QE | A399 Volume 91 FERN nite RP L January—March 2001 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY The Group of Adiantum gracile in Brazil and Environs David B. Lellinger and Jefferson Prado 1 Dryopteris X correllii hyb. nov. (D. sepsis x goldiana), a Rare Woodfern Hybrid from Vermont Warren H. Wagner, Jr. and Arthur V. Gilman 9 The Relictual Fern Genus Loxsomop Marcus Lehnert, Maren Ménnich, Thekla Pia Alexander Schmidt-Lebuhn, and Michael Kessler 13 Growth, Leaf Cl teristics, and Spore Prod in Native and Invasive Tree Ferns Hawaii Leilani Z. ova and Guillermo Goldstein 25 Shorter Notes Binomial for Dryopteris clintoniana X goldiana James H. Peck 36 Cryopreservation of Shoot Tips of Selaginella uncinata Valerie C. Pence 37 The American Fern Society Council for 2001 BARBARA JOE HOSHIZAKI, 557 N. Westmoreland Ave., Los Angeles, CA 90004-2210. President CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS 66045-2016. Vice-President W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI 53233-1478. ecretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-1110. Trea. DAVID B. LELLINGER, 326 West St. NW., Vienna, VA 22180-4151. robin Gece JAMES D. MONTGOMERY, Ecology III, R.D. 1, Box 1795, Berwick, PA 1860 RA ack Issues Curator GEORGE YATSKIEVYCH, Missouri Botanical Garden, P.O. Box 299, St. Louis, 0 63166-0299. Journal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166, Smithsonian Institution, Washington, DC 20560-0166. Memoir Editor CINDY JOHNSON-GROH, Dept. of Biology, Gustavus Adolphus College, 800 W. College Ave., St. Peter, MN 56082-1498. Bulletin Editor American Fern Journal EDITOR ' R. JAMES HICKEY y Departme ie Geet OH 45056 ph. (513) oa e-mail: hickeyrj @ muohio.edu ASSOCIATE EDITORS GERALD I. GASTONY ..3. i050 Dept. of Biology, Indiana egies Bloomington, IN 47405-6801 CHRISTOPHER H. HAUFLER .... Dept. of Botany, University of Kansas, Lawrence, 66045-2106 ROBBIN C. MORAN New York Botanical Garden, Bronx, NY 10458-5126 JAMES H. PECK Dept. of Biology, University of Ar Arkansas—Little Rock, 801 S. University Ave., Little Rock, AR 72204 “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the Sosa seat of ferns. ii is owned by the American Fern Society, and ares at 326 West St. NW., VA 22180-415 . Periodicals postage paid at Vienna, VA, and additional entry. aims adie’ - missing issues, made 6 nestic) to 12 sip pies ae after the date of issue, r back issues should be addressed to Dr. James D. Montgomery, Ecology III, R.D. 1, wrest Sg ay 18603-9801. Changes of address, dues, and applications for membership should be sent to the Membership General inquiries concerning fe the Secretary. Subscriptions $20.00 gross, $19.50 net if paid thro sol an Bg bs eg fee $0.50); sent gti - members of the Somer lage! en (annual dues, $15.00 + feng ailing surcharge beyo U.S.A.; life membershi + $140.00 raed a beyon Back volumes are cachet a most years as issues or on ce Please contact the Back Issues Curator for prices and availability. STMASTER: Send address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA 22180-4151. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society wel contributions from members and non-members, including miscellaneous notes, spi to — or purchase materials, personalia, horticultural notes, and reviews of non-technical books on fi SPORE EXCHANGE ae Stephen McDaniel, 1716 Piermont Dr., Hacienda Hts., CA 91745-3678, is Director. Spores changed and lists of available spores sent on request. _http://www.amerfernsoc -org/sporexy.html to m b and to others interested sifts qi the ety in ferns. Back issues of the Journal and cash or other gift J] it d are tax-deductible. MISSOURI BOTANICAL JUL 1 2 2001 GARDEN LIBRARY American Fern Journal 91(1):1—8 (2001) The Group of Adiantum gracil in Brazil and Environs DAvip B. LELLINGER Department of Botany, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0166 JEFFERSON PRADO Segao de Briologia e — Instituto de Botanica, Caixa Postal 4005, 061-970 Sao Paulo, SP, Brasil ABSTRACT.—The Adiantum gracile group of Brazil and adjacent Bolivia is a natural group distin- guishable from A. tetraphyllum and related species. We provide a key to the group a describe two new Brazilian species belonging to the group, A. cinnamomeum and A. dawson Adiantum is a large and widespread genus in the tropics, numbering ap- proximately 200 species in the Neotropics, 65 to 70 of them growing in South America. Brazil has about 50 species, about 75% of the total known for the continent. No recent monograph or revision exists for Adiantum in the Neotropics, nor has a comprehensive subgeneric classification of the genus ever been devel- oped. In preparing keys for floristic works, most authors have used as principal characters lamina venation (anastomosing vs. free) and shape of the ultimate divisions (roughly rounded and not dimidiate vs. more or less oblong and usually dimidiate). Secondary characters include lamina division (simple, pin- nate, bipinnate, or decompound), pinnule attachment (sessile vs. pedicellate), ultimate division articulation (articulate or not), and lamina apex shape (con- form vs. gradually attenuate). The species with bipinnate laminae, a conform terminal pinna, and oblong and usually dimidiate pinnules may be placed together conveniently as the A. tetraphyllum group. Delimited in this broad manner, the group includes about a third of the Neotropical species of Adiantum. Most, if not all, of the species of this group lack evenly and finely serrate sterile lamina margins that seem to predominate in other groups. Within the A. tetraphyllum Willd. group sensu Jato, four groups can be fairly readily distinguished, principally on the basis of their rachis and lamina in- dument. These groups are not yet fully characterized, nor are their constituent species known with certainty due to several names for which we have yet seen any specimens. Three of these groups are large and widespread in the Neo- tropics, but one small group is nearly restricted to Brazil: A. gracile Fée and two new species. The species of the A. gracile group differ from the other three groups in having more pinna pairs (5-7), more pinnules on a pinna (35—40 pairs), and pinnules that often are approximate or nearly so, narrower (2—4 mm wide) but 2 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) relatively longer, usually at least slightly lobed, and in lacking an obvious midrib, although they could be obscurely dimidiate. In addition, the larger scales on the abaxial surface of the pinnules, when present, are several cells wide and long-ciliate throughout, much like those of the rachises and costae. These scales are in contrast to those of A. tetraphyllum and its other allies, which are narrowed and sometimes even uniseriate distally. In the A. gracile group, the rachises are conspicuously covered with cinnamomeous, linear- lanceolate scales with long-ciliate margins and base, at least when young; these scales are most persistent in A. cinnamomeum. KEY TO THE SPECIES OF THE ADIANTUM GRACILE GROUP 1. Pinnules abaxially glaucous, the scales few or none, but rather inconspicuous, yellow- ish, sessile glands often present; pseudoindusia erose-ciliate at maturity, adaxially gla- brous. Rhizome scales narrowly lanceate, dark to medium brown, sparingly denticulate. Medial pinnules ca. 2 times longer than wide, usually acute at the sterile apex, nearl wounistl ot tee Fertile ays ee oe ed iy kag 3. A. dawsonii 1. Pinnules abaxially not glaucous or glandular, always sparsely scaly; pseudoindusia entire to erose, scarcely ciliate, adaxially glabrous or bearing toothed scales ........ 2(1). Medial pinnules ca. 2 times longer than wide, obtuse to acute at the sterile apex, angular at the fertile apex; pseudoindusia glabrous; rhizome scales medium to Ct AUTEN, WORE OS ee ee oe ya te ec ess a Fy 1. A. gracile 2(1). Medial pinnules ca. 3 times longer than wide, acuminate to acute at the always sterile apex; pseudoindusia bearing hairlike scales that are toothed at the base; thizome scales dark brown to blackish, the margins sometimes slightly paler, alwys sparsely dentioniate cc. 3. 5 Ske Ww eww ok 2, A. cinnamomeum 1. Adiantum gracile Fée, Gen. Fil. 116. 1852. Fig. 1A, B. Plants terrestrial. Rhizomes stout, short-creeping, ca. 4-5 mm in diam., scaly, the scales somewhat shiny, essentially concolorous, medium to dark brown, narrowly lanceate, entire. Fronds monomorphic, 2-pinnate, (30)60-100 cm long, the laminae (7)20—30 cm wide; stipes approximate, black, adaxially sul- cate, scaly, the scales appressed throughout or sometimes distally patent, con- colorous, cinnamomeous, 2~3 mm long, narrowly lanceate with a filiform apex, strongly denticulate proximally; rachises similar to the stipes and their indument similar; pinnae oblong-lanceate, slightly decreasing at the base, ta- pering at the apex, (7)15—22 cm long, 1.5—2.2 cm wide, the lateral pinnae (3)7- 10 pairs, patent, alternate, the terminal pinna conform, 1—1.5 times longer than the subtending pinnae 0.67—1 times as long as the medial pinnae; indument of the costae like that of the stipes and rachises; pinnules 27—43 pairs, ca. 2 times longer than wide, chartaceous, free-veined, without an evident midrib, the proximal pairs reduced, somewhat rounded or triangular, the medial pairs dimidiate, oblong to somewhat tapering, the acroscopic base truncate, the ster- ile apex obtuse to acute, the sterile margin irregularly and distantly serrate, except at the pinnule apex, the fertile apex angular, the distal pinnules ca. 1/2 as long as the medial pinnules, the adaxial surface of the pinnules glabrous or sparsely scaly near the sori, the veins slightly prominulous, the idioblasts LELLINGER & PRADO: ADIANTUM GRACILE GROUP 3 Acc¥ Spot Magn = wo 69. Roo TF a2 AccV¥ SpotMagn Det WD Ep F- 1WOkKV40 3027 105 6 IBUA. grac WOkV 40 3112x E 1069 IBA. grac Spot Magn Det WD Exp -————————1_ 50 um Spot Magn a wD Ep -— § 1366x SE 16.0 6 IBUA. cinnam. 10 OnvV40 1316x 160 3 IBA. cinnam 100 jm Spot Magn Det WD a ————————|, 2 Acc V agn «= Det WD Ep hooKy 40 2636x SE 105 (BUA. daws 100kV 40 7x SE 1054 BUA. daws Fic. 1. Scanning electron micrographs of Adiantum. Fics. 1A—B. Proximal eh mRaie views of A. gracile spores (Salino 393, UEC). Fics, 1C-D. Overall and distal v A a ype US). Fics. 1E-F. Suprabasal nee on ofa ance scale and an pseudoindusial scale of A. cinnamomeum (MacFarland et al. 283, holotype US). G-H. Proximal and distal views of A. dawsonii spores (Dawson 14868, holotype US). 4 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) inconspicuous or nearly so, the abaxial surface of the pinnules glabrescent, with patent, sparse, toothed scales ca. 0.5 mm long but otherwise similar to those of the stipes, glands absent, the veins slightly prominulous, the idioblasts conspicuous; sori oblong, up to 7 or 8 per pinnule; pseudoindusia glabrous, entire to erose; spores trilete, 25-30 4m in equatorial diameter, tetrahedral- globose without prolonged angles, the surface rugulate with a thin, fragment- ing outer layer, and so sometimes appearing finely cristate. TYPE: Brazil: sin. loc., Claussen s.n. (P, Morton photo 2596, photos GH, SP; possible isotype US1148302 [Claussen 177 ex C]). DISTRIBUTION: Endemic to central Brazil (Mato Grosso, Distrito Federal, and Goias). HABITAT: Slopes and stream banks in cerrado and campo vegetation, at ca. 50-850 m elevation. SPECIMENS EXAMINED: Brazit: Mato Grosso: Sararé, RADAMBRASIL, Serra de Pedra (S. Aguapei), 59°25'W, 16°10'S, Pires & Santos 16572 (UEC); Chapada dos Guimaraes, Véu-da-Noiva, Verardo 23668 (SP, UEC), same locality, Salino 393, 500 m (UEC), same locality, Hatschbach 37628 (MBM, UC). Goids: Ca. 1 km S of S. Joao de Alianga, ca. 850 m, Irwin et al. 31972 (NY not seen, SP, US); Serra do Caiap6, 50 km S of Caiaponia, Prance & Silva 59588 (GH, K, NY not seen, US). Distrito Federal: Brasilia, R. Windisch & Ghilldny 253 (HB). 2. Adiantum cinnamomeum Lellinger & Prado, sp. nov. Figs. 1C-F, 2. A specie A. gracili Fée paleis rhizomatis integris (vs. sparse denticulatis), pinnis 5—7 (vs. 8-10) jugatis patulis, pinnulis oblongis acuminatis, 3-plo (vs. 2-plo) longioribus quam latioribus differt. TYPE: Brazil: Rond6nia: 120 km SW of Pérto Velho-Highway BR 364, 15 km W of Mibrasa, 9°10’ S, 63°07’W, 29 May 1982, McFarland et al. 283 (holotype US photo SP; isotypes BM, GH). Plants terrestrial. Rhizomes stout, short-creeping, ca. 3 mm in diam. exclud- ing the protuberances, scaly, the scales shiny, slightly bicolorous, dark brown to blackish, narrowly lanceate, the margins lighter and sparsely denticulate. Fronds erect, monomorphic, 2-pinnate, 35-65 cm long, the laminae 20-25 cm wide; stipes ca. 5 mm distant, black, adaxially sulcate, scaly, the scales ap- pressed throughout or distally patent, concolorous, cinnamomeous, 4—5 mm long, linear-lanceate with a filiform apex, sometimes long-ciliate at the base, medially copiously toothed-ciliate; rachises more densely covered by scales than the stipes, the indument similar to that of the stipes; pinnae oblong- lanceolate, slightly decreasing at the base, tapering at the apex, 8-19 cm long, 2.0—3.5 cm wide, the lateral pinnae 5-7 pairs, patent, alternate, the terminal pinnae conform, 1—1.25 times longer than the subtending pinnae, ca. 1.33 times longer and wider than the medial pinnae, the indument of the costae = Fic. 2. Adiantum cinnamomeum. Fic. 2A. Frond (Steward et al. P20394, US). Fic. 2B. Long- creeping rhizome (MacFarland et al. 283, US). Fic. 2C. Rachis scales (MacFarland et al. 283, US). Fics. 2D-E. Abaxial surface of pinna and pseudoindusia showing scales (Irwin et al. 10180, US). LELLINGER & PRADO: ADIANTUM GRACILE GROUP SAAN aM ~*~. 2% a Wy ) Wh lee mil Wh ESS V/ 4 f\\ AS VES LS Ns \\ « WS . = CRS GN E Fe ae a Se Se Sl g a \\ SA Sif a SR 6 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) similar to that of the rachises; pinnules 35—40 pairs, ca. 3 times longer than wide, chartaceous, free-veined, without an evident midrib, the proximal pairs reduced, flabellate to subrhombic, the medial pairs dimidiate, oblong to some- what tapering, 0.5—-2 cm long, 2-4 mm wide, subarticulate (the dark color of the stalk terminating almost at the base of the pinnule), patent to erect, the acroscopic base truncate, roundish to straight, the apex subacute to acute, the sterile margins serrate to biserrate, except the basioscopic margin entire to the middle of the pinnule and dentate towards the apex, the distal pinnules ca. 1/4—1/5 as long as the medial pinnules, the adaxial surface of the pinnules glabrous, the veins not prominulous, the idioblasts conspicuous, the abaxial surface glabrescent, with sparse, cinnamomeous scales, the scales filiform, ca. 1 mm long, pectinate at the base, glands absent, the veins prominulous, the idioblasts inconspicuous; sori oblong, 7 or 8 per pinnule; pseudoindusia with filiform scales, bearing short basal processes, the margins entire to erose; spores trilete, 25-30 ym in equatorial diameter, tetrahedral-globose with pro- longed angles, the surface sparsely rugulate with a thin, fragmenting outer layer, and so sometimes appearing finely and sparsely cristate. DISTRIBUTION: North-central Brazil (Rondénia, Amazonas, Pard, Mato Grosso, Distrito Federal). HABITAT: On clay soil, in open canopy and dense understory, at 50-1000 m elevation. PARATYPES.—BRAZIL: Rondénia: Pérto Velho to Cuiabé Highway, vicinity of Santa Barbara, 15 km E of Km. 1117, Prance & Ramos 7157 (NY, UC). Amazonas: Manaus-Caracaraf Highway, Km. 160, Steward et al. P20394 (BM, MO, NY, US); Manaus-Caracarai Highway, Km. 185, Prance et al. 22691 (MO, NY, US); Rio Javari between Estirao do Equador and Rio Javarizinho, Prance et al. 24046 (NY, UC); Manaus, Rio Araras, SIDERAMA, Loueiro et al. s.n. (GH); Manaus, Pivetta 254 (HRCB); Barra, Prov. Rio Negro, Spruce s.n. (GH, US). Para: Municipio de Oriximind, Rio Trombetas, es- trada da Mineragao Santa Patricia, ramal 22, Cid et al. 1441 (NY, US); Rio Jamunda, Municipio de Faro, Sao Jorge, Black & Ledroux 50-10761 (HB). Mato Grosso: Vila Bela da Santissima Trindade, Faz. Cabixi, junto ao Rio Cabixi, 13°S, 60°10’W, ca. 12 km da divisa com Rondénia, Prado & Salino 8 (HB, SPF, UEC). Distrito Federal: Parque Municipal do Gama, Irwin & Soderstrom 5884 (NY, US); Irwin et al. 10180 (GH, K, MO, NY, US); Brasflia, Reserva Ecolégica do IBGE, Heringer et al. 3767 (MBM). Adiantum cinnamomeum is distinguished by its long, narrow pinnules that are acuminate to acute at the apex, presumably because the sori are always lateral. Adiantum cinnamomeum has been found only in Brazil, but may have a wider range in Amazonian South America. 3. Adiantum dawsonii Lellinger & Prado, sp. nov. Fig. 1G-—H, 3. A specie A. gracili Fée pinnulis adaxialiter glaucis, paleis paucis vel nullis, glandulis rotundis sessilibus luteolis differt. TYPE: Brazil: Goids: Southern Serra Dourada region, 20 km E of Formoso, 48°50'W, 13°45’S, on banks and margins of small stream running through hilly cerrado, 16 May 1956, E. Y. Dawson 14868 (US, 2 sheets). Plants terrestrial. Rhizomes stout, short-creeping, ca. 4-5 mm in diam., scaly, LELLINGER & PRADO: ADIANTUM GRACILE GROUP a” Gh ces BASS igh LE wdiiiler EASES SW jugs We MEE RR® avy MM le f x AAA! SESS s RASS SPLIT J 9 9 SSSidonoy Sadly lof A ws Phd iron = WYER TRE SO) OT T T T T T T 7. T 4.0 + 4 3.5} 4 3.0 + 4 q 2.5} ? & 20+ 4 e3 o Z 15+ 4 1.0} - 0.5+ 0.0 + b < 400} ; a) S 5 300 f ; a os 5 o 200+ 4 3 3 ‘s 100} : —] 0 Time (months) Fic. 2. Number of new fronds produced each month (a) and total surface area per plant (b), from November 1997 through October 1998. Symbols are mean + SE (n = 6 to 8). Symbols as in Fig. 1. Absence of error bar indicates that the SE is smaller than the symbol. number of fertile fronds produced per month, was between 20 and 50 percent higher in the native tree ferns (Table 2). DISCUSSION Native tree ferns on Oahu (this study) and the island of Hawaii (Walker and Aplet 1994, Wick and Hashimoto 1971) had relatively low annual height in- creases compared to the invasive tree fern. The Australian tree fern grew very little in January through March, when monthly precipitation was below 20 cm. Growth increased in April through August, apparently as a result of higher precipitation. Growth in the native tree ferns at the low elevation site also increased in April. At the high elevation site we noted that growth showed DURAND & GOLDSTEIN: NATIVE AND INVASIVE TREE FERNS 31 TABLE 2. Leaf mass per area (LMA), leaf life span, standard error (SE) of standardized leaf production de and of standardized s pore production for native and invasive tree ferns at three elevations. Standard error ard error of standardized al variability in spore production. Values are means + SE. Symbo (7) next to a grouping indicates statistical significance (P = 0.01) between the invasive and native ferns Pp within that grou Leaf life span SE of leaf SE of spore LMA (g m~”) (months) production production Low elevation Invader S. cooperi Be pena 6.0 O:15T 335 1.97 Native C. chamissoi 54.7 + 2.8 11.0 + 0.30 8.49 321 Mid elevation Natives C. chamissoi 509 + 3.1 12.0 + 0.71 11.7 3.66 C. menziesii 78.9 + 4.0 10.8 + 1.1 8.32 2.45 High elevation Invader S. cooperi 34.7 + LBF n/a n/a Natives C. chamissoi 1203) 7.2 10.6 + 0.42 9.02 3.46 C. menziesii 1275 = 38 12.0 + 0.50 8.72 2.63 C. glaucum 153.2 223:0 120 :220.73 8.94 4.04 less seasonal variation and may have been influenced by higher irradiance during drier months, when cloud cover is relatively low, as well as by rainfall atterns. Frond production of the native tree ferns also exhibited seasonal var- iation, with the majority of frond production occurring from February through April. There was virtually no frond production during the months of October, November, and December. This seasonal pattern was similar to the pattern of frond production observed in Cibotium on the island of Hawaii by Wick and Hashimoto (1971). In neither study did the pattern of frond production appear to correspond with precipitation. In contrast to the native tree ferns, frond production in the Australian tree fern was consistent throughout the year. The faster growth rate and greater and more consistent frond production of the Australian tree fern S. cooperi could contribute to its ability to outcompete the native Hawaiian tree ferns. When plants of the same growth habit are compared, plants with a higher leaf surface area tend to have a higher relative growth rate (Lambers et al. 1998). Sphaeropteris cooperi had over three times more fronds per plant than the native tree ferns, and over four times more leaf surface area (Fig. 2). A greater leaf surface area per plant, along with a greater annual height increase, could allow the Australian tree fern to intercept more light than the native tree ferns, and thus potentially fix more carbon. In a previous study, we ob- served that the photosynthetic rate at light saturation for shaded plants was between 5 and 7 pmol m~ s~ for the invasive tree fern, while the photosyn- thetic rate at similar light levels was between 3.4 and 4 pmol m~ s* for the native tree ferns (Durand and Goldstein 2001). The pattern of allocation of carbon to different plant organs can also affect the annual height increase of the plant. The native tree ferns have a high starch content in their trunk, an were once harvested for this starch (Ripperton 1924, Wick and Hashimoto 32 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) 16+ 14+ 5 ae 127 D & 10} n = oS Sr C= o = 6 a 4f 2+ $4 : L O Y y. = Z a aoe 1 | ==. —_: re 5s Re eM A UME A dA LS oa Time (months) Fic. 3. The number of fertile fronds (with spores) per plant of native and invasive tree ferns, measured from November 1997 through October 1998. Symbols represent + SE (n = 6 to 8). Symbols as in Fig. 1. 1971). While no data is available on the trunk starch content of S. cooperi, in areas where S. cooperi and C. chamissoi grow together, wild pigs preferentially feed on the trunk starch of the native tree ferns. Thus, another possible expla- nation for the low annual height increase of the native tree ferns is a relatively large allocation of carbon, in the form of starch, to the fern’s trunk. The native tree ferns exhibited seasonal changes in fertility, with the greatest number of fertile fronds present in the months of September through February. The invasive tree fern did not exhibit as much seasonal variation in fertility, though there were more fertile fronds per plant in the months of July through November. The difference in seasonal spore production may allow spores of the Australian tree fern to germinate at a time when spores of the native tree ferns are not available. It has been suggested that successful invasive species either utilize resources more efficiently, or at a time when natives are inactive or unable to access the resources (Vitousek 1986). Consistent monthly spore production could give the Australian tree fern a reproductive advantage over the native tree ferns. Ferns in the genus Sphaeropteris produce 64 spores per sporangium (Gas- tony 1974), as do ferns in the genus Cibotium (Sporne 1966). Dyer (1979) es- timated that an average fertile frond of Cibotium chamissoi produces 700,000,000 spores. With between 1 and 4 fertile fronds produced per plant per year (Fig. 3), spore production per year ranges between 700 million and 2.8 billion spores. Sphaeropteris cooperi produced between 22 and 27 fertile fronds per plant per year (Fig. 3), though the number of sporangia produced per frond is not known. However, this species has a greater frond surface area than Cibotium, and given the equal number of spores per sporangia in Cibo- tium and Sphaeropteris, assuming equal spore production per frond seems a DURAND & GOLDSTEIN: NATIVE AND INVASIVE TREE FERNS 33 conservative estimate. If S. cooperi does indeed produce an equal number of spores per frond as Cibotium chamissoi, this fern has the potential to produce 15.4 billion to 18.9 billion spores per year. While both species produce a large number of spores each year, S. cooperi can potentially produce 13 to 16 billion more spores per year. Studies of the spore production of the invasive tree fern, and spore viability of both the native and invasive tree ferns, need to be con- ducted in order to have a more complete understanding of the relative repro- ductive capacity of these ferns. Variation in leaf life span has been found to be an important predictor of numerous plant responses (Chabot and Hicks 1982, Reich et al. 1992, Reich et al. 1997). The leaf life span of higher plants can range from less than one month to 25 years (Chabot and Hicks 1982). When compared across diverse ecosystems and biomes, a short leaf life span is associated with factors such as higher photosynthetic rate, higher leaf nitrogen content, lower leaf construc- tion cost, and a lower LMA (Chabot and Hicks 1982, Reich et al. 1992, Reich 1993, Reich et al. 1997). The mean leaf life span of the Australian tree fern was significantly shorter than the mean leaf life span of the native tree ferns (Table 2). As expected for a plant with a shorter leaf life span, the Australian tree fern had a significantly lower LMA than the native tree ferns (Table 2). Generally, longer-lived leaves require more secondary compounds to protect the leaves against herbivores and pathogens, whereas shorter-lived leaves with a low LMA are able to maximize photosynthetic return per unit weight of the leaves (Lambers et al. 1998, Reich 1993). The Australian tree fern produced short-lived leaves with high leaf nitrogen and high photosynthetic capacity (Durand and Goldstein 2001) at a relatively low carbon cost to the plant com- pared to the native tree ferns. In a recent study of 34 native and 30 invasive higher plants in the Hawaiian Islands, it was found that the latter had lower LMA, higher leaf nitrogen, and higher photosynthetic rates (Baruch and Gold- stein 1999). The patterns observed with the ferns in this study were consistent with the pattern found in higher plants, suggesting that this suite of traits are important determinants of the success of invasive species in the Hawaiian Islands, for both ferns and higher plants alike. The difference in leaf life span, LMA, growth rate, and spore production between the invasive and native tree ferns suggests that there are differences in life history characteristics. Tree ferns in Costa Rica that grew as pioneer species in secondary forest had higher leaf turn-over rates, faster growth rates, and more rapid production of sori than tree ferns that grew in primary forest (Bittner and Breckle 1995). Similarly, the high leaf turn-over rate, high spore production, and rapid growth rate in S. cooperi suggests that it tends to behave as a pioneer, or r selected species. The native tree ferns, on the other hand, had a much lower leaf turnover rate and slower growth rates, as well as lower reproductive output. This difference in life history characteristics may par- tially explain why S. cooperi can become established and spread quickly in Hawaiian rainforests, particularly in areas of disturbance. 34 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) ACKNOWLEDGEMENTS We would like to thank the B. Krauss Fellowship in Botany and C. Smith and the — Research Program at the University ae Baweld for Sy hee funds for yee research. We also like to thank the Oahu Natural A sistance in acai Mt. Kaala every month for a year. We also gratefully acknowledge the a Souza Center at Kokee State Park and Lyon Arboretum for their support REFERENCES Barucu, Z., and G. GOLDSTEIN. 1999. Leaf construction cost, nutrient ae and net CO, assi = paige: of native and invasive species in Hawaii. Oecologia 121(2):183—192. BECKER, R. E. 1976. The phytosociological position of tree ferns (Cibotium spp. 5) in the montane rea on the Island of Hawaii. Ph.D. Dissertation. University of Hawaii-Manoa, Hono- lulu, HI BITTNER, J., and S. W. BRECKLE. rele a growth rate and age of tree fern trunks in relation to habitats. sg Fern J. 85(2):3 Buck, M. G. 1982. Hawaiian ie ‘pian affects forest regeneration and plant succession. Research note PWS-355. Pacific Southwest Forest and Range Experiment Station, United States Department of Agriculture Forest Service, Berkeley, California. CHABOT, B. F., and D. J. Hicks. 1982. The ecology of leaf life spans. Annu. Rev. of Ecol. and System. DeBacu, P., and D. ROSEN. 1991. Biological control by natural enemies. Cambridge University Press, Cambridge U DuRAND, L. Z., and G. GoLp. LDSTEIN. 2001. Photosynthesis, photoinhibition, ae earacane use effi- ciency in native and invasive tree ferns in Hawaii. Oecologia 126:345—3 k, Gastony, G. J. 1974. Spore morphology in the Cyatheaceae. I. The perine and sporangial capacity: general considerations. Amer. J Bot. 61(6):672-680. ERS, OORTER, and M. I. vAN VUREN. 1998. Inherent Variation in Plant Growth: Phys- iological Mechanisms and ea a Backhuys Publishers, The Netherlands. dy es aad) D: R-DOMBOIS. 1989. Ch aus ual invaded islands, with special fee to Hawaii. Pp. 257-280 A A. Drake, H. A. Mooney, F. di Castri, R. H. Groves, F. t, M Rejmanek and M. Williamson, eds. Biological Invasions: A Global Perspective. John Wiley & Sons, New York. LoopE, L. L., R. J. NAGATA, and A. C. MEDEIROS. 1992. Alien plants in Haleakala National Park. Pp. 551-576 in C. P. Stone, C. W. Smith, and J. T. Tunison, eds. Alien plant invasion in native ecosystems of Hawaii: Management and Research. Cooperative National Park Resources Study Unit, ee “ Hawaii ap oo ai MEDEIROS, A. C., L. L. LOopE, S. J. ANDERSON. 1993. D tial colonizati ti roaster spp.) and alien ‘eyed pana tree ferns in a Hawaiian 1 rain forest. Selbyana —74. a A e S. and G. P. McCase. 1993. Introduction to the practice of statistics. W. H. Freeman and Company, New York. PaAtMer, D. D., 1994. The Hawaiian species of Cibotium. Amer. Fern J. 84 (3):73-85. REICH, P. B. 1993. Reconciling apparent discrepancies among studies relating life span, structure and function of leaves in contrasting plant life forms and climates: ‘the blind man and the elephant retold’. Funct. Ecol. 7:721—725 REICH, P. B., M. B. WALTERS, and D. S. ELLSWORTH. 1992. Leaf life-span in in to ms plant, and stand characteristics among diverse ecosystems. _ Monogr. 62(3):365-39 REICH, P. B., M. B. WALTERS, and D. S. ELLSWORTH nies m tropics to tundra: global conver- gence in plant functioning. Ecology 94: 13730-1373 RIPPERTON, J. D. oe ae Hawaiian tree fern as a ce source of starch. Hawaii Agri. Exp. SPORNE, K. R. 1966. as Morphology of Pteridophytes. Hutchinson & Co., Publishers, London, U.K. DURAND & GOLDSTEIN: NATIVE AND INVASIVE TREE FERNS 35 STRATTON, L., G. GOLDSTEIN, and F. C. MEINZER. 2000. Temporal and spatial partitioning of water resources among eight woody species in a Hawaiian is ‘ates Pet a 124(3):309-317. VITOUSEK, P. M. 1986. Biological invasions and ecosystem properties: can species make a differ- ? in H. A. Mooney and J. A. Drake, eds. Ecology of Biological Invasions of North America and Hawaii. Springer-Verlag, New York. VITOUSEK P. M., G. GERRISH, D, R. TURNER, L. R. WALKER, and D, MUELLER-DOMBOIS. 1995. Litterfall and nutrient cycling in four Setemess montane rainforests. J. Trop. Ecol. 11:189-203 ViTousEK, P. M., L. L. Loope, and C. P. STONE. 1987. Introduced species in Hawaii: biological effects and ee for ecological research. TREE 2:224-227 WALKER, L. R. and G. H. APLET, 1994. Growth and fertilization responses of Hawaiian tree ferns. Biotropica ig 378-383. Wacner, W. H. 1995. Evolution of “isileieaad ferns and fern allies in relation to their conservation status. one ms 49(1):31-4 Wick, H. L., G. T. HASHIMOTO. “isl Frond development and stem growth of treefern in Hawaii. U.S. Forest Service Research Note PW-237, Pacific Southwest Forest and Range Experiment Station. Berkeley, CA. American Fern Journal 91(1):36—40 (2001) SHORTER NOTE(s) Binomial for Dryopteris clintoniana x goldiana.—One strikingly handsome Dryopteris hybrid, the largest of all North American temperate Dryopteris (Thorne, F. and L. Thorne. 1989. Henry Potter’s Field Guide to the Hybrid Ferns of the Northeast. Vermont Institute of Natural Science, Woodstock, VT) and an excellent fern for hardy gardening (Mickel, J. T. 1994. Ferns for American Gardens. Macmillan, New York, NY), is without a collective epithet: D. clin- toniana X goldiana. Dryopteris Xmickelii Peck, hyb. nov—TYPE: UNITED STATES, New Jersey, Sussex Co., west side of extensive swamp south of Big Spring, 3 km south of Springdale, 17 Oct. 1969, James D. Montgomery 90895n (NY). Planta hybrida inter Dryopterem clintonianam et D. goldianam, aliis char- acteribus inter parentes media, sporis abortivis The hybrid is named to honor Dr. John ee Mickel (1934—) for his many endeavors on North American fern taxonomy, floristics, and horticulture, par- ticularly his promotion of gardening with hardy ferns. The hybrid is distin- guished readily from its parents by its abortive and irregular sized spores, intermediate frond morphology, and shared characters with both parents (see table in Thorne and Thorne, 1989, page 34). Dryopteis < mickelii has fronds up to 160 cm long, stipe to frond length ratio of 1:4, and is covered moderately with brown to dark brown scales. The blade is about 120 cm long, with a width to length ratio of about 2:5, with a very long and broad outline. The pinnae are regularly spaced, ascending relative to the rachis, with pinna shape wide, long rectangular and acuminate. Sori are borne close to the pinnule midrib. Dryopteris Xmickelii shares with D. clintoniana the following features: dark coloration at base of scales, relative length of blade, pinnae ascending from rachis, relatively narrow blade, and intermediate sorus location. D. x mickelii shares with D. goldiana the following features: wide blade, length of pinnules, falcate pinnules, relatively dark scales, and shape of pinnae. This hybrid occurs in rare and local populations in southern Ontario, south to New York, New Jersey, and Pennsylvania, and westward as outliers in Mich- igan and Ohio. This relatively narrow geographic range reflects the geographic overlap of the two parental ranges in the northeastern region of North America. The habitat of wetland woods and swamps were more common in early post- glacial times, but have declined since then, particularly in the western one- half of the range. In nature, the hybrid warrants conservation efforts wherever it still occurs. PARATYPES.—CANADA. Ontario: in woods at Ottawa, Scott s. n. (NY). UNITED STATES: Michigan: Washtenaw Co.: deep yellow birch swamp in Irwin’s woods, Wagner 9458 (US); Tuscola Co.: south of Murphy Lake, Wagner 63051 (US); tamarack swamps at Oxford, Farnwell 6117 (US). New Jersey: 37 Sussex Co.: swamp west of Springdale at Big Spring near Newton, Dowell 4929 (NY, US), Dowell 5033 (NY); Sussex Co.: roadside 1 mi south of Greendell, Edwards s. n. (NY); Bearfort Mtn., W. D. Miller 1648 (NY); West Englewood, Carhart 2b (NY) 2 sheets. New York: Green Lake, Jamesville, W. R. Dunlop s. n. (NY); Harris Swamp near Pilot Knob, Benedict s. n. (US-2202218); Kirkville, L. M. Underwood s. n. (NY); Staten Island: Arlington Station, Dowell 2801a & b (US). Ohio: Geauga Co.: Hopkins s. n. (NY). Pennsylvania: Berks Co.: swamp along spring-run 1.5 mi ne of Bernharts, Wherry s. n. (US-1849217); Delaware Co.: valley of Cruise Creek, Poyser 1286 (NY); Wakefield, Lanbor, J. J. Carter s. n. (NY). Vermont: N. Lynnfret (Limfret?), A. P & L. V. Morgan s. n. (US-154517). In the garden, its vegetative propagation should be promoted and supple- mented with modern tissue culture techniques to meet horticultural demands. It is a large and vigorous fern that forms extensive colonies through vegetative expansion while under cultivation. One specimen from New York was planted at the New York Botanical Garden, Bronx, NY in 1960 by members of the American Fern Society. Thirty-five years later, that plant had formed a clone 4 m across with over 200 apices (Mickel, 1994). JAMEs H. Peck, Department of Biology, University of Arkansas at Little Rock, 2801 S. University Ave., Little Rock, AR 72204. Cryopreservation of Shoot Tips of Selaginella uncinata.—Cryopreservation, or storage in liquid nitrogen (LN), has been successful for a wide variety of plant tissues, including shoot tips of in vitro cultures of higher plants (Bajaj, In Y. P. S. Bajaj, Cryopreservation of Plant Germplasm I, Biotechnology in Ag- riculture and Forestry 32. Springer, Berlin, 1995). At the temperature of LN, —196°C, long-term, stable storage of rare, endangered, or other valuable plant germplasm can be achieved. In an attempt to extend this technology to pteri- dophytes, LN storage of shoot tips of in vitro grown Selaginella uncinata was tested using the encapsulation dehydration procedure (Fabre and Dereuddre, Cryo-Letters 11:413—426, 1990). Shoot cultures of Selaginella uncinata (Desv. ex Poir.) Spring. were estab- lished from a plant purchased from Carolina Biological Supply Co. (voucher specimens deposited at the University of Cincinnati Herbarium, CINN, and at the CREW Herbarium). Tissues were surface disinfested in a 1:20 dilution of commercial sodium hypochlorite for five minutes, followed by two rinses in sterile, ultrapure water. The tissues were then transferred to a basal medium consisting of half-strength Murashige and Skoog salts with minimal organics (MS medium; Linsmaier and Skoog, Physiol. Plant. 18:100-127, 1965) with 1.5% sucrose and 0.22% Phytagel (Sigma Chemical Co.). Cultures were main- tained on this medium in 60 X 15 mm plastic petri plates, approximately 15 ml of medium/plate, at 26°C under Cool White fluorescent lights with a 16:8 hr, light:dark cycle. For preculture, one week prior to freezing the tissues were transferred to fresh basal medium or to basal medium plus 10 pM abscisic acid (ABA), which was filter sterilized and added to the medium after autoclaving. 38 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) AB Survival of shoot tips of S. uncinata through the steps of the encapsulation dehydration technique, 3—15 clean shoot tips per treatment. Encapsulated shoot tips were pretreated in 0.75 M sucrose, dried and exposed to LN, and samples were cultured after each step. Results are the mean of three experiments (+SE), except for long term LN exposure, for which there was one trial. Numbers of samples per trial are given in parentheses. Percent survival! 1 hr 3:5 NC Preculture medium Pretreated Dried LN exposure LN exposure Basal fe Seg a 29 = 6 Lice gt 0 127593513) C1233, 530) (8,:15, 11) (14) Basal + ABA 90 + 6 92 ct Tekst ao C12 495) (2 sh, 3) G, 9, 6) (11) ' Values with different superscripts are significantly different to the 0.05 level. For cryopreservation, the procedure of Fabre and Dereuddre (1990) was fol- lowed. Shoot tips approximately 2-3 mm in length were excised and placed in a solution of 3% alginic acid with 0.75 M sucrose in calcium-free MS me- dium. The solution containing the tips was then added dropwise to a solution of 100 mM CaCl, to form the alginate beads. These were incubated for 18 hr in liquid MS medium containing 0.75 M sucrose and were then dried for 4 hours in open dishes on filter paper under the air flow of the laminar flow hood. The dried beads containing the shoot tips were then transferred to 2 ml polypropylene cryovials and immersed directly into LN. After 1 hr, the cryo- vials were removed from the LN and allowed to warm at room temperature for 20 minutes. The encapsulated shoot tips were then transferred to basal medium for rehydration and recovery growth. Survival was measured as re- growth from a shoot tip. Results were analyzed by ANOVA (Tukey HSD Test, StatSoft, QuickSTATISTICA software). Drying and LN exposure significantly decreased survival from shoot tips which were precultured on basal medium alone (Table 1). When shoots were precultured for one week on basal medium plus 10 »M ABA, there was no decrease in survival when the tissues were dried, although survival did de- crease to approximately half when the tissues were exposed to LN. However, survival of both dried and LN exposed tissues was significantly better than that of control tissues. Shoot tips which were prepared using the encapsulation dehydration pro- cedure with and without ABA preculture were also transferred to LN for long- term storage. When these were removed after 3 1/2 yrs of storage in LN and recultured on basal medium, there was no regrowth from shoot tips which had been isolated from tissues grown on basal medium. However, 44% of the shoot tips which had been precultured on basal medium plus ABA initiated good growth after eight weeks (Figure 1). These results demonstrate that shoot tips of S. uncinata can survive LN exposure and long-term storage in LN when they are prepared using the en- capsulation dehydration procedure in combination with preculture on ABA. FiGuURE 1. Shoot tip of S. uncinata growing out of alginate bead, 8 weeks after exposure to LN. 6.7 X Encapsulation dehydration is a technique which has been successful with shoot tips of a number of species of angiosperms (Touchell and Dixon, In B. G. Bowes. Atlas of Plant Propagation and Conservation., Manson Publishers, London, 1999) as well as with gametophytes of ferns and bryophytes (Pence, The Bryologist 101:278-281, 1998 and Amer. Fern J. 90:16—23, 2000), and it is not surprising that it can also be successful with shoot tips of the fern allies as well. Survival of S. uncinata through cryopreservation was low unless the tissue was precultured on ABA. Treatment with ABA has been shown to initiate desiccation tolerance in the fern Polypodium virginianum L. (Reynolds and Bewley, J. Exper. Bot. 44:921-928. 1993a and J. Exper. Bot. 44:1771-1779, 1993b), as well as in bryophytes (Hellwege et al., Planta 198:423-432, 1996; Pence, 1998), and ABA is involved in several aspects of desiccation tolerance and water stress in seed plants (Hartung and Davies, In W. J. Davies and H. G. Jones, Abscisic Acid. Physiology and Biochemistry. BIOS Scientific Publishers, Oxford, 1991). It appears that preculture on ABA improves tolerance of S. uncinata shoot tips to the stresses of drying and LN exposure in the encap- sulation dehydration procedure, perhaps mimicking a natural response to stress in this species. Several other species of Selaginella are known to be “resurrection plants,” with a remarkable ability to tolerate extreme water loss 40 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 1 (2001) (see Proctor and Pence, In M. Black and H. Pritchard. Desiccation and Plant Survival. CAB International, Oxon, U.K., in press). Although encapsulation dehydration was successful in preserving shoot tips of S. uncinata, preliminary studies in this laboratory indicate that the survival of some other species of Selaginella through this procedure may not be as high (unpublished results). Species differ in their ability to survive LN exposure, and other techniques, such as slow freezing or vitrification, may be more suc- cessful with other species in this genus (Fukai et al., Euphytica 56:149-153, 1991; Yamada et al., Plant Science 78:81—87, 1991). There are over 700 species of pteridophytes which are of conservation con- cern worldwide, including 23 species of Selaginella (Walter and Gillett, 1997 IUCN Red List of Threatened Plants. IUCN—The World Conservation Union, Gland, Switzerland 862 Pp., 1998). Further studies should reveal how broadly applicable the encapsulation dehydration procedure is to shoot tips of pteri- dophytes and how cryopreservation may be used for preserving Selaginella germplasm for research and conservation.— VALERIE C. PENCE, Center for Re- search of Endangered Wildlife, Cincinnati Zoo and Botanical Garden, 3400 Vine Street, Cincinnati, OH 45220. Wii INFORMATION FOR aes Authors are encouraged to submit manuscripts pertinent to pteridology for pub- lication in the American Fern Journal. Manuscripts should be sent to the Editor. Acceptance of papers for publication depends on merit as judged by two or more referees. Authors are encouraged to contribute toward publishing costs; however, the payment or non-payment of page charges will affect neither the acceptability of manuscripts nor the date of publication Authors should adhere to the following guidelines; manuscripts not so prepared may be returned for revision prior to review. Submit manuscripts in triplicate (xerocopies acceptable), including review copies of illustrations and originals of illustrations. After review, submission of final versions of manuscripts on diskette (in PC- or Mac-compatible formats) is strongly encouraged. Use standard 8% by 11 inch paper of good quality, not “erasable” paper. Double space manuscripts throughout, including title, authors’ names and addresses, short, informative ab- stract, text (including heads and keys), literature cited, tables (separate from text), and figure captions (grouped as consecutive paragraphs separate from figures). Arrange parts of manuscript in order just given. Include author’s name and page number in upper right corner of every sheet. Provide margins of at least 25 mm all around on typed pages. Do not submit right-justified copy, avoid footnotes, and do not break words at ends of lines. Make table headings and figure captions self-explanatory. Use S.I. (metric) units for all measures (e.g., distance, elevation, weight) unless quoted or cited from another source (e.g., specimen citations). For nomenclatural matter (i.c., synonymy and typification), use one paragraph per basionym (see Regnum Veg. 58:39—40. 1968). Abbreviate titles of serial publi- cations according to Botanico-Periodicum-Huntianum (Lawrence et al., 1968, Hunt Botanical Library, Pittsburgh) and its supplement (1991). References cited only as part of nomenclatural matter are not included in literature cited. For shorter notes and reviews, omit the abstract and put all references parenthetically in text. Use Index Herbariorum (Regnum Veg. 120:1—693. 1990) for designations of her- baria. Illustrations should be proportioned to fit page width with caption on the same page. Provide margins of at least 25 mm on all illustrations. For continuous-tone illustrations, design originals for reproduction without reduction or by uniform amount. In composite blocks, abut edges of adjacent photographs. Avoid com- bining continuous-tone and line-copy in single illustrations or blocks. Coordinate sequence and numbering of figures (and of tables) with order of citation in text. Explain scales and symbols in figures themselves, not in captions. Include a scale and reference to latitude and longitude in each map. Proofs and reprint order forms are sent to authors by the printer. Authors should send corrected proofs to the editor and reprint orders to the printer. Authors will be assessed charges for extensive alterations made after type has been set. For other matter of form or style, consult recent issues of American Fern Jour- nal and The Chicago Manual of Style, 14th ed. (1993, Univ. Chicago Press, Chicago). Occasionally, departure from these guidelines may be justified. Authors are encouraged to consult the editor for assistance with any aspect of manuscript eparation. Papers longer than 32 printed pages may be sent to the Editor of Preridologia (Memoir Editor, see cover 2). PTERIDOLOGIA ISSUES IN PRINT The following issues of Preridologia, 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 postpaid. 2A. Lellinger, David B. 1989. The Ferns and Fern-allies of Costa Rica, Panama, and the Chocé (Part 1: Psilotaceae through Dicksoniaceae). 364 pp. $32.00 postpaid. Send your order with a check or money order to: American Fern Society, Inc., c/o U.S. National Herbarium MRC-166, Smithsonian Institution, Washing- ton, DC 20560. AMERICAN FERN JOURNAL ON MICROFICHE Volumes 1—61 of the American Fern Journal are available as archival quality, silver positive microfiches. Single volumes or the entire run may be purchased. The fiches are easily read with 10X or greater magnification (using a dissecting microscope and transmitted illumination or a fiche reader). Silver negative micro- fiches of vols. 1-50 are also available. The price is $4.00 per volume or $244.00 per set of 61 volumes, postpaid. Send your inquiry or order with a check or money order to: American Fern Society, Inc., c/o Dr. James D. Montgomery, Ecology II, Inc., R.D. 1, Box 1795, Berwick, PA 18603. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://www.amerfernsoc.org/ AMERICAN aie FERN nant JOURNAL ae QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Systematics of the Northern Andean Isoétes karstenii Complex Randall L. Small and R. James Hickey 41 Shorter Note Additions and Corrections to the Pteridophyte Flora of Northeastern Argentina onica Ponce 70 The American Fern Society Council for 2001 BARBARA JOE HOSHIZAKI, 557 N. Westmoreland Ave., Los Angeles, CA 90004-2210. CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS e045: ce-Presiden W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI is 14 iets JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-11 = Treasurer DAVID B. LELLINGER, 326 West St. NW., Vienna, VA 22180-4151. Membership Secnaas JAMES D. MONTGOMERY, Ecology III, R.D. 1, Box 1795, Berwick, PA 18603-9801. k Issues Curator GEORGE YATSKIEVYCH, Missouri Botanical Garden, P.O. Box 299, St. Louis, MO 63166-0299, Journal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166, Smithsonian Seep Washington, DC 20560-0166. Memoir Editor CINDY JOHNSON-GROH, Dept. of Biology, Gustavus Adolphus College, 800 W. College ee St. Peter, MN 56082-1498. Bulletin Editor American Fern Journal EDITOR R. JAMES HICKEY al ReaD eae OH 45 45056 ph. (513) <5, 6000, e-mail: hickeyrj @muohio.edu ASSOCIATE EDITORS GERALD J. GASTONY: ... ss Bae of “pata Indiana naires Bloomington, IN 47405-6801 CHRISTOPHER H. HAUFLER .... Dept. o: ae University wie ansas, Lawrence, KS 66045-2106 ROBBIN C. MORAN w York Botanical Garden , Bronx, NY 10458-5126 JAMES H. PECK San of a acu of Arkansas—Little Rock, 1 S. University Ave., Little Rock, AR 72204 The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the Sg study of ferns. It is owned by the American Fern Society, and err at 326 West St. NW., Vienn VA oe 1. Periodicals 5 postage paid at Vienna, VA, and additional entry. . hs (domestic) to 12 ae (foreign) after the date of issue, and oes 8 for back issues should be addressed to Dr. James D. Montgomery, Ecology II, R.D. 1, Berwick, ag doe Eians dress, dues, and applications for membership should be sent to the Membership Secretary. eral inquiries concerning ferns should be addressed to the Secretary. por ncaa $20.00 gross, $19.50 net if paid through an agency (agency fee $0.50); sent free to members of the ne Fern Society (ann ual dues, $15.00 + $5.00 mailing surcharge beyond A.). ack volumes are sola for most years as printed issues or on microfiche. Please contact the POSTMASTER: Send address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA 22180-4151. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society welcomes contributions from members and non-members, including miscellaneous notes, offers to ~~ or purchase materials, personalia, horti notes, and reviews of non-technical books on ferns Mr. Stephen McDaniel, 1716 Piermont Dr., Hacienda Hts., CA 91745-3678, is Director. Spores Gifts dh t ee < eg a pn _ nee eial al say: soy cana pag inert dame! yalronpete 1 | eee tax-deductible Inquiries should be addressed to the Secretary. American Fern Journal 91(2):41-69 (2001) F > §& Se oi Systematics of the Northern Andean &@ WV & . s ne Oe \O Isoétes karstenii Complex Se & = = RANDALL L. SMALL B SS a Department of Botany, The University of Tennessee, Knoxville, TN 3306 a = Oo R. JAMES HICKEY Department of Botany, Miami University, Oxford, OH 45056 ABSTRACT.