiva 3:97-102. Harborne, J. B. 1984. Phytochemical Methods: A Guide to Modem Techniques of Plant Analysis. Chapman and Hall, London. Harborne, J. B. 1999. Recent advances in chemical ecology. Nat. Prod. Rep. 16:509-523. Martinez, E. Dellacassa, A. HernAndez-BarragAn and N. PiaiEz-HERNANDEz. 2010. Sfructure reassignment and absolute configuration of 9-epi-presilphiperfolan-l-ol. Tetrahedron Lett. 51:1963-1965. JuuANi, H. R., J. A. Zygadlo, R. Scrivanti, E. R. Sota and J. E. Simon. 2004. The essential oil of Anemia tomentosa (Savigny) Sw. var. anthriscifoha (Schrad.) Mickel. Flavour Fragr. J. Meirelles, S. T., E. a. deMattos and A. C. Silva. 1997. Potential desiccation tolerant vascular plants from Southeastern Brazil. Polish J. Environm. Stud. 6:17-21. Mendes, M. M., j. Dinis, J. Pais and E. M. Frhs. 2011. Early cretaceous flora from Vale Painho (Lusitanian basin, western Portugal): An integrated palynological and mesofossil study. Rev. Palaeobot. Palynol. 166:152-162. Mickel, J. T. 1962. A monographic study of the fern genus . State Coll. J. Sci. 36:349-^82. nia, subgenus Coptophyllum. Iowa Mickel, J. T. 1982. The genus Anemia (Schizaeaceae) in Mexico. Brittonia 34:388-^13. Miyazawa, M., E. Horiuchi and J. Kawata. 2007. Components of the Essential Oil from Matteuccia struthiopteris. J. Oleo Sci. 56:457-461. Moraes, M. G., a. a. Q. Oliveira, J. S. P. Barbosa and M. G. Santos. 2003. Efeito alelop4tico de extratos aquosos de Anemia tomentosa e A. villosa (PTERIDOPHYTA) na germinagao e crescimento de alface. Braz. J. PI. Physiol. 15:288. Moran, R. C. and J. T. Mickel. 1995. Anemia. Pp. 56-56 in R. C. Moran and R. Riba. Flora Mesoamericana. Universidad Nacional Autonoma de Mexico, Ciudad de Mexico. L Fern Journal 103(4);225-240 I Helical Cell Wall Thickenings in Root Cortical Cells of Polypodiaceae Species from Northwestern Argentina Marcela a. Hernandez, Lucrecia TerAn, and Marisol Mata Instituto de Morfologia Vegetal, Fundacidn Miguel Lillo, Migue l Lillo, 251, 4000, Tucumdn, Argentina, e-mail: marcela.alicia.hemandez@gmail.com Olga G. MartInez IBIGEO, Herbario MCNS, Facultad de Ciencias Naturales, Universidad Nacional de Salta, CP 4400, Salta, Argentina Jefferson Prado Herbdrio SP, Instituto de Botanica, Av. Miguel Est^fano, 3687, CEP 04301-012, Sao Paulo, SP, Brazil Abstract. — The occurrence of helical cell wall thickenings in fem roots is not well investigated and there are few records about it in the literature. To assess the presence of thickenings and their chemical composition, we studied all species of Polypodiaceae, which grow in northwestern Argentina, using light microscopy, scanning electron microscopy, and fluorescence microscopy. Twenty of the twenty-one species studied showed the thickening in the roots. Only in Melpomene peruviana are helical cell wall thickenings absent. All thickenings have cellulose as the main compound. The stmcture of the thickening may be classified as simple, furcate, or anastomosing. All data presented in this paper corroborate the same stmcture and chemical composition of thickenings previously reported for Aspleniaceae. Key Words. — Anatomy, Eupolypods I, ferns, epiphytes, cellulose The presence of a tissue similar to the velamen has been previously recorded in the outer cortical zone of roots for some Polypodiaceae. Geisenhagen (1901, apud Pande, 1935] was the first author to describe helical thickenings in cell walls of ferns, namely Niphobolus penangianus Hook. Also Pande (1935] reported thickenings in the root cortical parenchymatic cells of Niphobolus adnacens (Sw.) Kaulf. (= Pyrrosia lanceolata (L.) Farw.), and observed pits in the cell wall between these thickenings. These thickenings appear as strands internally in the cells. Ognra (1972) reported something similar for roots of Pyrrosia, but gave no details. Schneider (1997) studied the root anatomy of the Aspleniaceae and reported spiral wall thickenings in the outer cortex of some species. He mentioned that these thickenings can also be found in epiphytic ferns such as Grammitida- ceae, Polypodiaceae, and in some Vittariaceae, but he did not present additional details. He suggested that these cells with helical thickenings function like the cells of the velamen of Orchidaceae. Recently, Leroux et al. (2011) investigated the structure of these root cortical cells for 96 specimens (representing 72 