[ND AGS. N. MANCHESTER, THE HECKMAN BINDERY, iNC. AMERICAN FERN JOURNAL Volume 70 1980 PUBLISHED BY THE AMERICAN FERN SOCIETY EDITORS David W. Bierhorst Gerald J. Gastony David B. Lellinger John T. Mickel MERCURY PRESS, ROCKVILLE, MARYLAND 20852 CONTENTS Volume 70, Number 1, Pages 1-32, Issued March 28, 1980 Intersectional Hybrids in Isoétes BRIAN M. BOOM Notes on Some Pleopeltis and Polypodium Species of the Chihuahuan Desert Region TOM WENDT The Deletion of Vittaria graminifolia from the Flora of Florida GERALD J. GASTONY New Taxa and Combinations of Pteridophytes from Chiapas, Mexico ALAN R. SMITH Shorter Notes; A New County Record for Pilularia americana in Texas; Asplenium = gravesii Discovered in Arkansas; Pilularia americana New to Tennessee; New Names for Polypodium chnoodes and P. dissimile; A New Record for Pellaea atropurpurea in Maryland; An Atypical Athyrium from Eastern Tennessee Reviews 4, Volume 70, Number 2, Pages 33-80, Issued June 30, 1980 Equisetum * litorale in Illinois, lowa Minnesota, and Wisconsin JAMES H. PECK Gametophytes of Equisetum diffusum RICHARD L. HAUKE A Double Spore Wall in Macroglossum LUIS D. GOMEZ P. and KERRY S. WALTER Subdivision of the Genus Elaphoglossum JOHN T. MICKEL and LUCIA ATEHORTUA G. Notes on the Natural History of Stylites gemmifera ERIC E. KARRFALT and DALE M. HUNTER Reciprocal Allelopathy Between the Gametophytes of Osmunda cinnamomea and Dryopteris intermedi AYMOND L. PETERSEN and DAVID E. FAIRBROTHERS Shorter Note: Thelypteris torresiana in Venezuela Reviews 38, 68, bho 39 Volume 70, Number 3, Pages 81-112, Issued September 29, 1980 The Distribution and Ecology of Phyllitis scolopendrium in Michigan RICHARD P. FUTYMA 81 Supplemental Notes on Lesser Antillean Pteridophytes GEORGE R. PROCTOR 88 Additions to the Pteridophyte Flora of the Great Plains RALPH E. BROOKS 9] Flavonoid Synthesis and Antheridium Initiation in Dryopteris Gametophytes RAYMOND L. PETERSEN and DAVID E. FAIRBROTHERS 93 Date of Publication of Sodiro’s “Sertula Florae Ecuadorensis”’ DAVID B. LELLINGER 96 Reproductive Biology and Gametophyte Morphology of New World Populations of Acrostichum aureum ROBERT M. LLOYD 99 Shorter Notes: Diplazium japonicum New to Alabama; ths and Ferns; Three Additions to the Pteridophyte Flora of Escambia County, Florida 11 Reviews 92, 98 Volume 70, Number 4, Pages 113-140, Issued December 30, 1980 A Range Extension for Dryopteris filix-mas ERWIN F. EVERT 113 Differential Germination of Fern and Moss Spores in Response to Mercuric Chloride RAYMOND L. PETERSEN and PATRICK C. FRANCIS — 115 Differences in the Apparent Permeability of Spore Walls and ; . Prothallial Cell Walls in Onoclea sensibilis JOHN H. MILLER 11 Allelopathy and Autotoxicity in Three Eastern 74 North American Ferns WILLIAM E. MUNTHER and DAVID E. FAIRBROTHERS = |2 Shorter Notes: Sandstone Rock Crevices, an Exceptional New Habitat for Thelypteris simulata; A Second Alabama Locality for the Hart’s-tongue 8 American Fern Journal : Ee edt a at : Pied aieentos ee AM ERICAN Volume 70 FERN ee J 0 1 R NA L January-March, 1980 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Intersectional Hybrids in Isoétes BRIAN M. BOOM I — a Some Pleopeltis and ce ete Species f the Chihuahuan Desert Regi TOM WENDT 4 The — of Vittaria graminifolia e Flora of Florida GERALD J. GASTONY 12 New Taxa and Combinations of Pteridophytes from Chiapas, Mexico ALAN R. SMITH _ 15 seucien: ree A New ag Record for Pilularia americana as; Asplenium x gravesii Discovered in Arkansas; cats americana New to Tennessee; New Names for Polypodium chnoodes and P. dissimile; A New Record for Pellaea atropurpurea in Maryland; An Atypical Athyrium from Eastern Tennessee Reviews yaah GARDEN LIBRARY The American Fern Society Council for 1980 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, Ohio 45701. President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN 37235. Vice President LESLIE G. HICKOK, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, Va. 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, D.C. 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Berkeley, Calif. 94720 Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, N.Y. 10458. Newsletter Editor American Fern Journal EDIT DAVID B. LELLINGER Smithsonian Institution, Washington, D. C. 20560 ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington, Ind. 47401 JOHN T. MICKEL New York Botanical Garden, Bronx, New York 10458 The ‘‘American Fern Journal’’ (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and published at the Smithsonian Institution, Washington, DC 20560. Secon cians postage paid at Washington. Matter for publication g (made 3-6 months after the date of issue) should be addressed | to the Editor. Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog, Dept. of Biology, George Mason University, Fairfax, Va. 22030. Orders for back issues should be addressed to the Treasure General inquiries concerning ferns should be addressed to oe Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Feri Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or Gai: $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, New York 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter Dr. John T. Mickel, New York Botanical Garden, Bronx, New York Beit fe sisniges - _ agtine! ter . 'Fiddlehead ——_ if Fhog sees _™ oer g le hange or h materials, personalia reviews of scacckail books ¢ on 4s ferne: Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th Street, Seattle, Washington 98115, is Director. Spores exchanged and collection lists sent on request ena are Rearests Cur sll, Me + = no +h ‘. tan in ferns. Botanical books, back i jeiiana of the Journal, and cash or other gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secretary AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) 1 Intersectional Hybrids in Isoetes! BRIAN M. BOOM* Engelmann (1886) and Campbell (1891) have described the simple procedure for germinating /soétes spores in the laboratory to obtain micro- and megaga- metophytes. Perhaps the absence of hybridization studies in this genus can be explained partly by the fact that sexually mature living plants of more than one species were rarely assembled at one time, or that when they were, the primary aim was comparative embryology (La Motte, 1937). The present study was under- taken to test for genetic compatibility among four selected /soétes species in three sections of the genus (sensu Pfeiffer, 1922). All crossing combinations produced progeny, and in at least one species apogamy may occur. MATERIALS AND METHODS Spores for the crosses were obtained from populations of plants as follows: Isoétes (Reticulatae) macrospora Dur.—Monroe Co., TN, Boom 318, Shenan- doah Co., VA, 8 Nov 1978, Evans; I. (Reticulatae) engelmannii A. Br.—Polk Co., TN, Boom 317, Putnam Co., TN, Boom 267; I. (Tuberculatae) flaccida Shuttlew.—Dixie Co., FL, Boom 313, Wakulla Co., FL, Boom 314, 315; 1. (Cris- tatae) riparia Engelm.—Tyrrell Co., NC, Boom 316. Voucher specimens have been deposited in the Herbarium of the University of Tennessee (TENN). All plants were collected during the summer or autumn of 1978, and were grown in the greenhouse for a short time until the crosses were made in mid-January, 1979. The crossing technique was quite simple, yet rigorously controlled. Forty-eight glass vials were filled with about 1 ml of sterilized fine sand and 10 ml of sterilized pond water. Each sporangium was dissected out of the sporophyll base, washed in a sterile water bath, and then teased apart to release the spores into a vial, taking great care to insure that the vials were not contaminated with unwanted spores. Since microsporangia and megasporangia are usually found on the same plant, spores were taken from completely intact sporangia to avoid the possibility of using megagametophytes which already had been fertilized. The crosses among the eight populations were set up in such a way that the megaspores of each population were brought into contact with the microspores of every other population. To test for spore viability and self compatibility, one plant from each population was selfed. To provide controls for the crosses and to test for apogamy, one vial was set aside for each population in which only megaspores were placed. All vials were kept in the greenhouse at about 25° C, where they were exposed to normal ambient light fluctuations, and were not disturbed except for the occasional addition of sterile pond water. *Department of Botany, University of Tennessee, Knoxville, TN 37916. ‘Contribution from the Botanical Laboratory, The University of Tennessee, N. Ser. no. 516. Volume 69, number 4, of the JOURNAL was issued December 31, 1979. AMERICAN FERN JOURNAL: VOLUME 70 (1980) FIGS. 1-2. Sporophytes resulting from a cross between Isoétes flaccida and I. macrospora, two species thought to have had very different evolutionary histories. FIG. 1. Sporophytes in vial at two months after the cross was made. FIG. 2. Hybrid sporophytes * four months. B. M. BOOM: INTERSECTIONAL HYBRIDS IN ISOETES 3 RESULTS AND DISCUSSION About seven weeks after the crosses were made, the first green shoots of young Isoétes sporophytes were observed in a number of the vials (Figs. ] and 2). Although not every individual cross was successful, every possible crossing com- bination was successful within the first two months of the experiment. None of the control vials showed any signs of growth at that time, and the genetic compatibil- ity of the four species presumedly has been demonstrated. No new sporophytes were observed until 3.5 months after the experiment be- gan, when a single sporophyte appeared in the /. macrospora control vial (Boom 318). Since no microspores had been introduced into this vial, either the megagametophyte had somehow become fertilized before it was introduced into the vial or the species is capable of reproducing apogamously. Considering the precautions taken by using only fully intact sporangia as a source of spores, apogamy seems more likely. Eight months after the crosses were made, no sporophytes had developed in any of the control vials of the other three species. Easily hybridized species of Jsoétes means that hybridization followed by polyploidization may be a mode of evolution from time to time in the genus. The occurrence of facultatively apogamous taxa is consistent with such a process. If, as is suggested by the experimental observations, /. macrospora can be apogam- ous, this could help explain the Virginia and Tennessee populations disjunct from the typical northeastern range of this species (Dennis et. al., 1979). The reticulate distal face and the cristate proximal face of the megaspores of J. macrospora suggest a possible hybrid origin for this species. Some Jsoétes populations on the Coastal Plain of the Carolinas have various characters, primarily megaspore ornamentation, which clearly are intermediate between typical /. engelmannii and/. riparia; supposedly these plants are hybrids between the two. Specimens from such populations occasionally have been anno- tated as /. engelmannii var. georgiana Engelm. or var. caroliniana Eaton. The results of this study also support Matthews and Murdy’s (1969) interpreta- tion of the often confusing /soétes populations on the granite outcrops of the Piedmont of the southeastern United States. Introgression apparently is taking place in pools which are ecologically intermediate between the habitats typical of I. piedmontana (Pfeiffer) Reed and those of /. melanospora Engelm. For an alter- nate explanation, see Rury (1978), who suggests that intermediates represent de- velopmental stages of one polymorphic species. The naturalness of Pfeiffer’s (1922) sections of the genus is suspect now more than ever in light of the artificial intersectional hybridizations. The infrageneric classification of Isoétes should be reexamined by means of an extensive genetic, cytogenetic, and phytochemical survey, as well as by using traditional morpholog- ical characters. : This report of intersectional genetic compatibility need not necessarily affect Isoétes taxonomy at the species level, however. In natural circumstances, the taxa generally are isolated by geographic, ecological, or phenological barriers, and they can be distinguished morphologically from one another. The amount of gene va : AMERICAN FERN JOURNAL: VOLUME 70 (1980) flow between the typically isolated populations must be relatively small. If this is not the case, it remains a challenge to explain why selection has not favored the establishment of reproductive barriers between species. The present study was initiated to test the potential for genetic experimentation in /soétes. The preliminary results were very successful and indicate further and wider genetic studies would be beneficial. Such future hybridization research should take advantage of the artificial crossing technique recently described for Selaginella (Webster, 1979). The method appears to be well suited for /soétes crossing with little or no modification, and will permit more critical experimenta- tion than ever could be possible with the non-sterile technique employed in the present study. Appreciation is extended to Dr. A. Murray Evans for critically reviewing this paper. Field work was aided by a Grant-in-Aid of Research from Sigma Xi, The Scientific Research Society. LITERATURE CITED CAMPBELL, D. H. 1891. Contributions to the life-history of Isoétes. Ann. Bot. 5:231-358. DENNIS, W. M., A. M. EVANS, and B. E. WOFFORD. 1979. Disjunct populations of Isoétes macrospora in southeastern Tennessee. Amer. Fern J. 69:97-99. ENGELMANN, G. 1886. The genus Isoétes in North America. Trans. St. Louis Acad. Sci. 4:358—390. LA MOTTE, C. 1937. Morphology and orientation of the embryo of Isoétes. Ann. Bot. 1:695—715. MATTHEWS, J. F. and W. H. MURDY. 1969. A study of Isoétes common to the granite outcrops of he southeastern piedmont, United States. Bot. Gaz. 130:53—61. PFEIFFER, N. E. 1922. Monograph of the Isoétaceae. Ann. Mo. Bot. Gard. 9:79-232. RURY, P. M. 1978. A new and unique, mat-forming Merlin’s-grass (Isoétes) from Georgia. Amer. Fern J. 68:99-108. WEBSTER, T. R. 1979. An artificial crossing technique for Selaginella. Amer. Fern J. 69:9-13. REVIEW “THE ECONOMIC USES AND ASSOCIATED FOLKLORE OF FERNS AND FERN ALLIES,” by Lenore Wile May, Botanical Review 44:491-528. 1979-—As stated by the author, this paper is not taxonomic in nature, but discusses fern folklore and to a lesser extent their economic history. It provides an easily read text for the generalist and a good bibliography for those persons interested in pursuing this topic further. Some of the section titles include: Folklore, Fern Dyes, Fern Fibers, Fern Foods, Medicinal Uses of Ferns, The Male Fern, and The Bracken Fern. The section on medicinal uses occupies forty percent of this article, with related medicinal notes in the folklore portion. The author mentions the following about Ophioglossum vulgatum: ‘‘This plant is called adder’s tongue because out of every leaf it sendith forth a kind of pedestal like an adder’s tongue, it cureth the biting of serpents.’”-—J. Scott Peterson, Dept. of Botany & Plant Pathology, Colorado State University, Ft. Collins, CO 80523. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) 5 Notes on Some Pleopeltis and Polypodium Species of the Chihuahuan Desert Region TOM WENDT” In preparing a treatment of the Polypodiaceae s. str. for the forthcoming Chihuahuan Desert Flora (M. C. Johnston, et al.), I found several taxonomic changes to be necessary. A new variety of Polypodium thyssanolepis A. Br. ex Kl. is described, and Pleopeltis erythrolepis (Weath.) Pic. Ser. is lowered to varietal rank within Pleopeltis polylepis (Roem. ex Kunze) Moore. Material from GH, LL, NY, TEX, and US was consulted in preparation of the treatment. POLYPODIUM Polypodium thyssanolepis A. Br. ex KI. is a species of lithophilic fern which ranges from the southwestern United States to South America and the West Indies. Originally described from Colombian material (Klotzsch, 1847), it was not reported from the United States until 1913, when Maxon noted it among the collections of L. N. Goodding from the Huachuca Mountains of southeastern Arizona. Maxon (1913) stated that Goodding’s specimens were ‘“‘perfectly typical of the species as it exists from Mexico to the Andes and in Jamaica.’’ In his revision of several groups of squamate American polypodies (Maxon, 1916), the most recent revision in which P. thyssanolepis has been treated, he recognized no varieties within the species. However, in recording the species from Texas, Maxon (1923) noted that specimens from both Texas and Arizona “‘have fronds only scantily scaly beneath, in marked contrast to tropical material.”’ A number of characters are correlated with the sparser indument of the fronds of these northern populations. Material from western Texas, southeastern Arizona, northern Coahuila, and parts of Chihuahua appears to represent a strongly marked new variety of P. thyssanolepis (Fig. 1). None of the synonyms of P. thyssanolepis (see Maxon, 1916; Morton, 1973) refer to this new variety, which may be distinguished from the typical variety by the following key: Stipes sparsley scaly, the scales mostly ovate or lance-ovate, suborbicular scales few or none; scaly indument of the lower lami fi tsod to ot th ; venation mostly free, with fewer than 30%%(40%) of the sori within areoles; basal lobes of the lamina distinctly alternate. P. thyssanolepis var. riograndense Stipes densely scaly, the scales mostly suborbicular; scaly indument of the lower lamina surface dense, typically entirely obscuring the surface; venation mostly areolate, with more than 70% (usually nearly 100%) of the sori within areoles; basal lobes of the lamina opposite or subopposite. P. thyssanolepis var. thyssanolepis *Rama de Botanica, Colegio de Postgraduados, Chapingo, Edo. de México, México. ae !Work accomplished at University of Texas at Austin and Gray Herbarium of Harvard University. I thank Alan R. Smith of the University of California at Berkeley for unpublished data. 6 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Polypodium thyssanolepis var. riograndense Wendt, var. nov. A var. thyssanolepide stipitibus multo minus squamatis squamis suborbicularis paucis vel absentibus, laminis minus squamatis, venatione libera pro parte max- ima, et positione lobarum laminarum basilarium distincte alterna recedit. Small lithophilic ferns. Rhizome slender, wide-creeping, 1.5-2.5 mm thick, densely scaly; rhizome scales subulate to lanceolate-acuminate, (1.0) 1.5-3.0 mm long, imbricate, light brown with a darker central stripe composed of dark-walled cells with clear lumina, the margins slightly erose to irregularly ciliate. Fronds distant or occasionally a few somewhat crowded, to 15(20) cm tall, usually much smaller; stipes usually slightly shorter than to slightly longer than the laminae, but varying from one-half to twice as long, sparsely scaly, the scales mostly subulate to lanceolate-acuminate, peltate, to 3 mm long, erose to fimbriate, the larger ones usually brown with a blackish central stripe, these scales often continuing into the lower part of the rachis, with nearly orbicular, irregular, peltate scales scattered or absent; laminae oblong or ovate to triangular-oblong to deltate, to 10 cm long, to 5.5 cm wide, acuminate or acute, deeply pinnatifid into up to 9(11) segments on each side (usually fewer), glabrous above, sparsely to moderately densely scaly below but the scales usually not completely obscuring the green of the surface, the lamina scales peltate, ovate to lanceolate, usually attenuate-acuminate, 0.8- 2.0(3.0) mm long, light reddish-brown, darker at the point of attachment, weather- ing gray, the margin remotely toothed to lacinate, orbicular scales few or none, the lobes distant to fairly close, linear or spatulate to oblong, entire, obtuse to broadly acute, regular or irregular in length on the same frond, perpendicular or slightly ascending relative to rachis, the lowest pair distinctly alternate, venation mostly free, fewer than one-half of the sori (usually many fewer) within areoles; sori roundish, in a single row on each side of midvein of lobe, usually obscured by scales; spores 64 per sporangium. TYPE: Uncommon in crevices of cliffs and boulders in sheltered canyon with Quercus grisea, Juniperus sp., Ungnadia speciosa, Garrya ovata, etc., lower Indian Cave Canyon (side canyon of Dead Horse Canyon), north side of Chinati Mountains, Presidio Co., Texas, 16 Oct 1977, M. L. Butterwick & E. J. Lott 3897 (TEX; isotypes GH, MEXU). Polypodium thyssanolepis has generally been characterized as having areolate venation and pinnatifid fronds (Maxon, 1916). A rare form of the species with bipinnatifid fronds is known to show partial loss of areolation; this form occurs with the normal form in Central America (Maxon, 1916), but is not known from northern Mexico or the United States. It agrees in density of indument and all other characters with typical var. thyssanolepis. On the other hand, the new variety differs strongly and consistently from var. thyssanolepis in venation (Fig. 2). These venation characters are constant regardless of size; occasional speci- mens of var. thyssanolepis from Chihuahua in which the fronds are much reduced (laminae as small as 0.5 cm long), with many fronds nearly entire, nevertheless display the areolate venation (and all other characters) typical of much larger tropical plants of the variety. South American material, including all specimens seen from Colombia (at GH), the type locality of the species, agrees with var. thyssanolepis as here circumscribed. Fronds of var. thyssanolepis may reach much larger sizes (to 60 cm or more) than those of var. riograndense, but this probably is a direct environmental effect. T. WENDT: NOTES ON PLEOPELTIS AND POLYPODIUM 7 Specimens of var. thyssanolepis from northern Mexico are generally much smaller than tropical material; indeed, the reduction found in certain Chihuahuan speci- mens, noted above, is unparalleled in var. riograndense. Ss cea wre ee Pay AS ie ‘ ! FIG. 1. Distribution of Polypodium thyssanolepis varieties in the United States and Mexico excluding Chiapas. Black squares = var. thyssanolepis; open squares = var. riograndense; half-squares = intermediates. The chromosome number of P. thyssanolepis var. thyssanolepis has been re- ported as n=37 from South American material (Evans, 1963) and n=74 from a Jamaican population (Wagner & Wagner, 1975). Polypodium thyssanolepis (vari- 8 AMERICAN FERN JOURNAL: VOLUME 70 (1980) ety unknown) has been reported as n=ca. 72 from material from ‘‘Mexico’’ (Sorsa in Fabbri, 1965; Sorsa, 1966). A limited number of varietal intermediates is found in Chihuahua (Fig. /), but most specimens from this area are easily placed in one of the varieties. Several collections (e.g., Pringle 443) include fairly representative plants of both varieties. Further work, especially chromosomal, may reveal that var. riograndense would be better treated as a separate species. FIGS. 2 and 3. Venation of Polypodium thyssanolepis varieties. FIG. 2. Type of var. riograndense (Butterwick & Lott 3897, TEX); arrows indicate scattered areoles. FIG. 3. Var. t/ lepis from San Luis Potosi, Mexico (Johnston, Chiang & Wendt 12275, LL). Figure I shows the geographical origin of the United States and Mexican speci- mens examined. All Chihuahuan material for the species, along with representa- tive specimens of the new variety, is cited below. Polypodium thyssanolepis var. riograndense: MEXICO: Chihuahua: Mountains between Guadalupe y Calvo and Nabogame, 7200 ft, on large boulders in pine-oak forest, Correll & Gentry 23052, p. p. (GH, LL); 12 mi W of Cuauhtémoc, in crevices of cliff, Correll & Johnston 21610 (LL); Between Yepomera and Babicora, in crevices of boulders in pine-oak-juniper open mountain forest, Correll & Johnston 21624 (LL, US); Vicinity of village of Majalca, in crevices of boulders, Correll & Johnston 21780 (LL); 12 mi W of Cuauhtémoc, steep, rocky (granitic) slope, in pinyon pine-scrub oak association, Gould 8958 (LL); Rio Negro and vicinity, LeSueur 1273 (TEX); Rocky hills near Chihuahua, cold cliffs, Pringle 443, p. p. (GH, NY, US); 16 mi W of Cuauhtémoc, rolling terrain with scattered junipers, pinyons, and oaks, 7200 ft, in rocky i teep slope, Reeder, Reeder, & Soderstrom 3477 (GH). Coahuila: Sierra del Jardin, Canyon Hundido on N side of Pico Centinela, 8 km E of Rancho El Jardin by winding road, 1500-2250 m, steep canyon through igneous sierra, Johnston, Chiang, Wendt & Riskind 11803A (LL). Sonora: Loop of the Rio de Bavispe, S of Aribabi, Sierra de Huépari, 1495 m, Harvey 1706 (US). T. WENDT: NOTES ON PLEOPELTIS AND POLYPODIUM 9 UNITED STATES: Arizona: Cochise Co.: Conservatory Canyon, Huachuca Mts., vga Sept 1882, Lemmon s. n. (GH, NY, US); Chiricahua Mts., Peebles & Loomis 5415 (US). Pima Co.: Baboquivari Mts., Gilman 15 (US). Santa Cruz Co.: Sycamore Canyon, Patagonia Mts., 2800 ft, Ripley & Barneby 822 (NY). Texas: Brewster Co.: Chisos Mts., Boot Spring, 30 June 1932, Mueller s. n. (GH, NY, TEX). Jeff Davis Co.: Near Fort Davis, in clefts and crevices of porphyritic rocks, E. J. Palmer 32196 (TEX, U Polpodium thyssanolepis var. thyssanolepis: XICO: Chihuahua: Minas Nuevas, ca. 8 mi NW of Parral, 6000 ft, Correll & Gentry 22764 (GH, ee ai Ca. 5.5 mi NW of Parral, 5800 ft, Correll & Gentry 22723 (LL); Sierra de Santa Barbara, ca. 4 mi SW of Villa Matamoras, 6300 ft, Correll & Gentry 22802 (NY, LL); Along old railroad W toward Rancho Ojito, Correll & Johnston 21488 (LL, NY); 25 mi SE of Cuauhtémoc, Correll & Johnston 21597 (LL); 11 mi S of Matamoras (Cuevas), 1950-2100 m, Gentry & Arguelles 18037 (LL, US); Majalca (Pilares), 2075 m, Harvey 1463 (GH, US); La Bufa, on Rio Batopilas, Knobloch 578 (US); Canyon E of Hidalgo de Parral, Knobloch 751 (US); Cerocahui-Cuiteco Road, Knobloch 882 (US); Barranca Guerachic, between Agua Blanca and Guerachic, Knobloch 1849 (LL); Rocky hills near Chihuahua, cold cliffs, Pringle 443, p. p. (GH); Potrero Peak, Pringle 977 (NY); Between San Fran- cisco del Oro and Santa Barbara, near Arroyo de Granadefia, ca. 7000 ft, Soderstrom 894 (LL) Intermediates between var. thyssanolepis and var. riograndense: MEXICO: Chihuahua: Small mountain on NE edge of cal Correll 22688 (GH, LL); Mountains just SE of Nabogame, 6000 ft, Correll & Gentry 23033 (LL); Mountains between Guadalupe y Calvo and Nabogame, 7200 ft, Correll & Gentry 23052, p. p. (GH, LL, US); Majalea, Knobloch 329 (GH, US), LeSueur 476 (US). PLEOPELTIS A study of material of Pleopeltis erythrolepis (Weath.) Pic. Ser. and P. polylepis (Roem. ex Kunze) Moore throughout their ranges in Mexico and the United States has led to the conclusion that they represent geographical varieties of the same species. The following new combination is necessary: hers cose polylepis vi var. erythrolepis (Weath.) Wendt, comb. & stat. nov. Polypodium erythrolepis Weath. Contr. Gray Herb. 65:11. 1922. TYPE: Cold ape dea [Pot- rero] Peak, Chihuahua, Pha 825 (GH!; isotypes, GH!, LL!, NY-2 sheets!, US-3 sheets!). Phlebodium erythrolepis (Weath.) Conzatti, Fl. Tax. Mex. 1:95. 1946. Pleopeltis erythrolepis (Weath.) Pic. Ser., Webbia 23:189. 1968. Various characters have been used to distinguish the taxa. Weatherby (1922), in his description of Polypodium erythrolepis, emphasized the long stipe and imbri- cated, fimbriate-ciliate lamina scales of this ‘‘well-distinguished’’ species. How- ever, in his treatment of the ferns of north-central Mexico (Weatherby, 1943) he states: [These new collections] go very far to break down the differences between P. erythrolepis and P. peltatum [Pleopeltis polylepis]. In them, the abundant, ovate, deeply lacerate-margined scales of the former, which seemed so distinctive when it was proposed, nearly disappear and are replaced by suborbicular ones. The: surviv- ing distinctions are: P. erythrolepis, stipe nearly as long as the blade, costa green on the lower alee P. peltatum, stipe conspicuously shorter than the blade, costa black on lower surface. In addition, P. erythrolepis tends to have narrower rhizome-scales with narrower, more definitely erose-serrulate hyaline margins; but this is os a tendency. Furthermore, the collection here cited under P. pel- tatum . . . is also transitional . . . In all probability, P. erythrolepis would best be treated as a vay of P. palates. 10 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Knobloch and Correll (1962) ‘‘more or less avoided the taxonomic issue by recog- nizing as P. peltatum those plants that have a distinctly blackish costa and placing those plants that lack this characteristic into P. erythrolepis,’’ and in doing so recognized both species from Chihuahua. In the present study it was found that both costa color and stipe length are too variable within both taxa to be taxonomically useful. The characters for distin- guishing the varieties are given in the following key: Scales of the lower surface of mature laminae entire or merely erose, mostly orbicular, usually not densely imbricate; rhizome scales ovate to broadly lanceolate P. polylepis var. polylepis Scales of the lower surface of mature laminae fimbriate or ciliate and/or mostly acuminate, usually densely imbricate; rhizome scales lanceolate-acuminate to broadly lanceolate. P. polylepis var. erythrolepis Populations from Sonora, Chihuahua, northern Durango, and Texas fit easily within var. erythrolepis as here circumscribed, and almost all central and southern Mexican populations (southern Durango and Guanajuato south to Oaxaca, also Baja California Sur) are ‘‘good”’ var. polylepis. However, a broad range of inter- gradation between the varieties occurs in the northern Sierra Madre Oriental of southern Coahuila, Nuevo Leon, and San Luis Potosi, where individuals referable to both varieties as well as a preponderance of intermediates occur. Furthermore, northern Coahuilan populations (Muzquiz to Sierra del Carmen) include many intermediate types in additon to those referable to var. erythrolepis. The problem is compounded not only by intrapopulational variation in lamina scales, but also by the fact that scaliness of the fronds changes with age. Young fronds of all populations tend to have many acuminate scales; these are then apparently shed quite early in the case of var. polylepis, but are retained much longer in var. erythrolepis. Mature, preferably fertile fronds therefore are neces- sary for specimen identification. Only one glaring exception to the above-mentioned geographical pattern was found in the material studied. A specimen from the state of Mexico (Parque Nacional de Laguna Zimpoala, Barkley, Webster & Rowell 7420, TEX) is refera- ble by all characters to var. erythrolepis, although it is well south of the range of that variety. There also appears to be some problem in the differentiation of Pleopeltis polylepis var. polylepis from P. macrocarpa var. trichophora (Weath.) Pic. Ser. in central Mexico, particularly in the general area of Mexico City. The taxonomic problems involving P. polylepis and P. macrocarpa are emphasized by the fact that a variety originally described by Weatherby (1944) within P. polylepis (as Polypodium peltatum var. interjectum) appears to belong closer to Pleopeltis ee a (A. R. Smith, pers. comm.). Further studies are much needed in this complex. LITERATURE CITED EVANS, A. M. 1963. New chromosome observations in the Polypodiaceae and Grammitidaceae. Caryologia 16:671-677. T. WENDT: NOTES ON PLEOPELTIS AND POLYPODIUM 11 FABBRI, F. 1965. Secondo supplemento alle tavole cromosomiche delle Pteridophyta di Alberto iarugi. Caryologia 18:675—731. KLOTZSCH, J. F. 1847. Beitrage zu einer Flora der Aequinoctial-Gegenden der neuen Welt. Filices [Pt. 2]. Linnaea 20:337-445. KNOBLOCH, I. W. and D. S. CORRELL. 1962. Ferns and Fern Allies of Chihuahua, Mexico. Texas Research Foundation, Renner. MAXON, W. R. 1913. Some recently described ferns from the Southwest. Amer. Fern J. 3:109-116. —_—_——.. 1916. Studies of tropical American ferns-No. 6. Contr. U. S. Natl. Herb. 17:541-608. —_—. 1923. Notes on American ferns—XIX. Amer. Fern J. 13:73-75. MORTON, C. V. 1973. Studies of fern types, II. Contr. U. S. Natl. Herb. 38:215-281. SORSA, V. 1966. Chromosome studies in the Polypodiaceae. Amer. Fern J. 56:113-119. WAGNER, W. H., Jr. and F. S. WAGNER. 1975. A hybrid polypody from the New World tropics. Fern Gaz. 11:125-135. WEATHERBY, C. A. 1922. The group Polypodium lanceolatum in North America. Contr. Gray Herb. 65:3-14. . 1943. Polypodiaceae. Jn I. M. Johnston. Plants of Coahuila, eastern Chihuahua and adjoin- ing Zacatecas and Durango, I. J. Arnold Arbor. 24:306-339. . 1944. A southern variety of Polypodium peltatum. Amer. Fern J. 34:17-19. REVIEW ‘‘THE EXPERIMENTAL BIOLOGY OF FERNS,” by A. F. Dyer (ed.). Ex- perimental Botany: An International Series of Monographs, vol. 14, 657 pp. 1979. Academic Press, London and New York, ISBN 0-12-226350-2. $79.00—Over the past forty years a significant amount of research has been devoted to the experi- mental biology of pteridophytes. Although information has accumulated and de- velopmental and genetic problems have been better circumscribed, there has been no attempt to organize this into a fashion which would make the relevant ideas and literature easily available to experimental biologists and botanists. This volume attempts to review comprehensively nearly all of the significant studies in fern experimental biology in a context which stresses the controversies currently ex- tant and the problems and avenues of approach which promise to provide the most productive and interesting rewards. In essence, it is a technical introduction to the literature, with over 2000 reference citations. The textual contents reflect accu- rately the current state of knowledge with heavy emphasis on morphogenetic studies of the gametophyte generation. The 16 chapters, contributed by 16 au- thors, detail our knowledge of meiosis; spore initiation, morphogenesis, and ger- mination; structural, physiological, and biochemical aspects of the filamentous gametophytic stage; differentiation from one-dimensional to two-dimensional growth; antheridiogens; sporophyte development; apogamy, genetics, cytogenet- ics, and hybridization; and experimental ecology. Although many of the contribu- tions are excellent, some are superficial, reflecting in my estimation the lack of experimental studies in those fields. However, because of the comprehensive literature surveys and the emphasis on ideas and problems, this volume will be a very valuable source book for experimental biologists and pteridologists for many years to come.—Robert M. Lloyd, Department of Botany, Ohio University, Athens, OH 45701. 12 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) The Deletion of Vittaria graminifolia from the Flora of Florida GERALD J. GASTONY* Interest in new state and national records, range extensions, and the status of rare or endangered species of ferns is probably nowhere more keen than in Florida. Critical field and herbarium work to upgrade our knowledge of the dis- tribution and habitat requirements of ferns, especially in the subtropical southern region of the state, assumes political as well as scientific significance at a time when state and federal socio-economic decisions are influenced by the ecological status of species as humble as a lousewort or a snail darter. Efforts to increase the accuracy of our floristic records for Florida ferns have been quite evident in recent years, for example in the work of Messler (1974), Evans (1975), Ward and Hall (1976), Nauman and Austin (1978), Nauman (1978), Austin et al. (1979), Adams and Tomlinson (1979), and Nauman (1979a, 1979b). Such efforts, however, must include the deletion of erroneous records as well as the addition of new records. The deletion of one such erroneous record, the natural occurrence of Vittaria graminifolia Kaulf. in Collier County, Florida, is the subject of this report. The belief that V. graminifolia occurs in Collier County, Florida, its only re- ported occurrence in the United States, is based on a statement appended to the discussion of V. lineata (L.) J. E. Smith in Wherry’s (1964) Southern Fern Guide. Wherry stated that V. filifolia Fée was found in 1960 in Collier County, and he distinguished it from V. lineata by the weak iridescence and width of its scales. In another context, Tryon (1964a) showed that V. filifolia is an incorrect name for this species and that its correct name is V. graminifolia. In an effort to bring cytological evidence to bear on the identity of the Appala- chian gametophyte by counting its chromosomes and those of V. lineata and V. graminifolia Gastony, 1977), I undertook a search for V. graminifolia in Collier County. Dr. Wherry responded to my request for more information relating to his 1964 report by noting (in litt., 15 Aug 1976) that at the age of ninety and a half he was no longer able to recall more specific locality data or whether an herbarium voucher documented his report. He did recall, however, that he had visited the living fern collection assembled by John Beckner in St. Petersburg, Florida and that Beckner had there ‘‘two Vittarias,”’ one less winter hardy than the other. The less hardy one from Collier County was what Dr. Wherry took to be V. graminifolia (as V. filifolia). Wherry’s information enabled me to contact John Beckner, who agreed to take me to the site from which he had collected the Vittaria in question. In the com- clearly remembered having made the original collection, and we eventually found several specimens of what Beckner said was V. graminifolia if anything in that *Department of Biology, Indiana University, Bloomington, IN 47401. G. J. GASTONY: DELETION OF VITTARIA GRAMINIFOLIA 13 area was. These were immature, somewhat depauperate specimens epiphytic on the trunk of Persea palustris, unlike the usual Sabal palmetto epiphytism of V. lineata. They were not fertile and did not survive cultivation efforts in the Indiana University greenhouses. Beckner was certain, however, that Wherry had taken a specimen of his original collection back to Pennsylvania and that a voucher specimen documenting Wher- ry’s (1964) statement was to be found there. Subsequent inquiries led to a speci- men (number 0925236) at the herbarium of the Academy of Natural Sciences in Philadelphia (PH) bearing the stamp of the herbarium of the University of Pennsylvania, which is on permanent loan to PH. The label identifies the speci- men as Vittaria filifolia Fée and indicates that it had been cultivated from a plant collected by John Beckner west of Deep Lake in Collier County, Florida. The specimen was made on 3 January 1962, and has been annotated by Dr. Wherry as V. filifolia. Beckner (pers. comm.) has since assured me that the locality from which this specimen was taken is identical to the swampy locality we had visited west of Copeland and Deep Lake. On 4 January 1980, I returned to this site and established in a discussion there with Park Ranger Robert Goble that this locality is in the center of what is now the Fakahatchee Strand State Preserve protected by the Department of Natural Resources of the State of Florida. I have analysed Wherry’s specimen from PH, utilizing the characters employed by Tryon (1964b, pp. 212-215) in distinguishing V. lineata and V. graminifolia in the Ferns of Peru. Comparable or identical characters are used by Lellinger (pers. comm.) and by Stolze (pers. comm.) in distinguishing these species in their forth- coming treatments of the ferns of Costa Rica, Panama, and the Chocé and the ferns of Guatemala, respectively. Perhaps the most absolute criterion employed in the discriminatory sets of characters used in these three major floristic treatments is the incidence of tetrahedral-globose, trilete spores in V. graminifolia, as op- posed to reniform, monolete spores in V. lineata. In this regard and in the other characters examined, the specimen upon which the record of V. graminifolia in Florida rests is surely V. lineata. I sent Beckner a photocopy of Wherry’s her- barium specimen and he is certain (pers. comm.) that this specimen is the basis of Wherry’s (1964) report. It is interesting that Lakela and Craighead (1965), Long and Lakela (1971), and Lakela and Long (1976) discussed V. /ineata in their treatments of the ferns of south Florida but made no reference whatever to V. filifolia or V. graminifolia. The reason for omitting V. graminifolia from these three works is unknown and is particularly curious since the work by Long and Lakela (1971) does cite Wherry’s book (1964) as a selected reference on the ferns of Florida. Long is deceased and Craighead (pers. comm.) says that the decision as to what to include in their checklist was entirely that of Lakela. Lakela, now in retirement, does not recall the reason for this omission from any of these works (pers. comm.). There is no indication that any of these authors ever consulted the specimen at PH. Based on my experience with Beckner in revisiting the collection site of Wherry’s specimen in the Fakahatchee Strand, I suspect that the difference in the cold-hardiness of the ‘‘two Vittarias’’ was most likely due to the sub-optimal substrate of the 14 AMERICAN FERN JOURNAL: VOLUME 70 (1980) hardwood host tree and perhaps to a variant allelic constitution correlated with the occurrence of these individuals on this unusual host. It is always possible that V. graminifolia or any other species common in tropical America may be carried into southern Florida by hurricane winds or other means of dispersal and that such adventives may become temporarily or perma- nently established in subtropical Florida. Because of Wherry’s report, Austin and Nauman (pers. comm.) have searched extensively for V. graminifolia in the Fakahatchee Strand but have never found it and have concluded that it is not there. Critical examination of the morphology and ecology of the specimen dis- cussed above indicates that there is no longer any reason to believe that V. graminifolia ever did or does now occur in Florida. It should therefore be deleted from the floristic record for Florida and thus from the flora of the United States. I am grateful to Ross and Priscilla Stanley of Port Charlotte, Florida for their hospitality during the field work and manuscript preparation for this paper and to John Beckner for his aid in the field and in locating the specimen upon which Dr. Wherry’s report was based. I thank Dr. Michael Madison for help in the field, Dr. Wherry for help in interpreting his report, the officers of the herbarium of the Academy of Natural Sciences (PH) for the loan of the specimen discussed, and Ranger Robert Gobel for aid in interpreting the specimen locality data in terms of the Fakahatchee Strand. LITERATURE CITED ADAMS, D. C. and P. B. TOMLINSON. 