—The Isoétes karstenii complex includes those species characterized by laevigate mega- spores, acute to free ala apices, a highly reduced labium, and distributions in the high altitude ivelata and I. precocia. Chromosome counts revealed that I. karstenii and I. precocia are diploids (2n = 22) and that I. palmeri is tetraploid (2n = 44). Estimates of chromosome number based on spore size for I. fuliginosa and I. hemivelata indicate that they are polyploid (2n = 44) The laevigate-spored taxa of Isoétes L. from the northern Andes have long presented taxonomists with considerable difficulty. These difficulties are due largely to the overall morphological similarity of the species and the wide range of morphological variation within the species. Also, few collections of these species were available for study until the field work of A.M. Cleef and collaborators and J. and S. Keeley in the 1970s and 1980s. As part of continu- ing efforts to describe and characterize the Neotropical species of Isoétes we undertook a revisionary study of this difficult group of eRe spored spe- cies, the Isoétes karstenii A. Braun complex (Hickey, 1985). The goals of this study were to: (1) determine the number of taxa that should be included within the I. karstenii complex, (2) clarify the taxonomy of the species, and (3) document intra- and interspecific morphological variation. To accomplish these goals we have collected morphological data and conducted principal components analyses of those data to infer morphologically discon- tinuous groups. In addition we have determined or estimated chromosome numbers for the included species. Finally, we present a taxonomic treatment that formalizes our understanding of the delimitations and relationships among these species. TAXONOMIC HISTORY OF THE ISOETES KARSTENII COMPLEX The Isoétes karstenii complex was first delimited by Hickey (1985) in a mor- phologically based cladistic analysis of Neotropical Isoétes. It includes those 42 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) TABLE 1. Summary of the taxonomic recognition of the species historically associated with the /soétes karstenii complex. A “+” indicates the taxon was accepted as valid by the author(s); an “S” that the taxon was considered a synonym of /. lechleri; a ““V” that the taxon was considered a variety of 1. lechleri. Isoétes Isoétes Isoétes Isoétes Isoétes colombi- rimbachi- Isoétes Isoétes i ana palmeri lechleri karstenii — socia cleefii Mettenius (1859) aa Braun (1862) + + + Baker (1880) + Ss S Motelay & Vendryes (1882) =A S S Underwood (1888) + 5 S. Pfeiffer (1922) + S 5 eber (1922) + + S Plamer (1929, 1932) + S S y Vv Rodriguez (1955) + S S Vv Vareschi (1968 a5 Vv Vv Gomez (198 + 5 Ss s S Fuchs-Eckert (1982) + ~ - ~ oo ~ Hickey (1985) + + + + + taxa with laevigate megaspores, truncate to free ala apices and corneous (dark- ly sclerified) leaf apices. The complex is distributed throughout the péramos of Colombia and Venezuela at elevations above 3000 m, with a single disjunct location at Mt. Chimborazo, Ecuador. Hickey (1985) included five species in this complex: Isoétes cleefii H. P. Fuchs, Isoétes palmeri H. P. Fuchs, Isoétes karstenii A. Braun, Isoétes socia A. Braun (sensu Fuchs-Eckert, 1982), and Isoétes rimbachiana H. P. Fuchs. In a compendium of South American species, Fuchs-Eckert (1982) recognized eight laevigate-spored species for Venezuela, Colombia, and Ecuador and therefore may be associated with the I. karstenii complex: I. socia, I. karstenii, Isoétes steyermarkii (nom. nud.), I. palmeri, I. cleefii, I. rimbachiana, I. colombiana (T. C. Palmer) H. P. Fuchs, and Isoétes glacialis Asplund (including Isoétes arumiana [nom. nud.] pro syn.). Fuchs- Eckert (1982), however, did not recognize these species as a distinct group. Historically these names have been treated in the literature as distinct species, or more often, as synonyms or varieties of Isoétes lechleri Mett., a central An- dean (southern Peru and Bolivia) species (Table 1). MATERIALS AND METHODS MorPHOLOGY.—Approximately 200 herbarium sheets representing about 90 discrete collection numbers of Venezuelan, Colombian, and Ecuadorian lae- vigate-spored Isoétes were examined. We gratefully acknowledge the following herbaria for use of material: B, BM, COL, F, G, GH, MO, MU, NY, U, UC, US and VEN. Sixty-three of the collections were chosen and scored for eight quan- titative morphological characters; those collections that were excluded from the analyses consisted of immature or poorly preserved material. Characters SMALL & HICKEY: ISOETES KARSTENII COMPLEX 43 used in the ordination and statistical analyses were: leaf number, leaf length (mm), leaf width (mm) at the mid-length, ala length/leaf length ratio, length (mm) of corneous portion of leaf apex, velum coverage (scored as a discrete character; see below), megaspore diameter (um), and microspore length (ym). Additional characters were scored for a subset of the collections but were aban- doned due to non-discriminatory age-specific correlation (e.g., corm dimen- sions and sporangial dimensions) or extreme variation within individuals and collections (e.g., sporangial shape and ala apex morphology). Some characters used in the analyses are also correlated with age (e.g., leaf number), or showed variability within collections (e.g., leaf length). Despite this, these characters also appeared to show group specific and discriminatory ranges of variation and were therefore retained. Terminology for morphological characters follows Hickey (1985), and for two-dimensional shapes Radford et al. (1972). Velum coverage was scored as: (1) complete to nearly complete (> 90%), or as (2) less than half covered (25-50%). For ordination analyses, velum cov- erage was therefore scored as a discrete character, complete vs. incomplete. Although microspore length and width were measured, only microspore length was chosen for the ordination and statistical analyses because most previous workers have used microspore length, rather than width, some re- porting only length values. Correlation between these two characters is high for 55 collections analyzed (r? = 0.877), and either character should adequately represent the collection. Measurements were taken only from mature plants and mature organs of those plants. Each collection, regardless of number of sheets or individuals, was considered representative of a single population unless evidence of hy- bridization (e.g., aborted spores) or multiple taxa within a collection (e.g., gross discontinuities in morphology) were present. Several individuals per collec- tion were scored for each character (with the exception of microspore dimen- sions which were taken from a single sporangium) and a mean score for each character was calculated. The number of measurements per character differed from collection to collection. Some collections consist of only a single plant on a single herbarium sheet while others consist of numerous individuals on numerous sheets. In addition, the quality of materials varied greatly: many specimens were partially destroyed by insects, many were pressed with the substrate around the corm, and many were pressed in large clumps. As a re- sult, some characters were difficult to obtain for all collections. Ten measure- ments per character per collection (25 for megaspores and microspores) were recorded whenever possible. Two types of analyses of morphological data were conducted: (1) principal component analysis (PCA) and (2) character by character scatter plots. These analyses were performed in an attempt to reveal morphological discontinuities between groups of collections and were conducted without assigning collec- tions to specific taxa, except for type specimens which were so designated. This was done to determine, without bias, whether or not distinct clusters of collections would be revealed which could be designated as putative morpho- logical species and to determine which type specimen(s) would be associated ts AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) with which cluster. Once morphologically cohesive groups were recognized, descriptive statistics were calculated to document putative intraspecific vari- ation. Multivariate and statistical analyses were performed using NTSYS ver- sion 1.80 (Rohlf, 1993) and Statmost v. 2.50 (DataMost). CyToLoGy.—Methods and materials for chromosome counts generally follow those of Hickey (1984). Voucher specimens for the materials from which counts were made are deposited at LOC, MU, and BM (Table 2). All specimens from which root tips were collected were grown completely submerged and potted in either fine quartz sand or a 3:1 mixture of sphagnum peat: fine quartz sand (Taylor and Luebke, 1986). Cells with mitotic figures were best obtained in root tips collected in the late morning (11:00 a.m. + one hour). Young roots about 1 cm long were harvested and immediately placed in a saturated para- dichlorobenzene solution, incubated at room temperature for 4—5 hr, fixed in Farmer’s solution (3:1 95% ethanol: glacial acetic acid) and refrigerated until squashed. Root tips were hydrolyzed in 1N hydrochloric acid at 60°C for 20- 30 min, neutralized in a 95% ethanol series, stained in Whittman’s hematox- ylin for 25 min at room temperature, rinsed in glacial acetic acid to destain, and squashed in Hoyer’s mounting media. Squash slides were examined and photographed at 600 or 1000. Color photocopies of the photographic slides were then prepared. These photocopies were placed on a light box, covered with a plain piece of white paper and the chromosomes traced onto the white paper from which counts were made. Simultaneous viewing of the photo- graphic and/or microscope slides and the photocopy aided in interpreting slides in which chromosomes overlapped. When live material was unavailable, we attempted to estimate chromosome number by plotting mean megaspore diameter vs. mean microspore length for taxa of known chromosome number and then plotting the mean value of spore sizes of the unknown taxa onto that plot. A strong correlation exists between spore size and chromosome number in Isoétes (Cox and Hickey, 1984; Britton and Goltz, 1991). For northeastern North American taxa, Kott and Britton (1983) noted that ‘ploidy level can be determined with some certainty from spore measurements, even for dried herbarium material.” Mean values for spore sizes of 27 taxa of known chromosome number were taken from Rury (1978), Kott and Britton (1983), Britton and Goltz (1991), Brunton et al. (1994), Taylor et al. (1993), Hickey (1994), Musselman and Knepper (1994), Brunton and Britton (1996a, 1996b), and Watanabe et al. (1996). While such correlations are unlikely to definitively identify the chromosome number of an unknown species, they may be useful in obtaining an estimate, or as in our case (see below), in ruling out potential ploidy levels. RESULTS MorPHOLOGY.—Included among the collections examined were type speci- mens (holo-, iso-, and/or paratypes) of Isoétes arumiana (nom. nud.), I. cleefii, I. colombiana, I. glacialis, I. karstenii, I. lechleri, I. palmeri, I. rimbachiana, I. SMALL & HICKEY: ISOETES KARSTENII COMPLEX (DOD Z£EOrI Kajaay x» Kajaay RIQUIO[OD “eorBUeUIpUND ‘esIe’] eunse’] Jo § [ood [euls 7 = UZ p1209add "J (NED) Z8PLT Kuuar RIQUIO[OD ‘vorBUIeUIpUND ‘BoRSsIyD vunse’] Jeou ood [jews T= UZ (NED ZZPZ[ Kusar BIQUIO[OD ‘oreUBUIpUND ‘eoRsIyD euNnse’y seou [ood [yeuIS rr = UZ (JOO ‘AW) 2ajpz2u0yH = /Co[-G¢T jjows BIQUIO]OD “eoreUeUIpUND ‘voRsTYyD vuNse’] Jeou ood [jeuIS th = UZ tiauypd ‘| (WE) 6ZPZ[ Kuuar RIQUIO]OD ‘RoreUBUIpUND ‘BoRsIYyD vuNse’y Jeou jood [yews TT = UZ (TOO ‘AW) 2a/vp2u0yH x» pp ows BIQUIO[OD ‘vovhog ‘ANDO_D [ap BPRASN BLIOIG ‘epespeng eT eunse] TT = UZ (TOD ‘AW) 2a7p2u0yH » CpI-[ pI ows BIqUIO[OD ‘eovkog ‘And0_D Jap vpeaodyN vIOIS ‘epequlg eT vunse’] Co = UZ (TOD ‘AW) 24p2u0yH x» gz] jjows BIQuIO[OD ‘eoeAog ‘B1090g ap ordioruN| ‘BASIg Op OUIeIeg ‘puog TZ = UZ (NW) 4ax91H ~P g J]DUS PpANZIUDA ‘epLIg| ‘Ifeqnony eunsey CZ = UZ (NIN) 4ay217 3p Z jJpws BONZIUDA “EPLIO|] ‘BISON vuNse’] TT = UZ (AW) 4ay21H ~P 9 JJDUS Bjanzous, “epLsy ‘e[Inby ap oold Jo N ury S| ‘puog co = UC (AWD) 4ax91H P ¢ OWS Blanzoua, “eplLis| ‘e[Inby ap oolg JO N WY ¢] ‘puog CC = UW (AW) 4ay21H PY p% [JOU Pjanzous, ‘pug ‘epinby op OdIg JO N Wy g ‘puog Co = UZ NUaISADY “] uoneso’] Joyono,, uones0’] uonsa[[o9 qunod “Byep UsUTISads JOYONOA puR ‘UONRdO] UONDET]JOS ‘s]JUNOD WOSOWOIYD “7 -ATAV I, 46 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) TABLE 3. Eigenvectors for the first three principal components (PC) from the PCA. Character PC] Pe. 2 PC 3 Leaf length 0.21957 0.77593 0.21528 Leaf width 0.63580 0.16820 —0.61498 Microspore length 0.84544 —0.13050 0.28320 Megaspore diameter 0.76647 =—O.05274 0.14659 % ala coverag 0.18744 —0.75048 —0.42986 Sclerified portion of leaf tip 0.04654 0.53506 —0.72654 Leaf number 0.54509 0.51899 0.12401 Velum coverage —0.81467 0.36054 —0.04755 socia, and IJ. steyermarkii (nom. nud.). Examination of the holotypes of I. lech- leri, I. glacialis and I. socia (all from the Andes of southern Peru or Bolivia) was performed initially to determine whether or not these taxa should be in- cluded in the subsequent analyses. Both I. glacialis and I. socia were deter- mined to be synonyms of I. Jechleri, which in turn was determined to be dis- tinct from the northern Andean taxa. For a complete discussion of the vari- ability and synonymy of the I. Jechleri complex see Hickey (1994), but a brief discussion of our rationale for excluding these taxa follows. The synonymy of Isoétes socia and I. lechleri has been proposed by all authors except Fuchs- Eckert (1982, 1992) since the former’s publication. The type specimen of I. socia (Lechler 1937b, B!) consists of a single immature plantlet and a few isolated sporophylls. These materials were apparently removed from Lechler 1937 (the type of I. lechleri) and undoubtedly represent a vegetative offset from I. lechleri, a species in which vegetative reproduction is common (Hickey, 1994). The type of I. glacialis (Asplund 4041, isotype B!) is but one of the many morphotypes of the highly variable I. lechleri and is clearly distinct, both in morphology and distribution, from Cuatrecasas 19117, the basis for I. arumiana. Sixty-three collections were used in the PCA, including all of the type spec- imens except those of Isoétes glacialis, I. lechleri, and I. socia. Eigenvectors for the first three principal components are given in Table 3; the first two principal components accounted for 36% and 23%, respectively, of the total variation. A plot of principal components one vs. two (Fig. 1) reveals six po- tentially distinct groups, which will be referred to in subsequent discussion by number. Clusters 1, 2 and 3 contain single collections: the type of I. cleefii, the “type” of I. arumiana (nom. nud.) and the type of I. rimbachiana respec- tively. Cluster 4 contains the type collections of I. karstenii, I. palmeri, I. col- ombiana, and I. steyermarkii (nom. nud.), paratypes of I. palmeri and I. cleefii, and many additional collections. Cluster 5 consists of a group of eight collec- tions without an associated type specimen. Cluster 6 contains a group of five collections without an associated type specimen. The constituents of each of these groups were examined to determine which characters separated each the others and to determine whether or not these groups constitute morphologically definable taxa. SMALL & HICKEY: ISOETES KARSTENII COMPLEX 47 |. karstenii & H 7¢ fil i ‘ cleefii {e}2 PC2 0- _@@ e |. hemivelata |, fuliginosa oreo . |. karstenii var. anomala 2 y T -1 0 1 2 3 PC 1 Fic. 1. Scatter plot of principal component one vs. principal component two based on morpho- logical data. Clusters of associated specimens are indicated and numbered. 1 = I. cleefii; 2 = I. fuliginosa; 3 = I. karstenii var. anomala; 4 = I. karstenii + I. palmeri; 5 = I. precocia; 6 = I. hemivelata. Cluster 1 (the holotype of I. cleefii) appears quite distinct from the remaining clusters in Fig. 1, yet examination of the characters suggests that the separation between clusters 1 and 4 is tenuous. Cluster 1 is separated from cluster 4 along principal component (PC) 2 by leaf length, leaf width and leaf number, yet character by character, the type of I. cleefii fell within the range of values of the individuals in cluster 4. The type collection of I. cleefii has longer, wider, and more numerous leaves than the average member of cluster 4. However, some collections in cluster 4 have values greater than or equal to those of the type of I. cleefii for some of these characters, but not for all. It appears that the combination of high values for all of these characters serve to separate the type of I. cleefii from the collections in cluster 4. Cluster 2 contains the single collection, Cuatrecasas 19117, which is the basis of I. arumiana (nom. nud.). This collection is separated from the re- maining collections along PC 1 by differences in velum coverage, megaspore diameter, and microspore length. In addition, the plant is larger than any other collection in this group, and the ligule is also unusually large (ca. 7 mm high) relative to the other species in this complex. This combination of character states makes this collection the most distinctive of those analyzed, and there- fore we suggest that it deserves species status. To avoid further nomenclatural and taxonomic confusion caused by the invalid publication and subsequent synonymy of J. arumiana (nom. nud.), we propose this taxon be recognized as Isoétes fuliginosa sp. nov. Cluster 3 contains the single collection, Rimbach 171, the type of Isoétes rimbachiana (= I. lechleri var. anomala). Differentiation of this collection from I. karstenii (from which it is otherwise nearly indistinguishable) by Palmer (1932) was based primarily on velum coverage, which also separates it from the collections of cluster 4 along PC 1 in our PCA. While velum coverage is 48 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) often considered to be a highly plastic character, we have found velum cov- erage to be uniformly complete in all collections of I. karstenii. Thus, the presence of an incomplete velum, along with the disjunct geographical place- ment of this collection at Mt. Chimborazo, Ecuador suggests that this collection represents a potentially unique biological entity and thus should be main- tained as a distinct taxon. We do not feel, however, that the distinction is sufficient to warrant specific status. Because of its affinities with I. karstenii it should be recognized as I. karstenii var. anomala comb. nov. Cluster 4 is a large and heterogeneous collection. Along with a large number of other collections, it contains the holotypes of Isoétes karstenii, I. colombi- ana, I. steyermarkii (nom. nud.), and I. palmeri, as well as paratypes of both I. palmeri and I. cleefii. Due to this heterogeneity and the dubious separation of I. cleefii, we performed additional analyses on this group (see below). Cluster 5 contains a group of eight collections without an associated type specimen. While placed close to cluster 4 (Fig. 1), this group of collections exhibits a unique set of character states that set it apart from the remaining specimens. Specifically, the PCA separates them from other specimens based on their diminutive size (leaf length, leaf width, and leaf number) at maturity, and the reduction of the corneous leaf tip. These specimens are also ecologi- cally differentiated as they are found solely in ephemeral pools, not permanent lakes and ponds as are the other species of this group. We propose that this taxon be recognized as Isoétes precocia sp. nov. Cluster 6 contains a group of five collections without an associated type specimen. These collections are distinguished from the remaining collections along PC 1 by velum coverage, megaspore diameter, and microspore length. This group of specimens is also unique within the complex in having trigonal rather than terete leaves. Although similar to cluster 2 (Isoétes fuliginosa) in velum coverage and spore dimensions, the two are easily distinguishable based on leaf cross-sectional shape, leaf length, ala length / leaf length ratio, leaf number, and the length of the sclerified portion of the leaf tip. We propose that this unique taxon be recognized as Isoétes hemivelata sp. nov. To examine the possibility that additional undetected groups existed within cluster 4 and to determine the status of Isoétes cleefii, two-dimensional scatter plots of all combinations of characters were produced for those collections in clusters 1 and 4. The scatter plot of leaf length vs. leaf width at the mid-point (Fig. 2) reveals two distinct groups with different character trends: one dis- playing relatively long, slender leaves, the other with relatively short, thick leaves. The group with relatively long, slender leaves contains type and par- atype collections of I. palmeri and I. cleefii (Fig. 2) as well as a number of additional collections. The other group (Fig. 2), with relatively short, thick leaves, contains the type collections of I. karstenii, I. colombiana, and I. stey- ermarkii (nom. nud.) as well as a number of additional collections. Linear regression analyses of these specimens, individually and taken together sup- port our inferences that these two groups are distinct. If all specimens in Fig. 2 are included in a regression analysis, the r? = 0.0009 (P = 0.85) indicates a lack of significant correlation. If however, the two groups (above and below SMALL & HICKEY: ISOETES KARSTENII COMPLEX 49 400 _ ° o . 300 - = ea = = Boo D Ss 2004 oO a — oO © ro | ° 100 + * 2 Leaf Width at Mid-Length (mm) Fic. 2. Two-dimensional scatter plot of leaf length (mm) vs. leaf width (mm) for those specimens included in cluster 4 of the PCA (Fig. 1). The line divides the relatively long, narrow-leaved specimens (J. palmeri) from the relatively short, stout-leaved specimens (I. karstenii). the line in Fig. 2) are analyzed separately, values of r? = 0.70 (P = 1.2 X 10°-®) and r? = 0.25 (P = 0.012), respectively, are obtained. These values indicate that leaf length and width are significantly correlated within groups, thus lend- ing support to our hypothesis that they are biologically distinct. The separation of these two very similar groups appears to have been obscured in the initial PCA by the inclusion of such a morphologically diverse selection of speci- mens. The range of variation encompassed by the initial PCA was much larger than the variation between groups within cluster 4, therefore the entire range of variation encompassed by cluster 4 was confined to a small portion of the entire PCA space. A more fine-scaled analysis was then required to detect these differences. In addition to the separation of the two groups described above, this analysis clarifies the position of the type collection of I. cleefii as an ex- treme individual along a continuum (Fig. 2). On the basis of these analyses we delimit two additional species. The group with relatively long, slender leaves includes the type collections of Isoétes palmeri and I. cleefii. These two taxa were published simultaneously (Fuchs- Eckert, 1981a,b), and thus neither has priority. We propose that this taxon be recognized as I. palmeri because the type specimen is more representative of the species overall. The group with relatively short, wide leaves includes the type collections of I. karstenii, I. colombiana, and I. steyermarkii (nom. nud.). The name J. karstenii was the first of these to be published (Braun, 1862) and therefore has priority. In order to document the variation within these species, we calculated the mean and range of each character (Table 4). Although leaf length / leaf width at the mid-point ratio was not included as a character in the PCA (because both leaf length and leaf width were included as separate characters) it is an important character in discriminating Isoétes karstenii from I. palmeri and was therefore included in descriptive statistics. Additionally, as chromosome num- AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) br = UZ rr = UZ 7 = TZ = UZ Joquinu suwosowlo14 8V Orel + 6E)-LE Of + on 9 + ght 8E-(L'€ + O£)-S7 (wt) Sua] asodsossTy 00L—009 7Z9O-'6b + SOS)-EIS S6r7LE + Trh)-88E ISS66Il + 88P)-88E LOS-+6OII + 6Lr)-S6E (wit) sojourerp asodsesayy %SL-OS %SL-SE 006< %06< %06< (%) A3vI9AOD WIN]PA 0 JIXE0 + T1D-8'0 FO-1'0 + TOK-T0 y7HS'0 + FD-8'0 TE+(90'0 F BI-L'0 (wut) xede jee] Jo uonsod soaws0D S10 1 0-(S0'0 ¥ PE'O)-87'0 90-00 + OE0)-970 = 870-00 + 6I'0)-I 1.0: 8€'0-(90'0 + 97'0)-F1'0 ones YBue] Jea/psua] epy vel T8179 + ¥9)-9r LSIXS'be + 08)-9r Ose{T'bs + IP7)-6ll LEIP le + SL)-9€ ones Pp jeop/pSug] sea] 6'l v71S0 + OTT 60-470 + L'0)-S'0 bI-tr'0 = O'D-S'0 v7r'0 + 91-60 (wu) YISue]-pru we YYpIM Jeo] SST p8I-L'9s + 671)-Ss bv60'6I + TS)-LE Pr6E1{T'06 + OET)-€6 Isz-(8'es + SI1)-Or (wi) ySuey Jeo] Sc 91-81 + ELI HI-@O'T + L-9 t7-9'P + PIL IZ41S'€ + E18 JOQqUINN, JB9"T (I = (8 = N) Pivjaanuay ‘J (8 = N) »1909a«d ‘7 (@Z = N) Mauypd ‘7 (pZ = N) Muassapy J pmo yi ‘pojuasaid ase (sasayjuoied Ul) JOA prepuris | URW puke ddUBI dy) Ja}ORIRYO puR UOXe} YoRe JO,] ‘pafdues suondea]]O9 Jo Joquinu oy} = N UOXe) yore JO “xa[duIOD NUAaIsupy $ajaOS]T AY] JO SIAQUIDUI JO] Sayv}s Ja}OvIeYO Burysinsunsiq: “fp ATAVL SMALL & HICKEY: ISOETES KARSTENII COMPLEX 51 —= 1000+ 12 a = 900 how 2 800 4 ao = se cay (OO 10 c F g 600 4 4 Q 500; 455 & 4 g es Bo 4007 2 2 4 a. 2 4 = 300 3 Microspore length (um) Fic. 3. Two-dimensional scatter shee - whist ia length (wm) vs. megaspore diameter (ym) for species of Isoétes with known chromosome number. tonaes fA represents a single species with its chromosome number identified = habe (e.g., = 2x = 22, “4” = 4x = 44, etc.). The two unknown species are identified by letter (““H” = I. sedate! ‘F” = I. fuliginosa). bers are important in distinguishing taxa (see below), these data are also pro- vided in Table 4 CyToLocy.—Chromosome counts were performed for Isoétes karstenii, I. pal- meri, and I. precocia (Table 2). It was found that I. karstenii and I. precocia were diploids (2n = 22) and I. palmeri was tetraploid (2n = 44). In addition to numerous counts of 2n = 44 for I. palmeri, a single individual from one population of I. palmeri was found to be a triploid (2n = 33). This individual presumably represents a hybrid between I. palmeri and a diploid species, such as I. karstenii or I. precocia, both of which are sympatric with I. palmeri in the area of this population. These cytological data support the findings of the morphological analyses in further distinguishing I. karstenii and I. precocia from I. palmeri. Live material was not available for Isoétes fuliginosa or I. hemivelata. The relatively large size of the mega- and microspores of these taxa, however, sug- gest that they are most likely polyploid. We attempted to estimate the chro- mosome numbers of these taxa by comparing their spore sizes with taxa of known chromosome number. A scatter plot of megaspore diameter vs. micro- spore length was produced with individuals identified by chromosome num- ber (Fig. 3). Individual ploidy clusters show overlap with adjacent ploidy clus- ters and thus it is not possible to confidently characterize the chromosome numbers of the unknown species. Despite this, these data clearly indicate that these taxa are polyploid, both being well out of the range of known diploids. Isoétes fuliginosa appears to be a high-level polyploid, probably at least a hexa- ploid. Isoétes hemivelata is more difficult to characterize because it falls out near taxa of three different ploidy levels, yet the data suggest that it is at least a tetraploid, possibly higher. 52 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) DISCUSSION Taxonomy and Analyses As a result of our morphological and cytological analyses, we recognize the following taxa within the I. karstenii species complex: I. fuliginosa; I. hemi- velata; I. karstenii var. karstenii; I. karstenii var. anomala; I. palmeri; and I. precocia. The details of this taxonomic history are given in Table 1. The conclusions presented here are based on evaluation of morphological characters using multivariate analyses to derive distinct groups of populations. In the case of these taxa, traditional approaches using intuitive taxonomic as- sessments have led to undue lumping (placing all taxa in I. lechleri) or splitting (recognition of up to eight taxa within this complex by Fuchs-Eckert 1982). Many workers have discussed the morphological plasticity in Isoétes, yet few have emphasized the fact that circumscribing that variation is imperative for understanding species delimitations (but see Kott and Britton, 1985). Evidence presented in this study supports the utility of quantitative vegetative charac- ters in delimiting species, especially when those characters are examined us- ing a multivariate approach. Cytology and Speciation Speciation in Isoétes appears to occur primarily via two pathways: (1) al- lopatric speciation of populations of the same ploidy level, and (2) speciation via allopolyploidy (Taylor and Hickey, 1992). Cytological and morphological data suggest that both modes of speciation have occurred in the I. karstenii complex. Reticulate evolution involving diploids and allopolyploids has been well documented in Isoétes (Taylor et al., 1985; Hickey et al., 1989; Taylor and Hickey, 1992), yet autopolyploidy is apparently rare, having been reported only once (Rumsey et al., 1993). Allopolyploids in Isoétes are generally char- acterized by an intermediate morphology with respect to the parental species, especially in megaspore morphology, and additive chromosome numbers and allozyme banding patterns (Taylor and Hickey, 1992). Autopolyploids would be expected to strongly resemble the parental species in all characters except chromosome number and spore size. Because the tetraploid Isoétes palmeri has completely laevigate megaspores and Isoétes allopolyploids appear to always have intermediate megaspore sur- face morphology only those diploid taxa with laevigate megaspores should be considered as possible parents. Based on the results of this study only I. kar- stenii and/or I. precocia are likely to be involved. Although conclusive evidence to address the auto- vs. allopolyploid origin of I. palmeri is unavailable, the data do suggest that the allopolyploid hypoth- esis is more plausible for several reasons. First, as mentioned, autopolyploidy is rare in Isoétes, but allopolyploidy is frequent. Second, although I. karstenii, I. precocia, and I. palmeri are all morphologically similar, I. palmeri is inter- mediate for some characters (e.g., leaf width, length of corneous portion of the SMALL & HICKEY: ISOETES KARSTENII COMPLEX 53 Fic. 4. Hypothesized relationships among the species of the I. karstenii aia Circled X’s represent a hybridization event; boxed 2X’s represents a chromosome doubling ev leaf). Finally, the presence of a triploid in one of the populations of I. palmeri suggests that hybridization occurs between I. palmeri and one (or both) of the diploids. Although this does not bear directly on the origin of I. palmeri, it suggests that hybridization does occur in this group. Isoétes karstenii and I. precocia are both diploids and therefore have their origins in divergent (rather than reticulate) speciation. These two species exist in different habitats: I. karstenii is evergreen and is found in permanent lakes and ponds while I. precocia is ephemeral and is confined to small, seasonal pools which fill during the wet season and desiccate during the dry season. Given the phenetic resemblance and geographic sympatry of these two taxa they are probably closely related. Whether one gave rise to the other or they are each descendants of a now extinct taxon is beyond the scope of this study. Species Relationships Although further taxonomic and biosystematic work needs to be done, we offer a preliminary hypothesis regarding the relationships among the species in this complex (Fig. 4). The base of these relationships rest on the two diploid species, [. karstenii and I. precocia. Hybridization between these two species followed by chromosome doubling in the F, results in the allotetraploid, I. palmeri. The origin of J. hemivelata is less clear, but one possibility is hybridization between a member of the I. karstenii complex and a species outside the com- plex, possibly I. andina or one of several undescribed Colombian or Ecuador- ian species (Hickey, unpublished data). The evidence to support this includes: I. hemivelata’s possession of character states derived from the I. karstenii com- 54 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) plex and from the J. andina complex, along with its undoubtedly polyploid nature. The origin of I. fuliginosa possibly occurred by hybridization between I. hemivelata and another member of the J. karstenii complex. Isoétes fuliginosa is clearly polyploid, shares the character state of reduced velum coverage with I. hemivelata, and is larger than the other members of the I. karstenii complex, but in other respects maintains the characters of the complex. TAXONOMIC TREATMENT Isoétes karstenii Complex Corm globose to horizontally elongate, two lobed (rarely three), size variable; dichotomously branching roots arising from the circumbasal fossa(e). Leaves variable in number, length, and width, flexuous or stiff; alae chartaceous, brown to black, extending upward to 50% of total leaf length, ala apices at- tenuate, acute, truncate or free; subula terete to trigonal, olive green to brown; leaf apices corneous, typically spatulate. Scale leaves, phyllopodia, stomates and peripheral fibrous bundles not seen. Sporangia (ob)ovate to elliptic, tan, concolorous to spotted, basal. Velum incomplete to complete. Ligule deltate to trullate. Labium absent or reduced to a small transverse ridge on the lower lip of the foveola. Megaspores white, 350-700 ym in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 25—55 ym in length, echinate. Vegetative reproduction not observed. As discussed previously, these taxa have often been placed in synonymy with, or regarded as varieties of, I. lechleri. They can be distinguished from that species, however, on the basis of ala length/leaf length ratio (up to 50% in the I. karstenii complex vs. 50-75% in I. lechleri), ala apex morphology (attenuate to free in the I. karstenii complex vs. always attenuate in I. Jechleri), vegetative reproduction (common in I. Jechleri vs. absent in the I. karstenii complex), and distribution (J. karstenii complex in Ecuador, Colombia and Venezuela vs. I. lechleri in southern Peru and Bolivia). Additionally, the only member of the I. karstenii complex which strongly resembles J. Jechleri is I. karstenii. These two taxa can be further differentiated on the basis of chro- mosome number: I. karstenii has 2n = 22 and I. lechleri has 2n = 44 (Hickey, 1994). KEY TO SPECIES OF THE I. KARSTENII COMPLEX 1. Velum covering 25-75% of the sporangium 2. Alae extending 10-20% up the leaf length; leaves 150-310 mm long; velum covering 50-75% of the sporangium; microspores 45-55 wm long; megaspores 600—700 pm in CONE Si ci eS Ey atlanta RHEE pel oe ees pascal, Me Bo 1. I. fuliginosa 2. Alae extending 25-45% up the leaf length; leaves 40-200 mm long; velum covering 30-50% of the sporangium; microspores 25-45 ym long; megaspores 300-650 pm in diameter 3. Megaspores 300-450 pm in diameter; microspores 30—40 ym long; subula terete Sea a Se Ns a a re 4b I. karstenii var. anomala SMALL & HICKEY: ISOETES KARSTENII COMPLEX 55 3. Megaspores 500-650 ym in diameter; microspores 35-45 wm long; subula trigonal Be Re eee ag 8 aimee: ost ar es Cor yh Se Pec eT eee ee eS te 2. I. hemivelata 1. Velum covering >90% of the sporangium 4. Leaf apices not or only slightly sclerified (the sclerified part < 0.5 mm long); leaves = than 100 mm long, less than 1.0 mm wide at the middle; plants seasonally ephem- ON co ee he Ata Fe 4 1p. 5's ie Weds cy uy RiSMMR < es 4s Btw aoe oh eee 3. I. precocia 4. Leaf apices distinctly sclerified (the sclerified part 0.7—3.2 mm long); leaves pee ong, 0.5-2.4 mm wide at the middle; plants seasonally persistent . Leaves 25-200 mm long and 1.0-2.5 mm wide with a leaf ge width ratio ee ee Sa ica DES e ee ae cee I. karstenii var. oe . Leaves 100-400 mm reas and 0.5-1.5 mm wide with a leaf length /loaf width r ee ee i sic any wlonei kw lg «Se Mh tais wie eae are ae salvos 1) oa 1. Isoétes fuliginosa R. L. Small and Hickey, sp. nov. (Fig. 5-7; 14-15). Type: “Colombia. Departamento del Cauca: Cordillera Central, Lagunilla de las Casitas, 3700 m alt., 3 Dec 1944” Cuatrecasas 19117 (Holotype: F!; Isotypes: COL!, F!, G-2!, GH!, MO!, MU!, US!). Isoétes arumiana Fuchs-Eckert, nom. nud. Proc. Kon. Ned. Akad. Wetensch. C. 85(2):259. 1982. Based on Cuatrecasas 19117 (US!) Ob statura magna, velo incompleto et megasporis grandibus inter omnes Is- oétes species ab Andibus septentrionalibus peculiaris; Isoétes karstenii A. Braun tangit, ob megasporis laevibus et ob apicibus aliis truncatis versus lib- eris. Corm horizontally elongate, two-lobed, 9-10 (x = 9.7) mm high, 24-25 (x = 24.7) mm wide; dichotomously branching roots arising from the circumbasal fossa. Leaves 18-34, flexuous, to 310 mm long, 8-12 mm wide at base, 1.5— 2.5 mm wide at mid-length, leaf length/width ratio 134; alae chartaceous, brown to blackened, extending 11-17% of total leaf length, apices acute to truncate; subula terete in cross-section, olive green, apex obtuse, not or only slightly corneous at the tip; peripheral fibrous bundles, stomates, scale leaves and phyllopodia absent. Sporangia obovate to ovate, basal, unpigmented, 5.5— 7.8 (x = 7.2) mm long, 4.5—5.8 (x = 5.2) mm wide. Velum incomplete, covering 50—75% of the sporangial surface. Ligule deltate and auriculate, ca. 7 mm high, 5 mm wide. Labium absent or reduced to an extremely small transverse ridge on lower lip of the foveola. Megaspores immature, but ca. 600-700 pm in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 45-55 (x = 48.2) pm in length, 30-40 (x = 35.4) ym wide, echinate. Chromosome number unknown. Isoétes fuliginosa is known only from the type collection which was taken from a small lake or pond (‘‘lagunilla’’) at 3700 m altitude in the Cordillera Central, Departamento Cauca, Colombia (Fig. 8). This collection was made in December and contains both microspores and megaspores; however, the mega- spores are immature. Further collections are necessary to ina a the phe- nology and full range of morphological variation of this spec Despite the immaturity of this collection, Isoétes fuliginosa | is oe from 56 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) Fics. 5-7. Isoétes fuliginosa (Cuatrecasas 19117). Fic. 5. Habit view showing reduced alae (F #1351470). Bar = 5 cm. Fic. 6. Microphyll base showing reduced alae and partial velum (US). Bar = 1 cm. Fic. 7. Variation in ala apices: truncate, acute and free respectively (F # 1351471). Bar = 5 mm. other species of Isoétes and can be recognized by its incomplete velum, large growth habit and large mega- and microspores. It is clearly related to the other members of the J. karstenii complex given its laevigate megaspores and trun- cate to free ala apices, although the incomplete velum and apparently high ploidy level indicate that an outside influence is also present. The results of the spore analyses show that I. fuliginosa is polyploid (probably hexaploid or SMALL & HICKEY: ISOETES KARSTENII COMPLEX 57 ie an DS Fan 78°W : 0 4°N e bie ts Ps t N 4b Fic. 8. Distributions of the species of the I. karstenii complex. Open triangles = I. karstenii, —— arena = I. precocia, filled squares = I. palmeri, filled circles = I. hemivelata, and the fuliginos higher) and therefore is likely an allopolyploid derivative of other nothospe- cies. 2. Isoétes hemivelata R.L. Small and Hickey, sp. nov. (Fig. 9-11, 16—17) Type: “Colombia, Departamento del Valle, Cordillera Central, vertiente oc- cidental; cabeceras de los rfos Tulua y Bugalagrande: Paramo de Las Vegas, alt 3600-3800 m, 22 March 1946” Cuatrecasas 20301 (Holotype: F!; Isotypes: M!, NY!, UC!, US!) Haec species ab Sectione Isoete karstenii praeclare distinguitur subulis tri- gonis et velis 30-50% obtegentibus. Corm globose, two-lobed, 4—12 (x = 6.5) mm high, 6-11 (x = 8.1) mm wide; dichotomously branching roots arising from the circumbasal fossa. Leaves 10— 25, stiffly erect, to 200 mm long, 4-6 mm — at base, 1.2-2.4 mm wide at mid-length, leaf length/width ratio 45-82 (x = 64); alae chartaceous, dark brown to black, extending 28-51% (x = 37%) = leaf length, apices acute to 58 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) Fics. 9-13. Isoétes cases: (Cuatrecasas 20301, US) and I. precocia (Oberwinkler & Ober- sides 12968, M). Fic. 9. Habit view of several plants of I. hemivelata. Bar = 2 cm. Fic. 10. Leaf base of I. hemivlt phere partial — agg Bar = 2 mm. Fic. 11. Subula apices of I. pandas showing black, corneous apice: = 2 mm. Fic. 12. Habit view of a mature fertile specimen of I. pideesctosy Bar = 1 cm. Fic. a Leaf base of I. precocia agen a well filled mega- sporangium with a complete velum and the narrow, parallel alae. Bar = 2 m SMALL & HICKEY: ISOETES KARSTENII COMPLEX 59 Fics. 14-19. Isoétes fuliginosa, I. hemivelata, and I. precocia. Fics. 14-15—-SEM of spores of indies fuliginosa (Cuatrecasas 19117, US). Fic. 14. Equatorial view of immature megaspore. Bar = 20 Fic. 15. Proximal view of microspore showing one of the two proximal surfaces. Bar = 20 um. Fics. 16-17—SEM of spores of I. hemivelata (Cuatrecasas 20301, F). Fig. 16. Proximal view of megaspore. Scale as in Fig. 14. Fig. 17. Proximal view of microspore. Scale as in Fig. 15. Fics. 18-19—SEM of spores of I. precocia 2 (Oberwinkler & Oberwinkler 12968, M). Fig. 18. Near _ajesinegeee view of megaspore. Scale as in Fig. 14. Fig. 19. Equatorial view of microspore. Scale as in Fig. 1 60 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) free; subula + trigonal, olive green; leaf apices blunt, often spatulate, distal 0.75—1.75 mm of apex corneous; peripheral fibrous bundles, stomates, scale leaves and phyllopodia absent. Sporangia elliptic to obovate, basal, spotted, 1.5-3.2 mm long, 1-2 mm wide. Velum incomplete, covering 30-50 % of the sporangial surface. Ligule shallowly deltate to trullate, to 1.5 mm long, to 2 mm wide. Labium absent or reduced to a transverse ridge on the lower lip of the foveola. Megaspores white, 500-650 (x = 560) pm in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 35—40 (x = 38.5) um long, 25-30 (x = 27.6) wm wide, echinate. Chromosome number unknown. Isoétes hemivelata grows as a submerged aquatic in permanent lakes and ponds (occasionally streams) between 3300-4300 m. It is distributed through- out the central and southern Departamentos of the Colombian Andes including Meta, Risaralda, Huila, Cauca and Valle (Fig. 8). Although known from only a few collections, those included in this study were made in January, March, August and October and all contained both mega- and microspores. This suggests that this species is evergreen (as are most Andean Isoétes) and produces both mega- and microspores throughout the year. Isoétes hemivelata is one of the most distinctive elements within the I. kar- stenii complex given its trigonal subula and incomplete velum. It can be dis- tinguished from I. karstenii, I. palmeri, and I. precocia by these characters and its larger spores. It can be distinguished from I. fuliginosa by its trigonal sub- ula, smaller spores, and greater ala length / leaf length ratio. Analyses of spore size suggest that I. hemivelata is polyploid. The characters it shares with the I. karstenii complex indicate that one of the parental species comes from this complex; however, the trigonal subula and reduced velum coverage must have come from a source outside this complex. This suggests hybridization and polyploidization. One likely parent is I. andina Spruce ex Hook., a sympatric species with strongly trigonal leaves and an incomplete velum. One collection seen by us (Cuatrecasas 19058 F!, GH!, US!) apparently represents a collection of hybrids between a member of the I. karstenii com- plex and I. andina. This collection is highly variable for most characters and spore abortion is evident. While most megaspores of this collection are slightly to distinctly papillate, some are almost laevigate and all individuals of the collection have ~ trigonal leaves. Chromosome doubling in a hybrid such as this, or backcrossing to the J. karstenii complex parent could feasibly result in a taxon which strongly resembles I. hemivelata. Additional specimens examined: COLOMBIA. Departamentos Huila/Cauca: Macizo Colombiano; péramo de Las Papas, Cerros y alrededores de la laguna La Magdalena, alt. 3530 m, 16 Oct 1958, Idrobo, et al. 3307 (COL, NY). Departamento Meta: Péaramo de Sumapaz, Hoya Sitiales, Laguna La Primavera y alrededores, alt. 3510 m, 9 Jan 1973, Cleef 7555 (COL, U); Paramo de Sumapaz, Hoya El Nevado, Laguna La Guitarra y alrededores, alt. 3425 m, 23 Jan 1973, Cleef 8268 (COL, U). Departamento Risaralda: Cordillera Central; Municipie de Pereira; parque de los Nevados, alt 4250 m, Jaramillo, et al. 5700 (COL, U). Departamento del Valle: Péramo Pan de Azucar, alt 3300-3700 m, 23 Aug 1968, Espinal & Ramos 2471 (COL). SMALL & HICKEY: ISOETES KARSTENII COMPLEX 61 3. Isoétes precocia R.L. Small and Hickey, sp. nov. (Fig. 12-13, 18-19) Type: “Venezuela: Anden, Estado Mérida: Sierra de St. Domingo, Paramo de Mucubaji, Weg zur Laguna Negra, an stark vernaéBten Stellen, + 3500 m. 4. 10. 1968.” B & F Oberwinkler 12968 (Holotype: M!; Isotype: VEN!). Isoétes socia sensu Fuchs-Eckert 1982 pro parte, Proc. Kon. Ned. Akad. We- tensch. C. 85(2): 256 1982. non A. Braun. Species nova I. karstenii proxima, cujus megasporas laeves et velum com- pletum habet; differt staturo minore, apicibus foliis brevissimis corneis et ha- bitu ephemero. Corm globose to horizontally elongate, two lobed, 1.8—2.4 mm (x = 2.1) high, 2.5—4.6 mm (x = 3.6) wide; dichotomously branching roots arising from the circumbasal fossa. Leaves 6-11, flexuous, to 63 mm long, to 4 mm wide at base, 0.5 to 0.9 mm wide at mid-length, leaf length / width ratio 45-157 (x = 80); alae chartaceous, caramel-colored, extending 26-36% of total leaf length, apices acute to truncate; subula essentially terete in cross-section, pale green, apices blunt, corneous only at 0.1-0.2 mm of the tip; peripheral fibrous bun- dles, stomates, scale leaves and phyllopodia absent. Sporangia elliptic to cir- cular, basal, unpigmented, 1.6—2.7 mm (x = 2.1) long, 1.2-2.1 mm (x = 1.7) wide. Velum complete, covering 100% of the sporangial surface (rarely slightly less). Ligule ephemeral, not seen. Labium absent or reduced to an extremely small transverse ridge of tissue on lower lip of the foveola. Megaspores white, lustrous, 388—495 (x = 437) pm in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 27-30 (x = 29) ym in length 18.9—23.3 (x = 20.5) um wide, echinate. Chromosome number: 2n = 22. Isoétes precocia grows submerged in shallow, temporary pools at 3300—4400 m. It is currently known from Estado Mérida, Venezuela and Departamentos Boyaca, Cundinamarca and Meta in Colombia (Fig. 8). Although only eleven collections of this species have been found among the specimens examined, it is probably a common plant of seasonal pools but has not been collected due to its diminutive size and ephemeral nature. We visited the type locality in March of 1992, yet were unable to find this species because the pools were dry. Collections examined in this study were made in March, April, July, Oc- tober, November and December, and all contained mega- and microspores. Fuchs-Eckert (1982) recognized that a taxon resembling the type of I. socia existed, but failed to differentiate between the vegetative plantlets of I. lechleri in Peru and Bolivia and the distinct species (J. precocia) of Colombia and Venezuela. Hickey (1985) recognized that a distinct species existed, but did not study the associated specimens and based his acceptance of the name I. socia on the work of Fuchs-Eckert (1982). Hickey (1994) has subsequently placed I. socia in synonymy with I. lechleri. Additional specimens examined: COLOMBIA. Departamento Boyaca: Sierra Nevada del Cocuy, Alto Valle Lagunillas, 1.5 km NE de la Laguna Pintada, 4390 m, 30 Nov 1972, Cleef 5704 (U); Sierra Nevada del Cocuy, Boqueron de Cusiri, alt. 4310 m, 5 Mar 1973, Cleef 8785 (U); Paéramo de la Sarna entre Sogamoso y Vado Hondo, alt. 3330 m., 8 Apr 1973, Cleef 9522 (COL, U). Departamento Cundinamarca: Péramo de 62 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) Sumapaz, Chisaca, 100 m SW de la Laguna Larga, alt. 3700 m, 11 Dec 1971, Cleef 153 (COL, U); Péramo de Cruz Verde, alt. 3200 m, 16 Oct 1961, Idrobo & Hotheway 4174 (COL); Péramo de az, vicinity of Laguna Chisaca, pool 50 m S of Laguna Larga, alt. 3620 m, 7 Jul 1986, Keeley & Keeley 11032 (MU); Péramo de Sumapaz, vicinity of Laguna Chisaca, pool 50 m S of Lagun Larga, alt. 3620 m, 7 Jul 1986, Keeley & Keeley 11033 (MU). Departamento Meta: Pdéramo de Sumapaz, Cerro Nevado del Sumapaz, alt. 3725 m, 21 Jan 1973, Cleef 8185A (COL). VENEZUELA. Estado Mérida: Sierra de Sto. Domingo, Péramo de Mucubaji, alrededores de la Laguna Grande, alt. 3560-3600 m 19 Nov 1959, Barclay & Juajibioy 9588 (COL, GH). 4a. Isoétes karstenii A. Braun var. karstenii. Verh. Bot. Vereins Prov. Brandenberg 3/4:332. 1862. Figs. 20-21. Type: “Venezuela. Im Gebirge von Mérida 8000’ hoch in einem See ganz unter Wasser 1857.” Karsten s.n. (Holotype B!). Isoétes lechleri var. colombiana T. C. Palmer, Amer. Fern J. 19: 18. 1929. I. colombiana (T. C. Palmer) H. P. Fuchs, Proc. Kon. Ned. Akad. Wetensch. C. 85(2):259. 1982. Type: “Colombia, Dept. Santander: Laguna de Cunta, edge of Paéramo de Santurban; alt. 3,880 m, 21 Jan 1927” Killip & Smith 17964 (Holotype PH; Isotype: NY!), mixed collection. Isoétes steyermarkii Fuchs-Eckert, nom. nud. Proc. Koninkl. Ned. Acad. We- tensch. C85:257. 