species) of Aspleniaceae and designated them as helical cell wall thickenings (henceforth called ‘thicken- ings’). These authors defined them as simple helical thickenings that branch dichotomously, sometimes forming a net-like pattern. They also determined hernAndez ] iililiilillii 3 ^ 1 il |! |i II 232 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER ■ HERNANDEZ ET AL.: HELICAL CELL WALL THICKENINGS IN POLYPODIACEAE 233 Fig. 2. Mean and Standard deviation (95% confidence intervals) of (A) Thickening’s size; (B) Number of thickenings per cell. 234 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER ■ Histochemical results indicate that the inner cortex did not stain red with Phloroglucinol, indicating a lack of lignin (Fig. IB). This was further confirmed with the differential coloration Safranine-Astra Blue (Fig. IC). Observations made under the Fluorescence Microscope showed that this tissue has cellulose (Fig. ID). All tests for the presence of suberin were negative for the inner or outer cortex, but the overall result was positive for the endodermis, because of the presence of the Casparian bands. Measurements of all studied species were carried out and final results are presented in Table 2 and Fig. 2. The Campyloneurum species (C. aglaolepis, C. lorentzii, and C. tucumanense) do not show significant differences in the width of the thickenings. The number of helical thickenings per cell is quite variable among the species. For example in C. tucumanense, a rupicolous species, there are 9-45 thickenings per cell, whereas C. lorentzii (Fig. 3A), an epiphytic or rupicolous species 17^0 strands. Compared to other species, Lellingeria tunguraguae showed the narrowest thickenings (0.5-1. 2 pm) and least number of thickenings per cell. The species of Pecluma (P. barituensis, P. filicula, P. oranensis, and P. venturii), all epiphytes, have similar thickenings. HERNANDEZ ET AL.: HELICAL CELL WALL THICKENINGS IN POLYPODIACEAE Thickening types. (A) Simple and furcate; (B) Furcate and anastomosing: (C) Anastomosing, among strands there are primary field pits (p); (D) Detail of the cell wall showing irregular deposits of cellulose on the thickenings (d). (A, D) Campyloneumm aglaolepis; (B, C) Pleopeltis tweediana. resembling those of Microgramma. Phlebodium areolatum (also epiphytic) has thin thickenings (about 1.0 pm wide) and the biggest amount of thickenings per cell. The species of Pleopeltis [P. hirsutissima (Fig. 3B), P. bryopoda (Fig. 3C), P. macrocarpa, P. minima (Fig. 3D), P. pinnatifida, P. pleopeltidis, and P. tweediana), all of which are epiphytic, had the largest thickenings measure- ments, from 0.9-^.4 pm wide (Table 2). Polypodium chrysolepis and Serpocaulon gilliesii showed similar measure- ments between them in relation to the width and number of thickenings per cell. In general the number of thickenings per cell varies from 15-25 for most of the species studied. Campyloneumm lorentzii and Phlebodium areolatum, which are primarily epiphytic or rarely rupicolous, have about 30 thickenings per cell. Most of the species studied had thickenings that were simple or furcate (Table 2). Pleopeltis bryopoda, a saxicolous fern, has more furcate thickenings than simple ones, and there are several per cell. Some of the furcate thickenings joined each other, but did not form a net or reticulum. This kind of anastomosing is present in 50% of the analyzed taxa (Fig. 4A-C). mi AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014] 238 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) On the other hand, Leroux et al. (2011) demonstrated that thickenings support a monophyletic (BS 100%) group in Asplenium. Unfortunately, our data for Polypodiaceae are preliminary to conclude any relation among the groups into the family based on the thickenings. Because the regions of contact among the furcate thickenings are few (Figs. 4A, D), we called these thickenings ‘anastomosing’ (Figs. 4B, C) and not reticulate. On the other hand, the main kinds of thickenings observed were simple or furcated (Figs. 4A, D). Only in the epiphytic species of Micro- gramma, Phlebodium, and Pleopeltis are the anastomosing thickenings predominant; in Campyloneurum, another genus with epiph34;e species, these anastomosing thickenings are few. According to Pande (1935) and Ogura (1972), the thickenings in Polypodiaceae