1979. Acrostichum in Florida. Amer. Fern J. 69:42—46. AUSTIN, D. F., G. B. IVERSON, and C. E. NAUMAN. 1979. A tropical fern grotto in Broward County, Florida. Amer. Fern J. 69:14—16. EVANS, A. M. 1975. Cheilanthes in Florida. Amer. Fern J. 65:1-2. GASTONY, G. J. 1977. Chromosomes of the independently reproducing Appalachian gametophyte: A new source of taxonomic evidence. Syst. Bot. 2:43—48. LAKELA, O. and F. C. CRAIGHEAD. 1965. Annotated Checklist of the Vascular Plants of Collier, Dade and Monroe Counties, Florida. Fairchild Tropical Garden and the Univ. of Miami Press, Coral Gables, Florida. ; , and R. W. LONG. 1976. Ferns of Florida. An Illustrated Manual and Identification Guide. Banyan Books, Miami. LONG, R. W. and O. LAKELA. 1971. A Flora of Tropical Florida. A Manual of the Seed Plants and Ferns of Southern Peninsular Florida. Univ. of Miami Press, Coral Gables, Florida. Ce i 1974. The natural history of Ophioglossum palmatum in south Florida. Amer. Fern 33-39. NAUMAN, C. E. 1978. A checklist of the ferns an d primitive vascular plants of southeastern Florida. Castanea 43:155-162 I Amer. Fern J. 68:65-66. TRYON : = M. 1964a. Taxonomic fern notes. IV. Some American vittarioid ferns. Rhodora 66:110— —————. 1964b. The ferns of Pery Polypodi (D iti Ol 1 ) Meesiki ins ns ace Vieandreae). Contr. Gray Herb. WARD, D. B. and D. W. HALL. 1976. Re- introduction of Marsilea vestita into Florida. Amer. Fern J. 66:113-115. WHERRY, E. T. 1964. The Southern Fern Guide. Doubleday, Garden City, N.Y. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) 15 New Taxa and Combinations of Pteridophytes from Chiapas, Mexico ALAN R. SMITH* This is the second and, it is intended, final report on new taxa and new combina- tions (beginning on p. 26) of pteridophytes in the state of Chiapas, Mexico. For the first report, see Proc. Calif. Acad. Sci., Ser. IV, 40:209-230. 1975. All of the taxa are to be included in the treatment of the pteridophytes of Chiapas, a part of the ‘*Flora of Chiapas’’ project headed by Dennis Breedlove, California Academy of Sciences. English descriptions, fuller synonymies, and additional discussion will be found in the floristic account. Certain of the new names are also needed by Robert Stolze for his forthcoming treatment of the Polypodiaceae in the *‘Ferns and Fern Allies of Guatemala.”’ I thank Colleen Sudekum for preparation of the illustrations. Scanning electron micrographs of spores were made with a Coates and Welter 50 microscope, ob- tained by the Electron Microscope Laboratory at the University of California, Berkeley, under a grant from the National Science Foundation (GB-38359). Breed- love collections were made with the help of National Science Foundation grants GS-383, GS-1183, and GB-29483. I am grateful to A. M. Evans, who has collabo- rated in the description of Polypodium chiapense. yyy Asplenium insolitum A. R. Smith, sp. nov. Figs. 1-2. Rhizomata suberecta, caudices ca. 1 cm diametro; frondes 35-45 cm longae, stipitibus laminas fere aequantibus; stipites brunnei vel griseo-brunnei, non lus- trati, ca. 1.5 mm diametro, adaxialiter viridi-alati, glabri, basi paleis paucis atro- brunneis ovatis; paleae ca. 2 mm longae, obscure clathratae, parietibus crassis et luminibus congestis parvulis; laminae ovato-lanceolatae, 18-25 cm longae, bipin- natae, apice attenuatae sed nec flagelliformes nec proliferae; rhachides adaxialiter viridi-alatae, abaxialiter brunneolae, epaleatae; pinnae 15—20-jugae, usque 6 cm longae, 2.5 cm latae; pinnae infimae (1 vel 2 paria) aliquantum reductae, deflexae; pinnulae usque 8-jugae per pinnam, saepe lobo acroscopico, aliter dentatae vel bidentatae secus marginem, basi cuneatae, plerumque inaequilaterae (sub- dimidiatae), latere basiscopico exciso; pagina laminae et axes subter glabri vel pilis minutis (0.1 mm longis) adpressis capitatis; venae pinnatae, usque 4-jugae per pinnulam; sori 1.0—2.5 mm longi, indusia tenui albido. “TYPE: Terrestrial in montane rain forest, 11 km NW of junction of road to Motozintla along road to El Porvenir and Siltepec, southwest side of Cerro Mozotal, Munic. Motozintla de Mendoza, Chiapas, Mexico, 2100 m, 21 Nov 1976, Breedlove 41653 (DS). PARATYPE: Same locality, 27 Jun 1972, Breedlove 25760 (DS). 3, Asplenium insolitum has no obvious close relatives. In dissection, it is similar to some of the more divided members of the A. radicans complex (e.g., A. flabel- lulatum Kunze var. partitum Klotzsch), but the blade apex is neither flagelliform nor budding and the stipes and rachises are not shining. A closer relative is perhaps A. cuneatum Lam., but that species differs in the flabellate venation of the segments, longer sori, and obviously clathrate scales. Another possible rela- Berkeley, CA 94720. *University Herbarium, Department of Botany, University of California, AMERICAN FERN JOURNAL: VOLUME 70 (1980) | 5¢m FIGS. 1-4. New Asplenium species. FIGS. 1-2. Type of A. insolitum, habit and lower pinna. Breed- love 41653 (DS). FIGS. 3-4 . Paratype of A. sphaerosporum, habit and lower pinna, Breedlove 32512 (DS). Line scale for habit drawings. Jo3¢ A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 17 often a few minute clathrate scales in the pinna axils. Asplenium sphaerosporum A. R. Smith, sp. nov. Figs. 3-6. Rhizomata erecta; frondes plerumque 45-70 cm longae, usque 18 cm latae; stipites atri, ca. 2 mm diametro, glabri, longitudine 0.5-0.75 partes laminarum aequantes; rhachides atrae vel virides, distaliter viridi-alatae; laminae lanceolatae, ad apicem acuminatae; pinnae patentes, vulgo 25 vel plus, ad apicem acuminatae, paribus infimis plene bipinnatis, distaliter pinnis pinnatisectis, denique pinnis serrato-incisis; segmenta obovata, usque 1.3 cm longa, 5 mm lata, basi cuneata, apice dentata vel denticulata (dentibus 2-7), usque 12 paribus per pinnam, prope apices pinnarum segmentis adnatis et decurrentibus; paginae laminarum at- rovirides vel aeruginosae, crassae, glabrae; sori usque 2—3 per segmentum; sporae grandes, globosae (interdum ellipsoideae), 32 per sporangium. “TYPE: SE side of Cerro Baul (16 km NW of Rizo de Oro), Chiapas, Mexico, Breedlove 21805 with Smith (DS). tive is A. solmsii Baker ex Hemsl., but that species has tripinnatifid blades and S te oe G-— < 6) FIGS. 5-6. Scanning electron micrographs of spores of Asplenium sphaerosporum, Breedlove 28733 (DS), x 1500. PARATYPES: MEXICO: Chiapas: Lagos de Montebello, Breedlove 22303 with Smith (DS); Jct. of T: anaté River and river from Yochib, paraje Mahben Chauk, Munic. Tenejapa, Breedlove 6368 (DS, US); SE of Cerro Baul (16 km NW of Rizo de Oro), Breedlove 21810 with Smith and 31338 with Smith (both DS); 7 km NE of Bochil along road to Simojovel, Munic. Bochil, Breedlove 28723, 28733 (DS), Breedlove 32311 with Smith (DS); Near summit of Chuchil Ton, NE of Bochil, Munic. San Andres Larrainzar, Breedlove 29244 (DS); 7 km NE of Jitotol—-Pichucalco jct. on road from Bochil to Simojovel, Munic. El Bosque, Breedlove 32512 with Smith (DS); 6.5 km N of Jitotol, Munic. Jitotol, Breedlove 32758 with Smith (DS); Ocotepec, Rovirosa 1049 (PH); 20 km S of Ocozocoautla, Munic. Ocozocoautla - Espinosa, Breedlove 29138 p. p. (DS), Miinch s.n. (DS); Ghiesbreght 404 (K, NY, PH). Lenaenden Techolo, Sanchez 6 (UC, US); along Camino Real, near Jalapa, Weatherwax 176 (UC). —_ sarin Schnee s.n. (K); Bourgeau 2365 (K). GUATEMALA: Fuego ?, Salvin & Godman 368 (K); Heyde Lux [Donnell-Smith 3231] (K, US). 18 AMERICAN FERN JOURNAL: VOLUME 70 (1980) a y a Ly BLOND aN BS it ep PN \ en oe. SAN Sars a oem FIGS. 7-10. New Cheilanthes taxa. FIGS. 7-8. Type of C. complanata, habit and base of lowermost pinna, Breedlove 41747 (DS). FIGS. 9-10. Type of C. — var. fimbriata, habit and pinnule, Breedlove 39018 (DS). Line scale for habit drawings A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 19 This species resembles somewhat A. achilleifolium, but I do not think that they are closely related. The affinity seems closest to A. monodon Liebm. and A. cuspidatum Lam. The former has large, globose spores, 32 per sporangium, like those of A. sphaerosporum (Figs. 5-6); all specimens of A. cuspidatum that I have looked at have small, reniform spores, 64 per sporangium. It is possible that A. sphaerosporum arose through hybridization between some member of the A. auritum group and A. cuspidatum. Alternatively, it could have speciated from A. monodon. Additional studies are needed to understand the evolutionary relation- ships within this complex group. Asplenium auritum Swartz has often been applied in a broad sense, encompas- sing plants that are simply pinnate (sometimes with a basal auricle) to fully bipin- nate. I would restrict the application of the name to those plants of the complex that are simply pinnate; such plants also have tan, reniform, relatively small spores, 64 sporangium. In Chiapas (and apparently elsewhere in the range), A. auritum s. s. occurs only at low elevations, 200-500 m. Asplenium sphaerosporum occurs at higher elevations—(900)1250-2700 m—than any other member of the A. auritum complex in Chiapas, at elevations where A. cuspidatum can occur. The latter is chiefly from Montane Rain Forests, while A. sphaerosporum is most common in Pine-Oak-Liquidambar Forests. ? 084 Cheilanthes complanata A. R. Smith, sp. nov. Figs. 7-8. 701 . Differt a C. hirsuta Link paleis rhizomatis distincte bicoloris, ad marginem cinnamomeis, ad medium nigrescentibus; laminis pentagonis, latitudine lon- gitudinem fere aequantibus, planis, segmentis ultimis non pendulis; segmentis ultimis obovatis vel anguste ellipticis, 2—4-plo longioribus quam latioribus; laminis atroviridibus, utrinque glabris; indusiis membranaceis, non valde dissimilibus laminae, integris (sine glandibus vel pilis), non vel leviter ad axe decurrentibus. TYPE: North and west slope of Cerro Mozotal below microwave tower along road from Huixtla to El Porvenir and Siltepec, Munic. Motozintla de Mendoza, Chiapas, Mexico, 3000 m, Breedlove 41747 (DS). PARATYPE: Same locality, Breedlove 40335 (DS). The best characters for separating C. complanata from its nearest relatives, i marginata H.B.K., C. chaerophylla (Mart. & Gal.) Kunze, and C : hirsuta (synonym of C. pyramidalis Fée), are the planar blades and the thin, entire indusia that completely lack trichomes or marginal papillae. Another relative may be C. cuneata Link, but that species has black or nearly black stipes and rachises, larger blades, darker, concolorous rhizome scales, and more sharply differentiated in- dusia. Cheilanthes microphylla var. fimbriata A. R. Smith, var. nov. Figs. 2-18, Differt a var. microphyllo indusiis fimbriatis pilis usque 5 mm longis; trichomatibus numerosioribus albidis prope margines laminae supra; et laminis generaliter parvioribus bipinnatis subdimorphis. “TYPE: Along road to Ciudad Cuauhtemoc 6-8 km E of Frontera Comalapa, Munic. Frontera Comalapa, Chiapas, Mexico, Breedlove 39018 (DS). 20 AMERICAN FERN JOURNAL: VOLUME 70 (1980) || RE pepe ree EY ee eI cee Seti srs S| eee pentnse a eee EEE EE I sea 5 Sa YP | semeas Se Baa se gas ee se FIGS. 11-12. Type of Polypodium alavae, habit and base of pinna, Alava 1287 (UC). Line scale for habit drawings. 314! Te A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 21 PARATYPES: GUATEMALA: Petén: Lake Petén Itza, NW of San Andres, Contreras 3571 (US); Lake Petén Itza, along shore W of San Andres, Lundell 17251 (US). MEXICO: Chiapas: E] Carmen, Miinch 184 (DS); without locality, Miinch (DS); Same locality as type, Breedlove 26976 (DS); Munic. Ocozocoautla de Espinosa, Rio de la Venta at the Chorreadero near Derna, Breedlove 36556 (DS); Munic. Ocozocoautla de Espinosa, 13-18 km S of Ocozocoautla, Breedlove 37838 (DS); Villa de Yajalén, Rovirosa 971 p. p. (PH). Tamaulipas: 2 mi S of Tres Palos and 1 mi down road to Loreto, Johnston 4884 (TEX). Yucatan: San Anselmo, Gaumer 1238bis (US); Izamal, Gaumer 534 (UC, US), Gaumer 1409 (US); Chichan- kanab, Gaumer 1473 (US), P. Valdez 65 (US), Gaumer 533 (US); Ruins of Nojpat, Schott 686 (US); Mérida, Schott 135 (US). Variety microphylla is known from the Antilles, southeastern United States, and eastern and southern Mexico. I have not seen collections from the Yucatan peninsula, where var. fimbriata is common. In Chiapas, var. fimbriata seems to be more common than the type variety and does not grow with it. Diplazium drepanolobium A. R. Smith, sp. nov. Differt a D. lonchophyllo Kunze segmentis pinnarum magis obliquis et fal- catioribus; frondibus plerumque grandioribus, pinnis vulgo 20-25 cm longis, 3.5- 6.0 cm latis. -TYPE: 10 km above Rayon, Chiapas, Mexico, Breedlove 26122 (DS). PARATYPES: MEXICO: Chiapas: 26-28 km N of Ocozocoautla, Breedlove 22451 with Smith (DS); 45 km N of Ocozocoautla, Breedlove 20760, 32852 (DS); 2-4 km below Ixhuatan, Breedlove 24163 (DS); 32 km N of Ocozocoautla, Breedlove 38162 (DS); 46 km N of Ocozocoautla, Breedlove 38676 (DS); Without precise locality, Ghiesbreght 361 (K); Arroyo de Ona, cerca Ixtacomitan, Rovirosa 59 (K, PH). Veracruz: Schaffner 470 (P). It is possible that this is only an extreme variant of D. lonchophyllum Kunze, but D. drepanolobium has different blade dissection and seems at least as distinct as some other segregates of D. lonchophyllum (e.g., D. prominulum Maxon and D. subsilvaticum Christ). Relatively few specimens are intermediate between D. drepanolobium and D. lonchophyllum; one such specimen is Breedlove 21624-A (Chiapas, 13 km N of Berriozabal). This collection has abortive spores and may be a hybrid. More numerous are the intermediates between D. drepanolobium and a third species of the complex, D. franconis Liebm., which includes the following synonyms: D. camptocarpon Fée, D. hahnii (Fourn.) C. Chr., and D. donnell- smithii Christ. Several such intermediates from Chiapas are 22293, 22484, and 27480 (all DS). The best course in this difficult group seems to be to recognize all three species. The alternative is to recognize a single species, a treatment that has little merit. On the basis of certain collections identified by Fée (e.g., Schaffner 470), for- ‘merly I applied the name D. acutale Fée to this species. However, the type of D. acutale is referable to D. lonchophyllum. Figs. 11-12. : i : : ; : : est Rhizomata repentia, 5-8 mm diametro, paleis aurantiaco-fulvis, aliquan , patentibus, concoloribus, glabris, ca. 5-10 mm longis; ——. ne ; ee - oe gae; stipites straminei vel brunneoli, parce pubescentes, longitu JU. partes laminae aequantes, (1)3—5 mm diametro; 22 AMERICAN FERN JOURNAL: VOLUME 70 (1980) 5cm ‘e" 15 eat 13-17. New Polypodium species. FIGS. 13-15. Ty and rhizome, Breedlove 37155 (DS). FIGS. 16-17. Bebdins 27453 (DS). Line scale for habit drawi ings pe of P. fuscopetiolatum, habit, portion of Type of P. chiapense, habit and segment, A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 23 (7)14-30 cm latae, basi latissimae, pinnatae praeter apicem, apice Rep Te vel subconformi, confluenti; rhachides moderate vel conspicue pubescentes, sine 16 cm lo 7-10 m paleis; pinnae plerumque 7— ongae, 7-10 mm latae, integrae vel leviter crenulatae, ad apicem acutae, a rotundatae, Grecian parallels, pinnis ses- silibus dimidio proximali laminae, adnatis dimidio distali lam , Nunquam di- latatis; venae ie pean liberae; paginae laminarum — ne : aarviees vel olivaceae, subcoriaceae, utrinque glabrae vel sparse pubescentes; costae ven- aeque infra pilis onan is vel numerosis, laxis, septatis, 0.4-0.8 mm longis; sori rotundati, mediales vel ee vulgo ca. 2mm diametro; sporangia pilis 0 3 mm longis, laxis, ca. 6 pilis per capsulam “TYPE: About 3 mi from race ibe fecal from Sols to Chanal, Chiapas, Mexico, Alava 1287 (UC; isotype DS). PARATYPES: HONDURAS: Francisco Morazan: Km 24 on Tegucigalpa—Zamorano Rd., Kimnach 449 (UC). MEXICO: Chiapas: Lagos de Montebello, Breedlove 22337 with Smith and 32126 with Smith (both DS); same locality, Breedlove 37005 (DS). Polypodium alavae is related to P. adelphum Maxon, from which it differs in the narrower, more parallel-sided, mostly entire pinnae, long-hairy sporangia, and more coriaceous, dark-green leaf tissue. It is also related to P. puberulum Schlecht. & Cham., but it lacks the densely hairy leaf tissue on both surfaces and the lower pinnae are constricted at the base. B35 ghd as ane her igh on & Smith, sp. nov. Figs. 16-17. entia, usque 10 cm longa, 3-5 mm diametro; paleae rhizcmiatis’ cee e pasi fate rants dilatatae, renee acuminatae, incon- spicue como atae, a inem interdum denticulatae; frondes e sae, non r 12-40 cm longae, distantes (ca. 7-10 mm), paid 1-2(3) per plantam; stipites longitudine 0.1-0.2 partes laminae aequantes, 0.5-1.0 mm diametro, stra minei vel longae, 2.5— latae, anguste ovatae vel lanceolatae, basi subtruncatae vel truncatae; ewan sine paleis, utrinque laxe villosae pilis 1-2 mm longis; seg- menta 13—40 mm longa, 3—6 mm lata, patentia, lineari-lanceolata, apice obtusa vel subacuta, basi Satie ‘dilatata, integra vel versus apicem obscure undulata; venae l-furcatae, liberae; paginae laminarum dilute virides, pilis numerosis, 1-2 mm longis, laxis, argenteis, secus costas ene costae decurrentes ad rhachim; sori rotu ndati, inframediales; sporangia glabra “TYPE: Selva del Ocote, 32 km NW of Ocozocoauitla, Munic. Ocozocoautla de Espinosa, Chiapas, Mexico, Breedlove 27453 (DS). PARATYPES: MEXICO: Chiapas: Finca Mexiquito, Purpus 6754, p. p. (UC, US); US); 13 km N of Berriozabal, Breedlove 20281, 31231 (both DS); Same locality, Smith, 31537 with Smith (both DS). Polypodium chiapense is similar to P. hygrometricum stramineous to tan stipes, these villose with soft, silvery, lax, mm long; rachises stramineous with similar hairs 1-2 mm long; : perpendicular to the rachis; leaf tissue with numerous, lax, silvery, septate airs 1-2 mm long, these densest along the costa, and in the glabrous sporangia. Finca Irlanda, Purpus 7225 (UC, Breedlove 21671 with Splitg., but differs in the septate hairs ca. | 4 at 24 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Polypodium fuscopetiolatum A. R. Smith, sp. nov. Figs. 13- 15. Rhizomata repentia, 3-6 mm diametro, brunneola, paleis arte adpressis, lan- ceolatis; paleae fulvae, in medi vittatae, margine m plerumque castaneae, sine paleis; pinnae lanceolatae, apice acutae ve acuminatae, plerumque 9-15 mm latae; venae 3- vel 4-furcatae, anastomosantes, areolis 1-seriatis in quoque latere costae; pagina laminae utrinque pubescens vel glabrescens vel fere glabra; venae costaeque pubescentes; sori rotundi, ca. 1.0— 1.5 mm diametro, mediales vel inframediales; sporangia glabra vel setosa, setis minus quam 0.1 mm longis VTYPE: 6-8 km WNW of Soyalo, Chiapas, Mexico, Breedlove 37155 (DS). - PARATYPES: EL SALVADOR: San Salvador: Santa Tecla, Jaurequi 67 (UC); Same locality, Garcia 22 (UC). GUATEMALA: Santa Rosa: Jumaytepeque, Heyde & Lux [Donn. Smith 4090] (US). Solola: Near Pueblo San Jorge, Hatch & Wilson 317 (US). Suchitepequez: Cuyotenango, Rojas 146 (US). Guatemala: Sapoti barranca, Hayes (US); Pinala, Hayes (US); San Gerénimo, Salvin & Godman (K). MEXICO: Chiapas: Salto de Agua, Escuintla, Matuda 18397 (MEXU, DS); Cacaluta, Escuintla, Matuda 17005 (MEXU, DS); Finca Mexiquito, Purpus 6755, 6858 (UC, US); Huitla, Purpus 7223 (UC); El Sumidero, 22 km N of Tuxtla Guti Breedlove (all with Smith) 21584, 21593, 21595 (DS); 3 km N of Ocozocoautla, Breedlove 21923 with Smith (DS); 5-6 km W of Teopisca, Breedlove 22867 (DS); NW side of Cerro Vernal, 25-30 km SE of Tonala, Breedlove 25616 (DS); 6-8 km E of Frontera Comalapa, Breedlove 26978, 39083 (DS); 27 km NE of Huixtla, Breedlove 28612, 28653 (DS); 65 km S Hwy. 190 on rd. to Nuevo Concordia, Breedlove 37784 (DS); Cerro Vernal, 21 km S of Tonala, Breedlove 38135 (DS). Guerrero: 25 mi S of Chilpancingo, Storer 11] (US); Dist. Mina, Manchon, Hinton et al. 9451 (K); Dist..Montes de Oca, Vallecitos, Hinton et al. 11390 (K). Jalisco: Sierra del Halo, 7 mi SSW of Tecalitlan, McVaugh 16189 (US). México: Dist. Temascaltepec, Rinc6én del Car- Ps Hinton 1737 (K, UC, US). Michoacan: Dist. Coalcoman, Pto. Zarzamora, Hinton et al. 12249 (K). This appears to be one of the more common species of Polypodium at lower elevations (500-1350 m in Chiapas) on the Pacific slope of southern Mexico and northern Central America. Polypodium fuscopetiolatum is closely related to P. hispidulum Bartlett, but it can be distinguished by the narrower rhizome scales that are denticulate or papillate on the margin and filiform at the tip and also by the stipes and rachises usually castaneous and shining. There is also a resemblance to P. plesiosorum Kunze, which has broader, ovate rhizome scales and grows at higher elevations. Specimens of P. fuscopetiolatum have often been identified as P. plesiosorum in herbaria. Several Costa Rican collections are close to P. fuscopetiolatum but differ in having smaller, thicker-textured fronds, broader, squared sinuses, and broader and lighter-colored rhizome scales. I am not certain that they are conspecific with P. fuscopetiolatum, but they appear to be closely related: Pittier 904, Tonduz mine Standley 41904, Tonduz 8804, Mickel 2394, Stork 2987, and Allen 547 (all A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 25 Co mickelii A. R. Smith, sp. nov. 7st Rhiz erecta, a ae 1.5-3.0 cm diametro; frondes 45-110 cm longae; stipites aes 0.6—1. 0 partes laminarum aequantes, basi paleis ovatis mar- gine erosis usque rae ae usque 15 mm longis et 6 mm latis, bicoloribus, paleis fulvis in isa a guste brunneo-vittatis vel a ge oncoloribus et uni- formiter fulvis; lam ovato-attenuatae, 12-35 cm latae; rhachides ae pro- liferae, Slabieeseutes Gel paleis capillaceis praesertim basi pinnarum; pinnae pin- natae, pinnulae lobo parvo deltoideo acroscopico, aliter integrae, vu in frondis grinds pinnulis crenulatis vel grosse dentatis et lobo fere libero elliptico acro- scopico; costulae infra glabrescentes vel basi cee fibrillosis; pagina laminarum et venarum infra plus minusve oe Soria: supra nitida, atroviridis; sori parvi, ut videtur exindusiati, non con -TYPE: NE slope of Cerro Zhitibodbeepetl trail from La Candelaria to Zacatepec, Dist. Mixes, Oaxaca, Mexico, Mickel 4836 with Leonard (NY). PARATYPES: GUATEMALA: San Marcos: Near Aldea Fraternidad, between San Rafael Pié de la Cuesta and Palo Gordo, Williams et al. 26103 (F), 26299 (F, US). Above Finca El Porvenir, up Cerro de Mono, Volcan ree deguens as Steyermark 37397 (F). HONDURAS: Intibuca: Quebrada del Pelon de Guise, Molina R. 6375 (F). MEXICO: Chiapas: SE side of Cerro Tres Picos, Breedlove 25379, 34385 (DS); Without precise locality, Ghiesbreght 401 (YU). Oaxaca: Type locality, mn 4823 with Leonard (NY); Dist. Mixes, vicinity of Zacatepec, Mickel 1565 (NY); Dist. Choapan, Lovani to river toward La Selva, Hallberg 1577 (NY); Dist. Tuxtepec, 24km S of Valle = et “Mickel 5929 (NY); Dist. Ixtlan, 5km S of Vista Hermosa, Mickel 7189 (NY, UC); Dist. Ixtlan, 29 km S of Valle Nacional, Mickel 6363 (NY, UC). Veracruz: Munic. Yecuatla, El Haya, Ventura A. 3431 (NY); Munic. Yecuatla, El Cajon, Ventura A. 4812 (NY : In Oaxaca and Veracruz this species grows at rather low elevations, 450-1450 m; the Chiapas collections were made at 2100-2500 m. At the lower elevations, the only other Polystichums encountered in Mexico are P. platyphyllum (Willd.) Presl and occasionally P. muricatum (L.) Fée. Polystichum mickelii appears to be without close relatives in Mexico and Central America. It is possibly of the group of P. platyphyllum, as indicated by the exindusiate sori. It resembles somewhat species from southern Brazil, e.g., P. montevidense (Spreng.) Rosenst. Selaginella chiapensis A. R. Smith, sp. nov Species heterophylla ex affinitate S. idiosporae Alston, sed foliis intermediis Ovatis, acumnens (nec aristatis), foliis argenteis subtus, rhizophoris seacilnadiean .2-0.4 mm dia iffert. “TYPE: 18-20 ‘eon N of Ocozocoautla, Munic. Ocozocoautla de Espinosa, Chiapas, Mexico, 800 m, Breedlove 28159 (DS). PARATYPE: Above Rancho San Luis, ca. 2 mi N of Ocozocoautla, Chiapas, Mexico, Carlson 2127 (BM, US). In addition to the differences between S. chiapensis and S. idiospora mentioned above, the new species seems to have the median leaves more obviously in two ranks adjacent to. one another, the ranks scarcely or not at all overlapping. The paratype cited differs from the type in having acute median leaves, eciliate lateral leaves, and stouter rhizophores. It was originally annotated by Alston as “‘S. idiospora ined.’’, and presumably later in pencil, “‘sp. nov.’ In the sum of its characters, S. chiapensis seems more different from S. idiospora than the latter does from S$. guatemalensis Baker. 26 AMERICAN FERN JOURNAL: VOLUME 70 (1980) uy Diplopterigium bancroftii (Hook.) A. R. Smith, comb. nov. Li4¢@ Gleichenia bancroftii Hook., Sp. Fil. 1:5, t. 4A. 1844’ LECTOTYPE: Jamaica, Bancroft (K!; ee totype BM!), chosen by Proctor in apie ae Less. Antill. 60. 1977). Hooker may never have see the other syntype, which is Jamaica, Sw Nakai (Bull. Natl. Sci. Mus. 59: ‘50. 1950) regarded Gleichenia bancroftii as a synonym of Diplopterigium farinosum (Kaulf.) Nakai, but the original description and subsequent illustrations suggest that Mertensia farinosa Kaulf. is really a species of Sticherus. Love, Léve, and Pichi Sermolli (Cytotax. Atlas Pterid., 1977), apparently following Nakai, also adopt the name Diplopterigium farinosum for Gleichenia bancroftii. yb | Grammitis xiphopteroides (Liebm.) A. R. Smith, comb. nov. 25? = Polypodium xiphopteroides Liebm., Kongel. Danske Vidensk. Selsk. Skr., Naturvidensk. Afd., V. 1:196. 1849. MC ECTOTY PE sie chosen): Mexico, Veracruz, ‘‘Hac. de Mirador,’’ Liebmann [PI. Mex. 2548, Fl. Mex. 189] (C I regard Grammitis eee (Maxon) Proctor as a taxonomic synonym. 2%, Pellaea cordifolia (Sesse & Moc.) A. R. Smith, comb. nov. Tie me cordifolium Sessé & Moc., Naturaleza Seeoy City), ser. II, 1(App.):182. 1890. “TYPE: o, Cuyuacan and San Agustin near Mexico City =" rceaid Pellaea cordata (Cav.) J. Smith ye Fée, 1852), P. cardiomorpha Weath., and P. sagittata (Cav.) Link var. cordata (Cav.) A. Tryon as taxonomic synonyms. Pellaea cordifolia differs from P. sagittata in having segments rotundate- cordate (vs. ovate-triangular to sagittate), rachis and segment stalks glabrous (vs. usually puberulous), and spores tetrahedral-globose, 64 per sporangium (vs. ellip- soidal, 32 per sporangium) (A. Tryon, Ann. Missouri Bot. Gard. 44:125-193. 1957). Pellaea cordifolia is a sexual diploid (2n=29 II), whereas P. sagittata s. s. is an apogamous triploid (n =2n =87). The two species are partially sympatric, but P. cordifolia has a much more restricted distribution than P. sagittata. The mor- phological, chromosomal, and chorological differences between the two taxa seem of a magnitude to justify recognition of two species. que Pleopeltis macrocarpa var. interjecta (Weath.) A. R. Smith, comb. nov. 434 & Polypodium peltatum var. interjectum Weath. Amer. Fern J. 34:17. 1944. TYPE: Cerro de Tecpam near Santa Elena, Chimaltenango, Guatemala, Standley 60957 (F). As Weatherby stated, var. interjectum is closely related to var. trichophora (Weath.) Pic. Ser. and var. macrocarpa (Weatherby’s var. lanceolata) on the one hand and Pleopeltis polylepis (Roem. ex Kunze) Moore (= Polypodium peltatum Cav., according to Christensen, Dansk. Bot. Ark. 9(3):11. 1937) on the other hand. It is quite possible that var. interjectum is an evolutionary link to P. polylepis from P. macrocarpa. My reason for placing it with the latter is that I find that scale size and number (large and numerous in P. polylepis, small and sparse in P. macrocarpa including var. interjectum) on the abaxial surface of the blade are more consistent than characters of scale margin (entire in P. polylepis and entire to erose to denticulate in P. macrocarpa). Weatherby used characters of the scale margin to distinguish between P. macrocarpa and P. polylepis. A. R. SMITH: NEW PTERIDOPHYTES FROM CHIAPAS 27 7%) Polystichum fournieri A. R. Smith, nom. nov. 1712 Polystichum muelleri Fourn., Mex. Pl. 1:91. 1872 (non Schum., 1803) "LECTOTYPE (here chosen): ‘‘In pinetis prov. Chiapas,’’ Mexico, Linden (P!). The other syntypes (all from Mexico) are: “*San Luis de Potosi,’ Virlet d’Aost 46 (P-4 sheets!); ‘‘Orizaba,”’ F. Miiller 1496 (not found at P); and ‘‘Prope ignovomum Rio Frio,’ Bourgeau (P!). Of the syntypes I have seen, only the Linden specimen is ertile. 744 S§ticherus brevipubis (Christ) A. R. Smith, comb. nov. THIS Gleichenia brevipubis Christ, Bull. Herb. Boissier, II. 6:280. 1906. LECTOTYPE: Valle del Rio Navarro, Cartago, Costa Rica, Wercklé (P; isolectotype US), chosen by Lellinger, Proc. Biol. Soc. Wash. 89:713. 1977. 742? Tectaria transiens (Morton) A. R. Smith, comb. & stat. nov. 7428 Tectaria incisa subsp. transiens Morton, Amer. Fern. J. 56:133. 1966. TYPE: Cordoba, Veracruz, Mexico, Finck 57 (US). In the sum of its characters, T. transiens is intermediate between T. hera- cleifolia (Willd.) Underw. andT. incisa, and it is possible that it has arisen through hybridization. However, in Chiapas it has been collected in areas where neither of the suspected parents have been found. It differs from 7. heracleifolia in its nonpeltate indusia, greater number of lateral pinnae, shorter-stalked lower pinnae, and lack of strongly developed acroscopic lobes on lower pinnae. FromT. incisa, it differs in the pinnae being serrately incised most of their length with two or three large basal basiscopic lobes. From both species (Chiapas material only), 7. trans- iens differs in the presence of short, glandular hairs on the costae, veins, and leaf tissue below and also on the indusia. Spores of several collections appear well- formed. Throughout its range, T. transiens seems to grow at higher elevations than either 7. heracleifolia or T. incisa. Additional collections seen are Sanchez 63 (UC) from Veracruz; Tuerckheim 839 (UC) from Guatemala; and Brade 47 (UC), Stork 2107 (UC), and Stork 1546 (UC) from Costa Rica. Similar collections have been seen from Ecuador and Peru. 28 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) SHORTER NOTES A NEW COUNTY RECORD FOR PILULARIA AMERICANA IN TEXAS. — While collecting on the interesting granite outcrops of Central Texas on July 20, 1979 I came across a local, but extensive population of Pilularia americana A. Br. in a half-acre stock pond. The plants formed dense mats in mud at the margin of the pond, and in several places the mat extended onto the bare granite under the water surface. Livestock had uprooted many plants, and these were floating on the pond surface. Each rhizome measured 10-14 cm long. The plants were all in fruit, the sporocarps varying from olive-green to brown in color. Associated plants included Eleocharis sp., Bacopa rotundifolia, and Lindernia anagallidea. The pond in which the plants occurred had been enlarged from a previous vernal pool by means of an earth dam, which seems to have helped the Pilularia poplation. The population is the first known record for the species in Mason County. Other populations are known in Texas only from near Marble Falls, Burnet County, about 55 miles southeast, in similar habitats over granite. The Mason County outcrop is known locally as ‘‘Spy Rock’’ and is located on a private ranch 8.8 mi E on the north side of Ranch Road 1222 from the intersection with Hwy. 57, in Camp Air, near Fredonia. Specimens are being distributed to the following herbaria: B, F, GA, GH, K, MARY, MO, NCU, NY, SMU, TAES, TEX, US, VT.—Steven R. Hill, Department of Botany, University of Maryland, College Park, MD 20742. ASPLENIUM XGRAVESII DISCOVERED IN ARKANSAS.—In addition to serving as an addendum to the Pteridophytes of Arkansas by Taylor (Rhodora 81:503—-548. 1979), this note is intended to document the value of the A.F.S. annual fern forays as a stimulus to discovery. One of us (Werth) went to Arkansas this past summer not only for the pure joy of a fern foray, but also with the ulterior motive of obtaining Ozarkian and Ouachitan materials for his studies on genetic variation in Asplenium. After a very successful foray, and at the suggestion of the other of us (Taylor) and David Johnson of the University of Michigan, Werth ea Hot Springs National Park in Garland County, Arkansas, on 13 August It was known that A. pinnatifidum and A. bradleyi grew sympatrically in the park on the novaculite outcroppings of the Gulpha Gorge. This population was easily located, although fewer than thirty individuals of each species were found. (Asplenium platyneuron was also present and in much greater abundance.) Nonetheless a robust individual of A. x gravesii Maxon with leaves intermediate between the two parent species was discovered. Fronds and a portion of the rhizome were taken, and later examination of the spores showed them to be abortive. Comparison of the leaves with the drawings appearing in Wagner and Darling (Brittonia 9:57—63. 1957) confirmed that the plant was the hybrid. This hybrid is quite rare, having been reported previously from only five states, apparently as a consequence of the limited concurrence of the uncommon parent species. It is likely that plants of A. gravesii have appeared sporadically in Gulpha Gorge and died, never having been observed until the arrival one summer of an unsated fern forayer. Voucher specimens of Werth 39K8 have been deposited at MU and MIL.—Charles R. Werth, Department of Botany, Miami University, Oxford, OH 45056 and W. Carl Taylor, Department of Botany, Milwaukee Public Museum, 800 W. Wells St., Milwaukee, WI 53233. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) 29 PILULARIA AMERICANA NEW TO _ TENNESSEE.—The Pillwort, P. americana A. Braun, a diminutive, aquatic pteriodophyte in the family Mar- sileaceae, was first collected near Fort Smith, Arkansas by Thomas Nuttall in 1819. Since that time, its known range in the United States has been slowly expanded. Presently it is known in the United States from Crook (collected in 1894) and Lake Counties, Oregon; in California from Siskiyou and Modoc Counties south- ward to San Diego Co.; from Cherry Co., Nebraska; Reno and Harvey Counties, Kansas; Comanche Co., Oklahoma; Burnet Co., Texas; Conway, Faulkner, Gar- land, Logan, Sebastian, and Washington Counties, Arkansas; and Barrow, Wal- ton, Washington, and Oglethorpe (9.3 mi NE of Lexington at Echols Mill, 11 Nov 1962, D. Blake & F. Montgomery s. n., MO) Counties, Georgia. The range of P. americana has also been stated to include Louisiana by H. B. Correll (Amer. Fern. J. 57:31-32. 1967), but D. S. Correll (pers. comm. with D. B. Lellinger) has confirmed that this was an error On 12 Aug 1979, the second author discovered P. americana washed up on the shore of man-made Fall Creek Lake directly behind the Park Inn at Fall Creek Falls State Resort Park, Van Buren Co., Tennessee (A. J. Petrik-Ott 1379 & F. D. Ott, US). Following the initial discovery, we were amazed to find the shore line behind and ca. 100 yards to the east and west of the Inn literally covered with stranded plants of P. americana, and there were equally as many adrift in the water along the shore. These plants ranged from near perfect specimens to those in various stages of decay. In places along this shore, abundant stranded plants of P. americana formed drifts up to six inches wide and one inch deep. There were no rooted plants at this site. The first author found rooted plants on the southeast side of the dam, growing in sand and about six inches of water (4. J. Petrik-Ott 1380 & F. D. Ott, US). A search of the northwest side of the dam yielded an unbelievably large plant of P. americana which was floating and caught among the stem bases of a cattail popu- lation (A. J. Petrik-Ott 1381 & F. D. Ott, US). This plant consisted of an ex- tremely branched, continuous rhizome bearing numerous leaves (morphologically rachises and stipes) and tufts of roots. There was sufficient material from this one plant to make five rather crowded herbarium specimens. Other floating plants were caught among the rocks of the dam on the lake side in great quantity. Rooted plants of P. americana (only 1 to 2 cm tall) were found to bear several sporocarps, but those found floating and stranded on shore (up to 6 cm tall) bore only occasional sporocarps. Before our collections were made, there had been a good deal of rain and the lake appeared to be up about a foot above normal. Many of the stranded and floating plants may have been washed free from their rooted places; indeed many still had sediment clinging to their roots. However, it 1s difficult to believe that the large, extremely rhizomatous plant found on the northwest side of the dam had been uprooted. Its rhizome was quite green, not whitish like the rhizomes of rooted plants, and its roots free of sediment. This plant certainly would have broken into many pieces had it previously been rooted in the soil. 30 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) This site for P. americana extends its known range by approximately 400 miles ENE from Faulkner County, Arkansas and 170 miles NW of Walton County, Georgia.—Aleta Jo Petrik-Ott and Franklyn D. Ott, Department of Biology, Memphis State University, Memphis, TN 38152. NEW NAMES FOR POLYPODIUM CHNOODES AND P. DISSIMILE. —Among the species of Polypodium subg. Goniophlebium, few are more distinct than P. chnoodes Spreng. (Neue Entdeck. 3:6. 1822). Specimens of this species have very large (5-8 mm long), strongly clathrate, blackish, spreading rhizome scales and are weakly and evenly pilosulous on both lamina surfaces. The pinnae are fully to partially adnate to the rachis and, in the latter case, often have a conspicuous, basal auricle overlapping the rachis. Small, round sori are borne in 2 or 3 rows on each side of the pinna midrib. The veins anastomose in a typical goniophlebioid pattern. The species is found in the Antilles, on Trinidad, and from Guatemala to Venezuela and Colombia. In looking at type photographs and specimens of New World Polypodium, I was surprised to see that the type of P. dissimile L. (Syst. Nat. ed. 10, 2:1035. 1759), a specimen collected by Browne in Jamaica (LINN 1251.24), is exactly the same as P. chnoodes. Apparently every- body has adopted Sprengel’s name; nevertheless it must be put in synonymy under P. dissimile. The name P. dissimile has been applied consistently but incorrectly to a species of Polypodium subg. Polypodium which has small (2—4) mm long), non-clathrate, reddish-brown, appressed rhizome scales and which is glabrous on both lamina surfaces, except for minute hairs on the costae. Most pinnae are partially adnate to the rachis and are abruptly contracted at the base, often from a rather dilated supra-basal portion. Small to medium, slighty elongate sori are borne in a single row on each side of the pinna midrib. The veins are 2- or 3-forked and do not anastomose. The species is found over roughly the same range as *‘P. chnoodes,”’ and in addition extends to Mexico, Peru, and Suriname. Now that P. dissimile has to be used for what was called P. chnoodes, the next available name is P. sororium Humb. & Bonpl. ex Willd. (Sp. Pl. ed. 4, 5:191. 1810), based on a specimen collected by Humboldt and Bonpland near Caripe, Venezuela (B-Hb. Willd. 19684). The name P. sororium has been used occasionally in the past for some Venezuelan specimens, but generally has been thought to be a synonym of P. dissimile.—David B. Lellinger, U.S. Nat’l. Herbarium NHB-166, Smithsonian Institution, Washington, DC 20560. A NEW RECORD FOR PELLAEA ATROPURPUREA IN MARYLAND. — While collecting specimens for the Herbarium at Towson State University, I discovered a small colony consisting of six plants of the Purple-stemmed Cliff- brake, P. atropurpurea (L.) Link, growing in east-facing crevices of an old rail- road trestle at Rowlandsville, in Cecil County, Maryland. This is a new county record for Maryland, as well as the first record of this species for the Delmarva Peninsula. A voucher, Redman 3698, has been placed in the Towson State Uni- versity Herbarium (BALT).—Donn E. Redman, Herbarium, Towson State Uni- versity, Towson, MD 21204. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 1 (1980) 31 AN ATYPICAL ATHYRIUM FROM EASTERN TENNESSEE. — A recent bo- tanical exploration of the Doe River gorge in Carter County, Tennessee has led to the discovery on 8 August 1979 of an unusual growth form of Athyrium asplenioides (Michx). A. A. Eaton. A single population consisting of several frond-bearing rhizomes occurs along a moist, northeast-facing rock face at Pardee Point, adjacent to the abandoned roadbed of the East Tennessee & Western North Carolina Railroad, 1.5 miles southeast of Hampton (Wofford, Smith & Collins 79-222, TENN, US). The identity of the specimen was confirmed by A. M. Evans. These plants resemble vigorously growing parsley and have crispate, fasciated pinnae crowded toward the lamina apex (Fig. 1). Although this is a dramatic departure from the usual frond morphology of Athyrium, the fragile, pale brown scales of the stipe base, the sparsely glandular indusia, and the paired vascular strands uniting above the stipe base into a U-shaped midvein are distinguishing characters of the genus. Sori are infrequent, are borne intramarginally, and have typical hammate-asplenioid indusia attached near the lateral vein. Examination of the sori reveals aborted sporangia and a complete absence of spores. The apparent adaptation to the rock face, atypical growth form, incomplete sporogenesis, and FIG. 1. Specimens of Athyrium asplenioides mutant from Carter Co., Tennessee (Wofford, Smith & Collins 79-222). 