1982. Based on Steyermark 55904 (US!). Corm globose to horizontally elongate, two lobed (rarely 3), 2.0-4.0 mm (x = 2.8) high, 2.8-7.8 mm (x = 5.2) wide; dichotomously branching roots arising from the circumbasal fossa(e). Leaves 8—21, flexuous to stiffly erect, to 251 mm long, 5-10 mm wide at base, 0.9 to 2.4 mm wide at mid-length, leaf length/ width ratio 35-157 (x = 77); alae chartaceous, light-brown to blackish-brown, extending 14-38% of total leaf length, apices acute to truncate; subula essen- tially terete in cross-section, olive green, apices blunt, often spatulate, distal 0.5-3.2 mm of apex corneous; peripheral fibrous bundles, stomates, scale leaves and phyllopodia absent. Sporangia widely elliptic to circular, basal, unpigmented, 2.1—5.5 (x = 3.4) mm long, 1.8-3.9 (x = 2.6) mm wide. Velum complete, covering 100% of the sporangial surface (rarely slightly less). Ligule deltate, auriculate, 0.5—-1.5 mm high, 0.75—2.0 mm wide. Labium absent or reduced to an extremely small transverse ridge on lower lip of the foveola. Megaspores white, lustrous, 400-507 (x = 467) um in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 25-38 (x = 30) pm in length, 17.7—28.7 (x = 20.8) ym wide, echinate. Chromosome number: 2n = 22. Isoétes karstenii grows submerged in permanent lakes and ponds (occasion- ally streams) between 3300-4600 m. It is distributed throughout the eastern cordillera of the northern Andes from Estado Mérida, Venezuela, to Departa- mentos Arauca, Boyaca, Cundinamarca, Meta, Tolima and Narifio in Colombia (Fig. 8). The collections of this species were made from every month except August and October. With the exception of the single collection made in June, SMALL & HICKEY: ISOETES KARSTENII COMPLEX 63 Fics. 20-25. Isoétes karstenii var. karstenii, I. —— var. anomala, and I. palmeri. Fics. 20— O id ogee 13685, M). Bar = 20 pm. Fics. 22—-23—SEM of spores of I. karstenii var. anomala (Rimbach 71, US). Fic. 22. Distal view of ete Scale as in Fig. 20. Fic. 23. Equatorial view of micro- spore. Scale as in Fig. 23. Fics. 24-25—SEM of spores of I. palmeri (Cleef 8647, BM). Fic. 24. Distal view of megaspore. Scale as in Fig. 20. Fic. 25. Equatorial view of microspore. Scale as in Fig. 21 64 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) specimens from all other months contained both mega- and microspores, sug- gesting this species is evergreen and fertile year-round. This species is the most widely distributed (or at least most widely collected) member of this group. Isoétes colombiana (T. C. Palmer) H. P. Fuchs and I. steyermarkii H. P. Fuchs nom. nud. are synonyms of I. karstenii. Palmer (1929) originally separated I. lechleri var. colombiana from I. lechleri var. lechleri on the basis of the absence of an equatorial ridge on the megaspores, the small size of the microspores, and the corneous leaf apices. All of these characters agree with the current circumscription of J. karstenii. In addition, the type collection of I. lechleri var. colombiana is a mixture of fertile J. karstenii and plants of apparent hybrid origin. The hybrids are the specimens that contain ornamented and misshapen megaspores as noted in Palmer’s (1929) protologue. Isoétes steyermarkii was invalidly published by Fuchs-Eckert (1982) and no description was given. Ex- amination of the basis for this taxon (Steyermark 55904) reveals no characters that differentiate it from I. karstenii; in fact this collection is nearly identical to the type of I. karstenii. Finally, the type specimen of I. karstenii as cited by Braun (1862) is “In einem ungefahr 8000’ hoch See auf dem Péramo im Gebirge von Mérida (Ven- ezuela) ganz unter Wasser im Jahre 1853 von Prof. H. Karsten entdeckt,. . .” The specimen at Berlin which was annotated as the holotype by H. P. Fuchs bears the label “Jsoétes karstenii A. Braun. Venezuela. Im Gebirge von Mérida 8000’ hoch in einem See ganz unter Wasser. C. Dr. Karsten etc. 1857.” Al- though the locality information on this label closely matches that of Braun’s (1862) description, there is a discrepancy regarding the date of collection: 1853 in the description and 1857 on the specimen label. Stafleu and Cowan (1979) indicate that Karsten was in South America from 1848 to 1856, but returned to Berlin in 1856 where he remained until 1868. Thus, he could not have collected this specimen in 1857. A potential explanation for this discrepancy could be that 1853 was the date of collection, while 1857 was the date the material was accessioned into the Berlin herbarium. Additional specimens examined: COLOMBIA. Departamento Arauca: Cordillera Oriental, extremo sur de la Sierra Nevada del Co- cuy, Laguna el Amarillal, alt. 4065 m, 23 Mar 1977, Cleef & van der Hammen 10383 (U); Sierra Nevada del Cocuy, Laguna de la Plaza, alt. 4550 m, 20-30 Jan 1959, van der Hammen & Gonzalez 1374 (COL); Sierra Nevada del Cocuy, Laguna la Plaza, alt. 4150 m, Mar 1977, van der Hammen 4753 (COL, U). Departamento Boyaca: Sierra Nevada del Cocuy, Alto Valle Lagunillas, Laguna Cuadrada, alt. 4060 m, 26 Sept. 1972, Cleef & Florschtitz 5590 (COL, U); Sierra Nevada del Cocuy, Paramo Concavo, 3 km al N del morro Pulpito del Diable, alt. 4410 m, 27 Feb 1973, Cleef 8599 (U); Sierra Nevada del Cocuy, Alto Valle Lagunillas, medio km al WNW de la Laguna Pintada, alt. 4000 m, 4 Mar 1973, Cleef 8782 (COL, U); Municipio de El Cocuy, valle del rio Lagunillas, Sierra Nevada del Cocuy, Laguna La Pintada, alt. 3800 m, 23 May 1993, Small, Gonzalez & Ruiz 141- 143 (COL, MU); Municipio de El Cocuy, valle del rio Lagunillas, Sierra Nevada del Cocuy, Laguna La Cuadrada, alt. 3900 m, 23 May 1993, Small, Gonzalez & Ruiz 144 (COL, MU). Departamento Cundinamarca: Paéramo de Sumapaz, Chisaca, Laguna Negra, alt. 3750 m, 11 Dec 1971, Cleef 174 (COL, U); Km 31, south of Usme, Laguna Chisaca, alt. 3620 m, 3 Jan 1985, Keeley & Keeley 7908 (COL, MU); Km 31, south of Usme, Laguna Chisaca, alt. 3620 m, 3 Jan 1985, Keeley & Keeley 7909 (COL, MU); Small laguna adjacent to north side of road at km 31 between Laguna Chisaca and Laguna Larga, alt. 3650 m, 10 Dec 1985, Keeley & Keeley 10076 (MO, MU); Unnamed laguna on SMALL & HICKEY: ISOETES KARSTENII COMPLEX 65 opposite side of road from largest laguna at km 30.5, Paramo de Sumapaz, vicinity of Laguna Chisaca, alt. 3650 m, 10 Dec 1985, Keeley & Keeley vritt (MO, MU); Péramo de Sumapaz, laguna at km 30.5 opposite side of road from Laguna Chisaca, alt. 3700 m, 15 Dec 1985, Keeley & Keeley 10093 (MU); Paéaramo de Sumapaz, vicinity of Laguna Chisaca, laguna adjacent to road at km 31, alt. 3620 m, 7 Jul 1986, Keeley & Keeley 11037 (MU); Paramo, alt. 3700 m, Apr 1963, Saravia 2533 (COL). Departamento Meta: Paramo de Sumapaz, Laguna La Guitarra, alt. 3460 m, 21 Jan 1972, Cleef 832 (COL, U); sti de Sumapaz, Laguna La Primavera, alt. 3525 m, 25 Jan 1972, Cleef 1017 (COL, U); Paéramo de Sumapaz, Cerro Nevado del Sumapaz, alt. 4050 m, 30 Jan 1972, Cleef 1351 (COL, U); Péramo de Sumapaz, Cerro Nevado del Sumapaz, alt. 4090 m, 13 Jan 1973, Cleef 7766 (COL, U); Péaramo de Sumapaz, Laguna Los Sitiales, alt. 3620 m, 22 Jan 1973, Cleef 8248 (COL, U); Macizo de Sumapaz, Boqueron del Palacio, norte de Cerro del Nevado, alt. 3900 m, 5 Jul 1981, Diaz, et al. 2558 (COL); Macizo de Sumapaz, Laguna La Guitarra, alt. 3370 m, 2 Jun 1981, Diaz & Rangel 2390 (COL). Departamento Narifio: Cumbal, Laguna La Bolsa, alt. 3400 m, 21 Jan 1973, Hagemann & Leist 1968 (COL); Laguna Negra en vecindades de Pasto, alt. 3520 m, 17 Jul 1973, Leist & Méhle 2212 (COL) VENEZUELA. Estado Mérida: Laguna Grande de Mucubaji, alt. 3535 m, 23 Dec 1984, Keeley & Keeley 7861 (MU); Laguna Grande de Mucubaji, alt. 3535 m, 23 Dec 1984, Keeley & Keeley 7862 (MU); Laguna Negra, alt. 3430 m, 23 Dec 1984, Keeley & Keeley 7865 (MU); Laguna de los Patos, alt. 3670 m, 23 Dec 1984, Keeley & Keeley 7868 (MU); Laguna Negra, alt. 3400 m, 22 Nov. 1968, Oberwinkler & Oberwinkler 13685 (M); Laguna St. Barbara, alt. 3700 m, 11 Jan 1969, Oberwinkler & Oberwinkler 14272 (M); Laguna St. Barbara, alt. 3700 m, 11 Jan 1969, Oberwinkler & Oberwinkler 14274 (M); Small pond ca. 8 km N of Pico de Aguila on road to Pinango, alt. 3500 m, 15 Mar 1992, Small & Hickey 4 (MU); Small pond ca. 15 km N of Pico de Aguila on road to Pinango, alt. 3500 m, 15 Mar 1992, Small & Hickey 5 (MU); Small pond ca. 15 km N of Pico de Aguila on road to Pinango, alt. 3500 m, 15 Mar 1992, Small & Hickey 6 (MU); Laguna Negra, alt. 3500 m, 17 Mar 1992, Small & Hickey 7 (MU); Laguna Mucubaji, alt. 3500 m, 17 Mar 1992, Small & Hickey 8 (MU); Near upper limits of paramo, around small lake between Chachapo and Los Apartaderos, near El Aguila, 3930 m, 15 Apr 1944, Steyermark 55904 (F, US, VEN); Laguna Grande de Mucubaji, alt. 3500 m 19 Sep 1961, Tryon & Tryon 5839 (GH). 4b. Isoétes karstenii var. anomala (T. C. Palmer) R.L. Small and Hickey, comb. nov. Figs. 22-23. Isoétes lechleri var. anomala T. C. Palmer, Amer. Fern J. 22: 130-131. 1932. Isoétes rimbachiana H. P. Fuchs Proc. Kon. Ned. Akad. Wetensch. C. 85(2): 259. 1982. Type: “Ecuador. Mt. Chimborazo. Submerged in water of shallow pond. Rooting in mud. 4200 m.” Rimbach 171 (Holotype US!) Isoétes karstenii var. anomala differs from I. karstenii var. karstenii only in the character of velum coverage: 25-75% in var. anomala, >90% in var. kar- stenii. In Palmer’s (1932) description of this taxon (as I. lechleri var. anomala) he distinguishes it from other “I. Jechleri’” (of which he considered I. karstenii to be a synonym) based primarily on the presence of a two-layered velum which incompletely covered the sporangium. Examination of the holotype, however, revealed no indication of a two-layered velum on the sporophylls we observed. The presence of a “multilayered velum” has been reported else- where (e.g., Isoétes prototypus, Britton and Goltz, 1991). Initial investigations indicate that the apparent multilayered velum is actually due to disintegration of the mesophyll of the velum and separation of the upper and lower epider- mal layers resulting in what appears as a double velum (Hickey, unpublished data). Regardless, this taxon appears unique in the extent of velum coverage. 66 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) It is otherwise nearly identical to the type of I. karstenii, as noted by Palmer (1932). This taxon is known only from the type locality (Mt. Chimborazo, Ec- uador) although Rodriguez (1955) cites one specimen from Laguna Negra, Mé- rida, Venezuela as I. Jechleri var. anomala. 5. Isoétes palmeri H. P. Fuchs. Proc. Kon. Ned. Akad. Wetensch. C. 84 (2): 168-173. 1981. Figs. 24-25. Type: “Venezuela; Laguna Grande de Apartaderos, Mérida, floating on water, 21 January 1929” Pittier 13242 (Holotype US!; Isotypes G!, M!, MO!, VEN!). Isoétes cleefii H. P. Fuchs Proc. Kon. Ned. Akad. Wetensch. C. 84 (2):177-181. 981. Type: “Colombia, Cundinamarca: Péramo de Sumapaz, péramo y bosque alto-andino cerca de Lagunitas al S. de San Juan. Laguna Gober- nador: orilla W.” Cleef 8308 (Holotype COL!; Isotype COL!). Corm globose to horizontally elongate, two lobed, 1.3-6.0 mm (x = 2.9) high, 3.0-8.5 mm (x = 5.0) wide; dichotomously branching roots arising from the circumbasal fossa. Leaves 5-25, flexuous, to 350 mm long, 5-8 mm wide at base, 0.5 to 1.5 mm wide at mid-length, leaf length/width ratio 118-350 (x = 241); alae chartaceous, dark brown to blackened distally, pale proximally where leaf was imbedded in substrate, extending 10-27% of total leaf length, apices acute to truncate; subula essentially terete in cross-section, pale green, apices blunt, often spatulate, distal 1.0-2.5 mm of apex sclerified and black- ened; peripheral fibrous bundles, stomates, scale leaves and phyllopodia ab- sent. Sporangia elliptic to obovate, basal, unpigmented, 2.5-5.9 (x = 3.8) mm long, 1.9-4.0 (x = 2.8) mm wide. Velum complete, covering 100% of the spo- rangial surface (rarely slightly less). Ligule deltate, auriculate, to 1 mm high, to 1 mm wide. Labium absent or reduced to an extremely small transverse ridge on lower lip of the foveola. Megaspores white, lustrous, 380-550 (x = 460) jm in diameter, laevigate on both proximal and distal faces. Microspores cocoa brown, 26—36 (x = 30.4) um in length, 17.9-26.4 (x = 21.3) wm wide, echinate. Chromosome number: 2n = 44. Isoétes palmeri grows submerged in permanent pools, ponds, and lakes be- tween 3340-4435 m. It is distributed in the eastern cordillera of the northern Andes from Estado Mérida, Venezuela, through the Departamentos Arauca, Boyaca, Cundinamarca and Meta of Colombia (Fig. 8). The presence of both microspores and megaspores in all months for which collections are available (nine out of twelve) suggests that this species is ev- ergreen and fertile year-round. Isoétes palmeri'’s affiliation with the I. karstenii complex is apparent by its laevigate megaspores, acute to truncate alae apices, and corneous leaf tips. For the most part, I. palmeri and I. karstenii are easily differentiated morphologi- cally; I. palmeri has long, slender leaves which appear flexuous, and JI. kar- stenii has shorter, wider leaves which usually appear rigid and stiff. Occa- sional collections appear intermediate between these two extremes. These col- SMALL & HICKEY: ISOETES KARSTENII COMPLEX 67 lections often represent juvenile individuals of either species, but may also include hybrids. Isoétes cleefii is a synonym of I. palmeri. Although the holotype and one of the paratypes of I. cleefii have leaves that are slightly longer and wider than typical I. palmeri, these individuals are merely extreme variants within a wide range of intraspecific variation. Both I. palmeri and I. cleefii were published by Fuchs-Eckert (1981a,b) at the same time, thus neither name has priority. We select the name I. palmeri to represent this species because the type col- lection is more representative of the species. Additional speciemens examined: COLOMBIA. D t A Si Nevada del C , Cabeceras de la Quebrada E] Playon alt. 4245 m, 11 Mar 1973, ape’ 9088 (COL, U). Departamento Boyaca: Paéramos al NW de Belen, ca. 600 m NE de la Laguna el Alcohol, alt. 3850 m, 29 Feb 1972, Cleef 2059 (COL, U); Péramo de la Rusia, NW-N de Duitama, alt. 3570 m, 13, Dec 1972, Cleef 7144 (COL, U); Sierra ands del Cocuy, Péramo Concavo, alt. 4435 m, 28 Feb 1973, saocd 8647 (COL, U); Paramo de la Rusia, NW- N de Duitama, Laguna Negra, alt. 3745 m, 14 Dec 1972, Cleef 7221 (COL, U); Sierra act del Cocuy, Péramo Concavo, 2.5 km al N de morro Pali del Diablo, alt. 4435 m, 28 Feb 1973, Cleef 8648 (COL, U); Péramo entre Pefia Arnical y Alto de Mogotes; lagunita 0.5 km al SE de la Laguna Grande, alt. 3340 m, 1 Apr 1973, Cleef 9273 (COL, U); Paéramo entre Pena de Arnical y Alto de Mogotes, cerca de la Laguna Grande, alt. 3350 m, 10 Apr 1973, Cleef 9551 (COL, U); Péramos al Belen, vereda San Jose de la Montana, alt. 3725 m, 3 May 1973, Cleef 9699 (U); Péramo de Pisva, carretera Socha-La Punta km 77, Laguna Colorada, alt. 3425 m, 22 May 1973, Cleef 9898 (COL, U); Largest laguna, 17 km NW of Belen on road to San Jose de peace aes 3700 m, 1 Jan 1985, Keeley & Keeley 7887 (COL, MU). Departamento Cundinamarca: Paramo entre ean y San Cayetano, Laguna Verde, 3600 m, 30 Nov 1971, Cleef 80 (COL,U); Péramo de PUBS hoya la Laguna Larga, alt. 3750 m, 14 Dec 1 1971, Cleef & Jamarillo 262 (COL, U); Péramo de Palacio, Lagunas Buitrago, alt. 3600 m, 16 Dec 1971, Cleef 296 (COL, U); Paéramo de Sumapaz, Andabobos, alt. 3800 m, 9 Feb 1972, Cleef 1553 (COL, U); Péramo de Palacio, Lagunas de Buitrago, alt. 3580 m, 25 May 1972, Cleef 4112 (COL, U); Pa oO az, lagunitas al S de San Juan, La Gobernador, alt. 3815 m, 26 Jan 1973, Cleef 8308 (COL); Paramo de Chingaza, cerca de la “Palla” 10-20 Jan 1965, Huertas & Camargo 6025 (COL); Cordillera Oriental, Macizo de Sumapaz, Laguna de Chisaca, alt. 3680 m, 14 Apr 1958, Idrobo 2741 (COL); Laguna Seca, N of Zipaquire, alt. 3700 m, 30 Dec 1984, Keeley & Keeley 7876 (MU); Laguna Seca, N of Zipaquire, alt. 3700 m, 30 Dec 1984, Keeley & Keeley 7878 (COL, MU); Péramo de Sumapaz, vicinity of Laguna Chisaca, Laguna Larga, alt. 3650 m, 10 Dec 1985, Keeley & Keeley 10072 (MO, MU); Paéramo de Sumapaz, vicinity of Laguna Chisaca, Laguna Larga, alt. 3650 m, 10 Dec 1985, Keeley & Keeley 10073 (MO, MU); Pdéramo de Sumapaz, vicinity of Laguna Chisaca, small pools S of Laguna Larga, alt. 3650 m, 10 Dec 1985, Keeley & Keeley 10077 (MO, MU); Péramo de Sumapaz, vicinity of Laguna Chisaca, small pools ca. 100 m above SE end of Laguna Chisaca, alt. 3700 m, 13 Dec 1985, Keeley & Keeley 10088 (MO, MU); Paéramo de Sumapaz, Laguna Chisaca, alt. 3700 m, 13 Dec 1985, Keeley & Keeley 10090 (MU); Péramo de Sumapaz, vicinity of Laguna Chisaca, small pool ca. 200 m above SE end of laguna on NE side of road at km 30.5, alt. 3700 m, 13 Dec 1985, Keeley & Keeley 10091 (MU); Pdramo de Sumapaz, vicinity of Laguna Chisaca, laguna ca 100 m NE of agunita, alt. 3700 m, 13 Dec 1985, Keeley & Keeley 10092 (MO, MU); Paéramo de Sumapaz, agian of Laguna — aca, Laguna Larga, alt. 3620 m, 7 Jul 1986, Keeley & Keeley 11029 (MU); P. o de Sum vicinity of Laguna Chisaca, shallow pool on ridge SE of first lake, alt. 3620 m, 7 7 jal 1986, ee & Keeley 11039 (MU); Carretera Cogua-San Cayetano, Péramo de Laguna Seca, alt. 3660 m, 3 Aug 1970, Piedrahita 276 (COL). Departamento Meta: Péramo de Sumapaz, Laguna la Primavera, alt. 3510 m, 9 Jan 1973, Cleef 7556 (COL, U); Paramo de Sumapaz, Lagunas el Sorbedero y El Nevado, alt. 3635 m, 16 Jan 1973, Cleef re (COL, U); Péramo de Sumapaz, Laguna la Guitarra, alt. 3425 m, 23 Jan 1973, Cleef 8267 (COL 68 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) ACKNOWLEDGMENTS We thank the staff of the Herbario Nacional de Colombia, especially Favio Gonzalez and Maria pi Murillo, for assistance with field and herbarium work in Colombia; A. Clive Jermy for This work was supported by two sre se from the State of Ohio Academic Challenge Program and W.S. Turrell Herbarium Fund Grant # LITERATURE CITED BAKER, J. G. 1880. A synopsis of the species of Isoetes. J. Bot. Br. Foreign 18:65—70, 105-110. BRAUN, A. 1862. aie uber einige auslandische Arten der Gattung Isoétes. Verh. Bot. Ver. Prov. 333. BriTTON, D. M., and J. P. GOLTz. 1991. Isoetes prototypus, a new diploid species from eastern Canada. Can. J. Bot. 69:277—281. BRUNTON, D. F., and D. M. BrirTon. 1993. Isoetes prototypus (Isoetaceae) in the United States. BRUNTON, D. F., and D. M. BRITTON. 1996a. Taxonomy and distribution of Isoetes valida. Amer. Fern. J. 86:16—25. BRUNTON, D. F., and D. M. BRITTON. 1996b. The status, ns gepanet wie identification of Georgia Quillwort viatiny georgiana; Isoetaceae). Amer. Fern. J. 86:105—113. BRUNTON, D. F., D. M. BRITTON, and W. C. TayLor. 1994. ae i roar sp. nov. (Isoetaceae):a new pares from the southeastern United States. Castanea 2-21. Cox, P. A., and R. J. HICKEy. 1984. Convergent megaspore use a Isoetes. Amer. Naturalist 124:437—441., FUCHS-ECKERT, H. P. 1981a. Isoétes Palmeri H. P. Fuchs, eine neue Isoétes-Art des Péramo. Proc. Kon. Ned. Akad. Wetensch. C. 84 (2):165-174. FUCHS-ECKERT, H. P. 1981b. Isoétes psig H. P. Fuchs, eine weitere neue Isoétes-Art aus dem kolumbianischen Péramo. Proc. Kon. Ned. Akad. Wetensch. C. 84:175-182 FucHs-ECKERT, H. P. 1982. Zur dead annhes von Vorkommen und Verbreitng der siidamer- ee Isoétes-Arten Proc. Kon. Ned. Akad. Wetensch. C. 85:2 FUCHS-ECKER . 1992. Supplementum ad indicem Isoétalium. ee 12: 99-159. Gomez, L. D. 1980. Vegetative reproduction in a Central American Isoétes (Isoétaceae). Its mor- ho. ny ee reno and taxonomical implications. Brenesia 18:1-14 Hickey, R. J. 1 Chromosome numbers in neotropical Isoétes. Amer. HICKEY, R. J. si aon studies of neotropical Isoétes. Ph.D. aoniain, 8s sae of Connecticut, HICKEY, R. J. 1994. 28. solace, in Esebehicraas dey of Peru, part VI, (R. M. Tryon and R. G. Stoltze, eds.) aie g oy Bot. N.S. 3 Hickey, R. J., W. C. TAYLOR, ioe “s ma LUEBKE. 1989. The species concept in Pteridophyta with csgeoes sei to Isoétes. Am. Fern. J. 79:78-89. Korr, L. 5., D. M. BRITTON. 1983. Spore morphology and taxonomy of Isoétes in northeastern No 6 Maudie. Can. J. Bot. 61:3140-31 Kort, L. S., and D. M. BRITTON. 1985. Role of morphological pera ata of leaves and the — region in the taxonomy of Isoétes in northeastern North America. Amer. Fern. J. ‘44-55 ainan G. HL 1859. Filices Lechlerianae, fasc. 2 MOTELAY, L., and. A. VENDRYES. 1882. Seaiuahie fo Jandtonn: Act. Soc. Linn. Bordeaux 36:309-— 405. MUSSELMAN, L. J., and D. A. KNEPPER. 1994. Quillworts of per aci Amer. Fern. J. 84:48-68. PALMER, T. C. 1929. lncecions Lechleri Mett. Amer. Fern. J. 1 PALMER, T. C, 1932. More about Isoetes Lechleri Mett. ioe dss J. 22:129-132. PFEIFFER, N. E. 1922. Monograph of the Isoetaceae. Ann. Missouri Bot. Gard. 9:79-232. SMALL & HICKEY: ISOETES KARSTENII COMPLEX 69 RADFORD, A. E., W. C.DICKISON, and C. R. BELL. 1972. Vascular Plant Systematics. Univ. North Carolina, Chapel Hill. RODRIGUEZ, G. 1955. Revision del genero Isoetes en Venezuela. Bol. Mus. Cienc. Nat. I ROHLF, F. J. 1993. NTSYS-pc: numerical taxonomy and multivariate analysis system, ante 1. 80. Exeter seit Applied Biostatistics, es Setauket, NY. pied F. J., P. THOMPSON, and E. SHEFFIELD. 1993. Tiploid Isoetes echinospora (Isoetaceae; Pter- dophyta) in northern England. Fern cat 14:2 Rury, P. M. 1978. A new and unique, sateomibeay op s-grass (Isoétes) from Georgia. Amer. Fern STAFLEU, F. A., and R. S. Cowan. 1979. Taxonomic Literature, 2nd ed. Vol. II: H-Le. Regnum Vegetable Vol. 98. Bohn, Scheltema & Holkema, Utrecht dr. W. Junk b.v., Publishers, The ague. i W. C., and R. J. Hickey. 1992. Habitat, evolution and speciation in Jsoetes. Ann. Missouri. t. Gard. 79:613-622. es W. C., and N. T. LUEBKE. 1986. Germinating spores and growing sporelings of aquatic Isoétes. Amer. Fern. J. 76:21-—24. TAYLOR, W. C., N. T. LUEBKE, D. M. BRITTON, R. J. Hickey, and D. F. BRUNTON. 1993. seater) Pp. 64-75 in: Flora of North America, Vol. 2, Pteridophytes and Gymnosperms. Oxford U versity Press, New York TAYLOR, W. C., N. T. LUEBKE, sia M. B. SMITH. 1985. Speciation and arenes in North Amer- ican eee age Roy. Soc. Edinburgh Sect. B (Biol. Sci.) 86B: _ UNDERWOOD, L. M. 1 The distribution of Isoetes. Bot. Gaz. April: 8 VARESCHI, V. 1968. Sie Rn in: T. Lasser (ed.), Helechos, Flora de eo ae Vol 1(1 WaTANABE, M., M. TAKAMIYA, T. MATSUSAKA, and K. ONO. 1996. Biosystematic studies on the genus Isoetes (Isoetaceae) in Japan. III. Variability within qualitative and quantitative mor- phology of spores. J. Plant Res. 109:2 WenrR, U. 1922. Anatomie und Systematik a Gattung Isoétes L. Nova Hedwigia 63:219-262. American Fern Journal 91(2):70—72 (2001) SHORTER NOTE Additions and Corrections to the Pteridophyte Flora of Northeastern Argen- tina.—Two new records as well as material to confirm two non-vouchered records were collected by colleagues at the Instituto Darwinion and Parques Nacionales during botanic expeditions to Misiones in northeastern Argentina. Misiones is a strip of land between the Uruguay and Parana rivers, with lon- gitudinal mountains reaching approximately 800 m in elevation. This region is under exploration in order to study and document the floristic composition of the southernmost subtropical forests in South America. These records are elements of the Araucaria angustifolia forests and rocky “campos”, two pri- mary communities that should be maintained and preserved within national and state reserves. Megalastrum crenulans (Fée) A. R. Smith & R. C. Moran, Amer. Fern J. 77: 127. 1988. Aspidium crenulans Fée, Cr. vasc. Brés. 1: 139, t. 47, f. 1, 1869. Dryopteris crenulans (Fée) C. Chr. DESCRIPTIONS AND ICONOGRAPHIES.—Fée, op. cit.; Sehnem, Aspididceas, 186, f. 45, in R. Reitz (ed.) Flora Ilustrada Catarinense Parte I (ASPI) Itajaf, SC, Brasil, 1979. DISTRIBUTION AND EcoLoGy.—This fern grows in southern Brasil, Paraguay and has been cited also as occuring in the Venezuelan Guyana (Smith, Pterido- phytes pp. 1-334, in A. Steyermark, P. E. Berry & B. K. Holst (Gen. Eds.) Flora of the Venezuelan Guayana 2, Missouri Botanical Garden, 1995). Here it is recorded for the first time for Argentina, elevating the number of species of Megalastrum in this country to five (Ponce, Monogr. Syst. Bot. Missouri Bot. Gard. 60: 1-79. 1996). It inhabits subtropical Araucaria forests. This species is characterized by the occurrence of glandular trichomes mixed with simple trichomes, both of which are short and dense on the abaxial face, bulliform scales on the costule, and large, brown and glandular indusia. The latter character is not common in Megalastrum. The forms crenulans and glandulosa (Rosent.) C. Chr. were described by Christensen (Kongel. Danske Vidensk. Selsk. Skr. Naturvidensk. Afd. 8(6): 3-132. 1920); material from Mi- siones is referable to the latter. A revision of the genus is necessary to clarify species boundaries. SPECIMEN STUDIED.—ARGENTINA. Misiones, Dpto. San Pedro, Parque Provincial Cruce Caballero, 690 m, 26° 31’ S, 53° 59’ W, 15-IV-1996, F. O. Zuloaga, O. Morrone & M. Mulgura 5551 (SI). Elaphoglossum pachydermum (Fée) T. Moore, Index Fil. 12. 1857. Acrostichum pachydermum Fée, 2eme. Mém. Foug. 47. 1845. SHORTER NOTE 71 DESCRIPTIONS AND ICONOGRAPHY.—Sehnem, Aspididceas, 24, Fig. 5.2, in R. Reitz (ed.) Flora Ilustrada Catarinense Parte I (ASPI) Itajai, SC, Brasil, 1979. DISTRIBUTION AND ECOLOGY.—This species grows in Brasil, Paraguay and is here reported for Argentina. It is rupiculous, occurring on rocky promontories and steep slopes in montane areas. This species has erect to short creeping rhizomes and fimbriate scales from the base to the middle of the petiole. The sterile fronds are longer than the fertile fronds, the lamina is linear-oblong, attenuate-decurrently based, coria- ceous in texture, and with minute, reddish to brown, stellate trichomes on the abaxial surface. Elaphoglossum pachydermum is in the E. latifolium (Sw.) J. Sm. complex, and was thought to be confined to the Antilles (Mickel, pers. comm.). Previ- ously it was recorded for Argentina by Hassler (Trab. Inst. Bot. Farmacol. 45: 1-102. 1928) as E. latifolium, but the voucher specimem (Parodi 84) could not be located for verification. With this report, E. pachydermum is documented for the first time for Ar- gentina. The related species, E. macrophyllum (Mett. ex Kuhn) H. Christ and E. macahense (Fée) Rosenst., are known from Brasil. SPECIMEN STUDIED.—ARGENTINA. Misiones. Dpto. San Ignacio, Parque Provincial Teyti Cuaré Pefién del Teyti Cuaré, rupicola en zonas protegidas, 28/09/1998, Biganzoli, F. & D. Giraldo-Cafias 394 (SI). Cheiroglossa palmata (L.) C. Presl, Suppl. Tent. Pterid. 57. “1845” 1846. Ophioglossum palmatum L. Sp. Pl. 1063. 1753. DESCRIPTION AND ICONOGRAPHY.—Tryon & Stolze (Fieldiana, Bot. n.s. 20: 9, f. 2. 9) DISTRIBUTION AND EcoLocy.—This fern grows in southern Florida (USA), the Antilles, southern Mexico to Brazil, and northeastern Argentina. It is also dis- junct in Madagascar and southeastern Asia. It is dae cane cis suberect to pendant on trunks in subtropical primary and marginal fores Wagner’s treatment of the Ophioglossaceae See pp. 44-48, in G. Davise et al. (Eds.) Flora Mesoamericana. Vol. 1, Univ. Nac. Aut6noma de México, 1995) is accepted here. Wagner considers Cheiroglossa a genus justi- fiably separated from Ophioglossum. This species was cited by Molfino (Phy- sis, Buenos Aires, 8: 259-260. 1925) and Lichtenstein (Darwiniana 6: 380-441. 1944) as being in Misiones, but without documenting herbarium specimens. SPECIMEN STUDIED.—ARGENTINA. Misiones. Dpto. Gral. Manuel Belgrano: Deseado, Reserva Pri- vada Caé-pora, epifita sobre “Sotacaballo’’, invierno/1996, Chaves & Chebez s.n. (SI). Dicranopteris flexuosa (Schrad.) Underw., Bull. Torrey sige orig 34: 254. 1907. Mertensia flexuosa Schrad., Gott. gel. Anz. 1824: 863. DESCRIPTIONS AND ICONOGRAPHS.—Fée, Crypt. vasc. bras. 1: 199, t. 72, f. 1 (sub Mertensia scalpturata Fée); 200, t. 73, f. 2 (sub M. spissa Fée) 1869. Sehnem, 72 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 2 (2001) Gleiqueniaceas, in R. Reitz (ed.) Flora IJustrada Catarinense Parte I (GLEN): 25, f. 9. 1970 DISTRIBUTION AND ECoLoGy.—Southern Alabama (USA), the Antilles, southern Mexico, Central America, Colombia and Guyanas to Bolivia, southern Brazil, Paraguay, and Argentina. Terrestrial and rupiculous at shady and moist sites in rocky “campos” and marginal forests. Dicranopteris flexuosa is the correct name for the neotropical counterpart of the paleotropical D. linearis (Burm.) Underw. Roth (Faculty of Natural Scienc- es, University of Gotreborg, Sweeden, 1986), in her thesis on the neotropical Gleicheniaceae, suggested that D. flexuosa may represent a subspecies of D. linearis, but that combination has never been made. SPECIMEN STUDIED.—ARGENTINA. Misiones, Dpto. San Ignacio, Parque Provincial Teyui Cuaré. Pefién del Teyt Cuaré, rupicola, ocasional en zona protegida y htimeda, localmente abundante, 28/09/1998, Biganzoli, F. & Giraldo-Carias 397 (SI). Thanks to colleagues and institutions of the Inventario Florfstico de la Prov- incia de Misiones project (PMT-PICT 01511) directed by F. Zuloaga and sup- ported by the Consejo Nacional de Investigaciones Cientfficas y Técnicas (CONICET) Argentina.—MOonica PONCE, Instituto de Botanica Darwinion, La- bardén 200, B1642HYD San Isidro, Argentina. NT INFORMATION FOR AUTHORS: Authors are encouraged to submit manuscripts pertinent to pteridology for pub- lication in the American Fern Journal. Manuscripts should be sent to the Editor. Acceptance of papers for publication depends on merit as judged by two or more referees. Authors are encouraged to contribute toward publishing costs; however, the payment or non-payment of page charges will affect neither the acceptability of manuscripts nor the date of publication. Authors should adhere to the following guidelines; manuscripts not so prepared may be returned for revision prior to review. Submit manuscripts in triplicate (xerocopies acceptable), including review copies of illustrations and originals of illustrations. After review, submission of final versions of manuscripts on diskette (in PC- or Mac-compatible formats) is strongly encouraged. Use standard 8% by 11 inch paper of good quality, not “erasable” paper. Double space manuscripts throughout, including title, authors’ names and addresses, short, informative ab- stract, text (including heads and keys), literature cited, tables (separate from text), and figure captions (grouped as consecutive paragraphs separate from figures). Arrange parts of manuscript in order just given. Include author’s name and page number in upper right corner of every sheet. Provide margins of at least 25 mm all around on typed pages. Do not submit right-justified copy, avoid footnotes, and do not break words at ends of lines. Make table headings and figure captions self-explanatory. Use S.I. (metric) units for all measures (e.g., distance, elevation, weight) unless quoted or cited from another source (e.g., specimen citations). For nomenclatural matter (i.e., synonymy and typification), use one paragraph per basionym (see Regnum Veg. 58:39—40. 1968). Abbreviate titles of serial publi- cations according to Botanico-Periodicum-Huntianum (Lawrence et al., 1968, Hunt Botanical Library, Pittsburgh) and its supplement (1991). References cited only as part of nomenclatural matter are not included in literature cited. For shorter notes and reviews, omit the abstract and put all references parenthetically in text. Use Index Herbariorum (Regnum Veg. 120:1—693. 1990) for designations of her- Illustrations should be proportioned to fit page width with caption on the same page. Provide margins of at least 25 mm on all illustrations. For continuous-tone illustrations, design originals for reproduction without reduction or by uniform amount. In composite blocks, abut edges of adjacent photographs. Avoid com- bining continuous-tone and line-copy in single illustrations or blocks. Coordinate sequence and numbering of figures (and of tables) with order of citation in text. Explain scales and symbols in figures themselves, not in captions. Include a scale and reference to latitude and longitude in each ma Proofs and reprint order forms are sent to authors by the printer. Authors should send corrected proofs to the editor and reprint orders to the printer. Authors will be assessed charges for extensive alterations made after type has been set. For other matter of form or style, consult recent issues of American Fern Jour- nal and The Chicago Manual of Style, 14th ed. (1993, Univ. Chicago Press, Chicago). Occasionally, departure from these guidelines may be justified. Authors are encouraged to consult the editor for assistance with any aspect of manuscript preparation. Papers longer than 32 printed pages may be sent to the Editor of Pteridologia (Memoir Editor, see cover 2). 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 postpaid. 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 postpaid. Send your order with a check or money order to: American Fern Society, Inc., c/o U.S. National Herbarium MRC-166, Smithsonian Institution, Washing- ton, DC 20560. AMERICAN FERN JOURNAL ON MICROFICHE Volumes 1—61 of the American Fern Journal are available as archival quality, silver positive microfiches. Single volumes or the entire run may be purchased. The fiches are easily read with 10 or greater magnification (using a dissecting microscope and transmitted illumination or a fiche reader). Silver negative micro- fiches of vols. 1-50 are also available. The price is $4.00 per volume or $244.00 per set of 61 volumes, postpaid. Send your inquiry or order with a check or money order to: American Fern Society, Inc., c/o Dr. Jamés D. Montgomery, Ecology III, Inc., R.D. 1, Box 1795, Berwick, PA 18603. VISIT THE AMERICAN FERN SOCIETY’S WORLD WIDE WEB HOMEPAGE: http://www.amerfernsoc.org/ QR) A399 AMERICAN FERN nant J O U R N A ' July-September 2001 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY SPECIAL ISSUE The Evolution and Diversification of the Lycopods Proceedings of a Symposium Held August 1999 at the XVI International Botanical Congress Missouri Botanical Garden St. Louis, Missouri Jointly Sponsored by: VI International Botanical Congress Green Plant fgpicuate Research eRe resorts Group U.S. Department of Agricultur National Science Romndatinn. Department of Energy American Fern Society Symposium Organizers: Niklas Wikstr6m The Evolution and Diversification of the Lycopods W. Carl Taylor 73 Early Lycophyte Evolution Patricia G. Gensel and Christopher M. Berry 74 Isoetalean Lycopsid Evolution: from the Devonian to the Present Kathleen B. Pigg 99 The Triassic Lycopsids Pleuromeia and Annalepis: Relationships, Evolution, an Léa Grauvogel-Stamm and Bernard Lugardon 115 Diversification and Relationships of Extant Homosporous Lycopeds _—Niklas Wikstrém 150 The Utility of Nuclear ITS, a LEAFY H Intr d Chi last atpB-rbcL Spacer Region Data in Phylogenetic fe and Species Delimitation i in Isoétes Sara B. Hoot and W. Car! Taylor 166 The American Fern Society Council for 2001 BARBARA JOE HOSHIZAKI, 557 N. Westmoreland Ave., Los Angeles, CA 90004-2210. President CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS 66045-2016. oo W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI 53233-147 < etary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37996-1110. Treasurer VID B. LELLINGER, 326 West St. NW., Vienna, VA 22180-415 Membership Secretary a D. MONTGOMERY, Ecology III, R.D. 1, Box 1795, oe PA 18603-' ase Back Issues Curator R. JAMES HICKEY, Botany Dept., Miami University, Oxford, OH, 45056 Journal Editor DAVID B. LELLINGER, U.S. National Herbarium MRC-166, Smithsonian Institution, Basen er DC 20560-0166. Memoir Editor CINDY JOHNSON-GROH, Dept. of psi Gustavus Adolphus College, 800 W. College Ave., St. Peter, MN 56082-1498. Bulletin Editor American Fern Journal R. JAMES HICKEY y Departm ami University, Tae OH 45056 ph. (513) cae: paehera hickeyrj @muohio.edu ASSOCIATE EDITORS GERALD J. GASTONY 2.0.2. ares of ao Indiana University, Bl a IN cohesion CHRISTOPHER H. HAUFLER Dept. of — University of Kansas, Lawrence, KS 6604 ROBBIN C. MORAN New York Botanical cnc eee NY — se JAMES H. PECK Dept. of Biology, University of Arkansas—Little Rock, 2801 S. University ie. Little Rock, AR 72204 The “American Fern Journal” (ISSN ies sib is an illustrated quarterly devoted to the ee study of ferns. It is owned by the American Fern Society, and sas at ae West St. NW., VA 22180-4151. Periodicals postage paid at ee VA, and Claims for missing issues, made 6 months (domestic) to 12 sti (forei en) after the date = issue, and orders for back issues should be addressed to Dr. James D. Montgomery, Ecology III, Berwick, PA 18603-9801. Paes of address, dues, and applications for membership should be sent to the Membership eR Rae RAAT sh i h o the Secretary. Subscriptions $20.00 gross, “$19. 50 net if paid sane an 7 ag pega fee $0.50); sent free to of the American Fern Society (annual dues, $15.00 + $5.00 mailing surcharge beyond “life membership, $300.00 + $140.00 mailing surcharge a U.S.A Back volumes are available for most years as printed issues or on microfiche. Please contact the Back Issues Curator for prices and availability pac eee Send address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA 22180-4151. cS in a: FIDDLEHEAD FORUM The editor of the Ries of the American Fern Society welcomes contributions from members and non-members, includin scellaneous notes, offers to —— or purchase materials, personalia, horticultural notes, and reviews of non-technical books on fi SPORE EXCHANGE Stephen McDaniel, 1716 Piermont Dr., Hacienda Hts., CA sa De is Fad i Pit and lists of available spores sent on request. _http://www.amerfern: soc.org/sporex GIFTS AND BEQUESTS ifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Back issues of the Joumal and h ther gift 1 y 1 4 a tax -deductible. Inquiries should be addressed to the Secretary. American Fern Journal 91(3):73 (2001) The Evolution and Diversification of the Lypopods The lycopods, a diverse and ancient group of pteridophytes, form a major lineage of vascular plants. This lineage includes three orders, Lycopodiales, Selaginellales, and Isoetales. Each order contains a single extant family, but the group has a long and diverse fossil record beginning in the Devonian or possibly late Silurian. This fossil record is replete with many curious forms that have long captured the interest and imagination of those interested in plant evolution. Although lycopods have attracted the attention of paleobota- nists and morphologists for many years, it is only within the last few years that systematists have initiated studies to critically evaluate taxa and their modes of speciation. While there is agreement that the lycopods form a clade distinct from all other vascular plants, there is some argument in the interpre- tation of the diversity and evolution within the lycopod lineage. This is an exciting time for systematic biologists. We are beginning to uncover, discover, and explore the modes and mechanisms of divergent and reticulate evolution in the lycopods, revealing their incredibly long history and amazing diversity. The following five papers are facsimiles of oral presentations given on 4 August 1999, in a symposium entitled “The Evolution and Diversification of the Lycopods.” This symposium, a part of the XVI International Botanical Con- gress held in St. Louis, Missouri, brought leading researchers together to pre- sent their current findings relevant to the phylogeny of the lycopods. Niklas Wikstrém, Léa Grauvogel-Stamm, and Carl Taylor organized the symposium. The authors express their gratitude to Brent D. Mishler, the Green Plant Phylogeny Research Coordination Group, the U.S. Department of Agriculture, the National Science Foundation, the Department of Energy, and the organizers of the XVI International Congress for their assistance and support. The authors are also grateful to the American Fern Society, Inc. for sponsoring the publi- cation of their papers.—W. Carl Taylor, Department of Botany, Milwaukee Pub- lic Museum, Milwaukee, WI 53233 MISSOURI BOTANICAL nFC 3 1 2001 GARDEN LIBRARY American Fern Journal 91(3):74—98 (2001) Early Lycophyte Evolution PATRICIA G. GENSEL Department of Biology, University of North Carolina, Chapel Hill, NC 27599 CHRISTOPHER M. BERRY Department of Earth Sciences, Cardiff University, Cardiff, CF10 3YE, Wales, United Kingdom ABSTRACT.—Lycophytes, comprising the groups historically known as the lycopsids and zostero- phylls, have the longest history of any group of vascular land plants. The early evolution of the WHAT Is A LYCOPHYTE AND WHEN DID LYCOPHYTES APPEAR? The lycophytes ( = club mosses and related plants) are regarded as a distinct lineage of vascular plants with a lon evolutionary history. Considerable di- versity is evident for much of their existence, although only four (or 7+ if one accepts either Ollgard’s (1987) or Wagner and Beitel’s (1992) splitting of Ly- copodium into several genera) herbaceous genera are recognized today (Lyco- podium (and segregates), Phylloglossum, Selaginella, and Isoetes). Club mosses have long been defined as plants with microphyllous leaves, axillary or adax- ially borne reniform sporangia with dehiscence along their distal margin, and mostly exarch xylem maturation; such plants are now included in the class Lycopsida (“lycopsids”) by Kenrick and Crane (1997). Extinct taxa attributed to Zosterophyllopsida (and possibly including barinophytes) are considered closely related to lycopsids, either as ancestral or sister taxa (Gensel, 1992; Hueber, 1992; Bateman, 1992; Kenrick and Crane, 1997), being similar in spo- rangium shape and xylem attributes but differing from lycopsids in lacking leaves and leaf-borne sporangia. Additionally they have vascularized sporan- gial stalks (Fig. 1). REDEFINITION OF “LYCOPHYTES” AND “LYCOPSIDS”.—Several workers emphasize the monophyly of vascular plants (Mishler et al., 1994; Stewart and Rothwell, 1993; Kenrick and Crane, 1997). Many types of evidence, molecular and mor- phological, demonstrate that lycophytes sensu lato form a distinct lineage of vascular plants from early in their evolutionary history (Raubeson and Jansen, 1992; Kranz and Huss, 1996; Kenrick and Crane, 1997). In Kenrick and Crane’s recent, extensive morphological phylogenetic analysis of living and extinct plants, monophyly is supported, and zosterophylls and lycopsids are inter- preted as two classes in a larger group, subdivision Lycophytina (Fig. 1). This GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 75 lon t t Crenaticaulis t . ascidian Sawdoniaceae Deheubarthia t Sawdonia t +——®- Thrinkophyton t pennies Barinophyton citrulliforme + Barinophyton obscurum ¢ | Barinophytaceae ’ Protobarinophyton t Lycophytina [7 Euphyllophytina rere ia et he pe . *- -——@ Gosslingia t meds a te -[ Lycopsida ; Sere pee dl Zosterophyllopsida L___ Tarelia t |——3 Cooksonia cambrensis t —_—_——@- Zosterophylium divaricatum t¢ |_—a- Renalia t -—————- Zosterophyilum fertile t Math ae }———# Uskiella t }+————-#- Zosterophyltum lanoveranum t i -———# Yunia t +} Rebuchia t + Sartimana aaeereeeeey eee eapobaadunee! Zosterophyllum deciduum t 3 Cooksonia caledonica ¢ 4 Huperzia Rhyniopsida Asteroxyion t cineianban’ Lycopsida Lycophytina > suaedimasneatile ae Noma t a it Banks Acmeatogay t -——®- Zosterophylium myretonianum t +} Gumuia t \_____—» Huiat Hicklingia t Rhynia t Psilophyton ¢ A B @ Zosterophyllophytina sensu Banks Fic. 1. Strict consensus trees showing relationships among several polysporangiophyte taxa and among lycophytes according to analyses conducted by Kenrick and Crane (1997). A) Adapted from Kenrick and Crane, fig. 4.31. B) From Kenrick and Crane, fig. 5.25, with permission. Zosterophyl- lopsida and Lycopsida are highlighted. See text for discussion. Not all taxa discussed in the text are included, partly because of differences in preservation and therefore absence of characters for some taxa larger group is defined (p.237—238) by the following synapomorphies: “more or less reniform sporangia, marked sporangium dorsiventrality, isovalvate de- hiscence along distal sporangium margin, conspicuous cellular thickening along sporangial dehiscence line, sporangia on short, laterally inserted stalks, exarch xylem differentiation.” Within the Lycophytina, the strict consensus tree of their analysis (p. 172) shows a polytomy (Fig. 1B) from which emerge the Zosterophyllopsida, a clade largely consisting of lycopsids, and a large number of individual taxa. Zosterophyllopsida sensu Kenrick and Crane (Fig. 1B) are defined by only two synapomorphies, namely circinate growth and a unique, two-rowed arrangement of sporangia. Among the single unresolved taxa excluded from Zosterophyllopsida sensu stricto are several plant genera (Rebuchia, Discalis) originally placed among the zosterophylls, several species 76 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) of Zosterophyllum including the type species (Z. myretonianum Penhallow), and some recently discovered Chinese forms. Thus Kenrick and Crane regard plants traditionally considered zosterophylls as paraphyletic. The incomplete data among the taxa analyzed for all characters employed, and lack of reso- lution at lower levels in the cladogram, indicates further investigation of char- acters and character homologies is needed. Thus these conclusions should be regarded with caution until additional data are available. Kenrick and Crane (1997) define the second clade, the Lycopsida (Fig. 1B), by the synapomorphies of microphylls, stellate xylem strand, vascularized leaves, sporophylls (close association of sporangium and leaf), and loss of spo- rangial vasculature. Lycopsida includes a clade composed of three possibly transitional genera, Asteroxylon, Drepanophycus, and Baragwanathia, and a second larger clade of plants traditionally considered as lycopsids. The first clade is composed of taxa that do not possess all of the characters definitive of lycopsids—Asteroxylon and Drepanophycus have cauline sporangia with a vascularized stalk, and Asteroxylon lacks full vascularization of lateral ap- pendages. Neither has a sporangium associated with a leaf, as occurs in mem- bers of the lycopsid clade and most probably in Baragwanathia (Hueber, 1992). TIME OF APPEARANCE.—The occurrence of the genus Baragwanathia in the pre- sumed late Silurian (Ludlow; Garratt et al., 1984; Rickards, 2000) of Victoria, Australia suggests a possible pre-Devonian diversification of lycophytes. Zos- terophylls also are recorded from these strata, but detailed descriptions are lacking (Tims and Chambers, 1974). Most other occurrences of late Silurian plant remains are much smaller in size and simpler in organization than either Baragwanathia or Zosterophyllum (Fig. 2). An interesting exception is the as- semblage of zosterophylls and related plants from well-dated late Ludlow stra- ta of Bathurst Island (Kotyk et al., in prep.). These new forms support an early diversification of the lycophyte lineage. POSSIBLE PRECURSORS.—Most polysporangiate plants (i.e. plants with branched sporophytes) with round to reniform sporangia referred to either as rhynioph- yte?, cooksonioid, or rhyniophytoid (Taylor, 1988; Hueber, 1992) represent possible close relatives or precursors to the lycophytes but currently there are no obvious candidates. Outline drawings (Fig. 2) of several Late Silurian and Early Devonian rhyniophytoids show that these plants mostly are very small and exhibit variable sporangial shapes and modes of dehiscence. Their small size, simple construction, lack of anatomical preservation, and fragmentary preservation make more precise inference impossible, but those with globose- reniform sporangial construction may represent likely candidates for lycophy- te precursors. A few larger forms with similar sporangial morphology have been recorded in pre-Devonian or earliest Devonian deposits, such as the spec- imen from the Ludlow of Bathurst Island (Basinger et al., 1996; Kotyk, 1998) which resembles a pseudomonopodial Cooksonia or very simple Zosterophyl- lum (fig. 2 I). Renalia, and some Cooksonia species, inferred to have arisen from within the plexus of latest Silurian-earliest Devonian rhyniophytoids (but their rela- GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 77 Silurian plants (A - 1) B Cc @ = nt = A x3 D 2, E G | x250 F @ x32 GOs | \\ x3 \O x24 q Lochkovian plants (J - O) oo || *4 { . x64 M } * contains cryptospores ] a co dyads & obligate tetrad © ® | © trilete miospores N x23 (@) x25 Fic. 2. Selected late Silurian and earliest Devonian meso- and megafossils. A—I) Pridoli plants. A) from Cai et al. (1993); B—D) from Fanning et al. (1991); E—G) from Edwards (1996). H) from Kotyk (1998). I) from Basinger et al. (1996). J—O). Lowermost Devonian plants from Edwards (1996). tionship to these simple plants thus far is not extensively tested by cladistic analyses), also represent likely precursors. When Renalia and Cooksonia were included in the phylogenetic analyses of Kenrick and Crane, they formed part of the large basal polytomy of the polysporangiophytes and were interpreted as possible stem-lineage zosterophylls. Further investigation of these early plants is needed. 78 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) The present contribution notes major advances and addresses several areas of concern in regard to early lycophyte evolution and diversification, including current views of zosterophylls, early lycopsids, and evolution of major lyco- phyte structures. CURRENT DATA ON ZOSTEROPHYLLS As a probable sister group(s) or stem group(s) to lycopsids, zosterophylls sensu lato (Fig. 3A—3C) may provide evidence of early evolutionary patterns within these lineages. However, presently recognized taxa, while demonstrat- ing greater diversity of form and some new combinations of characters, do not alter conceptions of relationships between zosterophylls and lycopsids nor aid in clarifying ordinal or familial delimitation within zosterophylls. The two main zosterophyllopsid clades sensu Kenrick and Crane (1997) can be typified by two genera, Sawdonia ornata (Fig. 3A) and Oricilla bilinearis (Fig. 3C). Differences include sporangium orientation and presence/absence of emer- gences (multicellular appendages). The Sawdonia clade (Kenrick and Crane’s Sawdoniaceae) exhibits more upright orientation of sporangia and emergences of varying morphology and arrangement, while the Oricilla clade (Kenrick and Crane’s Gosslingiaceae) exhibits naked axes and sporangia stalks attached per- pendicular to the axis. The Zosterophyllum depicted in Fig. 3B exhibits heli- cally arranged sporangia and is most similar to those species of Zosterophyl- Jum that form part of the basal polytomy in Kenrick and Crane’s analyses. Beyond these two well- supported clades, we now have considerable new in- formation about some of the more basal lineages, discussed next. Plants which appear very similar to Zosterophyllum (Fig. 3B), except that sporangia are more scattered and borne on longer stalks than in many previ- ously described Zosterophyllum species, are found in Pragian deposits of Chi- na and Bathurst Island, Canadian Arctic Archipelago (Kotyk, 1998). These do not quite fit existing generic or species delimitations, but as they become more completely known, they might aid in a better understanding of the genus Zos- terophyllum or closely allied taxa. The separation of several Zosterophyllum species into separate clades in Kenrick and Crane’s analysis is based mostly on symmetry of sporangial arrangement and presence/absence of circinate ver- nation. However, these characters are not easy to evaluate and their coding may not reflect that data about them are incomplete; most particularly, several taxa lacked anatomy. It is difficult to document the presence or (known) ab- sence of circinate vernation (and true terminations). The character “symmetry of sporangial arrangement” may be difficult to ascertain because determining if sporangia are helically arranged or are subopposite in two rows is difficult in compression remains. A two-rowed pattern could be produced by a widely spaced, very simple helix coupled with twisting of either sporangial stalks or the axis (see Gerrienne, 1996b for an analysis of this feature), Current analyses indicate these new Zosterophyllum species and other Zosterophyllum-like plants may represent stem zosterophylls for which only a few distinguishing characters are present. GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 79 AN 1mm < = 2mm > | er Ii, Fic. 3. Selected putative basal lycophytes or zosterophyllopsids. A, C) Representatives of Zos- terophyllopsida sensu Kenrick and Crane. A) Sawdonia ornata (Dawson) Hueber, based on spec- Zosterophyllum with helically arranged sporangia, a stem group taxon according to Kenrick and Crane (1997). D) Gosferia curvata Gerrienne from the Early Devonian of Belgium. Note thickening near base of sporangium. Redrawn from Gerrienne (1991). E) Ensivalia deblondii Gerrienne, a probable zosterophyllopsid from the Early Devonian of Belgium. F) Faironella valentula Gerrienne, affinity uncertain, from the Early Devonian of Belgium. Sporangium is borne inside a recurved, thickened spiny structure. E, F) redrawn from Gerrienne (1996). 