are helical and simple, whereas Leroux et al. (2011) classified the thickenings of Aspleniaceae into three types: simple, furcate, and reticulate. Our results demonstrate the absence of pits in the thickenings. Leroux et al. (2011) also reported this for the Aspleniaceae. In the Polypodiaceae Pande (1935) cited the presence of primary pit fields only in the cell wall between the thickenings. About the chemical composition of secondary cell wall of roots, our results for different stains, fluorescence, and Phloroglucinol/HCl test revealed that the thickenings are basically formed of cellulose and that there is no deposition of lignin on them. Leroux et al. (2011) pointed out the same conclusions for Aspleniaceae. Additionally these authors commented that the thickenings present in the cell wall of the external cortical root zone are related to the mechanical protection of these small roots, acting as skeleton to the delicate roots of Aspleniaceae species to avoid root collapse during the dry periods in terrestrial, rupicolous, and epiphyte plants. The same interpretation might be applied to the species of Polypodiaceae here investigated. Among the analyzed species, all species of Pleopeltis (that are mainly epiphytic) showed pronounced development of the thickenings, probably as an adaptation to their revivescent ecology. Given our results and those previously published (Schneider, 1997; Leroux et al., 2011), thickenings are known only for Eupolypods I (in Polypodiaceae), Eupolypods II (in Aspleniaceae) (Smith et al., 2006), respectively. Therefore, this character might be better explored for all Polypodiaceae members. It is highly probable that these structures appear in other members of Polypodiales, but more studies are needed to confirm this hypothesis. In fact, the presence of this structure in other groups of ferns, their types, and its evolution into ferns and lycophytes remains obscure. This project was funded by Fundacidn Miguel Universidad Nacional de Salta. We thank to material: to Fabiana Rios, Technician of the fei locating specimens at LIL. We are a and by the Consejo de Investigaciones de la AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) Schneider, H., E. Schuettpelz, K. M. Fryer, R. Cranfill, S. Magallon and R. Lupia. 2004b. Ferns diversified in the shadow of angiosperms. Nature 428:553-557. Schuettpelz, E. and K. M. Fryer. 2007. Fern phylogeny inferred from 400 leptosporangiate species and three plastid genes. Taxon 56:1037-1050. Silva, G. B. da, M. Ionashiro, T. B. Carrara, A. C. Crivellari, M. A. S. Tine, J. Frado, N. C. Carpita and M. S. Buckeridge. 2011. Cell wall polysaccharides from fern leaves: Evidence for a mannan-rich Type IB cell wall in Adiantum raddianum. Fhytochemistry 72:2352-2360. Smith, A. R., K. M. Fryer, E. Schuettpelz, F. Koral, H. Schneider and F. G. Wolf. 2006. A I Fem Journal 103(4):241-244 (2014) Shorter Notes Hydrochemical Characterization of A Stand of the Threatened Endemic Isoetes malinvemiana. — Isoetes malinverniana Ces. & De Not. is an aquatic quillwort endemic to Northern Italy that, as with many other quillworts, is facing drastic changes in its habitat (Wen et al, J. Freshwater Ecol. 18:361-367. 2003.). Isomes malinverniana grows in running water canals used for rice field water supply (usually with Ranunculion fluitantis vegetation), but in the past it probably occurred in natural streams generated by springs and minor river branches, characterized by oligotrophic waters (Abeli et al., Aquatic Conserv: Mar. Freshw. Ecosyst. 22:66-73. 2012). During the last forty years the distribution range of /. malinverniana has been rapidly decreasing as a consequence of changes in the rice cultivation practices in Northern Italy. Particularly, the use of herbicides and fertilizers, the regimentation of water courses with water removal in winter, and the mechanical re-profiling of the canals are the major threats to the species (Barni et al., Aquat. Bot. 107:39-46. 2013). Although the species is protected at European and national levels, its conservation status is critical (Bilz et al., European Red List of Vascular Plants. Publications Office of the European Union, Luxembourg. 2011; Rossi et al., Lista Rossa della Flora italiana. 1. Policy Species e altre specie minacciate. Comitato Italiano lUCN e MATTM. 