32 AMERICAN FERN JOUNRAL: VOLUME 70 NUMBER 1 (1980) the fact that typical A. asplenioides occurs in more mesic sites along the base of the same rock face favors the soe pe that these plants are morphological mutants, rather than of hybrid ori In addition to this unusual atleirtiin. Sanguisorba canadensis, Scirpus ces- pitosus, Paronychia argyrocoma, and Drosera rotundifolia occur at the same locality. Except for D. rotundifolia, these species are of restricted occurrence in Tennessee and are included in the list of rare vascular plants of Tennessee (Com- mittee for Tennessee Rare Plants. 1978. J. Tenn. Acad. Sci. 53(4):128-133).— David K. Smith and B. Eugene Wofford, Department of Botany, University of Tennessee, Knoxville, TN 37916, andJ. L. Collins, Division of Forestry, Fisheries and Wildlife Development, TVA, Norris, TN 37828 Contributions from the Botanical Laboratory, University of Tennessee, n. s. No. 520. TAL SERVICE WNERSHIP p, MANAGEMENT AND CIRCULATION (Required b ry 39 U. 1. TITLE OF PUBLICATION Th 8 2. DATE OF FILING F 3. FREQUENCY OF ISSUE » NO. OF ISSUES PUBLISHED] ®. ANNUAL SUBSCRIPTION ANNUALLY PRICE uarter] 4 8,00 mem; $9.00 sui U.S. Nat'l. Herbarium, 10th & Constitution Aves. NW, Washington, DC 20560 U.S. Nat'l. Herbarium, 10th & Constitution Aves. NW, Washington, DC 20560 6. NAMES PUBLISHER (Name and Address) AMERICAN rama SOCIETY, INC., Smithsonian Institution, Washington, DC 20560 == DITOR (Name eee s 2 eile : MANAGING EDITOR (Name and Address) None 7. OWNER (If owned by a corporation, it: and address must be stated and also immediately thereunder the names and addresses of stock- holders owning or holding | percent or avi ag total amount of stock, If not owned by @ corporation, the names aaa addresses of the = pene owners must be given. If owned by a rebiegaecot — unincor, gesasnd vi , its mame csuaipsghioa _ well as that of each individual m Eames Uf the publication is published by a nonprofit organization, its name and address must be stat % KNOWN BONDHOLDERS, MORTGAGEES, AND OTHER SECURITY HOLDERS OWNING OR HOLDING 1 a OR MORE OF TOTAL AMOUNT OF BONDS, MORTGAGES OR OTHER SECURITIES (II there a1 AD me : | PSM) MAVE NOT CHANGED DURING MAVE CHANGED OURING (If changed, publisher at submit ination chan PRECEDING 2 MONTHS PRECEDING 12 MONTHS: with this statement) oe ne < AVERA ESEACH | ACTUAL NO. COPIES OF SINGLE . ISSUE DURING | Prece DING | ISSUE PUBLISHED NEAREST TO 12 MONT: FILING DATE A wn) be yes PAID CIMcULAT! Reagee \ Saugs ta Tumouer DORE AND COUNTER SALES UB cnt Lenin hr adaan ows OS LET 2. MAIL SUBSCRIPTIONS 1372 ¢. To 082) ee ©. FREE DISTRIBUTION BY MAIL, CARRIER OR OTHER MEANS. SAMPLES, COMPLIMENTARY, AND OTHER FREE COPIES = To ” rer eh F. COPIES NOT DISTRIBUTED 1. OFFICE use, Mae ‘ AFTER PRINTING _ re 2. RETURNS FROM NEWS AGENTS POP" 79 gr ae geruerseeERpinEseee ee RDI: | MEERA esac s. TOTAL (Sum of E. F1 and 2—should equal net press run shown ina) . Lcertify that th: SIGNATURE AND TITLE OF/eoIT: PUBLISHER, SUSINESS * 2 rig OR OWNER te above are correct and complete. ‘ Cas 20 ~ David B. Lellinger, Edit wuse pee w et such 4 x ai If eceordence with the provisions of this statute, | heres: fates presently authorized by 39 U.S. C. 3626. AT HENATURE AND TITLE OF EDITOR, PUBLISHER, BUSINESS MANAGER, OR OWNER Rue 1378 3526 (Page 1) (See instructions on reverse) TRIARCH Over 5@ Wears of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 AMERICAN FERN JOURNAL Volume 70 Number 2 April-June, 1980 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Equisetum x litorale in eae lowa Minnesota, and Wisco Gametophytes of Equisetum diffusum A Double Spore Wall in Macroglossum Subdivision of the Genus Elaphoglossum Notes on the Natural History of Stylites gemmifera Ls) a JAMES H. PECK RICHARD L. HAUKE 39 > a LUIS D. GOMEZ P. and KERRY S. WALTER = | JOHN T. MICKEL and LUCIA ATEHORTUA G. ERIC E. KARRFALT and DALE M. HUNTER 69 Reciprocal Allelopathy Between the Gametophytes of Osmunda cinnamomea a sy Dryopteris inte rmedia YMOND L. PETERSEN and DAVID E. FAIRBROTHERS Shorter Note: Thelypteris torresiana in Venezuela Reviews B 80 38, 68, 79 . i MITSOUM BOTAN RAT JUL 15 1980 GARDEN LIBRARY The American Fern Society Council for 1980 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, Ohio 45701. President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN 37235. Vice President LESLIE G. HICKOK, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916 Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, Va. 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, D.C. 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Berkeley, Calif. 94720 Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, N.Y. 10458. aie Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, D. C. 20560 ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington, Ind. 47401 JOHN T. MICKEL New York Botanical Garden, Bronx, New York 10458 The ‘American Fern Journal’’ (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and published at the Smithsonian Institution, Washington, DC 20560. Second-class postage paid at Washington Matter for publication ee claims for missing issues (made 3-6 months after the date of issue) should be addressed to the Edito Changes of address, ess and applications for membership should be sent to Dr. J. E. Skog, Dept. of Biology, George Mason University, Fairfax, Va. 22030. Orders for back issues should be addressed to the Treasure General inquiries concerning ferns should be addressed to the Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Fern Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or fess; $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, New York 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter r. John T. Mickel, New York Botanical Garden, Bronx, New York 10458, is editor of the newent ter **Fiddlehead Forum.”’ The editor welcomes Scntibubions from members and non-members, includ- ing miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural ches. and reviews of non-technical books on ferns. Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th rg Seattle, Washington 98115, is Director. Spores exchanged and collection lists sent on reque Gite and Bequests pets 4k : L mate Sr Clie ted in ferns. Botanical books, back issues of the Journal, and cash or ene gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secre AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 33 Equisetum X litorale in Illinois, lowa, Minnesota, and Wisconsin JAMES H. PECK* The cross between Equisetum arvense L. (Field Horsetail) and E. fluviatile L. (Water Horsetail) results in the hybrid E. x litorale Kuhl. (Shore Horsetail). It is distinguished from its parents by its abortive spores and intermediate anatomical and morphological traits (Hauke, 1978). The hybrid occurs less frequently in the Midwest than hybrids between unbranched species of Equisetum. The hybrid was reported from one county in Illinois (Mohlenbrock & Ladd, 1978) and four coun- ties in Wisconsin (Hauke, 1965). It was not reported from lowa (Peck, 1976) nor from Minnesota (Tryon, 1954), even though both parents occur together in those states. Its recent discovery in Minnesota (Peck & Swanson, 1978) initiated a field and herbarium search that uncovered a second locality in Illinois, first and second localities in lowa, second, third, and fourth localities in Minnesota, and eight new localities in Wisconsin. Most of these new localities are on or adjacent to the flood-plain of the Mississippi River or its tributaries. The 2-3 ha stand located in Allamakee County, Iowa, is probably the largest stand of the hybrid in North America (pers. comm. with Drs. Hauke and Wagner). Consquently, efforts were made to study the habitat, stand dynamics, and reproductive biology of the hybrid at this locality to identify factors which clarify the origin, abundance, and persis- tence of the hybrid in this locality. Observations also were made at 16 of the 20 localities of the hybrid in these four states, but in less detail. A summary of locality data is provided by Fig. ] and by citations from her- barium vouchers (new county records indicated by *). The original Illinois locality (Lee County) is accepted from the report by Mohlenbrock & Ladd (1978); voucher citation was not given and the specimen was not seen. ILLINOIS: *Carroll Co.: Wet grounds near entrance to Mississippi Palisades State Park, N of Savannah, Wunderlin 2668 (MWI). IOWA: *Allamakee Co.: Lansing Twp.: Lansing Wildlife Refuge, 2-3 ha marsh, T99N R4W S12, Peck 78-54 (ISTC, KIRI, MICH), Farrar 78-6-3-1 (ISC), Roosa 1759 (ISTC). Banks of Mississippi River, June 1900, Orr s.n. (EMNM). *Des Moines Co.: Huron Twp.: Near pumping station No. 4, lowa Slough, adjacent to Mississippi River, T72N R1W S4, Lammers 1542 and 2087 (1A, ISC, ISTC), Peck 78-300 (ISTC, KIRI, MICH). SOTA: *Houston Co.: Mississippi River flood plain in wetlands of Crooked Creek at Reno, Peck 79-734 (KIRI, MICH, MIN, UWL). Washington Co.: Confluence of Valley Branch Creek and St. Croix River in thicket of Salix interior, Swanson 2878 (MIN, UWL). *Wabasha Co.: Weaver Bottoms, floodplain of Mississippi River, 2 mi N of Weaver, Peck 79-709 (KIRI, MICH, MIN, UWL). *Winona Co.: On floodplain of Mississippi River, W of Red Oak Island in Lake Onalaska, Peck 79-727 (KIRI, MICH, MIN, UWL). WISCONSIN: *Buffalo Co.: Nelson Twp.: Nelson-Trevino Bottoms of Mississippi River floodplain, T22N R14W S36, Peck 79-824 (KIRI, MICH, WIS, UWL). *Crawford Co.: Emergent along shore near bridge over Swamp Creek, 0.5 mi E of Lynxville on County Road B, TIN R6W S23, 8 Jul 1973, Dawson s. n. (UWL). Grant Co.: Wilderness area of Wyalusing State Park, T6N R6W $20/21, 19 June 1959, Patman s. n. (WIS). Green Lake Co.: Marsh shore on S side of Lake Puckaway, Marquette, Fassett 8799 (WIS). *La Crosse Co.: Barre Twp.: Seepage area at roadside ditch along Swamp Road, *Department of Biology, University of Wisconsin-La Crosse, La Crosse, WI 54601. 34 AMERICAN FERN JOURNAL: VOLUME 70 (1980) TI6N R6W S820, Peck 79-803 (KIRI, MICH, UWL, WIS). *Pierce Co.: Clifton Twp.: Marshy thicket at mouth of Kinnicinick River under Salix interior, T27N R20W, Peck 79-824 (KIRI, MICH, UWL, WIS). Isabella Twp.: Marshy slough, backwater of Mississippi River. 0.5 mi W of Bay City, T24N RI7W S7, Peck 79-838 (KIRI, MICH, UWL, WIS). Richland Co.: Shallow water springhole in slough of Wisconsin River, 3 mi E of Gotham, T8N R2E S4, Hartley 5266 (IA, WIS). *Rock Co.: Marsh and lowlands east of cooling canal on Rock River near Beloit, Rice 1649 (UWJ, WIS). *Trempealeau Twp.: Marshy edge of backwater area of Mississippi River floodplain in Delta Fish and Fur Farm, T18N R1OW SII, Peck 79-814 (KIRI, MICH, UWL, WIS). *Vernon Co.: Genoa Twp.: Shore to Bad Axe River near Mississippi River, T12N R7W S12, Peck 79-813 (KIRI, MICH, UWL, WIS). Winnebago Co.: Springy shore of Fox River near Eureka, Fassett 13243 (WIS). The habitat of the hybrid is a shallow marsh or slough adjacent to a watercourse which has a fluctuating water level, periodically flooding or stranding the site where the hybrid occurs. The hybrid occurs in stands ranging from 1 m? to 2-3 ha, with many stands 0.05-0.1 ha in extent. Species diversity within the stand is very low compared to adjacent marshes. The most common associates are Sagittaria latifolia, Salix interior, and Typha latifolia. The parent species were found at the periphery of the hybrid’s stand, but rarely within the stand. The hybrid appears to be quite aggressive in a habitat which is subjected to repeated disturbance by flooding and sediment deposition. TABLE 1. CHANGES IN EQUISETUM x LITORALE STAND HEIGHT AND DENSITY IN ALLAMAKEE COUNTY, IOWA, DURING 1979. Stand height Stand density Survivorship Sample Date (¥ + sd) (stems/m?) (% of 21 April) 21 April 0.05 + 0.012 2809 + 61.8 100 12 May 0.49 + 0.049 2560 + 82.9 91 12 June 1.03 + 0.101 1792 2-133 64 30 July 1.52 + 0.082 1324 + 148 47 20 Aug! 0 0 0 15 Sep? 0.34 + 0.031 32. db 1 : Stand lodged; all aerial parts senescent. Aerial stems newly arisen from subterranean stems. Stand dynamics were monitored in 1979. A transect was established the length of the stand. Twenty-five 0.25 m2 quadrats were selected at random points along quadrat. Stand height was assessed by measuring the height of two corner plants per quadrat. Subsequent observations were made from a series of transects 5 m distant from and parallel to the previous transect. An average value for stand density (stems/m?) and stem height was calculated for each transect. Survivorship of aerial stems through the growing season was calculated based upon the average greatest in spring and declined until summer, when only 47% of its initial stand density was present. Considering the initial stand density (2,809 stems/m?), exten- sive self-thinning was expected. Stand height increased through time resulting in mean stand height of 1.52 m, with some exceptionally tall specimens over 2 m. J. H. PECK: EQUISETUM x LITORALE 35 The stand was flattened by a severe storm in early August 1979. Consequently, no measurements could be taken on August 20th. However, by September 15th, the stand had sprouted new aerial stems from subterranean stems. These new stems produced a comparatively sparse growth of limited stature. The new growth Wi FIG. 1. Distribution of Equisetum ~ litorale in the four states of the upper Mississippi River Valley. protruded through a 10-18 cm thick mat of senescent and decaying aerial stems that lodged following the storm. The stand also lodged in 1978 following a storm in July. The new growth that year attained a mean height of 1.2 m. The extent of recovery, therefore, is probably influenced by the time of lodging, with an earlier date favoring a stronger recovery. 36 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Observations on stand dynamics were also made on an adjacent pure stand of the parent species. Those stands were less densely stocked and did not recover following lodging to the extent the hybrid did. Whereas the hybrid occurs with stands of low species diversity, the parent species exist with many more species present within their stands. The lack of colonization or establishment within the hybrid’s stands and the extensive colonization in parent stands in autumn suggests that the thick mat of lodged stems in the hybrid’s stand physically prevents suc- cessful invasion. Consequently, the aggressive growth of aerial stems and capabil- ity to recover from mid-summer lodging contribute to the hybrid’s ability to main- tain stands of low species diversity. In considering the reproductive biology of the hybrid, it should be noted at the outset that the hybrid’s spores are unable to produce gametophytes and, thus, complete the sexual cycle. Furthermore, the general lack of strobilus production by the hybrid in the upper Mississippi River valley essentially precludes the rare event of production of an occasional viable spore. Consequently, sexual origin of the hybrid is restricted to hybridization between its parents. Frequent flooding along watercourses in the upper Mississippi River valley forms extensive sand and mud flats that provide suitable conditions for Equisetum gametophytes and ample opportunities for hybridization. The hybrid’s occurrence in small, discrete stands, with uniform growth within a stand and readily demonstrated physical connec- tions underground, suggests that each stand is the result of a single or a few successful hybridizations followed by establishment of a clone by vegetative growth. Therefore, the stand at each locality may represent an independent origin of the hybrid. Once formed, a hybrid plant can persist indefinitely as long as it avoids cata- Strophic factors, such as desiccation during drought years. An indication of the longevity of the hybird is given by the continued presence of the hybrid in Alla- makee Co., Iowa, for at least the last 80 years. Since 1900, when Ellison Orr first noted it on the flood plain of the Mississippi River, establishment of navigation lock and dams has changed the hydrology of the shoreline environment. The presence of the hybrid in that locality today suggests that its historical presence and persistence has not been eliminated by alterations to the shoreline. Vegetative proliferation of hybrid stands, on the other hand, was suggested by the occurrence of small stands and isolated stems 50-350 m downstream from the 2-3 ha stand in Allamakee Co., Iowa. Fragmentation followed by water dispersal of stems and their subsequent establishment downstream would result in new stands from the same original plant. The propensity for vegetative proliferation of Equisetum was discussed by Hauke (1963) and experimentally investigated by Wagner & Hammitt (1970). In early August 1979, an experiment was conducted to verify the hybrid’s ability to undergo vegetative proliferation and to contrast this ability with that of its parents. Aerial stems and subterranean stems of the hybrid in early September. The results (Table 2) indicate that burial increased the chance that a fragment would form a new plant, that subterranean parts withstood the J. H. PECK: EQUISETUM x LITORALE 37 stress of fragmentation, dispersal, and establishment better than aerial stems, and that the hybrid’s ability is equal to or superior to that of its parents. In summary, the origin, maintenance, and dispersal of the hybrid is favored by the physical results of flooding. Flooding can form mud flats where sexual forma- tion of the hybrid can occur, lodge aerial stems into a mat that prevents invasion of the hybrid’s stands by potential colonizers, and facilitate vegetative proliferation by physically breaking stems, dispersing them downstream and leaving them stranded or buried on mud flats where new stands can become established. Al- though vegetative proliferation probably has expanded the hybrid within its lo- cality, evidence of long distance dispersal is lacking, in that both parents occur with the hybrid in these states. TABLE 2. COMPARATIVE ABILITY OF E. x LITORALE AND ITS PARENT SPECIES TO REGENERATE PLANTS FROM AERIAL AND SUBTERRANEAN STEM FRAGMENTS AFTER BEING LODGED ON OR BURIED IN A MUD FLAT Species Cut and Lodged Cut and Buried Aeria Subterranean Aerial Subterranean E. arvense 0% 20% 60% 68% E. x litorale 0% 16% 40% 96% E. fluviatile 0% 8% 0% 20% These observations suggest that analysis and integration of habitat, stand dynamics, and reproductive ecology are feasible with Equisetum, and possibly with other pteridophytes. Additional observations are needed on changes in nutri- tional status (calories, macronutrients, and protein) of aerial and subterranean stems through the year. Additional measures on stand dynamics (height, density, biomass) are needed to contrast peripheral stands reported on here with stands to the north and east. Competitive experiments between the hybrid, parent species, and flowering plant associates would improve our understanding of Equisetum ecology. Dr. Richard L. Hauke, University of Rhode Island, and Dr. Warren H. Wagner, Jr., University of Michigan, are thanked for verifying hybrid material and for comments on the ecology of Equisetum. Curators of the following herbaria are thanked for permission to examine specimens: Effigy Mounds National Monu- ment, Marquette, [A (EMNM), Iowa State University (ISC), Mankato State Uni- versity (MANK), University of Iowa (IA), University of Minnesota (MIN), Uni- versity of Northern Iowa (ISTC), University of Western Illinois (MWI), Univer- sity of Wisconsin-Janesville (UWJ), University of Wisconsin-La Crosse (UWL), and University of Wisconsin-Madison (WIS). LITERATURE CITED HAUKE, R. L. 1963. A taxonomic monograph of the genus Equisetum subgenus Hippochaete. Nova Hedw. Beih. 8:1-123. . 1965. Preliminary reports on the flora of Wisconsin. No. 54. Equisetaceae—Horsetail Fam- ily. Trans. Wis. Acad. Sci. 54:331-346. . 1978. A taxonomic monograph of Equisetum subgenus Equisetum. Nova Hedw. 30:385—- 455. 38 AMERICAN FERN JOURNAL: VOLUME 70 (1980) MOHLENBROCK, R. H., and D. M. LADD. 1978. Distribution of Illinois Vascular Plants. Southern Illinois University Press, Carbondale, IL. PECK, J. H. 1976. The pteridophyte flora of Iowa. Proc. Iowa Acad. Sci. 83:143-160. ———.,, and S. D. SWANSON. 1978. Equisetum x litorale recorded for Minnesota. Amer. Fern J. TRYON, R. M. 1954. The Ferns and Fern Allies of Minnesota. University of Minnesota Press, Minneapolis, MN. WAGNER, W. H., Jr., and W. E. HAMMITT. 1970. Natural proliferation of floating stems of scouring-rush, Equisetum hyemale. Mich. Bot. 9:166-174. REVIEW “FLORA DE LA PROVINCIA DE JUJUY REPUBLICA ARGENTINA, PARTE Il. PTERIDOPHYTA,” by E. R. de la Sota. Angel L. Cabrera, general editor. Instituto Nacional de Tecnologia Agropecudria, Buenos Aires, Argentina, 1977. Ps. 22,000 (ca. $28.00).—INTA is publishing a series of very attractive Flora volumes for the provinces of Entre Rios and Jujuy (both in northern Argentina) and for the Patagonian region. This volume is the first on ferns for the three series. The endpapers have useful physiographic, phytogeographic, and political maps of the province. The book begins with a few pages concerning the morphology, cytology, reproduction, and systematics of the Pteridophyta. General keys lead to 23 families, which are used in a modern sense similar to that in Jermy and Mickel’s classification. Each species treatment includes a synonymy, description, notes, and list of specimens. Each is illustrated with a nicely drawn habit sketch plus sketches of morphological details. Some 244 taxa are treated, judging by the index that concludes the volume. The printing is mostly of high quality, although a few plates had some light areas across them and a few typographical errors can be found. The weakest point of the book is the binding. This book will be highly useful to all who need to know the pteridophytes of northwestern Argentina and vicinity. Orders should be sent to INTA Publicaciones, Chile 460, 1098 Buenos Aires, Argentina.—D.B.L. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 39 Gametophytes of Equisetum diffusum RICHARD L. HAUKE* Equisetum gametophytes have been studied for many years (Buchtien, 1887; Hauke, 1967, 1968, 1969, 1971, 1977; Duckett, 1970, 1972, 1973, 1977; and litera- ture cited in these references). The gametophytes of all Equisetum species except E. diffusum Don have been described. In a recent monograph of Equisetum subg. Equisetum (Hauke, 1979), I presented an evolutionary sequence of species ar- ranged according to gametophyte specialization which did not parallel the se- quence based on sporophyte specialization. Subsequently I was able to obtain viable spores of E. diffusum and to culture them. The purpose of this paper is to describe gametophytes of Equisetum diffusum and to discuss their evolutionary implications. MATERIALS AND METHODS On 24 February 1974, I collected living rhizomes of E. diffusum along the road between Chail and Kandaghat, near Simla, Himachal Pradesh state, India. These flourished in pots in the greenhouse at the University of Rhode Island. In August 1978, I first noticed cones developing on the ends of unbranched, new stems in two pots. The plants in these pots apparently had died back and regrown. The first cone was removed before it expanded, surface sterilized with 50% commercial sodium hypochlorite bleach, rinsed with sterile distilled water, and dissected in sterile distilled water. Ten drops of the spore suspension were inoculated onto petri dishes containing Bold’s Basal Medium (BBM) in 1.5% agar. Subsequent cones were allowed to open naturally, and the spores were shaken onto the sur- face of the solidified nutrient medium. The culture dishes were placed in a growth chamber on a 12 hr light/dark cycle at a temperature of 20/15°C under 40 watt cool white fluorescent tubes yielding 8000 ergs/cm?/sec radiant energy at the surface of the cultures, as measured with a YSI radiometer. Gametophytes were grown in isolation to determine crossability, self- compatibility, and the possible occurrence of apogamy. Spores that had begun to germinate in petri dishes were transferred aseptically into 15 mm diameter test tube slants of BBM agar capped with metal caps and placed in racks for incubation in the growth chamber. Initially the isolated gametophytes grew more slowly than those left in petri dishes, possibly because the test tube caps shaded the gametophytes; when the tubes were positioned to allow full light intensity on the agar surface, the growth rate accelerated. RESULTS The spores of E. diffusum, like those of all other species of Equisetum, are spherical, chlorophyllous, thin-walled, alete, and have two hygroscopic elaters attached at their middles, which form four strap-like arms with spoon-shaped tips. Under suitable conditions of moisture, temperature, and light, they germinate readily in 1-2 days by dividing eccentrically to produce a small rhizoid cell that *Department of Botany, University of Rhode Island, Kingston, RI 02881. 40 AMERICAN FERN JOURNAL: VOLUME 70 (1980) FIGS. 1+. Gametophytes of Equisetum a. FRG. i; Gametophyte a4 days old showing rhizoids, eal (m), and plates, x 70. FIG. 2 64davsold. x 1.8. G. 3. Archegonial gametophyte 50 Pee old, showing broad plates with. irregular margins, X nes FIG. Gametophytes 158 days old with sporophytes. Left =selfed bisexual gametophyte, sporophyte visible 20 days earlier. Right=crossed iia gametophyte, sporophyte visible <5, days earlier. Note smaller size of gametophyte on right, x R. L. HAUKE: GAMETOPHYTES OF EQUISETUM DIFFUSUM 4} loses its chloroplasts and a large, green somatic cell. Further division of the latter produces a flattened, linear gametophyte which branches to form several plates of cells. Eventually a parenchymatous cushion bearing plates dorsally and rhizoids ventrally is established by a marginal meristem. (Fig. 1). Sex organs begin to appear 35 days after inoculation on the basal cushion meristem of initially unisexual gametophytes. Young male and female gameto- phytes look alike, but with continued growth they become dimorphic (Fig. 2). The female gametophytes become larger than the male, develop numerous plates, and assume a grass-green color (Fig. 3). The plates of E. diffusum are up to 2 mm broad and have irregular margins and thickened bases. Archegonia develop at the base of the plates and consist of three tiers of four neck cells each. The egg is embedded in the cushion. The terminal neck cells elongate to four times longer than broad and spread apart in an arching manner. Seen from above, the spread terminal necks cells resemble a pinwheel. The male gametophytes remain smaller than the female, with only sparse plate development, and are yellowish to pinkish. Antheridia develop from the cushion and protrude somewhat at maturity, becoming twice as long as wide. They dis- charge sperm by two or four cap cells. The cap cells in this species are not so distinctive as in other species and apparently may divide anticlinally, so that at times there are three or five cap cells. Female gametophytes that are not fertilized within a certain time begin to pro- duce antheridia. The marginal meristem that has been producing archegonia grows out into an antheridial lobe. In petri dish cultures, where interaction between gametophytes is possible, the bisexual condition may become apparent within 50 days. In isolation tubes, where interaction is not possible, bisexuality is delayed to 90 days or later. Some isolated gametophytes 150 days old still appear only female. Gametophytes apparently cease growth when they begin bearing sporophytes (Fig. 4). If they are unisexual at that time, they never become bisex- ual. Unlike the situation usually seen in other pteridophytes, Equisetum gametophytes normally bear several sporophytes per gametophyte (Fig. 4). Whereas more than 50% of the gametophytes in plate cultures are antheridial, in isolation tubes, in the absence of interaction between gametophytes (but presum- ably with better nutrition), only 5 out of 42 (12%) were male. The other 37 were female. One tube inadvertently received two gametophytes initially, one of which became male and the other female. Three of the male gametophytes were trans- ferred to tubes with female gametophytes and flooded. The tube with two gametophytes was flooded (Fig. 4), as were 17 tubes with only female gametophytes. The four tubes with both male and female gametophytes all showed sporophytes a month later. None of those with only female gametophytes did. The tubes originally containing only female gametophytes were reflooded, and eventu- ally they became bisexual and selfed (Fig. 4). At 158 days post inoculation the experiment was terminated and the tubes refrigerated to stop growth. Later they were examined and discarded, at which time 14 of the 17 selfed gametophytes had visible sporophytes. In two cases, the sporophytes all looked achlorophyllous. 42 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Of the 17 tubes never flooded, one had a single visible sporophyte and another had four. In both cases, antheridial lobes had arched over to press antheridia against the archegonial meristem, and presumably sperm had been able to enter the archegonia in the absence of free water. The other 15 tubes showed no visible sporophytes. In general, flooded gametophytes grew slightly more than did un- flooded ones. DISCUSSION The gametophytes of E. diffusum are similar to those of other species of Equisetum in basic form. They consist of a parenchymatous cushion bearing rhizoids ventrally and photosynthetic plates dorsally. In sexual condition, they produce separate antheridial and archegonial gametophytes, with the latter sub- sequently becoming bisexual if unfertilized. Sexual dimorphism, cessation of growth in sporophyte-bearing gametophytes, and production of numerous sporophytes per gametophyte are all seen in other species of Equisetum. The last character is rather uncommon in archegoniates, where gametophytes usually have several archegonia, but normally produce only one sporphyte per gametophyte. Bruce and Beitel (1979) tabulated 476 Lycopodium gametophytes, of which 17% had more than one sporophyte and one had 13 sporphytes. Mesler (1977) reported up to 27 sporophytes on a wild gametophyte of Equisetum hyemale. The increased percentage of archegonial gametophytes in isolation culture and the delay in their becoming bisexual also occurs in other species of Equisetum (Hauke, 1977). Certain features of E. diffusum gametophytes indicate that they are relatively primitive and more closely allied to those of E. telmateia than to any other species. Their plates are broad, with irregular margins and thick bases. Their archegonia have terminal neck cells about four times longer than wide. Their antheridia have 2—4 cap cells and are about twice as long as wide. Plate develop- ment on antheridial gametophytes is more extensive than in other Equisetum species except E. telmateia. Duckett (1973) carefully described the morphology of several species of subg. Equisetum; using that information plus my own observations (Hauke, 1968), I aligned the gametophytes of subg. Equisetum in a presumed evolutionary se- quence (Hauke, 1979), with the most primitive species being E. telmateia: arvense telmateia __—_- pratense me = PALUStTe me =f Viatile ep = bOZOteNnse sylvaticum From my current observations of E. diffusum gametophytes, it is now obvious that in an evolutionary scheme based on gametophytes, E. diffusum is inter- mediate between E. telmateia and the three species of sect. Heterophyadica sensu Hauke: E. arvense, E. pratense, and E. sylvaticum. One problem raised by this sequence is that it does not correlate well with the apparent evolutionary sequence based on sporophyte character: fluviatile > palust > bogot > diff i sich p > gotense —P CITUSUM —eempe telmateia ————e pratense sylvaticum R. L. HAUKE: GAMETOPHYTES OF EQUISETUM DIFFUSUM 43 That sequence, it is true, placed E. diffusum and E. telmateia side by side, but it also considered E. bogotense to be most like E. diffusum. This is the greatest discrepancy between the schemes based on sporophyte and gametophyte evolu- tion. Perhaps there is no reason why the two generations should be correlated evolutionarily, since they presumably evolve for different environmental fitness, but it seems appropriate to consider the whole plant in a single evolutionary scheme. In that case, one might expect the gametophyte to reflect more conserva- tive traits, as the reproductive structures of other plants are assumed to do. For an example of independent selection for floral and vegetative fitness in flowering plants, however, see Wilken (1978). If taxonomy is intended to be phylogenetic, should it emphasize gametophyte or sporophyte? One factor which diminishes the usefulness of the gametophytic stage in pteridophyte taxonomy is the paucity of characters it possesses. A compromise scheme, utilizing characters of both gametophytes and sporophytes might be: arvense pratense palustre sylvaticum bogotense diffusum telmateia It injures my sense of the fitness of things to look at two species with sporophytes as similar as those of E. bogotense and E. diffusum and to separate them widely in a classification. Yet the gametophyte as well as the sporophyte must be consid- ered in arriving at any taxonomy which claims to be phylogenetic. In fact, if it is more conservative in evolution, then it should be given greater emphasis in taxonomy. Isolation experiments showed that the gametophytes of E. diffusum outcross readily. They also self readily in most cases, but the absence of any detectable sporophytes on three of 17 selfed gametophytes, and the chlorotic sporophytes on two others, indicates some lethal load (see Lloyd, 1974). The absence of sporophytes on individuals that were not flooded indicates the absence of apogamy in E. diffusum. Ease of selfing and absence of apogamy are also found in the other species of Equisetum. I wish to thank Dr. Stoddard Malarky for helping me collect E. diffusum and Dr. Roger Goos for reading the manuscript. ADDENDUM In a paper which appeared while this article was awaiting publication, Duckett (1979) made several observations which are pertinent here. He reported that pro- longed culture and numerous attempted fertilizations were required to obtain maximal sporophyte frequencies, and suspected ‘‘leaky lethals,”’ but on the basis of crossing and selfing tests discounted that possibility. He noted that the initia- tion of sporophytes is accompanied by cessation of gametophyte growth, and attributed this to allelopathic substances from the sporophyte. He reported that polyembryony is present in all species, but is uncommon and occurs mostly in ot AMERICAN FERN JOURNAL: VOLUME 70 (1980) Species with rapid sex change from female to bisexual, or with numerous archego- nial lobes. | have observed polyembryony to be common in isolated, selfed gametophytes not only of E. diffusum, but also of E. fluviatile, E. hyemale var. affine, and E. arvense, and suspect that they were all larger and therefore with more receptive archegonia when flooded than were those of Duckett’s e xperi- ments LITERATURE CITED BRUCE, J. G., and J. M. BEITEL. 1979. A community of Lycopodium gametophytes in Michigan. Amer. Fern J. 69:33-41. BUCHTIEN, O. 1887. Entwicklungsgeschichte des Prothallium von Equisetum. Biblio. Bot 2(8):1-49. DUCKETT, J. G. 1970. Sexual behaviour of the genus Equisetum, subgenus Equisetum. Bot. J. Linn. Soc. 63:327-352. . 1972. Sexual behaviour of the genus Equisetum, subgenus Hippochaete. Bot. J. Linn. Soc. 65: 87-108. . 1973. Comparative morphology of ai gametophytes of the genus Equisetum, subgenus auisetian, Bot. J. Linn. Soc. 66:1-2 . 1977. Towards an understanding of sex | detienination j in Equisetum: an analysis of regenera- tion in gametophytes of the subgenus Equisetum. Bot. J. Linn. Soc. 74:215-242. . An experimental study of the reproductive biology and hybridization in the European and North American species of Equisetum. Bot. J. Linn. Soc. 79:205- HAUKE, R. L. 1967. Sexuality in a wild population of Equisetum arvense gametophytes. Amer. Fern 57:59-66. . 1968. Gametangia of Equisetum bogotense. Bull. Torrey Bot. Club 95:341-345. . 1969. Gametophyte development in Latin American horsetails. Bull. Torrey Bot. Club 96: 568-577. - 1971. The effect of light quality and intensity on sexual expression in Equisetum gametophytes. Amer. J. Bot 73-377. 1977. Experimental studies on growth and sexual determination in Equisetum Co Amer. Fern J. 67:18-31. 979. A taxonomic monograph of Equisetum subgenus Equisetum. Nova Hedwigia 30:385— od LLOYD, R. M. 1974. Reproductive biology and evolution in the Pteridophyta. Ann. Missouri Bot. Gard. 61:318-331. MESLER, M. R., and K. L. LU. 1977. Large California. Amer. Fern J. 67: 97-98. 1978. Vegetative and floral relationships among western North American popula- tions of Collomia linearis Nuttall (Polemoniaceae). Amer. J. Bot. 65:896-901. gametophytes of Equisetum hyemale in northern AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 45 A Double Spore Wall in Macroglossum LUIS D. GOMEZ P. and KERRY S. WALTER* Since the original description of Macroglossum alidae by Copeland (1909a, p. 343; 1909b, p. 9), little has been learned about this strange and primitive member of the Marattiales. Campbell (1911, 1914a, 1914b) has dealt with the anatomical aspects of the gametophytes and sporophytes but, surprisingly, he gives no de- scription of the spores. In Erdtman (1957, p. 79), the proximal and distal faces of M. alidae are illustrated both in surface view and in optical section; the spore is globose but distally flattened, with an irregular and sparsely tuberculate exine. Kremp & Kawasaki (1972, p. 4) described the spores from seven specimens as rounded-triangular (32.4 x 28.7 wm), trilete, and scabrate. The observation of spores of M. alidae (Molesworth-Allen 3197, US) under the SEM shows a regularly and densely tuberculate-subbacillate perispore (Fig. 1), the shape and dimensions of which are in agreement with the foregoing authors. An unusual case of double-walled spores is shown in Fig. 2, a preparation from the same specimen. In it, the outer exine layer is cracked to reveal a smaller but morphologically perfect spore inside, one per *‘parent spore.’’ This phenomenon is hitherto unreported in the fern literature but may not be rare, for we have observed a ‘“‘parent spore’ of Botrychium sp. containing four ‘‘daughter spores,” and it is quite possible that such ‘‘angiospores’’ occur in other pteridophytes. The biosystematic implications of angiospore production are, as yet, unknown. Research is needed to elucidate various questions that come to mind, including: (a) What, if any, percentage of angiospores is viable? (b) What is their genotype and resulting phenotype? (c) What is the ploidy level of angiospores in relation to ‘‘parent spores’’? (d) Do angiospores represent a reduction mechanism for poly- ploidal pteridophytes? (e) Within a sporangium, what percentage of ‘parent spores’’ contain angiospores? (f) What effect would this ratio have on the popula- tion structures of the resulting gametophytic and sporophytic generations? (g) Does the smaller size the angiospores have any effect on the range and pattern of their dispersibility? (h) Is angiospory a primitive trait only to be found in eu- sporangiate pteridophytes? pore size and shape, of themselves, are not indicative of viability. Recent literature abounds with examples of the germination of supposedly non-viable abortive spores in hybrids and of larger than normal diplospores formed through ameiotic apogamy. At present, questions (b) and (d) are unanswerable due to the lack of appropriate materials. The fact that the cytology of Macroglossum has never been investigated prevents speculation on whether angiospores represent any change in ploidy level, be it reduction or augmentation. The Marattiales have high chromosome numbers as do the Ophioglossales, the only other instance in which we have as yet observed angiospory. Until a large enough quantity of both ‘‘parent spores’ and angiospores are cultured, questions (e) and (f) also remain unanswerable. *Museo Nacional de Costa Rica, Apartado 749, San José, Costa Rica. 46 AMERICAN FERN JOURNAL: VOLUME 70 (1980) The Marattiales show a high degree of endemism; Macroglossum, for instance, is confined to Borneo. It is logical to assume that smaller spores, such as angio- spores, might be more easily dispersed. On the other hand, smaller spores could be short-lived, reducing their dispersibility and enhancing endemism. The genus Botrychium is cosmopolitan, but throughout its range its species show complex patterns of geographically separated cytological races. This genetic variation may be partly responsible for the taxonomic chaos within the genus. It may be that angiospory is related to these races. - (1) FIG. 1. Normal spore of Macroglossum alidae, x 3075. FIG. 2. 1535. 2] Double-walled spore of M. alidae, x LITERATURE CITED CAMPBELL, D. H. 1911. The Eusporangiatae. Carnegie Inst. Wash. Publ. 140. . 1914a. The genus Macroglossum Copeland. Phil. J. Sci. 9:219-226, t. I. : — The structure and affinities of Macroglossum alidae, Copeland. Ann. Bot. 28:651- See ae B. 1909a. New genera and species of Bornean ferns. Phil. J. Sci. 3:343-352, t. 1909b. The ferns of the Mala ; y-Asiatic region. Part I. Phil. J. Sci. 4:1— 7 ERDTMAN, G. 1957. Pollen and Spor. hil. J. Sci. 4:1-66, 1. I-XXI € Morphology/Plant Taxonomy. Almavist & Wiksell, Stock- holm. KREMP, G. O. W. and T. KAWASAKI. 1972. The Spores of the Pteridophytes. Hirokawa, Tokyo. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 47 Subdivision of the Genus Elaphoglossum JOHN T. MICKEL and LUCIA ATEHORTUA G.