80 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Another possible basal taxon, Gosferia curvata Gerrienne (1991, 1999), from the Early Devonian of Belgium, is unique in having long, basipetally recurving stalks, each terminated by a reniform sporangium (Fig. 3D). The closest simi- larity of this plant is to some unidentified zosterophyll-like plants in China and Bathurst Island and to some Silurian-Pragian rhyniophyte-like plants. The occurrence of a swelling near the sporangium base where it attaches to the stalk also is of interest; similar swellings are known in some Drepanophycus species (Li and Edwards, 1995) and in Kaulangiophyton (Gensel, pers. obs.). What it represents in any of these plants is currently unknown (but see below). Ensivalia deblondii Gerrienne (1996a) from the Early Devonian of Belgium has spiny, anisovalvate sporangia terminating stout stalks. The abaxial sporan- gial valve is enlarged and spiny, inside of which the more delicate, smooth adaxial valve rests (Fig. 3E). Odonax borealis Gerrienne (1996b), also from Belgium, and some unnamed zosterophylls from the Early Devonian of New Brunswick also exhibit modification of the sporangium/stalk in the form of enlargement of either stalk or distal sporangium valve and presence of emer- gences, features which perhaps help protect the sporangium. Further evaluation of the variability of emergence types (see Gensel, 1992 and Kenrick and Crane, 1997 for examples) among zosterophylls is needed as well as, perhaps, clarification of the homologies of different types of emer- gences. For example, in Nothia aphylla (Lyon, 1964) El Saadawy and Lacey (1979), the so-called emergences are produced by raised areas of axial tissue (cortex and epidermis), resulting from expansion of existing cells, with api- cally located stomata (Kerp et al., 2001). Are these structures truly emergences or do they represent short, discontinuous ridges of the stem? The unusual prismatic emergences of Serrulacaulis (Berry and Edwards, 1994) and those now known to occur in Bathurstia (Kotyk and Basinger, 2000) also add to diversity of those structures but do not provide any further tests of homology between emergences and microphylls. Some Early Devonian plants in which sporangia are arranged in an obvious strobilus may relate to the lycophyte clade, either among zosterophylls or ly- copsids. Demersatheca, based on only a few leafless fertile specimens from the Pragian of China, exhibits reniform sporangia deeply sunken into the axis (Fig. 4D) and forming tight strobili (Li and Edwards, 1996). Other plants with stro- bili include Distichophytum and Bathurstia, both attributed to the zostero- phylls, and some un-named plants of uncertain affinity from Bathurst Island. This morphology blurs the distinction between zosterophyll and barinophyte to some extent. Barinophytes are defined, according to Kenrick and Crane, by a very com- pact, unbranched strobilus, indehiscent sporangia, intrasporangial heterospo- ry, and sporangia borne on a horizontally extended axis which wraps around the stem (their “clasping” sporangium orientation). However, their relation- ships are unclear as the Barinophytaceae form part of a polytomy that includes some major groups of zosterophylls in Kenrick and Crane’s analysis. More data are needed about sporangium shape, attachment, and dehiscence in both the early strobilate plants and some plants usually attributed to barinophytes, such GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 81 Fic. 4. Plants from the Pragian of China that may be related to lycophytes. A) Adoketophyton subverticillatum Li and Edwards with a leaf-like appendage subtending each sporangium. Re- drawn from Li and Edwards (1992). B, C) Eophyllophyton bellum Hao, in which lobed leaf-like structures terminate lateral axes. Globose to reniform sporangia attach on the upper surface of the lobes. Redrawn from Hao and Beck (1993). D. Demersatheca contigua Li and Edwards, with spo- rangia sunken in the axes. Redrawn from Li and Edwards (1996). as Protobarinophyton. Faironella valentula Gerrienne (1996a) is also of interest in this discussion as a possible link between zosterophylls and barinophytes. Its sporangia are attached to the inside of a curved appendage (Fig. 3F), and while affinities presently are uncertain, it thus bears some resemblance to Bar- inophytaceae. Adoketophyton, a genus described from the Pragian of China by Li and Ed- wards (1992), demonstrates fertile regions with leaf-like structures that bear sporangia in their axils (Fig. 4A). Crane and Kenrick (1997) suggest the leaf- like structures might represent sterilized sporangia that became modified as a “leaf” as a result of changes in developmental processes such as multiplication and co-option. Another plant from the Pragian of China, Eophyllophyton bel- lum (Hao) Hao and Beck (1991) has the habit of a typical Early Devonian plant, but bears lobed, laminate leaf-like structures (Fig. 4B, 4C). Whereas this plant is at the base of the euphyllophyte clade of Kenrick and Crane, largely because 82 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Fic. 5. Line drawing of Drepanophycus qujingensis Li and Edwards. Note cauline sporangia not associated with leaves. Redrawn from Li and Edwards (1995). of its centrarch protostele, its globose-reniform sporangia are located adaxially on leaves and its tracheids possess G-type wall thickenings like those present in many zosterophylls and stem-lycophytes. In our opinion, this plant pos- sesses an odd suite of characters for basal euphyllophytes and it is perhaps more appropriately considered in the context of zosterophylls or lycopsids. Both Adoketophyton and Eophyllophyton thus provide idiosyncratic evidence of the development of laminate appendages, but little information on their homologies. PRE-LYCOPSIDS This term has been applied to plants with many lycopsid characters, but lacking the consistent association of sporangium with a leaf or lacking fully vascularized leaves. Genera included are Drepanophycus, Kaulangiophyton, Asteroxylon, and in some instances, Baragwanathia (for another interpretation, see Hueber, 1992). They form a sister clade to other lycopsids in Kenrick and Crane’s analysis and are treated there as a stem-based group of Lycopsida, the Drepanophycales. However, some aspects of their structure remain controver- sial, and re-evaluation of these features may influence their position relative to lycopsids or the interpretation of their evolutionary significance. Based on current knowledge of Drepanophycus, the following appears cer- tain: two species of Drepanophycus, D. qujingensis Li and Edwards and the type species D. spinaeformis (Goeppert) Li, Hotton and Hueber, exhibit both rhizomatous and aerial axes covered with vascularized microphyllous leaves and cauline sporangia (Fig. 5; Li and Edwards, 1995; Li et al., 2000). Among GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 83 vegetative remains assigned to Drepanophycus, leaf morphology varies con- siderably, from long and falcate to short and thorn-like. It has been suggested that the axes with the former represent aerial regions, those with the latter rhizomatous axes. Sporangia borne on comparatively short stalks (usually shorter than leaves) are interspersed among the microphyllous leaves, with no apparent association with them. Kaulangiophyon akantha exhibits a similar growth habit to that of Drepan- ophycus, but with all axes covered in stout, thorn-like emergences (Gensel et al., 1969). Studies are underway to test assertions that Kaulangiophyton is very similar to or congeneric with Drepanophycus (Kasper, 1972; Schweitzer and Giesen, 1980), or that Kaulangiophyton has sporangia associated with leaves (Hueber, 1992). Careful examination of original Maine collections and new material ha New Brunswick shows no indication of vascularization in the emergences. Sporangia were originally described as cauline, terminating a elas long nak and extending beyond emergences (Gensel et al., 1969). Hueber (1992) noted features of the sporangial stalk that led to his hypothesis that a leaf extends beneath the sporangium but is obscured by the sporangium and thus is not visible on impression remains. We note that the stalks of some sporangia attach to a thickened pad or ring of coalified material that itself is borne on an axial protrusion. It resembles the pad at the base of sporangia in Stockmansella and Huvenia. A similar thickening may occur in Drepanophy- cus qujingensis. However, in contrast to Hueber’s ideas, no extra appendage, emergence, or leaf lamina has been demonstrated in attachment to this thick- ening in Kaulangiophyton. Additional information on these and other char- acters of Kaulangiophyton and assessment of relationships will be presented in a later paper. ROOTING STRUCTURES IN EARLY LYCOPHYTES Information about rooting structures, other than rhizoids, in basal lycophy- tes is increasing, being recorded for the zosterophylls Zosterophyllum, Ba- thurstia, Crenaticaulis (Gensel et al., 2001), and Hsua (Li,1992), for Drepano- phycus (Rayner, 1984; Li and Edwards, 1995; Gensel et al., 2001) and for the rhyniacean Stockmansella (Taeniocrada) langii Fairon-Demaret (1985) by Schweitzer (1980). Two patterns are recognized by Gensel et al. (2001) and may correlate with habit. The first pattern occurs in tufted forms, such as some Zosterophyllum species, which produce downward trending axes from a cen- tral region that appear root-like. Upward trending axes departing from the same point are interpreted as shoots. In the other pattern, slender smooth axes depart from the rhizomatous regions, and in some cases divide to varying de- grees. These frequently lie across, not on, the bedding plane. Also, they are sometimes produced during the formation of K-branches as in Bathurstia and Drepanophycus (Gensel et al., 2001). In these plants, an axis departs, then orks. One of the resultant axes is oriented towards the rhizome apex and appears identical to shoots and the other, lacking either emergences or leaves and directed away from the apex, appears root-like (Figs. 6A, 6B). Based on 84 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) 1cm Fic. 6. A) Line drawing of the zosterophyll Bathurstia denticulata (Hueber) Kotyk and Basinger, showing numerous K-branches in which one portion is a rooting structure and the other a shoot- like structure, frequently in crozier form. B) Line drawing of Drepanophycus from the Pragian of Bathurst Island, showing several K-branches in which one part is a shoot and the other a smooth, root-like organ. r = rooting organs. Redrawn from Kotyk (1998), with permission and Gensel et al. (2001). current evidence, these rooting structures appear to be similar and perhaps homologous to shoots. Gensel et al. (2001) postulate that among lycopsids, and perhaps most non-seed plants, roots arose via a dichotomy of the shoot system, with one apex transformed into a root apex. This condition occurred over time progressively earlier in ontogeny, until the root arose in the embryo. The po- sition of the root primordium lateral to a shoot apex in many pteridophyte embryos may be significant in this regard. Furthermore, roots and shoots thus are homologous structures. None of the rooting structures among zosterophylls or lycopods have any kind of emergences or leaves on them except perhaps some fine hairs on those in Crenaticaulis. These differ from any ornament on aerial axes. Thus, part of the transformation to a root apex may have been to lose any potential to develop leaves or emergences. GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 85 PROTOLEPIDODENDRALES (PROTOLEPIDODENDRACEAE, HASKINSIACEAE, AND ARCHAEOSIGILLARIACEAE): EARLY LYCOPSID DIVERSIFICATION Following the Late Silurian-Early Devonian establishment of zosterophy]- lopsids and lycopsids, diversification of lycopsids is evidenced by abundant and widespread remains of lycopsids with distinctive anatomical and mor- phological features. Many of these form a group currently recognized as the Protolepidodendrales. These plants were prominent in Middle to early Late Devonian times and show more advanced and variable leaf morphology and vascular anatomy than pre-lycopsids. Sporangia where known are epiphyllous. Two families and a satellite taxon (Thomas and Brack-Hanes, 1984) are rec- ognized within protolepidodendralean lycopsids. The Protolepidodendraceae are well known from several parts of the world in Middle Devonian times. They consisted of lycopsids with helically ar- ranged, forked leaves, and globose to elongate sporangia located on the adaxial surface of the leaf. Their characteristic leaves, which are divided into three (Colpodexylon, Minarodendron) or five or more (Leclercqia) segments, are the most striking features of the plants. In many of these taxa, sporophylls resem- ble vegetative leaves. Sporangia in Leclercqia attach by a single discrete layer of cells to the leaf just proximal to its division, and the free part of the spo- rangium is directed towards the axis. A single line of dehiscence extends the length of the sporangium. Mode of sporangial attachment in other protolepi- dodendrids is not as clear, but they apparently lack stalks. Permineralized remains of some of these plants show a solid xylem column with several nar- row ridges of protoxylem around the periphery (Leclercqia, Minarodendron), resulting in a shallowly ribbed protostele, or a more deeply lobed xylem strand (Colpodexylon). Leclercqia is notable for being a ligulate, homosporous taxon (Grierson and Bonamo, 1979). Berry (1996) summarized the considerable new data and major features of these plants and other representatives of the Pro- tolepidodendrales (Table 1). Pre-Middle Devonian Protolepidodendraceae have been found. Kasper and Forbes (1979) recorded Leclercqia sp. from the late Early to Early Middle De- vonian of Maine and Kasper (1977) noted the presence of a new species of Leclercqia in the Emsian of New Brunswick. In both of these occurrences, axes bear five-segmented leaves, but all segments are oriented in the same direction (Fig. 7B) rather than being three-dimensionally arrayed as in L. complexa Banks, Bonamo and Grierson. At the time of recording them, the age deter- minations of sediments were uncertain. Since then, dispersed spore data in- dicate the New Brunswick sediments are undoubtedly Emsian. Recent collec- tions from these Emsian New Brunswick outcrops by Gensel document a sec- ond occurrence of the genus, these plants being very similar in leaf morphol- ogy and sporangial attachment to L. complexa (Fig. 7A), differing mainly in their smaller size. Remains of a plant with leaves possessing a total of nine segments also have been found in the Emsian of New Brunswick. In these forms, four segments are oriented in an upwardly directed, flattened lamina to each side of a long, AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) “paqo] a[91s010.1g ‘UMOUYUN SdINjION.NS 9.104 “sSuruoyory) euapidaqns Aq papunodns saseq Jeo] [RUOSeXOY ,,poyeYUl,, ‘,YI9e1, poyuatO AqyeIstp ‘yesoyey, ‘uouTWOI’d Jo Jed eB YIM ‘“TeUTWR] SOAR] L661 SpareMpy 19 ALlog puaua (/€6] ploury uosydvogy 1061 volsply (l4addoy) nian sisoapys4sy (UOXB} 991998S) TVAOVRLV TIDISOAVHOUV ‘ayv[NoNuap [asoj01g ‘umOUyUN saiods !ploaogo 0} asogoys eIsuRsods ‘QyBISeY JO aVNISES ‘ayejooouR] opeyq ‘ayejoned saava7y IL61 YIAoyuag uopydynyrsjay 9661 SpremMpy Jo ALieg pusuls ¢gG] SyUeY Ja UOSIOUID ,vIsuLysDET AVAOVISNIASVH ‘gyefnonuap 10 (uojAxapodjoy) paqoy 1ayita 9f91S0101g ‘UMOUY doyM SNoJodsowoY Sploao aesuoja vIsuRIOdS ‘[BUOISUDLIp-daiy) JO ayeuR]d ‘syusWISs MOLIEU 6 O} dn OVUT PapIAIp sdArey ZEH| PurlAaM Ja jasne.ry Uvospuapopidajojosg prol Syurg uazAxapodjoz 0661 V1 “ospuaposvUpy ZL6| UOSIALID 19 OWKUOg ‘syueg ,.o1ba4aj9a'7 JANRIOSOA 0} [RONUAPI ssa] JO a10Ur s{|Aydosods ‘sjoymopnesd Jo sadtjay SuIjquiasas suJayed Ul payasul sacra] ‘spleyoen panid pasopsog YIM WafAx ‘yorrxa ‘g1aisojoid suiays :(qUaisIsied soARe]) ssvds Jeo] OU ‘sanssH Asepuodeas ou “IIqey SuIdaes9 ‘snoyUOzIYys Ajquqoid ‘saxe poyouesg {A[SNOWOJOYSIP YIM sIUe[d @ (MO]aq YSLAISe Se UMOYS-URTUOAAG A[Iey o1e[) URIS UT SddUaLINIZO BOS Ing URTUOASG ae] AjIee puL o[PpIy UL JUepUNge A[[eIDedsy @ SHTVAGNAGOCIda TOLOUd ‘spisdoo4] uvajespuapopidajojoid jo stsdoukg ‘| aIaV GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 87 Fic. 7. Illustrations of canes from the Emsian of New Brunswick. A) Leclercqia collected by Gensel. This is most similar to L. complexa, 4X. B) Leclercqia sp. described in abstract by Kasper (1974). Oval bodies represent sporangia (arrows). 3.6X. C) New leclercqioid plant wi 8- segmented leaf. Note long distal segment, upward trending lateral ones in side view on upper left, an imprint of a sporangium associated with leaf between arrowheads to lower right, other sporangia at arrows, 2.4X. D. New leclercqioid plant with 9-segmented leaf, lateral leaf segments in surface view, 4X. 88 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) central, horizontal to downward- extended one (Figs. 7C, 7D). Elongate spo- rangia attach, perhaps at one end, to the upper surface of the undivided por- tion of the leaf (Fig. 7C, arrows). Another genus with segmented leaves, Cer- vicornus wenshanensis, was reported from the Pragian of Yunnan, China (Li and Hueber, 2000). This genus has leaves that are divided into two groups of four (eight segments total) segments, arrayed three-dimensionally. Although lacking important diagnostic evidence such as sporangia and anatomy, Cervi- cornis may be the oldest record of Protolepidodendraceae. Thus, more diver- sification of this group prior to the Middle Devonian is evident. In terms of younger examples, a lycopsid of similar size as all examples of Protolepidod- endraceae, and superficially identical in appearance, has been reported from Lower Carboniferous sediments of Argentina. Frenguellia Arrondo et al. (1991) has leaves with a long median segment, and two pairs of shorter opposite lateral tips, two of which point backward towards the stem. Sporangia are borne on unmodified leaves. However, we are concerned about the lack of conclusive and direct evidence for the present dating of these fossils; they may be substantially older. A second protolepidodendralean type is represented by the Haskinsiaceae (Figs. 8A-8C). This family differs from the Protolepidodendraceae in having more laminate, hastate or sagittate leaves and perhaps in sporangium shape. Leaf morphology varies less within this group than in Protolepidodendraceae. Epiphyllous, globose or flattened ellipsoidal sporangia have been demonstrat- ed in two species of Haskinsia from Venezuela, although mode of dehiscence and spores are unknown (Berry and Edwards 1996). Haskinsia from New York State has a primary xylem column with peripheral ridges of protoxylem sim- ilar to those of Leclercqia and Minarodendron (Grierson and Banks 1983). Archaeosigillariaceae, including the genera Archaeosigillaria and Gilboaph- yton, are sometimes regarded as members of the Protolepidodendrales but fer- tile examples are lacking. Archaeosigillaria has recently been restricted to the type specimen (Berry and Edwards 1997). Archaeosigillariaceae sensu Berry and Edwards (1997) includes Middle-Late Devonian plants which have a hex- agonal pattern of swollen leaf bases on the stem surface, and permanent lan- ceolate (more or less hastate) leaves (Figs. 8E, 8F). A permineralized specimen of Gilboaphyton goldringiae Arnold from Gilboa (Grierson and Banks 1963, PI. 37, figs 4, 7) demonstrates a protostele with approximately eight peripheral lobes in transverse section. Specimens of Gilboaphyton griersonii from Vene- zuela and New York State have similar leaf bases, and Berry and Edwards (1997) were able to infer the presence of elongate-oval, possibly parenchy- matous, regions beneath each leaf base (Fig. 8D). These are suggestive of pos- sible greater differentiation of stem tissue and may represent a prototypical form for one or more types of parenchyma/aerenchyma (e.g. parichnos) asso- ciated with leaf traces found in Late Devonian and Carboniferous lycopsids. Among the Protolepidodendrales, where xylem anatomy is known, second- ary wall pitting patterns include simple and bordered pits as well as annular and scalariform pitting. Minarodendron tracheids also exhibit pitlet sheets which resemble to some extent pitting found in younger lycopsids, termed GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 89 3mm { ion Dp \2| Fic. 8. Haskinsiaceae and Archaeosigillariaceae. A) Haskinsia- line drawing of sterile specimen of H. hastata from Venezuela. B) Fertile leaf and sporangium of H. sagittata Edwards and Bene- detto emend. Berry and Edwards fromVenezuela. C) Fertile leaf and sporangium of H. hastata. D, F) Gilboaphyton griersonii from the Devonian of Venezuela. E) Gilboaphyton goldringiae from New York State. A—C) Redrawn from Berry and Edwards (1996), D—F) redrawn from Berry and Ed- wards (1997). Williamson’s striations or fimbrils. Fertile examples, where known, were ap- parently homosporous. No basal regions or rooting structures have been de- scribed. A ligule has been documented only in Leclercqia. Few protolepidodendralean taxa were included in cladistic studies largely because many of them are not completely enough known. Leclercqia and Min- arodendron, the most completely-known examples, consistently form a clade that is sister, or paraphyletic, to Selaginellaceae or as part of a polytomy with 90 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Selaginellaceae and the rhizomorphic lycopsids (Kenrick and Crane, 1997; Bateman et al., 1992). Further, earlier putative lycopsids such as Drepanophy- cus, Asteroxylon, and Baragwanathia usually are sister to all of these clades. Most troubling in Protolepidodendrales is our current lack of knowledge about their rooting structures. This feature has the potential to reveal crucial infor- mation about relationships of the Protolepidodendrales to other groups, in- cluding the arboreous forms. The lack of well-preserved rooting structures is worrisome in the context of the large amount of above-ground material known and the excellence of some of the descriptions. One possible implication is that these plants may not have had specialized rooting structures. Alternative- ly, rooting structures may have existed, but were not preserved in the same horizons where the stems occur. INNOVATIONS IN GROWTH HABIT, ROOTING, AND REPRODUCTIVE STRUCTURES IN DEVONIAN LYCOPSIDS Until recently, the Protolepidodendrales, or certain less well-known re- mains, were the most likely group from which the Late Devonian and Carbon- iferous arborescent, ligulate, heterosporous lycopsids may have arisen. A cru- cial difference between Protolepidodendrales and aborescent forms however, is that the latter have a bipolar shoot system and erect growth habit. New discoveries from China have provided important evidence about changes in growth habit and rooting structures among Devonian lycopsids. SMALL TREE-LIKE LYCOPSIDS OF THE MIDDLE DEVONIAN.—The earliest examples of true bipolar growth, resulting in upright growth of a trunk and downward growth of a dichotomous rooting system, are to be found in the late Middle Devonian (Givetian) and earliest Late Devonian (Frasnian) of China. The two plants in question are Longostachys (Cai and Chen, 1996) and Chamaedendron (Schweitzer and Li, 1996). Both are small in height (0.5-1.5m), having a slen- der trunk that branches isodichotomously at the top to form a crown of narrow branches and terminal cones, and at the base to form a downward-pointing, conical- shaped, rooting system which divides into four major segments (Fig. 9). Roots may dichotomize further but lack any sign of attched rootlets or stigmarian-type rootlet scars. Longostachys is the larger of the two plants and has secondary growth in both the trunk and the rooting system. Sterile leaves are long and narrow, with spiny margins. Fertile leaves are spoon-shaped (the bowl containing the spo- rangium nearest the axis), and also have spiny margins (Fig. 9). Both of these plants are interesting because they have major features characteristic of rhi- zomorphic lycopsids (habit, dichotomous rooting system, bipolar growth) but lack rootlets. LATE DEVONIAN ROOTING STRUCTURES.—The famous ‘Naples tree’ (Lepidosigil- laria whitei) from the Late Devonian of New York State has previously been interpreted to have a swollen base covered in stigmarian rootlets (White, 1907; Pigg, 1992, 2001). This specimen is poorly preserved, and suffering pyrite de- GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 91 15cm Fic. 9. Reconstruction of Longostachys, showing the four-parted — structure and leafy stems, and (inset) a single sporophyll. After Cai and Chen 1996, with permis cay, and the presence of rootlets and rootlet scars can no longer be verified. The poor quality of White’s photographs showing the basal region of the plant compared to his other photographs of the upper trunk would suggest it was no better preserved when he worked on it. Thus we cannot accept this spec- imen as showing the earliest example of a lycopsid rhizomorph. Earlier Middle Devonian stumps named Eospermatopteris (Goldring 1924) which several peo- 92 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) ple considered to be possible lycopsid bases (as noted by Pigg, 1992), are far more likely to be cladoxylalean in affinity (Boyer and Matten 1996; Berry and Fairon-Demaret in progress). Clevelandodendron, from the Late Devonian (Fa- mennian) Cleveland Shale, has an isoetalean habit but again stigmarian root- lets were not demonstrated (Chitaley and Pigg, 1996). Branched rooting struc- tures are reported, but not described in any detail, for the Upper Devonian Cyclostigma kiltorkense (Johnson, 1913). Leptophloeum from the Upper De- vonian (Famennian) of South Africa has a bulbous base from which many thick dichotomous roots emerge all around (Gess and Hiller 1995). These roots are up to 10 mm wide in comparison with the 45 mm bulbous base width and seem too large to be rootlets. This suggests an absence of rootlets in this plant. Lastly, cormose lycopsid plant bases, some with attached rootlets, have been reported from the Upper Devonian (7late Famennian age) Red Hill, Pennsyl- vania deposits (Cressler, 1998, 1999; Scheckler et al., 1999). The previous earliest western occurrence of an undisputed lycopsid rhizo- morph (Pigg, 2001) is the report by Jennings et al. (1983) of Protostigmaria eggertiana from the Mississippian of Virginia, USA. Chinese occurrences in- clude Stigmaria rugulosa Gothan from the Upper Devonian Wutung Formation near Nanjing (Gu & Zhi 1974). Berry and Wang Yi collected dichotomous stig- marian roots with characteristic rootlet scars from Nanshan Quarry, in Jiangsu Province, also from the Wutung Formation (unpublished) where other fossils included Eviostachya and Hamatophyton (Li, Cai and Wang 1995). The facts, as outlined above, demonstrate that the emergence of the lycopsid rhizomorph remains a mystery. According to Rothwell and Erwin (1985) ‘there now can be little doubt that the rooting structures of all rhizomorphic lyco- phytes are shoot systems modified for rooting’ and implied rootlets therefore were homologous to leaves (p. 95). Bateman et al. (1992) however recognize that the stigmarian rhizomorph ‘is a shoot-like developmental system’, but prefer to regard it as a unique organ reflecting limited developmental options within the arborescent lycopsid bauplan’. The new evidence from China sug- gests that the earliest lycopsids with bipolar growth, although possessing a dichotomous, downwards-pointing rooting system, did not possess stigmarian rootlets, supporting Bateman et al.’s hypothesis. Such rootlets are not found until the uppermost Devonian in China and either uppermost Devonian or Lower Carboniferous in the U.S.A. By the time tree-shaped lycopsids appeared in the Middle Devonian, leaves were well developed as structures, and it seems inconceivable that leaves might have been involved in the development of a downward-directed rhi- zomorph as implied by Rothwell and Erwin’s hypothesis. If the K branching recognized above in zosterophylls and pre-lycopsids (Gensel et al., 2001) does in fact represent the earliest stages of bipolar growth in the Lycophytina (sensu Kenrick and Crane 1997) we can propose homologies between structures in zosterophylls and arborescent lycopsids (Fig. 10). In some zosterophylls and in Drepanophycus, K-branching occurs, the posterior (or downwards)-directed branch being naked and the apically (or upwards)-directed branch bearing ena- tions or leaves. The two branches are homologous structures, deriving from a GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 93 Zosterophyll or Lycopsid Longostachys Lepidodendrid Early Devonian Middle Devonian Latest Devonian / Carboniferous Fic. Hypothesis of homology of rooting structures between creeping early lycophytes (zos- ail or lycopsid), Longostachys, and a Carboniferous arboreous lepidodendrid. Shoot and rooting structures are homologous, being derived by a branching event. Early rooting structures lack lateral pita rootlets arose later, possibly in early arboreous forms. See text for further explanatio regular dichotomous branch as found in all zosterophylls, but the posterior- directed branch having developed the characters of roots and loss of enations/ leaves. In the earliest pseudobipolar lycopsids (Middle Devonian of China) the K branching occurs precociously at the initial growth of the sporophyte, caus- ing upward growth of the bifurcate, leafy aerial axes and downward growth of the bifurcate, naked rooting system. Final development of the arborescent lycopsid body plan occurs when rootlets evolve de novo on the rooting system to increase efficiency. INNOVATIONS IN FERTILE STRUCTURES AMONG MIDDLE AND LATE DEVONIAN Ly- copsips.—Middle Devonian Protolepidodendrales bear spores in single adaxial sporangia where known. Characteristically the fertile leaves are unmodified compared with the sterile ones. Often this means that the sporangia are borne more or less exposed on narrow leaf pedicels before the leaf forks (e.g. Leclerc- gia, Colpodexylon). In Haskinsiaceae, however, the laminate nature of the leaf and its upcurved pedicel (Figs. 8B, C) means that the sporangia are slightly more protected by the leaf (Berry and Edwards 1996). In those Protolepidod- endrales where spores are known, the plants are homosporous. Sporophylls are often found concentrated into areas of the axes, but show no closer inser- tion nor increased overlap to areas of sterile leaves. In Chamaedendron, leaves are slightly broader near their base, but in Lon- gostachys sterile leaves are narrow and fertile leaves are distinctly enlarged near their base in the area of attachment of the sporangia (Fig. 9), the shape being suggestive of an upturned spoon. The leaf margins of these Chinese lycopsids are spiny. The enlarged bases of the sporophylls are curved around 94 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) sporangia which contain large megaspores, demonstrating that the plant is probably heterosporous (microsporangia are unknown). Leaves are not well enough preserved to know for certain if ligules were present or not. In these small, bipolar, tree-shaped lycopsids the sporophylls are crowded closer to- gether forming distinctive fertile regions, and individual leaves overlap to some extent. Such fertile regions are therefore intermediate between the lax protolepidendrid fertile areas and the compact cones of many more advanced lycopsids. Lycopsid strobili from Kazakhstan, correlated to the upper part of the Giv- etian on the basis of plant fossils, are reported to contain a random mixture of mega- and micro-sporangia (Senkevich et al., 1993). These demonstrate an early record of compaction of sporophylls to form a tight, protective strobilus in lycopsids. A new probably arborescent lycopsid from the uppermost Devonian of Wuxi, China (Berry, Wang, and Cai, in prep), with branches slightly wider than those of Longostachys, demonstrates the earliest known occurrence of a lycopsid that has distinct micro- and megasporangiate cone-like structures. Megasporangiate ‘strobili’ contain leaves similar to those of Longostachys (but up to 90 mm long) whereas the microsporangiate strobili contain closely packed sporo- phylls that are much narrower. The megasporangiate strobili of this lycopsid are not terminal, but occur at dichotomies on the distal axes. Thus some mod- ification and compaction of sporophylls is well underway prior to the end of the Devonian. CONCLUSIONS Important new information has been obtained in recent years about the early history of lycophytes, particularly concerning anatomy and morphology of shoots, the occurrence of rooting structures, and greater diversification and differentiation of fertile regions. Recent broadly based phylogenetic analyses of vascular plants present interesting hypotheses of relationships, the most significant result being recognition of zosterophyllopsids, lycopsids, and of several basal lineages not yet well known. These analyses increase our under- standing of lycophytes, zosterophyllopsids and lycopsids as natural groups, but do not at this stage help significantly to clarify relationships within and between these groups. Moreover, some characters and possibly character states employed in these phylogenetic studies require modification or re-interpreta- tion in order to more fully resolve relationships. For example, the phylogenetic positions of several plants more traditionally viewed as zosterophylls remain unresolved as do the relationships of presumed basal lycophytes from China which show extremely interesting combinations of characters. Interpretation and understanding of important morphological characters and character-state transformations which appear to be critical in delineation of major clades are presently particularly unsatisfactory, and remain a major obstacle to resolution of early lycophyte phylogeny. Important examples include helical versus lin- ear insertion of sporangia, presence/absence of circinate vernation, and inter- GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 95 pretation of sporangial shape, orientation, and mode of dehiscence. Re-eval- uation is also particularly needed in linking early lycopsids (?protolepidod- endrids) to the bipolar, increasingly arborescent Late Devonian/Carboniferous forms. Although we can view a simple transformation sequence of, for exam- ple, the Middle Devonian protolepidodendralean fertile leaf via the Middle / Upper Devonian Longostachys sporophyll to the compacted Carboniferous cone, we do not see a corresponding pattern in the evolution of the stigmarian rooting system using the conventional homologies of rootlets and leaves. With dramatic advances in our knowledge of fossils as likely in the next decade as has occurred during the past three decades, and perhaps better understanding of the developmental basis for evolutionary change, we anticipate greater un- derstanding of the early radiation of lycophytes in the years to come. LITERATURE CITED ARNOLD, C.A. 1937. Observations on fossil plants from the Devonian of eastern North America. Ill. Gilboaphyton goldringiae, gen. et _ nov. from the Hamilton of eastern New York. Con- trib. Mus. Palaeontol. Univ. Mich. 5:7 ARRONDO, O., S. N. CESARI and P. R. as 1991. Frenguellia a new As of lycopods from the Early Carboniferous of Argentina. Rev. Palaeobot. Palynol. 70:187-197. BANKS, H. P. 1944. A new Devonian lycopod genus from ee Ses York. Amer. J. Bot. 31: 649-659 BANKS, H. P., P. M. BoONAMO, and J. D. GRIERSON. 1972. Leclercgqia complexa gen. et sp. nov., an lycopod from the late Middle Devonian of eastern New York. Rev. Palaeobot. Palynol. ad 19—40. BASINGER, J. F., M. E. Koryk, and P. G. GENSEL. 1996. Early land plants from the Late Silurian- Early Devonian of Bathurst Island, irene Arctic Archipelago. In: Current Research 1996B; Geological piers of Canada, pp. BATEMAN, R. M. 1992. Morphogenetic eae ppampenes and phylogeny of Oxroadia gracilis png emend., and O. conferta sp. nov.: orn ically preserved lycopods from Oxroad Bay, SE Scotland. Palaeontographica B228:2 BATEMAN, R. M., W. A. DIMICHELE and D. A. WILLARD. 1902 experimental cladistic analysis of anatomically preserved arborescent lycopsids from th on cap as phylogenetics. Ann. Missouri. Bot. Gard. 79:500-559. Berry, C. M. 1996. Diversity and distribution of Devonian Protolepidodendrales (Lycopsida). The Palaeobotanist 45: ile” 16. vo Berry, C. M. and D. Epwarps. 1994. New data on the morphology and anatomy of the Devonian a hana Hueber and Banks from Venezuela. Rev. Palaeobot. Palynol. 81: 41-150. eae e M. and D. Epwarps. 1996. The herbaceous lycophyte genus Haskinsia Grierson and Banks from the Devonian of western Venezuela, with observations on the leaf morphology and fertile specimens. ee J. Linn. Soc. 122:103-122 Berry, C. M. and D. Epwarps. 1997. A new species of the lycopsid Gilboaphyton Arnold from th erouites of coeeanet and New York State with a revision of the closely related genus Archacosiillara Kidston. Rev. Palaeobot. Palynol. 96:47—70. nd BOYER, J. S. a ATTEN. 1996. Anatomy of ee eriana from the upper Middle Devonian ( = srncan of New York. Abstracts—I.0.P.C. V — 1996, p. 13. Santa Barbara: Department of Geological Sciences. CAI, ai and L. CHEN. 1996. On a Chinese Givetian lycopod, L llus ZI and Feng, seis se morphology, anatomy and vontaclrniiiad: oo Abt. " porate 238 Cal, C. Y., Y. Dou and D. oe. 1993. New observations on a Pridoli plant assemblage from 96 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) north Xinjiang, northwest China, with comments on its evolutionary and palaeogeographic significance. Geol. Mag. 130:155-170. CHITALY, S. and K. Picc. 1996. Clevelandodendron ohioensis, gen. et sp. nov., a meaigptet upright lycopsid from the Late Devonian Cleveland Shale of Ohio. Amer. J. Bot. 83 CRANE, P. R. 1990. The phylogenetic context of microsporogenesis. In Hitinay. if asi Knox, R. B., eds. Microspores: stig and ontogeny. Academic Press, London CRANE, P. R. and P. KENRICK. 1997. Diverted development of reproductive organs: a source of ho loge alee in land plants. Pl. Syst. Evol. 206:161-174. an ae W. L . A diverse Late rina flora preserved in a fluvial setting, north-central Pen sii LH a S.A. Amer. J. B CRESSLER, W. L. 1999. Site-analysis ea floristics of the Late Devonian Red Hill locality, Pennsyl- vania, an Archaeopteris-dominated plant community and og tetrapod site. Unpublished Ph.D. Dissertation, Univ. of Pennsylvania, Philadelphia PA, 5 EpDWarbps, D. 1996. New gon into early land ecosystems: a ive jl a Lilliputian world. Rev. Palaeobot. Palynol. 90:159-174 EL Saapawy, W., and W. es Lacey. 1979. Observations on Nothia aphylla Lyon ex Heeg. Rev. Palaeobot. Palynol. 27:119-147. FAIRON-DEMARET, M. 1985. Les ser fossils de l’Emsien du Sart Tilman, Belgique. I. Stockman- sia langii (Stockmans) comb. nov. Rev. Palaeobot. Palynol. 44:243—260. FANNING, U., J. B. Ri RICHARDSON and D. ante 1991. A review of in situ spores in Silurian land plants. pp. 25-47 in S. Blackmore and S.H. Barnes, eds, Pollen and spores, patterns of di- versification. The Seatsenutics Association Special Volume No. 44. Oxford: Clarendon Press. GarraTT, M. J. 1984. The appearance of Baragwanathia (Lycophytina) in the Silurian. Bot. J. Linn. Soc. 89:355—358. GENSEL, P. G. 1982. Oricilla, a new genus referable to the a oe a the late Early evonian “ hye ser n New Brunswick. Rev. Palaeobot. Palynol. 37:3 GENSEL, P. G., 1992. Phylogenetic relationships of the zosterophylls and ean evidence from net el paleoecology, and cladistic methods of inference. Ann. Missouri Bot. Gard. 79: 450-473. GENSEL, P. G., A. E. KAsPER, and H. N. ANDREWS. 1969. Kaulangiophyton, a new genus of plants from the ates of Maine. Torrey Bot. Club Bull. 96:265—276. GENSEL, P. G., E. Koryk and J. F. BASINGER. 2001. Morphology of above- and below-ground structures “a Early Devonian (Pragian-Emsian) plants. In Gensel, P. G. and D. Edwards, eds. Plants Invade the Land: Evolutionary and Environmental Perspectives. Columbia University rat GERRIENNE, P. 1991. Les plantes emsiennes de Marchin (bord nord du Synclinorium de Dinant, ee Il. Forgesia curvata gen. et sp. nov. Compte Rend. Acad. Sci. Paris, Sér. II 313: Prete 2 pm Contribution a |’étude paléobotanique du Dévonien inférieur de Belgique. Les genres nouveaux Ensivalia et Faironella, Mém. Acad. Royale Sci. Belgique Publ., Cl. Sci. Coll. 4°, 3 Ser., T. 1:1-94 GERRIENNE, P. 1996b. Lower Devonian plant remains from Marchin (northern margin of Dinant en Belgium). IV. Odonax borealis, gen. et sp. nov. Rev. Palaeobot. Palynol. 93:89-— GERRIENNE, P. 1999. spies a new name for Forgesia (fossil plants). Taxon 48:61. Gess, R and N. HILLer. . A preliminary catalogue of fossil algal, plant, arthropod, and fish remains fron a “jin mee black shale near Grahamstown, South Africa. Ann. Cape Prov. Mus. (Nat. Hist.) 19:225-304. GOLDRING, W. 1924. The Upper Devonian forest of seed ferns in Eastern New York. New York State us. Bull. 251:5-92. GRIERSON, he D. and H. P. BANks. 1963. Lycopods of the Devonian of New York State. Paleontogr. Amer. 4: 217-295 RIERSON, I a and H. P. BANKS. 1983. A new genus of lycopods from the Devonian of New York State. Bot. J. Linn. Soc. 86:81-101 GENSEL & BERRY: EARLY LYCOPHYTE EVOLUTION 97 GRIERSON, J. D. and P. M. ener ree Leclercqia complexa: earliest ligulate lycopod (Middle Devonian). Amer. J. Bot. 66: Gu and Zui. 1974. Paleozonic wince Pl Ching Fossil Plants of China, 1. Sci. Press, Beijing, 262 pp. lin Chinese] Hao, S. and C. B. Beck. 1993. Further observations on Eophyllophyton bellum from the Lower Devonian — of Yunnan, China. Palaeontographica, Abt. B. Palaophytol. 230:27—41. HUEBER, F, M. 1992. Thoughts on the early lycopsids and zosterophylls. Ann. Missouri Bot. Gard. Feb ee JENNINGS, J.R., E. E. KARRFALT and G. W. oMpigtta 1983. Structure and affinities of Protostig eggertiana. Amer J. JOHNSON, T. 1913. On Sdnndon is kiltorkense Haughton sp. Sci. Proc. Roy. Dub- lin Soc. 13:500-528. Kasper, A. E, 1974. The reproductive organs of Kaulangiophyton and Drepanophycus. Amer. J. ot. 61:16. KASPER A. E. 1977. Anew vhs sot of the Devonian lycopod genus Leclercqia from New Brunswick, anada. Amer. J. Bot. 154:39. aperenen A. E. and W. H. nie 1979. The Devonian lycopod OE ON oe the Trout Valley Formation of Maine. Geol. Soc. of Maine, Maine Geology Bull. 1:49- KENRICK, P. and P. R. CRANE. 1997. The origin and early sedi ay of land plants: a cladistic study. Smithsonian Series in Comparative Evolutionary Biology. Washington: Smithsonian Institution Press. 441 p. Kerp, H., H. Hass, and V. MosBRUGGER. 2001. New data on Nothia aphylla Lyon 1964 ex El- ete et Lacey 1979, a poorly known plant from Lower Devonian Rhynie Chert. In: Gen- sel, P. G. and D. Edwards, eds. Plants po oon the Land: Evolutionary and Environmental escehiienh Columbia University Press. P. 52-82. KipsTOoN, R. 1901. ae lycopods and sphenophylls. Trans. Nat. Hist. Soc. Glasgow Trans. Ser. 3, 6:25—14 Kortyk, M. E. 1998. ae Silurian and Early Devonian Fossil Plants us apne Island, Arctic Canada. M.Sc. thesis, University of Saskatchewan, Saskatoon, Can. Kotyk, M. E. and J. F. BASINGER. 2000. The Early Devonian (Pragian) cme Bathurstia denticulata Heuber. Can. J. Bot. 78:193—207. KRaAnz, H. D. and V. A. R. Huss. 1996. Molecular evolution of pteridophytes and their Payee 4 to seed plants: bau from gr 18S rRNA gene sequences. Plant Sys. Evol. 2 KRAUSEL, R. and H. WEYLAND, 1932. Pflanzenreste aus dem Devon. IV. pee an ar Seackesibetgiarik 14:391—403. Li, C. S. 1990. Minarodendron cathaysiense (gen. et comb.nov.), a lycopod we wa late Middle Devonian of Yunnan, China. Palaeontographica, Abt. B, Palaophytol 220:9 Li, C. S. 1992. Hsiia robusta, an Early Devonian plant from Yunnan Province, Gin a its bearing on some structures of early land plants. Rev. Palaeobot. Palynol. 71:121-147 Li, C. S. and D. Epwarps. 1992. A new genus of Early Devonian land plants with novel strobilar construction from the Lower Devonian Posongchong Formation, Yunnan Province, China. Palaeontology 35: 257-272. Li, C. S. and D. Epwarps. 1995. A re-investigation of Halle’s Drepanophycus spinaeformis Goepp. from the Lower Devonian of Yunnan Province, southern China. Bot. J. Linn. Soc. 118:163- 192. Li, C. S. and D. Epwarps. 1996. Demersatheca Li et Edwards, gen. nov., a new genus of early land plants from the Lower Devonian, Yunnan Province, China. Rev. easing agi oe 93: 77-88. Li, C. S. and F. M. Hueser. 2000. Cervicornus wenshanensis, gen. a Pragian (Early re Nasiaie plant with forked leaves from Yunnan, China. Rev. cae it Palynol. 109:113- 19, Li, C. ;. F. M. Hueser and C, L. Hotron. 2000. A neotype for Drepanophycus spinaeformis Gép- pert 1852. Can. J. Bot. 78:889-902. Li, X., C. Cat and W. Yi. 1995. Hamatophyton verticillatum (Gu & Zhi) emend. A primitive plant of sphenopsida from the Upper Devonian-Lower Carboniferous in China. Palaeontographica, Abt. B, Paliophytol 235:1-22. 98 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Lyon, A. G. 1964. — fertile region of Asteroxylon mackiei K. and L. vionyesd 203:1082—1083. MISHLER, B. D., L. A. LEwis, M. A. BUCHHEIM, K. A. RENZAGLIA, D. J. GARBARY, C. F. DELWICHE, F. W. T. . KANTz and R. L. CHAPMAN. 1994. per eri ae of the “green algae” and “bryophytes. ” Ann. Missouri. Bot. Gard. 81:4 QMLLGAARD, B. 1987. A revised classification of the hie el s. lat. Opera Botanica 92:153- Picc, K. B. 1992. Evolution of isoetalean lycopsids. Ann. Missouri Bot. Gard. 79:589-612. Ficc RB. wa Isoetalean lycopsid evolution: from the Devonian to the present. Amer. Fern J. 91: 99-11 RAUBESON, L. m4 and R. K. JANSEN. 1992. se ea gran evidence on the ancient evolutionary split in one land plants. Science 255:1 RAYNER, R. J. 1 New finds of ea, spinaformis Goppert from the Lower Devonian of Scie? Tena: Roy. Soc. Edinb. E 363. RickarDs, R. B. 2000. The age of the earliest club mosses: the Silurian Baragwanathia flora in Victoria, Australia. Geol. Mag. 137:207—209. ROTHWELL, G.W, and Erwin, D. M. 1985. The rhizomorph apex of Paurodendron; implications for homologies among the rooting organs of lycopsida. Amer. J. Bot. 72:86—98. KLER, S. E., W. L. CRESSLER, T. CONNERY, S. KLAVINS and D. L. PostNikoFr. 1999. Devonian shrub and tree epee eaperiig 2 XVI International Botanical Congress, paseatiie 13. SCHWEITZER, H. J. and P. Gre 0. Uber Taeniophyton inopinatum, Protolycopodites devoni- cus, und Cladoxylon scoparia aus dem Mitteldevon von Wuppertal. Palaeontographica, Abt. B, Palaophytol 17 SCHWEITZER, H. J. and C.S. . on Chamaedendron nov. gen., eine ae a epg lycophyte aus dem Frasnium siidchinas. Palaeontographica Abt. B, Palaophytol. 2 SENKEVITSCH, M. A. 1971. New Devonian Lycopodiales. Mat. Geol. Miner. ees .. “Nau- ka” Kazakhskoi SSR, 4:88-92. SENKEVITSCH, M. A., A. L. JURINA and A. D. ARKHANGELSKAYA. 1993. On fructifications, morphology and anatomy of Givetian lepidophytes in Kazakhstan (USSR). Palaeontographica Abt. B, Pa- laophytol 230:43-58 STEWART, W. N. and G. Ro OTHWELL. eae Paleobotany and the evolution of plants. Cambridge Uni rp cs Cambridge. 52 TAYLOR, T. N 8. The origin of lend plants: some answers, more questions. Taxon 37:805-833. THOMAS, B. A. ghee S. D. BRACK-HANES. 1984. A new approach to family groupings in the lyco- phytes. Taxon 33:247—2 sees Tims, J. D. and T. C. CHAMBERS. 1984. Rhyniophytina and Trimerophytina from the early land flora of Victoria, Australia. Palaeontology 27:265-279. WacneRr, W. H. and J. BEITEL. 1992. Generic classification of modern North American Lycopodi- aceae. Ann. Missouri “a Gard. 79:676-686. Wuire, D. 1907. A remarkable fossil tree trunk from the middle Devonic of New York. Bull. New York State Museum, (Geol. Pap.) 107:328-340. American Fern Journal 91(3):99—114 (2001) Isoetalean Lycopsid Evolution: from the Devonian to the Present KATHLEEN B. PIGG Department of Plant Biology, Box 871601, Arizona State University, Tempe, AZ 85287-1601 .s fs. p eS ABSTRACT.