2013), and urgent conservation actions are needed to stop the decline of the species, by reinforcing some extant populations and/or reintroducing new populations within the historical range. The major problem with respect to reintroduction actions is the fact that the original habitat of I. malinverniana has been greatly modified. This reduces the probability of successful translocations and also affects the possibility to study the real ecological requirements of the species. However, a site of /. malinverniana with about 30 plants discovered a few years ago in a natural stream in the Ticino river basin at La Sforzesca near Vigevano (voucher specimen in PAV) still has many of the characteristics of the original habitat. Here we have analyzed the ion concentrations of surface water and, for the first time, sandy sediment pore water along a 30 m long transect crossing the population. Pore water samples were collected in three points along the transect at about 10 m from each other, while a single surface water sample was collected in the middle of the transect. Sediment pore water samples were collected at a depth of about 10 cm using ceramic cups (Eijkelkamp Agrisearch Equipment, Giesbeek, the Netherlands), connected to 100% vacuum PVC syringes (50 ml) by means of a PVC tube, according to Van Der Welle et al. (Freshwater Biol. 52:434-447. 2007). For each water sample, pH, electrical conductivity and temperature were immediately recorded, while ion concen- trations were analyzed at the laboratory of the Radboud University (Nijmegen, The Netherlands). Surface water pH, electrical conductivity and temperature were identical among the three samples (6.8; 271 pS/cm; 11°C, respectively). This pH was AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) lower than in populations within channels used for rice field water supply (Bami et ah, 2013; Table 1), but higher than the mean pH values found in Dutch and Norwegian populations of /. lacustris and I. echinospora (Table 1). The pH of the sediment pore water was remarkably high, reaching a mean value of 8.7 ± 0.9, which also results in very low pore water carbon dioxide concentrations. Mean pH values in Dutch and Norwegian Isoetes stands are around 6, whereas pore water carbon dioxide concentrations are typically (much) higher than surface water concentrations (Table 1). Being able to take up carbon dioxide by the roots, this situation provides a competitive advantage for isoetid species compared to non-isoetid species, which lack this ability (Smolders et al., Aquat. Bot. 73:325—350. 2002). Therefore Isoetes species thrive very well in soft water lakes and streams where carbon availability in the water layer is low and species dependent on carbon uptake from the water layer are unable to grow. The high calcium, magnesium and bicarbonate concentrations of the water layer and sediment pore water indicate that the water in the stand from the Ticino river basin is relatively well-buffered compared to the Dutch and the Norwegian Isoetes stands, which were found in weakly buffered soft waters. The Ticino population is probably dependent on the uptake of carbon dioxide from the water layer where carbon dioxide concentrations are much higher than in the sediment pore water. Pedersen et al. (J. Exp. Bot. 62:4691—4700. 2011.) showed the potential gas exchange via the leaves to be substantial for /. australis, although the resistance to gas exchange was up to three times higher than for roots. The uptake of CO2 via the roots may have further lowered CO2 concentrations and indirectly increased the pH of the sediment pore water in the Ticino stand. Regarding the surface water, the concentration of phosphate, total-P and ammonia are at the lower end of the ranges found for Dutch and SW Norwegian Isoetes stands and are also lower than the mean values found for other /. malinverniana stands (Table 1). However, the nitrate concentration is unnaturally high (Table 1), suggesting water nutrient enrichment probably due to the presence of rice and cornfields just a few hundred meters upstream from the population. High concentration of nitrate was also evident in sediment pore water. Although for the parameters analyzed in the sediment pore water, it was not to possible to make a comparison with other stands of I. malinverniana, pore water nitrate was very high compared with values measured in Dutch and Norwegian Isoetes stands (Table 1) and values reported for isoetid lakes in Spain (Catalan et al., Hydrobiologia 274:17-27. 