* The genus Elaphoglossum is one of the largest and most complex of fern genera. It is composed of perhaps 600 species, which frequently are difficult to distin- guish. The genus has not been afforded careful study in the past. As a result, some names have been misapplied and others not used at all, and sometimes new names have been made for old species. Consequently, much herbarium material is either misidentified or unidentified. ‘‘Elaphoglossum is more in need of a taxonomical revision than other fern genera, all the more so as many species are imperfectly known or badly delimited in comparison with their allies.’’ (Pichi Sermolli, 1968). Elaphoglossum is a remarkably uniform genus of mostly simple-bladed ferns with acrostichoid sori. The veins are free (with two exceptions), the rhizomes scaly, and the blades densely scaly to nearly glabrous. The taxonomy of the genus is based to some extent on frond form and rhizome habit, but more importantly on the scales of the rhizome and blade. Until now there has not been a useful treatment that conveniently breaks the large number of species into smaller, more coherent units. Modern keys have been made for a few areas, such as tropical Africa (Schelpe, 1969), Brazil (Alston, 1956a), Guatemala (Mickel, 1980b), Malaysia (Holttum, 1978), and India (Sledge, 1967), but these do not provide insight into treatment of the genus as a whole. The purpose of this paper is to take the first step in revising the genus by breaking it into infrageneric units that subsequently can be monographed. Even so, this treatment is provisional because a complete understanding of the genus can be had only after the species are better known. HISTORY OF THE GENUS Linnaeus (1753) described Elaphoglossum crinitum under Acrostichum, which included all ferns with sporangia covering the dorsal blade surface. Schott (1834) first proposed the name Elaphoglossum, but it was not formally described until later by John Smith (1841, p. 148), and was not widely accepted until the end of the century. The first broad treatment of the genus was prepared by Fée (1845) under Acro- stichum. He divided the elaphoglossoids into two groups, Oligolepideae and Polylepideae. These in turn were subdivided on the basis of frond size and scale characters. Later, Fée (1852) used four primary groups without further subdivi- sion: Oligolepideae, Polylepideae, Pilosellae, and Chromatolepideae. Moore (1857-1862) was the first to use the genus name Elaphoglossum exten- sively; he made many new combinations under it. He utilized a generic breakdown of Oligolepidum (‘‘fronds naked, or with but few scales’’) and Polylepidum (‘fronds clothed with numerous scales’’). Sodiro (1897, under Acrostichum) used more group names (Glabra, Setosa, Oligotrichia, Polytrichia, Squamosa, Oligolepidia, Laciniata, Polylepidia), but without designating nomenclatural rank. *New York Botanical Garden, Bronx, NY 10458. 48 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Diels (1899) made two sections of the genus: Eu-Elaphoglossum for the bulk of the species and Hymenodium for the distinct, broad, net-veined E. crinitum. His informal subsectional groups largely followed those of Sodiro. In the same year, Christ (1899) published his classic Monographie des Genus Elaphoglossum, which stands as the most detailed treatment of the genus, cover- ing 142 species (and numerous synonyms) in 32 groups. The primary division was based on venation: ordo Stenoneura with veins running all the way to the margin without thickened vein ends and ordo Condyloneura with veins ending just short of the margin with swollen vein ends (hydathodes). Christ further subdivided these into sections, subsections, and divisions, using characters such as blade scales, frond size, stipe articulation, and rhizome habit and thickness. Basically, his species groups were natural and recognizable (nearly all of our ultimate divi- sions are based on them). Unfortunately he placed what we believe to be unrelated divisions and subsections together and did not provide a usable key to the groups. For example, Christ placed some of the species with black, marginal, subulate scales in each of the two ordos, when in fact they are very closely related and none shows the hydathodes of ordo Condyloneura. By the same token, his subsections Dimorpha, Petiolosa, Pilosa, and Ovata in Condyloneura show no signs of hydathodes and belong with close relatives in ordo Stenoneura. This is not to say that hydathodes are a poor character; we would have adopted Christ’s two ordos as subgenera were it not for misgivings about possible convergent evolution of subulate scales and hydathodes or lack thereof. As generally accepted, the species of Elaphoglossum have narrowly elliptic, undivided fronds, but several species of elaphoglossoid affinity with divided fronds are known. Several small, flabellately divided species were first generically segregated by Link (1841) as Peltapteris and pinnately divided plants by Presl (1851) as Microstaphyla. Although some of the American species had the pinnate form of Microstaphyla and had been placed in that genus, Gomez (1975) saw that their true relationship was with Peltapteris. He left only the type species (from St. Helena) in Microstaphyla. One large, pedately divided species, E. cardenasii, was described from Bolivia by Wagner (1954); strangely, no one has proposed a sepa- rate genus for it. Christ placed the species of Peltapteris (as Rhipidopteris) and Microstaphyla under Elaphoglossum and, after showing that frond architecture was the sole character for distinguishing these groups from species of Elaphoglos- sum, Mickel (1980a) concurred in keeping them all in Elaphoglossum. RELATIONSHIPS OF THE GENUS The relationships of the elaphoglossoid ferns to other ferns are unclear. Elaphoglossum is placed close to the lomariopsid genera by some authors because of the acrostichoid sori, largely epiphytic habitat, monolete spores, and chromo- some number of x=41. Holttum (1947) placed it in the Lomariopsidoideae of his Dennstaedtiaceae, and Alston (1956b) placed it in the Lomariopsidaceae. Crabbe, Jermy, and Mickel (1975) placed it in the Elaphoglossoideae of the Aspleniaceae and close to the Lomariopsidoideae. Pichi Sermolli (1968) pointed out the dis- tinctness of the group from the lomariopsids and erected a new family for it, the MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 49 Elaphoglossaceae, but placed this next to the Lomariopsidaceae. Christensen (1938) gave it its own subfamily Elaphoglossoideae in the Polypodiaceae, but could not determine its relationship with certainty. Sporne (1975) and Holttum (1947) have suggested a davallioid relationship be- cause of similarities of stele and chromosome number. We have found some spores which closely resemble those of Bolbitis appendiculata of the lomariopsids (Hennipman, 1977, pl. 4). Many other spores are similar to those of various species of Oleandra and Arthropteris (illustrated by Liew, 1977). This is especially interesting because Oleandra, like Elaphoglossum, has simple blades, straight, free veins, phyllopodia, hydathodes, and chromosome number of x=41. MORPHOLOGICAL CHARACTERS Rhizome habit.—Most species have short-creeping rhizomes, but they range from moderately to long-creeping and from ascending to erect, and are 1-12 mmin diameter. Members of a single subsection usually have a similar rhizome habit, but the entire spectrum can be found within subsection Pachyglossa. Aerophores.—Elaphoglossum aerophores are pale, aerenchymatous out- growths from the lenticel line on the stipe base or adjacent rhizome, one on each side of the stipe. They may appear as fleshy wings about 1 mm broad along the phyllopodia or as tongue-shaped emergences from the stipe base or adjacent rhizome. They are not conspicuous on dried specimens, especially in those species with fasciculate fronds, which makes a survey based on herbarium mate- rial difficult. Lloyd (1970) made a study of living specimens in Costa Rica and found aerophores in 92% of the species studied. More extensive systematic and morphological study of this structure is needed. Phyllopodia.—These dark, sclerified stipe bases are not found in all species and are concentrated in a few subsections, particularly those with coriaceous, sub- glabrous fronds (esp. subsects. Pachyglossa and Huacsaro and sect. Amyg- dalifolia). Pichi Sermolli (1968) distinguished Peltapteris from Elaphoglossum on the basis of the former’s lacking phyllopodia, but Peltapteris’ closest relatives, which are in Elaphoglossum sect. Squamipedia, also lack them. Most sections or subsections either lack phyllopodia or have them only poorly developed. Frond size.—Christ (1899) distinguished some of his glabrous fasion by frond size. To some extent this is valid. Section or subsection members are consistently approximate in size. Subsections may have small and medium fronds or medium and large fronds, but generally not both small and large, although there are di- minutive specimens of nearly all species. There are also very long individual specimens in groups with normally medium-sized fronds (e.g., E. vestitum to 1.7 m long and E. herminieri to 2 m long). Blade shape.—Although the blades are basically simple and unlobed, there is some variation in form. They are fairly uniformly ovate-elliptic in the E. lindenii complex (subsect. Setosa), lanceolate with an obtuse apex in subsect. Muscosa, narrowly elliptic with an obtuse apex in subsect. Huacsaro, and linear-elliptic in subsect. Eximia. Blade shape seems to be more consistent in species groups or series than in sections or subsections. 50 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Venation.—Elaphoglossum fronds are basically free-veined. Elaphoglossum crinitum has net venation and has been set aside as the genus Hymenodium by some authors, but E. crassifolium, which also has net venation, does not seem to be closely related. Elaphoglossum decoratum has an occasional anastomosis, and in several species the vein endings are laterally extended and even fused occa- sionally (e.g., E. acutissimum) to form a commissural vein. In most species the veins diverge from the midvein and go to the margin at an angle of 70—80°. In several sections (esp. Undulata, Setosa, Eximia, and Amygdalifolia), they diverge at a narrower angle (40-60°), are farther apart (ca. 2 mm), and terminate about 1 mm short of the margin, often ending in a round thickening or hydathode. Appar- ently in all species the veins end slightly short of the margin, but in many species the blade is coriaceous with slightly recurved margins, making the actual vein endings difficult to distinguish. Therefore, some authors (including Christ) have claimed that the veins go to the margin. Hydathodes.—The veins in the blade generally end near the margin and are at least slightly swollen. In certain groups, the veins end 1-2 mm short of the margin and are greatly enlarged to form hydathodes. Whether these function differently from those not enlarged is not known. Hydathodes are found in all species of sects. Setosa, Eximia, Undulata, and Amygdalifolia, and are not known from any species of the other sections. These sections comprise what Christ designated ordo Condyloneura. We do not recognize these as subgenera here because it is unclear whether there are close relationships between members of the two ordos, as perhaps between sects. Setosa (with hydathodes) and Polytrichia (lacking them) Blade texture.—Among members of a subsection and closely related groups, texture seems to be nearly uniform. Coriaceous fronds generally are nearly gla- brous, but whether all species with such fronds are related is questionable. The absence of blade scales makes it difficult to assess relationships. Most species have herbaceous to firm laminae, but those with hydathodes are usually thin. Rhizome scales.—The rhizomes are generally densely scaly, with linear, lan- ceolate, ovate, or rarely even round scales, which are attached at a cordate to peltate base. Scale color ranges from bright orange to maroon, brown, or black. Texture varies from thin to sclerotic. Scale margins in most species are entire or have weak teeth, but in sect. Lepidoglossa they often are fringed with slender, hair-like teeth. Differences in cell patterns have not yet been examined systemat- ically, but probably will be helpful taxonomically. In some groups within sect. Elaphoglossum, the rhizome scales may be deciduous, leaving in some cases completely glabrous creeping rhizomes. In a few species the naked rhizome is black and glutinous. Blade scales.—There is great diversity in scale morphology, but the blade scales are basically lanceolate. They may be erose-margined (E. muscosum), entire (E. decoratum), or deeply ciliolate-toothed (subsect. Polylepidea). Some are so deeply ciliolate-toothed there is barely any scale body (E. vestitum) or are reduced to stellate hairs (E. pilosum). In E. tectum, the stipe and upper blade scales are MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 51 round-peltate with erose margins, whereas those of the lower blade surface are stellate hairs. In E. petiolatum, E. huacsaro, and their relatives, the blade scales are further reduced to resinous dots. Coriaceous, subglabrous groups often have reduced, minute, black, linear or stellate blade scales. There are no multicellular hairs in Elaphoglossum; those thought to have hairs (sects. Setosa and Polytrichia) have in reality hair-like subulate scales which are linear-lanceolate with inrolled margins. It is not clear whether all groups with subulate scales are related or not. Although they have distinctive, minute, glandu- lar hairs in common, they differ in their spores. Stipe scales.—They are like those of the rhizome at the base and like those of the blade near the apex. Sometimes these are transitions from one to the other type, but in other cases, two distinct types are intermixe Glandular hairs.—Minute, erect, unicellular, sland. tipped hairs are located on the stipe in sects. Polytrichia, Setosa, and Eximia. Their presence seems to be correlated with the presence of subulate scales. In subsect. Apoda, these hairs also are found on the blade surface. Fertile fronds.—In most species, the fertile fronds have narrower and shorter blades and proportionally longer stipes than the sterile ones. The fertile fronds may be longer, essentially equal, or noticeably shorter than the sterile ones. The relationship does not seem to be consistent in whole sections or subsections, but is so in species groups or occasionally in subsections (fertile longer in subsects. Huacsaro and Pilosa). In some species, the fertile frond is folded in half lengthwise along the midvein until maturity. This conduplicate fertile blade is found in E. lindenii and E. pilosel- loides of sect. Setosa, but not in all species of the section; in subsect. Petiolosa, both species show this condition. The indument of the fertile frond generally is the same on the upper surface as on that of the sterile blade. However, there are a few species that have scales on the lower surface among the sporangia, and their presence is especially notewor- thy. It is not consistent with subsections, occurring in E. villosum, E. muscosum, E. siliquoides, all of different subsections, but seems to be common to all mem- bers of subsect. Plumierana. In certain species, there seems to be a conspicuous sterile margin, that forms a pale border 1-2 mm wide outside the sporangial mass. It is especially noticeable in some thin-textured species such as E. albomarginatum. Although it is not readily apparent in most other species, close examination shows that the sterile margin is underrolled and thus concealed. Probably some sort of sterile margin exists in most or all species, which is to be expected since the veins do not reach the margin. Spores.—Elaphoglossum spores usually have been described as having narrow crests with minute spicules on the surface (Erdtman, 1957, based on E. vieillardii from New Caledonia), but our survey of spores using the scanning electron micro- scope has shown considerable variation within the genus. We have examined 163 species, including representatives of all sections and subsections. Figures 1-18 are based on specimens in the New York Botanical Garden herbarium. 52 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Spores of most species have slender crests or low folds or ridges. Section Squamipedia (Fig. 7) also has low ridges, but the entire surface is densely or- namented with distinctive spicules found also in the flabellately or pinnately di- vided subsect. Peltapteris. High slender crests are found in sects. Amygdalifolia, Decorata, and in some members of sects. Setosa and Elaphoglossum. Subsection Pachyglossa, probably the largest in the genus, displays several different spore types: some species have tall, nearly spine-like, smooth or perforated crests, to highly fenestrated, lace-like crests; others have long, slender crests lacking holes; and others have low, broad ridges. Spores with low, broad ridges are found in sects. Lepidoglossa (Fig. 9), Poly- trichia (Fig. 12), and part of sect. Elaphoglossum subsect. Pachyglossa (Fig. 3). The ridges tend to be short (Figs. 9 and /2) rather than elongate, and often have small verrucae in the valleys between the ridges, which suggests a possible rela- tionship among these groups. On the other hand, the lack of other ornamentation could indicate either a common or a primitive spore type or a condition arrived at independently. Correlations of additional characters are needed to resolve this. Spores of subsect. Muscosa are unique in having verrucate surfaces (Fig. //). The spores of sects. Eximia, Undulata, and part of sect. Setosa subsect. Setosa lack ridges or crests of any sort. Instead, they have a dense covering of short spines which branch at the base to form a reticulum (Figs. 14 and /6-/8). In sect. Undulata, the spines occasionally fuse to form small crests (Fig. 18). In E, beaurepairii (sect. Eximia, Fig. 17) the reticulum is very Open, whereas in sect. Setosa subsect. Setosa (E. crinipes et aff.) the branches often fuse laterally to make a more dense covering on the spore surface (Fig. 14). Spores of some members of subsect. Setosa (E. lindenii et aff.) and Alpestria have low, slender crests which often have holes of various sizes in them (Fig. 13), somewhat like those mentioned for subsect. Pachyglossa. Holes also occur in the spore surface, which in some cases resembles somewhat the spore surface reticulum of sects. Undulata and Eximia. Perforations, at least in the body of the spore, also are found in some species of subsect. Pachyglossa, but whether this indicates a rela- tionship has not yet been determined. Not every species fits into the spore pattern for its group. Some species of subsect. Pilosella have holes in the surface and others do not. In subsect. Plumierana, E. buchii (Fig. 15) and E. plumieri have solid crests, whereas the morphologically similar E. lanceum has highly fenestrated crests. In general, however, spore morphology has been quite useful in confirming relationships presumed on other grounds and in Suggesting new phyletic interpretations. CONSPECTUS OF THE GENUS ELAPHOGLOSSUM (INCLUDING CHRIST’S INFRAGENERIC GROUPS IN PARENTHESES) sect. Elaphoglossum (sect. Craspedoglossa) MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 93 sect. hi emia Mickel & Atehortuta sect. Squamipedia Mickel & Atehortua (subsect. Pachyglossa div. Squamipedia) pa oa aghiapten (Link) Mickel & Atehorttia (subsect. Pachyglossa div. Rhipidopteris) subsect. Ova subsect. Bt Christ sect. Decorata Mickel & Atehortta (subsect. Platyglossa div. Decorata) - sect. Lepidoglossa Christ (sect. Gymnoglossa) subsect. Polylepidea Christ (subsect. Polylepidea div. Auricoma) a Microlepidea Christ (subsect. Microlepidea div subsect. Pilosa Christ (subsect. Pilosa div. Grata; subsect: Microlepidea div. Viscosa; subsect. cans lepidea div. Stipitata; subsect. Dimorpha) subsect. Petiolosa Chris subsect. Huacsaro Ee & Atehortua aba Muscosa Mickel & Atehortia (subsect. Polylepidea divs. Muscosa and Bellerman- niana) sect. eter Chris subse emeilouiok (Fée) Christ eee Hybrida Christ (subsect. Platyglossa div. Melanolepidea; subsect. Hybrida by type and description but not by other included species) subsect. Apoda Mickel & Atehorttia sect. Setosa (Christ) Mickel & Atehortua subsect. Alpestria Mickel & Atehortta ec Plumierana Mickel & Atehortta sect. Eximia Mickel & Atehortua subsect. Eximia Mickel & Atehorttia (subsect. Hybrida by included species but not by type or escription) subsect. aca Mickel & Atehortua sect. Wali sect. Amygdalifolia (cist Mickel & Atehortta The following of Christ’s groups are of unknown relationship: subsect. Glos- soides, subsect. Coespitosa, subsect. Pachyglossa div. Micradenia, subsect. Polylepidea div. Argyrophylla, subsect. Polylepidea div. Fimbriata, subsect. Pilosa div. Boraginea, and subsect. Pilosa div. Gardneriana. KEY TO THE SECTIONS OF ELAPHOGLOSSUM 1. Veins ending short of the margin, enlarged at the tip to form generally conspicuous pneeene 2. Rhizome long-creeping; blade glabrous; phyllopodia present, but short. ...... sect. mall 2. Rhizome short- to long-creeping; blade scaly, only rarely glabrous; gral aaa 3. Blade linear-elliptic or pedately divided: blade scales very small, dark, lanceolate, not ig ane sect. Eximia 3. Blade narrowly iced or ovate-lanceolate; blade scales subulate or cordate- sr once pale. 4. Blade scales subulat t. Setosa 4. Blade scales tdi ene. erose or toothed sect TF lakes 1. Veins ending at or very close to si margin, not ending in hydathodes. 5. Blade glabrous or subglabro 6. Rhizome long-creeping, font very small (2-20 cm long), usually lacking lg anes blade simple to finely dissected. Squamipedia 6. Rhizome erect or short- to clei ade fronds small to large (12-200 cm m lone) phyllopodia distinct or indistinct but always present t. Elaphoglossum 94 AMERICAN FERN JOURNAL: VOLUME 70 (1980) ong 1-6. Spores of Elaphoglossum sect. oe FIG, 1. E. pteropus ipo 19180), x ac it leg a e598 ai Pa Ae IG. 3. E. dussii rie ay $.n.), FIG. 4.E ra ; Wacket 7849), pg ), * 1200. FIG. 5. E. bicolor (Wacket 121), x 1560. FIG. i E. wacketii MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 95 be Pee aly. 7. Blade scales subulate. sect. Polytrichia 7. Blade scales lanceolate, reniform, round-peltate, or stellate, not subulate. 8. Blade scales only at frond margin and along midvein. sect. Decorata 8. Blade scales not restricted to frond margin and midvein sect. Lepidoglossa 215" Elaphoglossum sect. Elaphoglossum 258}° Sect. ee ossa Christ, Monogr. Elaph. 20. 1899. e long- to short-creeping or suberect; rhizome scales linear to lanceo- late; shvllosodie distinct or indistinct; blade glabrous or subglabrous, coriaceous to very thin; hydathodes lacking; blade scales minute, laciniate-pectinate to stel- late; spores With various forms of crests or low ri YPE SPECIES: Acrostichum conforme Swartz [= ean conforme (Swartz) Schott ex J. Smith] KEY TO THE SUBSECTIONS OF SECT. ELAPHOGLOSSUM 1. Blade coriaceous; veins obscure; phyllopodia conspicuous subsect. Pachyglossa 1. Blade very thin, translucent; veins evident; phyllopodia inconspicuous........... subsect. Tenuifolia _ 21585 Elaphoglossum subsect. Pachyglossa Christ, Monogr. Elaph. 20. 1899. ued 28% \Subsect. cineeie! Christ, Monogr. Elaph. 20. 1899. CECTOTYPE (chosen here): Acrostichum latifolium Swartz [=Elaphoglossum latifolium (Swartz) J. Smith]. Christ did not designate a t species for pie subsection, but chose E. latifolium as the type of his first and most typical division, atotula. 2597 — Flaccida Christ, Monogr. Elaph. 20. 1899. Subsect. Herminieriana Christ, Monogr. Elaph. 21. 1899. eee of the section, with the blade coriaceous and the phyllopodia espe- cially distinct. AYPE SPECIES: Acrostichum conforme Swartz [=Elaphoglossum conforme (Swartz) Schott ex J. Smith]7~ !°97%° SELECTED SPECIES EXAMINED: Elaphoglossum acrostichoides (Hook.) Schelpe, E. acutifolium Rosenst., E. affine (Mart. & Gal.) E. angulatum (Blume) Moore, E. angustatum (Schrad.) Hieron, *E. bicolor Rosenst., E. callifolium (Blume) Moore, E. chartaceum (Baker) C. Chr., E. crassifolium (Gaud.) Anderson & Crosby, E. crassinerve (Kunze) Moore, *E. funckii (Fée) Moore, *E. gayanum (Fée) Moore, E. glabellum J. Smith, E. glaucescens Rosenst., E. glaucum (Fée) Moore, *E. glossophyllum Hieron., E. guatemalense Klotzsch, E. herminieri (Bory & Fée) Moore, E. hoffmannii (Mett.) Hieron, E. hymenodiastrum (Fée) Brade, E. inaequalifolium (Jenm.) C. Chr., E. latifolium (Swartz) J. Smith, E. a (Fée) Moore, *E. lingua (Presl) Brack., E. longifolium (Presl) J. Smith, E. maxonii x Morton, E. pteropus C. Chr., E. rigidum (Aubl.) Urban, E. schiffneri Christ, E. schom- sehr (Fes) Moore, E. simplex Nagi J. Smith, E. sporadolepis (Kunze) Moore, E. oo (Liebm.) Moore, E. tovarense (Moritz ex D. C. Eat.) Moore, E. tuckerheimii Brause, *E. vagans (Mett.) Hieron., and E. wawrae (Luerss yc Chr. This is the largest and taxonomically most complex subsection of the genus. There are relatively few available diagnostic characters, and the varia- tion within and between species is difficult to interpret. Several groups are vaguely discernible, but whether they are all closely allied or have evolved to the coriaceous, glabrous condition independently is still in question. We have re- frained from distinguishing taxonomic groups until more information is gathered. One group has very short to suberect, stout (4-12 mm diam.) rhizomes and spores with many short crests that resemble broad spines (Fig. 2). In some of 56 AMERICAN FERN JOURNAL: VOLUME 70 (1980) oage ane oe of Elaphoglossum. Sect. Squamipedia. FIG. 7. E. squamipes (Wurdack 692), ; roi eet ow 540), x 1200. Sect. Lepidoglossa. FIG. 9. E. paleaceum (MacBride 1uacsaro ey nct Annie FOF) 1200. FIG. 11. E. bellermannianum (Camp E 3938B), x 960. FIG. 12. E. auripilum (Mickel 2657), x 1800. ate sae MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 37 these there are perforations in the crests and in the spore surface as well (Figs. 4 and 5). On the other hand, some species (e.g., E. pteropus) have spores with tall, slender, solid crests (Fig. /). Many species have more slender rhizomes (2-4 mm diam.) that have dark, sclerotic rhizome scales which may become deciduous, as in E. guatemalense and E. glabellum. These have spores with low, broad ridges (Fig. 3 There is a species complex, including E. affine, E. tenuifolium, E. leptophyllum, and E. schiffneri, which has ovate-lanceolate, orange to tan rhizome scales with varying degrees of dark, sclerotic streaks in them. Their rhizomes range from short-creeping to rather long-creeping. Quite possibly this complex is related to what looks like a distinct group with very long-creeping, slender, cord-like rhizomes with black, ovate-lanceolate rhizome scales (species marked with an asterisk in the above list). Elaphoglossum hoffmannii has the most slender, nearly naked rhizomes and smallest fronds of the group and may not belong here. Elaphoglossum subsect. Tenuifolia Mickel & Atehortua, subsect. nov. Rhizomata gracilia, saepe nuda; phyllopodia inconspicua; lamina angusta tenuissima; nervi visibiles; sporae dense spiculatae cristatae TYPE SPECIES: Elaphoglossum acutissimum Christ. - Fibs SELECTED SPECIES EXAMINED: Elaphoglossum burchellii (Baker) C. Chr., E. slay dase (Jenm.) Urban, E. praelongum (Fée) C. Chr., E. sherringii (Baker) C. Chr., and E. wacketii Rose This group is distinct in its very Gis eshived blades and often naked or even black and glutinous rhizome. The glabrousness of the blade induces us to place these species in sect. Elaphoglossum, but they may well belong elsewhere, perhaps nearer to subsect. Pilosa. At least one member of this group has resinous dots on the dorsal surface of the blade and spiculate spores (Fig. 6), such as are found in E. huacsaro (Fig. 10) in subsect. Huacsaro and in some members of subsect. Pilosa, but the naked rhizome and glabrous blade resemble subsect. Pachyglossa. Elaphoglossum gramineum has a glabrescent rhizome and resin-dotted blade, but the blade is coriaceous, unlike that of other members of this subsection. Subsection Flaccida was intended by Christ to include the thin-bladed species we treat as subsect. Tenuifolia, but unfortunately he selected as type E. flac- cidum, which generally is regarded today as a synonym of E. rigidum, a member of the coriaceous subsect. Pachyglossa. ge eam sect. Squamipedia Mickel & Atehortua, sect. nov.” Rhiz a gracillima longe repentia; rhizomatis stipitisque paleae ovato- iliceolntes. eaten vel lacerato-pectinatae; phyllopodia rara; lamina parva; laminae paleae parvae fuscae reductae, saepe hastatae; nervi inconspicul; hydathodi nulli; sporae plerumque multispiculatae cristis demissis latis anas- tomosantibus ornatae. ? Although the sectional names Squamipe edia, Setosa, Ber Eximia repeat subsectional names, Latin — oses are given for both levels since at the subsect type subsections, and some botanists might not seam he subsectional names as properly described without separate Latin diagnoses. 58 AMERICAN FERN JOURNAL: VOLUME 70 (1980) FIGS. 13-18. Spores of Elaphoglossum. Sect. Setosa. FIG, | 1402. crinipes (Standley 85982), x 1200. FIG. 15. E. scan (Jenman. s.n.), x 1200. Sect. Eximia. - E. eximium (Mickel 3417), x 1080. FIG. Sect 17. E. beaurepairei (Brade 20937), x 1200. FIG. 18. E. hirtum (Mickel 2978), x 1200. E. moritzianum (Fendler 362), x 1440. FIG. U, adulata arty MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 59 TH TY PE SPECIES: Acrostichum squamipes Hook. |=Elaphoglossum squamipes (Hook.) Moore}. KEY TO THE SUBSECTIONS OF SECTION SQUAMIPEDIA 1. Stipe scales deeply lacerate-pectinate subsect. Ovata 1. Stipe scales entire, ovate- aes lat 2. Sterile blade pinnately or Aabelieesl? divided, rarely undivided but then es: flabellat bsect. Aciedies 2. Sterile blade entire, linear-elliptic or oblanceolate. 3. Blade margin crenulate; blade thin; spores with narrow crests subsect. Feeana 3. Blade margin entire; blade coriaceous; spores with low folds or oy ot e — icules. sect. fe area Elaphoglossum subsect. Squamipedia Mickel & Atehortua, bie no Rhizomatis stipitisque paleae ovato-lanceolatae peltatae cakes lamina simplex; sporae spiculatae (Fig. 7). ATYPE SPECIES: Acrostichum squamipes Hook. |=Elaphoglossum squamipes (Hook.) Moore}. OTHER SPECIES EXAMINED: Elaphoglossum cardiophyllum (Hook.) Moore, E. craspedariiforme (Fée) Brade ex Alston, E. deltoideum (Sod.) Christ, E. lloense (Hook.) Moore, and E. revolutum (Liebm.) Moore This is one of the most distinctive groups in the genus, with its small fronds, broadly ovate rhizome and stipe scales, and lack of phyllopodia. In blade form E. cardiophyllum seems to belong here, but phyllopodia are present and spore spicules are lacking. This subsection is extremely closely related to subsect. Pel- tapteris and agrees in all characters except frond architecture. 5%) [Elaphoglossum subsect. ane oo Mickel & ee stat. nov. ~st}4 Peltapteris Link, Fil. Sp. Hort. Reg. eet Berol. em 147. 1841 Rhipidopteris Schott ex Fée, ag gf ta . 1845, nom. illeg. Similar to subsect. RE iiecatia: but at Serie blade pinnately or flabellately divided, rarely undivided but flabellat TYPE SPECIES: Osmunda ante Swartz [=Elaphoglossum peltatum (Swartz) Urban]. OTHER SPECIES EXAMINED: Elaphoglossum ie seer rr este dager - moorei (E. Britt.) Christ, E. peruvianum (Gomez) Mickel, and E. tripartitum (Hook. & Grev.) M This group is usually treated as a ect genus (Gomez, 1975; Morton, 1955), but can be distinguished from species of subsect. Squamipedia only by the frond dissection. Elaphoglossum subsect. Ovata Christ, Monogr. Elaph. 23. 1899. Similar to subsect. Squamipedia, but the scales of the blade, stipe, and rhizome deeply lacerate-pectinate; phuilapedia distinct; spores with broad folds or ridges but lacking spicules (Fig. 8). TYPE AND SOLE SPECIES: Acrostichum ovatum_ Hook. & Grev. [=Elaphoglossum ovatum (Hook. & Grev.) Moor rs, ~ 12974 nd asing sect. D d sub dia and Muscosa on Christ's divisions egies sts some Sot tanists would argue that “ cited divisions pa have been described in Latin to be valid. Since Christ’s descriptions were in German, we here supply Latin po to avoid any confusion. 60 AMERICAN FERN JOURNAL: VOLUME 70 (1980) In frond form and slender rhizome this looks strikingly like E. sguamipes, but the deeply lacerated scales are quite distinct. oP Elaphoglossum subsect. Feeana Christ, Monogr. Elaph. 22. 1899, ronds linear-lanceolate or oblanceolate, thin, with crenulate margin; phyl- lopodia short but distinct; spores with crests and sparse to abundant spicules. TY PE SPECIES: Acrostichum feei Bory [=Elaphoglossum feei (Bory) Moore]. OTHER SPECIES EXAMINED: q ro88t Elaphoglossum procurrens (Mett.) Moore and E. wrightii (Mett.) Moore. These species are close to subsect. Squamipedia in the very slender, long- creeping rhizome and entire scales, but the thin blade with crenulate margin and spores with narrow ridges or crests distinguish them as a separate group. Rg sect. Decorata Mickel & Atehortua, sect. noy.? hizomatis erecti crassi paleae lineares; phyllopodia brevia; stipitis paleae magnae patulae; laminae magnae ellipticae praeter costam Mmarginesque paleis imbricatis aureis vestitas glaberrimae; nervi conspicui hinc inde anastomosantes; hydathodi nulli; sporae breviter cristatae. TYPE AND SOLE SPECIES: Acrostichum decoratum Kunze | =Elaphoglos- sum decoratum (Kunze) Moore]. Lays This species is remarkably distinct and without Close relatives. \ ee Elaphoglossum sect. Lepidoglossa Christ, Monogr. Elaph. 21. 1899, ger Sect. Gymnoglossa ea ey sash Elaph. 22. 1899. LECTOTYPE (chosen here): Osmunda bifur- cata Jacq. [=Elaphoglossum ifurcatum (Jacq.) Mickel]. Christ did not designate a type for sect. Gymnoglossa, but he did choose E. furcatum ( syn. E. bifurcatum) as type of his subsect. Dimorpha, which is the first named and most typical subsection of sect. Gymnoglossa. 7 ad . s toothed, ciliolate or rarely entire, or round-peltate, or modified to stellate hairs; veins evident to obscure, free; hydathodes lacking; spores with low ridges, rarely ridges. 4148 ee paleaceum (Hook. & Grev.) Sledge] (syn. Acrostichum sSquamosum KEY TO THE SUBSECTIONS OF SECTION LEPIDOGLOSSA 1. Blade densely scaly; scales lanceolate, toothed, not stellate or reduced to resinous dots. 2. Blade Ovate-lanceolate, coriaceous, often obtuse at apex; blade scales erose; spores verruculate. 2. Blade linear-lanceolate or narrowly elliptic, acuminate a Spores with low, non-verruculate Leip. ek noe: subsect. Polylepidea modified to stellate hairs and/or resinous dots Itate. ha oeatheslecdeveseesaty rts oi subsect. Microlepidea MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 61 4. Blade glabrous, elliptic with a caudate apex. subsect. Petioloa 4. Blade scaly to subglabrous, linear to lanceolate, acuminate to obtuse at apex 5. Rhizome short-creeping; blade acute to acuminate, rarely crenulate to pinnate divided. sect. Pilosa 5. Rhizome long, ascending; blade obtuse. See Huacsaro 5k? Elaphoglossum subsect. Polylepidea Christ, Monogr. Elaph. 21. 1899. Rhizo short-creeping to ascending; rhizome scales dark, ciliolate; blade scales a ciliolate; blade narrowly elliptic, densely ay spores with low ridges, usually with small verrucae between the ridges (Fig. 9). “LECTOTYPE SPECIES (chosen here): Acrostichum paleaceum, Hook. & Grev. [=Elaphoglossum paleaceum (Hook. & Grev.) Sledge]. — #!9 SELECTED SPECIES EXAMINED: Elaphoglossum acuminans C. Chr. ex Urban, E. auricomum Seapey Moore, E. casanense Rosenst., E. deckenii (Kuhn) C. Chr., E. dombeyanum Moore & Houlst., E. eatonianum (E. Britt.) C. eggersii (Baker) Christ, E. fuertesii Brause, E. vain ype E. kuhnii Hieron., E. laminarioides (Bory) Moore, E. langsdorfii (Hook. & Grev.) Moore, E. meridense ies Moore, E. orbignyanum (Fée) Moore, E. plumosum (Fée) Moore, E. eens (Kuhn) Christ, E. rufescens (Liebm.) Moore, E. vestitum (Schlecht. & Cham.) Schott ex J. Smith, and E. wagneri (Kunze) Moore This is one of the two largest, most complex, and unmanageable groups in the genus. The toothed or ciliolate scales on the blade and usually also on the rhizome are the principal distinguishing features. 2589" Elaphoglossum subsect. Microlepidea Christ, Monogr. Elaph. 22. 1899. Rhizome scales Bneay auceoine. entire; vac’ scales stellate below, round and peltate above; spores with low, oth r idge “LECTOTYPE SPECIES Taina here): yee tectum Humb. & Bonpl. 7445 ex Willd. [=Elaphoglossum tectum (Humb. & Bonpl. ex Willd.) Moore]. Christ selected this species as type of his div. Tecta, which is the first-named and most typical of the divisions under subsect. Microlepidea. OTHER SPECIES EXAMINED: Elaphoglossum furfuraceum (Mett.) Christ. This subsection is very close to subsect. Pilosa, but the round, peltate scales are not found in any other group. 258aC- Teper subsect. Pilosa Christ, Monogr. Elaph. 23. 1899. staphyla Presl, Epim. Bot. 160. 1851 LSev -Elapodlssum subsect. Dimorpha Christ, Monogr. Elaph. 22. me short-creeping; rhizome scales usually on. entire to pinnately di- vided; bhade salen resembling stellate hairs or lanceolate-toothed; blade often with resinous dots; spores with low ridges or narrow crests. -LECTOTYPE SPECIES (chosen here): Acrostichum pilosum Humb. & Bonpl ex Willd. [=Elaphoglossum pilosum (Humb. & Bonpl. ex Willd.) Moore]. Elaphoglossum pilosum is the type species of Christ’s div. Grata, which was the first and most typical division of subsect. Pilosa. SELECTED SPECIES EXAMINED: Elaphoglossum bifurcatum (Jacq.) Mickel, E. dimorphum (Hook. & Grev.) Moore, E. gratum (Fée) Moore, E. lagesianum Rosent., E. lepidotum J. Smith, E. mathewsii (Fée) Moore, E. nervosum (Bory) Christ, E. petiolatum (Swartz) Urban, E. rosenstockii Christ, E. salicifolium (Willd. ex Kaulf.) Alston, and E. viscosum (Swartz) J. Smith. "920 | 62 AMERICAN FERN JOURNAL: VOLUME 70 (1980) The degree of scaliness and variation from scales to resinous dots within one species is not well understood. This subsection is very closely allied to subsects. Huacsaro and Microlepidea, and the demarcation between them is not clear. Elaphoglossum dimorphum displays an intermediate frond morphology be- tween simple fronds, such as those of E. nervosum, and the pinnately dissected fronds of E. bifurcatum. All three species occur on St. Helena, and Mickel (1980a) has shown that they differ essentially only in dissection. Whether these are in fact three distinct species or all forms of one species is still a question. aoe Elaphoglossum subsect. Petiolosa Christ, Monogr. Elaph. 23. 1899. Rhizome short-creeping; rhizome scales dark, linear, bristle-like; blade elliptic with caudate tip; fertile blade folded; blade scales lanceolate or lacking, occasion- ally with resinous dots; spores with low ridges. ¥e iver VTYPE SPECIES: Acrostichum petiolosum Desv. [=Elaphoglossum petiolosum (Desv.) Moore}. OTHER SPECIES EXAMINED: Elaphoglossum trianae Christ. The relationships of this subsection are not clear. The resinous dots are re- miniscent of those of subsects. Pilosa and Huacsaro, the bristle-like rhizome scales are similar to those of some members of subsect. Hybrida, and the condup- licate fertile blades are similar to those found in some members of subsect. Setosa. The blade shape and spore details are unique. 9G) ae Elaphoglossum subsect. Huacsaro Mickel & Atehortua, subsect. nov. zomata adscendentia longa; rhizomatis paleae fuscae integrae; lamina ellip- tica, apice obtusa; laminae saepe resinoso-punctatae; paleae lanceolatae den- ticulatae vel pilis stellatis conspersae; sporae dense spiculatae late cristatae. “ae PE SPECIES: Acrostichum huacsaro Ruiz [=Elaphoglossum huacsaro c (Ruiz) Christ]. e 7" Pica oe SPECIES EXAMINED: aphoglossum alfredii ili dence Ce ten Roses and acne eke en ml Mose This group probably is closely related to subsect. Pilosa, as shown by the scales and resinous dots, but the long ascending rhizome, obtuse blade apex, and the highly spiculate spores are distinctive (Fig. 0). ase Elaphoglossum subsect. Muscosa Mickel & Atehortua, subsect. noy.? paar ar a Magee repentia; phyllopodia nulla; stipites laminaeque dense ovate lance ane Is paleae latae patulae; lamina anguste elliptica ad lanceolata vel a uy COrlacea, apice plerumque obtusa; paleae saepe cum sporangiis atae. a1 “TYPE SPECIES: Acrosti : chum muscosum Swa (Swartz) Moore]. as SELECTED SPECIES EXAMINED: Elaphoglossum aschersonii Hieron., E. bellermannianu e. corderoanum (Sod.) Christ, E. decipiens Hieron., E. Christ, and E. plicatum (Cav.) C. Chr. Although the scales of E. engelii approach those of the subsect. Polylepidea, the spores are unique in the genus in having verruculae covering the surface (Fig. 1/). bi he =Elaphoglossum muscosum m (Klotzsch) Moore, E. blandum Rosenst., engelii (Karst.) Christ, E. leamannianum | si 25345, MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 63 Elaphoglossum sect. Polytrichia Christ, Monogr. Elaph. 22. 1899. Rhizome short- -creeping to erect; rhizome scales linear to linear-lanceolate; phyllopodia inconspicuous or lacking: stipe and blade, especially blade midvein and margin, with subulate scales and also bearing minute glandular hairs; hy- dathedee lacking; spores with low ridges and often small verrucae between them. “LECTOTYPE SPECIES (chosen here): Acrostichum crinitum L. [=Elapho- > glossum crinitum (L.) Christ]. This was the type of Christ’s first-named and most typical subsection, Hymenodium. KEY TO THE SUBSECTIONS OF SECTION POLYTRICHIA 1. Blade broadly elliptic and fleshy; veins netted, obscur subsect. Hy see ee 1. Blade pee et narrowly elliptic, or ovate- ete subcoriaceous or papyraceous; vein free, usually eviden 2. Stipe short to si blade scales dark brown to black, located mostly on the margin and midvein. subsect. Hybrida 2. Stipe nearly lacking; blade scales orange, distributed subuniformly over the blade surfaces. subsect. Apoda Elaphoglossum subsect. Hymenodium — Christ, Monogr. Elaph. 23. 1899. Hymenodium Fée, Mém. Fam. Foug. 2: 20. 1845. eae ne fleshy, oe ‘slip veins obscure, netted; spores with owr TYPE eae SOLE SPECIES: Acrostichum crinitum L. [=Elaphoglossum crinitum (L.) Christ]. Although this is occasionally distinguished as a separate genus, it agrees very Closely with subsect. Hybrida. 