—The evolution of the isoetalean id t vascular plants, from Late, (or possibly Middle), Devonian to the. current day genus Isoetes. The fhe ate fossil members of this group are the arborescent lepidodendrids that dominated the Late Carboniferous coal swamps. Simpler unbranched isoetaleans with elongate stems also predated, coexisted sae and postdated the coal swamp trees, extending well into the Mesozoic. Whereas certain syna morphies such as stigmarian rootlets, bipolar growth and secondary tissues unite the clade, tains features characterize smaller subgroups of differing age, growth form and possibly, evolutionary lineage. Although some of these features are well known for plants of given time periods, partic- ularly the Carboniferous, trends in character evolution have never been adequately docume nted olete microspores, sunken sporangia and elaborate ligules with glossopodia occur within elongate- stemmed Triassic forms. The dominant plant habit of modern Isoetes, a reduced cormose form that lacks appreciable stem elongation, originated at least by the Jurassic and typifies late Mesozoic and Cenozoic isoetaleans. The lycopsids have always held a particular fascination for paleobotanists and neontologists alike because of their diverse morphologies, long strati- graphic record and their phylogenetic position as the outgroup to the rest of the vascular plants (Gifford and Foster, 1989; Kenrick and Crane, 1997). Whereas most lycopsids are similar to other pteridophytes in their growth habit and life history, isoetalean plants are characterized by distinctive mor- phological innovations, including secondarily derived bipolar growth, second- ary cortical (and possibly vascular) tissues and arborescence. This group is generally thought to make its first appearance in the Late Devonian, or perhaps earlier, and reach its greatest diversity and ecological importance in the Late Carboniferous. Isoetaleans continue into the Early Mesozoic, where they are represented worldwide by numerous species of Pleuromeia Corda, Annalepis Fliche and several other forms. In the Jurassic, plants quite similar to modern Isoetes L. first appear. They continue to diversify throughout the Cretaceous and Tertiary and up to the present day. The best known members of the Isoetales are the arborescent lepidoden- drids, a group that dominated the Late Carboniferous ( = Middle-Upper Penn- sylvanian) coal swamp of Euramerica. These heterosporous trees had multi- branched crowns of varying forms, long microphyllous leaves attached to char- acteristic diamond- or hexagonal-shaped leaf cushions, thick secondary cor- tical tissues, well developed stigmarian rooting systems and often monosporangiate cones with reduced numbers of megaspores and specialized 100 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) modifications for dispersal (Phillips, 1979; Bateman et al., 1992; Phillips and DiMichele, 1992). Most of the arborescent forms became extinct by the Late Carboniferous of Euramerica concomitant with the loss of the most extensive coal swamp habitats (Bateman, et al., 1992; Stewart and Rothwell, 1993). How- ever, growing alongside the Middle Pennsylvanian tree lycopsids and surviv- ing after them into the Late Carboniferous (Upper Pennsylvanian) were smaller (1-2 m tall), unbranched or rarely branched isoetaleans such as Chaloneria Pigg & Rothwell and Sporangiostrobus Bode (Pigg and Rothwell, 1979, 1983a, 1983b, 1985; Wagner, 1989). It is now known that isoetaleans with this elon- gate, unbranched growth habit originated by the Late Devonian (e.g., Cleve- landodendron Chitaley & Pigg, Chitaley and Pigg, 1996) and were widespread throughout the Triassic where they belonged to at least two broad groups, Pleuromeia-like and Annalepis-like forms and their relatives (Pigg, 1992; Grau- vogel-Stamm and Lugardon, 2001). Today the order Isoetales is represented by a single genus, Isoetes, which remains one of the most unusual living pteri- dophytes and retains many morphological, anatomical and developmental fea- tures that reflect its origins (e.g., Stewart, 1947; Gifford and Foster, 1989; Bate- man, et al., 1992; Pigg, 1992). Recent studies have provided new information about the origin, evolution and diversity of the Isoetales. In this contribution, four phases of evolution in isoetalean lycopsids are addressed: (1) presumed Devonian origins of the Is- oetales; (2) their maximum diversity in the Late Carboniferous, (3) Mesozoic plants with the growth habit of modern Isoetes, (4) new occurrences of Cre- taceous and Tertiary forms. Lastly, problems of relating fossil taxa to extant Isoetes are addressed and suggestions are outlined for a better understanding of the isoetalean lineage through time. ORIGIN OF ISOETALEAN LYCOPSIDS Historically, the Isoetales, the rhizomorphic heterosporous lycopsids, have been assigned to a variety of orders including the Lepidodendrales, Pleuro- meiales, Isoetales and some forms occasionally to Protolepidodendrales (Chal- oner, 1967; Pigg and Rothwell,1983b; DiMichele and Bateman, 1996; Kenrick and Crane, 1997), and even, probably mistakenly, the Selaginellales (see Roth- well and Erwin, 1985). In this paper I follow DiMichele and Bateman (1996), who recognize the entire group within the order Isoetales, based on a suite of synapomorphies including bipolar growth, secondary tissue production, and stigmarian rootlet formation. Whereas much is known about both vegetative and reproductive structure of fossil isoetaleans, particular attention has been given to the basal parts, or rhizomorphs of the plants, especially the size and shape of their lateral lobes or horns. A great deal of diversity has been documented for these rhizomorphic features which vary from elongate, dichotomizing rooting systems of Stigmaria Brongniart to smaller bilaterally symmetrical lobed or rounded forms (e.g., Pleuromeia), to radial unlobed forms (e.g., Paurodendron Fry). All of these thizomorphic forms can be homologized (Rothwell and Erwin, 1985; Grauvo- PIGG: ISOETALEAN LYCOPSIDS 101 gel-Stamm and Lugardon, 2001). It is now clear that the shape of the rhizo- morph, or even its symmetry, is not nearly as significant as its production of stigmarian appendages or “rootlets” with their characteristic anatomy and the likelihood of their secondary derivation from an aboveground shoot system. This fascination with the basal portions of the plants has been particularly centered around the stigmarian rooting system and its ubiquity among Late Carboniferous coal swamp lepidodendrids. There has been interest for some time in finding evidence for the origin of Stigmaria, with the underlying as- sumption that this would lead toward understanding the origin of the isoeta- lean clade as a whole. Devonian and Early Carboniferous plants that obtained tree stature and/or those which produced rhizomorphic plant bases have been sought as the ancestors of lepidodendrids. The best known candidates have been the enigmatic ‘Naples tree” Lepidosigillaria whitei Kraéusel & Weyland from New York State and Cyclostigma kiltorkense Haughton, from Late De- vonian strata in several areas of Ireland and Bear Island in the Arctic, and hina. Lepidosigillaria whitei Krausel & Weyland, a large stem with a plant base that appears rounded and lacking a stigmarian rooting system, occurs in the Late Devonian of New York State (Grierson and Banks, 1963). The original specimen was over 5 m tall and demonstrated that well developed, arborescent lycopsids had clearly evolved by the Late Devonian (White, 1907; Krausel and Weyland, 1949; Grierson and Banks, 1963; Chaloner, 1967). Because of lateral compression during fossilization, the exact nature of the plant base of Lepi- dosigillaria is unknown. However, there is some increased thickening of basal cortical tissues, suggesting secondary tissues were produced. Several distinc- tive decortication layers are recognized on the stem surface, and small (less than 3 cm long), very narrowly attached leaves occur at higher levels. Al- though Lepidosigillaria has been described as having rootlets attached to the plant base (White, 1907), preservation is poor and their presence cannot be verified (Gensel and Berry, 2001). A second Late Devonian lycopsid, Cyclostigma kiltorkense Haughton, has also been difficult to interpret. This plant was originally described from the Upper Devonian of Ireland, and later found in Bear Island, in the Arctic, and China (Johnson, 1913; Chaloner, 1967, 1968; Schweitzer, 1969; Cai and Wu, 1994). Cyclostigma is described as a large, heterosporous plant with charac- teristic rounded leaf scars (Chaloner, 1967) and is reported to have a bilobed plant base (Johnson, 1913). Specimens showing plant bases may be known (Chaloner, pers. comm. 1999) but they have never been clearly illustrated and have yet to be studied in detail. It also remains unclear whether all plant remains assigned to this genus represent the same plants. Clearly, reinvesti- gation of this problematic genus is in order. Despite the currently limited information about Lepidosigillaria and Cyclos- tigma, their occurrences have established the presence of large isoetalean plants at least as early as the Late Devonian. In recent years additional Late Devonian and older, Middle Devonian lycopsids with a presumed relationship to the Isoetales have been reported. The Middle Devonian Longostachys latis- 102 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) @ ia Fic. 1. Line diagram of the rootstock of Longostachys latisporophyllus. Redrawn from Cai, Chon- gyang, and Lizhu Chen, 1996. porophyllus Zhu, Hu & Feng emend. Cai & Chen was described from Givetian strata of southern China (Cai and Chen, 1996; Gensel and Berry, 2001). Lon- gostachys is reconstructed as a small, branching heterosporous lycopsid tree that was up to 1.5 m tall with a stem 3.5 cm wide and a plant base that had a three-dimensional funnel-like structure with thick, rootlike appendages (Fig. 1). Because of its bipolar growth and centralized rooting system, this Middle Devonian plant may have some significance to isoetalean origins. The rooting structure of Longostachys bears some resemblance to stigmarian appendages, however it apparently lacks the dichotomizing stigmarian axes and character- istic stigmarian rootlet scars. Given the plant’s Middle Devonian age it may be that these rooting organs are modified branching axes that cannot be clearly defined as stem or root, and may be more comparable to the H & K branching characteristic of zosterophylls (Gensel and Berry, 2001). However, well-defined leaves and strobili are produced by Longostachys, indicating that true leaves and stems had evolved by then. Whatever their exact homologies, the rooting structures of Longostachys suggest a potential intermediate step in the evolu- tion of bipolar growth and the origin of stigmarian rooting systems in isoeta- leans. Additional Late Devonian isoetaleans have been discovered in Pennsylvania and Ohio. A flora from Red Hill, Pennsylvania contains cormose lycopsid plant bases (Cressler, 1998, 1999; Scheckler, et al., 1999). Five specimens with at- tached rootlets have been recovered from oxbow lake facies of this fluvial deposit. One specimen has a stem width of 10 cm and a flared base of 12 cm, and with two lobes visible and is estimated to have four lobes. A second spec- imen is approximately 2 cm wide with a flared base. Roots are borne in or- thostichies, with largest roots adjacent to one another. Associated, usually un- PIGG: ISOETALEAN LYCOPSIDS 103 branched, stems have various decortication patterns including Knorria Stern- berg, Helenia Zalessky and Cyclostigma, as well as Lepidodendropsis Lutz-like pseudowhorls. Reproductive structures are not yet known (Walter L. Cressler, University of Pennsylvania, pers. comm. 1999). The Elkins, West Virginia lo- cality that provided the earliest seed plant, Elkinsia, is of comparable age to Red Hill, and also contains lycopsid stems but plant bases are not yet known from Elkins (Gar W. Rothwell, Ohio University, pers. comm.1999). Another recently recognized Late Devonian plant with an elongate isoeta- lean plant habit is Clevelandodendron ohioensis Chitaley & Pigg from the Cleveland Shale (Chitaley and Pigg, 1996). The genus is based on a single specimen which represents the entire plant. Clevelandodendron is an un- branched, slender plant with a compact terminal bisporangiate cone at the apical end and a partially preserved plant base proximally. The specimen is 125 cm long from base to apex and 2 cm wide for most of its length. Although details of the plant base are obscure, it appears to bear several thick append- ages that taper distally and an elongate stigmarian root system is lacking. The terminal bisporangiate cone contains trilete megaspores that are laevigate (smooth) and trilete microspores that conform to either Calamospora Schopf, Wilson & Bentall or Punctatisporites (Ibrahim) Potonié & Krempf. Ligules are not known. Clevelandodendron is of particular significance because it dem- onstrates that elongate, slender isoetalean lycopsids with an unbranched plant habit, more common in later isoetalean forms, were present as early as the Late Devonian. Plant remains with similarities to Clevelandodendron are also known in younger, Lower Carboniferous strata at several localities in eastern North America, France and Ireland. For example, anatomically preserved stems of similar dimensions are present in Lower Carboniferous localities in France (Brigitte Meyer-Berthaud, Université Montpellier, pers. comm. 1996) and ad- ditional permineralized stems showing cortical and stelar features transitional between the traditional Protolepidodendrales and Lepidodendrales occur at a variety of localities in North America, (e.g., Andrews, Read and Mamay, 1971; Cichan and Beck, 1987; Roy and Matten, 1989), Ireland (Matten, 1989; Roy and Matten, 1989), Scotland (Beck, 1958), and France (Meyer-Berthaud, 1984). Several bisporangiate cones with spore types similar to those of Clevelandod- endron have been described from Upper Devonian and Lower Carboniferous strata of France and eastern North America (Arnold, 1933, 1935, 1939; Ma- thews, 1940; Meyer-Berthaud, 1984). These bisporangiate cones with laevigate (smooth) megaspores are unlike the stratigraphically younger bisporangiate cones of Flemingites Carruthers. The genus Flemingites bears Lycospora Schopf, Wilson & Bentall microspores with an equatorial flange, and Lageni- cula (Bennie & Kidston) Potonié & Krempf or Lageniosporites Potonié & Krempf megaspores with an apical spore extension called a gula (Brack-Hanes and Thomas, 1983). Thus, not all Paleozoic bisporangiate lycopsid cones can be assigned to Flemingites (Chitaley and Pigg, 1996). Species of Flemingites are thought to have been borne by small lepidodendrid trees such as Paraly- copodites brevifolius (Williamson) DiMichele (e.g., Brack, 1970; DiMichele, 104 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) 1980). Thus there may be at least two lineages of Paleozoic lycopsids that produced bisporangiate cones: those related to Clevelandodendron and those bearing Flemingites cones. One interesting Lower Carboniferous plant of isoetalean affinities is the Pro- tostigmaria/Lepidodendropsis plant. Protostigmaria Jennings plant bases are reported to have had up to 13 lobes and a maximum width of 32 cm. This multilobed rooting structure was initially described by Jennings (1975) and later studied by Jennings, Karrfalt and Rothwell (1983). These authors sug- gested that Protostigmaria grew like Isoetes, with new lobes being produced by forking at the ends of furrows. The presumed above-ground parts of Pro- tostigmaria, assignable to Lepidodendropsis Lutz are found both at the same and at several additional nearby sites in the Mississippian Price Formation of Virginia (Patricia G. Gensel, personal communication, 1999). Above-ground parts of Lepidodendropsis suggest the plant was a tree that branched dichot- omously and may have had determinate growth. It was apparently eligulate but heterosporous and had linear leaves. Middle Devonian fossils assigned to Lepidodendropsis may represent a mix of several different plants that may not be equivalent to this Lower Carboniferous plant. Study and correlation of Pro- tostigmaria plant bases and vegetative stems at several decortication layers of Lepidodendropis and reproductive remains from the same localities will pro- vide the opportunity to reconstruct this interesting Lower Carboniferous ly- copsid (Gensel and Pigg, in progress). LATE CARBONIFEROUS DIVERSITY During the Late Carboniferous, diversity and ecological importance of isoe- talean plants reached their maximum. Arborescent lepidodendrids are exten- sively well known as both permineralizations in coal ball floras, and as coal- ified compressions. Whole plant reconstructions, reproductive biology and pa- leoecological aspects of these plants have been reviewed elsewhere (e.g., Phil- lips, 1979; Bateman, et al., 1992). The extensive information available about Carboniferous lepidodendrids has allowed for the whole plant reconstruction of several species including the bisporangiate Paralycopodites, and larger monosporangiate trees with special- ized dispersal units that have been modified presumably for water dispersal (Phillips, 1979). These include Diaphorodendron DiMichele and Synchysiden- dron DiMichele & Bateman, genera with two different crown branching forms that both produce Achlamydocarpon varius (Baxter) Taylor & Brack-Hanes, a megasporangium with a specialized, thickened wall surrounding a single func- tional megaspore (DiMichele and Bateman, 1992). Another tree lycopsid, Lep- idophloios Sternberg, produces Lepidocarpon Scott, a cone that breaks up into individual megasporophyll units that contain the megasporangium with a sin- gle functional megaspore surrounded by an elongate and elaborated sporophyll (Phillips and DiMichele, 1992). Another tree form, Sigillaria Brongniart, bore monosporangiate cones on lateral cauline peduncles. Permineralized megas- porangiate cones of Sigillaria, Mazocarpon (Scott) Benson produced 4-8 large PIGG: ISOETALEAN LYCOPSIDS 105 megaspores that are found each associated with a portion of sporangial wall and a pad of sterile tissue (Schopf, 1941; Pigg, 1983; Phillips and DiMichele, 1992). Parallels of these specialized forms with the seed habit have been com- monly made (e.g., Phillips, 1979; Thomas, 1981). We now know that in addition to the large arborescent lepidodendrids, which themselves were quite diverse, isoetaleans had a variety of other plant habits. These included the small scrambling “‘pseudoherbs” Oxroadia Alvin and Paurodendron Fry, as well as the unbranched monopodial plants Chalo- neria Pigg & Rothwell and Sporangiostrobus Bode. Oxroadia, a plant of Lower Carboniferous age, is known from a number of localities in England and Scot- land (Bateman, 1992), while Paurodendron has been described from Middle and Late Pennsylvanian coal ball floras of eastern and midcontinent North America (Rothwell and Erwin, 1985). Comparative developmental studies of apical and lateral growth of rhizomorphs in Paurodendron, Stigmaria, Na- thorstiana Richter and modern Isoetes have been instrumental in understand- ing homologies across the isoetalean clade (Karrfalt, 1984; Rothwell, 1984; Rothwell and Erwin, 1985; Rothwell and Pryor, 1991). The genus Chaloneria is reconstructed as a whole plant based on perminer- alized stems, fertile areas with in situ spores, decortication surfaces, and mega- gametophytes (Pigg and Rothwell, 1979, 1983a, 1983b, 1985). In contrast to the well known lepidodendrids, Chaloneria was more like its Devonian rela- tive, Clevelandodendron, in having an unbranched, elongate stem. Instead of compact cones Chaloneria exhibits a less well differentiated fertile area that occurs either apically with alternating mega- and microsporangiate zones (C. cormosa Pigg & Rothwell) or alternating vegetative and fertile zones (C. per- iodica Pigg & Rothwell; DiMichele, Mahaffy and Phillips, 1979; Pigg and Roth- well, 1983a, 1983b). These and presumably related forms share the megaspore type Valvisisporites Ibrahim (Gastaldo, 1981) and microspores of Endosporites Wilson & Coe (e.g., Brack and Taylor, 1972; Hanes, 1975). Compressed lycopsid reproductive structures bearing the same spore types as Chaloneria are given the name Polysporia Newberry (Chaloner, 1958; Dra- bek, 1976a, 1976b; DiMichele, et al. 1979; Pigg and Rothwell, 1983b). Previous reports of Polysporia were from incompletely preserved specimens, and it was difficult to tell whether these structures were compact cones or fertile areas such as those of Chaloneria. A recent study of well-preserved, complete cones of a new species of Polysporia from France, P. doubingeri Grauvogel-Stamm & Langiaux shows that Polysporia was at least sometimes a compact cone rather than an undifferentiated fertile area (Grauvogel-Stamm and Langiaux, 1995). This study brings to question whether all plants that produced these charac- teristic spore types, well known from Middle Pennsylvanian coals (Pigg, 1992), had the same type of growth habit (Pigg and Rothwell, 1983b, Grauvogel- Stamm and Langiaux, 1995). Sporangiostrobus Bode is known from coal ball floras of the Middle Penn- sylanian of Kansas and compressed floras from the Stephanian ( = Upper Pennsylvanian) of Spain (Pigg and Rothwell, 1983b; Wagner, 1989 and refer- ences cited therein). This plant parallels the Chaloneria/Polysporia plants in 106 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) stem anatomy, size, unbranched to rarely branched habit and presence of fer- tile areas rather than compact cones. These plants grew in nearly monotypic stands, a situation also common in later pleuromeian forms (e.g., Retallack, 1975; Fuchs et al., 1991; Pigg, personal observation, 1991). In contrast to Chal- oneria/Polysporia plants, Sporangiostrobus bore the highly variable Densos- porites (Loose) Schopf, Wilson & Bentall microspores, characterized by a prom- inent equatorial ridge, the cingulum, and Zonalesporites (Bartlett) Leisman megaspores. Whereas Pigg and Rothwell (1983b) suggested that Sporangiostro- bus may be assignable to the same family as Chaloneria, the Chaloneriaceae, similarities of the two groups may be more the result of convergence in growth habit. Several other Carboniferous forms which, for various reasons, have been allied with the isoetaleans are reviewed by Pigg (1992). ORIGIN OF MODERN ISOETES PLANT HABIT AND MORPHOLOGICAL FEATURES As the Paleozoic ended and most of the large, specialized lepidodendrid plants died out, a variety of several smaller, elongate, unbranched forms re- mained. These are represented in the Mesozoic by the genera Pleuromeia Cor- da, Annalepis Fliche and several other diverse forms (Pigg, 1992; Skog and Hill, 1992; Retallack, 1997; Grauvogel-Stamm and Lugardon, 2001). Around a dozen species of Pleuromeia have been named from numerous localities throughout the world including Germany, Russia, China, Japan and Australia (Pigg, 1992). Some pleuromeian genera, such as Cylomeia (White, 1981) and Lycomeia (Dobruskina, 1985) have been proposed on basis of the geographic occurrence, morphology and mono- vs bisporangiate nature of plants. How- ever, whereas numerous species of Pleuromeia have been described and some are well known (e.g., Grauvogel-Stamm, 1993), the overall diversity and inter- relationships of these taxa remain obscure. In addition to pleuromeians there are a suite of Triassic plants assigned to Annalepis and related taxa which have been described by Grauvogel-Stamm and colleagues (Grauvogel-Stamm and Duringer, 1983; Grauvogel-Stamm and Lugardon, 2001). This latter group appears distinct from pleuromeians on the basis of a number of features including sporophyll morphology and spore char- acters. In some features Annalepis may resemble some later isoetaleans of Cre- taceous-Tertiary ages (Grauvogel-Stamm and Lugardon, 2001). A third source of Triassic lycopsids is the discovery of several permineral- ized forms. These include Takhtajanodoxa Snigirevskaya from Russia (Snigir- evskaya, 1980a, 1980b), forms attributed to Pleuromeia, also from Russia (Sni- girevskaya and Srebrodelskaya, 1986) and the more recently reported permi- neralized lycopsids from the Lower Triassic of Australia (Cantrill and Webb, 1998). The affinities of permineralized Triassic forms and their relationships to the more widely known compressed remains have not been fully resolved. In the early 1980’s Snigirevskaya described the permineralized Triassic plant Takhtajanodoxa from Russia (Snigirevskaya 1980a, 1980b). This plant has stem, leaf and megaspore anatomy similar to Chaloneria and other Pennsy]- vanian lycopsids. What is surprising, however, is the presence of elaborately PIGG: ISOETALEAN LYCOPSIDS 107 shaped ligules with curving basal extensions of the glossopodium (Snigirev- skaya, 1980a, 1980b; Pigg, 1992). This type of ligular structure is lacking in Paleozoic permineralized lycopsids but present in extant Isoetes (Gifford and Foster, 1989). Permineralized stems described as Pleuromeia were also char- acterized from Russian Lower Triassic strata (Snigirevskaya and Srebrodel- skaya, 1986). These specimens show small, protostelic stems lacking second- ary tissues with mesarch leaf traces, and fimbriate metaxylem tracheids with scalariform and reticulate wall thickening patterns. More recently, perminer- alized forms have been described from the Lower Triassic Bowen Basin of eastern Australia by Cantrill and Webb (1998). These authors describe a suite of disarticulated lycopsid plant organs that include one bisporangiate and one monosporangiate cone type, two types of stems, one with prominent aeren- chyma in the cortex, and one type of rooting structure that bears stigmarian- like rootlets. From the presently known information, at least two types of ly- copsids thus appear to be present in these remains. As with late Carboniferous coal ball lycopsids, for the first time the exciting potential exists for “whole- plant” reconstructions of Lower Triassic isoetaleans, based on the co-occur- rence, anatomical similarity and organic interconnections between these sep- arate plant organs. It is during the Mesozoic that we see the origin of a number of morphological features present in modern-day Isoetes. These features include: 1) a change in microspores from trilete to monolete suture (and probably also change in tetrad arrangement); 2) sporangia that are sunken into the adaxial surface of the spo- rophyll; 3) an elaboration of the basal portion of the ligule into a glossopo- dium; 4) the origin of a velum and/or a labium, covering the proximal, distal, or entire upper surface of a sporangium. All of these features can be seen among various Triassic-aged plants. Micro- sporophylls of Annalepis are characterized by monolete microspores of the Aratrisporites (Leschik) Playford & Dettman type (Grauvogel-Stamm and Lu- gardon, 2001). Sporangia are notably sunken into the sporophyll surface in several species of Pleuromeia such as P. rossica Neuburg and P. longicaulis (Burges) Retallack (Neuburg, 1960, 1961; Retallack, 1975). Ligules with an ex- tended basal glossopodium much like that present in many extant species of Isoetes occur in the permineralized Triassic Russian plant Takhtajanodoxa (Snigirevskaya 1980a, 1980b; Pigg, 1992). Whereas it appears that no single currently known Triassic isoetalean taxon possesses all of these features, all are found within various representatives of Triassic lycopsids. As the Triassic and later Cenozoic lycopsids are better doc- umented, it will be possible to better understand character evolution in the various lineages of Triassic lycopsids (Grauvogel-Stamm and Lugardon, 2001). This suggests the Triassic was a time of important radiation and change in several key morphological characters prior to the appearance of the extant, reduced Isoetes forms. The origin of the modern Isoetes plant habit with a nonelongated stem was previously thought to occur in the Triassic (Pigg, 1992). However, evidence for this type of plant habit is scant, as the best documented Triassic Isoetes-like 108 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) ~, ¥ a ee ss a \ j Fics. 2-5. Tertiary and extant isoetalean lycopsids. Fig. 2. cf. Isoetites from Dennison Gap, Wy- oming. Florida Museum of Natural History. UF 18120-12604. Fig. i ni s PIGG: ISOETALEAN LYCOPSIDS 109 Although it is clear that all of these forms are isoetalean plants, none show undisputed evidence of a non-elongating stem such as is usually characteristic of modern Isoetes. They should certainly be included within the Isoetales, based on their particular morphological features, yet I am reluctant to accept them as belonging to the modern genus. The earliest strong evidence of the modern plant habit is that seen in the Jurassic Isoetites rolandii Ash & Pigg from western North America (Ash and Pigg, 1991). Concomitant with this change in plant habit is the loss of vege- tative leaves: modern Isoetes and its younger fossil relatives typically produce only fertile leaves, thereby minimizing the vegetative phase of its life history. CRETACEOUS AND TERTIARY—SOME NEW OCCURRENCES Over a dozen Isoetes-like plants have been described in Cretaceous and Ter- tiary strata (see Pigg, 1992, and Skog and Hill, 1992 for review). Several newly recognized forms have been recently described. One unusual example is Mon- ilitheca Krassilov & Makulbekov described from the Upper Cretaceous of Mon- golia by Krassilov and Makulbekov (1996). The genus consists of fragmentary linear megasporophylls up to 18 mm long and 1 mm wide that bear within the sporangium, a single row of megaspore tetrads. These tetrads are borne along almost the entire length of the megasporophyll, which has a short sterile tip. Megaspores are trilete, and reticulat Isoetalean plants described from the ‘Lower Cretaceous of Tunisia were named Isoetites daharensis Barale (1999). These fossils are represented by corms 10 cm long X 6 cm wide with attached sporophylls and roots. Sporo- phylls are ligulate and bear sporangia containing trilete megaspores 480 pm in diameter. Isoetalean plants have also been recognized in the lower Creta- ceous of Brazil, in the Santana Formation (Dilcher et al., 2000). In this flora, Isoetes-like plants are represented by clusters of sporophylls that occur along with ferns, conifers, angiosperms, and abundant gnetalean remains. In the Tertiary, plants from the Paleocene Joffre Bridge locality in central Alberta, Canada originally identified by Hoffman as lilean monocots are clearly isoetalean (Hoffman, 1995, Plate 14, figs. 101, 102). The specimen is a plant base approximately 5.5 cm long and around 3 cm high with a crown of elon- gate leaves at the top of the specimen. At the base dense, helically arranged, rounded stigmarian rooting scars are present, a feature diagnostic for the is- oetalean clade (Fig. 4). Around twenty stigmarian rootlets, each measuring 3 mm wide and over 7 cm long, are found in attachment to the plant base. Additional isoetalean specimens are known from the Lower Eocene Denni- son Gap locality of the Wasatch Formation near Sweetwater, Wyoming, where they occur with platycaryoid foliage, fructifications and Azolla (Steven R. Manchester, Florida Museum of Natural History, written communication, 1992; Fig. 2, 3). These specimens are probably assignable to either I. horridus or I. serratus of Brown, depending on details of the leaf margin that are not readily visible, and show the characteristic air channels that are often found in Isoetites leaves (Fig. 3; Brown 1939, 1958, 1962). 110 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) CONCLUSIONS AND SUGGESTIONS FOR FURTHER RESEARCH In recent years, valuable new types of information have appeared that prom- ise to help clarify long-held questions about the evolution of the isoetalean clade. Notable are newly discovered 1) Middle Devonian forms from China that may have some bearing on the origins of this order, 2) permineralized Triassic fossils that provide a missing “middle” to our knowledge of anatom- ical features between Paleozoic forms and extant Isoetes, and 3) Cretaceous and Tertiary forms that document additional diversity of the group. Phyloge- netic analyses of anatomically preserved Carboniferous forms have been suc- cessful in delimiting interrelationships among the most complex isoetaleans which have been documented as “whole plants” and are understood ecologi- cally (e.g., Bateman et al., 1992; Phillips and DiMichele, 1992). However, to date it has not been possible to extend these types of studies beyond that stratigraphic range because of the limitations of preservation and difficulty of homologies. Whereas researchers studying Devonian, Pennsylvanian or Triassic aged fos- sils or extant Isoetes have made partial comparisons, homologies of characters through geological time have been more difficult. Part of the problem has been the use of terminology developed for one group of lycopsids for a different group. In some cases, terms and concepts have had to be “stretched” to ac- commodate real differences between isoetalean lycopsids of different ages. What is needed is an analysis of overall diversity within given characters, such as sporophyll morphology, without terminology imposed a priori. Lastly, with all of the work done on fossil and extant isoetaleans to date, the question of how to define the modern genus Isoetes in relation to its fossil relatives has never been resolved. In the present contribution I have referred to the extant Isoetes plant habit as having a reduced, non-elongating stem. However, there is considerable morphological variation in extant species. In addition to “typical” plants there are forms with branching axes such as I. andicola (formerly Stylites Amstutz), and the unusual rhizomatous mat-form- ing I. tegetiformans Rury (Fig. 5). Some species of Isoetes have even been pg to have small rhizomorphic structures (W. Carl Taylor, pers. comm. 1999). However, despite this diversity, it is well known among neontologists that morphology has been of little value in understanding speciation in Isoetes because of the high degree of homoplasy and convergence to habitat (e.g., Taylor and Hickey, 1992). Despite numerous discussions, it is still unclear what is basal in the group based on morphology because of the strong imprint of ecological influence. Molecular studies that clarify extant species relation- ships may help us resolve this problem and offer clues to the relationships of the extant and relatively recent fossil isoetaleans. Further developmental stud- ies of Isoetes may also aid in a better understanding of homologies. Neverthe- less, the fossil record still promises to resolve much more about diversity and evolution within this interesting group throughout its long evolutionary his- tory. PIGG: ISOETALEAN LYCOPSIDS a1 ACCKNOWLEDGEMENTS I thank W. Carl Taylor, N. Wickstrum, and Lea Grauvogel-Stamm for inviting my ee to this symposium and the American Fern Society for travel funds to the symposium; Char M. Christy, Walter L. Cressler, Melanie L. DeVore, Else Marie Friis, Patricia G. Gensel, judith - Gordon, David M. Jarzen, Steven R. Manchester and Ruth A. Stockey for access to material and field localities and assistance; Melanie L. DeVore, Patricia G. Gensel, Lea Grauvogel-Stamm, R. — Hickey, Gar W. Rothwell and Wesley C. Wehr for their comments on the manuscript; Ste- 9980388 REFERENCES ANDREWS, H. N., C. B. READ, and S. H. MAMAy. 1971. A Devonian lycopod stem with well-pre- served cas tissues. Palaeontology 14:1-9. aaa ns A. 1933. A lyc aaa strobilus from the Pocono Sandstone of Pennsylvania. r. J. bot. 20:114-11 pay . A. 1935. Notes on some American species of Lepidostrobus. Amer. J. Bot. 22: 23-25. ARNOLD, C. A. eis Observations of fossil plants from the Devonian of eastern North America. ant remains from the Catskill Delta deposits of northern Pennsylvania and southern ae York. Contrib. Mus. Paleontol., Univ. Mich. 5:271-314 ASH, S. R., and K. B. PicG. 1991. A new Jurassic Isoetites [feockales) from the Wallowa Terrane in Hells Canyon, Oregon and Idaho. Amer. J. Bot. 78:1636—1642. BARALE, G. 1999, Sur la présence d’une nouvelle espéce d’Isoetites dans la flore du Crétacé infér- ieur de la région Bote (Sud tunisien): implications paléoclimatique et phylogéné- tiques. Can. fe = 196. BATEMAN, RM gv roan’ reconstruction, pperiiberait and phylogeny of Oxroadia gracilis oa panes and O. conferta. sp. nov.: Anatomically preserved egal ly- copsids from the Dinantian of Oxroad Bay, SE Scotland. eee 228B:29— BATEMAN, R. M., W. A. DIMICHELLE, and D. A. WILLARD. 1992. Experimental cladistic oer of anatomically preserved arborescent lycopsids from the Carboniferous of Euramerica: an essay on paleobotanical phylogenetics. Ann. Missouri Bot. Gard. 79:500—559. BEcK, C. B. 1958. “Levicaulis arranensis”’, gen. et sp. nov., a anaes axis from the Lower Car- Bock, W. 196 tudy on fossil Isoetes. J. Paleontol. 36:53— Brack, S. D. 1970. On a new structurally preserved acicxtecont lycopsid fructification from the Lower Pennsylvanian of North America. Amer. ot. 57:317-330. BRACK, S. D., and T. N. eae 1972. The alteeicoctare and organization of Endosporites. Mi- mae on 18:101—10 BRACK-HANES, S. D., and B. A. te 1983. A re-examination of Lepidostrobus Brongniart. Bot. J. Linn. Soc. 86:125-133. BROWN, r hy 1939. Some American fossil plants belonging to the Isoetales. J. Wash. Acad. Sci. —269. BROWN, - om 1958. New occurrences of the fossil quillworts called Isoetites. J. Wash. Acad. Sci. 48:358—-361. BROWN, R. W. 1962. Paleocene flora of the Rocky Mountains and Great Plains. Prof. Pap. U. S. Geol. Surv. 375:1-119. Cal, C, and X. Wu, 1994. First discovery of Cyclostigma- ‘gine stem ent — pit from the Upper Devonian of Chaeohu i Anhui. Acta Palaeontol. Sin. 33:81— Cai, C., sees L. CHEN. 1996. On a Chinese Givetian Icopod, Longostachys ar ee ng and Feng, emend.: = morphology, anatomy an CANTRILL, D. J., and J. A. Wess. 1998. Permineralized pleuromeid lycopsid remains from the Early Triassic Arcadia Formation, Queensland, Australia. Rev. Palaeobot. Palynol. 103:189-211. 112 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) CHALONER, W. G. 1958. Polysporia mirabilis Newberry, a fossil lycopod cone. J. Paleontol. 32:199- 209. CHALONER, W. G. 1967. ery Pp. 437-845, in: E. Boureau, ed. Traité de Paléobotanique. I. Masson et Cie, Pari CHALONER, W. G. 1968. The cone ir ae ir ap kiltorkense Haughton, from the Upper Devonian of Ireland. J. Linn. Soc., Bot. CHITALEY, S, and K. B. Picc. 1996. Sah ade as ohioensis, gen. et sp. nov., y abeuey iain lycopsid from the Late Devonian Cleveland Shale of Ohio. Amer. J. Bot CICHAN, M. A., and C. B. BECK. 1987. A woody lycopsid stem from the New pea aa ‘ixaat Mi ssp 4 Kentucky. Amer. J. Bot. 74:1750-1757 CRESSLER, W. 1 ead Late ee flora ceeerved in a fluvial setting, north-central sue. U.S.A. Amer. J. Bot. 8 CRESSLER, W. L. 1999. Site- saahee and ris of the Late Devonian Red Hill locality, Pennsyl- vania, an Archaeopteris-dominated plant community and early ye te site. Unpublished Ph.D. dissertation, University of an. Philadelphia, PA, 1 DILCHER, D. L., A. F. ARIM-DE-LACERDA, A. M. F. BARRETO, and M. E fe BERNARDES-DE-OLIV- EIRA. ome bagigrs fossils from the Santana Formation, Chapade do Araripe, Brazil. Amer. J. Bot. DIMICHELE, “ a “1980. Paralycopodites Morey & Morey, from the Carboniferous of Euramerica. A reassessment of the generic affinities and evolution of “Lepidodendron” brevifolium Wil- liamson. Amer. J. Bot. 67:1466—1 ston DIMICHELE, W. A., and R. M. BATEMAN. 1992. Diaphorodendraceae, fam. nov. (Lycopsida: Carbon- iferous): systematics and Lonel relationships of Diaphorodendron and Synchysiden- ron, gen. nov. Amer. J. Bot. 79:605-617 DIMICHELE, W. A., and R. M. BATEMAN. 1996. The bara eye lycopsids: a case-study in paleo- bietaniesl classification. Syst. Bot. 21:535-5 DIMICHELE, W. A., J. F. MAHAFFY, and T. L. ee 1979. Lycopods of Pennsylvanian age coals: Polysporia. Can, J. Bot. 57:1740-1753. DOoBRUSKINA, I. A. 1985. Some problems of the systematics of the Triassic lepidophytes. Palaeon- tological Journal 19:74—88 DRABEK, K. 1976a. Polpsporic mirabilis ronal 1873 from the Nyfany locality (Westphalian D). Cas. Nar. Mus., Odd. Prir. 145:93-9 DRABEK, K. eating Polysporia robust de n. from the Carboniferous of Central Bohemia. Cas. Nar. Mus., Odd. Prir. 145:204— Fucus, G., L. GRAuVvOG uot, and D. MADER. 1991. Une remarquable flore a Pleuromeia et Anomopteris in situ du Buntsandstein moyen (Trias inférieur) de l’Eifel (R. is Allemagne). pre ae pe dap caisson et paléogéographie. rao sam aoc 222B:89-120. GASTALDO, R. - An ultrastructural and taxonomic study of Valvisporites nik (Zerndt) Bhardwaj, a yop ‘eimai from the mide Pennsylvanian of southern I]linois. Micro- paleontology 27 GENSEL, P. G. and C. hr sca 2001. Early lycophyte evolution. Amer. Fern J. 91:99— Girrorb, E. M., and A. S. FosTER. 1989. Morphology and evolution of vascular plants. aa edition. W. H. Freeman & Co., New York. emer hens cs ace L. 1993. Pleuromeia sternbergii (Miinster) Corda from the Lower Triassic of any t observations and comparative morphology of its rooting organ. Rev. Pa- nee Palynol. 77:185—212. GRAUVOGEL-STAMM, L., and P. DuRINGER. 1983. Annalepis zeilleri FLICHE 1910 emend., un organe reproducteur de lycophyte de la Lettenkohle de |’Est de la France. Morphologie, spores in et paléoécologie. Geologische Rundshau 72: be GRAUVOGEL-STAMM, L, and J. LANGIAUX. 1995. Polysporia doubingeri n. sp., un nouvel organe reproducteur de lycophyte du Stérhanion (Carbon eg she ane de Blanzy-Montceau (Mas- sif Central, France). Sci. Géol. Bull. Strasbourg 48:6 rine TAMM, L. and B. LuGaRDON. 2001. The waite cops Pleuromeia and Annalepis: relationships, evolution, and origin. Amer. Fern J. 91:115-14' PIGG: ISOETALEAN LYCOPSIDS 113 GRIERSON, J. D., and H. P. BANKs. mes esac of the Devonian of New York state. Palaeonto- graphica Araceae 4 (31):220- Hanes, S. D. 5. Structurally staan lepidodendracean cones from the Pennsylvanian of Nort Asin Unpublished PhD hic’ Ohio University, Athens, Ohio HorrMaN, G. L. 1995. Paleobotany and paleoecology of the Joffre Bridge readout locality (Paleo- cene), Red Deer, Alberta. Unpublished Master’s thesis, University of Alberta, Edmonton, Can- JENNINGS, Re 1975: sion ae a new plant organ from the Lower Mississippian of Virginia. pe pons 18 eh oe Z RE: E aed LT, and G. bs huihipsiniaes 1983. Structure and affinities of Protostig- a eggertiana. Amer. J. Bot. 70:9 ice. T 1913. On ms NyccaoraNg pamela kiltorkense Haughton sp. Sci. Proc. Roy. Dub- lin Soc, 13:500-52 KARRFALT, E. E. 1984. is observations on Nathorstiana (Isoetaceae). Amer. J. Bot. 71:1023— 30 1030. KENRICK, P, and P. R. CRANE. eign lip origin and early diversification of land plants. Smithsonian Institution Piosa, Washing peorcacnnld V. A., and N. M. Sane 1996. Isoetalean Sitio sganea ag a megaspores m the Upper Cretaceous of Mongolia. Rev. Palaeobot. Palynol. 94: siidtoas, R., and H. WEYLAND. 1949. Gilboaphyton und die a cern met Seawiadick 0:129-15 MATHEWS, G. B. 1940. New lepidostrobi from central United States. Bot. Gaz. 102:26—49. MarTEN, L. C. 1989. A petrified lycopod from the Uppermost Devonian of Hook Head, County Wexford, Ireland. Bot. Gaz. 150:323-336. MEYER-BERTHAUD, B. 1984. Les axes de lycophytes a structure anatomique conservée du carboni- fere basal (Tournaisien) de la Montagne Noire: Trabicaulis gen. nov. et Landyrodendron gen. nov. Palaeontographica 190B:1-36. NEUBERG, M. F. 1960. Pleuromeia Corda from the Lower Triassic deposits of the Russian Platform. Trudy Inst. Geol. Nauk. Ser. Stratigr. 43:63-94 [In R an]. NEUBERG, M. F. 1961. New data on the morphology of Pree Corda from the Lower Triassic of the Russian Platform. Dok. Akad. Nauk. SSSR 136:445—448. [In Russian]; 200-203 [English summary]. Puituips, T. L. 1979. Reproduction of heterosporous arborescent lycopods in the Mississippian— Pennsylvanian of Euramerica. Rev. Palaeobot. Palynol. 27:239-289 PHILLIPS, T. L, and W. A. DIMICHELE. 1992. Comparative ecology and life-history biology of “en ey lycopsids in Late Carboniferous swamps of Euramerica. Ann. Missouri Bot. Gard. 0-588. PicG, 4 B. 1983. es oe and reproductive biology of the sigillarian cone Mazocarpon. t. Gaz. 144: 600-61 PiGG, K . 1992. Evolution 2 isoetalean lycopsids. Ann. Missouri Bot. Gard. 79:589-612. Picc, K. B., and G. W. ROTHWELL. 1979. Stem-root transition of an Upper Pennsylvanian woody lycopsid. Amer. J. Bot. 66:914—924. Pic, K. B., and G. W. ROTHWELL. 1983a. Chaloneria, gen. nov.: heterosporous lycophytes from the Pennsylvanian of North America. Bot. Gaz. 14 —147. Pica, K. B., and G. W. ROTHWELL. 1983b. Mesapaietaphyto development in the Chaloneriaceae fam. nov., permineralized ae Iseotales (Lycopsida). Bot. Gaz. 144:295-302. Picc, K. B., and G. W. ROTHWELL. 1985. Cortical development in Chaloneria cormosa (Isoetales), and the biological derivation of compressed lycophyte decortication taxa. Palaeontology 28: 545-553. RETALLACK, G. 1975. The life and times of a Triassic lycopod. Alcheringa 1:3-29 RETALLACK, G. J. 1997. Earliest Triassic origin of Isoetes and quillwort evolutionary radiation. J. Paleontol. 71: icatiser ROTHWELL, G. W. . The — of Stigmaria (Lycopsida), rooting organ of Lepidodendrales. Aine J. Bot. is pirohe 114 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) ROTHWELL, G. W., and D. M. ERwIN. 1985. The rhizomorph apex of Same implications for homologies among the rooting organs of Lycopsida. Amer. J. B 6-98. ROTHWELL, G. W., and J. S. PRvyor. 1991. Developmental dynamics cs jee nbneaete lycophytes— apical and lateral growth in Stigmaria ficoides. Amer. J. Bot. 78:1 Roy, . A., and L. C. MaTTEN. 1989. Lycopods from the New Albany oe Palaeontographica 12B:1—45. Pa aig S. E., W. L. CRESSLER, T. CONNERY, S. KLAVINS, and D. L. POSTNIKOF. 1999. Devonian shrub and tree dominated landscapes. XVI International Botanical Congress, Abstracts: 13. SCHOPrF, J. M. 1941. Contributions to Pennsylvanian paleobotany: Mazocarpon oedipternum sp. nov. and sigillarian pies oh Illinois Geological Survey, Report of Investigations 75:1-40 ae H. J. 1969. Die Oberdevon-Flora der Bareninsel. 2. Lycopodiinae. Pelvouiaabsod ees SkoG, J. E., and C. R. HILL. 1992. The Mesozoic herbaceous lycopsids. Ann. Missouri Bot. Gard. 79:648-675 merninas N. S. 1980a. Takhtajanodoxa Snig., a new link in evolution of lycopsids. Pp. 4 5 «oy oy me Systematika i evolyutsiya vysshikh rastenii, Nauka, Leningrad e Hus sian]. sais Ug on N. S. 1980b. A new fossil genus of Isoetopsida in the early Triassic of East Siberia. Zarn. SSSR 65:95—96 [In Russian]. ena N. S., and I. N. SREBRODELSKAYA. 1986. Petrified stems rf pe (Lycopo- diophyta) with a preserved anatomical structure. Bot. Z*urn. 71:4 STEWART, W. N. 1947. Comparative study of stigmarian ahah ah Pe see roots. Amer. J. STEWART, W.N., and G. W. ROTHWELL. 1993. Paleobotany and the evolution of plants. 2nd edition, Cambridge University Press, Cambridge. TayYLor, W. C., and R. J. Hickey. 1992. Habitat, evolution, and speciation in Isoétes. Ann. Missouri. Bot. Gard. 79:613-622. THOMAS, B. A. 1981. Structural adaptations shown by the Lepidocarpaceae. Rev. Palaeobot. Pa- lynol. 32:377-388. WaGneR, R. H. 1989. A late Stephanian forest swamp with ee ee cogs ae by volcanic ash fall in the Puertollano Basin, central Spain. Int. J. Coal Geol. 1 WANG, Z. 1991. Advances on the Permo-Triassic lycopods in North ae : ae rae from the mid-Triassic in northern Shaanxi Province. Palaeontographica 222B:1 White, D. 1907. A remarkable fossil tree trunk from the middle Devonic of Now York. Bull. New Wnire, M. E. 1981. Cylomeia undulata (Burges) gen. et comb. nov., a lycopod of the Early Triassic strata of ai South Wales. Rec. Austral. Mus. 33:723-734. American Fern Journal 91(3):115-149 (2001) The Triassic Lycopsids Pleuromeia and Annalepis: Relationships, Evolution, and Origin LEA GRAUVOGEL-STAMM EOST-Géologie, Université Louis Pasteur, 1 rue Blessig, 67084 Strasbourg, and ISEM, UMR 5554 CNRS, Montpellier, France BERNARD LUGARDON Université Paul Sabatier, Laboratoire fae Biologie Sa 39 Allées Jules Guesde, 31000 oulouse, Franc ABSTRACT.—Two kinds of isoetalean lycopsids widely prevailed in the Triassic, the Pleuromeia- type and the Annalepis-type, the latter including a plexus of closely related genera. Comparative terconnected and closely related to Isoetes. Morever they suggest that Annalepis is p n- tral to Isoetes, via Isoetites. Besides several of the morphological and ultrastructural features of the Triassic lycopsids and Isoetes also appear to be present in some of the most anci ycop- sids, suggesting that the lineage including the modern Isoetales has a very remote origin. The lycopsids have the longest evolutionary history of any vascular land plants, covering about 400 million years and spanning every geological period from the Siluro-Devonian to the present. Although their structural diversity has decreased considerably since their peak in the Carboniferous, they were still widespread in the Triassic of both hemispheres. However the Mesozoic lycopsids were not arborescent like many of the Carboniferous ones but rather consisted of slender, herbaceous or pseudoherbaceous plants having an un- branched habit. Two kinds of lycopsids widely prevailed in the Triassic, the Pleuromeia-type and the Annalepis-type, the latter including a plexus of close- ly related genera, which are all regarded as belonging to the isoetalean lycop- sids (Pigg, 1992). Pleuromeia Corda was long regarded as one of the intermediates between the Carboniferous arborescent lepidodendrids and living Isoetes and thus thought to be part of a reduction series (Solms-Laubach, 1899; Potonié, 1904; Magdefrau, 1931). This concept was questionned by Jennings (1975) who no- ticed that the lycopsids with a cormose rhizomorph already existed in the pper Devonian, long before the lepidodendrids with a dichotomously branched Stigmaria rooting system. However DiMichele and Bateman (1996) also suggested, on cladistic foundations, that the isoetaleans evolved from the lepidodendraleans and that the bilateral condition emerged from the radial one in the Devonian. Nevertheless Bateman (1994, 1996) recognized that the precise relationships between both groups are ambiguous The comparative study of Pleuromeia and Annalepis presented here, using new macromorphological and ultrastructural data, shows that these Triassic isoetalean lycopsids appear closely related to Isoetes and that Annalepis seems to be closer than Pleuromeia to the living genus. Moreover this study consti- 116 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) tutes a new basis for comparison with other lycopsids and for evaluating and possibly clarifying the relationship between the isoetaleans and the lepidod- endrids. THE TRIASSIC LYCOPSID PLEUROMEIA: COMPARATIVE STUDY Although the lycopsid Pleuromeia Corda was long regarded as being phy- logenetically highly significant, it was never estimated at its true value because many characteristics of this widespread Triassic genus were not known with enough accuracy. The progress in knowlege of its rhizomorphs, growth habit, reproductive organs and spores now allows to propose a comprehensive com- parative analysis of this outstanding genus. THE RHIZOMORPH OF PLEUROMEIA—As in the other rhizomorphic lycopsids (Bateman 1996), the rooting organ of Pleuromeia has proven to be, phyloge- netically, highly informative since it shows many similarities with that of Is- oetes, suggesting that these lycopsids are closely related. The reinvestigation and morphological analysis of several fairly well-preserved rhizomorphs of P. sternbergii (Miinster) Corda in light of new data on Isoetes permit us to dem- onstrate the structural and developmental correspondence between both. Thus most of the features of the rhizomorph of P. sternbergii can be interpreted in terms of those of the living genus (Grauvogel-Stamm, 1993). Like the rhizomorph of Isoetes, that of Pleuromeia is lobed and shows a bilaterally symmetrical furrow system, when seen from below (Figs.1a-f). The four-lobed rhizomorph of P. sternbergii has a central furrow bifurcating distally into two peripheral furrows which run along the midline of the lobes and extend upward into their extremities (Fig. 1c). Its roots have the same structure as those of Isoetes and Stigmaria. As in Isoetes but unlike Stigmaria, they are arranged on both sides of the furrows according to two intersecting row sys- tems, one roughly parallel to the furrows (series) and the other diverging from them at a relatively high angle (orthostichies) (Fig. 1f). In the two-lobed rhi- zomorph of P. epicharis Wang and Wang (1990) from the Lower Triassic of China, the roots are also arranged in series and orthostichies on both sides of the straight furrow, as in the two-lobed rhizomorph of Isoetes (Figs.1a,b,e). This bilateral arrangement suggests that the roots were produced as in the living genus. That is, they are produced by a basal linear root-producing mer- istem underlying the lobed stele and running parallel to the furrow, and the roots emerged at the furrows and were progressively displaced by continued meristematic activity. The lobing of the rhizomorph developed in relation to this meristem and this process of root production. However the difference in cortical development in Pleuromeia and Isoetes resulted in a difference in lobe and furrow arrangement (Figs.1a-d). Indeed in Isoetes where the cortex is strongly developed, the peripheral furrows of a four-lobed rhizomorph seem to run between the lobes (Fig. 1d). In fact, the correct developmental corre- spondence requires comparing the lobes of Pleuromeia with those of the root- bearing stele in Isoetes, without its cortical lobes. Observations on the diameter GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 117 --O- - -- Oo rath ‘ = es ’ - * On. . . . a Org ’ Fic. 1. Comparative morphology of the rhizomorphs of Pleuromeia and Isoetes. a—Side-view (above) and underside view (below) of a two-lobed rhizomorph of Pleuromeia, such as those of P. epicharis (Wang and Wang, 1990) or P. sanxiaensis (Meng, 1995) from the Triassic of China. b— Cross section of a two-lobed rhizomorph of Isoetes at the level of the basal meristem and furrow (inferred from Karrfalt and Eggert, 1977). This rhizomorph has a straight furrow like the two-lobed rhizomorph of Pleuromeia in Fig. 1a, but its two lobes (stippled) are more developed due to the great cortex production. c—Side-view (above) and underside view (below) of a four-lobed rhizo- morph of Pleuromeia, like that of P. sternbergii (Grauvogel-Stamm, 1993) or P. rossica (Neuburg, 1960; Dobruskina, 1982). The bilaterally-symmetrical furrow system consists of a central furrow lobes. d—Cross section of a four-lobed rhizomorph of Isoetes at the level of the meristem and furrow system (inferred from Karrfalt and Eggert, 1977). The farrow-system is at to that of Pleuromeia in Fig. 1c, but the high cortex production resulted in the development of cortical lobes (stippled) between those of the root-bearing stele. Thus the correct developmental correspondence requires comparing the lobes of Pleuromeia with those of the root-bearing stele in Isoetes (cortical rotrusions excluded). e—Root arrangement on both sides of the straight furrow in a two-lobed high angle (orthostichies) (modified from Karrfalt and Eggert, 1978). f—Root arrangement in series and orthostichies on both sides of the bifurcated furrow system in the four-lobed rhizomorphs of Pleuromeia or Isoetes (modified from Karrfalt and Eggert, 1978). of the root scars of P. sternbergii in relation to their place on the rhizomorph demonstrate that root production proceeded in an acropetal direction as in Isoetes. The significant increase of the root scar diameter from youngest to oldest (also observed in P. epicharis, unpublished observations, L. G.