1994; Gacia et al., J. Limnol. 68:25—36. 2009) and in Scandinavia (Vestergaard & Sand-Jensen, Aquat. Bot. 67:85— 107. 2000). Isoetes species generally grow on mineral, usually sandy, sediments with low oxygen consumption rates and actively maintain the sediment in an oxidized state by leaking oxygen via the roots (Pedersen et al., 2011). Due to the oxidized conditions nitrate, iron and sulphate reduction, are normally not important in isoetid stands. This results in low iron concentra- tions in sediment pore water, as iron is not mobilized by iron reduction (Table 1), which also results in a low mobility of phosphorus, which is 244 AMERICAN ] [ JOURNAL: VOLUME 103 NUMBER 4 (2014) efficiently bound to oxidized iron(hydr)oxides (Smolders etal., 2002). Also the observation that at our location nitrate and sulphate concentrations do not differ between water layer and pore water suggest the lack of microbiological consumption of nitrate or sulphate in the sediment, indicating oxidative conditions in the sediment. The chemical characterization of the I. malinverniana habitat shows that the species is growing on an oxidized sediment with a relatively low availability of phosphorus. The water layer and sediment are relatively well buffered and characterized by a high pH and a very low availability of carbon dioxide in the sediment pore water. Although the surface water phosphorus concentrations are still low, the nitrate enrichment due to the intensive agricultural activity of the area is evident even in a site apparently less impacted than other stands. As a consequence even a temporary increase of the phosphorus availability in the water layer might easily lead to an excessive grovvdh of algae. This poses serious threats for the conservation of this endemic species for which a suitable site for translocation is at the moment unavailable. Flowing water may strongly benefit the species under more eutrophic conditions because in flowing water algae are less likely to become dominant and uptake of CO2 from the water layer is facilitated. Nevertheless increased nutrient levels in the water layer will at least lead to the growth of epiphytic algae as has been observed in many of the remaining 1. malinverniana populations (e.g. Arborio, Vigevano), which may depress growth by shading and by depletion of inorganic carbon and nutrients at the leaf surface.— T. Abeli (e-mail: thomas.abeli.it@gmail.com), S. Orsenigo and N. M. G. Ardenghi, DSTA, Department of Earth and Environmental Sciences, University of Pavia, via S. Epifanio 14, 27100, Pavia, Italy, E.C.H.E.T. Lucassen and A.J.P. Smolders, Department of Aquatic Ecology and Environmental Biology, Radhoud University, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands. Fem Journal 103(4):245-250 (2014) Review Flora of China, vols. 2-3, Lycopodiaceae through Polypodiaceae, by Wu Zhengyit and Peter H. Raven, Co-chairs of the Editorial Committee and Hong Deyuan, Vice Co-chair of the Editorial Committee. Science Press (Beijing) and Missouri Botanical Garden Press (St. Louis). ISBN 978-0-915279-34-0 (series), ISBN 978-1-935641-11-7 (vols. 2-3). Publication date: 6 June 2013. 959 pages; English, with Chinese authors’ names and Chinese common names in both roman and Pinyin, and roman and Pinyin indices of Chinese common names (no common names are given in English). No standard format for citing this work is suggested, but the MBG Press website indicates that the author of the series is the Flora of China Editorial Committee. Available from the Missouri Botanical Garden Press (www.mbgpress.info). $190 plus shipping. Text also available on-line at www.efloras.org. The final text volume of the great new Flora of China, published as a cooperative venture by a consortium of Chinese botanists and international experts, has arrived. Inspired by Peter Raven, it was originally proposed at a joint meeting of Chinese and American botanists held at the University of California, Berkeley, in 1979. It was formally initiated in 1988 with the appointment of a joint Sino- American Editorial Committee, with Raven and Wu as Co-chairs. Further details of its inception are given in the Foreword to the first published volume, 17: vii-viii, 1994. It is a testament to