25246 Elaphoglossum subsect. Hybrida Christ, Monogr. Elaph. 23. 1899. izome short-creeping or ascending; stipe long; blade usually 1 stool ous; blade scales especially on the margin and midvein, black or da rk bro spores with low ridges (Fig. /2). — 1S ey “TY PE SPECIES: Acrostichum hybridum Bory | =Elaphoglossum hybridum (Bory) Moore]. SELECTED SPECIES EXAMINED: Elaphoglossum albomarginatum A. Reid Smith, E. auripilum Christ, E. cordifolium Rosenst., E. denudatum (Jenm.) Maxon ex Morton, E. erinaceum (Fée) Moore, E. lindbergii (Mett.) Rosenst., E. melanopus (Kunze) Moore, E. prestonii J. Smith, E. scolopendrifolium (Raddi) J. Smith, E. spannagelii Rosenst., and E. tambillense (Hook.) Moore. This subsection is very complex, and the species limits are not well under- stood. The variation in rhizome scales is especially perplexing. Most species have linear, orange rhizome scales, and others have bristle-like, maroon scales, but the differences are not always clear-cut. At least two species have a glabrous or subglabrous blade. This subsection is composed mostly of Christ’s ‘‘divisio Melanolepidea™ of ordo Stenoneura, since they lack hydathodes. Christ referred subsect. Hy- brida to ordo Condyloneura, although E. hybridum, the type species, lacks hydathodes and belongs to div. Melanolepidea. All other species Christ in- cluded in subsect. Hybrida have hydathodes and make up our subsect. Eximia. _ 2244 64 AMERICAN FERN JOUNRAL: VOLUME 70 (1980) nid o> Elaphoglossum subsect. Apoda Mickel & Atehortua, subsect. nov. Stipites fere nulli; laminae paleae aurantiacae subulatae, per laminae ssa ficiem regulariter conspersae; sporae breviter cristatae. ras AY PE SPECIES: Acrostichum apodum Kaulf. [=Elaphoglossum apodum (Kaulf.) Schott ex J. Smith]. OTHER SPECIES EXAMINED: Elaphoglossum cubense (Mett. ex Kuhn) C. Chr. and E. siliquoides (Jenm.) C. Chr. Members of this subsection closely resemble those of subsect. Setosa in their orange to brown subulate blade scales, but seem to belong to sect. Polytrichia on the basis of no hydathodes and spores with low ridges and .. perforated crests. They also differ from sect. Setosa in their very short stipes. iW Elaphoglossum sect. Setosa (Christ) Mickel & Atehortua, stat. nov.” 1b">) Elaphoglossum subsect. Setosa Christ, Monogr. Elaph. 23. 1899. izome short- to long-creeping or erect; rhizome scales linear; phyllopodia lacking; stipes with minute, erect, glandular hairs; plants mostly small; veins evident, spaced well apart, ending well short of the margin in distinct hydathodes; scales subulate, orange to brown; spores with many low crests and usually with a perforate surface, or not crested and the surface echinate-» reticulate. 25 TYPE SPECIES: Acrostichum villosum Swartz | =Elaphoglossum villosum (Swartz) J. Smith]. KEY TO THE SUBSECTIONS OF SECTION SETOSA 1. Rhizome long-creeping; rhizome scales dark brown to black subsect. Alpestria 1. Rhizome ascending to erect: rhizome scales pale to dark brown. 2. Blade margin usually crenulate. 2. Blade margin entire. 3. Plants very small (2-15 cm tall): blade spatulate, rarely sublinear; hydathodes inconspicuous. subsect. Pilosella ); blade narrowly elliptic to linear-lanceolate: veins Pisin tees seine tev akan andenstescaey wik iG ur) itso subsect. Setosa 2¢\ 4) Elaphoglossum subsect. Setosa Christ, Monogr. Elaph. 23. 1899. hizome ascending to erect; blade narrowly elliptic to linear-lanceolate; veins and hydathodes evident: Spores with low crests (Fig. 13) or spines (Fig. /4). “TYPE SPECIES: Acrostichum villosum Swartz [=Elaphoglossum villosum (Swartz) J. Smith]. L_ 3sst\ SELECTED SPECIES EXAMINED: Elaphoglossum costaricense Christ, E. lindenii (Bory ex Fée) Moore, E. moritzianum (Klotzsch) oore, E. ocoense C. Chr., E. omphalodes (Fée) Brade, E. Palorense Rosenst., and E. setosum Moore. ree thicss tiseb eles see Oro subsect. Plumierana 3. Plants small to medium-sized (540 cm tall and hydathodes evident. Some other species, such as E. crinipes C. Chr., E. oblanceolatum C. CAG, Ee. papillosum (Baker) Christ, and E. setigerum (Sod.) Diels, are included here be- cause they look like species of subsect. Setosa in their external morphological characters, but they differ significantly in Spore architecture. Rather than having crests, their spores are densely covered with Short spines whose bases branch to form a reticulum (Fig. 14). Elaphoglossum fluminense Brade may belong here also. It looks like a slender member of subsect. Pilosella, but has perforate spores with very low crests like those of most members of subsect. Setosa. ast 2890 MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 65 0 Elaphoglossum subsect. Pilosella Christ, Monogr. Elaph. 23. 1899. hizome erect; scales subulate; hydathodes inconspicuous; plants especially small (2-15 cm tall); spores non-perforate with the ridges low and broad, usual lacking spicules. ~ “TYPE SPECIES: Acrostichum piloselloides Pres| |=Elaphoglossum pilosel- loides (Presl) Moore]. OTHER SPECIES EXAMINED: Elaphoglossum hayesii (Mett.) Maxon, E. horridulum J. Smith, E. jamesonii (Hook. & Grev.) Moore, E. pusillum (Mett.) C. Chr., and E. spatulatum (Bory) Moore. Elaphoglossum horridulum and E. jamesonii look like they belong here, but are different in their spores having spicules. This subsection is close to subsect. Setosa, although its spores have ridges rather than crests and hardly any are perforate, while the hydathodes are less conspicuous than in subsect. Setosa. Elaphoglossum subsect. Alpestria Mickel & Atehortua, subsect. nov. Rhizomata longe repentia; rhizomatis paleae fuscae; phyllopodia inconspicua, nervi raro furcati; hydathodi conspicui; stipitis laminaeque paleae subulatae, raro lanceolatae; paleae inter sporangia nullae; sporae perforatae spiculatae breviter cristatae. 3% L-TYPE SPECIES: Acrostichum alpestre Gardn. [=Elaphoglossum alpestre (Gardn.) Moore}. OTHER SPECIES EXAMINED: Elaphoglossum barbae Rosenst., E. chiapense A. Reid Smith, E. hirtipes (Fée) Brade, and E. leptophlebium (Baker) C. Chr. This subsection is close to subsect. Setosa in the subulate scales and perforate spores, but differs in the rhizome scales and habit. Elaphoglossum yatesii (Sod.) Christ seems to fit here, except that its blade is densely clothed with lanceolate scales. 759°” Elaphoglossum subsect. Plumierana Mickel & Atehortua, subsect. nov <5 rn. e qo i hizomata breviter repentia usque erecta; lamina margine crenulata; laminae paleae subulatae aurantiacae; sporae cristatae (Fig. 15). ATY PE SPECIES: Elaphoglossum plumieri Moore.~ ! OTHER SPECIES EXAMINED: Elaphoglossum buchii C. Chr., E. lanceum Mickel, and E. smithii (Baker) Christ. In their thin, crenulate blades these species resemble somewhat the species of subsect. Feeana, but are distinct in rhizome habit, rhizome and blade scales, phyllopodia, and hydathodes. There is considerable spore variation within the few species of this group. Elaphoglossum lanceum has highly perforate, lace-like crests, whereas the other species have solid, slender, nonperforate crests. a37F Elaphoglossum sect. Eximia Mickel & Atehortua, sect. nov.” site Rhizomata breviter repentia vel adscendentia; phyllopodia nulla; stipitis paleae lanceolatae minimae vel saepe subulatae; nervi distantes, angulo 40—60° abe untes; hydathodi conspicui; laminae paleae sparsae minimae, non subulatae; sporae reticulato-echinatae ecristatae. a5 YPE SPECIES: Acrostichum eximium Mett. [=Elaphoglossum eximium (Mett.) Christ]. Cy ly P20 0 97°F 66 AMERICAN FERN JOURNAL: VOLUME 70 (1980) KEY TO THE SUBSECTIONS OF SECTION EXIMIA 1. Blade entire, linear to linear-elliptic; stipe scales subulate. subsect. Eximia 1. Blade pedately divided; stipe and rhizome scales small, lanceolate. .......... subsect. Cardenasiana is Elaphoglossum subsect. Eximia Mickel & Atehortua, subsect. nov. Lamina integra, linearis vel lineari-elliptica; stipitis paleae subulatae; laminae paleae sparsae minusculae. ! YPE SPECIES: Acrostichum eximium Mett. [=Elaphoglossum eximium (Mett.) Christ]. OTHER SPECIES EXAMINED: Elaphoglossum aubertii (Desv.) Moore, E. beaurepairii (Fée) Brade, E. brachyneuron (Fée) J. Smith, E. gracile (Fée) Christ, E. lineare (Fée) Moore, and E. stenopteris (Klotzsch) Moore. In the subulate scales, hydathodes, and reticulate-echinate spores (Figs. 16 and 17), this group is similar to some members of subsect. Setosa and to some extent to sect. Undulata. scat Elaphoglossum subsect. Cardenasiana Mickel & Atehortua, subsect. nov. A subsect. Eximia paleis stipitis rhizomatisque minoribus, rhizomate magis carnoso, laminaque pedatim divisa diversa. 4G /TYPE AND SOLE SPECIES: Elaphoglossum cardenasii Wagner. This species is unique in the genus in its pedately divided fronds, but in other characters shows close relationship to subsect. Eximia. ot Elaphoglossum sect. Undulata Christ, Monogr. Elaph. 24, 1899. hizome short-creeping to erect; phyllopodia lacking; blade ovate-lanceolate; blade scales subulate to deltate-lanceolate, erose or toothed; hydathodes con- Spicuous; spores without ridges, openly reticulate-echinate, the spine bases di- verging and forming a reticulum occasionally with irregular verrucae or perforated crests (Fig. 18). TYPE SPECIES: Acrostichum hirtum Swartz [=Elaphoglossum_ hirtum (Swartz) C. Chr.] (syn. E. undulatum (Willd.) Moore). OTHER SPECIES EXAMINED: Elaphoglossum bakeri (Sod.) Christ, E. boryanum (Fée) Moore, and E. proliferans Maxon & Mor- ton ex Morton. This subsection is closely related to subsect. Setosa and possibly to subsect. Eximia, with which it has subulate scales and echinate spores in common. Elaphoglossum castaneum (Baker) Diels is similar to E. hirtum in size and shape and In spores with very small, crest-like projections perforate at the base (Fig. 18). This could represent a condition intermediate between sects. Undulata and Setosa, but the rhizome is longer-creeping, the rhizome scales are small, _ Sclerotic, and resinous, and blade scales are lacking. oa tO rs Bes secs te Amygdalifolia (Christ) Mickel & Atehortua, stat. nov. AGA a saree mas oe obs Monogr. Elaph. 22. 1899. : ‘a eate cae, 2 A ia dy ender; phyllopodia short; rhizome scales round Ee : €s linear-lanceolate; veins evident: hydathodes conspicu- ous; blade subglabrous with irs: Bie meee tt : g ith minute stellate hairs; spores with narrow crests with 1\F% TYPE AND SOLE SPECIES: Acrostichum amy gdalifolium Mett. [=Elapho- glossum amygdalifolium (Mett.) C hrist]. — “\4y This species is quite distinct and has no close relatives. MICKEL & ATEHORTUA: SUBDIVISION OF ELAPHOGLOSSUM 67 ACKNOWLEDGMENTS This study was supported by National Science Foundation grant DEB 77-25582 to the senior author. We gratefully acknowledge the kind help of Dr. Rupert Barneby for his assistance with the latin and his criticism of the manuscript, Mr. Joel Huang for his technical work with the scanning electron microscope and for preparing the photographs, and Drs. Arthur Cronquist and David Lellinger for nomenclatural advice. LITERATURE CITED ALSTON, A. H. G. 1956a. The Brazilian species of Elaphoglossum. Bol. Soc. Brot. 32:1-32. _______, 1956b. The subdivision of the Polypodiaceae. Taxon 5:23-25. CHRIST, H. 1899. Monographie des Genus Elaphoglossum. Neue Denkschr. Allg. Schweiz. Naturf. . Gesammten Naturwiss. 36:1-159, t.1-4. CHRISTENSEN, C. 1938. Filicinae. Jn F. Verdoorn. Manual of Pteridology. Nijhoff, The Hague. CRABBE, J. A., A. C. JERMY, and J. T. MICKEL. 1975. A new generic sequence for the pteridophyte herbarium. Fern Gaz. 11:141-162. DIELS, L. 1899. Polypodiaceae. In A. Engler & K. Prantl. Die Natiirlichen Pflanzenfamilien 1(4): 139-339. ERDTMAN, G. 1957. Pollen and Spore Morphology/Plant Taxonomy. Almgvist & Wiksells, Uppsala, Sweden. FEE, A. L. A. 1845. Histoire des Acrostichées. (Mem. Fam. Foug. 2.) Veuve Berger-Levrault, Strasbourg. —_—___—.. 1852. Genera Filicum. (Mém. Fam. Foug. 5.) Veuve Berger-Levrault, Paris & Strasbourg. GOMEZ, L. D. 1975. Contribuciones a la pteridologia costarricense. VI. El género Peltapteris Link en osta Rica. Brenesia 6:25-31. HENNIPMAN, E. 1977. A monograph of the fern genus Bolbitis (Lomariopsidaceae). Leiden Bot. Ser. 2:1-331. HOLTTUM, R. E. 1947. A revised classification of leptosporangiate ferns. J. Linn. Soc., Bot. $1:123-158 ————. 1978. Elaphoglossum. Flora Malesiana II, 1(4):289-314. LIEW, F. S. 1977. Scanning electron microscopical studies on the spores of pteridophytes. XI. The family Oleandraceae (Oleandra, Nephrolepis, and Arthropteris). Gard. Bull. Singapore 30:101-110, ¢.J-VI. LINK, H. F. 1841. Filicum Species in Horto Regio Botanico Berolinensis Cultae. Veit, Berlin. LINNAEUS, C. 1753. Species Plantarum, vol. 2. Salvi, Stockholm. LLOYD, R. M. 1970. A survey of some morphological features of the genus Elaphoglossum in Costa Rica. Amer. Fern J. 60:73-83. MICKEL, J. T. 1980a. Relationships of the dissected elaphoglossoid ferns. Brittonia 32:109-1 17. —_— . 1980b. Elaphoglossum. /n R. G. Stolze. Ferns and Fern Allies of Guatemala. Fieldiana, Bot. 39: in : MOORE, T. 1857-1862. Index Filicum. Pamplin, London. MORTON, C. V. 1955. Notes on Elaphoglossum, III. The publication of Elaphoglossum and Rhipidopteris. Amer. Fern. J. 45:11-14. PICHI SERMOLLI, R. E. G. 1968. Adumbratio Florae Aethiopicae. 15. Elaphoglossaceae. Webbia PRESL, C. B. 1851. Epimiliae Botanicae. Abh. Konigl. Bohm. Ges. Wiss. V, 6:361-624. SCHELPE, E. A. C. L. E. 1969. Reviews of tropical African Pteridophyta, 1. Contr. Bolus Herb. 1:1-132. SCHOTT, H. 1834. Genera Filicum. Wallishausser, Vienna. SLEDGE, W. A. 1967. The genus Elaphoglossum in the Indian peninsula and Ceylon. Bull. Brit. Mus. (Nat. Hist.), Bot. 4:79-96. 68 AMERICAN FERN JOURNAL: VOLUME 70 (1980) SMITH, J. 1841. An arrangement and definition of the genera of ferns, with observations on the affinities of each genus. J. Bot. (Hook.) 4:38-70, 147-198. SODIRO, A. 1897. Cryptogamae Vasculares Quitenses. Typis Universitatis, Quito. SPORNE, K. R. 1975. The Morphology of Pteridophytes, ed. 4. Hutchinson, London. WAGNER, W. H., Jr. 1954. A Bolivian Elaphoglossum of unique leaf structure. Bull. Torrey Bot. Club 81:61-67. REVIEW “HOW TO KNOW THE FERNS AND FERN ALLIES,” by John T. Mickel, 1979. Wm. C. Brown Company Publishers, Dubuque, Iowa. $7.95 hardcover, $5.95 wire coil bound.—In the standard format of the Pictured Key Nature Series, this book provides the information needed to identify North American ferns and fern allies. Introductory chapters on structure, life history, hybridization, cultiva- tion, collection, and nomenclature provide a basic understanding of ferns and fern allies and the terminology needed to identify these plants successfully. This also makes the work useful as a textbook or handbook for amateur pteridologists. The greatest part of the volume consists of bracketed keys which lead the user to the appropriate genus and species. Diagnostic characters for each species, provided in the annotated keys, are supplemented by a brief description, habitat preference, frequency of occurrence, and distribution map, as well as an illustra- tion for nearly every species. Limited synonymy is also included. Hybrids and infraspecific taxa are mentioned with related species. Problems in taxonomy are explained so that the basis for confusion can be understood. The genera are listed alphabetically, enabling the experienced pteridologist to turn to the appropriate genus and begin keying at that point. The uninitiated can begin with the generic key that is found near the beginning of the book. Edgar Paulton’s line drawings are quite good and in general capture the distinguishing characteristics of each species. The book concludes with a listing of state and regional identification pteridologists alike have long awaited.—W. C : : -—W. Carl Taylor, D ; Milwaukee Public Museum, Milwaukee, WI 53233. ne lado ls AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 69 Notes on the Natural History of Stylites gemmifera ERIC E. KARRFALT and DALE M. HUNTER* Several unexpected observations were made recently while collecting plants of Stylites E. Amstutz for anatomical study. The plants were collected near the end of the growing season (18-19 April 1979) so as to obtain young plants which had just completed various numbers of growing seasons. In the type locality, Srylites is described as forming pure colonies which stand just above the water level in the lacustrine bog in which they are found (Rauh & Falk, 1959); but the plants we collected at 4100 m altitude (Karrfalt & Hunter 22, NY) near Lago Junin, 14 km north of Junin, Peru, were invariably growing in association with various flower- ing plants (Figs. 1 and 2) and were frequently submerged. The colonies at the Junin locality were generally in the form of radially symmetrical, dome-shaped hummocks (Fig. 1), but various other rounded shapes occurred as well. The hummocks ranged in diameter from 20 to about 200 cm. The larger hummocks generally were found to contain a larger proportion of other plants in addition to Stylites than did the smaller ones. The plants in the hummocks were extremely densely packed and usually stood above the water level, but some hummocks were partially or completely submerged. The submerged portions of these hum- mocks were populated nearly exclusively by Stylites, but the emergent portions included other plants as well (Fig. 5). These plants usually were rather small (with stems a few millimeters in length), but some were quite good-sized (stems 2-4 cm long) and bore about 40 leaves up to 7 cm long, as well as numerous gemmae. Their leaves did not have the typical flattened form with deflexed tips, but rather were subtriangular to terete in cross section and ascending. All intermediate forms between these atypical leaves and those described by Rauh and Falk (1959) were also seen; the variation in leaf morphology will be described in detail in a sub- sequent report. The plants with the atypical leaf form always were submerged and not densely crowded. On the other hand, plants bearing typical leaves occurred both above and below the water level in the bog, but these plants always were densely crowded. Leaf form correlates with population density rather than with emergence or submergence. The nature of this correlation is not certain, but experiments in progress suggest that it is largely or entirely environmental. In contrast to its very limited geographical range, Stylites is extremely vigorous and abundant where appropriate conditions for its growth exist. The J unin locality is a bog which has been used as a pasture for many years. It is heavily grazed by sheep and llamas, as indicated by the cropped herbage (Figs. 1 and 4) and abun- dant llama dung. The Stylites plants, however, very rarely show any evidence of even accidental damage by the animals. The Junin locality occupies at least sev- eral acres; we were unable to determine its full extent due to our anoxemia and consequent lack of energy. The leaves of Stylites are coated with considerable quantities of mucilage, as are the basal parts of the leaves of all /soétes species of which we have seen living *Department of Biology, Brooklyn College, Brooklyn, NY 11210. 70 AMERICAN FERN JOURNAL: VOLUME 70 (1980) material. Also as in /soétes, the leaves are replaced annually (Rauh & Falk, 1959). As the new leaves grow and expand within the hummock, the dead, mucilage- coated leaves of the largest plants are extruded en masse onto the surface of the fie a at i FIG. 1. A typical hummock. FIG. 2. Close up of part of the hummock Gouden e ig ? os Submerged plants with atypical leaves. FIG. 4. A hummock and surrounding g . . 5. A partially submerged hummock. Unlabelled arrows = Stylites plants; E = pat- rice ea (Fig. 2, E). Once out on the surface of the hummocks, the individual ea sporophylls become separated from one another (Figs. 1, 5, S). The extru- sion of the old sporophylls would seem to be advantageous for spore dispersal. Indeed, an analogous process has been shown to be involved in spore dispersal in KARRFALT & HUNTER: NATURAL HISTORY OF STYLITES 71 Isoétes drummondii (Osborn, 1922). However, examination of large numbers of extruded sporophylls of Stylites never revealed any discernible evidence of the establishment of gametophytes by the spores carried with the extruded sporophylls. Field observations of gametophytes were necessarily limited to those which could be made with a hand lens; that is, only megaspores were examined and these only for the opening of the trilete scar. Any megagametophytes which were contained within unopened spore walls were not distinguished from un- germinated spores. Rauh and Falk (1959) found very few megagametophytes and no mi- crogametophytes of Stylites. In our material, gametophytes were likewise very infrequently encountered. Only megagametophytes were identified in the field and these were found only in association with adult sporophytes which showed some evidence of recent damage or injury, such as relatively few leaves or a reduced stem diameter near the apex. Six gametophytes were found, all of which bore sporophytes with one or two leaves and roots. The occurrence of the rare megametophytes only in association with the rare, injured sporophytes suggests that the absence of gametophytes from other locations is the result of unequal competition between the gametophytes and the much larger, densely crowded adult sporophytes and gemmae. It was not possible to determine the specific source of the spores which gave rise to the gametophytes we collected. Our gametophytes were probably derived from the massive quantities of spores produced by the immediately adjacent sporophytes, but the possibility cannot be excluded that the successful spores may be have been transported with old sporophylls which had been extruded onto the surface of the same or some other hummock. Many of the plants which we collected bore abundant gemmae and therefore must be assigned to S. gemmifera W. Rauh, inasmuch as S. andicola E. Amstutz has no vegetative reproduction. The other criteria by which Rauh and Falk (1959) distinguished the sporophytes of S. andicola from those of S. gemmifera are merely quantitative and are of questionable value. For example, the leaves of S. andicola are said to be 5—7 cm long, whereas those of S. gemmifera are said to be 3.5-5 cm long, but as noted above some gemma-bearing plants in our collection had leaves as long as 7 cm. Although these longest leaves did not have the mor- phology typical of Stylites, the leaves of plants collected from hummocks invari- ably had the typical form, and some of these were as long as 5.5 cm. According to Rauh and Falk, the stem of S. andicola is mostly unbranched and up to 20 cm long, whereas that of S. gemmifera is frequently branched and not more than 8 cm long; obviously these characters would be of no use in identifying an unbranched plant whose stem was not more than 8 cm long. Also, S. andicola is supposed to form hummocks in which all individuals are the same age, whereas colonies of >. gemmifera contain both old and young plants. Although they did not explicitly state their method, Rauh and Falk seem to have used size as an indicator of relative age. In any case, the hummocks which contain unbranched plants of a uniform large size and no gemmae (i.e., hummocks of *’S. andicola’’) may simply 72 AMERICAN FERN JOURNAL: VOLUME 70 (1980) be relatively old colonies in which the intense competition for space has been resolved in favor of the largest and most vigorous individuals which neither branch nor produce gemmae. The only qualitative distinction between the sporophytes of the two species of Stylites is the presence of gemmae in S. gem- mifera and their absence in S. andicola. However, the number of gemmae on a plant is highly variable. In our material, from one to eight were seen, and many plants had no gemmae at all. According to the criteria given by Rauh and Falk, sporophytic specimens without gemmae and with stems shorter than 8 cm may be distinguished as to species only by the length of their leaves. Unfortunately, as noted above, we have gemma-bearing plants, obviously assignable to S. gemmi- fera, which have leaves longer than 5 cm. Thus it appears to us that the distinct- ness of the two species of Stylites is in sufficient doubt that a critical reexamina- tion of these two taxa is in order. Moreover, inasmuch as the separation of Stylites from /soétes already has been questioned (Kubitzki & Borchert, 1964; Bierhorst, 1971), this reexamination also should review the generic assignment of these species. LITERATURE CITED BIERHORST, D. W. 1971. Morphology of Vascular Plants. Macmillan, New York. KUBITZKI, K. and R. BORCHERT. 1964. Morphologische Studien an Isoétes triquetra A. Braun und Bemerkungen uber das Verhaltnis der Gattung Stylites E. Amstutz zur Gattung Isoétes L. Ber. Deutsch. Bot. Ges. 77:227-233. OSBORN, T. G. B. 1922. Some observations on Isoetes Drummondii, A. Br. Ann. Bot. 36:41-54. RAUH, W. and H. FALK. 1959. Stylites E. Amstutz, eine neue Isoetacee aus den Hochanden Perus. I. Teil: Morphologie, Anatomie, und Entwicklungsgeschichte der Vegetationsorgane. Sitz- ungber. Heidelberger Akad. Wiss. 1959:1-83. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) 73 Reciprocal Allelopathy Between the Gametophytes of Osmunda cinnamomea and Dryopteris intermedia RAYMOND L. PETERSEN* and DAVID E. FAIRBROTHERS** Allelopathy is the chemical inhibition of growth and/or development of one organism by another. The literature on allelopathy is very extensive and has been summarized in several reviews (Muller, 1970; Pickett & Baskin, 1973; Rice, 1974). See also Swain (1977) for a synoptic review of secondary compounds as al- lelopathic agents. In the life cycle of any species, one portion, designated by Petersen & Fair- brothers (1973) as the weakest link, is likely to be the most vulnerable to al- lelopathic interactions. As an evolutionary strategem, allelopathy would be de- veloped most effectively against the weakest link in an organism’s life cycle, such as germinating spores and developing prothalli of ferns or germinating seeds and seedlings of higher plants. Furthermore, it is at these critical points that one ought to be able to best detect allelopathy. Intra- and interspecific allelopathic interactions occur in four ways: (1) sporophytes acting on sporophytes; (2) sporophytes acting on gametophytes; (3) gametophytes acting on sporophytes; and (4) gametophytes acting on gametophytes. Seed plants reduce gametophytic vulnerability structurally by en- closing the gametophyte and ovule and functionally by pollination and fertiliza- tion. Therefore, the weakest link in seed plants is shifted from their gametophytes to their germinating seeds and seedlings. In pteridophytes, presumably the gametophyte generation—spore, prothallus and gametes—is the most vulnerable portion of the fern life cycle with regard to interspecific interactions for habitat maintenance (allelopathy and competition), for a smaller amount of chemicals would be required to eliminate a gametophyte from effective competition than a sporophyte. Of the four interactions, gametophytes acting on gametophytes of different species permits study of the weakest link hypothesis in an experimental design where interspecific competition is minimal. Fern gametophytes were selected as experimental organisms because they are easily cultured and amenable to the experiments’ requirements of control, replication, manipulation, and minimiza- tion of competition. Six fern species usually found growing in similar natural habitats were selected for the initial survey experiment which was conducted to determine if allelopathic effects were detectable between any of the species. By similar natural habitats we mean that, among the species selected, there is some overlap between their respective ecological amplitudes so that it is possible that they would be in competition for the same space. The literature on fern allelopathy is limited but increasing. Froeschel (1953) reported that water extracts of Polypodium aureum and Lycopodium clavatum *Department of Botany, Howard University, Washington, D.C. 20059. **Department of Botany, Rutgers University, New Brunswick, NJ 08903. AMERICAN FERN JOURNAL: VOLUME 70 (1980) 74 FIGS. 1-3, Gametophytes of Dry. growth. FIG. 1. Gametophytes o opteris intermedia and Osmunda cinnamomea after 30 days of f D. intermedia (2 prothallial cell Stage) and O. cinnamomea (multi- al D. intermedia gametophyte from a control monoculture. FIG. 3. phyte from a control monoculture. PETERSEN & FAIRBROTHERS: RECIPROCAL ALLELOPATHY 75 decreased the growth rates of gametophytes of four fern species. Bell (1958) found that an aqueous extract of Dryopteris filix-mas gametophytes stimulated spore germination and prothallium growth in D. borreri, but that the gametophytic de- bris of D. filix-mas, when incorporated into agar medium, prevented D. borreri spore germination, an example of gametophyte—gametophyte allelopathy. Fukuzumi (1971) reported allelopathic effects from Preris japonica frond extracts on Impatiens balsamina root growth. Gliessman and Muller (1972) investigated the phytotoxic effects of compounds from Preridium aquilinum on the surround- ing vegetation. (See Miller, 1968 for an excellent, though now dated, comprehen- sive review of fern gametophyte literature.) Davidonis and Ruddat (1973, 1974) reported sporophyte—gametophyte al- lelopathy in ferns. The roots and fronds of Thelypteris normalis produce two allelopathic chemicals, thelypterin A and B, which inhibit cell division in the gametophytes of TJ. normalis, Pteris longifolia, and Phlebodium aureum. Thelypterin A has been tentatively identified as an indole derivative. Davidonis (1976) found that 7. normalis gametophytes produce thelypterin A; the sporophytes of other ferns also contain these compounds: thelypterin A and B in T. dentata roots, thelypterin A in T. noveboracensis leaves, and thelypterin B in the roots of two Pteris species. Davidonis (1976) also reported the presence of an unidentified growth inhibitor in the leaves of Osmunda cinnamomea. Star and Weber (1978) reported that sporophyte exudates from Pityrogramma species in- hibit spore germination and gametophyte development in P. calomelanos. In- hibitors were identified as a dihydrochalcone and a flavonol. Munther and Fair- brothers (1980), testing leaf leachates and extracts of Dennstaedtia punctilobula, Osmunda cinnamomea, and Osmunda claytoniana obtained from New Jersey and Vermont populations, demonstrated geographic differences in allelopathic and autotoxic responses among these species based on the amount of spore germina- tion. MATERIALS AND METHODS Mature spores were collected from the following species: Osmunda cin- namomea, O. claytoniana, O. regalis, Matteuccia struthiopteris, Onoclea sen- sibilis, and Dryopteris intermedia. White's minimal nutrient medium at one-half strength at pH 5.5 was used to culture the gametophytes. Both liquid and 1% agar cultures were used. Spore density was approximately 1,000 spores per 90 mm diam. plate. All cultures were grown under axenic conditions (Steeves, 1955), ina growth chamber with fluorescent light (300 ft-c) at 20° C under a diurnal cycle of 12/12 hr. In the initial survey experiment, spores of the six species were sown on agar plates in paired strips adjacent to one another in all possible combinations. Con- trol plates containing spores from one species also were sown. Plates were examined daily at the interfaces of adjacent species for symptoms of allelopathy or competition such as decreases from the control plates in percent germination or growth. 76 AMERICAN FERN JOURNAL: VOLUME 70 (1980) PETERSEN & FAIRBROTHERS: RECIPROCAL ALLELOPATHY 77 On the basis of this experiment, D. intermedia and O. cinnamomea were selected as the most promising taxa for further experimentation because the gametophytes of these species appeared to inhibit each other’s growth. A minimum of 10 replicates was run for each experiment. Initially, control plates and experimental plates containing a mixture of Dryopteris and Osmunda spores were prepared. The next phase was designed to eliminate the possibility of competition. Separate liquid cultures of D. intermedia and O. cinnamomea gametophytes (0.5g spores/liter of half-strength White’s Medium at pH 5.5) were initiated and grown for two weeks. The gametophytes were then filtered off and the supernatants were conserved. Agar plates were prepared as above. Control plates consisted of spores of one species sprayed with one ml of the supernatant of the same species; experimental plates contained spores of one species sprayed with one ml of the supernatant of the other. This experimental design eliminates the possibility of interspecific competition (e.g., differential nutrient assimilation by one species over the other). RESULTS AND DISCUSSION In the survey of six species for allelopathic symptoms, a clear area was detected on the plates at the interface between O. cinnamomea and D. intermedia gametophytes. This was the result of progressive inhibition of D. intermedia gametophyte growth, which was proportional to the proximity of Osmunda gametophytes. In the first phase of the Dryopteris—Osmunda growth rate analyses, Dryopteris spore germination was initially lower on the experimental plates (60% germination after 7 days) compared to the control plates (90% germination after 7 days). But after 10 days, 90% germination was reached on the experimental plates. Post- germination rate data was discontinued because it soon became apparent that the gametophytes of both species on the experimental plates were growing very slowly (Fig. 1) and no longer had significantly different growth rates. In contrast, the control plate gametophytes developed normally (Figs. 2 and 3). The gametophytes of D. intermedia and O. cinnamomea inhibit the growth and development of one another, but one cannot distinguish from these results whether the inhibition is the result of allelopathy or competition, and so a set of supernatant experiments was designed to do so. These experiments essentially reproduced the results of the preceeding set of experiments. Control plates sprayed with supernatant from the same species produced normally de veloped gametophytes (Figs. 4 and 5). But experimental plates sprayed with supernatant from the other species yielded gametophytes having a severely retarded growth rate; after germination and a few cell divisions, development essentially ceased (Figs. 6 and 7). This reciprocal supernatant experiment proves that the gametophytes of each species were suppressing cell division of the gametophytes of the other species and that this was done through the release of inhibitory compounds, rather than by competition. 78 AMERICAN FERN JOURNAL: VOLUME 70 (1980) This is the first recorded example in ferns of reciprocal allelopathy, in which two antagonistic species act on one another, that has been demonstrated in vitro between gametophytes. Previous investigators working with ferns have demonstrated unidirectional allelopathy in the following systems: sporophyte acting on gametophyte (Froe- schel, 1953; Davidonis & Ruddat, 1973, 1974; Star & Weber, 1978; Munther & Fairbrothers, 1980) and gametophyte acting on gametophyte (Bell, 1958). We wish to acknowledge financial aid from NSF Grant GB-13202 and a Rutgers Research Council Grant awarded to D. E. Fairbrothers. LITERATURE CITED BELL, P. R. 1958. Variations in the germination-rate and development of fern spores in culture. Ann. Bot -5S11. DAVIDONIS, G. H. 1976. The occurrence of thelypterin in ferns. Amer. Fern J. 66:107-108. —_—_——., and M. RUDDAT. 1973. Allelopathic compounds, thelypterin A and B in the fern Thely- pteris normalis. Planta (Berlin) 111:23-32. UDDAT. 1974. Growth inhibitor in gametophytes and oat coleoptiles by the thely plerin A and B released from roots of the fern Thelypteris normalis. Amer. J. Bot. 1:925-930. FROESCHEL, P. 1953. Remstoffen by lagere plantaardige Organismen. Natuwuv. Tydschr. 35:70-75. FUKUZUMI, K. 1971. cr a effect of Pteris japonica on Impatiens balsamina root growth from leaf extracts. Biol. J. Naro Women’s Univ. 21:8-9. GLIESSMAN, S. R., and C. H. MULLER, 1972. The phytotoxic potential of bracken, Pteridium aquilinum (L.) Kuhn. Madrono 21:299-304. MILLER, J. H. 1968. Fern gametophytes as experimental material. Bot. Rev. 34:361-440. MULLER, C. H. 1970. The role of allelopathy in the evolution of wea In K. L. Chambers, ed. Biochemical Coevolution. Oregon State Univ. Press, Corv MUNTHER, 4s E., and D. E. A 1980. Allelopathy iad autotoxicity in three species of ferns. Amer. Fern J. 70: i S: PETERSEN, . L. and D. E. FAIRBROTHERS. 1973. Allelopathy: Gametophytic chemoantagonism between Dryopteris and Osmunda or supporting the ‘‘weakest link’’ hypothesis. Amer. J. Bot. 60:32. ( Abstr.) PICKETT, S. T., and T. M. BASKIN. he Allelopathy and its role in the ecology of higher plants. Biologist (Phi Sigma Soc.) 55:49-73. RICE, E. L. 1974. Allelopathy. pn a Press, New York. STAR, A. E., and P. WEBER. 1978. Sporophytic exudate inhibition of pe development in Pityrogramma calomelanos. Plant Sci. Conf. VPI, Blacksburg, Va. (Abs STEE VES, T. A., I. M. SUSSEX, and C. R. PARTANEN. 1955. In vitro sides on | abnormal growth of prothalli of the bracken fern. Amer. J. Bot. 42:232-245. SWAIN, T. 1977. Secondary compounds as protective agents. Ann. Rev. Plant Physiol. 28:479-501. AMERICAN FERN JOURNAL: VOLUME 70 (1980) 79 REVIEW “FLORA DEL AVILA” by Julian A. Steyermark and Otto Huber. Publication Especial de la Sociedad Venezolana de Ciencias Naturales, Caracas. 1978. 971 pp. + 308 plates. Bs. 150 (ca. $35.00).—This is a flora of the Parque Nacional ‘El Avila,’ part of a small mountain range lying between the city of Caracas and the Caribbean Sea. The area is about 130 square kilometers and the major peaks reach an altitude of 2000 to 2700 meters. Introductory chapters describe the history of botanical exploration of the re- gion, the soils, geology, climate, principal vegetation types, and the geographic relations of the flora. There is also a series of color photographs of flowers, the vegetation, and three pteridophytes. The flora consists of 127 families of flowering plants, 809 genera, and 1741 species. Each family treatment has an illustrated key to the species and a list of them with ecological notes. There are also line drawings of a selection of species. The ‘Flora of Avila’ is a major addition to neotropical floras, and its utility extends well beyond the region covered. The ferns and fern-allies are all treated under the Pteridophyta, rather than by families, with 52 genera and 151 species included. The key to these is rather long, but it is well organized and accurate. Marginal illustrations of characters aid in the use of the headings. Following the list of species, there are excellent drawings of 99 species. The largest genus is Polypodium (sens. lat.) with 19 species, followed by Elaphoglossum with 15 and Asplenium with 13. About a third of the genera are represented by a single species. The fern flora consists mostly of rather common and widely distributed species of the montane forest and cloud forest of northern South America. There are no ednemics, but about 25 species are of restricted distribution or are otherwise of geographic interest. For example, the rare Lycopodium caracasicum, restricted to the coastal Cordillera, reaches its easternmost station here, and both Phanerophlebia juglandifolia and Pellaea ovata are represented by disjunct sta- tions at the eastern limit of their range. About two-thirds of the species grow in the cloud forest zone and some are restricted to it, for example the five species of Cyatheaceae and several species of Elaphoglossum and Polypodium. There is a subparamo zone on the highest parts of the mountains, and it is here that the strongest relations with the flora of the high Andes are found. Asplenium monanthes and Lycopodium vestitum are examples among the pteriodphytes. The ‘‘Flora de Avila’’ may be obtained from: Sociedad Venezolana de Ciencias Naturales, Calle Arichuna, Apartado de Correos 76771, Urb. El Marques, Caracas 107, Venezuela. The price if 150 bolivares (ca. $35.00).—Rolla Tryon, Gray Herbarium, Harvard University, C ambridge, MA 02138. 80 AMERICAN FERN JOURNAL: VOLUME 70 (1980) SHORTER NOTE THELYPTERIS TORRESIANA IN VENEZULEA. — For several years, the senior author has been intrigued by the sudden appearance and persistence in his Caracas garden of a terrestrial, acaulescent fern, most attractive with its pale green, large, gracefully thrice-cut fronds. Specimens sent to Dr. John Mickel were identified as Thelypteris torresiana (Gaud.) Alston, a species originally described from Guam and found native elsewhere in the Asian tropics. In the New World, this species has become introduced and naturalized in the southeastern United States, Cuba, Jamaica, the Lesser Antilles, Trinidad, Tobago, Honduras, Ven- ezuela, Brazil, and Argentina. Vareschi in Lasser (Fl. Venez. 1:439. 1969) treats this species under the in- validly published name ‘‘Lastrea setigera.’’ Morton (Amer. Fern J. 52:27-29. 1962) gives a correct synonymy and has shown that the species should not be confused with the rare Old World T. setigera (Blume) Ching. Leonard (Amer. Fern J. 62:97-99. 1972) observed the preference of 7. tor- resiana for moist ravines and stream banks in the southeastern United States. It occurs in similar habitats in Venezuela, in such places as moist forests along roads and trails at 400-1400 m altitude. It is common in cool cloud forests, but also grows in warmer zones, both in deciduous and evergreen tropical forests. In the senior author’s garden in Caracas, it is aggressive, weedy, and often invasive, characteristics which have facilitated its spread in natural habitats. In the lo- calities where it has become naturalized, it appears to be part of the native vegeta- tion. According to specimens in the Herbario Nacional de Venezuela (VEN), T. torresiana was first found in Venezuela in 1943 in the Parque Nacional Pittier, Estado Aragua (Killip & Lasser 37797, US, VEN). Since then it has spread in the Coastal Cordillera throughout northeastern Venezuela to the states of Portuguesa, Yaracuy, Carabobo, Guarico, Miranda, Sucre, and Monagas, and to the Distrito Federal.