-S.), their numerical increase with the aging and thickening of the rhizomorph, and the maintenance of their arrangement give evidence for long retention of the roots, 118 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) possibly their persistence throughout the life of the plant as in Selaginella. This is unlike Isoetes in which the roots are regularly lost and replaced. The sequence of root initiation, the orientation of the root traces inside the lobes and thus the direction of their emergence and the marked lengthening of the peripheral furrows which occurred during ontogeny show how the rhizo- morph of Pleuromeia developed laterally to the detriment of its downward growth. This new restatement of the structure of the rhizomorph of P. stern- bergii and the comparative study with the living Isoetes accurately emphasizes the numerous points of structural and developmental correspondence between these plants. Fossil lycopsids with a cormose rhizomorph that is furrowed, lobed and bilaterally symmetrical are rather rare in the literature. Besides P. sternbergii, such a rhizomorph has been described in P. rossica Neuburg from the Lower Triassic of the Russian platform (Neuburg, 1960; Dobruskina, 1982) and P. ep- icharis from the Lower Triassic of China (Wand and Wang, 1990). The rhizo- morphs of P. sternbergii and P. rossica are four-lobed and quite similar (Fig. 1c). In contrast, the rhizomorph of P. epicharis which is two-lobed, shows a straight, unbranched furrow (Fig. 1a). Also P. sanxiaensis Meng (1995) from the early Middle Triassic of South China has a two-lobed rhizomorph. Besides Pleuromeia, three other lycopsids with lobed, furrowed, and bilaterally sym- metrical rhizomorphs have been described: Protostigmaria eggertiana Jennings from the Lower Carboniferous of Virginia, USA (Jennings, 1975; Jennings et al., 1983), Nathostianella Glaessner and Rao (1955) from the Lower Cretaceous of South Australia, and Nathorstiana Magdefrau from the Lower Cretaceous of Germany (Karrfalt, 1984). Unlike P. sternbergii, they have no prominent lobes. In the lobed rooting organ of Cylomeia White (1981) from the early Triassic of Australia, no furrows have ever been mentioned and the other data are too imprecise to interpret this feature. In Chaloneria cormosa Pigg and Rothwell from the Pennsylvanian of North America, which is said to have a rounded, slightly lobed, and bilaterally symmetrical cormose rooting organ (Pigg and Rothwell, 1979; 1983; Pigg, 1992), the furrow system has not been mentioned and the root arrangement has not been described. The lobe and furrow system of Chaloneria was not fully documented because of the incomplete preserva- tion of cortical tissue on plant bases of available specimens (Pigg, personnal communication, 1999). Among the cormose lycopsids, there are others in which the rhizomorph is radially or nearly radially symmetrical, such as Pau- rodendron fraiponti Fry from the Upper Pennsylvanian of Ohio, USA (Roth- well and Erwin, 1985) and Oxroadia gracilis Alvin from the Lower Carbonif- erous of Scotland (Stewart and Rothwell, 1993). Moreover, several other cor- mose lycopsids clearly lack an extensive branched rooting system but are not well enough preserved to permit a precise study of their rhizomorph. These are Lepidosigillaria whitei Krausel and Weyland, Cyclostigma kiltorkense Haughton, possibly Eospermatopteris erianus (Dawson) Goldring from the De- vonian (see Pigg, 1992, 2001), Bodeodendron Wagner/Sporangiostrobus feist- mantelii (Feitmantel) Némejc from the Late Stephanian of Spain (Wagner, 1989) and Clevelandodendron ohioensis Chitaley and Pigg (1996) from the GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 119 Late Devonian of Ohio, USA. In all of the cormose lycopsids cited above, the structure of the rhizomorph, its development, and its root arrangement are not known with as much precision as in species of Pleuromeia. The bilaterally symmetrical cormose rhizomorphs of Pleuromeia and Isoetes seem to differ greatly from the radially symmetrical stigmarian rooting organs of the lepidodendrids which are dichotomously branched and in which the roots are helically arranged. However, Rothwell and Erwin (1985) demonstrat- ed that the radially and bilaterally symmetrical rhizomorphs are homologous and that they are merely growth variations rather than indicators of two major lines within the rhizomorphic lycopsids as suggested earlier by Jennings (1975) and Jennings et al. (1983). These homologies were the basis for includ- ing both the Isoetales and Lepidodendrales of the traditional classifications in the rhizomorphic or isoetalean clade (Rothwell and Erwin, 1985; Stewart and Rothwell, 1993; DiMichele and Bateman, 1996). GROWTH HABIT AND REPRODUCTIVE ORGANS OF PLEUROMEIA.—The similarity in rhizomorphic structure which suggests a close relationship between Pleuro- meia and Isoetes strongly contrasts with the differences in growth habit and reproductive structure of these plants. In Pleuromeia (Fig. 2c-g) the sterile leaves differ greatly from the fertile ones which form a well defined cone at the apex of the stem whereas in Isoetes (Fig. 2h, 7d), the fertile and sterile leaves are alike and do not form a cone (Jermy, 1990). However such differ- ences seem to be frequent in related lycopsids. The genera Polysporia New- berry and Chaloneria which belong to the Chaloneriaceae and produce mor- phologically similar microspores and megaspores, are characterized by well- defined terminal cones in Polysporia (Grauvogel-Stamm and Langiaux, 1995) and by fertile zones in Chaloneria in which the sporophylls resemble the ster- ile leaves (Pigg and Rothwell, 1983). Likewise, in the extant genera Phyllog- lossum and Huperzia of the Lycopodiaceae which produce comparable spores, there is a compact strobilus in the first genus while the second genus has either fertile zones or compact strobili (Ollgaard, 1990). All the Pleuromeia species, which have been discussed by Dobruskina (1985) and Pigg (1992), consist of unbranched plants having basal cormose rhizomorphs and terminal, well defined cones. The sizes of the plants are rather variable but many species are relatively small. Even among specimens of P. sternbergii from the type-locality (Bernburg, Germany) for which Mag- defrau (1931) indicated a height of 2—2.5 m, there are fertile specimens which only reach 23 cm (Fig. 2d) (Grauvogel-Stamm, 1999). Those from the Eifel in Germany (Fig. 2c) are 1-1.5 m tall (Mader, 1990; Fuchs et al., 1991). Pleuro- meia rossica from the Lower Triassic of Russia was no more than 1 m high (Neuburg, 1960) and P. jiaochengensis Wang and Wang (1982) from the Lower Triassic of China usually reached 20-30 cm high, sometimes 50 cm (Fig. 2e). Recently, several new species have been described from the early Middle Tri- assic of southern China (Meng, 1995,1996) which consist of plants mostly 50 cm high. However, P. marginulata Meng (Fig. 2f) includes specimens more AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Sy = r hig Sf eS5t ET ete ed aS itty! seecitainel a = ove © ® Fic. 2. Comparative growth habit of the Paleozoic and Triassic Pleuromeia-like lycopsids and the extant Isoetes. (scale-bar = 10 cm, except in Fig. 1g,h = 1 cm). a—Clevelandodendron ohioensis; Devonian of Ohio, USA (redrawn from Chitaley and Pigg, 1996). b—Chaloneria cormosa; Penn- ,1983). c—Pleuromeia sternbergii; Lower Triassic (Olenekian) of the Eifel, Germany (from Fuchs et al., 1991). d—P. sternbergii from the Lower Triassic of Bernburg, the type-locality, (reconstructed after the illustrations of Bischof, 1853, Figs.1,2). e—P. jiaochengensis; Lower Triassic (Induan) of Shanxi, China (redrawn from structed after the illustrations of Meng, 1995: Pl.1 fig. and Meng, 1996: Pl.1 fi iaensis; early Middle Triassic (Anisian) of South China (reconstructed after the illustrations of Meng, 1995, 1996, Pl.1 Fig. 9). h—Isoetes brochoni, showing leaves 3-7 cm long (redrawn from Motelay, 1893). than 1 m high. In contrast, P. sanxiensis Meng is only 5 cm tall, with a 2,5 cm long cone (Fig. 2g). The trunk of Pleuromeia is usually devoid of leaves but is covered with helically arranged leaf scars. However some of the specimens of P. sternbergii from Eifel had trunks covered with leaves (Fig. 2c). These leaves have a strong midvein and, according to Magdefrau (1931), they were succulent and contain transverse ridges extending perpendicularly on both sides of the midvein. As suggested by Grauvogel-Stamm (1999, Fig. 4), these transverse ridges might correspond to air chambers, as in Isoetes and Isoetites phyllophilla Skog, Dilcher and Potter (1992). However this feature requires further investigation. Cone size is variable in Pleuromeia. According to Magdefrau (1931), the cone of P. sternbergii reached 20 cm long. However, other cones of that species GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 121 are only 5 cm long (Bischof, 1853, Figs.1,2; Grauvogel-Stamm, 1999). Likewise according to Magdefrau (1931), the cones were monosporangiate, containing either microspores or megaspores. Nevertheless, the small cone figured by Bis- chof (1853) is clearly bisporangiate (Grauvogel-Stamm and Lugardon, in prep- aration). Its microsporophylls and megasporophylls are inter-mixed as in P. rossica (Neuburg, 1960). The cones of P. sternbergii appear to be mono- or bisporangiate depending perhaps on the size and the age of the plants. The two cones on which Magdefrau (1931) relied were rather large whereas that of Bischof (1853) is small. In Isoetes too, the size of the plants has an influence on the production of microsporophylls and/or megasporophylls (Williams, 1943). The sporophylls of the different Pleuromeia species show a remarkable sim- ilarity in shape and structure since they consist of a circular to oval lamina and a large, globose sporangium which covers nearly all of the adaxial surface (Figs.3a-j). Distally, the margin has a slight notch or a tiny pointed tip at the apex. The ligule is attached immediately distal to the sporangia and is often obvious in the apical border of the sporophylls (Figs.3a,c,d,h-j). Trabeculae similar to those of Isoetes have also been observed in the sporangia of several species. According to Snigirevskaya (1989), the sporophylls of Pleuromeia are similar to Isoetes leaves having lost their distal sterile portion. Indeed, as sug- gested by this author, they had a tip which was longer than usually depicted but which dried and decayed in maturing cones. However, examination by one of us (L.G.-S.) of sporophylls of P. sternbergii and P. rossica in side view shows that they have a tip extending from the lamina without direction change and that this tip is a little longer than the border of the sporophylls in face view (Grauvogel-Stamm, 1999), but never as long as suggested by Snigirevskaja (1989). Lycopsids with the same growth habit as Pleuromeia already existed in the Paleozoic: Clevelandodendron ohioensis Chitaley and Pigg (1996) from the Late Devonian of Ohio, USA and Chaloneria cormosa Pigg and Rothwell (1983) from the Pennsylvanian of North America. Clevelandodendron ohioen- sis consists of 1.25 m tall unbranched plants covered with helically arranged leaf scars, and possessing a well defined bisporangiate cone containing trilete microspores and megaspores (Fig. 2a). Its partially preserved rhizomorph is said to bear “thick appendages”. Chaloneria cormosa is also unbranched, and is 2 m tall with a bilaterally symmetrical cormose rhizomorph (Fig. 2b). Its reproductive organs consist of a terminal fertile zone up to 21 cm long, with alternating megasporangial and microsporangial regions resembling the sterile nes. Meyen (1987) has suggested that Pleuromeia evolved from Chaloneria, and Stewart and Rothwell (1993) state: ‘““Chaloneria seems to bridge the gap be- tween the Upper Devonian treelike lycopsids with their cormose rhizomorphs and the cormose Triassic Pleuromeia’’. The discovery of Clevelandodendron ohioensis may indicate that a lineage comprising Pleuromeia-like plants al- ready existed in the Late Devonian and was contemporaneous with the begin- ning of the radiation of arborescent lepidodendrids. This suggests that the di- 122 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) WE ¢ WOO OC) ® Fic. 3. Comparative shape of the sporophylls in Pleuromeia. Note the globose sporangium oc- cupying most of the upper surface; note also the ligular pit (Ip) distal to the sporangium, often visible in the narrow apical border of the sporophyll. (scale-bar, between h and i = 1cm). a—P. sternbergii; Lower Triassic of Germany (redrawn from Solms-Laubach, 1899; Potonié, 1904). b— Paltinis; Lower Triassic of North China (redrawn from Wang and Wang, 1989). c—P. sternbergii; Lower Triassic of North China (redrawn from Wang and Wang, 1989). d—P. rossica; Lower Triassic of Russia (redrawn from Neuburg, 1960; Dobruskina, 1982). e—P. epicharis; Lower Triassic of North China (redrawn from Wang and Wang, 1989, 1990). f—P. patriformis; Lower Triassic of North China (redrawn from Wang and Wang, 1989). g—P. jiaochengensis; Lower Triassic of North China (redrawn from Wang and Wang, 1982). h—P. sanxiaensis; early Middle Triassic of South China (redrawn from Meng, 1995, Fig. 3b,c). i—P. marginulata; early Middle Triassic of South China (redrawn from Meng, 1995, Fig. 3a). )j—P. hunanensis; early Middle Triassic of South China (re- drawn from Meng, 1995, Fig. 3d). versification within the Isoetales was well established prior to the Carbonif- erous (Chitaley and Pigg, 1996). It may also question the statement of Di- Michele and Bateman (1996) which suggests that the isoetalean line evolved from the lepidodendrids. THE SPORES OF PLEUROMEIA.—Like the structure of the rhizomorphs, the ultra- structure of the spores of Pleuromeia appears to be phylogenetically infor- mative, suggesting a close affinity with Isoetes. The microspores of both genera, especially, show many ultrastructural sim- ilarities (Lugardon et al., 1997, 1999). This is all the more noteworthy since Isoetes microspores have a rather large number of distinctive characteristics which clearly differentiate them from all the other extant spore types, as stressed below. In order to underscore the similarities in ultrastructure and GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 123 correspondences of the homologous walls of Pleuromeia and Isoetes micro- spores, a description of both is given hereafter (Fig. 4). Isoetes microspores are monolete spores easily recognizable, through light microscope, by an apparently bipartite sporoderm consisting of an ovoid, un- adorned inner part wrapped in a wide, thicker, usually slightly ornamented outer part (Fig. 4a). The inner part is generally regarded as the exospore in morphological studies, while the outer part is variously interpreted and termed (perine, sexine. . .). TEM studies (Lugardon, 1973, 1980) have shown that these microspores possess three distinct aceto-resistant walls (Fig. 4b). The inner- most one, which really represents the exospore, is a thin, almost plain-sur- faced, bilayered wall that is characterized mainly by its aperture and by special areas of the wall called “laminated zones’ ( = “zones pluristrates” in the initial description of Lugardon, 1973). The aperture is distinguished by a sim- ple reduction of the exospore thickness, without any other special modifica- tion. The laminated zones are small, appreciably thickened areas of the exo- spore arranged on both sides of the aperture, in which the exospore inner layer is tangentially cleft into several irregularly segmented laminae. The middle wall, that is called “para-exospore”’, consists of interconnected elements com- posing a spongy layer which is detached from the exospore all around the spores except in some points of the lateral regions of the proximal face; more- over this wall forms a high projection, usually divided apically into two lips, above the aperture. The outermost wall, which is termed perispore, is a rather thick, complex wall closely applied to the para-exospore surface. Ontogenet- ical fine studies (Lugardon, 1990; Tryon and Lugardon, 1991) showed that the para-exospore elements develop at the same time as the exospore outer layer, through simultaneous accumulation of the same sporopollenin on the exo- spore inner layer and on the tenuous framework of the para-exospore that is formed very early in the course of sporogenesis. Moreover, these studies showed that the perispore develops after the para-exospore and exospore com- pletion, with materials of different chemical nature, like the typical perispore of homoporous fern spores (it is generally admitted that a genuine perispore is formed only in pteridophytes having a plasmodial tapetum; nevertheless, TEM studies proved that, besides the microspores of recent Isoetaceae, those of a number of Selaginellaceae, as well as the spores of rather numerous Ly- copodiaceae, have a perispore comparable to that of the fern spores, although the lycopsids possess a secretory tapetum; Lugardon, 1990). Thus the wide sporoderm outer part of Isoetes microspores comprises two quite distinct com- ponents, i.e an uncommon structure, the para-exospore, that is ontogenetically related to the exospore, and a veritable perispore. The microspores of Pleuromeia are trilete and rounded or rounded-trian- gular in equatorial outline (Fig. 4c). The morphological studies achieved with light and scanning electron microscopes (Neuburg, 1960; Yaroshenko, 1975) showed that these microspores have an outer envelope largely separated from a much thinner inner part often called inner body. The outer envelope appears spongy and faintly ornamented. The inner body is roughly smooth-surfaced and reveals under the light microscope, especially in specimens having lost 124 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) . 4, Comparative ultrastructure of the apertural area in the microspores of Isoetes and Pleu- romeia (scale-bar = 10m in a, c and 2 pm in b, d). a—Isoetes—Diagrammatic representation of the microspore proximal face showing the outer envelope (OE) and the inner body (IB). The dotted spots on both sides of the monolete aperture mark the location of the laminated zones observed with TEM, but indiscernible in intact spores observed with light microscope. b—Isoetes. Ultra- structural features of the apertural area of a microspore sectionned along x-y, as shown in Fig. 4a. Note that there are three distinct resistant walls: the exospore (E) with the aperture (a) consisting of a thinner, non-folded wall portion, bordered with laminated zones (z); the para-exospore (PE) broken above the apical para-exospore lips. c—Pleuromeia—Diagrammatic representation of the microspore proximal face showing the outer envelope (OE) and the inner body (IB) with three interradial papillae (pap) arranged between the arms of the trilete aperture and visible with light microscope only in particular conditions of preservation. d—Pleuromeia—Ultrastruct features microspore apertural area sectionned along x-y, as shown in Fig. 4c. Two distinct walls are preserved: the exospore (E) with the aperture (a) and laminated (z) similar to those in Isoetes; e para (PE) which is thicker but structurally analogous with that of Isoetes microspores. in? th tuo GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 125 the outer envelope, three rounded papillae arranged between the rays of the trilete mark near the proximal pole (Fig. 4c). Ultrastructural investigations in Pleuromeia rossica (Lugardon et al., 1997, 1999) and P. sternbergii (Grauvogel- Stamm and Lugardon, in prep.) proved that the inner body is a two-layered exospore, uniformly solid and thick except in the proximal region where it shows two notable characteristics (Fig. 4d). In the aperture area indeed, the exospore simply appears thinner without other noticeable differentiation, as in Isoetes. Secondly, in each of the three areas lying between the aperture rays, the exospore shows a well-defined laminated zone which is quite similar to those of Isoetes microspores, and which obviously corresponds to one of the three interradial papillae observed with light microscope. The outer envelope wholly consists of interconnected elements which show the same electron per- meability and reactions to chemical reagents as the outer layer of the exospore, and are surely made up of the same sporopollenin. It is attached to the exo- spore in the peripheral regions of the proximal face. In the other spore regions, this outer envelope is usually separate from the exospore, and forms a well- marked protrusion divided apically into two lips above each of the three ap- erture rays. Thus, the ultrastructural features of the exospore are very similar in Pleu- romeia and Isoetes microspores, in spite of the fact that those of Pleuromeia are trilete whereas those of Isoetes are monolete. Likewise, the main features of the outer envelope of Pleuromeia microspores remarkably resemble those of the para-exospore in Isoetes, so that both walls obviously appear homolo- gous, and can be equally called para-exospore. Besides the trilete or monolete condition, the only noteworthy differences lie in the lesser development of the para-exospore and the presence of an extra wall, the perispore, in Isoetes. The combination of the exospore and para-exospore features which is quite identical in Isoetes and Pleuromeia microspores is also that of the dispersed spore genera Endosporites and Aratrisporites which are respectively produced by Chaloneria/Polysporia and Annalepis (Lugardon et al. 2000b, Grauvogel- Stamm and Lugardon, in prep.). The three interradial papillae have been clear- ly shown in Endosporites by Brack and Taylor (1972). This combination of features differentiates these microspores from all the other spore types of ex- tant, and probably fossil, pteridophytes. The aperture consisting of a simply thinner, unfolded exospore area appears unique, and contrasts with the well- marked prominent fold showing various outlines and more or less intricate structures that characterizes the aperture of most pteridophyte spores, includ- ing the microspores of Selaginellaceae (Lugardon, 1972, 1980, 1986; Tryon and Lugardon, 1991). Moreover, although the laminated zones are present in the microspores of many Selaginellaceae, they are totally missing in a rather im- portant number of species of this large family (Tryon and Lugardon, 1991: Figs. 231.21, 25, 26,28). Likewise, the para-exospore appears to exist only in very few species of Selaginellaceae in most of which the exospore is devoid of laminated zones, and the para-exospore is almost solid and does not form a sharp protrusion above the aperture (Tryon and Lugardon, 1991: Figs. 231. 26, 28). Among the recent and fossil lycopsids in which the spore ultrastructure 126 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) is known to date, the microspores which most resemble those of Pleuromeia and Isoetes are those of the present Selaginella selaginoides (Lugardon, 1972) and those from the Pennsylvanian lycophyte cone Selaginellites crassicinctus (Taylor and Taylor, 1990). In both, indeed, the microspores have an exospore with laminated zones and a fairly spongy para-exospore forming prominent lips above the aperture rays. However, these spores are unambiguously distin- guishable by the specially narrow, strongly marked apertural fold of the exo- spore (known only in Selaginellaceae microspores among recent pteridophyte spores, as it can be seen in Tryon and Lugardon, 1991). Thus, not only the microspores of Isoetes and Pleuromeia, as well as those of Annalepis (Aratris- porites) and Chaloneria/Polysporia (Endosporites), show very close ultrastruc- tural similarities, but also they differ markedly from the spores of all the other pteridophyte groups, including the especially variously structured micro- spores of Selaginellaceae. It is regrettable that there are no adequate data on the fine features of the microspores of other fossil rhizomorphic lycopsids, particularly the lepidod- endrids, which could be compared to those of Pleuromeia. The different TEM studies on the lepidodendrid microspores Lycospora, achieved by Thomas (1988), Taylor (1990), Hemsley and Scott (1991), Scott and Hemsley (1993), suggest that these microspores also have a thin exospore partly separated from a thicker outer envelope presumably equivalent to a para-exospore. However the Lycospora studied by these authors, like those presently investigated (un- published observations, B.L.), are rather poorly preserved and have not clearly shown the ultrastructural characteristics of their apertural area which, very likely, would provide significant information on the relationships of the lepi- dodendrids with the other lycopsid lineages. The megaspores of Pleuromeia are also trilete with a slight ornamentation and a spongy aspect rather comparable to that of the microspores (Zhelezkova, 1985; Marcinkiewicz and Zhelezkova, 1992). Ultrastructurally (Lugardon et al., 2000a; Grauvogel-Stamm and Lugardon, in prep.), their exospore consits of a very thin, solid inner layer and a thick spongy outer layer which constitutes most of the wall and is usually divided by a tangential gap in its innermost area, except near the aperture. These features, as well as those of the aperture, are comparable to those of the Isoetes megaspores (Lugardon, 1986; Taylor, 1993 ; Tryon and Lugardon, 1991). However they are also analogous to those of all the other, recent or fossil, lycopsid megaspores, so that they do not pro- vide any special phylogenetic information. Nevertheless, the ultrastructural study of especially well-preserved megaspores of P. rossica (Lugardon et al., 2000a) demonstrates that the spongy outer layer consists of elements which usually show a rod-like shape, are mostly arranged parallel to the wall surface and are rather widely spaced in moderately crushed spores. These features of the exospore outer layer, which are quite obvious and unvarying throughout the wall in those P. rossica megaspores, precisely represent the few character- istics regarded as distinctive of the isoetalean megaspores (Kovach, 1994). Thus the megaspores indicate also, although in a less demonstrative way than GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 127 the microspores, the relation joining the Triassic Pleuromeia to the living Is- oetes. Furthermore, the exospore inner layer of Pleuromeia megaspores includes, along both sides of the three apertural arms, a number of laminated zones which are structurally quite similar to those of the microspores. Such lami- nated zones, that have not been observed in Isoetes megaspores, undoubtedly correspond to the “papillae, pustules, cushions, nipple-like projections. . .” described in several morphological studies of fossil lycopsid megaspores. In particular, rather abundant papillae, i.e. 15-20 on either side of each apertural ray, can be observed with LM in the megaspores of P. sternbergii (Grauvogel- Stamm and Lugardon, in prep.). However, for the moment these structures do not provide any reliable information on the affinities of Pleuromeia since their distribution among the lycopsids is not clearly characterized. THE TRIASSIC LYCOPSID ANNALEPIS AND RELATED TAXA: COMPARATIVE STUDY. Besides Pleuromeia, there is another noteworthy Triassic lycopsid, Anna- lepis Fliche, which has long been misunderstood. The type-species, A. zeilleri Fliche from the Middle and Upper Triassic of eastern France, is represented by isolated “scales” described by Fliche (1910) who noticed their resemblance to seed scales of the conifer Araucaria but finally placed them as an incertae sedis. Grauvogel-Stamm and Duringer (1983) showed that the scales of A. zeil- leri are sporophylls belonging to the lycopsids and containing microspores assignable to Aratrisporites. Such sporophylls are often termed “scales” in descriptions, likely on the basis of their shape and maybe on the analogy of the wording used in the conifer seed scales. Undescribed new material from the Ladinian (late Middle Triassic) of Germany now allows to give some further details, particularly about the rest of the plant (Grauvogel-Stamm and Lugar- don, in preparation). GROWTH HABIT AND REPRODUCTIVE ORGANS OF ANNALEPIS ZEILLERI FLICHE.—The sporophylls of A. zeilleri are preserved as compressions having 2.5—5 cm long and being slightly trapezoidal in shape (Fig. 5a, 6a). They consist of a long and wide proximal portion flaring abruptly at the base which is somewhat thickened, widening distally and extending by a thick and short, roughly tri- angular distal portion ending with a small pointed tip. The presence of a trans- verse wide groove followed by a bulge, distal to the sporangium, between the proximal and distal portions in the sporophylls in compression and adaxial view, indicates that their junction was more or less at right-angle. Indeed, the groove corresponds to the zone of juncture of the two portions and represents exactly the place where the distal portion was straightened up (Fig. 5d-e). This right-angled juncture remained partly preserved and appears as a groove be- cause of the thickness of this zone while the thinner distal portion has been folded back by the sediment pressure and appears in the same plane as the proximal portion in the fossilized sporophylls. Sometimes the groove+ bulge appears as a transversal fold (Fig. 5f). Thus, when the sporophylls were packed 128 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) Fic. 5. The sporophyll of Annalepis zeilleri Fliche in adaxial view (a-c): (a) in compression, entire; (b) in compression but lacking the lateral wings; (c) in life condition (modified from Grau- the course of fossilization: (d) sporophyll in life condition, with both portions roughly perpen- dicular to each other (the arrow indicates the direction of flattening during fossilization); (e) flat- tened sporophyll with a bulge distal to the transverse groove; (f) flattened sporophyll showing a small fold above the groove. (pp = proximal portion; dp = distal portion, upturned in life position as in c and d; g = transverse groove; b = transverse bulge; f = transverse fold; j = thickened junction zone; sp = sporangium; w = wing; lp = ligular pit; pl = rod-like placenta, located on the abaxial side but often visible on the adaxial side of the sporophylls in compression, especially when the sporangium is empty. GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 129 Ch) SS : Mt Ot bia Oe N op Gx: @ A © Fic. 6. Comparative morphology of the sporophylls of the Annalepis-type and related genera. (Scale-bar = 1 cm). a—Annalepis zeilleri Fliche (from Grauvogel-Stamm and Duringer, 1983). The arrow shows a scale having lost its lateral wings. b—A. Jatiloba Meng (redrawn from Meng, 1998). c—A. angusta Meng (redrawn from Meng, 1995). d—A. sangzhiensis Meng (redrawn from Meng, 1995). e—A. brevicystis Meng (redrawn from Meng, 1998). f—Annalepis sp. (unpublished). Note the range of variation in shape, size and apex length. The arrows indicate two scales having lost their lateral wings. g—Isoetes ermayinensis Wang (redrawn from Wang, 1991). h—Tomiostrobus radiatus Neuburg (redrawn from Dobruskina, 1985). i—Tomiostrobus belozerovii, T. fusiformis, T. bulbosus, T. convexus, T. gorskyi and T. migayi, from left to right (redrawn from Sadovnikov, 1982). j—Lepacyclotes ellipticus Emmons (redrawn from Fontaine, 1900). k—Cylostrobus sydneyensis Helby and Martin (redrawn from Helby and Martin, 1965). |—Skilliostrobus australis Ash (redrawn from Ash 1979). m—Isoetites daharensis Barale (redrawn from Barale, 1999). n—Isoetites horridus Brown from the Tertiary of North America (redrawn from Hickey, 1977). o—Isoetites choffatii (Saporta) (redrawn from Teixeira, 1948) . 130 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) together in the cones, their proximal portion was perpendicular to the cone axis whereas their distal part was upturned and more or less parallel to this axis (Fig. 5c). The adaxial surface of the proximal portion was occupied, in the middle and on the whole length, by a long, oval and well-delimited spo- rangium. The borders of the proximal portion which project on both sides of the sporangium appear to be membranous, as also noticed by Fliche (1910). The boundary between the median part bearing the sporangium and its mem- branous borders is not always very clear but it is particularly obvious when the imprints are covered with coaly organic matter which is thinner on the borders and therefore less dark than on the other parts of the scales. These membranous borders, called wings or alae according to the authors, likely tended to fall or to be damaged. In some specimens they are partly detached and in others they are completely missing (Fig. 5b), so that it appears clearly that the proximal part is widened distally and that the wings are restricted to the area below this distal widening (see Grauvogel-Stamm and Duringer, 1983; compare figs. 2 and 3 of Pl. 4). The long oval sporangium seems divided by a longitudinal midline which likely corresponds, as in Isoetes (Hall, 1971), to the trace of a rod-like placenta which was located on the abaxial side of the proximal portion of the sporophyll and to which the sporangium wall was attached. The presence of a ligule, which was not recognized when Grauvogel- Stamm and Duringer (1983) described these scales, is clear but this ligule is usually decayed and represented only by its attachment point, which is lo- cated immediately distal to the sporangium. Undescribed new material from the Ladinian (late Middle Triassic) of Ger- many shows that the distal portion of these sporophylls was fleshy. The coal- ified organic matter covering their imprint is thick and the cell imprints are usually quite visible on its brilliant surface. These sporophylls were arranged in cones reaching at least 8-10 cm in diameter. The occurrence, together with these scales and cones, of several rounded furrowed rhizomorphs, probably 4- lobed, each still attached to a stem (10 cm long preserved) which is covered with oval leaf-scars (the leaves are unknown) is informative since it shows that this plant had the same growth habit as Pleuromeia and consisted of an un- branched monopodial axis with a bilaterally symmetrical rooting base and a large, terminal and compact cone (Fig. 7a). The impressive and numerous roots, 3-7 mm wide and more than 30 cm long, which are still attached to the rhizomorphs, may indicate that this plant was robust and strongly anchored. As shown by Grauvogel-Stamm and Duringer (1983), the sporophylls of the cones were monosporangiate, containing either monolete microspores (Ara- trisporites Leschik) or trilete megaspores (Tenellisporites marcinkiewiczae Reinhardt and Fricke). No specimens offer evidence as to whether the cones were bisporangiate or monosporangiate. The ultrastructural features of the monolete microspores are comparable to those of Pleuromeia and Isoetes (Grauvogel-Stamm and Lugardon, in preparation). THE OTHER ANNALEPIS SPECIES.—Five further species of Annalepis have been described from the Anisian (early Middle Triassic) of South China: A. furo- GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 131 \\ wittiga/// \\" iy / Fic. Comparative growth habit of Annalepié, ienetiiee and Isoetes. (Scale-bar = 5 cm). a— peel cies zeilleri Fliche. Note the apical d the basal rounded rhizomorph bearing numerous long roots, more than 30 cm long, arranged in two sets on both sides of the furrow. The leaves are unknown but their scars cover helically the stem. Ladinian of Germany (unpublished). b—lIsoetites daharensis Barale. Note the basal rounded rhizomorph and the stocky stem covered ly were packed in a cone as suggested by the specimens figured by Barale, which are at the origin of this modified reconstruction. Lower Cretaceous of Tunisia. (modified from Barale, 1999). c— et al., 1992). d—Isoetes boryana, having 20-30 lacunate leaves 10—20 cm long (redrawn from Motelay and Vendryés, 1882 ngqiaoensis Meng, A. latiloba Meng (Fig. 6b), A. angusta Meng (Fig. 6c), A. sangzhiensis Meng (Fig. 6d) and A. brevicystis Meng (Fig. 6e). The last three species are shown to have a long, narrow leaf tip emerging abruply from a rounded apex (Meng, 1995, 1996, 1998, 2000). The sporophylls representing A. brevicystis are by far bigger than the others and greatly differ in shape. Meng (1998) moreover described several further sporophylls with a long leaf tip con- taining transverse partitions which he assigned to A. zeilleri but which prob- ably correspond to another species. Indeed none of the specimens correspond- ing to the original material of A. zeilleri described from the Ladinian (late Middle Triassic) of France show such features (Grauvogel-Stamm and Durin- ger, 1983). In all the sporophylls described by Meng, the sporangium is tongue- shaped and is longitudinally divided by a midline. Unfortunately none of the photographic illustrations of the Chinese specimens clearly shows the struc- ture of these sporophylls, such as it is shown in the rough reconstructions figured by Meng (1995, 1998, 2000). Furthermore, an assemblage of numerous undescribed sporophylls from the Ladinian (late Middle Triassic) of Germany which is here provisionally as- +32 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) signed to Annalepis sp. (Fig. 6f), represents a further new species which is still under study (Grauvogel-Stamm and Kelber, in preparation). Indeed, these sporophylls which originally have been attributed to A. zeilleri by Kelber and Hansch (1995) contain features which distinguish them from this species. They show a great variability in shape, size and apex length, which probably de- pends on their place within the cone. Transverse partitions have never been observed in any of the long apex of these sporophylls. However, they seem to have lateral membranous borders as in A. zeilleri (Fig. 8a,b) and the tongue- shaped sporangium shows a longitudinal midline as in the other Annalepis- like sporophylls. In spite of a roughly similar organization, the sporophylls of Pleuromeia and Annalepis differ in several respects. Their general outline and the shape of their sporangium are quite different. The wings bordering the sporophyll prox- imal part and the longitudinal midline running along the sporangium in An- nalepis are missing in Pleuromeia. In contrast to the sporophylls of Annalepis that are clearly differentiated into a proximal and a distal portions which were very likely more or less perpendicular to each other in lifetime, those of Pleu- romeia do not show such a morphological partition and are approximately on the same plane from the base to the apex. Moreover, in Pleuromeia the micro- spores and megaspores are trilete whereas in Annalepis only the megaspores are trilete while the microspores are monolete (Grauvogel-Stamm and Durin- ger, 1983), as those of Isoetes. THE ALLIED GENERA: FROM ANNALEPIS TO ISOETITES.—While it is easy to recognize a sporophyll as belonging to the genus Pleuromeia, there are difficulties in identifying those of the Annalepis-type. Most of the latter have been long mis- understood, being first regarded as araucarian seed scales and therefore often assigned to the genera Araucarites Pres] or Pseudoaraucarites Vladimirovitch (Dobruskina, 1985). Now, eventhough they are regarded as lycopsids, these sporophylls are assigned to different genera as Tomiostrobus Neuburg, Skil- liostrobus Ash, Cylostrobus Helby and Martin and Lepacyclotes Emmons, and classified, together with Annalepis, as satellite-taxa in the Pleuromeiaceae (Thomas and Brack-Hanes, 1984). However their taxonomic position is greatly in flux. Dobruskina (1985) stressed the similarity between Annalepis and To- miostrobus whereas Retallack (1997) synonymized Annalepis with Lepacyclo- tes. Moreover Sadovnikov (1982) and Retallack (1997) synonymized Skillios- trobus with Tomiostrobus. Nevertheless, since the genus Annalepis, particu- larly the type-species A. zeilleri, is the best known, it is here regarded as the type of this kind of sporophyll. Lepacyclotes was described earlier by Emmons (1856) but the precise structure and the in situ spores have never been studied. Other taxa for which the structure and the shape of the sporophylls are un- known include Lycostrobus scotti Nathorst and Austrostrobus ornatum Mor- belli and Petriella. As for Lycostrobus chinleana Daugherty from the Triassic of southwestern USA, it has proven to be an equisetalean reproductive organ (Grauvogel-Stamm and Ash, 1999) and not a lycopsid, as previously believed, or an arborescent quillwort, as recently asserted by Retallack (1997). GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 133 f a s @)— \ o “-\-. Oe @ wie Fic. 8. Comparative morphology and terminology of the — of Annalepis, Lepidoden- drids and Isoetes in adaxial views. (scale-bar = 5mm, except in d,e = m). a—Annalepis zeilleri Fliche (pp, proximal portion; dp, distal portion; lp, lieular pit; sp, ee w, wing). Note the clearly delimited wings (stippled) of the proximal portion (modified from Grauvogel-Stamm and Duringer, 1983). b, c—Annalepis sp. (undescribed): two sporophylls within the range of variation. The wings (stippled) are clearly delimited in b, while they are not in c. d—Isoetes coromandelina Lin. F ils: a scale-leaf with an aborted sporangium (asp) (redrawn from Srivastava and Wagai, 1996, pasa from Bochenski, 1936). Sab Sanger ps jenneyi (White) Abbott. Sporophyll showing a proximal pedicel below the sporangium and an apically directed lamina (la). Due to its strong distal widening, the pedicel has been ea T-shaped by Abbott (modified from Abbott, 1963). i—Lepidocarpopsis lanceolatus (Lindley and Hutton) Abbott. The pedicel clearly bears well-de- stippled), sporangium (sp), ligula (li), and the terminal subula (s) 3,4 cm long (modified from Kubitzki and Borchert, 1964; terminology after Hickey 1985) Since the precise structure and shape of the sporophylls are essential for identifying and comparing these lycopsids, only the genera for which these features are known with considerable accuracy and which resemble the An- nalepis-type sporophylls are listed in Table 1 and shown in Fig. 6, Some Is- oetites are also listed but their inclusion raises some questions on the exact interpretation of this fossil genus, as commented below. 134 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) TABLE 1. The crap -like sporophylls. Data taken from the following sources: 'Fliche (1910); *Meng (1995); 3Meng (1998); Meng (2000); ‘Dobruskina (1985); “Sadovnikov (1982); 7Wang (1991); ‘Emmons (1856); °Ash (1979); tletby and Martin (1965); ''Teixeira (1948); '?Barale (1999); Brown (1958), Hickey (1977) Taxon Age Occurrence Figure si nrtcande (type-species) Middle Triassic NE France; Germany Fig. 6a A. ine ensis* Middle Triassic Southern China A. lati Middle Triassic Southern China Fig. 6b A. angusta* Middle Triassic Southern China Fig. 6c A. sangzhiensis* Middle Triassic Southern China Fig. 6d A. brevicystis? Middle Triassic Southern China Fig. 6e i vat ohana Middle Triassic Germany Fig. 6f oetes erma Middle Triassic North China Fig. 6g Tmicarrabee le (type-species) Lower Triassic South Siberia, Russia Fig. 6h . belozerovii® ower Triassic East Siberia, Russia Fig. 61 T. fusiformi Lower Triassic East Siberia, Russia Fig. 61 bulbosus® Lower Triassic East Siberia, Russia Fig. 61 T. gorskyii® Lower Triassic East Siberia, Russia Fig. 61 T. pees Lower Triassic East Siberia, Russia Fig. 61 T. convexus® Lower Triassic East Siberia, Russia Fig. 61 Lepeveoe Sera (type-species), Upper Triassic North Carolina, USA Fig. 6j yn. L. elliptic ae ior fer neyensis"® (type-species) Lower Triassic Australia Fig. 6k Skilliostrobus emeieg (type-species) ower Triassic Australia Fig. 61 Isoetites daharens Lower Cretaceous Southern Tunisia Fig. 6m Isoetites cevichas™ Tertiary North America Fig. 6n Isoetites choffati'' Lower Cretaceous Portugal Fig. 60 The sporophylls characterizing these taxa share a remarkable similarity with Annalepis. They have a comparable shape and structure (Fig. 6a-o). Most of them (Annalepis, Tomiostrobus, Isoetes ermayinensis, Skilliostrobus, Cylostro- bus) contain microspores assignable to Aratrisporites (Balme, 1995). In all, the sporangium is elongate-oval and shows a longitudinal midline which has been variously interpreted. In Annalepis zeilleri, Grauvogel-Stamm and Duringer (1983) suggested that it might represent the trace of the vascular bundle (Fig. 6a). In Isoetites daharensis (Fig. 6m), this line, which is obvious too and quite resembles that of Annalepis, has been interpreted as being the internal bound- ary of the velum (Barale, 1999). In Isoetes ermayinensis (Fig. 6g) it has been regarded as a strong keel or a vascular bundle (Wang, 1991). In Tomiostrobus (Figs.6h,i), Sadovnikov (1982) interpreted it as a keel whereas Dobruskina (1985) regarded it as the vascular bundle. Fontaine (1900) also regarded it as a keel in Lepacyclotes ellipticus (Fig. 6j). The fact that this line appears in variably the same in the different scales indicates that it is a eicatiette feature of this kind of sporophyll and that it has the same significance in all. A comparison with Isoetes shows that it does not represent the velum which is a thin membrane covering partially or entirely the sporangium and which never has a longitudinal slit-like free edge in the middle. In contrast, the lon- gitudinal midline more likely indicates that, as in Isoetes (Hall, 1971), a rod- GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 135 like placenta including the vascular bundle existed on the abaxial side of these sporophylls, to which the sporangium wall was attached, and that it became apparent on the adaxial side of the compressions particularly when the spo- rangium was empty. According to J. Hickey (pers. com), in Isoetes this placenta is represented by a pad of vascularized tissue which often intrudes into the sporangium and runs lengthwise. Hall (1971) moreover noticed that in dried specimens of I. tenuifolia and I. pitotii, the rod-like shaped placenta shrinks and leaves an especially deep and profound groove. In some specimens of Tomiostrobus (Figs. 6h,i) and Skilliostrobus (Fig. 61) which one of us (L. G.-S.) observed, and according to the illustrations of Isoetes ermayinensis (Fig. 6g) and Cylostrobus sydneyensis (Fig. 6k), most of the scales seem to have a proximal portion with lateral wings, as in A. zeilleri (Fig. 6a). These lateral wings are clearly shown in the transverse section of the sporo- phyll of Tomiostrobus figured by Sadovnikov (1982, Fig. 2, at right), though not noticed by him. They are comparable to those shown in transverse sections of Isoetes sporophylls (Pitot, 1959; Hall, 1971). In some specimens the wings seem to be missing or partly destroyed, as in Skilliostrobus (see Ash, 1979, Figs. 7 A—D) or I. ermayinensis (see Wang, 1991, Pl. 1 figs. 9-11, 13, 14). How- ever the question of the wings would require further investigations since their presence has not been noticed in most of the sporophylls listed in Table 1. The long, oval and contiguous depressions, often visible on both sides of the sporangium in most of the scales, such as Annalepis sp., A. zeilleri, A. latiloba and Lepacyclotes circularis, correspond to the imprint of the sporangia of the underlying and overlying sporophylls when they were packed together in the cone. The wide transverse groove visible at the junction of the proximal and distal, thick portions of the scales of L. circularis, as seen in adaxial view and as figured by Bock (1969, Figs. 93-94), is quite similar to that observed in A. zeilleri (Grauvogel-Stamm and Duringer, 1983, Pl. 4 figs. 2,3); this strongly suggests that the proximal and distal portions were originally more or less perpendicular to each other in Lepacyclotes too. The sporophylls of Tomios- trobus which one of us (L.G.