the foresight of Raven and Wu (who passed away in June 2013), and a proud achievement for China and the world botanical community. Planned as an updated and revised English language edition of the Flora Reipublicae Popularis Sinicae (FRPS, 1959-2004), it is remarkable that this Flora - now 24 volumes bound in 21 books and awaiting only the introductory volume - has been brought to fruition in such a short time. Just coordinating the contributions from 59 authors for the current volume (26 Chinese authors and 33 others) must have been a difficult task. The work was undertaken at eleven botanical institutions, four in China and seven elsewhere, with overall organization and coordination at the Missouri Botanical Garden. Details may be seen at the Project website, flora.huh.harvard.edu/china. Throughout the entire Flora, the Editorial Committee has been under the unbroken and steady leadership of both Raven and Wu. The undertaking was massive, its execution nearly flawless. This combined volume is dedicated to the Chinese pteridologist Ching Ren-Chang (1898-1986), who was Secretary General of the FRPS project. For pteridologists, last is best, given the substantial progress made in our understanding of the phylogenetic relationships of ferns and lycophytes since the project began, not to mention the additional floristic knowledge and discoveries. We commend the editors and authors, and recommend the book to anyone seriously interested in Chinese botany or ferns. 246 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER 4 (2014) China is enviably wealthy in pteridophytes, particularly ferns, and this wealth is arrayed for study by anyone who takes up this book. As stated in the preface, “[The pteridophyte flora of China] includes 38 families, 177 genera, and 2,129 species, among which three genera and 842 species are endemic to China, and one genus and four species are introduced to China.” This is eye opening for North American fern lovers: we have only 96 genera and 554 species (FNA 1993). Various historical events, especially the Himalayan orogeny and the recent glaciations of North America, have set the two areas on different paths, and China is the richer by far. One has only to contemplate its 123 species of Athyrium vs. our two species, its 167 species of Dryopteris to our fourteen, or its 207 species of Polystichum to our fifteen (we share four), to wonder how different things could have been. Our two regions share approximately 60 indigenous species but for the most part, these do not occur in the classic East Asia - Eastern North America pattern so well known in angiosperms. Instead, most of our shared species are circum-north temperate or circumboreal, e.g., several lycopods, eight horsetails, and such species as Cryptogramma stelleri, Woodsia ilvensis, W. glabella, Gymnocarpium jes- soense, and Dryopteris fragrans. The volume has only short prefatory remarks, without discussions of geography, climate, or floristic associations. The heading of the main text is labeled “Pteridophytes (Lycophytes and Ferns)” which, for ferns, transitions away from the recently popular “monilophytes.” Families are presented in taxonomic order, based on the linear classification of Christenhusz et al. (2011), itself based largely upon Smith et al., (2006). Following Christenhusz et al., Equisetaceae are treated first (vs. Ophioglossaceae in the scheme of Smith et al. 2006), as sister to the rest of the ferns. Treatments are in standard format and include dichotomous keys and full descriptions of families, genera, and species. Subspecies, varieties and, in a very few instances, forms are treated where authors deem necessary. Hybrids are mentioned, at least in some genera, but are not treated in depth. Keys and descriptions show the high standards that one anticipates from previous volumes. The descriptions are extremely detailed, most involving 75-100 morphological characters and often running from 25 to 35 or even 40 lines, although the treatments of Hymenophyllaceae are notably shorter. Short or full synonymies are given, but types are not reported. Each species description is followed by a statement of habitat, a list of provinces (including Taiwan and islands of the South China Sea) where it occurs, and a statement of worldwide range. Following that are notes about relationships, variability, medicinal