—Julian A. Steyermark, Instituto Botanico, Apartado 2156, Caracas, Venezuela and Francisco Ortega, Estacion Biolégica ‘‘Pozo Blanco,’’ Apartado 116, Acarigua, Edo. Portuguesa, Venezuela. TRIARCH Over 5@ Years of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 alice mnehaineonmne AMERICAN volume 7 FERN ee July-September, 1980 JOURNAL cee en QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY The Distribution and Ecology of Phyllitis scolopendrium in Michi RICHARD P. FUTYMA Supplemental Notes on Lesser Antillean Pteridophytes GEORGE R. PROCTOR Additions to the Pteridophyte Flora of the Great Plains RALPH E. BROOKS Flavonoid Synthesis and Antheridium Initiation in Dryopteris Gametophytes RAYMOND L. PETERSEN and DAVID E. FAIRBROTHERS Date of Publication of Sodiro’s **Sertula Florae Ecuadorensis”’ DAVID B. LELLINGER % Reproductive Biology and Gametophyte Mor of New World Populations of Acrostichum aureum ROBERT M. LLOYD 99 Shorter Notes: Diplazium japonicum New to Alabama; Moths and Ferns; Three Additions to the Pteridophyte Flora of Escambia County, Florida MISSOUR) BWR 111 OCT 16 GARDEN LIBRARY The American Fern Society Council for 198 ROBERT M. LLOYD, Dept. of Botany, Ohio L University, Athens, Ohio 45701. President DEAN P. WHITTIER, Dept. of Biology, ‘Vanderbilt University, Nashville, TN 37235. Vice President LESLIE G. HICKOK, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916 Se ecretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, Va. 22030. ecords Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, D Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Howey i 94720 Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, N.Y. 10458. seaiiee Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, D. C. 20560 ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington, Ind. 47401 JOHN T. MICKEL New York Botanical Garden, Bronx, New York 10458 The ‘‘American Fern Journal’ (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and published at the Smithsonian Institution, Washington, DC 20560. REC One Class postage paid at bara oes Matter for publi g (made 3-6 months after the date of issue) should be addressed to the Editor. Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog, Dept. of Biology, George Mason University, os Va. 22030 Orders for back issues should be addressed to the ‘Treasure er. General inquiries concerning ferns call be addressed to the Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Fern Society (annual dues, $8. 00; life membership, $160.00 Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or ae $1.25; 65-80 pages, $2. 00 each; si 80 fa oe $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; ingle b ach, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, New York 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter r. John T. Mickel, New York a ious Bronx, New York JOSS, is editor of a newslet- ter ‘‘Fiddlehead Forse ” The edi includ- membe es, offers ige or purchase materials, personalia, horticultural 1 notes, and reviews of paageaNy mth on n ferns. Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th Street, Seattle, Washington 98115, is Director. Spores exchanged and collection lists sent on request. ‘Gifts and Bequests ye : wtarected in ferns. Botanical books. t ack te issues of the Journal, and cash or biker gifts are always ni IS and are tax-deductible. Inquiries should be addressed to the Secretary AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 3 (1980) 81 The Distribution and Ecology of Phyllitis scolopendrium in Michigan RICHARD P. FUTYMA* The Hart’s-tongue, Phyllitis scolopendrium, has been noted for its circumboreal and North American disjunct distributions (Fernald, 1935; Wagner, 1972). On this continent P. scolopendrium var. americana Fern. is known to occur in Ontario (Soper, 1954), Michigan (Hagenah, 1954, 1956), New York, Tennessee, and Alabama (Short, 1979). By far the majority of the Hart’s-tongue sites are associ- ated with the limestones and dolomites of the Niagara escarpment. This geological formation can be traced from central New York westward into Ontario, where it turns northwestward near the head of Lake Ontario, and through the Bruce Penin- sula and Manitoulin Island in Lake Huron, into the upper peninsula of Michigan. From there it arcs southwestward through Wisconsin’s Door Peninsula to the east of Green Bay and disappears to the south. This paper will deal principally with the northernmost American Hart’s-tongue colonies, those in upper Michigan, and with some ideas concerning the factors determining its distribution in that region. A NEW LOCALITY IN MICHIGAN On 3 August 1978, I discovered a previously unreported locality for P. scolopendrium in Mackinac County, Michigan, while I was botanizing along the slopes of a bedrock knob on the Niagara escarpment. The site is strewn with low, moss-covered dolomite boulders under a tree canopy almost completely domi- nated by Acer saccharum, with only minor numbers of other hardwood species (Fig. Ty At the time of the discovery of the site, two fronds of P. scolopendrium were collected as a voucher specimen and deposited in the herbarium of the University of Michigan Biological Station (UMBS). Also found were Polystichum lonchitis and Geranium robertianum (Fig. 2), two plants frequently associated with Phyl- litis scolopendrium (Hagenah, 1956). Walking Fern, Camptosorus rhizophyllus, occurs locally at the site, densely covering large boulders at least 1.5 m above the ground surface. Dr. W. H. Wagner, Jr. was among those visiting the site shortly after its discovery. He compiled the following list of pteridophytes on 29 August 1978: Asplenium trichomanes, Botrychium virginianum, Camptosorus rhizophyl- lus, Cystopteris bulbifera, C. fragilis, Dryopteris filix-mas, D. intermedia, D. spinulosa, Equisetum arvense, E. scirpoides, E. sylvaticum, Matteuccia struthiopteris, Onoclea sensibilis, Polypodium virginianum, Polystichum braunii, and P. lonchitis. All were found within 50 m of the Phyllitis colony. Asplenium viride also has been reported at the site (D. Henson, pers. comm. This new Hart’s-tongue site is situated within the Hiawatha National Forest, and will be referred to here as the ‘East Lake station.”’ In the Fall of 1978, the U.S. Forest Service decided to survey and map the extent of the Phyllitis colony. *Department of Botany, University of Michigan, Ann Arbor, MI 48109. Volume 70, number 2, of the JOURNAL was issued June 30, 1980. 2 AMERICAN FERN JOURNAL: VOLUME 70 (1980) 8 tube (horizontal) in the right foreground is 65 cm long. FIG low boulder, along with Polystichum lonchitis and Geraniu . 2. Phyllitis scolopendrium growing on a m robertianum on top of boulder. R. P. FUTYMA: PHYLLITIS SCOLOPENDRIUM IN MICHIGAN 83 In all, 64 individuals were counted, their locations mapped, and the tree nearest to each group of individuals was marked. In combination with long-term observa- tion, this information has potential for use in studying the population ecology of this rare fern. During the summer of 1979, the U.S. Forest Service undertook an inventory of all prospective localities within the borders of the Hiawatha National Forest in hopes of uncovering other unreported Hart’s-tongue sites. Unfortunately, no suc- cess was reported. DISTRIBUTION IN UPPER MICHIGAN Hagenah (1954, 1956) reported on the first two upper Michigan sites for P. scolopendrium. The East Lake station is situated between these two, which are 30 km. apart (Fig. 3). The population at the Trout Lake station in Chippewa County apparently is extinct. The easternmost station, known as the Hagenah site, has recently been acquired by the Michigan Nature Association as a plant preserve. The location of these sites and the major North American concentration of P. scolopendrium in Ontario, south of Lake Huron, are shown in Figure 3. All three Michigan localities are similarly located on prominent hills that are part of the Niagara escarpment. Along much of its length in Mackinac and Chip- pewa counties, the escarpment is obscured by thick deposits of glacial drift. The position of the escarpment is manifested mainly by a series of bedrock knobs scattered from east to west across the region. These hills rise 30-100 m above the surrounding plain and range in area from 150 to over 3000 hectares. It is highly unlikely that they were ice-free nunataks during the Wisconsinan glaciation, as suggested by Fernald (1935) in explaining the occurrence of Hart’s-tongue on the highest outcrops of the same escarpment in Ontario. Another similarity shared by the three localities is that the Hart’s-tongue col- onies are situated at elevations near or above that of the ancient shoreline of Lake Algonquin. In fact, the East Lake station was discovered in the course of floristic reconnaissance along one such shoreline. The plants are growing on boulders uncovered by wave action. Lake Algonquin, a precursor of lakes Huron and Michigan, covered much of upper Michigan immediately upon retreat of the con- tinental ice sheet about 11,000 years ago. At that time the bedrock hills that define the Niagara escarpment formed an archipelago in the lake. By about 10,400 years ago Lake Algonquin ended when the waters fell to lower levels and more of the present land area was uncovered. The fact that P. scolopendrium has been found in upper Michigan only in places that were islands in Lake Algonquin may have some special significance. Throughout Mackinac and parts of Chippewa counties, there is a large area which was inundated by Lake Algonquin but now is covered by deciduous forests of the sort preferred by the fern, in which there are limestone and dolomite outcrops (e.g., Drummond Island) or concentrations of glacially-transported boulders. Al- though actively sought, Hart’s-tongue has not been seen in those places. This distribution pattern may be explained in several ways; here we will consider three hypotheses. 84 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Hypothesis 1.—Phyllitis scolopendrium was dispersed to upper Michigan, pre- sumably from the south, at the time of the existence of Lake Algonquin, ca. 10,500 years ago. These rocky islands with their depauperate flora may have offered suitable substrates and conditions of low competition favorable to the establish- f= TL ee : eu rignlent ide — vet ‘um in the upper Great Lakes region. The three stations of sine re indicated by dots. The general : pe shown by stippling. (modified from Soper 1904: g area where the fern occurs in Ontario is ment of this fern. When more land area was uncovered by the recession of lake levels and closed forests covered the region, P. scolopendrium may not have been a es aggressive Colonizer to spread from its territory acquired earlier. Se deed tat dp pcroer of P. scolopendrium to the former islands may be a of environmental differences between hilltop rock outcrops and those at R. P. FUTYMA: PHYLLITIS SCOLOPENDRIUM IN MICHIGAN 85 lower elevations. The hills along the escarpment have thin soils with more out- crops, and the fact that they were high enough to escape submergence under Lake Algonquin may be only a coincidence. Fewer favorable sites for the fern exist at lower elevations because thicker deposits of glacial till and lake sediments cover the bedrock. At those lowland sites where rock surfaces are available, other environmental factors may be unfavorable. Hypothesis 3.—P. scolopendrium had (or has) a wider distribution in upper Michigan than is known at present. Logging of forests may have opened the vegetation at many former localities of the fern, making the sites unsuitable and leading to its extinction. Therefore, the original distribution of the Hart’s-tongue in upper Michigan prior to European settlement had little to do with the geography of Lake Algonquin. DISCUSSION The suggestion that P. scolopendrium first reached upper Michigan during the existence of Lake Algonquin has some appeal. Such an hypothesis could explain why extensive areas of limestone outcrops which are situated between the Michi- gan stations and the main North American concentration of Hart’s-tongue in Ontario, and which were inundated by Lake Algonquin, such as Manitoulin Island and Drummond Island, are devoid of the fern. The Bruce Peninsula, where many of the Ontario stations are located, also was completely submerged at that time, but one may propose that its connection to the mainland at a point where many non-submerged Hart’s-tongue localities exist facilitated its colonization at a later date. However, fossil pollen studies by the present author and others (Brubaker, 1975; Saarnisto, 1974) indicate that the late-glacial forests of the region, during and after the existence of Lake Algonquin, comprised mainly spruce (Picea spp.) and Jack pine (Pinus banksiana). A salient feature of the ecology of Phyllitis scolopen- drium var. americana is that it is never found in coniferous forests, even when adjacent tracts of deciduous forests contain the fern. In Ontario it is seen under deciduous canopies ranging from successional poplar stands to climax maple- beech forest, but never under conifers (A. Reznicek, pers. comm.). If the Hart’s- tongue is a strict associate of the northern hardwoods forest, then it might have reached northern Michigan only within the past 5000 years, which is when this vegetation type became most widespread in the region. Thus, unless the ecology of this species was different 10,500 years ago from what it is today, we should be safe in rejecting the first hypothesis. The effect of forest clearance on populations of P. sclopendrium is poorly known. Most of the localities in Ontario and Michigan have been logged at one time or another. In the case of the East Lake station, logging did take place, but perhaps not to the extent of clearcutting. The example of the Ontario Hart’s- tongue colonies found in early successional Populus woods shows that it can be an aggressive colonizer little affected by logging, provided that spore sources exist nearby. There is still the possibility that small, isolated colonies could become extinct and not be recolonized after logging and forest regrowth. 86 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Most North American P. scolopendrium sites appear to be associated with moist slopes or hillsides, such as bouldery talus slopes and crests of escarpments (Soper, 1954) and sinkholes (Short, 1979). Along the Michigan outcrops of the Niagara escarpment, there are very few places where there is a high, steep cliff below which a rocky talus has accumulated. The Michigan Hart’s-tongue colonies are found where bedrock just breaks through the surface on a moderately steep hillside or on slopes with a high concentration of low boulders that are separate from the bedrock. At the East Lake station, and possibly the second, easternmost locality described by Hagenah (1956), the boulders on which the ferns are growing represent a lag deposit formed by the removal of the surrounding sandy till by the action of the waters of Lake Algonquin. The boulders themselves had been quar- ried from nearby outcrops by the glacier and carried only a short distance before being deposited. Extensive flat areas with limestone bedrock near or outcropping at the surface, such as Manitoulin and Drummond islands, do not appear to be suitable. Possibly these present too dry a habitat (Hagenah, 1956) or the forests are too open for the Hart’s-tongue. Another indicator of this fern’s requirement for moist conditions is its prefer- ence for growing on low boulders no more than 30 or 40 cm above the forest leaf litter (Fig. 2) or in crevices of limestone pavement. The plants would be more exposed and subject to desiccation on sheer cliff faces or higher on boulders. Low position may also be a consequence of insulation and protection from desiccation provided by winter snow at northern localities. Large, glacially transported dolomite and limestone boulders exposed above the soil are scattered throughout much of Mackinac County south of the Niagara escarpment. Most are close to | m in diameter, but individuals 2-3 m in diameter are not unusual. Seldom do these boulders occur in concentrations similar to those seen in typical Hart’s-tongue habitat. Although the overall forest setting may seem Suitable, these boulders may present desiccation problems and may be too few at any given location to provide a sufficient number of microsites for a viable colony of P. scolopendrium to become established. CONCLUSIONS Phyllitis scolopendrium has now been reported from three localities on the Niagara €sCarpment in upper Michigan. All three stations are similar in that they are situated on hills that were islands in Lake Algonquin, which existed ca. 10,500 years ago. Despite assiduous searching by botanists, P. scolopendrium has never been found in this regi This interesting distribution pattern probably does not indicate that P. scolopendrium first reached these sites while Lake Algonquin was in existence, €getation was coniferous forest, a vegetation type at present this fern. A more likely explanation is that these hillside sites vironmental requirements in terms of topography, moisture, and microsite abundance more adequately than other sites where limestone out- crops and boulders are available. not associated with R. P. FUTYMA: PHYLLITIS SCOLOPENDRIUM IN MICHIGAN 87 We can never be certain that P. scolopendrium did not occur in a greater number of localities in upper Michigan at some time in the last 10,000 years. More intensive botanical exploration in the region may eventually confirm or refute the apparent correspondence between Hart’s-tongue fern localities and island areas in Lake Algonquin. In this regard it may be productive to pay particular attention to boulder concentrations along the ancient shorelines of this glacial lake. With respect to determining the time of immigration of P. scolopendrium to upper Michigan, we can only say that it arrived less than 10,000 years ago and possibly only within the last 5000 years. There is little hope of being able to pinpoint this date more exactly, for it is unlikely that the spores or other parts of this rare plant will be found in the fossil condition. Therefore, the first hypothesis is the least likely and the second is the most plausible explanation for the distribution of Hart’s-tongue in Michigan. We do not have sufficient information to reject the third hypothesis. In order to understand the factors determining the geographic distribution of rare plant species such as P. scolopendrium, we must take into account the vege- tational history of the region, as well as the ecological relationships between rare species and the biotic and abiotic components of their immediate environment. Further contributions in this regard can be made by studying the population dynamics of known colonies of P. scolopendrium. Such a study will be possible at the East Lake station, where an entire Hart’s-tongue colony has been counted and mapped. I would like to thank the following people for their help in providing information at various times during my research and for their comments and criticisms of the manuscript: Joseph Beitel, William S. Benninghoff, Don C. Henson, Anton Reznicek, Charlotte Taylor, Edward G. Voss, and Warren H. Wagner, Jr. The work was supported in part by a National Science Foundation Doctoral Disserta- tion Improvement Grant. LITERATURE CITED i coc atc LB 1975, sidecases forest patterns associated with till and outwash in northcentral L Michigan. Quaternary Res. 5:499-527. FERNALD. M. L. 1935. Critical plants of the Upper Great Lakes region of Ontario and Michigan. Rhodora 37:197—222, 238-262, 272-301, 324-341. a cae D. J. 1954. The Hart’s-tongue in Michigan. Amer. Fern J. 44:2-7. ————. 1956. More Hart’s-tongue in Michigan. Amer. Fern J. 46:70-74. SAARNISTO, M. 1974. The deglaciation history of the Lake Superior region and its climatic implica- tions. Quarternary Res. 4:316-339. SHORT, J. W. 1979. Phyllitis scolopendrium newly discovered in Alabama. sy Fern J. 69:47-48. SOPER, J. H. 1954. The Hart’s-tongue fern in Ontario. Amer. Fern J. 44: 129-14 WAGNER, W. H, Jr. 1972. Disjunctions in homosporous vascular plants. Ann. panos Bot. Gard. 59:203-217. 88 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) Supplemental Notes on Lesser Antillean Pteridophytes GEORGE R. PROCTOR* During the several years since publication of the author’s pteridophyte volume in R. A. Howard’s “Flora of the Lesser Antilles” (vol. 2, 1977), a number of errors and a few omissions have come to light. For the convenience of persons interested in the ferns of this geographic area, it seems desirable to bring together and make available the more important corrections and augmented facts. I am indebted to Dr. David Lellinger of the Smithsonian Institution for placing many of these data at my disposal. The information is presented according to the page-numbers of the volume. Emendations of geographic range, minor corrections of spelling, etc., are omitted. . 7.—In citing collectors for the various islands, the name of Stehlé, 1953-1954, was unfortunately omitted from the lists for Guadeloupe, Les Saintes, and Marti- nique. My apologies to Dr. Stehlé! . 16.—The correct citation for this plant is Psilotum nudum (L.) Beauv., Prodr. Fam. Aethéog. 112. 1805. Also, the correct date of J. Bot. Schrader 1800(2) should be 1802, not 1801. This correction should be made at numerous places where it appears throughout the volume. p. 18.—H. & M. Stehlé (Mém. Soc. Bot. France 1953-54, p. 45) reported Psilotum complanatum Swartz from Martinique on the basis of unpublished reports by Fée and Urban. No specimens were cited. The Urban record (Symb. Ant. 9:391. 1925) merely cites the Fée reference. Fée’s record (Mém. Foug. 11:133. 1866) cites a collection by Mlle. Rivoire, “sur un carapa, prés de Saint-Pierre.” This report should be considered doubtful unless it can be substantiated by an authentic specimen. pp. 75-78.—In subgenus Mecodium, two possible additional species should be noted. One of these, known only from Guadeloupe, was given the illegitimate name Hymenophyllum caespitosum by Fée (1866, non Gaud., 1825) and an unpublished one by Maxon and Morton. This entity, based on a L’Herminier specimen at Paris, ra pee be considered a diminutive form of H. undulatum, but requires further study. The other species is Hymenophyllum I’ herminieri Mett. (Linnaea 35:392. 1868). This plant has the frond shape of H. fucoides (subg. Hymenophyllum), but its entire margins place it in subg. Mecodium. Although the original L’Herminier specimens were found in Guadeloupe, more recently similar material was collected in Domin- ica (G. Proctor Cooper 106, US, collected Jan. 31, 1933, “on rocks”). This may well be a valid species. pp. 132-133.—The correct generic name of this plant is Lonchitis L., Sp. Pl. 2:1078. 1753, and the correct species citation is Lonchitis hirsuta L., loc. cit. The name Anisosorus falls into synonymy, but the relevant typification remains the same. 136.—the correct citation for the Pteridium is P. aquilinum var. arachnoideum (Kaulf.) Brade (Zeitschr. Deutsch. Ver. Wiss. Kunst. S. Paulo 1:56. 1920). *Arnold Arboretum Herbarium, 22 Divinity Avenue, Cambridge, MA 02138. G. R. PROCTOR: LESSER ANTILLEAN PTERIDOPHYTES 89 pp. 136—138.—The plant described and illustrated here should bear the generic name Blotiella R. Tyron (Contr. Gray Herb. 191:96. 1962), and the species should be known as Blotiella lindeniana (Hook.) R. Tyron, based on a type from Venezuela. The name Lonchitis should not be associated with this taxon. As the only record of this species from the Lesser Antilles has been shown by Lellinger (Taxon 26:578-580. 1977) to be based on a misidentification, the genus Blotiella cannot be listed as occurring in the Lesser Antilles, and this entire entry should be deleted from the book. p. 176.—The type species of Cheilanthes, by conservation, is C. micropteris Swartz, of South America. p. 183.—The correct authority of Adiantum lucidum is “(Cav.) Sw.”, with the basionym Pteris lucida Cav., based on a specimen collected by Née in Ecuador; this was cited by R. Tryon (Contr. Gray Herb. 194:148. 1964). The phrase “not Preris lucida Cav.” on line 1 should therefore be deleted. p. 194.—The correct name of this plant is Oleandra articulata (Swartz) C. Presl, Tent. Pterid. 78. 1836, based on the same type cited for O. nodosa. The latter name should be considered illegitimate. This question was thoroughly discussed by G. J. de Joncheere (Taxon 18:538—-541. 1969). p. 219.—Bolbitis aliena (Swartz) Alston (Kew Bull. 1932:310. 1932) was record- ed from the Lesser Antilles by Hennipman (Leiden Bot. Ser. 2:135. 1977). He cited the following specimens: St. Eustatius: Boldingh 44B (U); Guadeloupe: L’ Hermin- ier 21 (CAL); and “Leeward Is.,” Holme s. n. (K). This species can be distinguished from B. nicotianifolia by the lobed pinnae and the absence of a separate terminal pinna. From B. portoricensis it is distinguished by the vein-areoles lacking (or nearly lacking) included free veinlets and by the Sterile blades being neither elongate nor proliferous. p. 288.—The correct citation for the species on this page should be Nephrolepis rivularis (Vahl.) Mett. ex Krug in Urban (Engl. Bot. Jahrb. 24:122. 1897). p. 265. Species 8, listed as Diplazium limbatum, should be removed from Diplazium and restored to the monotypic genus Hemidictyum. The correct name of this species is therefore Hemidictyum marginatum (L.) C. Presl, based on Asplenium marginatum L. p. 292.—The correct name of species no. 20 is Thelypteris opulenta (Kaulf.) Fosberg (Smiths. Contr. Bot. 8:3. 1972), based on Aspidium opulentum Kaulf. The type is Chamisso s. n. (LE), from Guam. p. 295.—The correct name of species no. 22 is Thelypteris kunthii (Desv.) Morton (Contr. U. S. Natl. Herb. 38:53. 1967), based on Nephrodium kunthii Desv. The type is a Venezuelan specimen without stated collector, ex. herb. Desvaux (P.) p. 296.—The correct name of species no. 24 is Thelypteris hispidula (Dene. ) Reed (Phytologia 17:283. 1968), based on Aspidium hispidulum Dene. (Nouv. Ann. Mus. Hist. Nat. 3:346. 1835). The type is said to be a Riedlé or Guichenot specimen from Timor (P. . p. 297.—Note 24a, Thelypteris hispidula var. hispidula and 24b, Thelypteris hispidula var. inconstans (C. Chr.) Proctor, comb. nov., based on Dryopteris dentata var. inconstans C. Chr. (Kungl. Sv. Vet. Akad. Handl. 16(2):27. 1936). The 90 AMERICAN FERN JOURNAL: VOLUME 70 (1980) lectotype of the latter is Ekman H 10524 (S, isolectotype US), selected by A. R. Smith (Univ. Calif. Publ. Bot. 59:67. 1971). p. 329.—The correct name of species no. 3. is Polypodium sororium Humb. & Bonpl. ex Willd. in L. (Sp. Pl. 5:191. 1810). The type is Humboldt 424 (B—Herb. Willd. 19684-1) from near Caripe, Venezuela. p. 331.—The correct name of species no. 6 is Polypodium dissimile L., the type being P. Browne (LINN 1252.24) from Jamaica. p. 338 or 339. Polypodium palmeri Maxon (Contr. U. S. Natl. Herb. 17(7):600. 1916) should be added to the Lesser Antilles list. The type is Palmer 308 (US 572544 from Mexico. This member of subg. Microgramma was collected long ago in Barbados by Jenman (NY, US), unfortunately without further data. This species is largely Central American in distribution, but also has been found once in Jamaica. This species is somewhat similar to P. lycopodioides, but differs in its larger size (sterile fronds 5—20 cm long, 2-4 cm broad) and thicker texture, and in the thicker, rope-like, whitish-scaly rhizomes. p. 344.—Polypodium decurrens Raddi (Opusc. Sci. Bol. 3:287. 1819; Pl. Bras. 1:23, t. 33. 1825), based on material from Brazil, has been confirmed as occurring in the Lesser Antilles. There is a good Martinique specimen of Duss 1568 at US, whose label says, “Terrestre, rare. Piton Marcel, entre la montagne Pelée et le Precheur; tres rare dans les pitons de Fort de France,” collected July 1885. The Plumier plate cited on p. 344 (Tr. Foug. 99, t. 1/4) unquestionably belongs to the same species. Polypodium decurrens is referred to subg. Campyloneurum, and is unique among the lesser Antillean species of this subgenus in having pinnate instead of simple fronds. The individual pinnae are not unlike the entire blade of P. repens in shape, texture, and venation. p. 344.—Polypodium recurvatum Kaulf. can definitely be excluded from the Lesser Antilles list; the record was based on Duss 4093, which was referred to P. y Weenes & Bonpl. ex Willd. by A. M. Evans (Ann. Mo. Bot. Gard. 55(3): p. 361.—In the caption for Figure 60b, the name for “e” should be G. taxifolia (not G. taenifolia). _ p. 366. A recently described species of Cochlidium from the Lesser Antilles is C. jungens L. E. Bishop (Amer. Fern J. 68:84. 1978), based on Nicolson 1975 (US) from Morne Micotrin, St. George Parish, Dominica. Among the Lesser Antillean species of Cochlidium, C. jungens is distinguished from C. seminudum by the smaller size of the fronds (2-8 cm long vs. 8-20 cm), but also by the non-contracted and non-acuminate fertile fronds; the sterile blades are also narrower, being mostly less than 3 mm wide. From C. rostratum it is distinguished by the superficial sori, not immersed in a deep central groove. p. 368.—The name of species no. 2. should be Cochlidium rostratum (Hook. ) Maxon ex C. Chr. (Dansk. Bot. Ark. 6(3):23. 1929), based on Wright s. n. (K., weed US) from Omotepe Island, Nicaragua. Additional Lesser Antillean records of this species include Stehlé 341 and 1096 (both US) from Guadeloupe. Pp. 374. The correct citation for Vittaria lineata is (L.) J. E. Smith (Mém. Acad. Turin 10:421, ¢. 9, f. 5. 1793). AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 3 (1980) 91 Additions to the Pteridophyte Flora of the Great Plains RALPH E. BROOKS* Recent herbarium studies made while preparing manuscript for the forthcoming manual of the Great Plains flora have led to the discovery of several specimens representing new state records or significant range extensions apparently overlooked by Petrik-Ott in “The Pteridophytes of Kansas, Nebraska, South Dakota and North Dakota” (Beih. Nova Hedwigia 61:1—332. 1979). Botrychium lunaria (L.) Swartz var. lunaria.—This taxon has been reported previously for the Great Plains from North Dakota by Petrik-Ott (1979, p. 37); however, that specimen is B. minganense Vict. The South Dakota collection represents a southern range extension for this circumboreal species. SOUTH DAKOTA: Lawrence County: Northern Black Hills: Old Balmoral Mine, NW from Crown Hill, shrubby glade on plateau at old mine, 6100 ft alt., June 1930, Mrs. F. L. Bennett s.n. (BHSC). Botrychium lunaria var. onondagense (Underw.) House.—This plant previous- ly was known from scattered localities in the northwestern and northeastern United States. NORTH DAKOTA: Burke County: 12 mi SE of Lignite, N-exposed wooded ravine, 11 June 1971, G. D. Hegstad 7855 (NDA). Botrychium matricariifolium A. Braun.—Petrik-Ott (1979, p. 294) stated that she had seen no collections of this species from South Dakota, and so she excluded it from the Great Plains flora. In 1978 I visited the U.S. National Herbarium and found the specimen cited by Clausen (Mem. Torrey Bot. Club 19:87. 1938) to be determined correctly. This was verified by Dr. David Lellinger (pers. comm., 1980), and so B. matricariifolium must remain a part of the Great Plains flora. SOUTH DAKOTA: Custer County: Black Hills: Custer, 5500 ft alt., 15 Aug 1892, P. A. Rydberg 1186 (US). Botrychium minganense Vict.—Petrik-Ott (1979, pp. 34-36) annotated, de- scribed, and illustrated this specimen as typical -B. /unaria. Of the five plants on the cited sheet, two are immature. The remaining three are typical B. minganense, they have distinctly pinnatifid or pinnate lower pinnae, with only the uppermost pinnae flabellate. This determination was kindly verified by Dr. Warren Wagner, Jr. (pers. comm., 1979). The collection represents a slight southern range extension since the species previously was known from Labrador west to Alaska and south to Michigan, Minnesota, Colorado, Nevada, and California. NORTH DAKOTA: McHenry County: Towner, sandy prairie, 11 June 1955, O. A. Stevens 1530 (NDA) Ophioglossum vulgatum var. pseudopodum (Blake) Farw.—This specimen was first reported as O. vulgatum L. by Clausen (1938, p. 126), who did not recognize any infraspecific taxa within this species. Petrik-Ott (1979, p. 295) listed “Herbarium, University of Kansas, Lawrence, KS 66044. 92 AMERICAN FERN JOURNAL: VOLUME 70 (1980) the record as unverified since she had not examined the specimen. I examined it in 1978, and found it to be the northern var. pseudopodum. The collection was made at the southernmost limit of the range in our region. The variety previously was reported from southern Canada south to Virginia, Indiana, Illinois, Nebraska, and California. I have visited the Kansas locality in recent years and, although the habitat is suitable for this plant, the chances are that it is now extirpated. Road construction and housing developments have drastically altered the area since 1929. NSAS: Crawford County: Pittsburg, | mi W of Broadway, in woods on low, rich slopes and in draws, rare clusters, 15 June 1929, F. A. Riedel s.n. (NY). Polystichum lonchitis (L.) Roth.—This collection is a slight southern and eastern range extension from northeastern Wyoming. Both the South Dakota and northeastern Wyoming sites are disjunct from the primary range of this circumboreal species, which lies more than 150 miles to the west and many more miles to the ortn. SOUTH DAKOTA: Lawrence County: Black Hills, SW1/4, Sec. 36, T5N, RIE, S of Roughlock Falls, mossy loam underwoods, over talus below limestone bluff, N-facing slope, 22 July 1971, C. A. Taylor, W. Casper & A. Glynn 10918 (SDC). I wish to thank the following curators for the loan of specimens and for aid in various other ways: Dr. William T. Barker, North Dakota State University (NDA); Dr. Gary Larson, South Dakota State University (SDC); Dr. John Mickel, New York Botanical Garden (NY); and Dr. Joseph Thomasson, Black Hills State College (BHSC). REVIEW “DAS BUCH DER FREILANDFARNE,” by R. Maatsch. 196pp. illustr. Paul Parey, Berlin and Hamburg. 1980. ISBN 3-489-61422-4. DM. 68. (ca. $40.00).— This book is intended for serious hardy fern growers. An introductory portion contains notes on nomenclature, taxonomy, morphology, and fern habitats illustrated with black-and-white photographs and line drawings. About half the book is a useful alphabetical list of fern species and cultivars, concentrating on those grown in Europe, and giving Latin and common names, a brief description of the plant, and other useful notes. The last quarter of the book concerns fern culture. Unusual and helpful information on flowering plants suitable for growing with ferns is included. | hope the publisher will prepare an English edition so that Prof. Maatsch’s book receives the wide circulation it deserves in the English-speaking world.—D.B.L. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 3 (1980) 93 Flavonoid Synthesis and Antheridium Initiation in Dryopteris Gametophytes RAYMOND L. PETERSEN* and DAVID E. FAIRBROTHERS** There is now a fairly extensive vascular crytogam flavonoid literature and there are a number of researchers actively engaged in this research field (Swain & Cooper-Driver, 1973). All vascular crytogam flavonoid work has been done on the sporophyte generation, with one exception: Laurent (1966) determined that Blechnum brasiliense gametophytes produced the flavonoid kaempferol, which is one of the flavonoids produced by the B. brasiliense sporophyte. Reasons for the paucity of information on fern gametophyte flavonoids include the easy accessibil- ity of sporophytes and the now disreputed opinion that flavonoids, being associ- ated with lignin synthesis, are exclusive to vascularized plant bodies. This exclusivity has been lost because flavonoids have been isolated and identified in various non-vascular plant groups: certain algal divisions, bryophytes (Swain, 1974), and fern gametophytes (Laurent, 1966). Initially our investigation was undertaken to determine if Dryopteris intermedia A. Gray and D. marginalis A. Gray gametophytes produce flavonoids and, if so, were these the same flavonoids produced by their sporophytic counterparts (Petersen, 1976). Because of the unusual results of this first portion of the re- search, the inquiry was amplified to include an analysis of flavonoid content along a developmental profile of the gametophytes. Half-strength White’s minimum nutrient medium adjusted to pH 6.0 was used to culture the gametophytes. Liquid cultures were prepared by placing 0.25 g of spores in a 41 flask and adding 2 | of nutrient solution. Separate cultures of D. intermedia and D. marginalis were grown at 22°C under 300 ft-c of illumination from cool-white fluorescent lights in a 12/12 hr diurnal cycle. Gametophytes were harvested and assayed for flavonoids at three de velopmen- tal stages: (1) pre-antheridial initiation (0 antheridia/gametophyte), (2) antheridial initiation (0 or 1 antheridia/gametophyte), and (3) post-antheridial initiation (4-6 antheridia/gametophyte). Ten-gram samples of harvested gametophytes were immediately extracted in methanol and re-extracted repeatedly until a colorless supernatant was obtained. Concentrated extracts were spotted onto Whatman 3MM chromatography paper (42 x 55 cm). Chromatograms were developed in two dimensions employing the two standard solvent systems for the separation of flavonoids: t-butanol, acetic acid, water (3:1:1) for the first dimension and 15% acetic acid, water for the second dimension. Completed chromatograms were inspected under UV light, both in the presence and absence of NH. Spot color changes under both condi- tions were noted and R; values determined. Spots were excised, eluted in spectral grade methanol, and UV spectral data obtained using standard procedures (Mabry *Department of Botany, Howard University, Washington, D.C. 20059. **Department of Botany, Rutgers University, New Brunswick, New Jersey 08903. 94 AMERICAN FERN JOURNAL: VOLUME 70 (1980) et al., 1970). Positive determinations of isolated flavonoid aglycones were done by co-chromatographing them against authentic compounds. Quantitative scoring for flavonoid content was done by comparative visual inspection of spot intensity and size. In the initial experiment, D. marginalis gametophytes were cultured for 30 days and then harvested. Because of extreme crowding, most of the gametophytes formed filaments rather than plates, and most bore a number of antheridia lat- erally. Flavonoids were detected in these gametophytes, and they were the same ones that occur in D. marginalis sporophytes (Table 1). TABLE 1. IDENTIFICATION DATA FOR FLAVONOID GLYCOSIDES OF QUERCETIN AND KAEMPFEROL FOUND IN DRYOPTERIS INTERMEDIA AND D. MARGINALIS GAMETO- PHYTES AND SPOROPHYTES. Quercetin (A) Kaempferol (B) Kaempferol (C) CHROMATOGRAM SPOT APPEARANCE UV Violet Violet Violet UV/NH3 Yellow Yellow Yellow CHROMATOGRAM SPOT Rr VALUES TAB 0.50 0.60 0.63 HOAc 0.44 0.56 0.45 UV SPECTRAL DATA (A max., nm) MeOH 256, 282, 306, 357 266, 292, 347 265, 302, 349 NaOMe Zils 325, 412 274, 324, 401 275, 325, 401 AICIs sige sp 362sh, 274, 302, 349, 395 273, 302, 348, 394 AIChs/HCI 268, 300sh, 358, 273, 302, 343, 392 273, 298, 343, 392 NaOAc 273, 415 273,302, 381 273, 301, 375 maoeres of inadequate material for spectrometric analysis, new cultures of D. marginalis were Started, and cultures of D. intermedia were begun to determine if they likewise produced the same flavonoids as D. intermedia sporophytes. After three weeks, the cultures were harvested and assayed for flavonoids. No flavonoids were detected on the chromatograms, and examination showed that the ns Jas had not produced antheridia. Therefore, more gametophytes were ema wea flavonoid content could be analyzed at three developmental The last experiments showed that the same flavonoids are produced by the gametophytes of these two species as are produced by their respective sporophytes (Table 1). They are quercetin and kaempferol glycosides. (See Mabry et al., 1970, for data comparisons and structural details.) Dryopteris intermedia (Co rophytes and gametophytes produced two flavonoids: a quercetin glycoside ompound A) and a kaempferol glycoside (Compound B). Dryopteris marginalis produced these two compounds, a wa: é (Compound C), , as well as an additional kaempferol glycoside PETERSEN & FAIRBROTHERS: ANTHERIDUM INITIATION IN DRYOPTERIS 95 The experiments (Table 2) also show flavonoids to be absent (—) during the pre-antheridial stage. They were first detected as faint spots (+) at the onset of antheridium formation. Flavonoid concentration is much higher (+ +) in the post- antheridium initiation stage than at the onset of antheridium formation, as re- vealed by greater intensities of spot fluorescence. Antheridium formation and TABLE 2. FLAVONOID CONTENT OF DRYOPTERIS GAMETOPHYTES AT THREE DE- VELOPMENTAL STAGES. Pre-antheridial Antheridial Post-antheridial initiation initiation initiation COMPOUNDS (A) 4B) 6 = (Cf @EAP 2ONBP AB) ay ay €C) D. intermedia = = as ae ee i at oh a D. marginalis - _ ue - iy =e aie a7. vet — = compound not detected; + = compound detected but at a low concentration relative to ++. flavonoid synthesis clearly are associated, but whether the two events are interre- lated (cause and effect) or merely a coincidence remains to be determined. Al- though nothing comparable has been discovered in flowering plants, Barber (1956) identified a glucose-rhamnose glycoside of quercetin in staminate squash flowers that was absent from the pistillate flowers. We wish to acknowledge financial aid from NSF Grant GB-13202 and a Rutgers Research Council Grant awarded to D. E. Fairbrothers. LITERATURE CITED BARBER, G. A. 1956. Flavonoids of staminate and pistillate squash flowers. Arch. Biochem. Biophys. 