-S.) observed, also show that the junction of the proximal and distal portions was more or less right-angled. Likewise, in An- nalepis sp., the distal portion also seems to have been upturned. Also Meng (1998) noticed the presence of a transverse ridge in A. latiloba which suggests that the sporophyll distal portion was more or less perpendicular to the prox- imal one. However various interpretations and/or misinterpretations have occurred because these sporophylls often have a more or less blunt, seemingly incom- plete apex and/or because the scale tip has been considered as being the ligule, making the interpretation of this area confusing. Moreover, some of these spo- rophylls have been regarded as having lost their leafy distal portion. The spo- rophylls of Isoetes ermayinensis Wang (Fig. 6g) are so similar to those of A. zeilleri (Fig. 6a) that the inclusion of that taxon in Isoetes is surprising. In fact, Wang (1991) interpreted the numerous, associated elongate leaf imprints as their detached apical parts. These imprints contain transverse partitions that 136 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) he supposed to be collapsed air chambers, similar to those of Isoetes. However no specimens show them in organic connection. A similar leaf imprint found associated with sporophylls of A. brevicystis in the Anisian of South China has also been regarded as their tip (Meng, 1996, Pl. 2 Fig. 14). Sterile parti- tioned leaves have never been found associated with A. zeilleri from the Tri- assic of eastern France (Grauvogel-Stamm and Duringer, 1983). Retallack (1997) noticed that Wang’s sporophylls of I. ermayinensis are “unlikely to have borne the wider undulose leaf fragments on their tips as in Isoetes’” and re- assigned I. ermayinensis to the genus Lepacyclotes. Likewise, the reassignment of Lepacyclotes circularis and L. ellipticus to the genus Isoetites, as proposed by Brown (1958), does not seem to be justified and has been questioned by Chaloner (1967) and Bock (1969) since these scales have a short and well delimited triangular apex and are devoid of any traces of a long sterile leafy portion comparable to that of Isoetes. Skog and Hill (1992) also questioned this reassignment and noticed the resemblance between Lepacyclotes circu- laris and Annalepis. Among the fossils attributed to the genus Isoetites, there are also isolated scales with a small pointed apical tip resembling the scales of Annalepis, which have been supposed to be the fertile base of Isoetes leaves. It is the case, especially, of Isoetites horridus (Fig. 6n) from the Tertiary of North America (Brown 1958, Figs. 1,3; 1962, Pl. 9 Fig. 1) and I. choffati Saporta (Fig. 60) from the Cretaceous of Portugal (Teixeira, 1948, Pl. XXVI figs. 7-8). However the small apical tip is not preserved in some specimens, as those of I. horridus from the early Tertiary of Dakota figured by Hickey (1977, Pl. 1 figs. 6,7) and described as being dumb-bell shaped by Collinson (1991). Moreover, according to the illustration of Collinson (1991, Fig. 7.3h), which shows a circlet of such supposed fertile leaf bases, one can see that Hickey (1977) rep- resented these sporophylls upside down and that the narrower extremity cor- responds to the base while the broader one corresponds to the distal extremity to which the long leafy part was supposed to be attached. It is noteworthy that this circlet closely resembles those consisting of Annalepis sporophylls. Since the leaf tip is often missing in the Isoetites remains, Collinson (1991) suggested that this tip was deciduous in the fossil genus. In Isoetes, the long leafy part is usually not deciduous, except in some terrestrial species where the distal portion of the leaves erodes away leaving only a leaf base, typically with a sporangium (J. Hickey, pers. com.). However, the fossils attributed to Isoetites call for several remarks. An over- view of these fossils shows that the precise shape and structure of the fertile parts are usually poorly described and illustrated and that their relationship with the sterile leaves is not clear. For most taxa, the fertile and sterile parts are not shown in organic connection and several rhizomorphs and stems seem to contain only sterile leaves, e.g., Isoetites phyllophila (Skog et al., 1992), I. horridus (Brown, 1962; Hickey, 1977), I. indicus (Bose and Banerji, 1984), I. serratus (Brown, 1939), I. rolandii (Ash and Pigg, 1991) and Cylomeia (White, 1981) which is now reassigned to Isoetes beestonii (Retallack, 1997). It is sur- prising that none of them shows any trace of a sporangium at the base of the leaves. Likewise, the late Cretaceous and Tertiary specimens assigned to Is- GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 137 oetites by Pigg (2001) do not show clearly if there were sporangia at the base of the leaves. Long sterile leaves, without partitions, still attached to an axis have also been found associated with the sporophylls of Annalepis sp. in the Triassic of Germany and have been assigned to Isoetites sp. (Kelber and Hansch, 1995). In Isoetes, a great number of leaves are usually fertile (Jermy, 1990) and when once the plant becomes fertile, all subsequently produced leaves will either be fertile or at least initiate a sporangium (J. Hickey, pers. com.). During the ontogeny of the Isoetes leaves, the sporangium develops very early (Fig. 8e), is already discernible in the leaf primordia (Bhambie, 1963; Kubitzki and Borchert, 1964), and even the vegetative leaves have an aborted sporangium at the base (Pitot, 1959; Rauh and Falk, 1959; Kubitzki and Borch- ert, 1964). Thus the real nature of many Isoetites leaves remains quite ambig- uous. Moreover, some of the described Isoetites species do not represent ly- copsids. According to Skog and Hill (1992), Isoetites gramineus (Ward) Bock more likely is an osmundaceous fern and J. bulbosus Drinnan and Chambers more probably corresponds to “pentoxylalean fertile short shoots”’. Likewise, the interpretations about the scales of Isoetites having a small tip and resembling the Annalepis sporophylls are still questionable. Do they rep- resent fertile leaf bases which have lost their leafy extremity, and is the small tip visible in some of them actually a ligule? In Isoetes, the ligule is a delicate ephemeral parenchymatous flap which is mostly deciduous and tends to dis- appear at maturity (Pitot, 1959; Bhambie, 1963; Goswami, 1976; Kott and Brit- ton, 1985). Therefore, it seems unlikely that the ligule would persist if the long leafy extremity of the sporophyll was destroyed, as for example in I. horridus where a specimen representing a circlet of sporophylls still attached to an axis has been interpreted by Brown (1962, Pl. 9 Fig. 1,3) as being a “corm with sporangia tipped by ligules’”’. The Isoetites scales with a small apical tip seem more likely to represent entire sporophylls rather than the basal part of fertile leaves, and their small apical tip more likely represents their actual extremity and not the ligule. Moreover, the fact that these Isoetites sporophylls with short pointed or blunt apices are often found associated with long, apparently ster- ile, leaves and are still closely packed (see Brown, 1958 Fig. 3; Hickey, 1977; Collinson, 1991, Fig. 7h; Barale, 1999, Fig. 5)—as in the circlets of sporophylls of I. horridus, A. zeilleri, Annalepis sp. and Lepacyclotes ellipticus - suggests that they represented distinct organs, occupied a place distinct from that with sterile leaves, and became detached or abcised from an axis. Isoetites dahar- ensis (Fig. 7b) is particularly interesting in this respect since it has a rounded rhizomorph with roots and a short stem bearing distal scale-like sporophylls and elongate sterile leaves (Barale, 1999). Such a plant does not conform to Isoetes (Fig. 7d) in which the sterile and fertile leaves are morphologically similar (Jermy, 1990) but more closely resembles Annalepis or Pleuromeia in which the fertile and sterile leaves differ and are located in distinct places on the plant. Isoetites daharensis (Fig. 6m) is all the more interesting because its sporophylls are of the Annalepis-type, as are those of I. horridus (Fig. 6n). However, contrary to the latter, in J. daharensis the leaves are not lacunate. In most of these fossil lycopsids, the length of the corm plus stem is unknown. 138 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) In I. daharensis (Fig. 7b), the rounded corm + stem are 11 cm tall (Barale, 1999) as in Annalepis zeilleri (Fig. 7a). In Isoetites phyllophila (Fig. 7c), the rhizomorph is elongate, conical, and 6 cm high (Skog et al., 1992). However, as noticed by Skog and Hill (1992), there are no broadly ranging studies on the shape of the stems at various stages in development of extant species of Isoetes stripped of their appendages and de-corticated, so it is difficult to com- pare fossil and living plants. In Stylites for example, the stem reaches 20 cm long (Rauh and Falk, 1959). Likewise, there are no broadly ranging studies on the distribution of the fertile and sterile leaves in extant Isoetes. Clearly, most of the Isoetites specimens need further investigation and better illustrations, all the more since this genus is regarded as representing primitive Isoetes and is included in the subgenus Euphyllum, together with three extant species of Isoetes (Hickey, 1986, 1990; Taylor and Hickey, 1992). Nevertheless, this comparative analysis of the sporophylls of Annalepis and related taxa (Tomiostrobus, Lepacyclotes, Skilliostrobus, some Isoetites) dem- onstrates that all these sporophylls have a similar basal unit which is narrow at the base, widens distally and extends by a short, roughly triangular distal portion ending with a more or less long pointed apical tip. In all there is a long oval adaxial sporangium and a ligule immediately distal to it. In most of them moreover, the proximal portion appears to bear lateral wings which are restricted to the area below the widened portion of the leaf. The sporophyll of Isoetites horridus figured by Hickey (1977, Pl. 1 figs. 6,7) is especially in- teresting in this respect because it shows particularly well the distal widening (Fig. 6n), suggesting that initially it also had lateral wings restricted to the proximal part. RESEMBLANCES WITH ISOETES AND LEPIDODENDRID SPOROPHYLLS.—Besides the ul- trastructural similarities of the spores of Annalepis and Isoetes (Lugardon et al. 2000b; Lugardon and Grauvogel-Stamm, in prep.), there are still some other morphological resemblances between both genera which deserve to be noticed. Indeed, some Annalepis sporophylls (Fig. 8c) closely resemble the devel- oping leaf primordia of Isoetes (Fig. 8e), which have a short pointed apex devoid of air chambers and an almost fully developed ligule, proximal to this apex (Kubitzki and Borchert, 1964). Such a resemblance shows how an An- nalepis sporophyll may have developed into an Isoetes sporophyll. It would be instructive to know the shape and growth sequence of the wings in young Isoetes leaves but this development has not been studied (J. Hickey, personal information, September 1999). Likewise it is noteworthy that the Annalepis sporophylls much resemble the scale leaves which occur in many Isoetes species and differ greatly from the typical Isoetes long leafy sporophylls. These scale leaves which are mostly unknown to Isoetes non-specialists are present at the growing apex and at the base of the corm, but are also said to occur below the spirally arranged leafy sporophylls, sometimes intermingled with them, and to form two or three whorls in old corms (Duthie, 1929; Stolze and Hickey, 1983; Hickey, 1986; Srivastava and Wagai, 1996). Thus for example, within the range of variation GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 139 of Annalepis sp. there are sporophylls (Fig. 8c) which are very similar to the scale leaves of Isoetes montezumae A.A. Eaton (Stolze and Hickey, 1983, Fig. 10c) or I. coromandelina Lin. fils (Srivastava and Wagai, 1996, Pl. 4 Fig. 1) (Fig. 8d). Likewise, the sporangium bearing scales with a small pointed apical tip (Fig. 60) which occur in Isoetites choffati (Teixeira, 1948) among the long leafy Isoetes-like sporophylls, and which resemble the sporophylls of Anna- lepis, might also represent such scale leaves. According to Hickey (pers. com.), the scale leaves of Isoetes are environmentally induced and occur in distinct zones (alternating with typical fertile leaves) as a result of seasonality. Also according to Hickey (1986), they have a protective function for the corm and represent leaf primordia in which the development has become arrested. How- ever, their resemblance with Annalepis sporophylls suggests that they also might have a phylogenetical significance and might represent vestigial rem- nants of ancestral sporophylls, indicating that the genus Annalepis is an an- cestor of Isoetes. The sporophylls which have been described recently from the Triassic of South China and which have been assigned to several new Annalepis species, e.g. A. angusta (Fig. 6c), A. sangzhiensis (Fig. 6d), A.brevicystis (Fig. 6e) and A. zeilleri (sensu Meng 1998) are also particularly interesting in the compari- son between Annalepis and Isoetes. Indeed, these sporophylls which are char- acterized by a long, narrow (spur-like) outgrowth, arising distal to the ligular pit, running across the short, wide, triangular to rounded distal portion and extending far beyond it, also closely resemble the scale leaves which occur in Isoetes and which are characterized by a spine-like apex. In I. mahadevensis for instance, the scale apices are said to appear as spur-like outgrowths and to emerge abruptly from an almost rounded region (Srivastava and Wagai, 1996). In Annalepis zeilleri (sensu Meng 1998) moreover, this apical outgrowth is described as containing transverse partitions. Meng (1998) interpreted it as the leaf tip of the sporophylls. According to Hickey (pers. com.), it is comparable to the ‘“‘subula” of the Isoetes leaves, i.e. their distal awl-like shaped outgrowth. Meng (1998, 2000) also noticed the similarity of the Annalepis sporophylls with Isoetes leaves, and suggested that A. brevicystis, especially, might be an- cestral to the living genus. The new Chinese Annalepis species are moreover particularly interesting because they allow to understand some unexplained features of Tomiostrobus radiatus (Fig. 6h). Indeed, the two parallel lines shown to arise distal to the ligule and to extend apparently out from the fragmentary apex of the sporo- phylls in the rough drawings of Dobruskina (1985, Fig. 1b-e, g, i, n), more likely represent partially preserved apical spine-like outgrowths comparable to those described by Meng (1995, 1998, 2000) in several Chinese Annalepis species and compared by Hickey (pers. com.) to the subula of the Isoetes leaves. In a sporophyll of T. radiatus figured by Yaroshenko (1988, Fig. 11), the long apex appears moreover to contain transverse ridges in the middle, though not noticed by the author, which might represent collapsed air cham- bers. However, none of the fragmentary sporophylls attributed by Sadovnikov (1982) to several distinct Tomiostrobus species (Fig. 61) seems to contain such 140 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) partitions. Likewise, in none of the specimens of Tomiostrobus from the Lower Triassic of East and North Siberia (Russia) which one of us (L. G.-S.) could observe, the apex shows such features. Comparable unexplained spur-like out- growths seem to exist in Annalepis sp. and A. zeilleri but they are badly pre- served and have not yet been described (Grauvogel-Stamm and Lugardon, in preparation; Grauvogel-Stamm and Kelber, in preparation). However, in spite of these resemblances there are some differences in the shape of the fertile area and lateral wings between the Annalepis sporophylls and the typical Isoetes long fertile leaves. In Isoetes the proximal portion is wide at the base, narrows distally and extends upwards in a long leafy extrem- ity (Fig. 8j), whereas in Annalepis and allied genera the proximal portion is narrow at the base, widens distally and extends by a short more or less tri- angular portion ending in a small pointed apical tip (Fig. 8a, b). In Annalepis moreover the lateral wings are restricted to the area below the distal widening whereas in Isoetes the winged borders, now commonly termed alae, extend upwards over a variable length (Kott and Britton, 1985). It is worthy of note that in most of the extant Isoetes species, these alae are also restricted to the proximal portion of the leaf (Hickey, 1986; Taylor and Hickey, 1992) but, in contrast to Annalepis, they are not limited by a distal widening of that prox- imal portion. In fact the shape of the proximal part and lateral wings of Annalepis spo- rophylls more resembles that of the basal unit of the sporophylls of the Car- boniferous lepidodendrids Lepidostrobus Brongniart, Lepidostrobopsis Abbott and Lepidocarpopsis Abbott, as already noticed by Grauvogel-Stamm and Dur- inger (1983, Figs. 3a, b). These lepidodendrid sporophylls that have usually a long, wide, upturned leafy distal portion termed lamina in most descriptions, also have a basal unit consisting of a proximal portion bearing an adaxial sporangium and sometimes lateral wings, which is termed pedicel (Fig. 8f-i). It is noteworthy that the proximal portion of these sporophylls widens distally as in Annalepis or Isoetites horridus (Hickey, 1977) and that the wings, when present, are restricted to the area below the distal widening of the proximal portion. The sporophyll of Lepidophyllum jenneyi White (1899, Pl. 59 Fig. 2) which shows very clearly the distally widened proximal portion was regarded as having lost its lateral wings while it was interpreted by Abbott (1963) as having a T-shaped pedicel and was transferred by this author to the genus Lepidostrobus (Fig. 8h). In Lepidocarpopsis lanceolatus (Lindley and Hutton) Abbott on the contrary, the pedicel is clearly alate, exceeding the width of the sporangium (Fig. 8i). The sporophylls of Cyclostigma kiltorkense Haughton (Fig. 8f) from the Upper Devonian of Ireland (Chaloner, 1968) and Lepidostro- bus bohdarowiczii Bochenski (1936) from the Carboniferous of Poland (Fig. 8g) also have such a T-shaped pedicel and a long distal lamina, but their ped- icel seems to be devoid of lateral wings. However, according to Taylor and Hickey (1992), the alation which characterizes the sporophylls of the arbores- cent lepidodendrids and those of the isoetaleans is a plesiomorphy and there- fore is not phylogenetically informative. GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 141 THE TERMINOLOGY IN THE LYCOPSID REPRODUCTIVE ORGANS Comparisons of the Annalepis sporophylls with those of the arborescent lycopsids and those of Jsoetes underscore a problem of terminology. Indeed, as there is no standardization for the different groups of lycopsids, various terminologies occurred which make rather complex and confusing the com- parative studies. As regards to the lepidodendrids, the proximal and distal portions of the sporophylls are respectively termed ‘‘pedicel” and “lamina” by most of the authors (Abbott, 1963; Chaloner, 1967; Emberger, 1968; Bateman et al. 1992; Stewart and Rothwell, 1993). Emberger (1968, p.179) noticed that the proximal portion is equivalent to a petiole bearing a sporangium. In con- trast, Taylor and Taylor (1993, p.270) defined the lepidodendrid sporophylls (Lepidocarpon Scott) as consisting of two lateral laminae and a distal exten- sion. The similarities between the Annalepis sporophylls and those of the lep- idodendrids led Grauvogel-Stamm and Duringer (1983) to use for Annalepis the terminology applied to the latter by most authors. However the resem- blances between Annalepis and Isoetes emphasized for the first time in this study, required to change the terminology employed previously, as it has been done above. In contrast to the lepidodendrids, in Isoetes the proximal portion of the sporophyll is termed “vagina” while its distal portion is called “subula”. The current use in the terminology of Isoetes of these Latin words has been intro- duced by Hickey (1985) whereas, previously, they were employed only in the Latin diagnosis (Braun, 1868; Duthie, 1929; Wanntorp, 1970). However the use of these Latin words complements those which were already in use for Isoetes, such as “ala”, “ligula”, “‘velum” and “labium’”. As for the lateral wings (or alae) which are present in the proximal portion of the Isoetes leaves, they are regarded as the restricted remnants of the lamina (Hickey, pers. com.). In fact the meaning of these Latin words describes the features of the different por- tions of the leafy sporophylls. Thus, the portion containing the alae is called “vagina” which means “sheath”, because the wings are slightly enveloping, as shown for example in the drawings of Wanntorp (1970). On the other hand, as the distal outgrowth of the Isoetes sporophylls is awl-like shaped, it has been called “subula” which means “awl”. The terminology such as applied to Isoetes seems to be applicable to some of the Chinese species of Annalepis (Meng 1995, 1998, 2000), as noticed by Hickey (pers. com. 1999) who recognized a subula in some of them, but it does not seem applicable to all the Annalepis-like sporophylls, particularly those in which there is no subula. However, the leaf primordium of Isoetes also seems to be devoid of a subula. Thus, only the shape and growth sequence of the wings and the subula in young Isoetes leaves would allow to establish the exact homologies with Annalepis but this development has not yet been stud- ied (Hickey, pers. com., September 1999). CONCLUSIONS The comparative analysis of the Triassic lycopsids Pleuromeia and Anna- lepis using new macromorphological and ultrastructural data provides signif- 142 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) icant support for their close relationship with each other and with the living genus Isoetes and more ancient lycopsids, some of which extend back to the Devonian. A close affinity between Pleuromeia and Annalepis appears unquestionable. Preliminary observations of some rather well preserved specimens of Anna- lepis indicate that they have furrowed, bilaterally symmetrical rhizomorphs as in Pleuromeia. The sporophylls, with their adaxial sporangium and distal lig- ule, are comparable in both genera, except that the sporophylls of Annalepis are wedge-shaped and have well-marked membranous wings which are seem- ingly missing in Pleuromeia. Moreover, although they are monolete, the mi- crospores of Annalepis clearly show the same ultrastructural features as the trilete microspores of Pleuromeia, i.e. a thin exospore with a thinner, non- folded apertural area combined with laminated zones, and a thicker spongy outer wall largely dissociated from the exospore and divided into two pro- truding lips above the aperture. Most of the features shared by Pleuromeia and Annalepis also appear ve similar to those of the living Isoetes. The comparative study of the rhizo- morphs of Pleuromeia and Isoetes demonstrates that these organs have many points of structural and developmental correspondence. The ultrastructural features of the exospore and the spongy outer wall in the Pleuromeia and Annalepis microspores are remarkably similar and homologous with the walls of the Isoetes microspores (the latter having in addition a third, superficial wall, the perispore). This similarity between fossil and recent microspores is all the more striking because the fine structure of the Isoetes microspores dif- fers markedly from that of all the other extant pteridophyte spores, including those of Selaginella. Indeed, despite some ultrastructural resemblances reflect- ing the relationships between Isoetes and Selaginella, the microspores of the latter show a number of unambiguous differences which strongly suggest that Isoetes is much more closely related to Pleuromeia and Annalepis than it is to Selaginella. The resemblances in morphology of the sporophylls also de- serve to be noted. Snigirevskaya (1989) stressed the similarity between the sporophylls of Pleuromeia and the basal fertile region of Isoetes leaves. How- ever, the clearly winged sporophylls of Annalepis show a stronger resemblance to Isoetes. In fact, the most important difference lies in the growth habit of the Triassic genera and that of Isoetes. Like some lycopsids from the Carboniferous (Chaloneria cormosa) and the Devonian (Clevelandodendron ohioensis), the Triassic lycopsids Pleuromeia and Annalepis have an elongated stem covered with entirely sterile leaves and a terminal, well-defined cone with short spo- rophylls, whereas the living genus consists of a short corm with long lacunate leaves, most of which bear a basal sporangium. However these differences do not seem to be very significant since they are rather common in the lycopsids where they are regarded as simple variations among related genera. The fact that in Isoetes many or all the leaves are fertile can be explained either by the loss of sterile leaves (Pigg, 2001) or, more likely, by the evolutionary reduction and transfer of the reproductive function to the vegetative leaves, as in the Gymnosperms and the Angiosperms (Nozeran, 1955). GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 143 All the similarities shown by both Triassic genera and Isoetes surely indicate close connections between them. Nevertheless, at first glance, they are not clearly indicative of a sister-group or an ancestor-descendant relationship and do not reveal the exact role played by Pleuromeia and Annalepis in the evo- lutionary history of Isoetes. However, there are some further noteworthy re- semblances between Annalepis and Isoetes which strongly suggest that An- nalepis is closer to the extant genus than Pleuromeia and that it is ancestral to Isoetes. Recently Meng (1998, 2000) also made this assumption. Besides their winged sporophylls/leaves noted above, there is, indeed, a striking sim- ilarity between the Annalepis sporophylls and the leaf primordia and scale leaves of the extant genus. Especially the scale-leaves which occur in many Isoetes species, and appreciably resemble Annalepis sporophylls, might be vestigial structures and might represent remnants of ancestral sporophylls, which might indicate that originally Isoetes had distinct sterile and fertile leaves as in Pleuromeia and Annalepis. The microspores of Annalepis are monolete as in Isoetes, whereas they are trilete in Pleuromeia. Moreover, the similarities between Annalepis and some Cretaceous and Tertiary fossils attri- buted to the genus Isoetites give further support to an ancestor-descendant relationship between Annalepis and Isoetes. The newly described species Is- oetites daharensis from the Lower Cretaceous of Tunisia (Barale 1999), which is the best known among these plants, shows quite clearly that its sporophylls are remarkably similar to those of Annalepis and differ morphologically from the sterile leaves. Furthermore, its rhizomorph is cormose as in the Triassic genus, although not bilaterally symmetrical. As discussed above, Isoetites hor- ridus from the Tertiary of North America (Brown 1939, 1958; Hickey 1977; Collinson 1991) was probably organized similarly, bearing its Annalepis-type sporophylls and sterile lacunate leaves in distinct areas, unlike Isoetes. These similarities strongly suggest that Isoetites and Annalepis are closely related and that Isoetites may have evolved from Annalepis. Thus the Triassic genus Annalepis seems to have played a more direct role in the evolutionary history of Isoetes than Pleuromeia. As regards to Isoetites daharensis and I. horridus, they may represent a group of lycopsids intermediate between Annalepis and Isoetes, and may indicate which structural changes occurred during this evo- lution. Thus the evolutionary series ‘Annalepis—Isoetites—Isoetes’ appears very likely. Unfortunately, most of the Isoetites species are poorly known. They are usually identified on the sole basis of their lacunate leaves whereas the most important features, such as the precise shape and structure of the fertile parts and the microspore ultrastructure, remain unknown. The fine structure of Minerisporites mirabilis, the megaspore of I. horridus (Collinson, 1991), shows rather clearly the features of the isoetalean megaspores, but it does not provide any more precise informations on the affinities of Isoetites. Further and more thorough investigations of all the Isoetites material would be essen- tial for definitely establishing the evolutionary series from the Triassic lycop- sids to Isoetes. Besides, the fact that Annalepis and Pleuromeia are contem- poraneous suggests that an ancestor-descendant relationship cannot exist be- tween them. Therefore it is inferred that Pleuromeia either became extinct, or 144 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) originated within the nonalate species of Isoetes. However the latter hypoth- esis seems unlikely in view of the apparent homogeneity of the living genus which strongly suggests a monophyletic origin for it. Some of the typical features of Pleuromeia and Annalepis are also present in more ancient lycopsids, providing some evidence of their ancestral affini- ties. The bilateral symmetry of the rhizomorph is known from the Devonian (Pigg, 2001). The sporophylls of the Carboniferous lepidodendrids, especially those with lateral wings, and those of more ancient lycopsids such as Cyclos- tigma kiltorkense from the Devonian of Ireland (Chaloner, 1968) are compa- rable in some respects to Annalepis. The special ultrastructural features of the microspores shared by the Triassic Pleuromeia and Annalepis, as well as by the extant Isoetes, also occur in Polysporia (Lugardon et al., 2000), Chaloneria, and other Carboniferous lycopsids. Moreover, it is of note that the trilete mi- crospores of Bisporangiostrobus harrisii Chitaley and McGregor (1988) from the Upper Devonian of Pennsylvania contain proximal papillae as do the mi- crospores of Chaloneria/Polysporia, Annalepis, Pleuromeia and Isoetes. This is all the more interesting since Cyclostigma kiltorkense, which has cones sim- ilar to those of B. harrisii (Stewart and Rothwell, 1993), also has a bilobed, bilaterally symmetrical rhizomorph as in Pleuromeia, Annalepis and Isoetes. Moreover several Frasnian-Famennian (Late Devonian) dispersed spores have ultrastructural characteristics comparable to those of the microspores of the Triassic genera and the extant genus (unpublished observations, B.L.). Thus it appears that the main features characterizing the rhizomorph and the micro- spores of living Isoetes appeared very long ago and remained outstandingly unchanged from the Devonian to the present. It is likely that these features characterize a particular group among the rhizomorphic lycopsids and that this group probably has a very remote origin. However, this group cannot be defined at the present time, notably because there are no accurate data on the microspore ultrastructure of most of the fossil lycopsids, mainly the arbores- cent lepidodendraleans. Methodical fine studies of in situ microspores of these lycopsids would surely supply significant information on the origin, relation- ships and delimitation of the different groups of isoetalean lycopsids. ACKNOWLEDGEMENTS. We thank W. Carl Taylor for inviting one of us (L.G.-S) to coorganize the symposium on “Evo- lution and diversification of the Lycopods” in the XVI International Botanical Congress held in St. Louis, Missouri, 1-7 August 1999, and Jim Hickey, Kathleen Pigg and an anonymous reviewer REFERENCES ABBOTT, M.L. 1963. Lycopod fructifications from the Upper Freeport (no 7) coal in southeastern Ohio. Palaeontographica, Abt.B, 112: 93-118. GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 145 ASH, : R. 1979. peoevirayng gen. nov., a new lycopsid cone from the Early Triassic of Australia. cheringa ASH, [ R., and : - an 1991. A new Jurassic agen nerean from the Wallowa terrane in Hells Canyon, Oregon and Idaho. Amer. J. Bot. 78(12): 1636-1642. BALME, B.E. 1995. Fossil in situ spores and orate grains: an oie catalogue. Rev. Palaeobot. Palynol. 87: 81-323. BARALE, G. 1999. Sur la présence d’une nouvelle espéce d’Isoetites dans la flore du Crétacé infér- ieur de la ar Ge = Tataouine spe Tunisien): implications paléoclimatiques et phylogéné- tiques. Can. J. B : 189- BATEMAN, R M. 1994. ile aidan change in the growth architecture of fossil rhi- zomorphic lycopsids: scenarios constructed on cladistic foundations. Biol. Rev. Cambridge Philos. ye 527-597. BATEMAN, R.M. . An overview - lycophyte phylogeny. iii 405-415 in Pteridology in Per- spective, os Ia Camus et al., Royal Botanic Gardens, BATEMAN, R.M., ina and a ‘4 WILLARD. 1992. oatenenind cladistics of anatomically preserved hie rescent lycopsids from the oe of Euramerica: an essay on paleo- botanical phylogenetics. Ann. Missouri Bot. Gard. 79: 500-559. BHAMBIE, S. 1963. Studies in Pteridophytes. III On the structure and eng sa! of = leaf and sporophyll of Isoetes coromandelina. Proc. Indian acad. Sci., Sect. B, 29: 169-19) BISCHOF, 1853. Abbildungen der Sigillaria Sternbergii. Zeitschr. Ges. eeniiriady = 1; 257; Bd.2: PI.VIII, Halle. BOCHENSKI, T. 1936. Uber a (bliiten) 0 Lepidophyten aus dem produktiven Karbon Polens. Ann. Soc Bock, W. 1969. The American pens flora = lb -aiecennibiae, Geologic Center Research Series 3-4, North Wales, Pennsylvania Bose, M.N. and J. BANERJI. 1984. The fossil pra ‘of Kachchh. I—Mesozoic megafossils. The Pa- sacrteingts 33: 1-189. BRACK, S. D. and T.N. TAYLOR. 1972. The ultrastructure and organization of Endosporites. Micro- agape 18: 101-109. BRAUN, A. 1868. Uber die ee Arten der Gattung Isoétes. Monatsber. KGnigl. Preuss. Akad. eae Berlin, 523-545. Brown, R.W. 1939. Some rinicoe fossil plants belonging to the Isoetales. J. Wash. Acad. Sci. 261-269. BROWN, R. W. 1958. New occurrences of the fossil quillworts called Isoetites. J. Wash. Acad. Sci. 48: 358-361. BROWN, R.W. 1962. oe flora of the Rocky Mountains and Great Plains. Geol. Surv. Prof. Paper 375: 1— CHALONER, W.G. 1967, -Lycophya, Pp.435-781 in Traité de Paléobotanique, ed. E.Boureau II Mas- son et Cie., Par ane W.G. ie The cone of Regan renng iai Haughton, from the Upper Devonian Ireland. J. Linn. Soc., Bot., paises, S. and D.C. McGREGOR. ae ‘Bisprangistobs harrisii gen. et sp. nov., an eligulate lycopsid cone with Duosporites megaspores and Geminospora microspores from the Upper Devonian of Pennsylvania, U.S.A. Pa te sseinuutiags B 210: 127-149. CHITALEY, S., and K.B. PicG. 1996. Clevelandodendron ohioensis, gen. et sp. nov., a slender upright lycopsid from the Late Devonian Cleveland Shale of Ohio. Amer. J. Bot. 83(6): 781-789. COLLINSON, M. 1991. Diversification of modern heterosporous pteridophytes. Pp. 119-150, in: Pol- n and spores, eds. S. Blackmore and S.H. Barnes. Systematic Association Special vol. 44, Clarendon Press, Oxford. DIMICHELE, W.A. and R.M. BATEMAN. 1996. The rhizomorphic lycopsids: a case-study in paleo- tanical classification. Syst. Bot. 21(4): 535-552 DOBRUSKINA, I.A. 1982. Triassic floras of Eurasia. Akad. Sci. USSR Trans., 365:1—193 (in Russian). DosRUSKINA, I.A. 1985. Some problems of the aise of the Triassic lepidophytes. Paleont. ly shee 74-88 (transl. from Paleont. Zhurn., 3: 90-104, in Russian). 146 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 3 (2001) DuTHiE, A.V. 1929. The species of Isoetes found in the union of South Africa. Trans. Roy. Soc. South Africa 17: 523-545. slag L. 1968. Les plantes fossiles dans leurs rapports avec les végétaux vivants. Masson & aris. diana, E. 1856. Geological Report of the Midland counties of North Carolina. G. Putnam, New York, 352 p- FLICHE, P. 1910. Flore fossile du Trias en Lorraine et Franche-Comté (avec des considérations sigs _ M.R. Zeiller). Berger-Levrault, Paris-Nancy, 297 p.; Bull. Soc. Sci. Nanc cy Ill, XI, a0 eee "1900. Notes on fossil plants collected by Dr. Ebenezer Emmons from the older Mesozoic rocks of North Carolina. U.S. Geol. S urvey, Ann. abit 20: (2): 277-315. Fucus, G., L. GRAUVOGEL-STAMM and D. MApER. 1991. Une remarquable flore a Pleuromeia et nan rnin in situ du Buntsandstein moyen (Trias inféiews) de l’Eifel (R. hore Morphologie, ogee et paléogéographie. Palaeontographica Abt. B, 222(4—-6): 89-1 GLAESNER, M.F. and V.R. RAO. 1955. Lower Cretaceous plant remains from the vicinity of Sn cus Goswami, H.K. 1976. A revision of ligule and labium in Isoetes. Acta Soc. Bot. Palawine. XLVI, (1-2): 69-76. GRAUVOGEL-STAMM, L. 1993. Pleuromeia sternbergii (Miinster) Corda from the Lower Triassic =f any—further observations and comparative morphology of its rooting organ. Rev. laeobot. Palynol. 77:185-212 GRAUVOGEL-STAMM, L. 1999. Bivenimeie sternbergii (Miinster) Corda, ein characteristische Pflanze des deutschen Buntsandsteins. Pp. 271-281, in Trias—Eine ganz andere Welt. Europa im GRAUVOGEL-STAMM. L. and S. AsH. 1999. “Lycostrobus” chinleana, an equisetalean cone from the Upper Triassic of the southwestern United States and its phylogenetic implications. Amer. J. Bot. 86(10): 1391-1405. GRAUVOGEL-STAMM, L, and P. DurRINGER. 1983. Annalepis zeilleri Fliche 1910 emend., un organe reproducteur a Lycophyte de la Lettenkohle de |’Est de : France. Morphologie, spores in situ et paléoécologie. Geologische Rundschau 72(1): 23-5 GRAUVOGEL-STAMM, L. and J. LANGIAUX. 1995. Polysporia pies n.sp., un nouvel organe producteur ae lycophyte du pci , fére supérieur) de Blanzy-Montceau (Massif Central, France). Sci. Géol., Bull. 4 HALL, J.B. 1971. a on an ei a. Bot. J. Linn. Soc. 64: 117-139. HELBy, R. and A.R.H.MarTIN. 1965. Cylostrobus gen. nov., cones ‘e oe plants from the Bot Whetorn North D ota. The Geol. Soc. America, Mem. 150: 1-183. HICKEY, R.J. 1985. series studies of Neotropical Isoetes. Ph.D. dissertation, The University of Connecticut. Storr HICKEY, R.J. 1986. The oak evolutionary and ae ree ntay oe of Isoétes, with descriptions of two new neotropical species. Syst. Bot. 11(2): 3 HICKEY, R.J. 1990. Studies of neotropical Isoétes L. I. aie a new subgenus. Ann. Missouri Bot. Gard. 77: 239-245. JENNINGS, JR. 1975: eee a new plant organ from the Lower Mississippian of Virginia. adage. 18(1): 1 JENNINGS, J.R. R. KARRPALT = G.W. ROTHWELL. 1983. Structure and affinities of Protostigmaria eggertiana. heat er. J. Bot. 70(7): 963-974 JERMY, A.C. 1990. Isoetaceae. Pp. 26-31, in The families and genera of vascular plants. I Pteri- dophytes and gymnosperms. Eds. K.U. Kramer and P.S. Green, Springer—Verlag, Berlin KARRFALT, E.E. 1984. Further observations on Nathorstiana (Isoetaceae). Amer. J. Bot. 71(8): 1023- 1030. GRAUVOGEL-STAMM & LUGARDON: TRIASSIC LYCOPSIDS 147 KARRFALT, E.E. and D.A. EGGERT. 1977. The comparative morphology and ee = Isoetes L. I. Lobe and furrow development in I. tuckermanii A.Br. Bot. Gaz. 138(2): 236 neem. E.E. and D.A. EGGERT. 1978. The comparative morphology and ribet af Ionia L. Ill. The sequence of root initiation in Lape and four-lobed plants of J. tuckermanii A. Br. and J. nuttallii A. Br. Bot. Gaz oy 2 KELBER, K.-P. and W. HANscu. 1995. si spliaisoare Die Entratselung einer tiber 200 Millionen ie Kort, L.S. and D.M. BRITTON. 1985. Role of morphological characteristics of leaves and the spo- jig region in the taxonomy of Isoetes in northeastern North America. Amer. Fern J. 75(2): 44— KOVACH, e ‘. 1994. A review of Mesozoic oka ultrastructure. Pp.23—37, in Ultrastructure of fossil spores and pollen, eds. M.H. Kurmann and J.A. Doyle—R.B.G., “sear K. and R. BORCHERT. 1964. Morphologische Studien an Isoetes cain a 7% Braun und Bemerkungen tiber das Verhaltnis der Gattung Stylites E. Amstutz zur Gattung Isoetes L.. Ber. Deutsch. Bot. Ges. 77: 227-233. LUGARDON, B. 1972. Sur la structure fine et la nomenclature des parois microsporales de Selagi- nella denticulata (L) Linck et faa (L) Link. Comptes Rend. Acad. Sci. Paris, Sér. D, Sciences Lippi 274: 1 menclature a anes fine des parois acéto-résistantes des microspores d’'Is patie, Compt Rend. Acad. Sci. Paris, Sér. D, Sciences Naturelles 276: 3017-3020. LUGARDON, B. 1980. Isospore and microspore walls of living Pteridophytes; identification possi- bilities ne different pre ig instruments. Proceedings IV Internat. Palynol. Conf., B.S.I.P. Lucknow, India, 1: 199-206. LUGARDON, B. 1986. Données a tig tales sur la fonction de l’exospore chez les oo Pp: wee 264, in sind and spores: Form and function. Eds. S. Blackmore and I.K. Ferguson (Linn Symp. Ser. N°12), Academic Press, London LUGARDON, . err Pinvidoplivte sporogenesis: a survey of spate wall ontogeny and fine structure in a polyphyletic plant group. Pp. 95-120, in Microspores: Evolution and Ontogeny, eds.S. Blackmore and R.B. Knox. Academic ieee, London. LUGARDON, B., L. GRAUVOGEL-STAMM AND I. DOBRUSKINA. 1997. Sur |’ultrastructure des micro- spores et mégaspores de Pleuromeia rossica (Lycophyta) du Trias inférieur de Russie. XV Symposium APLF, Lyon 1997, Abstract: 41. LUGARDON, B., L. GRAUVOGEL-STAMM AND ‘I DoOBRUSKINA. 1999. The microspores of Pleuromeia rossica Neuburg (Lycopsida; Triassic): comparative ultrastructure and phylogenetic impli- cations. Comptes Rend. Acad. Sci. Paris, Sér. Ila, Sciences de la Terre et des Planétes 329: 435-442. LUGARDON, B., L. GRAUVOGEL-STAMM AND I. DOBRUSKINA. 2000a. Comparative ultrastructure of the megaspores of the Triassic lycopsid Pleuromeia rossica rhe 4 Comptes Rend. Acad. Sci. Paris, Sér. Ila, Sciences de la Terre et des Planétes 330: 50 LUGARDON, B., L. GRAUVOGEL-STAMM, R. COQUEL AND C. coe ak 2000b. Spores a ultrastructure isoétalienne ches des Lycophytes triasiques et plus anciennes (Carbonifére, Dévonien). Colloque OFP, Lyon 2000, Abstract: 11 and OFP Informations Maper, D. 1990. Palaeoecology of the flora in aera and Keuper in the Titinde ot Middle rope. 2 i 1-1582; Gustav Fischer Verlag, Stuttgar eben ta K. 1931. Zur Morphologie und pylogentschon Bedeutung der fossilen Pflanzgat- tung Saal Beih. Bot. Centralbl. 48: MARCINKIEWICZ, T. and E.V. ZHELEZKOVA. 1992. a compatitoit between dispersed megaspores of m =I Q > a6! fs g m8 MENG, F. 1995. Flora of the Badong Formation. Pp. 6-27, in Nonmarine biota ane sedimentary ge ¢ the Badong Formation in the Yangzi and its neighbouring area, an oo N + P. aquilinum D. expansa A. filix femina A. phyllitidis species Fic. 1. Histogram of spore germination plotted as % differences between sucrose treatments (0.05M—dotted; 2M—diamond shading) and 0M sucrose controls. Order of species reflects in- creasing age of spores, youngest far left, oldest far right. specific processes, or it may again be significant that Platycerium is epiphytic, and therefore very different from the terrestrial ferns tested in this study. The spores of Platycerium are also much larger than those of the species tested in the present study, so this difference may reflect different levels of storage car- bohydrate. All the spores of Platycerium tested by Camloh may have had suf- ficient storage carbohydrate to germinate, in which case no increase in ger- mination with added sucrose would be expected. It may also be that a different sugar or medium component would promote germination in this species. The spores Camloh used were stored before use, and it is known that levels of proteins as well as soluble sugars decline with storage of fern spores (Beri & Bir, 1993). It could be speculated, therefore, that the spores of Platycerium that fail to germinate in basal medium including sugar, might germinate if provided with amino acids. Our finding that sucrose did not promote germination of Anemia spores as strongly as the other species may also reflect storage-related decline (see Fig. 1). The Anemia spores used here had been stored for many years longer than the other species. A pilot experiment carried out with the same spore sample five years earlier showed far greater promotion with sucrose (Douglas, 1994). In that experiment the germination percentages (s.e. in parentheses, n = 4) were as follows: OM sucrose 53.1(6.7); 0.05M sucrose 72.5 (6.3); 0.2M sucrose 74.8 (3.4). The percentage germination without sucrose was very similar to that seen here (54.0), but the promotion of germination by sucrose was far more pronounced. This may suggest that during storage another cellular com- ponent had declined to limiting levels by the time the spores were retested. SHEFFIELD ET AL.: SPORE GERMINATION AND GAMETOPHYTE GROWTH 185 The interaction between promotive effects of sucrose and length of time in storage should clearly be tested for individual spore samples, growth condi- tions and species. However, the results indicate that sucrose promotes germi- nation in many ferns and would therefore be worth adding to media used to initiate cultures of a previously untested species. This study examined the effects of sucrose on germination only. Effects of sucrose on subsequent growth of the species studied were not tested and ap- pear to differ from one species to another. Camloh (1993) reported a stimula- tion of growth (in terms of cell number) using sucrose in medium used to culture Platycerium, while Douglas (1994) found evidence of growth inhibition (in terms of numbers achieving 2D growth) in Anemia and Pteridium. Fernan- dez et al. (1997) reported that sucrose inhibited growth (fresh weight and gem- ma formation) in gametophytes of Osmunda regalis. Further experimentation with a wide range of species may reveal some reliable trends, but the current weight of evidence is on the side of deleterious effects of sucrose on growth of terrestrial species. This, coupled with the enhancement of growth of micro- bial contaminants which attends the use of media containing sugars, prompts us to suggest a 2-stage cultivation process in order to optimise yield/growth of cultured fern gametophytes. Germination should be effected in media con- taining sucrose, subsequent growth should employ immobilised or air-lift cul- ture in media with no sucrose. ACKNOWLEDGMENTS We are grateful to Sue Rumsey and Dave Robson for technical support. GED was in receipt of a BBSRC studentship, JMW is in receipt of a NERC CASE studentship. We thank Bob Callow and Russ Drakeley for their help with statistical analysis, and an anonymous referee for suggestions which vastly improved this manuscript. LITERATURE CITED Bert, A., and S. S. Bir. 1993. Germination of stored spores of Pteris vittata. Amer. Fern. J. 83: ba M. 1993. Spore germination and early gametophyte development of moe pind m. Amer. Fern. J. 83:79-85. ———. 1999. Spore age and sterilisation affects nape and early gametophyte development een 2 nigel bifurcatum. Amer. Fern. 1124-132. DEBERGH, BC. Effects of agar brand one" ition a on the tissue culture medium. Phy- iol. se. cent 278. niarus om E. 1994. An investigation into the growth, development and ultrastructure of fern gametophytes in existing and novel culture systems. PhD dissertation, University of Man- aly UK. Douc tas, G. E., and E. SHEFFIELD. 1992. The investigation of ony and novel pear growth systems for the production of fern gametophytes. Pp. 183-187, in J. M. Ide, A. C. Jermy, and A. M. Paul (eds). Fern Horticulture: past, present and future perspectives Intercept Andover. Dyer, A. F. 1979. The culture of fern gametophytes for -Roots and Rhizomes | & am bt ace oe hee ES a es Ey, z a 1. 0.5 0.0 J 40 > 2b. - sf a : es o 3 8 25; b = b SB 20+ : b b b = Sn at ae me sf 2. | b in heels Limi | dl b b b 54 o> 180 ~— 2c. 160 ; 140 + ts] a 2120 | 5 oe aa a Eon} @ +" sagen © eR io! a = 4 & vs] 60 4 ie a id — a b | 40 + oa b b PRES KE : b ‘oi teenie cal ‘jae 0 : j «© Se Se 02-Apr 12-Apr 22-Apr = «O2-May. = 12-May. = 22-May.—O1-Jun date AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) 250 + ance eC, 3a. 200 | — -Old Fronds a 8 — New Fronds a 5 150 | & 4 aialid S$ an IZOMES | + ea Roots and Rhi : a i a. 100 - 4 ee A Me ‘a aoe . ae d d 2 d| era ont ee ee 0 3.5 3b. 3.0 - a $ 25 - § ab 2 20- s Qa Bis. a ¢ E be Cc 2 1.0 cael See S -o Cc 0.5 + ¢ as. Cc Cc Cc Cc 0.0 UJ t qT T mit 3 ae 02-Apr 12-Apr 22-Apr = s«O2-May-=—s«12-May = 22-May 01-Jun date Fic. 3a. Phosphorus content of fern components through spring. 3b. Phosphorus concentration of fern components through spring. Means with the same letter were not significantly different at a = 0.05. Error bars indicate one standard error above and below the mean. net photosynthesis are observed early in the season (Figure 1), it is possible that photosynthesis may be occurring in these fronds during periods of the winter as well (Salisbury 1984, 1985), so long as light sufficiently penetrates the snow pack and temperatures permit metabolic activity. The capacity to photosynthesize during spring may be important to annual net energy capture in Dryopteris intermedia. It may be critical to an understory species that usually receives low light levels in the shade of canopy species during the summer season (Anderson, 1964: Federer and Tanner, 1966; Hutch- ison and Matt, 1977) to take advantage of high light conditions prior to canopy leaf out. Spring ephemerals, species that develop above ground early in spring and senesce with canopy leaf out, are capable of high rates of photosynthesis during spring as compared to summer rates (Harvey 1980) and are adapted to TESSIER: WINTERGREEN FRONDS OF DRYOPTERIS 193 growth and reproduction during this period of high light prior to canopy leaf out. Also, some disturbance-adapted species can acclimate to periods of high light (Brach et al., 1993) and increase their photosynthetic rate. Since Dryop- teris intermedia does not respond to these periods of high light (Brach et al., 1993), maintaining year-round photosynthetic capacity may be critical to op- timizing productivity. A comparison of spring and summer photosynthetic rates would be valuable to test this hypothesis. It is not clear, however, that the C fixed by the wintergreen fronds is trans- located to the new fronds or to any other part of the fern. Carbon gained during the spring period may be lost to soil microbes as the fronds decompose and not be of any immediate benefit to the plant. The non-significant pattern of increased belowground biomass in early spring (Figure 2c) suggests that carbon may be moving from the photosynthesizing fronds into the belowground tis- sue. A critical next step would be to conduct a tracer study to isolate the intermediate and eventual sinks for the spring-fixed carbon. NITROGEN AND PHOSPHORUS POOLS.—As is typical for many herbaceous peren- nial plants (Zavitkovski, 1976), much more biomass is present belowground in Dryopteris intermedia than in aboveground components (Figure 2c). The mean root:shoot for biomass is roughly 4:1 through the spring. Unlike biomass, the concentration of N and P does not differ between mature parts of the plant. In fact the vernal constancy in N and P concentration and content in the win- tergreen fronds suggests that no net mineral nutrient retranslocation to the rest of the plant is occurring with frond senescence. This contrasts with nutrient retranslocation patterns seen during senescence in other herbaceous understo- ry plants (DeMars and Boerner, 1997). Short senescence times can result in greater retranslocation efficiency (del Arco et al., 1991). Slow senescence of the wintergreen fronds (from snowfall through mid spring) may preclude ef- ficient retranslocation in this species. Therefore, maintaining wintergreen fronds as storage organs may not benefit the plant in the retranslocation and subsequent use of the nutrients. PHOTOSYNTHETIC NUTRIENT USE EFFICIENCY.—Maintenance of the wintergreen fronds does, however, increase the nutrient retention time of N and P, thereby increasing nutrient use efficiency (Escudero et al., 1992) of these often limiting nutrients (Chapin, 1980; Vitousek and Howarth, 1991; Yanai, 1992). How much of an increase in nutrient use efficiency does the maintenance of wintergreen fronds in Dryopteris intermedia provide? Often nutrient use efficiency is de- fined as the inverse of the nutrient concentration (Vitousek, 1982; Shaver and Mellilo, 1984; Hirose and Werger, 1994; Minotta and Pinzauti, 1996; Fisk et al. 1998). This definition does not take into account energy expended for re- productive and maintenance efforts, only those expended for growth. There- fore, simply examining the inverse of the nutrient concentration may ignore significant uses of nutrients. A more useful assessment of nutrient use effi- ciency examines total carbon fixed per unit nutrient termed the Potential Pho- tosynthetic Nutrient Use Efficiency (Field and Mooney, 1986; Hirose and Wer- ger, 1994). 