uses (if any), and nomenclatural problems. These notes offer insight into historical and current taxonomic concepts, problems in the flora, and often additional comparative details of identification. They will be of great interest to current and future pteridologists as our knowledge improves and will no doubt prompt studies for many years to come. Inevitably, some families (e.g., Ophioglossaceae) lack recent research focused on China, while others (e.g., Aspleniaceae) benefit from recent and ongoing scrutiny, so the treatments are slightly uneven by subject, but they (2014) Erratum AFJ volume 103 issue 2, pp. 112-117 (April - June 2013) Figure 1 in the American Fern Journal article entitled Myriopteris windhamii sp. Nov., a New Name For Cheilanthes villosa (Pteridaceae) by Amanda L. Grusz was originally published incorrectly. The correct figure is shown below. Fc.l. American Fern Journal 103(4):252-254 | Erratum AFJ volume 103 issue 3, pp. 175-181 (July-September 2013) Figures 1 and 2 in the American Fern Journal article entitled A New Hybrid of Serpocaulon (Polypodiaceae) from Costa Rica by Alexander Fco. Rojas- Alvarado and Jose Miguel Chaves-Fallas were originally published incorrectly. The correct figures are shown on the next two pages. ERRATUM 254 AMERICAN FERN JOURNAL: VOLUME 103 NUMBER ■ for 2013 Table of Contents for Volume (A list of articles arranged alphabetically by author) Abeu, T., S. Orsenigo, N. M. G. Ardenghi, E. C. H. E. T. Lucassen, and A. J. P. Smolders. Hydrochemical Characterization of a Stand of the Threatened Endemic Isoetes malinvemiana 241 Aguraiuja, R. (See D. H. Lorence) 166 Arakaki, M. (See B. LeOn) 40 Ardenghi, N. M. G. (See T. Abeli) 241 Arunachalam, a. (See D. Balasubramanian) 49 Arunachalam, K. (See D. Balasubramanian) 49 Balasubramanian, D., K. Arunachalam, and A. Arunachalam. Occurrence of Critically Endangered Pteridophyte Helminthostachys zeylanica (L.) Hook, in Burachapori Wildlife Sanctuary, Northeast India 49 Barrington, D. S. (See J. P. S. Condack) 118 Batke, Sven Peter, and Nicholas Hill. First Record of Serpocaulon lasiopus (Polypodiaceae) from Mesoamerica 182 Bogonovich, Marc, Scott Robeson, and Maxine Watson. Patterns of North American Fern and Lycophyle Richness at Three Taxonomic Levels 193 Chaves-F ALLAS, J. M. (See A. F. Rojas-Alvarado) 175 Christenhusz, Maarten J. M., Mirkka Jones, and Samuli Lehtonen. Phylogenetic Placement of the Enigmatic Fern Genus Dracoglossum 131 Condack, JoAo Paulo S., Monique A. McHenry, Rita E. Morero, Lana S. Sylvestre, AND David S. Barrington. Polystichum montevidense Demystified: Molec- ular and Morphological Data Reveal a Cohesive, Widespread South American Species HB Fernandes, C. P. (See M. G. Santos) 215 Fernandez, H. (See E. L. Peredo) 27 Garrett, R. (See M. G. Santos) 215 Geiger, Jennifer M. O., Petra Korall, Tom A. Ranker, Annabelle C. Kleist, and Christine L. Nelson. Molecular Phylogenetic Relationships of Cibotium and Origin of the Hawaiian Endemics 141 Giannangeu, JuuAn Mostacero. The Identity of the Type of Polypodium gamerianum Vareschi (Polypodiaceae) 53 Gilman, Arthur V., and Michael A. Sundue. Flora of China, vols. 2-3, Lycopodiaceae through Polypodiaceae 245 Grusz, Amanda L. Myriopteris windhamii sp. nov., a New Name For Cheilanthes villosa (Pteridaceae) 112 256 Guo, Zm-You, and Hong-Mei Liu. Gametophyte Morphology and Development of Three Species of Cyrtogonellum Ching (Dryopteridaceae) He, H. (See L.-B. Zhang)- Hernandez, A. Marcela, Lucrecia TerAn, Marisol Mata, Olga G. MarUnez, and Jefferson Prado. Helical Cell Wall Thickenings in Root Cortical Cells of Polypodiaceae Species from Northwestern Argentina Hill, N. (See S. P. Batke) Hirai, R. Y. (See J. Prado) Jones, M. (See M. J. M. Christenhusz) Kelecom, a. (See M. G. Santos) Kleist, a. C. (See J. M. O. Geiger) Korall, P. (See J. M. O. Geiger) Kumari, a. (See S. G. E. Reddy) Lal, B. (See S. G. E. Reddy) Landoni, Gabriela Fausti, and Paulo G. Windisgh. Predation of Bracken Fern [Pteridium aquilinum (L.) Kuhn var. arachnoideum (Kaulf.) Brade) in Southern Brazil by Moths of the Genus Erupa Lehtonen, S. (See M. J. M. Christenhusz) Le6n, Blanca, Carl J. Rothfels, M6nica Arakaki, Kenneth R. Young, and Kathleen M. Pryer. Revealing a Cryptic Fern Distribution Through DNA Sequenc- ing: Pityrogramma trifoliata in the Western Andes of Peru Liu, H.-M. (See Z.