64:401-411. LAURENT, S. 1966. Contribution a I’étude des tanins et des autres substances phénoliques hy- drosolubles, élaborées par les prothalles de filicinées. Thése Docteur I’ Université de Paris. MABRY, T. J., K. R. MARKHAM, and M. B. THOMAS. 1970. The Systematic Identification of Flav onoids. Springer-Verlag, ek York, Heidelberg, and Berlin. PETERSEN, R. L. 1976. Chemical research in the genus Dryopteris Adanson: systematics, mor- aoe and allelopathy. Ph.D. Thesis. Rutgers, The State University, New Brunswick, Jer New SWAIN, T. 1974. | Biochestica) evolution in plants (Chapter II). In nai Florkin and E. H. Stotz, eds. Comprehensive Biochemistry, vol. 29A. Elsevier, Amster » and G. COOPER-DRIVER. 1973. Biochemical anes in the Filicopsida. Jn A. C. Jermy, J. A. Crabbe, and B. A. Thomas, eds. The Phylogeny and Classification of the Ferns. Bot. J. Linn. Soc. 67, Suppl. 1. Academic Press, New York and London. 96 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) Date of Publication of Sodiro’s ‘‘Sertula Florae Ecuadorensis”’ DAVID B. LELLINGER* Several years ago, Morton (Amer. Fern J. 62:57-64. 1972) published on the dates of original publication of the parts of Father Luis Sodiro’s ‘*Cry ptogamae Vasculares Quitenses.”’ In that case, original publication in periodicals antedated the reprint ‘‘Cryptogamae,”’ sometimes by several years. Sodiro’s **Sertula Florae Ecuadorensis’’ exhibits similar, but shorter differences between original and reprint publication dates. The portions concerning ferns first were published in four parts in the ‘‘Anales de la Universidad Central’ in Quito, Ecuador. These parts were reprinted a few months later with changes of pagination in two parts under the general title ‘‘Sertula Florae Ecuadorensis.’’ The details are presented in Table 1. Some pages apparently were not reset to produce the reprint, but others obviously were. For instance, the bottom line of page 191 of the first part of the ‘‘Anales’’ ends ‘‘linea-’’, but in the repaged reprint this is six lines from the bottom and has been misspelled ‘‘ilnea-’’! This might lead one to think that the reprint actually is a preprint. However, the ‘‘Anales’’ issue is dated January 1905 and the reprint 1905 with no indication of month, and so it is unlikely that the reprint was issued before the ‘‘ Anales.”’ TABLE 1. DATES OF PUBLICATION OF “SERTULA FLORAE ECUADORENSIS”’ PARTS CONTAINING FERNS. Original publication in Repaged and reprint publication as Anal. Univ. Central Sert 19(135)191-200. Jan 1905. I. Acrosticha. pp. 1-12, t. J. 1905. 22(158-159):21-30. Jan/Feb 1908. II. Pteridophyta. pp. 1-12(part). June 1908. Ps Raa Mar 1908. II. Pteridophyta. pp. 12(part)-27(part). June 1908. (161):161-176. Apr 1908. II. Pteridophyta. pp. 27(part)-42. June 1908. Fortunately, not much time elapsed between the original and reprint publication dates; probably no botanical names will be affected by the adoption of the earlier dates. Many authors, including Christensen in his ‘‘Index Filicum,”’ have cited pages and dates from the reprints, rather than from the original publication, as should be done. Because the original publication is rare, the names and biblio- graphic data for correct citation of new fern names and taxa are given in Table 2. I am indebted to Dr. Dan Nicolson and Mr. James Zarrucci for calling this easy to my attention and for obtaining xerocopies of the original ‘‘Anales”’ * , . ° . U. S. Nat’L. Herbarium, Smithsonian Institution, Washington, DC 20560. D. B. LELLINGER: SODIRO’S “SERTULA FLORAE ECUADORENSIS” 97 TABLE 2. NEW FERN SPECIES AND VARIETIES PUBLISHED BY SODIRO IN THE ‘**ANALES DE LA UNIVERSIDAD CENTRAL.” Acrostichum actinolepis Sodiro, op. cit. 19(135):199. Jan 1905. angamarcanum Sodiro, op. cit. 19(135):193. Jan 1905. antisanae Sodiro, op. cit. 22(161):164. Apr 1908. chodatii Sodiro, op. cit. 22(161):174. Apr 1908. christii Sodiro, op. cit. 19 135):192. Jan 1905. cinereum Sodiro, op. cit. 22(161):172. Apr 1908. cladotrichum Sodiro, op. cit. 19(135):197. Jan 1905. diversifolium Sodiro, op. cit. 22(161):166. Apr 1908. ellipsoideum Sodiro, op. cit. 22(161):164. Apr 1908. engleri Sodiro, op. cit. 22(161):167. Apr 1908. fulvum Sodiro, op. cit. 22(161):168. Apr 1908. gossypinum Sodiro, op. cit. 22(161):172. Apr 1908. guamanianum Sodiro, op. cit. 22(161):169. Apr 1908. hieronymi Sodiro, op. cit. 19(135):199. Jan 1905. hikenii Sodiro, op. cit. 22(161):169. Apr 1908. litanum Sodiro, op. cit. 19(135):198. Jan 1905. longissimum Sodiro, op. cit. 19(135):191. Jan 1905. molle Sodiro, op. cit. 22(161):171. Apr 1908. muriculatum Sodiro, op. cit. 22(161):174. Apr 1908. oleandropsis Sodiro, op. cit. 19(135):195. Jan 1905. pangoanum Sodiro, op. cit. 19(135):194. Jan 1905. pellucidum Sodiro, op cit. 19(135):195. Jan 1905. pichinchae Sodiro, op. cit. 22(161):173. Apr 1908. pruinosum Sodiro, op. cit. 22(161):166. Apr 1908. pteropodum Sodiro, op. cit. 19(135):196. Jan 1905. rupicola Sodiro, op. cit. 22(161):175. Apr 1908, as ‘*rupicolum.”’ subsessile Sodiro, op. cit.. 19(135):194. Jan 1905. trichophorum Sodiro, op. cit. 19(135):198. Jan 1905. urbani Sodiro, op. cit. 22(161):170. Apr 1908. viscidulum Sodiro, op. cit. 22(161):165. Apr 1908. Alsophila bilineata Sodiro, op. cit. 22(160):90. Mar 1908. A. christii Sodiro, op. cit. 22(160):89. Mar 1908. Asplenium anomalum Sodiro, op. cit. 22(160):95. Mar 1908. chimboanum Sodiro, op. cit. 22(160):102. Mar 1908. costale Sodiro, op. cit. 22(160):95. Mar 1908. crassifolium Sodiro, op. cit. 22(160):97. Mar 1908. heterolobum Sodiro, op. cit. 22(160):98. Mar 1908. hieronymi Sodiro, op. cit. 22(160):99. Mar 1908. humile Sodiro, op. cit. 22(160):99. Mar 1908. melanosorum Sodiro, op. cit. 22(160):101. Mar 1908. oxylobum Sodiro, op. cit. 22(160):96. Mar 1908. procerum Sodiro, op. cit. 22(160):96. Mar 1908. tungurahuae Sodiro, op. cit. 22(160):97. Mar 1908. Mt a aM eg gl gd) okt at al all cl le aaa ot Preer rr? ee 98 AMERICAN FERN JOURNAL: VOLUME 70 (1980) A. vesiculosum Sodiro, op. cit. 22(160):100. Mar 1908. Cyathea asperata Sodiro, op. cit. 22(158-159):27. Jan/Feb 1908. asperata var. minor Sodiro, op. cit. 22(158-159):28. Jan/Feb 1908. brachypoda Sodiro, op. cit. 22(158-159):26. Jan/Feb 1908. canescens Sodiro, op. cit. 22(158-159):22. Jan/Feb 1908. furfuracea Sodiro, op. cit. 22(158-159):25. Jan/Feb 1908. muriculata Sodiro, op. cit. 22(158-159):28. Jan/Feb 1908. nitens Sodiro, op. cit. 22(158-159):21. Jan/Feb 1908. ochroleuca Sodiro, op. cit. 22(158-159):29. Jan/Feb 1908. oxyacantha Sodiro, op. cit. 22(158-159):24. Jan/Feb 1908. parvifolia Sodiro, op. cit. 22(158-159):25. Jan/Feb 1908. subinermis Sodiro, op. cit. 22(158-159):28. Jan/Feb 1908. . tungurahuae Sodiro, op. cit. 22(158-159):30. Jan/Feb 1908. Nephrodium cinereum Sodiro, op. cit. 22(160):103. Mar 1908. N. cinereum var. intermedium, Sodiro, op. cit. 22(160):104. Mar 1908. N. longipilosum Sodiro, op. cit. 22(160):103. Mar 1908. Polypodium scutulatum Sodiro, op. cit. 22(161):163. Apr 1908. Pteris aspidioides Sodiro, op. cit. 22(160):91. Mar 1908. biternata Sodiro, op. cit. 22(160):94. Mar 1908. esmeraldensis Sodiro, op. cit. 22(160):92. Mar 1908. falcata Sodiro, op. cit. 22(160):93. Mar 1908. procera Sodiro, op. cit. 22(160):94. Mar 1908. rigida Sodiro, op. cit. 22(160):92. Mar 1908. rimbachii Sodiro, op. cit. 22(160):91. Mar 1908. robusta Sodiro, op. cit. 22(160):93. Mar 1908. aanaanaaaaa MO OO TO pe REVIEW ‘TAXONOMY OF THELYPTERIS SUBGENUS STEIROPTERIS, INCLUD- ING GLAPHYROPTERIS (PTERIDOPHYTA), by Alan R. Smith, Univ. Calif. Publ. Bot. 76:1-38, t. 1-4. 1980.—Among the large neotropical fern genera, none 1s more complex than Thelypteris. Even though subg. Steiropteris is but a small portion of the genus (22 species, some divided into varieties), this monograph is very welcome. Over half the taxa are new species or required a new name or combination, which is a measure of the confusion which the author has resolved. Spores and some details of morphology are illustrated with photographs, and distributions are shown by means of maps. The author has seen many type specimens, and so the synonymies are doubtless accurate. A list of exsiccatae is included but, unfortunately, an index to scientific namies is not. The work is available for $5.00 from University of California Press, 2223 Fulton St., Berkeley, CA 94720.—D.B.L. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 3 (1980) 99 Reproductive Biology and Gametophyte Morphology of New World Populations of Acrostichum aureum ROBERT M. LLOYD* The majority of homosporous ferns are characterized by a life-cycle which permits the production of a genetically homozygous zygote following self- fertilization of a single gametophyte (intragametophytic selfing). This homozygos- ity leads to the expression of recessive deleterious or lethal genes (genetic load, as defined here) present in the genotype, unless this expression is buffered by the polyploid system. Sporophytes expressing such genes will be eliminated rapidly and the spore genotypes produced individually by the remaining viable sporophytes will be genetically uniform, barring mutation and meiotic ir- regularities (e.g., homeologous pairing; Klekowski, 1979). In species which regu- larly undergo selfing, genetic load will be absent or will be expressed at low levels (Klekowski, 1979). Thus, analysis of genotypes for genetic load allows for an estimate of the genetic variability in a population. The fern life-cycle also permits reproduction which is genetically analogous to inbreeding and outbreeding in angiosperms, the latter facilitating the storage of recessive deleterious and lethal genes (Wallace, 1970). Although none of the above patterns of reproduction are mutually exclusive, work of the past decade has led to the hypothesis that specific morphological and developmental features of the gametophyte generation will increase the probability of selfing or crossing (intergametophytic mating) and that these probabilities can be correlated with estimates of heterozygosity in the form of genetic load (Lloyd, 1974). However, more recent work with Ceratopteris (Lloyd & Warne, 1978) and Acrostichum (Lloyd & Gregg, 1975) suggests that the past hypotheses are insufficient to explain the genetic diversity expressed in these species and that other factors are in- volved. This paper summarizes our most recent work on the gametophyte mor- phology, reproductive biology, and genetic diversity in a number of populations of Acrostichum aureum distributed from Florida to the northern coast of South America and attempts to circumscribe the current problems in this field. The genus Acrostichum consists of at least three species: A. danaeifolium Langsd. & Fisch., a New World endemic which is widely distributed in fresh water and slightly saline swamps (Adams & Tomlinson, 1979); A. aureum Ais circumtropical in distribution and usually most abundant in mangrove habitats where it can withstand partial tidal immersion (Holttum, 1954; Small, 1938); and A. speciosum Willd., a species of tropical Asia and Australia which is abundant in mangroves through Malaya in areas frequently inundated by tides (Holttum, 1954). These types of habitats are extreme; few species of plants have evolved the necessary physiological and morphological features to successfully colonize them. Previous work on the gametophyte generation in Acrostichum includes mor- *Department of Botany, Ohio University, Athens, OH 45701. 100 AMERICAN FERN JOURNAL: VOLUME 70 (1980) FIGS. 1-3. Spores of A. aureum. FIG. 1. Papua, Brass 518 (UC), x 1000. FIG. 2. Florida, Curtiss 5463 (UC), x 10000. FIG. 3 3. Papua, Brass 518 (UC), x 8000. R. M. LLOYD: REPRODUCTIVE BIOLOGY OF ACROSTICHUM 101 phological studies of A. speciosum by Stokey & Atkinson (1952) and a study of the morphology and reproductive biology of Mexican populations of A. danaeifolium by Lloyd & Gregg (1975). MATERIALS AND METHODS Spores of A. aureum were collected from 39 plants from eight populations as follows: Culture No. 146: 1.1 mi W of U. S. Highway 41 on State Highway 92, Collier Co., Florida. 148: 2.1 mi W of Westlake on State Highway 27, Everglades National Park, Dade Co., Florida. 150: 30.5 mi SW of entrance station on State Highway 27 at road to Westlake, Everglades National Park, Monroe Co., Florida. 190: 0.25 mi E of Negril on road to Savana la Mar, Westmoreland Parish, Jamaica. /9/: Mile post 57, 57 mi E of Georgetown on Public Road East, Guyana. 192; 8 km N of Governor’s Palace, Parimaribo, near end of road to Leonsburg, Suriname. /93: 0.2 mi from road to Colon on road to Coco Solo, Canal Zone, Panama. 194: Lowland area near Pacific Ocean at N end of the Bridge of the Americas, Canal Zone, Panama. Spores were sown and gametophytes grown aseptically on sterile inorganic nutrient medium solidified with 1% agar (for composition see Klekowski, 1969) in 100 x 15 mm petri dishes. Gametophytes were grown under continuous illumina- tion by fluorescent and incandescent lamps at an intensity of 210 to 290 ft-c at temperatures of 19-24° C. Prothallial morphology was studied using living mate- rial as well as that mounted in Hoyer’s medium mixed with acetocarmine. Spore sizes were determined by mounting spores in diaphane and calculating their equatorial diameters with a calibrated ocular micrometer. Other methods utilized in specific experiments are described below. Spores observed by the scanning electron microscope were dry-mounted on double-stick tape, coated with gold ca. 10 nm thick, and observed at 20 kv ac- celerating voltage with a Hitachi HHS-2R scanning electron microscope. RESULTS Spores of A. aureum are tetrahedral and are (37)45—-72 (mean + s.d. = $6.7 = 4.58) um in diameter. The spore surface is minutely tuberculate (Fig. 1). The tubercle-like structures on the surface bear varying numbers of projecting papil- lae. Spores from plants from Papua (Figs. 1, 3) and Fiji exhibit numerous but somewhat irregularly shaped and oriented papillae. Spores from Florida plants (Fig. 2) and other Fijian plants appear to have more numerous papillae as well as other types of superficial deposits. In contrast, spores examined from plants from Australia and Trinidad are tuberculate but appear either to lack papillae or to have thickened superficial deposits which more or less obscure their presence. Spores examined of A. speciosum from Papua exhibit surface features highly similar to those from Florida plants of A. aureum. eS Spore germination is usually initiated by the emergence of a rhizoid five days following sowing. Gametophytes produce a one-dimensional filament up to 10 cells in length before initiation of two-dimensional growth (Fig. 4). In some in- Stances, cells near the base of the filament will divide, producing a second one- dimensional filament (Fig. 5). 102 AMERICAN FERN JOURNAL: VOLUME 70 (1980) ait wd Tt) ee * ; eee r) ee a ee? QS ta ec c | cc ce < / FIGS. 4-11. Stages in gametophyte development and sexually mature gametophytes ot A. aureum. FIG. 4. One-dimensional filament with initiation of two-dimensional growth in basal cell, 185 wm long, days after sowing. FIG. 10. Y ancia. 4.0 mm long, = days after sowing. FIG. 11. Mature gametophyte showing sequential pattern of archegonial- antheridial-archegonial production and branched meristem, 7.5 mm long, 96 days after sowing. R. M. LLOYD: REPRODUCTIVE BIOLOGY OF ACROSTICHUM 103 Two-dimensional growth is initiated by a longitudinal division in a central or more basal cell of the one-dimensional filament. Further longitudinal divisions in the filament may follow one or more patterns: cells throughout the filament, except for the terminal and basal cells, may divide longitudinally and produce a two-dimensional filament two cells wide (Fig. 6); less frequently, central cells of the 1-dimensional filament may divide sequentially, producing an area up to four or more cells in width before further divisions occur in the more basal and terminal cells of the filament. In both pathways, 2-dimensional growth ultimately results in a broadly linear or spatulate gametophyte four to six cells wide. Further divisions of cells along one of the lateral margins of these prothalli will produce a lateral meristematic region located near the basal region of the gametophyte. Subsequent growth produces an asymmetrical ovate prothallus with different sized wings (Fig. 8). In older gametophytes, growth frequently produces more or less symmetrical wing tissue on both sides of the lateral meristem, resulting in a mature prothallus which appears to have an apical meristematic notch (Figs. 9 and /0)). This notch area remains shallow in most prothalli observed; in some, however, the meristem exceeds the wing tissue and no notch is evident. In some older gametophytes, the meristematic region becomes quite broad, and, rarely, may divide into two sepa- rate regions (Fig. 11) with non-meristematic tissue between. Gametangia initiation in culture was rapid. All cultures, except 193-K!, exhib- ited a female to hermaphroditic gametangial sequence of development (Fig. 8) with the exception of occasional gametophytes which precociously initiated an- theridia (Fig. 7). Of the 20 cultures studied in detail, seven produced some male gametophytes, but the percentage of such gametophytes in culture (except 193-K) was less than 6.0 (Tables ] and 2). In all cases these prothalli rapidly became hermaphroditic. The length of the unisexual female gametophytic stage varied from culture to culture. In one culture (190-D), hermaphroditic prothalli were produced simulta- neously with female prothalli. In other cultures (150-I, 193-E, 193-L), gametophytes remained unisexual and female throughout the culture period. In the remaining cultures, hermaphroditic prothalli were produced (2)6—28 (mean = s.d. = 15.6 + 5.8) days following appearance of female prothalli (Table 1). Gametangial sequences of individual gametophytes are diverse, but the vast majority of gametophytes exhibited a female to hermaphroditic sequence. Arche- gonia were initiated on the cushion immediately behind the young lateral meristem and were produced continuously until the gametophytes were fully cordate witha pronounced elongate cushion with numerous senescent gametangia (Fig. 8). An- theridia were initiated on wing tissue near the apical notch, initially along the margin of the cushion and later outward toward the wing margins (Figs. 9 and / 0). Fully mature hermaphroditic gametophytes exhibited antheridia covering both not been observed in the distal portions of the prothallus. In culture 190-A, sam- phyte ‘Here and elsewhere in this paper the letter following the culture number designates a gameto population originating from a specific sporophyte. 104 AMERICAN FERN JOURNAL: VOLUME 70 (1980) ples of gametophytes 46 days after sowing indicated that some of them had pro- duced up to 90 antheridial initials. These prothalli exhibited up to 28 or more senescent, 8 mature, and 50 immature archegonia. Six days later (52 days follow- ing sowing), fully mature hermaphroditic prothalli were present, exhibiting up to 60 senescent, 10 mature, and 35 immature archegonia, and over 325 antheridia per wing of which about 12% contained mature spermatozoids. Gametophytes which were initially male produced only one or two antheridia prior to the initiation of archegonia. In most cases, following maturation and dehiscence of these antheridia, gametophytes became functionally unisexual and female and their subsequent ontogeny paralleled that described above. TABLE 1. DAYS FROM SOWING TO APPEARANCE OF GAMETOPHYTE-TYPES IN CULTURES OF A. AUREUM. Culture Days to appearance of number Density/cm? Male Female Hermaphroditic 146-C 27.7 28 2 38 146-G 6.5 35 29 (35)50 146-L 6.9 : 50 146-M 21:3 28 46 148-B 21 28 44 150-A 18.0 51 3 (35)56 150-B 29.7 44 150-C 28.2 56 150-E 3.9 } 50 150-I 8.1 3 -! 150-K 14.1 3 44 190-A 6.1 ] 49 se i | : 2. 37 : 6 190-F 31.5 59 oe 193-A 17.6 28 56 193-B 98.3 44 ) _* re 45.7 29 3 5 191 193-L | id : “er ‘sampled up to 76 days ?sampled up to 85 days 3sampled up to 90 days Some of the gametophytes expressed sequential patterns of functional unisexu- ality. In 146-C, about 15% of the sampled gametophytes expressed a sequence of archegonial initiation, maturation, and senescence, followed by initiation and maturation of antheridia. Other gametophytes appear to have gone through func- tional stages in sexual ontogeny from archegoniate to hermaphroditic to an- theridiate to archegoniate. This sequence was noted in several older prothalli in 193-A (Fig. 11) and in one gametophyte of 150-K. In gametophytes producing proliferations near the base, such as those with two 1-dimensional filaments aris- ing from the single basal cell, some of the proliferations were covered with an- theridia, whereas others were only archegoniate. Culture 193-K was unique among those studied in its expression of large num- bers of male prothalli (Table 2). The initial ratio of male to female prothalli, excluding asexual prothalli, was 1:1. As the culture developed, male gameto- R. M. LLOYD: REPRODUCTIVE BIOLOGY OF ACROSTICHUM 105 phytes increased in frequency to a 3:1 ratio; this; in turn, was followed by an increase in both female and hermaphroditic gametophytes. It is of interest to note that at the end of the observation period, there was a 1.1:1 ratio of male and hermaphroditic prothalli to female prothalli. Male gametophytes in this culture frequently were highly elongate and irregularly formed. Many of them initiated antheridia in the early ontogenetic stages following the attainment of a 2-dimensional morphology. In contrast, female prothalli were larger and appeared to be similar in all respects to the female prothalli of the other cultures. TABLE 2. SEXUAL ONTOGENY IN AGAR CULTURES OF A. AUREUM. ‘Sexual expression (%)- Days from sowing Neuter Male Female Hermaphroditic Culture No. 146-C 21 97.5 25 28 48.6 5.4 45.9 35 26.4 74.6 38 8.7 78.3 13.0 44 10.5 2.6 68.4 18.0 47 : J 40.0 34.2 Culture No. 190-A 33 00.0 AY 64.7 bo 43 22.2 77.8 49 29.4 47.1 235 52 28.6 31.4 40.0 55 11.1 40.7 48.1 58 6.7 33.3 60.1 76 14.3 85.7 87 100.0 Culture No. 193-K 30-33 8.3 29. 1.9 39-45 217 56.7 19.1 25 48-54 15.9 32.9 50.0 1.2 63 24.4 46.3 29.3 Culture No. 193-L 30 28.6 40 58.3 41.7 42 38.5 61.5 100.0 90 100.0 Parameters of gametophyte morphology and ontogeny discussed above suggest that intergametophytic mating should be prevalent in gametophyte populations of Acrostichum aureum. Specific factors exhibited by the gametophyte generation which increase the probability for intergametophyte mating include the female to hermaphroditic gametangial sequence in most gametophytes studied, the dioeci- ous condition expressed in some populations, and the sequential functional uni- sexuality expressed in some gametophytes of some populations. Additional sup- 106 AMERICAN FERN JOURNAL: VOLUME 70 (1980) port for this assessment comes from observations on sporophyte production in composite cultures and in timing of the appearance of sporophytes in isolate vs. composite cultures. For example, in culture 146-M, after 83 days 90.9% of the prothalli were unisexual female and 9.1% were hermaphroditic. Examination of sampled prothalli indicated that all of the unisexual female gametophytes had produced one or two young embryos, which would only be possible through intergametophytic mating. In addition, in nearly all cultures sampled sporophyte production in composite culture occurred between 17 and 31 days earlier than in isolate culture. As the gametangial sequence is from female to hermaphroditic in these prothalli, in composite cultures spermatozoids produced by the early her- maphroditic gametophytes will fertilize many of the unisexual female prothalli. In contrast, in isolate culture, each of the gametophytes must become hermaphrodi- tic prior to sporophyte production. This sequence of events has been well documented in studies on Ceratopteris by Klekowski (1970a). TABLE 3. FREQUENCY OF DELETERIOUS SPOROPHYTIC GENOTYPES IN IN- TRAGAMETOPHYTICALLY SELFED, ISOLATED HERMAPHRODITIC GAMETOPHYTES OFA. AUREUM. Number No. (%) No. (%) No. (%) late No. (%) Culture tested No. (%) zygotic embryonic sporophytic lea number prothalli normal lethals lethals lethals lethals 46-N 19 18(94.7) 1(5 148-B 20 19(95.0) 1(5.0) 150-A 20 19(95.0) 1(5.0) 150-B 20 17(85 .0) 1(5.0) 2(10.0) 150-I 18 17(94.4) 1(5.6) 190-F 19 15(78.9) 3(15.8) 1(5.3) 191-B 19 17(89.5) 1(5.2) 1(5.2) 192-A 40 39(97.5) 1(2.5) 193-C 20 17(85.0) 1(5.0) 2(10.0) 193-E 20 11(55.0) 9(45.0) 193-F 8 7(87.5) 1(12.5) 193-M 20 19(95.0) 1(5.0) All others:! 511 $11(100.0) Totals: 754 726(96.3) 20(2.65) 2(0.26) 1(0.13) 5(0.66) ‘Includes 27 cultures: 146-C, 146-G, 146-I, 146-J, 146-L, 146-M, 146-O, 150-C, 150-E, 150-H, 150-K, Ee ee oe OS, 193-D, 193-G, 193-H, 193-1, 193-K, 193-L, Additional information relative to reproductive biology can be obtained by analyzing frequency of deleterious or lethal genes (genetic load) expressed in sporophytes (Klekowski, 1979). To analyze genetic load in A. aureum, from 20 to 40 gametophytes per sporophyte (= a gametophyte family), prior to the attainment of sexual maturity, were selected at random from stock cultures and individually isolated in 60 x 20 mm petri dishes containing nutrient agar. Following growth, cultures were watered twice weekly to facilitate fertilization, and the resultant sporophytes were allowed to develop to the third frond stage. Results of these studies are given in Table 3. R. M. LLOYD: REPRODUCTIVE BIOLOGY OF ACROSTICHUM 107 Genetic load was determined as the percentage: of the hermaphroditic gametophytes per gametophyte family which failed to yield normal sporophytes. Families exhibiting genetic load in a portion of the gametophytes tested are con- sidered to be expressing heterozygosity in their gametophytic genotypes. Expres- sions of genetic load were in the form of zygotic lethals (in 2.65% of the 754 gametophytes tested), embryonic lethals (in 0.26%), late sporophytic lethals (in 0.13%) and leaky lethals (in 0.66%) (see Klekowski, 1970b, 1979, for complete discussion of these genetic expressions). It is significant to note that 96.3% of the tested prothalli did not exhibit any deleterious or lethal genotypes. Of the 39 sporophytes tested, 27 (69.2%) were devoid of genetic load (Table 4). In the 12 sporophytes expressing load, it varied from 5.0% (cultures 148-B, 150-A, 193-M) to 45.0% (culture 193-E). The mean genetic load for all plants tested was 3.7%. TABLE 4. GENETIC LOAD IN A. AUREUM RELATIVE TO SIZE AND LOCATION OF THE POPULATION. Population oe Size (est. Range (x) % No. plants No. (%) plants number Location no. plants) genetic load tested with genetic load 193 Panama 3000 0-45.0 (5.96) 13 4(30.8) 190 Jamaica 1000 021.1 (4.22) > 1(20.0) 191 Guyana 400 0-10.4 (5.2) 2 1(50.0) 150 Florida 75-100 0-15.0 (3.65) 7 3(42.8) 146 Florida 25-50 0-5.3 (0.66) 8 1(12.5) 148 Florida 20 5.0 1 1(100.0) 194 Panama 15 0 1 000.0) 192 Surinam 8 0-2.5 (1.25) 2 1(50.0) Total: 0-45.0 (3.24) 39 12(30.8) Although the number of plants tested from each population is insufficient for Statistical comparison, it is of interest to note that those sporophytes exhibiting the higher genetic load values are found in the larger populations and that the small populations (with 50 or fewer individuals) have very low levels of recessive deleterious or lethal genes (Table 4). Leaky lethal expression (Klekowski, 1970b) was noted in three gametophyte families. In 150-B, normal sporophytes appeared on the two prothalli 165 days after sowing and 38 days following normal sporophyte production of the remaining prothalli tested. Each of these two prothalli exhibited several abortive embryos, indicating that previous selfing had occurred involving lethal genetic combina- tions. In 193-M, the first sporophyte which appeared was abnormal and exhibited a long, cylindrical, tubular growth with ruffled margins. Subsequent sporophytes from other fertilizations produced normal fronds. ; Apomictic proliferations were noted on only one gametophyte in 147-L. Ninety days after sowing, this prothallus proliferated a blade of tissue bearing rhizoids on one margin and small epidermal cells similar to those found on young sporophytes. Irregularly organized vascular tissue was present near the base of this blade, but no roots or stomata were noted. 108 AMERICAN FERN JOURNAL: VOLUME 70 (1980) DISCUSSION Gametophyte Morphology.—The gametophyte morphology and ontogeny of A. aureum is remarkably similar to that of A. danaeifolium (Lloyd & Gregg, 1975) and agrees in most respects with that of A. speciosum Willd. (Stokey & Atkinson, 1952). Spores of A. aureum are almost identical in size and shape to those of A. danaeifolium; however, there are minute differences in spore surface markings, especially the more pronounced tuberculate pattern exhibited by A. aureum. Other gametophyte features which are qualitatively similar between the two species are formation of the 1-dimensional filament, the lateral meristematic re- gion of the 2-dimensional prothallus, the relatively shallow apical notch region (however, protruding beyond the wing tissue in some prothalli of A. aureum), the female to hermaphroditic gametangial sequence, and the sexual expression in gametophyte families grown in composite culture. The major difference between gametophytes of the two species is the distribu- tion of antheridia, which are mostly restricted to the apical wing and meristem region of A. aureum, but also are found in more basal regions along the cushion margins and among the rhizoids in A. danaeifolium. In addition, the sequential production of archegonia-antheridia-archegonia in some prothalli of A. aureum is unknown in the other species. It is apparent from both sporophyte and gametophyte studies that these two species are closely related. Further evidence in support of this is their ability to freely hybridize in culture and to produce normal viable F: sporophytes, although these sporophytes have not yet been grown to maturity to measure chromosome homology (Lloyd, unpubl.). Reproductive Biology.—Sex ontogeny in most cultures of Acrostichum aureum sampled in this study is female to hermaphroditic or initially dioecious. The length of the unisexual stage prior to the attainment of bisexuality is sufficient to facili- tate intergametophytic mating. The facility for such mating is also evidenced in culture by the rapidity of embryo formation in unisexual prothalli following the initiation of antheridia on just one gametophyte in a composite culture. Thus, the gametophytic developmental pathway must be considered as one which has a higher probability of intergametophytic mating than of intragametophytic selfing. However, correlative heterozygosity in the form of genetic load is insufficient in naturally occurring sporophytes to suggest that outbreeding is a normal occul- rence. For example, of the tested plants 69% exhibited no heterozygosity for recessive deleterious genes and 15% exhibited such genes in less than 6% of the genotypes sampled. As intergametophytic mating is strongly suggested by the culture experiments, if the assumption is made that these plants are genetically homozygous due to the lack of genetic load expression, other factors must be superimposed upon the hypothesized mating system which are more significant in determining the genetic composition of the populations as a whole. First and foremost, the culture methodology as used in these experiments may be insufficient to document with accuracy the gametangial sequences as they are realized in nature. In parallel experiments on A. danaeifolium, gametophytes R. M. LLOYD: REPRODUCTIVE BIOLOGY OF ACROSTICHUM 109 grown on soil exhibit greater antheridial production (Lloyd & Gregg, 1975). Al- though some of these gametophytes undergo a male to hermaphroditic gametan- gial ontogeny, dioecism in cultures was still highly prevalent, suggesting that soil grown gametophyte populations in nature would have higher probabilities of in- tergametophytic mating. As gametophytic ontogenies on agar cultures of A. au- reum and A. danaeifolium are highly similar, it is reasonable to assume that the gametophytes of A. aureum would present similar responses to soil culture. How- ever, the habitat of A. aureum is at least partially inundated by tides, suggesting that the soil component for gametophyte populations will contain higher levels of salts. Brief experiments by Stokey & Atkinson (1952) using dilute sea water as part of the culture medium induced restricted growth of gametophytes of A. speciosum. This type of reduced growth under less than optimal conditions fre- quently leads to the initial production of antheridia and can prevent formation of viable archegonia (Page, 1979). Thus, it is possible that the gametophytic on- togenies in the culture experiments reported here do not represent gametophytic ontogenies as realized in nature. Other factors which undoubtedly have a significant influence are population size, spore output per plant, the influence of the specific aquatic habitat, and the genetic system. It is of interest to note that the highest levels of genetic load were found in the larger populations, suggesting that the frequency and success of recombinants increases with number of individuals as well as age of the popula- tion. As spore production by each individual of A. aureum is massive, it is proba- ble that inbreeding (in this case, intergametophytic selfing) will occur until such time as there is sufficient spore intermixing to increase the likelihood of outbreed- ing. The influence of the aquatic habitat may play an important role in the selection of specific genotypes, perhaps perpetuated by intragametophytic selfing. It is significant to note that work to date on other aquatic species, including Acro- stichum danaeifolium, Ceratopteris thalictroides and C. pteridoides, has provided highly similar results. These species are all characterized by a gametophyte on- togeny which favors intergametophytic mating (including an antheridogen in Ceratopteris spp.), but the vast majority of individuals tested express little or no heterozygosity in the form of genetic load. In this regard, Baker (1965) cites seashores and the margins of salt marshes as open habitats where species which are inbreeding with ‘‘general purpose genotypes’? may be advantageous. Angio- sperms which occupy these open and disturbed types of habitats are generally found to be autogamous or apomictic and so are unable to build up recombinants in the population rapidly. : Lastly, the genetic system of pteridophytes must be considered. We still have little understanding of the polyploid system and the maintenance and expression of heterozygosity in these organisms. It is possible that most of them are highly heterozygous and that genetic load is effectively screened from expression. If so, our current methodology for analysis for heterozygosity is insufficient. It is obvious from these studies that we have little understanding of fern mating systems as they operate in nature and much further work, especially that oriented 110 AMERICAN FERN JOURNAL: VOLUME 70 (1980) toward the genetic system and natural populations of gametophytes and sporophytes, is required before we will be able to circumscribe adequately these phenomena as they operate in nature. This work has been supported by National Science Foundation Grants Nos. GB-36923, BMS 75-07191, and DEB 79-05079. I would like to thank T. R. Warne, D. Buckley, and S. Buckley for their assistance in the laboratory. LITERATURE CITED ADAMS, D. C. and P. H. TOMLINSON. 1979. Acrostichum in Florida. Amer. Fern J. 69:42-46. BAKER, H. G. 1965. Characteristics and modes of origins of weeds. Jn H. Baker and G. L. Stebbins, eds. The Genetics of Colonizing Species. Academic Press, New York. HOLTTUM, R. E. 1954. Flora of Malaya, Vol. II. Ferns of Malaya. Gov't. Printing Office, Singapore. KLEKOWSKI, E. J., JR. 1969. Reproductive biology of the Pteridophyta. III. A study of the Blechnaceae. Bot. J. Linn. Soc. 62:361-377. . 1970a. Reproductive biology of the Pteridophyta. IV. An experimental study of mating selene in Ceratopteris thalictroides (L.) Brongn. Bot. J. Linn. Soc. 63:153-169. . 1970b. Populational and genetic studies of a homosporous fern—Osmunda regalis. Amer. a Bot. 56:1122-1138. . 1979. The genetics and reproductive biology of ferns. In A. F. Dyer, ed. The Experimental Biology of Ferns. Academic Press, London LLOYD, R. M. 1974. Reproductive biokiay and évohiion in the Pteridophyta. Ann. Missouri Bot. Gard. 61:318-331. , and T. L. GREGG. 1975. Reproductive =a and gametophyte morphology of Acro- stichum danaeifolium from Mexico. Amer. Fern J. 65:105—120. , and T. R. WARNE. 1978. The absence of aid load in a morphologically variable sexual cee a cigar sae aia ce ponpanen “ie Bot. 3:20-36. PAGE, C. N. 1979. Experimental as of _ ecology. In A. F. Dyer, ed. The Experimental okey of Ferns. Academic eg Lon SMALL, J. K. 1938. Ferns of the Southeastern ee Reprint ed., 1964. Hafner, New York. STOKEY, A. G. and L. R. ATKINSON. 1952. The gametophyte of Acrostichum speciosum Willd. Phytomorphology 2:105-11 WALLACE, B. 1970. Genetic Load, Its Biological and Conceptual Aspects. Prentice-Hall, En- glewood Cliffs, NJ. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 3 (1980) 111 SHORTER NOTES DIPLAZIUM JAPONICUM NEW TO ALABAMA.—In the summer of 1977 while I was a student at Auburn University, J. I. Glick, another student, asked me to confirm his identification of a fern frond he had collected from a plant growing in a steep-sided, wooded ditch on the Auburn campus in an area that had been allowed to remain wild by the university groundskeepers. Glick had identified the frond as Athyrium thelypteroides. However, Auburn is in Lee County of east-central Ala- bama, far south of the range of that fern, and the fern did not look exactly like A. thelypteroides. 1 identified the frond as Diplazium japonicum (Thunb.) Bedd., a native of eastern Asia (Glick s.n., Short 979, both AUA). According to Wherry (Southern Fern Guide, 1964), this fern had been found previously in one Florida locality. Wherry was doubtful whether the fern was cultivated locally and did not know the source of the spores which produced the plants. The species is not known to be in cultivation in the Auburn area, and the source of these plants is equally unknown. To my knowledge, D. japonicum has not been reported from any other localities in the southeastern United States. The Auburn population consisted of two mature, spore-bearing plants and several juveniles. The plants seem to be estab- lished well enough to be considered naturalized. The winters previous to and subsequent to the plants’ discovery were among the most severe on record, but they had no adverse effect on the plants. Other ferns found in the ditch include Asplenium platyneuron and the naturalized Lygodium japonicum, Pteris multifida, and Thely- pteris torresiana.—John W. Short, 905 McKinley Ave., Auburn, AL 36830. MOTHS AND FERNS.—In a previous paper (Amer. Fern J. 61:166-170. 1971), 1 reported on a nymphalid-like moth that oviposits on C. 'yathea holdridgeana in such a fashion that the eggs mimic the immature sori of the fern. Recently, I have found the larvae of a microlepidopteron predating the laminar tissue of Cnemidaria mutica, C. choricarpa, and Sphaeropteris brunei, of the family Cyatheaceae. In their last stages, the caterpillars weave cocoons with silk, spores, and sporangia. The adults obtained from the three ferns are the same species of moth, a species yet to be determined. Another moth lays eggs on the fronds of Botrychium dissec- tum. The larval stages feed on it and later spin a cocoon by sewing together two segments or pinnules. The most interesting relationship between moths and ferns I have encountered So far is that of Hymenophyllum myriocarpum and a microlepidopteron whose larvae feed on the filmy fern and pupate in a case made up of the folded segments, which then resemble mature involucres in their size, color, and position. It is surprising to find that the insects only use the basal, lower, and middle pinnae of the fern fronds to build their cocoons, perhaps to guarantee themselves an appro- priate relative humidity among the mosses or a more efficient camouflage. Cer- tainly, a careful survey of tropical ferns will reveal that they are not so impervious to insect attack as they commonly are thought to be.—Luis D. Gomez P., Her- bario Nacional, Museo Nacional de Costa Rica, Apartado 749, San Jose, Costa ica. 112 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 2 (1980) THREE ADDITIONS TO THE PTERIDOPHYTE FLORA OF ESCAMBIA COUNTY, FLORIDA. — In 1978 a number of students from the University of West Florida conducted a field survey of the pteridophytes of Escambia County, Florida, concentrating on the botanically largely unexplored northern part. Just north of where Florida Highway 4 crosses Canoe Creek (R31W, TSN, Sec. 8), eight fern species were found along the west bank, including two which we had not observed previously in the county: Thelypteris torresiana (Gaud.) Alston (Burk- halter & Booker 5844, 5912) and Athyrium asplenioides (Michx.) A. A. Eaton (Burkhalter & Booker 5845, 5913). Vouchers have been deposited at the Univer- sity of Florida, Gainesville (FLAS), and the University of West Florida, Pen- sacola (UWFP). Dr. Daniel B. Ward (pers. comm., 26 Feb. 1979) verified that these species had not been collected previously in Escambia County. The collection site is an open area with mostly clay and sand soil. In addition to a number of weedy flowering plants like Carex spp. and Boehmeria cylindrica, the following ferns were found: Asplenium platyneuron, Lygodium japonicum, Os- munda regalis, O. cinnamomea, Thelypteris normalis, and Woodwardia areolata. Thelypteris torresiana and A. asplenioides are not found in the shaded swamp forest understory somewhat to the north of the collection site; they were, how- ever, observed at a few other scattered locations in northern Escambia County. Later, a small, spontaneous colony of Nephrolepis cordifolia (L.) Presl was discovered in downtown Pensacola on an old, brick building at the northeast corner of Palafox and Main Streets. The plants were securely anchored in the brickwork about seven feet above the sidewalk, and were situated below the broken end of a raingutter downspout, from which they received plentiful water. This is not an unusual habitat for certain ferns in urban areas like New Orleans and Mobile, and the colony appeared to be quite old and healthy. Clifton Nauman (pers. comm., 31 Aug 1978) commented that the specimen (Burkhalter 5919, UWFP) constituted a new record for Escambia County.