194 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) In this study, ambient light intensities were used to assess net photosyn- thetic rates since these rates are a better representation of the actual benefit of maintaining wintergreen fronds than would be light saturated net photosyn- thetic rates. Photosynthetic nutrient use efficiencies (PNUE) were 9.80, 1.44, and 2.96 pmols CO, / g / second for nitrogen on April 3, April 31, and May 14, respectively. For phosphorus PNUE were 0.19, 0.03, and 0.06 pmols CO, / mg / second on April 3, April 31, and May 14, respectively. Vitousek (1982) points out that one way to improve nutrient use efficiency is by fixing C for a longer period of time. In this case, springtime photosynthesis in wintergreen fronds clearly improves nutrient use efficiency since without this springtime photosynthesis PNUE would be zero until the new fronds began photosynthe- sizing. In summary, vernal net photosynthesis does occur in the wintergreen fronds of Dryopteris intermedia, however it is not clear that the fixed carbon is trans- located to the rest of the plant, and a tracer study would be useful to test this possibility. No evidence of N and P retranslocation from senescing wintergreen fronds to the rest of the fern plant was observed. Maintenance of wintergreen fronds may simply increase photosynthetic nutrient use efficiency of nitrogen and phosphorus. Future studies should examine the long-term fate of carbon and mineral nutrients held within senescing fronds and compare spring and summer photosynthetic rates in this species. ACKNOWLEDGMENTS Several people and organizations assisted with this work and the author thanks each of them for their assistance. The Frost Valley YMCA permitted use of their forested property and the New York City Department of Environmental Protection provided funding. Sam McNaughton permitted the use of his LCA-3. Lisa Tessier, Kim Anderson, Tim Schreiber, Karl Didier, Steve Fuller, Dave Kubek, Scott Jones, Kim Keirnan, and Tom Touchet assisted with work in the field. Don Bickel- haupt assisted with the Kjeldahl digestions and Marlene Braun and Deb Driscoll assisted with the ICP. Dudley Raynal, Scott Heckathorn, Ruth Yanai, and an anonymous reviewer commented on earlier versions of the manuscript and provided many helpful suggestions. Lastly, the SUNY Col- lege of Environmental Science and Forestry’s Ecolunch group provided intellectual input and support. LITERATURE CITED Chamber Analysis System. Analytical Development Co. Ltd., Hoddesdon, England. 33pp. ANDERSON, M. C. 1964. Studies of the woodland light climate II. Seasonal variation in the light regime. J. Ecol. 52:643-663. Avery, T. E. and H. E. BURKHART. 1994. Forest Measurements. 4‘ Edition. McGraw-Hill., New York. 40 BICKELHAUPT, D. H. and E. H. Wuire. 1982. Laboratory Manual for Soil and Plant Tissue Analysis. State University of New York College of Environmental Science and Forestry. Syracuse, New York. 67pp. BRACH, A.R., S. J. MCNauGHTon, and D. J. RAYNAL. 1993. Photosynthetic adaptability of two fern species of a northern hardwood forest. Amer. Fern J. 83:47-53. CuHasor, B. F. and D. J. Hicks. 1982. The ecology of leaf life spans. Ann. Rev. Ecol. Syst. 13:229- 259. TESSIER: WINTERGREEN FRONDS OF DRYOPTERIS 195 CHAPIN F. S. 1980. The mineral nutrition of wild plants. Ann. opie Ecol. Syst. 11: peciige CREED, I. F., L. E. BAND, N. W. Foster, I. K. Morrison, J. A. NICOLSON, R. S. SEMKIN, and D se te 1996. ani of nitrate-N release fein temperate ieee A test of the N fvikine hypothesis. ieee Resour. Res. 32:3337-3354. DEL ARCO, J. M., A. ESCUDERO, and M. V. GARRIDO. 1991. Effects of site characteristics on nitrogen ea from senescing ee Foolagy 7 08. DeMars, B. G. and R. E. J. BOERNER. 1997. Foliar phosphors and nitrogen resorption in three woodland herbs of contrasting eat Castanea 62:43-54. ESCUDERO, A., J. M. DEL Arco, I. C. SANZ, and J. AYALA. ee Effects of leaf longevity and re- ean ragae’e efictacy on the ae time of nutrients in the leaf biomass of different oody species. Oecologia 90:80-87. oe - ing 1989. Photosynthesis and nitrogen relationships in leaves of C, plants. Oecologia 78: Feo, ie A. and C. B. TANNER. 1966. Spectral distribution of light in the forest. Ecology 47: 55-560 ie Gs ad 8 A. Mooney. 1986. The photosynthesis-nitrogen relationship in wild plants. In: Tey. Givininti; ed., On the Economy of Plant Form and Function. Cambridge University Press. New York. 7 Fisk, M.G.,.S. K. SCHMIDT, and T. R. SEASTEDT. 1998. Topographic patterns _ naeeiet and below- ground production and nitrogen cycling in alpine tundra. Ecology 79:2 2266. GILMAN, L. B. 1988. General Guidelines for Microwave Sample calle ee Corporation. atthews, North Carolina. Pe B. S. and D. W. LARSON. 1992. Seasonal changes in oe biemaiaaaae in the desiccation- tolerant fern Si virginianum. Oecologia 89:38 Harvey, G. W. 1980. Seasonal alteration of aay sizes in three herb layer compo- nents of a deciduous forest community. Amer. J. Bot. 67:293—299. Hirose, T. and M. J. A. WERGER. 1994. Photosynthetic capacity and nitrogen a among species in the canopy of a — plant community. Oecologia 100:203—21 HuTcHIson, B. A. and D. R. Matt. 1977. The distribution of solar radiation within a ne forest. Ecol. Monogr. 47:185—20 LuLow, C. J. and F. T. WoLF. 1975. aeons and respiration rates of ferns. Amer. Fern J. MAHAL, B. E. aad F. H. BORMANN. 1978. A apagetocaged description of the vegetative phenology herbs in a northern hardwood forest. Bot. Gaz MINOLETTI, M. L. and R. E. J. BOERNER. 1993. Seasonal photosynthesis, nitrogen and peste namics and resorption in the wintergreen fern Polys Bull. Torrey Bot. Club 120:397—404. Minotta, G. and S. PINZAUTI. 1996. Effects of light and soil fertility on growth, leaf chlorophyll content and pid use efficiency of beech (Fagus sylvatica L.) seedlings. Forest Ecol. Man- agem. 86:61— MITCHELL, M. J., D i" Risen, and C. T. DRISCOLL. 1996. Biogeochemistry of a forested watershed in the central Adirondack Mountains: Temporal changes and mass balances. Water, Air, and Soil Pollution 88:355-369. ONK, C. D. 1966. An ecological significance of evergreenness. Ecology 47:504—505. Mook, P. D. 1984. Why be an evergreen? Nature Murpocu, P. S. and J. L. SroppARD. 1992. The role of nitets in bee acidification of streams in the Catskill b Mountains of New Water Resour. Res. 28:2 20. Murpoc#, P. S. and J. L. SroppARD. 1993. Chemical sas iret ie temporal trends in eight streams of the Catskill Mountains, pte York. Water, Air, and Soil Pollution 67:3667—395. REIFSNYDER, W. E., G. M. FURNIVAL, and J. L. Horowitz. 1971. Spatial and ie ag distribution of solar radiation beneath forest canopies. Agricultural Meteorology 9:2 SAS INSTITUTE INC. 1997. Selected SAS ies Bi for APM 620: Analysis = Vari (Spring 1998). SAS age Inc. Cary, North Car venerrtand F.B . Light condition s and a nee under snow. In: J. F. Merritt. at Ecology sf Sunk Mammals. Carnegie Museum of Natural History. Pittsburgh. 380 196 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) . Plant growth ane snow. Aquilo, Ser. Bot. 23:1 SHAVER, : ei and J. M. MELLILO. 1984. Nutrient budgets of oan plants: Efficiency concepts and relation to availability. eooléey 65:1491-1510. VAN ae J. and J. Eowarps. 1995. Contribution bs hears leaves to early spring growth n the wood fern Dryopteris intermedia. Amer. Fern J. 85:5 Sacenae P.M. 1982. Nutrient cycling and porn use at eotie vue Naturalist 119:553-572. baleen P. M. and R. W. Howartu. 1991. Nitrogen limitation on land and in the sea: How can cur? Biogeochem. 13:87-115. Yavat R A 1992. Phosphorus budget of a 70-year-old northern hardwood forest. Biogeochem. a ee y 1976. Ground vegetation biomass, production, one serie of energy utilization in some northern Wisconsin forest ecosystems. Ecology 57: 06. ZHANG, Y. and M. J. MITCHELL. 1995. Phosphorus cycling in a ‘ede’ forest in the Adirondack Mountains, New York. Canad. J. Forest. Res. 25:81-87 American Fern Journal 91(4):197—213 (2001) Phylogenetic Placements of Loxoscaphe thecifera (Aspleniaceae) and Actiniopteris radiata (Pteridaceae) Based on Analysis of rbcL Nucleotide Sequences GERALD J. GASTONY and WILLIAM P. JOHNSON Department of Biology, Indiana University, Bloomington, IN 47405-3700 ABSTRACT.—Nucleotide sequences of the chloroplast-encoded rbcL gene were determined for Lox- oscaphe thecifera and Actiniopteris radiata and used in maximum parsimony cladistic analyses to determine their phylogenetic positions in the context of a broad range of advanced fern taxa. Loxoscaphe nested firmly within the Aspleniaceae, and Actiniopteris was placed with Onychium in the Pteridaceae. To help resolve conflicting contemporary treatments that either subsume Lox- oscaphe species within Asplenium or segregate them as an independent genus, the rbcL sequence of L. thecifera was subjected to a more focused analysis involving all rbcl. sequences available to represent the taxonomic diversity of Aspleniaceae. Loxoscaphe thecifera was sister to Asplenium griffithianum+ | po hea robustly and surprisingly nested within the clade of Asplenium spe- cies recognized as A splenium section Thamnopteris, a group accepted by some as the segregate accepting Loxoscaphe as a genus independent of Asplenium. Similarly Actiniopteris radiata, re- cently moved from the cheilanthoid to the taenitidoid group of Pteridaceae, was subjected to a more focused analysis in the context of an expanded set of cheilanthoid and taenitidoid species that included the first use of an rbcL sequence from the genus Anogramma and newly sequenced species of Onychium and Pteris. Actiniopteris is robustly grouped with two Onychium species in a Clade sister to traditional taenitidoids and deeply separated from the cheilanthoids, supporting affinities previously suggested by spore morphology. Loxoscaphe T. Moore is a genus of epiphytic species segregated from As- plenium L. by many contemporary authors (e.g., Crabbe et al., 1975; Pichi Sermolli, 1977; Smith, 1981; Mickel and Beitel, 1988; Lellinger, 1989), al- though others (Tryon and Tryon, 1982; Kramer and Viane, 1990; Tryon and Stolze, 1993) reject this segregation, instead subsuming Loxoscaphe species within the genus Asplenium. Mickel and Beitel (1988) expressed the segre- gationist view in stating that ‘““Loxoscaphe is a genus of four or five species, one in the neotropics, two in Africa (one sometimes considered only varietally distinct from the neotropical species), and two in the South Pacific and In- donesia. Although it is most closely allied to and sometimes placed in syn- onymy under Asplenium, Loxoscaphe superficially more closely resembles Odontosoria in the pocketlike sori terminal on the narrow blade segments, but the small fronds, heavy texture, and epiphytic habit readily distinguish it from that genus.” Its sori in pouches near the margin of the leaf segments (Fig. 1) give it the appearance of a Davallia, in which genus most of the older species of Loxoscaphe were first described, as noted by Copeland (1947). In com- menting on the generic placement of Loxoscaphe thecifera (Kunth) T. Moore (as Asplenium) in their treatment of the ferns of Peru, Tryon and Stolze (1993) 198 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) Fics. 1-3. Illustrations of Loxoscaphe thecifera and Actiniopteris radiata. 1. Apex of a pinna of Loxoscaphe thecifera drawn from Holm-Nielson et al. 4110, Ecuador (F) to show the pouch-like 2 late lamina of this xeromorphic fern. Scale bar = 5 cm. 3. Magnified view of fascicle of segments from the right half of the lamina in Fig. 2, showing continuous, membranous, marginal indusia partly covering the submarginal sori. Scale bar = 2 cm. noted that the group of Asplenium species often placed in the genus Loxo- scaphe are quite distinctive in their indusia, which are generally pouch- or cup-shaped and occur near the segment margin. They noted that by means of this indusial character state, Asplenium theciferum (Kunth) Mett. cannot be confused with any other Asplenium in Peru, their area of coverage. Their re- tention of the Loxoscaphe group within Asplenium was influenced by their observation that the sori of a few Old World species in the group intergrade to a more typical Asplenium sorus. The global analysis of fern phylogeny by Hasebe et al. (1995) based on rbcL nucleotide sequences did not include a representative of the Loxoscaphe spe- cies group, but the results of that study do provide a context in which the generic relationships of Loxoscaphe can now be assessed. Doing that is one goal of this paper. To focus the reader’s attention on Loxoscaphe and other relevant groups sometimes proposed for generic segregation from Asplenium, the following text and figures 5 and 6 refer to these taxa by their names in segregate genera. This is simply an attention-focusing device and does not signal our taxonomic conclusions, which are sometimes contrary to the seg- GASTONY AND JOHNSON: LOXOSCAPHE AND ACTINIOPTERIS 199 regate nomenclature. Alternative nomenclatures for these groups are noted in the lists of species in MATERIALS AND METHODS and in Table 1. Actiniopteris Link was regarded as an Old World genus of cheilanthoid af- finities by Tryon and Tryon (1982) who placed it (as Actinopteris) in Pterida- ceae tribe Cheilantheae. Later, Tryon et al. (1990) shifted Actiniopteris from cheilanthoid to taenitidoid affinity by placing it in Pteridaceae subfamily Taen- itidoideae, without explanation. Actiniopteris radiata (Sw.) Link, with an un- usual fan-shaped lamina (Fig. 2) bearing linear, submarginal sori with contin- uous, membranous, marginal indusia (Fig. 3), is widespread within the Afro- Indian distribution of its genus (Tryon et al., 1990). It occurs, for example, throughout tropical Africa where hot and dry conditions prevail (Jacobsen, 1983). Continuing the ongoing efforts of the first author to decipher phyloge- netic relationships among the lineages of cheilanthoid ferns (Gastony and Rol- lo, 1995, 1998), a second goal of the present study is inferring the subfamilial affinities of Actiniopteris on the basis of rbcL data. The first author encountered epiphytic plants of Loxoscaphe thecifera on the rim of the Ngorongoro Crater and terrestrial plants of Actiniopteris radiata in fissures of the limestone cap bordering Olduvai Gorge, both during a recent trip to Tanzania. Materials suitable for DNA-level analyses of these species were obtained. The resulting rbcL-based phylogenetic analyses determine the phylogenetic relationships of these species and provide a fresh perspective for addressing the generic disposition of Loxoscaphe. MATERIALS AND METHODS Total genomic DNA of Actiniopteris radiata and Loxoscaphe thecifera was extracted by the DNeasy method (Qiagen, Inc., Valencia, CA) from silica-gel dried leaves of sporophytes collected in nature (Table 1), purified with the Elu-Quick DNA purification kit (Schleicher & Schuell, Keene, NH), and quan- tified to 20 ng/yl with a Hoeffer Scientific fluorometer. DNA of the other spe- cies in Table 1 was obtained as follows: Anogramma lorentzii from gameto- phytes grown from spores of the Brazilian sporophyte voucher, Pteris cretica from fresh, field-collected sporophytes, and Onychium lucidum from sporo- phytes grown from spores of the Chinese voucher. Polymerase Chain Reaction (PCR) amplification of the rbcL gene was as reported in Gastony and Rollo (1995) except that for Loxoscaphe thecifera the hot start PCR method of Gas- tony and Ungerer (1997) was used with the following specifications: 0.8 pl @20 ng/pl of genomic template DNA was added to 8.62 yl distilled H,O, 3.6 wl of Master Mix (see components below), 0.24 wl of 10 wM forward primer 1F, and 0.24 pl of 10 wM reverse primer 1351R (primers 1F and 1351R as in Gastony and Rollo, 1995). In PCR, these reagents were heated to 96°C for 5 min to denature the templates before adding 1.5 wl Taq polymerase (“hot start”) during a 5 min period at 72°C. The final concentration of the hot start Master Mix was 2 mM MgCl, 30 mM tricine, 50 mM KCI, 100 pM of each dNTP, and 5% acetamide. In all cases PCR was carried out in an MJ Research Thermal Cycler programmed as follows (after the initial 96°C denaturation or AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) 200 L O9E09EHV 807-68 AUOIsVD >? YoOKaarysinx OOIX9|] ‘OSTepIH ] VINAAD SMANT a: POLOESAV 767-68 Auo|svpyH ? YyokaalysivX ODTXI|Y ‘VOVXBOC) UOALL “WY CD) pipyofial punupssosdiy i COLOEEAHV [SZ-6§ Auolsvy »? YoaarysivZ OOTXOJAL ‘BVIBXBO yur] (J) Sounjawojpps pUUDASOAN ig 6SEO9EHV 78-§ duplx »? 17 eulyd ‘ueuuny “Suaids umpiony wniyscud L‘? LOLSECAV u's yorupdyAry W'S’ ‘erlusoy yea ‘NIA (YOoH) vousofijvs sayjunpiayy 9‘S ‘PY 6609€E4V 101-Z6-Z] Auolsvy viueZUR], “1dIeID O10SUOIOS NY O00, ‘L (yUNY) Buafioays aydvIsoxOT =: Noy (IUNY) Wndafioays wniuajds7 L COL9eCAV TL/OL/OL “U's wauyas [Ize1g [Ng Op spurIH ory SPI (UOsaTH) 12iuMa10] DuUD.ssoury L‘v OOT9ECAV 2O1-L6-Z] Auoispp vruezuRy ‘adIOH wAanpIO AUNT CMS) vinipes stsaidoquna7 Apmis sty} “OU ~UOIssadoR (49ySNOA,) UOTIDATJOD DOULUIAOI saiads UI Soins yur guan “CNI 18 pasoyonoa are Mojaq suotdaT[oo [Ty “(satoads asoyy Jo sist] Icy Spoypoy] pur speLiayey oes) osoy poyadeas jou ase (q6G6G]) “[e 19 HUIRYRINYY JO | BQRL JO (PGGH]) [B19 TwIRYRINYY JO Z AIqQeL ‘(SH6I) ‘Te 19 aqaseH Jo xIpuaddw ay) ur poysiqnd a1am eyep yorym joj sotoadg “Apnys sty) ul pasn exei Jo saouanbas oq! JO sIaquuNU UOTSsa00R YURGUaDH pue [eLIa}VUI JO sdoOINOg “| ATAVL GASTONY AND JOHNSON: LOXOSCAPHE AND ACTINIOPTERIS 201 after the hot start addition of Taq): 40 cycles of 94°C for 1 min, 45°C for 1 min, 72°C for 2 min, followed by a 6 min final extension at 72°C and storage at 4°C. PCR yielded a 1381 base pair fragment of the rbcL gene, including the forward and reverse primer regions. With the suite of forward and reverse sequencing primers listed in Table 2 of Gastony and Rollo (1995), we were able to read 1325 bp of sequence between the PCR primers in both directions. Sequencing was carried out with an ABI automated sequencer according to manufacturer’s protocols, and the contigs were assembled and read with Sequencher™ 3.1.1 (Gene Codes Corporation, Ann Arbor, MI). To determine the phylogenetic relationships of Loxoscaphe thecifera and Actiniopteris radiata, their 1325 bp rbcL sequences (Table 1) were analyzed cladistically with PAUP* 4.0b4a (Swofford, 1998) in three separate analyses as follows. In the first analysis, the Loxoscaphe and Actiniopteris sequences were com- bined with those of a large subset of the species used in the global analysis of fern phylogeny of Hasebe et al. (1995) to determine where Loxoscaphe and Actiniopteris would probably have been positioned if their rbcL sequences had been included in that 1995 analysis. In reference to Fig. 3 of Hasebe et al. (1995), the rbcL sequences of all taxa from Orthiopteris through Loxogramme were obtained from GenBank (except for Cornopteris crenulatoserrulata, De- paria bonincola, and Hypodematium crenatum, which were provided by Mit- suyasu Hasebe) and used for the ingroup for this first analysis along with the new sequences of Loxoscaphe and Actiniopteris and that of Cheilanthes cali- fornica (Table 1), an American cheilanthoid sometimes segregated into the genus Aspidotis and shown by Gastony and Rollo (1998) to have Asian affin- ities. This broad spectrum of taxa includes dennstaedtioids, lindsaeoids, Dav- allia, taenitidoids, cheilanthoids, and other advanced fern taxa in order to permit free positioning of Loxoscaphe and Actiniopteris wherever their rbcL affinities might reasonably lie. As outgroup for this analysis the GenBank rbcL sequences of four taxa (Loxoma, Cyathea [represented by Cyathea lepifera, more appropriately named Sphaeropteris lepifera as in our Fig. 4], Sphaer- opteris, and Metaxya) from the clade sister to this ingroup in Fig. 3 of Hasebe et al. (1995) were selected. Thus, 69 taxa (65 ingroup plus 4 outgroup, Fig. 4) were used in this first analysis with PAUP*4.0b4a (Swofford, 1998) on a Mac- intosh PowerPC G4 350MHz with the following specifications: maximum par- simony, heuristic search, all characters equally weighted, starting tree by step- wise addition, random addition sequence with 1000 replicates, random start- ing seed number generated by the program, TBR branch swapping, MulTrees, and steepest descent. Of the 1325 characters, 504 were parsimony-informative. The positions of Loxoscaphe and Actiniopteris determined by this first anal- ysis permitted subsequent selection of appropriate new ingroups and out- groups for each of them in more detailed analyses of their respective relation- ships in the second and third analyses. The first analysis placed Loxoscaphe with Asplenium and its relatives, and Actiniopteris in the clade of Pteridaceae plus Vittariaceae (Fig. 4). For a more focused analysis of the rbcL relationships of Loxoscaphe within Aspleniaceae 202 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) oxogramme grammitoides lata tumohra adiantiformis ‘laphoglossum hybridum tty >s> OSs a 2 gS Dryopteris cristata esatichum ih saline tis e Ke > =) = =) s 386 Ne aS Q > = > a 3 = s = S 8 wB =f S23 ie oS, S>SfkS Ss & 3 eS ral “A l Pellaea tandromedifo lia ‘esta ic concolor demionitis levyi iste rose Fe pk anti edatum rus angustifolius >>PoOSPODOO ay gE SORES — = i=) ‘ Hue s22F Sc is) as 8 = 5 8 indsaea odora ata nhen rthiopteris Lingii OUTGROUP as ee SAAMO), aS 5 & Ss 3 % be . Ds s v SI “sh & S élaxya rostrata GASTONY AND JOHNSON: LOXOSCAPHE AND ACTINIOPTERIS 203 the clade from Asplenium through Camptosorus of Fig. 4 was therefore se- lected as a new ingroup, to which were added Aspleniaceae sequences avail- able in GenBank (27 total ingroup taxa, Fig. 5), and a subset of seven taxa in the sister clade from Loxogramme grammitoides through Thelypteris beddomei (Fig. 4) was selected as the outgroup (Fig. 5). Most of these ingroup species were used in a recent rbcL study of the phylogeny of Aspleniaceae (Murakami et al., 1999b), the results of which are seen in Fig. 1 of that study. Use of generic names Camptosorus, Neottopteris, Phyllitis, and Hymenasplenium in discussing our results in Figs. 5 and 6 below facilitates comparison with Fig. 1 of Murakami et al. (1999b). Boniniella ikenoi of Murakami et al. (1999b) Fig. 1 is reported here under its name as a species of Hymenasplenium, H. cardi- ophyllum. Of the 1325 characters in this data set, 275 were parsimony-infor- mative, probably reflecting the fact that the GenBank sequences of many of the Asplenium species used in the ingroup for Figs. 5 and 6 were much shorter than those generated in our lab, and their 5’ and 3’ end truncations relative to ours were treated as missing data. Similarly, for a more focused analysis of the rbcL relationships of Actinio- pteris, the clade from Platyzoma microphyllum through Haplopteris flexuosa of Fig. 4 was selected as a new ingroup. To these were added (Table 1) species of the putatively taenitidoid (Tryon and Tryon, 1982; Tryon et al., 1990) genera Pityrogramma (P. calomelanos and P. trifoliata, both previously used by Gas- tony and Rollo, 1998) and Anogramma (A. lorentzii, new here), and additional species of Onychium, Pteris, and Cheilanthes (O. lucidum and P. cretica, both new here; C. Janosa previously used by Gastony and Rollo, 1995, 1998), yield- ing 26 total ingroup taxa (Fig. 7, Table 1). As outgroup to this new Actiniopteris ingroup the sister clade Coniogramme of Fig. 4 was used, plus two basal taxa from the next most sister clade, viz. Monachosorum arakii and Microlepia strigosa (Fig. 4). Of the 1325 characters in the data set, 374 were parsimony- informative. Both of these more focused analyses were carried out in the same manner as the first analysis. The bootstrap option in PAUP* was used to assess support for clades re- vealed in each of the three analyses. Bootstrapping for the first analysis was Fic. 4. Single most parsimonious tree placing Loxoscaphe thecifera and Actiniopteris radiata (bold) within the context of taxa seen in Fig. 3 of Hasebe et al. (1995), with the addition of of synapomorphies supporting each clade and autapomorphic branch lengths for terminal taxa are indicated above the lines. Bootstrap percentages based on 1000 replicates of five random addition sequence replicates each are provided below the lines. The placements of Loxoscaphe thecifera and Actiniopteris radiata determined here directed choices of more focused ingroups and out- groups for the analyses reported in Figs. 5—7, as explained in the text. 204 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) Asplenium normale 100 Asplenium oligophlebium Asplenium trichomanes 100 Asplenium tripteropus 88 Asplenium incisum Asplenium sarelii Fo Asplenium tenuicaule Camptosorus sibiricus 63] [7 Asplenium wrightii bi Asplenium ritoense Asplenium wilfordii a Ag. Asplenium pseudo-wilfordii Saige a Asplenium ensiforme Neottonteric + P 100 [i Neottopteris antiqua a See Neottopteris nidus Phyllitis scolopendrium Hymenasplenium hondoense Hymenasplenium cardiophyllum 97 Hymenasplenium obliquissimum 49 53 Hymenasplenium cheilosorum 100 A iacaaedl stag Hymenasplenium laetum Hymenasplenium riparium Oleandra pistillaris Nephrolepis cordifolia Rumohra adiantiformis Hypodematium crenatum Thelypteris beddomei OUTGROUP Athyrium filix-femina Diplazium esculentum - 5. Strict consensus of 15 equally most parsimonious trees based on the new rbcL sequence of Loxoscaphe thecifera (bold) plus all rbcL sequences of species in Aspleniaceae presently avail- lasicum = Neottopteris australasica by Murakami et al. [1999a]) would be placed in Neottopteris if that genus were accepted and the combination were made. GASTONY AND JOHNSON: LOXOSCAPHE AND ACTINIOPTERIS 205 18 Asplenium normale Asplenium oligophlebium Asplenium incisum Camptosorus sibiricus 12 : i ae 5 [— Asplenium wrightii 79 Lo Asplenium ritoense 22 wu Asplenium wilfordii ae 14 fc Asplenium pseudo-wilfordii Bi sa ot Gee Asplenium ensiforme S Neottopteris australasica NEOT Asplenium setoi Loxoscaphe thecifera Asplenium griffithianium Asplenium prolongatum 2 Neottopteris antiqua 99 = a Neottopteris nidus Phyllitis scolopendrium Hymenasplenium hondoense Hymenasplenium cardiophyllum Hymenasplenium obliquissimum Hymenasplenium cheilosorum Hymenasplenium laetum Hymenasplenium riparium Oleandra pistillaris Nephrolepis cordifolia Rumohra adiantiformis Hypodematium crenatum Thelypteris beddomei OUTGROUP | Athyrium filix-femina 22 Diplazium esculentum Fic. 6. One of the equally most parsimonious trees randomly selected from 15 represented by the consensus tree of Fig. 5 in order to present branch lengths (synapomorhies and autapomor- phies) above the lines and bootstrap percentages as in Fig. 5 below the lines. The position of Loxoscaphe thecifera is given in bold. Branch lengths for the subclade Neottopteris australasica through N. nidus, in which Loxoscaphe is nested, are identical for all 15 equally parsimonious trees. 206 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) 11 Platyzoma microphyllum Taenitis blechnoides Anogramma lorentzii Pityrogramma calomelanos Pityrogramma trifoliata Actiniopteris radiata Onychium japonicum : Onychium lucidum 31 F ie. 10 pa Preris fauriei i ee ee Pteris cretica 89 : ’ . 39 [7 Ceratopteris thalictroides 99 Se Acrostichum aureum 7 = Argyrochosma delicatula Pellaea andromedifolia Cheilanthes allosuroides Cheilanthes lanosa Doryopteris concolor Hemionitis levyi Cheilanthes californica Notholaena rosei Bommeria ehrenbergiana Adiantum pedatum Ananthacorus angustifolius Polytaenium lineatum Antrophyum reticulatum Haplopteris flexuosa OUTGROUP Coniogramme japonica 58 ka Monachosorum arakii 69 ai Microlepia strigosa Fic. 7. Single most parsimonious tree based on new rbcL sequence of Actiniopteris radiata (bold), with other ingroup taxa drawn from the clade of Platyzoma through Vittaria in Fig. 4 supple- GASTONY AND JOHNSON: LOXOSCAPHE AND ACTINIOPTERIS 207 carried out via a maximum parsimony heuristic search of 1000 bootstrap rep- licates, with the starting tree obtained by stepwise addition based on a random addition sequence of five replicates, the random starting seed number gener- ated by the program, and with TBR branch swapping, MulTrees, and steepest descent in effect. Bootstrapping for the two more focused analyses used the same settings and parameters as the first, except that for the more focused Actiniopteris analysis the 1000 bootstrap replicates were based on a random addition sequence of 10 replicates each. To avoid unnecessary repetition of previously published voucher and GenBank data for species used in our analyses and figures, that information is not presented in Table 1 but can be found as follows (authors of names are given only when not found in the respective following references). See Ap- pendix of Hasebe et al. (1995) for the following species: Acrostichum aureum, Adiantum pedatum, Ananthacorus angustifolius, Antrophyum reticulatum, Arachniodes aristata, Argyrochosma delicatula, Arthropteris beckleri, Asplen- ium antiquum (= Neottopteris antiqua (Makino) Masam.), Asplenium nidus (= Neottopteris nidus (L.) Hook.), Asplenium ruprechtii (= Camptosorus si- biricus), Athyrium filix-femina, Blechnum orientale, Blotiella pubescens, Bom- meria ehrenbergiana, Ceratopteris thalictroides, Cheilanthes allosuroides, Cheilanthes lanosa, Coniogramme japonica, Cornopteris crenulatoserrulata, Ctenitis eatonii, Cyathea lepifera (see Sphaeropteris lepifera below), Cycloso- rus opulentus, Cystopteris fragilis, Davallia denticulata, Dennstaedtia puncti- lobula, Deparia bonincola, Diplazium esculentum, Doodia maxima, Doryo- pteris concolor, Dryopteris cristata, Elaphoglossum hybridum, Gymnocarpium dryopteris, Haplopteris flexuosa (Fée) E. H. Crane as Vittaria flexuosa in Has- ebe et al. (1995), Hemionitis levyi, Histiopteris incisa, Hypodematium crena- tum, Hypolepis punctata, Lindsaea odorata, Lonchitis hirsuta, Loxogramme grammitoides, Loxoma cunninghamii, Matteuccia struthiopteris, Metaxya ros- trata, Microlepia strigosa, Monachosorum arakii, Nephrolepis cordifolia, Noth- olaena rosei, Oleandra pistillaris, Onoclea sensibilis, Onychium japonicum, Orthiopteris kingii, Paesia scaberula, Pellaea andromedifolia, Phyllitis scolo- pendrium, Platyzoma microphyllum, Polystichum tripteron, Polytaenium Ii- neatum, Pteridium aquilinum, Pteris fauriei, Rumohra adiantiformis, Sadleria pallida, Sphaeropteris lepifera (Hook.) R. M. Tryon as Cyathea lepifera in Has- ebe et al. (1995), Sphaeropteris cooperi, Sphenomeris chinensis (note that the GenBank accession for this uses the synonym Odontosoria chinensis), Sten- ochlaena palustris, Taenitis blechnoides, Tectaria gaudichaudii, Thelypteris beddomei, Vittaria flexuosa (see Haplopteris flexuosa above), Woodsia polys- tichoides. See Table 2 of Murakami et al. (1999a) for the following (when more than one accession of a species is listed in that table, the GenBank accession number is given here to indicate which accession was used in our analysis): Asplenium australasicum AB013249 (= Neottopteris australasica J. Sm.), As- plenium griffithianum, Asplenium setoi AB013243 (segregated from A. austra- lasicum by Murakami et al. [1999a]; would be a species of Neottopteris if that genus were accepted; combination apparently not yet made). See Table 1 of Murakami et al. (1999b) for: Asplenium cardiophyllum (= Hymenasplenium 208 AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) cardiophyllum = Boniniella ikenoi), Asplenium cheilosorum (= Hymenas- plenium cheilosorum), Asplenium ensiforme, Asplenium hondoense (= Hy- menasplenium hondoense), Asplenium incisum, Asplenium laetum (= Hy- menasplenium laetum), Asplenium normale, Asplenium obliquissimum (= Hymenasplenium obliquissimum), Asplenium oligophlebium, Asplenium pro- longatum, Asplenium pseudo-wilfordii, Asplenium riparium (= Hymenasplen- jum riparium), Asplenium ritoense, Asplenium sarelii, Asplenium tenuicaule, Asplenium trichomanes, Asplenium tripteropus, Asplenium wilfordii, Asplen- ium wrightii. RESULTS The first analysis, designed to place Loxoscaphe and Actiniopteris within the context of the taxa in Fig. 3 of Hasebe et al. (1995), resulted in a single most parsimonious tree of 3573 steps (Fig. 4) whose general topology is similar to that of Fig. 3 in Hasebe et al. (1995). Loxoscaphe thecifera is positioned sister to Asplenium antiquum (23 synapomorphies and 100% bootstrap con- fidence) in the robustly supported (65 synapomorphies, 100% bootstrap value) clade of Aspleniaceae (Asplenium antiquum through Camptosorus sibiricus) near the center of the tree. This is far removed from genera noted in the lit- erature (Mickel and Beitel, 1988; Copeland, 1947) as having superficially sim- ilar soral pouches—Odontosoria and its relatives toward the base of the tree (where Odontosoria chinensis is reported as Sphenomeris chinensis, as it was in Fig. 3 of Hasebe et al. 1995) and Davallia near the top. Actiniopteris radiata nests deeply within Pteridaceae+ Vittariaceae (the clade from Platyzoma mi- crophyllum through Haplopteris flexuosa) in the lower half of the tree, where its sister relationship to Onychium japonicum is strongly supported by 17 synapomorphies and a 97% bootstrap value. Actiniopteris is further placed by 22 synapomorphies and a 90% bootstrap value within the clade (Platvyzoma through Acrostichum) that is sister to the clade containing the deeply sepa- rated and strongly supported cheilanthoids (Argyrochosma delicatula through Bommeria ehrenbergiana) that are themselves united by 19 synapomorphies and a 98% bootstrap value. The more focused analysis designed to position Loxoscaphe within a context of all species of Aspleniaceae for which rbcL sequences were available in GenBank yielded 15 equally most parsimonious trees of 871 steps whose con- sensus topology (Fig. 5) is comparable to that in Fig. 1 of Murakami et al. (1999b). Fifteen equally most parsimonious trees were also produced by the similar but not identical data set used by Murakami et al. (1999b), with all variation both there and here attributable to unresolved polytomies in the clade from Asplenium normale through Camptosorus sibiricus (Fig. 5). The uncertainties of that clade, however, do not affect the positioning of Loxo- scaphe thecifera, which nests within the fully resolved sister clade from As- plenium wrightii through Neottopteris nidus in all 15 equally most parsimo- nious trees. Within that clade, Loxoscaphe is placed within the subclade from Neottopteris australasica through N. nidus by 14 synapomorphies with a boot- AMERICAN ce FERN 2001 JOURNAL QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Editor R. James Hickey Botany Department, Miami University, Oxford, OH 45056 Associate Editors Gerald J. Gastony, Department of Biology, Indiana University, Bloomington, IN 47405-6801 Christopher H. Haufler, Department of Botany, University of Kansas, Lawrence, KS 66045-2106 Robbin C. Moran, New York Botanical Garden, Bronx, NY 10458-5126 James H. Peck, Department of Biology, University of Arkansas—Little Rock, Little Rock, AR 72204 The American Fern Society Council for 2001 BARBARA JOE HOSHIZAKI, 557 N. Westmoreland Ave., Los Angeles, CA 90004- — esident CHRISTOPHER H. HAUFLER, Dept. of Botany, University of Kansas, Lawrence, KS 05 "2016. ce-President W. CARL TAYLOR, 800 W. Wells St., Milwaukee Public Museum, Milwaukee, WI Pat i cretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916-1110. Treasurer DAVID B. LELLINGER, 326 West St. NW., Vienna, VA 22180-4151. Membership Secretary JAMES D. MONTGOMERY, Ecology III, R.D. 1, Box 1795, Berwick, PA 18603-9801. Issues Curator GEORGE YATSKIEVYCH, Missouri Botanical Garden, P.O. Box 299, St. Louis, MO 63166-0299. Journal Editor DAVID B. LELLINGER, U.S. National Herbarium a 166, 0-0 Smithsonian Institution, Washington, DC 2056 Memoir Editor CINDY JOHNSON-GROH, Dept. of Biology, Gustavas Adolphus College, 800 W. College Ave., St. Peter, MN 56082-1 Bulletin Editor American Fern Journal EDITOR R. JAMES HICKEY y Department, mi SRA a? aed: OH 45056 ph. (513) Sn: 6000, e-mail: hickeyrj@ muohio.edu ASSOCIATE EDITORS GERALD J. GASTONY.......... Dept. of Biology, Indiana University, ig asi IN 47405-6801 CHRISTOPHER H. HAUFLER ....Dept. of Botany, University of Kansas, Lawrence, KS 66045-2106 ROBBIN C. MORAN New York Botanical Garden, soa ‘NY 10458-5126 JAMES H. PECK Dept. oe ology, University of Arkansas—Little Rock, 1 S. University Ave., Little Rock, AR 72204 **American Fern seve (ISSN 0002-8444) is an illustrated quarterly devoted to the eae study a ferns. It is owned by the American Fern Society, and published at on West St. NW., Vien VA 22180-4151. Periodicals. seco paid at Vienna, VA, and additional e Claims for missing issues, made 6 months (domestic) to 1D months prs after the 7 ee issue, and orders for back issues should be addressed to Dr. James D. Montg gomery, Ecology I Berwick, PA 18603-9801. ao of address, dues, and applications for membership should be sent to the Membership Secr maaatet inquiries concerning ferns should be addressed to the Secretar ry. ro! members of the American Fern Society (annual dues, $15.00 + $5.00 mailing surcharge beyond AS i membership, $300.00 + $140.00 mailing surcharge beyond U.S.A Back volumes are available for most years as printed issues or on microfiche. Please contact the Back pete Curator for prices and availability. POSTMASTER: ik address changes to AMERICAN FERN JOURNAL, 326 West St. NW., Vienna, VA 22180-4151. FIDDLEHEAD FORUM The editor of the Bulletin of the American Fern Society welcomes contributions from members and bers, including miscellaneous notes, offers to exchange o or purchase materials, personalia, horticultural notes, and reviews of non-technical books on fer: SPORE EXCHANGE Dr. Stephen McDaniel, 1716 Piermont Dr., Haein basse CA 91745-3678, is Director. Spores exchanged and lists of available spores sent on request f / 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 Secretary. er Table of Contents for Volume 91 (A list of articles arranged alphabetically by author) AmBrOSIO, S. T. & N. E DE MELO. New Records of Pteridophytes in the Semi-Arid Region of Brazil DE MELO, N. E (see S. T. AMBROSIO) DoucLas, G. E. (see E. SHEFFIELD) DurRAND, L. Z. & G. GOLDSTEIN. Growth, Leaf Characteristics, and Spore Production in Native, and Invasive Tree Ferns in Hawaii GastTony, G. J. & W. P. JOHNSON. Phylogenetic Placements of Loxoscaphe thecifera (Aspleniaceae) and Actiniopteris radiata (Pteridaceae) Based on Analysis of rbcL Nucleotide Sequences GENSEL, P. G. & C. M. Berry. Early Lyocophye Evolution GILMAN, A. V. (see W. H. WAGNER, JR.) GOLDSTEIN, G. (see L. Z. DURAND) GOMEZ-PIGNATARO, L. D. (see B. PEREZ-GARCiA) GRAUVOGEL-STAMM, L. & B. LUGARDON. The Triassic Lycopsids Pleuromeia and Annalepis: Relationships, Evolution, and Origin HEARNE, S. J. (see E. SHEFFIELD) HICKEY, R. J. (see R. L. SMALL) Hoot, S. B. & W. C. TAyLor. The Utility of Nuclear ITS, a LEAFY Homolog Intron, and Chloroplast atpB-rbcL Spacer Region Data in Phylogenetic Analyses and Species Delimitation in /soétes JOHNSON, W. P. (see G. J. GASTONY) KESSLER, M. (see M. LEHNERT) LaBIAK, P. H. (see J. PRADO) LEHNERT, M., M. MOnniIcH, T. PLEINES, A. SCHMIDT-LEBUHN, AND M. KESSLER. The Relictual Fern Genus L LELLINGER, D. B. & J. PRADO. The Seon of Adiantum gracile in Brazil and Envi- rions LUGARDON, B. (see L. GRAUVOGEL-STAMM) MeENDOozPolypodium vulga- re *Polypodium pectin- atum *Polypodium plumu- la >Polypodium chnoo- des °Polypodium lepido- richum is ng crassi- rongqiaoensis, 130, 134 eee 129, 130, 135 ere gi 129, 4, 139 sp., 129, ag 134-138 ae 127-129, 131-1 a on 203 lorentzii, 199, 203, 206, as reticulatum, 202, 207 pak ae a Arachnioides aristata, 202, 207 Aratrisporites, 107, 125-127, 130, 134 Araucaria, 127 5- lca 202, 206-209 pe manne Set 202, 20 Arthromeris, 224 ce 214, 216 walli- chiana, ; Artschaliphyton, spidium crenulans, 70 . 1 Asplenium, 197-212 antiquum, 202, 207, 208 ralasica, 204, 207, 210 cardiophyl- 7 1G 208 ensiforme 204, 205, 208 laetum, 208 nidus, 207 nor- male, 204, 205, 208, 209 obliquissimum, ie oligophlebium, ae 205, 208 prolon- gatum, 197, 204, 205, 208, 210, 211 pseu- 208 rit , 204, ’ ruprechtii, 207, 209 sarelii, 204, 205, 208 enium, 209 s am- 4, 205, 208 sin ig 204, 205, 208 ae 204, 205, Assamiasporites, 160 Asteroxylon, 75, 76, 82 Athyrium filix-femina, a 182, 184, 202, 204, Azolla, 109 microphylla, 228 oe a Fibs 90 nophytace ep eareen staliforme 75 obscurum, 75 feline , 80, 83 denticulata, Binomial jee Hahei clintoniana X goldi- ana, 36 Bisporangiostrobus harrisii, 144 lec m orientale, : §, 202, 20 B Blottiela pubescens aii ghee 118 Bommeria, 212 volte ae 202-209 Fonniella. Sime 203, Calamospora, 1 Camptosorus, a sibiricus, 202, 204, 207-209 s, 209 Campyloneurum, 225 angustifolium, 214, 216 ol 215 haloes, 202, 206, 207 corn pore 88 chore an en 179, 122.925; 126, 144 ¢ osa, ee 118, 120, 122, 142 per- iodic si a. 90, 93 Cheilanthes, 203, 227 access: 202, 206, 2203, 206 botium, 25, 32 menziesii, 25-27, 29, 31 glauc- . ie = 31 nealii, 25 chamissoi, sons hae net 92, as 105 ohioensis, 18, , % Colpo etioaglon 5, — is, 214, | 28 elit, 214, 216 hem- nitidea, 214, pedunculata, 214, 216 Panevan 203, on per 202, 206, 207 234 sei ea 76, 77 caledonica, 75 cambrensis, 75 pertonii, 75 Cornopteris crenulatoserrulata, 201, 202, 207 Coronatispora ee Crenaticaulis, 7 Cryopres Cage * aes Tips of Selaginella unci 3 SN 224, era 214, 216 has- us, 214, pas eatonii, 7, 207 Cyathea, 201 cooperi, 25 lepifera, 201, 203, 207 mamillata, 229 Cyclosorus opulentus, 202, 207 Cyclostigma, 103 kiltorkense, 92, 101, 118, 132, a 136 134 feat ears 134, 135 Cystopteris fowilis 202, DE MELO, N. F. es S. T. AMBR Deheubarthia Demersatheca, on contigua Dennstaedtia ae Zi arent 202, 207 De A ORIN vs Densopori Deparia ete 201, 202, 20 Sasa of the Sexual ele of Pseudo- sis oe (Polypodiaceae), 214 Ton, 104 Davallia, 197, 208 hip nee 202, 207 10) 13, 21 oe reas 71, 72 linearis, 72 azium esculentum, 202, 204, 205, 207 07 Doryopteris donasibon 202, 206, 207 Douc as, G. E. (see E. SHEFFIELD) Drepanophyeus, 75, 76, 80, 82-84, 90, 92 qujin- gensis, 82, 83 sane 82 aria, tam 220, Staak, PME 11 carthusi- eager 9; 187, 188, mh oo x mickelii, 36 X po , 11 Xtriploidea Dryopteris Soon hyb. nov. a carthusiana AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) x goldiana), a Rare Woodfern Hybrid from are ont, 9 DURAND, L. & G. GOLDSTEIN. Growth, Leaf at te and Spore Production in ative, and Invasive Tree Ferns in Ha- waii, 25 Early Lyocophye Evolution, 7 Ela app atliesryti 70 hybridum, 202, 207 latifol- 71 macahense, 71 macrophyllum, 71 golsiocniaask 70, 71 Elkinsia, 103 Endosporites, 105, 125 Enhancement of Fern Spore Germination and Gametophyte Growth in Artificial Media, Ensivalia deblondii, 79, 80 h Eviostachya, 92 Evolution and Diversification of the Lycopods, 73 Faironella valentula, 79, 81 Flemingites, 103, 104 Frenguellia, 88 GasTONY, G. J. & W. P. JOHNSON. ie Seen ic einen Based on Analysis of rbcL ucleotide ig i 19 eee ' oo M. Berry. Early Lyocophye Evo n, ae ie 88 goldringiae, 88, 89 grier- beige i, 88, 89 GILMAN, A. V. (see W. H. WAGNER, JR.) Chiisads 221, 234 Gleicheniaceae, 21 Gnetum, 168 GOLDSTEIN, G. (see L. Z. DURAND) GOMEz- PIGNATARO, L. D. (SEE B. PEREZ-GARC{A) , L. & B. LuUGARDON. The Tri- s Pleuromeia and Annale- pis: Relationships, Evolution, and Origin, 115 Group of Adiantum gracile in Brazil and Envi- rions, Growth, Leaf Characteristics, and Spore Pro- INDEX TO VOLUME 91 duction in ier and Invasive Tree Ferns in Hawai Gumuia, 75 Gymnocarpium dryopteris, 202, 207 Hamatophyton, 92 asin exes 202, 203, 206-208 89 sagittata, 89 0 —9 S. J. (see E. SHEFFIELD) Hemionitis levyi, 202, 206, 207 tomentosa, 228 HIc KEY, R. J. (see R. L. SMALL) Hicklingia, 75 Histiopteris incisa, 202, Hoot, S. B. & W. C. eel The Utility of Nu- clear ITS, a LEAFY Homolog Intron, and Chloroplast atpB-rbcL Spacer Region Data in Phylogenetic Analyses and Species De- limitation in Isoétes, 166 Hsua, 75, 83 Huia, 75 Huperi, 75, 119, 150, 153, 155-157, 161, 162, 68 atte = 7 tetragona, 157 un compen Sed i 156, 157 wilsonii, 152, Huvenia, 83 Hymenasplenium, 203, 209, 210 cardiophyl- 03-205, 208 cheilosrum, 204, 205, 208 hondoense, 204, 205, 208 eee 204, 205, 208 ee 204 208 riparium, 204, 205, 208 ole ae ey 202, 204, 205, , Sensteias 21 punctata, 202, 207 Isoetaceae, 150 Isoetales, 100 235 Isoetes, 74, 99, 100, 105, 107, 109, 110, 115- 0-55, 57-60 japonica, 166, 169-171 gens a1 42, 44-54, 56, 57, 174 abudevends, 139 m melano. ospora es 72 mon- orcuttii, ; 171, 174 palm , 60, 63, 66, 67 pitotii, 135 precocia, pe 47, 50-53, 55, gers 65, 169-171 rimbachiana, = Ss » LOG, 1 ee iguetro, 132 valida, 169-172 velata, 166, Isoetites, aes 110, pe 10a, 136, 137, 138 bul- lophilla, 120, 131, 137 rolandii, 109, 136 serratus, 109, 136 sp., 137 JOHNSON, W. P. (SEE G. J. GASTONY) langiophyton, 80, 82, 83 akantha, 83 Kaulinia, 215, 220, 224 Kessler, M. (see M. Lehnert) Knorria, 103 Konioria, 75 LABIAK, P. H. (see J. PRADO) Lagenicula Leclercqia, ‘85-89, 93 complexa, 85, 87 236 LEHNERT, M., M. MONNICH, T. PLEINES, A. SCHMIDT-LEBUHN, AND M. KESSLER. The Relictual Fern Genus Loxsomopsis, 13 rope ea D. B. & J. PRADO. The Group of Ad- iantum gracile in Brazil - Envirions, 1 = Es aE =] 3 wn = 5 NS idophyllum 40 Lepdesana white, nh 101,218 Lepidostrobopsis, 140 Le pidotrobus, 140 bohdarowiczii, 132, 140 jen- neyi, 1 Lepisorus a 215, 218 thunbergianus, Leptochils, ri Leptophloeum, pe ea 202, 207 ii Loxomopsis, 1-2 pearcei, 13, 15, 18, 19 leh- pondenge 2 notabilis, 13, 22 costari- yh eceeeeca pei thecifera, 197-202, 204, LUGA ARDON, B. (see L. GRAUVOGEL-STAMM) 150, 151 150, 153-155, 158-161 Section ostachys, 159 Section Carolini- ana, Ae Section Lateristachys, 159 Sec- tion Lycopodiella, 159 154, 157, 159 car i ola, alopecuroides, 152, 5G pendulinis, 154, 157, 159 copie, pet 158, 159 podi 153-156, 158-161, 168 an- pinum, 154, 157 alpinum, 159 annotin- AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) um, 154, 157, 159 casuarinoides, 154, 5 160 donsien 152, 154, 159 fastigiatum, 154, 157, 159 jussiaei, 154, 159 mage icum, 154, ay 159 epi san 152, 154, 15, 1597 scariosu 9 vestitum, 154, 157, 150 vl, ies, 155, 157, 159 wightianum, 154, 159 igticpeticapod ae sh, fe Lycopsi Lycospora, 103, 126 Lyc costrobus scotti, 133 Speen 133 Lycoxylon indicum, 150, 9 Marchantia ae 152 Marsilea quadrifolia, 2 nat Matteucia struthiopteris, 202, 207 Mazocarpon, 104 wagon astrum, 70 crenulans, 70 crenulans f. nulans, 70 crenulans f. glandulosa, 70 ama A. (SEE B. PEREZ-GARCiA) Mertensia flexuosa, 71 scalpturata, 71 Metaxya, 201 yan 202, ig a, 215, 143 ona cael arakii, 202, 203, 206, 207 cia agg nich, M. iS M. Lehnert) Nathorstiana, 105, 118 Ponmetesaiaee age 118 sehr 160 eottopteris, 197, 203, 204, 210 antiqua, 204, nf 5, 207, 210 ep agen 204, 205, 207, nidus, 204, 205, 207, Nephrol crodifolia, 202, i, 205, 207 Records of Pteridophytes in the Semi-Arid egion of Brazil, 227 Niphidium, 225 geet 215, 218, 224 heer 75 aphylla, otholaena rosei, a 206, 207 Odonax borealis, Odontosoria, 197, 208 chinensis, 207, 208 Oleandra ie rary 202, 204, 07 , 209, 211, 212 japonicum, 206-208 lucidum, 199, 200, 203, 206, hioglossum, 71 palmatum, 71 Osmunda, 14, 215 regalis, 185 INDEX TO VOLUME 91 Oxroadia, 105 gracilis, 118 Paesia scaberula, 202, 207 Paraleptochilus, 224 ikdad nina 104 oe 103 rodendron, 100, ae stes inom a Dryopteris clintoniana ae ee. 202, 206, 207 Pence, V. C. Cryopreservation of shoot Tips of Selaginella uncinata, 37 EREZ-GARCIA, B., A. ny se A, R. RBA, & L. D. Phase of eer bradeorum (Polypo ee Phlebodium, 215, 224, Ao araneosum, 224 de- cumanuin, 218, 224 pseudoaureum, 218, ahiisuer 203, 209, 210 scolopendrium, 202, 205, 207 Phylglossum, 74, 119, 151-155, 157, 158, 162 52 tonal ponesiaes nts of Loxoscaphe theci- era ( aioe Nucleotide Sequences, 19 Picea Pid: os 214 edulis, 1 t Pityrogramma, 206 ae a 200, 203, 206, 212 calomelanos var. calomelanos, 228 tri Ped 200, ie ing 2 jogyri eae 184, 185 fe 179, 180, 183 soap 206, 208, 209, 212 microphyllum, 202, 203, 206—208 S, T. (see M. LEHNERT) nea te 214, ig! excavata, 214, 216, 224 normalis, 214, 216 Pleuromeia, 99, 100, 106, 107, 115-117, 119- 127, 130, 133, 141-143 epicharis, 116— 118, 121 hunanensis, 121 jiaochengensis, 119-121 longicaulis, 107 marginulata, 119-121 patriformis, 121 rossica, 107, 117-119, 121, 122, 125, 126 sanxiaensis, 117, 118, 120, 121 sternbergii, 116-122, Pleuromeiales, 100 ‘olypodium, 214, 224, 225 amoemum, 214, 218 chnoodes, 215, 218, 224 dissimile, 224 215, 218 neg repens 228 arc ira 9, 187, 224, 2 Ig eg rete 105, Fe 130, 126. 126, 144 doub- ingeri, 105 237 Polystichum acrostichoides, 187, 190 tripteron, 202, 207 Polytaenium lineatum, 202, 207 Ponce, dine Additions and Corrections to the Pte- ophyte Flora of Northeastern Argenti- Nicuaet . & P. H. Labrak. The Typification, Iden- tity, and Distribution of Cyathea mamil- lata Fée, 22 PRADO, J. (see D. B. LELLINGER) Protobarinophyton Protoleidodendraceae, '85- 90 Protolepidodendrales, 85-90, 100, 150 Protolepidodendron, 86 P : Psuedodrynaria piogern 218, 224 85 aquilinum, 179, 180, ae Peri, 17, aos 209, 212 cretica, 199, 200, 203, 12 fauriei, 202, 206, 207 aia 103 Quercus, 214 Rebuchia, 75 Relictual ye Genus Loxsomopsis, 13 Renalia, 75, 7 huiulatsportes, 159, 160 Retitrilet Review: Flora of Florida, Volume I, Pterido- phytes and Gymnosperms, 230 Rhynia, 75 Rhyniophytina, 7 Riba, R. (SEE B. sul -GARC{A) Rumohra adiantiformis, 202, 204, 205, 207 Sadleria pallida, 202, 207 Sartilmania Sawdonia, 75, As ornata, 78, 79 Sawdoniaceae, SCHMIDT-LEBUHN, A. (see M. LEHNERT Slag, 74, 118, 142, 168 apoda, 152 8 rupestris, 152 ade fg uncinata, 3 elaginellaceae, 89, 124-126, 150, 151 Selaginellites crassicinctus, 126 errulacaulis, 75, 80 Sestrosporites, 160 SHEFFIELD, ts - E. Douc.as, S. J. HEARNE, & J. Enhancement of Fern Spore pneiaer ion and ae Growth in Artificial Media, 1 238 Sigillaria, Siltrobus 133-135, 138 australis, 129, 134 wre J. Hickey. Systematics of the a Andean a Complex, 41 Spaheropers 32, 201 gardneri, 229 horrida, Sphaeropters 201 cooperi, 25-29, 31-33, 202, 07 lepifera, ag i 07 Sphagnan palustre, Sphe Tis pein 202, 207, 208 Psunencerea 100, 105, 106 feistmantelii, 118 Stenochlaena palustris, 202, 207 Sti se60" 21 Stigmaria, 100, 101, 105, 115, 116 rugulosa, 92 Sea tacite 83 langii, 83 Stylites, 110, 138 Synchysidendron, 104 Systematics of the Northern Andean Isoétes Complex, 41 Taeniocrada, Taenitis, 212 betes 202, 206, 207 akhtajanodoxa, 106, Tarella, 75 Ta AYLOR, W. C. The Evolution and Diversifica- OR taria gaudichaudii, 202, 207 PP ot marcinkiewiczae, 130 IER, J. T. Vernal Pho ans and Nutri- ent nar oa amen in Dryopteris inter- Thelypt eris afar taeae 202-205, 207 Thelypteris interrupta, 228 Thrinko Tae che 133-135, 138, 139 belozerovii, 34 fusiformis, 129, 134 bulbosus, 129, 134 convexus, 129, 134 gorskyi, 129, AMERICAN FERN JOURNAL: VOLUME 91 NUMBER 4 (2001) 134 migayi, 129, 134 radiatus, 129, 134, 139 Tree Ferns in Hawaii, 25 Triassic Lycopsids Pleuromeia and Annalepis: ener aig Evolution, and Origin, 115 imerophytina Ei ieee and Distribution of Cy- athea mamillata Fée, 229 Uskiella, 7. Utility of Nalidt ITS, a LEAFY Homolog In- Species Delimitation in Isoétes, 166 Valvisporites, 10 hice ai toynthsis and Nutrient ee n Dryopteris intermedia a. oe 224 flexuosa, 203, 207 WAGNER, JR., W. H. & A. V. GILMAN. Dryopteris