-Y. Guo) Lobo, j. F. R. (See M. G. Santos) Lorence, David H., Kenneth R. Wood, and Ruth Aguraiuja. Taxonomic Reassessment and Conservation Status of Three Kaua’i Species of Aspleniiun in the LUellia Alliance Lucassen, E. C. H. E. T. (See T. Abeu) MartInez, O. G. (See M. A. HernAndez) Mata, M. (See M. A. HernAndez) McHenry, M. A. (See J. P. S. Condack) M^ndez-Couz, M. (See E. L. Peredo) Morero, R. E. (See J. P. S. Condack) Nelson, C. L. (See J. M. O. Geiger) Orsenigo, S. (See T. Abeu) Peredo, E. L, M. MfiNDEzAiiuz, M. A. Revdlla, andH. FernAneez. Mating System in Blechnum spicant and Dryopteris affinis ssp. affinis Correlates with Genetic Variability . . . 153 191 21 215 141 141 185 185 139 40 153 225 225 118 141 241 27 257 Prado, Jefferson, Eric Schuettpelz, Regina Y. Hirai, and Alan R. Smith. Pellaea flavescens, a Brazilian Endemic, is a Synonym of Old World Pellaea viridis Prado, J. (See M. A. HernAndez] Pryer, K. M. (See B. Le6n) Ranker, T. A. (See A. L. Vernon) Ranker, T. A. (See J. M. O. Geiger) Reddy, S. G. E., A. Kumari, andB. Lal. First Report of the Aphid, Amphorophora ampullata (Homoptera: Aphididae) on the Fern, Hypolepis polypodioides (Hypolepidaceae) from Western Himalayas (India) Revilla, M. A. (See E. L. Peredo) Robeson, S. (See M. Bogonovich) Rocha, L. (See M. G. Santos) Roe- Andersen, Susan M., and Darlene Southworth. Microsite Factors and Spore Dispersal Limit Obligate Mycorrhizal Fern Distribution: Habitat Islands of Botrychium pumicola (Ophioglossaceae) Rojas-Alvarado, Alexander Fco., andJos^ Miguel Chaves-Fallas. A New Hybrid of Serpocaulon (Polypodiaceae) from Costa Rica Rothfels, C. j. (See B. Le6n) Santos, Marcelo Guerra, Caio Pinho Fernandes, Luis Armando Candido Tietbohl, Rafael Garrett, Jonathas Felipe Revoredo Lobo, Alphonse Kelecom, and Leandro Rocha. Chemical Composition of Essential Oils from Two Fern Species of Anemia Schuettpelz, E. (See J. Prado) Smith, A. R. (See J. Prado) Smolders, A. J. P. (See T. Abeu) Southworth, D. (See S. M. Roe-Andersen) SuNDUE, M. A. (See A. V. Sundue) Sylvestre, L. S. (See J. P. S. Condack) TerAn, L. (See M. A. HernAndez) Tietbohl, L. A. C. (See M. G. Santos) Vernon, Amanda L., and Tom A. Ranker. Current Status of the Ferns and Lycophytes of the Hawaiian Islands Watson, M. (See M. Bogonovich) WiNDiscH, P. G. (See G. F. Landoni) Wood, K. R. (See D. H. Lorence) 258 40 59 185 27 193 215 1 175 40 215 21 241 245 118 225 215 59 139 166 Wyatt, Robert, and Graham E. Wyatt. Lygodium japonicum (Thunbei^) Swartz in the Piedmont of Georgia Wyatt, G. E. (See R. Wyatt) Yatskievych, George. Ferns of Southern Africa: A Comprehensive Guide Young, K. R. (See B. LeOn) Zhang, Li-Bing, and Hai He. Ferns and Fern Allies of Taiwan Volume 103, number 1, January-March, pages 1-58, issued 15 July 2013 Volume 103, number 2, April-June, pages 59-140, issued 05 November 2013 Volume 103, number 3, July-September, pages 141-192, issued 12 March 2014 Volume 103, number 4, October-December, pages 193-259, issued 17 April 2014 INFORMATION FOR PTERIDOLOGIA ISSUES IN PRINT e available for purchase: 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Choco (Part 1 : Psilotaceae through Dicksoniaceae). 364 pp. $32.00 plus postage and 3. Lellinger, David B. 2002. A Modem Multilingual Glossary for Taxonomic Pteri- dology. 263 pp. $28.00 plus postage and handling. 4. Hirai, Regina Y., and Jefferson Prado. 2012. Monograph of Moranopteris (Polypo- diaceae). 1 13 pp. $28.(X) plus postage and handling. For orders and more information, please contact our authorized agent for sales at: Mis.souri Botanical Garden Press, P.O. Box 299, St. Louis, MO 63166-0299, tel. 314-577- 9534 or 877-271-1930 (toll free). For online orders, visit: http://www.mbgpress.org. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fem Society welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns. SPORE EXCHANGE Mr. Brian S. Aikin, 3523 Federal Ave, Everett, WA 98201-4647 (spores.afs@comcast. net), is Director. Spores exchanged and lists of available spores sent on request, http:// amerfemsoc.org/sporexy.html GIFTS AND BEQUESTS Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Back issues of the Journal and cash or other gifts are always welcomed and are tax-deductible. Inquiries should be addressed to the Membership Secretary. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://amerfernsoc.org/