—James R. Burkhalter, 3703A W. Brainerd St., Pensacola, FL 32504. TRIARCH Over 5©@ Wears of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED .O. Box 98 Ripon, Wisconsin 54971 ee a anise tence rene era AMERICAN ae FERN ~ yma MISSOURY ROTENICAT October oe December, 1980 JOURNAL say 3 hed V5 GARDEN LIBRARY coos QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY A Range Extension for Dryopteris filix-mas ERWIN F. EVERT 113 Differential Germination of Fern and Moss Spores in Response to Mercuric Chloride RAYMOND L. PETERSEN and PATRICK C. FRANCIS |! a Differences in the Apparent Permeability of Spore Walls and Prothallial Cell Walls in Onoclea sensibilis JOHN H. MILLER 119 Allelopathy and Autotoxicity in depen Easter d North American Ferns LLIAM a MUNTHER and DAVID E. FAIRBROTHERS 124 Shorter Notes: Sandstone Rock Crevices, an Exceptional New Habitat for Thelypteris simulata; A Second Alabama Locality for 136 the Hart’s-tongue is ‘ 138 American Fern Journal 138 Index to Volume 70 “ Errata es _ A34 The American Fern Society Council for 1980 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, Ohio 4570 President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN ee Vice President LESLIE G. HICKOK, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, Tenn. 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, Va. 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, D.C. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, han Pair 94720 Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, N.Y. 10458. Newsletter Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, D. C. 20560 ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Sem 34 Indiana University, Bloomington, Ind. 47401 JOHN T. MICKEL w York Botanical Garden, Bronx, New York 10458 The “American Fern Journal’’ (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. Itis owned by the American Fern Society, and published at the Smithsonian Institution, Washington, DC 20560 - Second-class postage paid at Washington: Matter for publication and g 3-6 months after the date of issue) should be addressed to the Editor. Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog, Dept. of — George Mason University, Fairfax, Va. 22030. rs for back issues should be addressed to he Treasurer. Sey inquiries concerning ferns should be addressed to the Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Fern Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or Sei. $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, New York 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter - John T. Mickel, New York ee aera HOO, mer York Jee. - editor of the newslet- ter "Fiddlehead —— ogc edito members, includ- rites or purchase materials, personalia, tarietcn notes, and reviews of mages! pata on Pais Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th — Seattle, Washington 98115, is Director. Spores exchanged and collection lists sent on reque sits ‘al Peanesit fichc sun tk ee eke i ee ree ‘: om 1 * eorected in ferns. Botanical books, back i Satine of the Journal, and cash or other gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secretary. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) 113 A Range Extension for Dryopteris filix-mas ERWIN F. EVERT* On 14 December 1979, I collected Dryopteris filix-mas (L.) Schott in a mesic ravine in Glencoe, Cook County, Illinois. Two large plants and a smaller one were growing within a few feet of each other on a steep, relatively undisturbed northwest- facing slope of the ravine, about 400 feet west of Lake Michigan in T42N, R13E, SE1/4, Sec. 6. Associates were Acer saccharum, Aralia nudicaulis, Hamamelis virginiana, Quercus rubra, and Trillium grandiflorum. A search of other ravines in the vicinity failed to reveal any other plants of D. filix-mas. This fern has not been reported previously for Illinois (Mohlenbrock & Ladd, 1978). It is newly reported for Wisconsin (Brown County) by Peck and Taylor (1980). The nearest stations appear to be in Marquette County, Michigan (Billing- ton, 1952, p. 177) and in Brown County, Wisconsin, about 270 and 160 miles distant. Are these plants spontaneous on this site or is their occurrence due to the actions of man? Although this question cannot be answered with absolute certainty, the following points support natural occurrence: (1) The plants grow in an environment typical for the species, in a relatively undisturbed area with autochthonous associates. (2) Pepoon (1927), Moran (1978), Swink and Wilhelm (1979), and personal observation document the presence of many other uncommon plants for this area in the ravines along Lake Michigan. Some of these are: Dryopteris intermedia, D. marginalis, Equisetum scirpoides, Fagus grandifolia, Lycopodium lucidulum, Mit- chella repens, Pinus resinosa, Polystichum acrostichoides, Shepherdia canadensis, and Thelypteris hexagonoptera. (3) Personal observation indicates that D. filix-mas is not commonly cultivated in this area at present. Therefore, it is not likely that these plants have escaped from cultivation. ‘Although the Male Fern formerly was grown for its medicinal properties, the colony does not appear to have been established for a long time, and so it is unlikely to have originated from plants that were cultivated in the past. It is also unlikely that plants on such a steep and inaccessible site were deliberately planted. (4) Single plants or small colonies of other ferns with northern distributions are known to occur widely disjunct in southern Michigan (Wagner, 1972, p. 205 and pers. comm.). Examples include Botrychium minganense, Gymnocarpium dryopter- is, and Polystichum braunii. (5) Dryopteris filix-mas, a homosporous pteridophyte, apparently is capable of intra-gametophytic selfing (W. H. Wagner, pers. comm.), and so one wind-borne Spore could produce a new, disjunct colony. ee (6) The distribution pattern of D. filix-mas (Hultén, 1962, p. 119), with its widely disjunct stations in California, Mexico, South America, Hawaii, Greenland, Iceland, and Africa, indicates that this species is capable of wide dispersal. *1476 Tyrell St., Park Ridge, TL 60068. Volume 70, number 3, of the JOURNAL was issued September 29, 1980. 114 AMERICAN FERN JOURNAL: VOLUME 70 Specimens of Evert 1651 have been deposited at the Morton Arboretum Herbari- um (MOR) and the University of Michigan Herbarium (MICH). LITERATURE CITED BILLINGTON, C. 1952. Ferns of Michigan. Cranbrook Institute of Science, Bloomfield Hills, MI. HULTEN, E. 1962. The Circumpolar Plants 1, Vascular Cryptogams, Conifers, and Monocotyledons. Kungl. ‘Ape Vetenskapsakad. Handl. 8(5):1-275. MOHLENBROCK, R. and D. M. LADD. 1978. ae of Illinois Vascular Plants. Southern Illinois pula cals Carbondale, IL. MORAN, R. C. 1978. Vascular flora of the ravines along Lake Michigan in Lake County, Illinois. 0 M : PECK, J. H. and W. C. TAYLOR. 1980. Check list and distributions of Wisconsin ferns and fern allies. Michigan Bot. 19:252-268. PEPOON, H. S. 1927. An Annotated Flora of the Chicago Area. Academy of Sciences, Chicago, IL. SWINK, F. and G. WILHELM. 1979. Plants of the Chicago Region, rev. & expanded ed. Morton Arboretum, Lisle, IL. WAGNER, W. H., Jr. 1972. Disjunctions in homosporous vascular plants. Ann. Missouri Bot. Gard. 59:203-217. AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) 115 Differential Germination of Fern and Moss Spores in Response to Mercuric Chloride RAYMOND L. PETERSEN and PATRICK C. FRANCIS* The pervasive use of toxic substances and the synthesis of myriad new such compounds requires that toxic substance limits be set and that environmental monitoring be maintained in order to insure the health and integrity of the biosphere. Bioassays are the primary mechanism for meeting these two require- ments. A bioassay designed to determine the environmental impact of a particular toxic substance, for example divalent mercury, should have the following attributes: ease of performance, low cost, brevity, sensitivity, and high overall reflectivity of environmental stress induced by the presence of the toxic substance. The appropri- ateness of a bioassay is based on the selection of its two primary components—the living system employed and the parameter measured. Petersen et al. (1980) have demonstrated the feasibility of using the germination of Onoclea sensibilis spores to gauge the toxicity of various heavy metal ions. It was found that for the three metal ions tested, toxicity was directly proportional to atomic weight. Hg** is twice as toxic as Cd** and four times as toxic as Co**. The present study is a comparison of the germination responses of different fern and moss spores to divalent mercury. There are few investigations on the effects of metal ions and other potential pollutants on fern spores and gametophytes that yield data pertinent to pollution research and control. Nakazawa and Tsusaki (1959) determined that the fern spore cytoplasm associated with rhizoid differentiation has a marked affinity for, metal ions. Nakazawa and Otaki (1962) demonstrated the affinity of developed rhizoids for metal ions. Fern rhizoids apparently function like the root hairs of vascular plants in absorbing water and minerals from the soil. Therefore, their affinity for metal ions is not surprising. Co** and Ni** were shown to prolong filamentous (one- dimensignal) growth in Lygodium smithianum gametophytes (Parés, 1958) resulting in a retardation of their development sequence. LiCl caused a precocious differentia- tion of terminal papillae (gland-like hair cells produced by some fern gametophytes) in Dryopteris varia (Nakazawa, 1960a, b). Several metal chlorides at concentrations of 0.005—-0.08M decreased the period of fern sperm motility (Igura, 1958). A few papers on ferns with direct application to pollution monitoring have been published. Klekowski (1976), Klekowski and Berger (1976), and Klekowski and Poppel (1976) found that meiotic chromosome behavior during fern sporogenesis was correlated with the presence of toxic substances in the environment. Howar and Haigh (1972) studied the effects of increasing doses of X-radiation on the first mitotic division of Osmunda regalis spores. Edwards and Miller (1970, 1972a, b) Studied the quantitative effects of ethylene on Onoclea sensibilis spore germination and gametophyte growth. Fern gametophytes also have been successfully employed In bioassay procedures for the plant hormones kinetin, gibberelic acid, and antheridogen (Bopp, 1968; Brandes, 1973). *Department of Botany, Howard University, Washington, DC 20059. 116 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Mosses are especially adapted to absorb and concentrate substances present in the atmosphere, and so moss gametophytes have proven to be sensitive indicators of airborne pollutants (Huckabee, 1973: Skaar et al., 1973; Little & Martin, 1974). Francis and Petersen (1980) reported on the synergistic effects that metal ion combinations had on the germination of Polytrichum commune spores. pores of Onoclea sensibilis, Osmunda cinnamomea, and Osmunda claytoniana and the moss Polytrichum commune were selected for investigation because they are of wide distribution and frequent occurrence and all produce copious quantities of easily collected spores. In addition, O. sensibilis had been employed in the development of the fern spore heavy metal bioassay (Petersen, et al., 1980). The two Osmunda species represented a fern family distinct from that of Onoclea and would yield data on intrageneric differences in response to heavy metal ions. Polytrichum commune was selected to compare its heavy metal response with those of the ferns. TABLE 1. SPORE GERMINATION IN VARIOUS Hg* * CONCENTRATIONS EXPRESSED AS A PERCENTAGE OF CONTROL GERMINATION. ' Hg* + ppm O. sensibilis O. cinnamomea O. claytoniana P. commune 0.0 100 100 100 100 0.2 101 98 96 77 0.4 98 91 89 30 0.6 98 Ti 33 8 0.8 88 60 20 0 1.0 85 13 14 0 LS 75 8 8 0 2.0 68 1 1 5 3.0 55 0 0 r 4.0 30 zi z 6.0 9 a a 8.0 l w * 10.0 0 i i 20.0 0 1 * Tests were not run, or the experiment was lost, where no numerical value is given. MATERIAL AND METHODS _Approximately 10,000 spores of each species were cultured in a petri dish 60mm diam. in 8 ml of full strength liquid Knudson’s medium at a pH 5.5 with 0-20 ppm of divalent mercury ion (Hg* *) added as HgCl). The dishes were sealed in clear Plastic sandwich bags and cultured in a Sherer Growth Chamber at 20°C in 300 ft-c of continuous light from cool-white fluorescent lamps. Spore germination was considered to have occurred when the spores produced a rhizoid or the first prothallial cell divided. After eight days, 500 spores were examined per plate. Three replicates were run for each species tested at each Hg* * concentration. PETERSEN & FRANCIS: DIFFERENTIAL SPORE GERMINATION 117 RESULTS AND DISCUSSION The germination of the controls (0 ppm HG**) was uniformly high; O. sensibilis had 88%, O. cinnamomea 92%, O. claytoniana 96%, and P. commune 94%. Spore germination at each Hg* * concentration except one was lower than in the controls. Germination was expressed as a percentage of each species’ control germination (Table 1). These results are summarized in terms of the standard bioassay toxicity values of LCs9 and LCj99, which are presented both in terms of ppm and wM (Table 2). LCs is the concentration of a toxic substance necessary to kill 50% of the organisms from a control value, and LCjo9 is the minimum concentration necessary to kill 100%. TABLE 2. Hg* * LCs) AND LCjo99 VALUES FOR FOUR TAXA. Toxicity i O. sensibilis O. cinnamomea O. claytoniana P. commune LCso ppm 3.2 0.82 0.53 0.30 uM 16 4.1 2.6 1.5 LCi ppm 10 3.0 3.0 0.8 uM 50 15 15 4.0 A comparison of spore germination responses shows that P. commune is the taxon most susceptible to HG** and is ten times more sensitive than O. sensibilis, which is the least susceptible. The two Osmunda species have intermediate and similar HG* * toxicity values. They are approximately four times more sensitive to HG** than is O. sensibilis (Table 2). Therefore, based on the sensitivity of response, P. commune would be the species of choice in a mercury ion bioassay. We wish to acknowledge financial aid from NSF Grant SM177-05566. LITERATURE CITED BOPP, M. 1968. Control of differentiation in fern-allies and bryophytes. Ann. Rev. Pl. Physiol. 1-380 19:3 ; BRANDES, H. 1973. Gametophyte development in ferns and bryophytes. Ann. Rev. Pl. Physiol. 24:115-128. EDWARDS, M. E. and J. H. MILLER. 1970. Inhibition of cell division by ethylene in fern gametophytes under various light conditions. Pl. Physiol. (Lancaster) 46(Suppl.):32. , and J. H. MILLER. 1972a. Growth regulation by ethylene in fern gametophytes. Il. Inhibition of cell division. Amer. J. Bot. 59:450-457. . and J. H. MILLER. 1972b. Growth regulation by ethylene in fern gametophytes. III. Inhibition of spore germination. Amer. J. Bot. 59:458-467. oe FRANCIS, P. C. and R. L. PETERSEN. 1980. Synergistic behavior of heavy metal combinations by the fern and moss spore germination bioassay. Assoc, Southeastern Biol. Bull. 27(2):46 (abstract). HOWARD, A. and M. V. HAIGH. 1972. Radiation responses of fern spores during their first cell-cycle. Int. J. Radiat. Biol. 18:147-158. ; HUCKABEE, J. W. 1973. Mosses: sensitive indicators of airborne mercury pollution. Atmosph. Environ. 7:749-754. 118 AMERICAN FERN JOURNAL: VOLUME 70 (1980) IGURA, I. ASS. Cytological and morphological apa on ie gametophytes of ferns. XI. The ife of the fern spermatozoid. Bot Mag. Tokyo 71:37-42. pecan E. J., Jr. 1976. Mutational load-in a fern population growing in a polluted environment. r. J. Bot. 63:1024-1030. ‘ — B..B. BERGER. 1976. reenine mutations: ina =m ar growing in a polluted envi t r. J. Bot. 63:239-246. ; _ D. M. POPPEL. 1976. Ferns: potential in-situ. bioassay ile for aquatic-borne tagens. ene Fern. J. 66:75-79. LITTLE, . and oe . MARTIN. 1974. Biological monitoring of heavy metal pollution. Environ. Pollut oe NAKAZAWA, s. ne Cytodifferentiation patterns of Dryopteris protonema modified by some chemical agents. Cytologia 25:352-361 1960b. Aceleracion y retardo en la morfogénesis de la fase protonema del protalo de los helelchos. Anal. Inst. Biol. Univ. México 32:191-1 , and A. TSUSAKI. 1959. Appearance of * ‘metallphilic cytoplasm” as a prepattern to the differentiation of rhizoid in fern protonema. Cytologia 24:378-388. I. 1962. A prepattern to the papilla differentiation in Dryopteris gametophytes. 3 Phyton (Buenos Aires) 18:113-120. PARES, Y. 1958. Etude experimentale de la morphogenése du gamétophyte de quelques Filicinées. Ann. Sci. Nat. Bot., Il, 19:1-120. PETERSEN, R. L., D. ARNOLD, D. E. LYNCH, and S. A. PRICE. 1980. A heavy metal bioassay on percent spore germination of the sensitive fern, Onoclea sensibilis. Bull. Environ. Con- tam. Toxicol. 24:489-4 SKAAR, H., E. OPHUS, and B. M. GULLVAG. 1973. Lead accumulation within nuclei of moss leaf cells. Nature 241:215-216 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) 119 Differences in the Apparent Permeability of Spore Walls and Prothallial Cell Walls in Onoclea sensibilis JOHN H. MILLER The spores of Onoclea sensibilis L. undergo marked changes in their apparent permeability during germination. For example, spores do not stain with an acetocarmine-chloral hydrate mixture until germination has proceeded for about 20 hr (Towill & Ikuma, 1975; Fisher & Miller, 1978), whereas after that time stain is absorbed readily. During the early stages of germination, the spores of Onoclea, Matteuccia struthiopteris (L.) Tod., and other species are difficult to prepare for electron microscopy because embedding resins fail to penetrate adequately (Mar- engo, 1973; Gantt & Arnott, 1976; Raghavan, 1976). Spores in later stages can be processed with no difficulty. This impermeability seems to be associated with the inner spore coat (intine). Fisher and Miller (1978) noted that when the intine of Onoclea spores was artificially ruptured during the early stages of germination and the protoplast was directly exposed to acetocarmine-chloral hydrate, the protoplast stained rapidly, whereas no intact spores could be stained. The time at which Matteuccia and Onoclea spores normally become penetrable by embedding resins coincides with the time the intine ruptures naturally during germination (Gantt & Arnott, 1965; Marengo, 1973). If the intine of Onoclea is caused to open by treatment with sodium hypochlorite, even dormant spores and those in early stages of germination may be infiltrated readily with embedding resin (Bassel, Kuehnert & Miller, 1981). Dormant spores of both Onoclea and Matteuccia have a loose outer spore coat and a thick intine, along which there is a longitudinal seam, the raphe (laesura), on the flattened, proximal face of the spore; the spore protoplast is naked within the intine (Gantt & Arnott, 1965; Bassel, Kuehnert & Miller, 1981). Germinating spores synthesize a new wall around the protoplast inside the intine between 8 and 16 hr. At the time the intine ruptures and is cast off, this new wall becomes the bounding wall of the young protonema. Clearly there is a difference in which materials will penetrate the spore intine and which will cross the normal prothallial cell wall. Carpita et al. (1979) published a method for obtaining quantitative information about the apparent capillary pore size of plant cell walls, which limits the passage of solutes. The cells are placed in a solution of a non-ionic solute having a water potential lower than that of the cells, which causes water to leave the cells. If the solute can pass through the wall, and thus the solution can be in contact with the plasma membrane, one observes plasmolysis (retraction of the protoplast from the cell wall). If, however, the solute particles are too large to penetrate the wall, exit of water from the cell causes cytorrhysis (collapse of the cell wall around the protoplast as it shrinks). I applied this technique to spores and young gametophytes of Onoclea and observed differences in the apparent capillary pore sizes of the spore intine, Prothallial cell walls, and rhizoid walls. “Department of Biology, Syracuse University, Syracuse, NY 13210. 120 AMERICAN FERN JOURNAL: VOLUME 70 (1980) MATERIALS AND METHODS Sporophylls of Onoclea sensibilis were collected in the vicinity of Syracuse, NY in the fall of 1978. They were stored in plastic bags and refrigerated until needed. The methods by which spores were isolated from the fertile fronds and general cultural methods are described in Miller and Greany (1974). An additional step was taken in which the outer spore coat was removed from the dry spores by brief treatment with sodium hypochlorite, following which the spores were redried and stored (Vogelmann & Miller, unpubl.). Removal of the brown outer spore coat made it easier to observe plasmolysis or cytorrhysis. Dormant spores were hydrated before they were used by floating them on the surface of distilled water for three hours. Young gametophytes were grown by floating spores on the surface of Knop’s solution for three days in an air-conditioned growth chamber where the temperature was 26 + 1°C and the light intensity was about 800 ft-c of continuous cool-white fluorescent illumination. After three days, the gametophytes had 3—5 vegetative cells and one primary rhizoid. The solutes tested included NaCl, ethylene gylcol, glycerol, glucose, sucrose, polyethylene glycol-600 (PEG-600) and PEG-1000 (all chemicals from Fisher Scientific Co.). All compounds were dissolved in distilled water. The concentrations of the solutions, which are given in Table ], were selected in preliminary experi- ments so that responses were observed within 1-5 min. Spores or prothallia were mounted directly in a drop of solution on a microscope slide and were covered with a cover slip. Observations were made with a microscope at a magnification of approximately 300 x. Photographs were made of representative cells. TABLE 1. INDUCTION OF PLASMOLYSIS (P) OR CYTORRHYSIS (C) BY DIFFERENT SOLUTES. Substance and and olecular | Prothallial Concentration diameter (nm)? Spores cells Rhizoids Ethylene glycol (50, 15) 0.45 P P Glycerol (50, 15) 0.55 P P 4 Glucose (50, 20) 0.88 . P Sucrose (50, 20) 1.03 € P i PEG-600 (50, 20) 2.9 c P and C r PEG-1000 (—, 40) 35 — Cc C % concentration (w/v). First concentration in parenthesis is for spores; second concentration 1s for prothallia and rhizoids. *Values for ethylene glycol and glycerol from Goldstein and Solomon , fe ei (1960); for glucose and sucrose rom Durbin (1960); for PEG-600 and PEG-1000 from Carpita, et al. (1979). ' RESULTS AND DISCUSSION The phenomena of plasmolysis and cytorrhysis occurred in Onoclea spores, prothallial cells, and rhizoids (Table 1). Collapse of the spore wall through cytorrhysis was demonstrated in 50% sucrose (Figs. ] and 2 show two views of a spore at different focal levels). Typically the spores became indented and bowl- shaped; the indentation always occurred on the proximal face (Fig. 1). No plas- molytic retraction of the protoplast took place at any point; even at the rim of the J. H. MILLER: DIFFERENCES IN APPARENT WALL PERMEABILITY 121 - 1-10. Examples of plasmolysis and cytorrhysis in spores, rhizoids, and prothallial cells of O. sensbils 50 um scale is the same for all photographs. FIGS. | and 2. Cytorrhysis in a spore shown at different focal levels. FIGS. 3 and 4. Plasmolysis in spores shown at different focal levels. FIG. 5. Control rhizoid. FIG. 6. Plasmolysed rhizoid. FIG. 7. Ribbon-like rhizoid, collapsed as a result of oo FIG. 8. Control prothallus. FIG. 9. Plasmolysed prothallus. FIG. 10. Cytorrhysis in Prothallus 122 AMERICAN FERN JOURNAL: VOLUME 70 (1980) bowl the granular cytoplasm was in contact with the intine (Fig. 2). A 15% NaCl solution clearly induced plasmolysis (Fig. 4). Spore plasmolysis also was accompa- nied by an indentation of the proximal spore face (Fig. 3), since there appeared to be a firm adhesion of the protoplast to the intine in the area of the raphe, and when the protoplast became plasmolysed, the intine was drawn in at that point. A control rhizoid is shown in Fig. 5, and plasmolysis in 20% glucose is illustrated in Fig. 6. Cytorrhysis in a rhizoid in 40% PEG-1000 resulted in the collapse of the cell into a ribbon form (Fig. 7). The vegetative portion of a normal prothallus is pictured in Fig. 8. Plasmolysis gave the appearance shown in Fig. 9, whereas the collapse and crumpling of the cells through cytorrhysis is shown in Fig. 10. The figures give the pictorial definition of the terms plasmolysis and cytorrhysis as they are used in this paper. The main results are summarized in Table /. Prothallial cells showed only plasmolysis with compounds up to the size of sucrose. PEG-600 caused both cytorrhysis and plasmolysis, whereas PEG-1000 produced pure cytorrhysis. Follow- ing the reasoning of Carpita, et al. (1979), the limiting pore size of prothallial cell walls appears to be between 2.9 and 3.5 nm. This is somewhat smaller than the values found by Carpita, et al. (1979) for the cells of several species of angiosperms. Rhizoids appear to have a slightly larger wall pore size than prothallial cells, since PEG-600, which caused both cytorrhysis and plasmolysis in prothallial cells, caused only plasmolysis without cell collapse in rhizoids. The overall permeability of rhizoids was shown by Smith (1972) to be greater than the permeability of prothallial cells of Polypodium vulgare. His measurements were made by following the uptake of a vital dye into the protoplast, and thus reflect the permeability of the plasma membrane. The results of the present study indicate that some of the difference between the permeability of rhizoids and prothallial cells may be caused by differences in the permeabilities of their walls. The spore intine clearly was much more impermeable than the cell walls of the gametophyte. Glycerol was the largest molecule which caused plasmolysis of the spore protoplast; glucose induced pure cytorrhysis. The capillary pore size of the intine appeared to be less than 0.8 nm. Spores were visibly affected only by higher concentrations of each of the substances than were required to produce effects in rhizoids or prothallial cells. This may reflect the fact that the spore cytoplasm is very dense and non-vacuolate, as ‘seen in electron micrographs (Bassel, Kuehnert & Miller, 1981). More of the water of the spore may be bound in the hydration of proteins, for example, and relatively little available for free osmotic exchange. The same concentrations of osmotic solutions which plasmolysed or collapsed prothallial cells within 1-5 min acted much more rapidly on rhizoids. In each test, rhizoids were affected in less than 30 sec, the tume necessary to prepare the sample and make the first observation. This rapidity is probably another reflection of the greater permeability of rhizoids. When young gametophytes were placed in plasmolysing solutions, the basal cell of the plant was affected first, followed by the intermediate and more apical cells. Deplasmolysis occurred in prothallial cells and rhizoids which were plasmolysed in certain of the solutions. Both cell types deplasmolysed completely within 15 min after immersion in ethylene glycol. Prothallial cells were deplasmolysed completely in two hr in glycerol and were partially deplasmolysed in the same time in glucose. J. H. MILLER: DIFFERENCES IN APPARENT WALL PERMEABILITY 123 Rhizoids: showed only a partial recovery in either of these two compounds. Plasmoly- sis appeared to be permanent in solutions of any of the large substances. The instances of deplasmolysis indicate that the plasma membranes of the cells were permeable to the plasmolysing solute, and enough was taken up eventually to reverse the flow and cause an influx of water into the protoplasts. The results which are presented in this paper support the idea that the low permeability of Onoclea spores results from the permeability properties of the intine. The estimated capillary pore size of the intine is only about one quarter that of prothallial cells and rhizoids. The intine should play a major role in determining the entry and exit of materials into and from the spore during the first stages of germination before the intine is ruptured. Some aspects of permeability seem not to be explicable on this basis. Vogelmann (1980), for example, showed that colchi- cine, griseofulvin, and isopropyl N-chlorophenyl carbamate produce striking effects on spore germination, and each appears to enter the spore before the time of intine rupture, although one would expect them to be excluded on the basis of their size. One possible explanation for this type of anomaly is the suggestion by Carpita, et al. (1979) that a small number of larger pores might provide access to the protoplast by larger molecules, whereas osmotic effects may be governed by the more abundant smaller pores. This research was aided by grant PCM-7904593 from the National Science Foundation. LITERATURE CITED BASSEL, A. R., C. C. KUEHNERT and J. H. MILLER. 1981. Nuclear migration and asymmetric cell ees | in eaiaes sensibilis spores: an ultrastructural and cytochemical study. Amer. J. Bot 68:(in press). CARPITA, N.. D. SABULARSE. D. MONTEZINOS and D. P. DELMER. 1979. Determination of the pore size of cell walls of living plant cells. Science 205:1144 —114 DURBIN, R. P. 1960. Osmotic flow of water across permeable cellulose membranes. J. Gen. Physiol. 44:315-326. FISHER, R. W. and J. H. MILLER. 1978. Growth regulation by ethylene in fern gametophytes. V. Ethylene we the early ree of spore germination. Amer. J. Bot. 65:334 —339. GANTT, E. and H. J. ARNOTT. 1965. Spore germination and development of the young gametophyte of the ane fern (Matteuccia struthiopteris). Amer. J. Bot. 52:82-94. GOLDSTEIN, D. A. and A. K. SOLOMON. 1960. Determination of oe pore radius for human red cells by osmotic pressure measurement. J. Gen. Physiol. 44:1 MARENGO, N. P. 1973. The fine structure of the dormant spore of crac struthiopteris. Bull. orrey Bot. Club 100:147-150. MILLER, J H. and R. H. GREANY. 1974. Determination of rhizoid orientation by light and darkness N germinating spores of Onoclea sensibilis. Amer. J. Bot. 61:296—302. RAGHAVAN, V. 1976. Gibberellic acid-induced germination of spores of Anemia phyllitidis: Nucleic acid and protein synthesis during germination. Amer. J. Bot. 63:96 60-972. SMITH, D. L. 1972. Staining and osmotic properties of young gametophytes of Polypodium vulgare L. and their bearing on rhizoid formation. Protoplasma 74:465—497. TOWILL, L. R. and H. IKUMA. 1975. Photocontrol of the germination of Onoclea — II. Analysis of the germination process by means of anaerobiosis. Plant Physiol. 55:150—15 VOGELMANN, T. C. 1980. Nuclear migration, asymmetric cell division and rhizoid aaa: in germinating spores of the sensitive fern, Onoclea sensibilis L. Ph.D. dissertation, Syracuse University. 124 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) Allelopathy and Autotoxicity in Three Eastern North American Ferns WILLIAM E. MUNTHER and DAVID E. FAIRBROTHERS* Undoubtedly gametophytes are the most vulnerable stage in the fern life cycle. Although individual sporophytes produce from 50 to 100 million haploid spores (Shaver, 1954), few of the small, delicate gametophytes, which in most ferns are less than 1.25 cm wide, survive and produce sporophytes. Conditions governing spore germination are critical to gametophyte establishment. Conway (1953) found that, in Scotland, few gametophytes or young sporophytes of Pteridium aquilinum occurred under field conditions, despite heavy spore production by mature plants. This suggests that the gametophyte and young sporophyte stages may be limiting in the establishment of this species. Many studies (cited in Hill, 1971) suggest that fern gametophytes and sporophytes can possess quite different habitat requirements and that one stage of the life cycle can be more sensitive than the other to physical and biological conditions, and so limit the success of the species. Although environmental (abiotic) conditions are extremely important, biotic conditions, particularly those created by the sporophytes or gametophytes, may also be of critical importance in the establishment of new individuals of the same or another species. Therefore, these factors may regulate both population density and community composition. Chemical inhibition of one plant by another, or allelopathy, has been known for over a century (Muller, 1966) and has been much studied in flowering plants and conifers. The phenomenon of antibiosis is also well known to microbiologists. However, relatively little research has been devoted to rite: inhibitors produced by non-seed plants, other than microorganisms (Rice, Bohm and Tryon (1967) reported that many species of ferns produce phenolic compounds. They examined 46 species for the presence of hydroxylated cinnamic and benzoic acids and found a basic complement of cinnamic acids (p-coumaric, caffeic, and ferulic) in the ferns they tested. Also generally present were p-hydroxybenzoic, protocatechuic, and vanillic acids. Sinapic, syringic, and o-coumaric acids were reported to be less common. In a follow-up study, Glass and Bohm (1969) found similar phenolic compounds in 46 additional species. The presence of a basic complement suggests that the well established pathways of phenolic metabolism in the seed plants also function in ferns. Many of the phenolic compounds found in ferns are known to be allelopathic in many species of higher plants, either directly or indirectly, such as after microbial decomposition (Rice, 1974). Most phenolic acids are at least slightly soluble in water. With the increasingly acidic rainfall in the northeastern United States (Li- kens et al., 1970; Bormann & Likens, 1977), weak organic acids such as the phenolics may be leached quite readily either from the leaves during the growing season or from senescent plants. A wide variety of organic and inorganic sub- *Department of Botany, Rutgers University, New Brunswick, NJ 08903. MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 125 stances known to be allelopathic are capable of being leached from the leaves of a number of species of higher plants (Tukey, 1966, 1969; Rice, 1974). Of the few allelopathic studies concerned with ferns, most have involved Pteridium aquilinum. Del Moral and Cates (1971) reported that aqueous leaf ex- tracts and litter extracts of P. aquilinum inhibit seeds of Hordeum vulgare, Bromus tectorum, and Pseudotsuga menziesii. Gliessman and Muller (1972) found that P. aquilinum inhibited germination and subsequent radicle growth of seeds of Bromus rigida and Avena fatua. The toxic principle was detected in the fern leaf leachates. Because the toxic principle was transported by some form of precipita- tion, water-soluble phenolic acids were suspect, and cinnamic acid was tentatively identified. In another study, Stewart (1975) found that water-soluble extracts from Western Bracken (P. aquilinum var. pubescens) delayed germination of Rubus spectabilis seeds and inhibited germination of R. parviflorus seeds, but had no effect on the seeds of Pseudotsuga menziesii. Glass (1976) prepared a solution of phenolic acids having the same composition as was detected in the soil associated with the roots of P. aquilinum and tested its effect on the growth of barley, wheat, oats, rye, rye-grass, barley grass (Hordeum murinum), Clover, and Agropyron repens. Growth was inhibited in all species investigated except A. repens. Whether P. aquilinum sporophytes in some way inhibit the germination of their own spores or those of another species of fern was not investigated in Glass’s or any other study reported in the literature. Davidonis and Ruddat (1973) found that Thelypteris normalis sporophytes in- hibit the growth of T. normalis gametophytes, as well as those of Preris and Phlebodium. The inhibitors, which they termed thelypterins a and b, were similar in many ways to indoleacetic acid (IAA) and were found to be exuded from sporophyte roots. Davidonis and Ruddat (1974) also reported that gametophytes grown in the immediate vicinity of a mature sporophyte of 7. normalis had a reduced number of cells and an altered gross morphology. They observed the greatest inhibition of T. normalis gametophytes when the thelypterins were added before spore germination had occurred. : Davidonis (1976) reported that T. noveboracensis, Pteris multifida , and P. vit- tata also contained thelypterins. Since the thelypterins were detected in leaf diffu- Sates of T. normalis and T. noveboracensis, she postulated that foliar leaching may be one mechanism by which these inhibitors are released into the environ- ment Petersen (1976) found that Dryopteris intermedia and Osmunda cinnamomea gametophytes inhibited each other’s development when grown together in culture. € also reported that gametophytes of these two species “lock” each other into Perpetual juvenility. Petersen was unable to isolate and indentify the inhibitory Substances since they occurred in very small amounts or were highly volatile. Horsley (1977) reported that the presence of Dennstaedtia punctilobula and T. noveboracensis sporophytes were correlated with reduced numbers of Prunus Serotina, Acer rubrum, and A. saccharum seedlings under field conditions. — Very limited research has been conducted either to detect possible allelopathic interactions between species of ferns or to demonstrate the inhibition of one 126 AMERICAN FERN JOURNAL: VOLUME 70 (1980) generation by the other within a species (autotoxicity). It is not known whether allelopathy and autotoxicity are widespread in natural fern communities. Davidonis and Ruddat (1973, 1974) and Davidonis (1976) examined a small number of species for inhibitors, but in many instances the species tested for allelopathic interactions were those that are maintained in greenhouses, and not those which occur together in natural situations. Petersen (1976) has shown that fern gametophytes of one species produce substances that inhibit gametophytes of others, the classic allelopathic inhibition of one species by another. Under field conditions, however, the dominant and normally perennial sporophytes probably produce much greater quantities of allelopathic compounds capable of affecting spore germination or gametophyte growth than would the much_ smaller gametophytes. Petersen’s research suggests that sporophytes might also inhibit gametophytes, as they were found to have produced the same flavonoids, but no experimental evidence to support such a hypothesis was presented. There are reports suggesting allelopathy between fern generations in both natural populations and the greenhouse. In greenhouses, potted ferns often pro- duce large numbers of spores, but gametophytes are seldom found growing on the soil surface directly underneath the sporophytes. It is also unusual to find gametophytes growing underneath or even near mature sporophytes under field conditions, even when spore production is heavy. In this research, aqueous leaf extracts, leaf leachates, and leaf litter infusions from Osmunda cinnamomea, O. claytoniana, and Dennstaedtia punctilobula were examined for their allelopathic and autotoxic potential. Leaves of all three species were also tested for the production of volatile inhibitors. All three species occupy much the same habitat, although O. cinnamomea is often found in moister areas than the other two. The geographical ranges of all three species also are approximately the same, and all are native and common in the northeastern United States. Although these three species occupy almost iden- tical habitats, they seldom occur in close proximity to one another within a given area. MATERIALS AND METHODS Mature spores of O. cinnamomea, O. claytoniana, and D. punctilobula were collected in 1976 and 1977 from wild populations which were located in Sandy- stone Township, Sussex County, New Jersey, and in the Town of Goshen, Addi- son County, Vermont. Spore-bearing leaves of both species of Osmunda were collected during the first two weeks of May from New Jersey populations and during the last week of May and first week of June from Vermont populations. Sporangia were allowed to dehisce at/room temperature. Spores were refrigerated at 4.4” C as soon as possi- ble after they were shed from the sporangia. Although Osmunda spores contain chlorophyll and remain viable for only a few days at room temperature (Cobb, 1963), they will retain their viability for over a year if refrigerated (Stokey, 1951)- Spore-bearing leaves of D. punctilobula were collected during the last two weeks of July from the New Jersey populations and during the first week of MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 127 August from Vermont populations. After dehiscence of the sporangia, spores were stored dry and at room temperature. The spores of D. punctilobula lack chlorophyll and have a thicker, more resistant spore wall than Osmunda spores, and so require no refrigeration to remain viable. Fresh fronds from all three species were used to prepare aqueous extracts and leachates. They were collected from both Vermont and New Jersey populations throughout the growing season. Fronds were refrigerated immediately after col- lection and were then stored at 4.4° C. The extracts and leachates were prepared from the fronds within three or four days after collection. All aqueous leaf extracts were prepared by first cutting into pieces 5 g of leaves (fresh weight) and then placing the pieces in 80 ml of distilled water. The leaves were boiled briefly (3 min) to stop enzyme activity and, it was hoped, to destroy some of the microorganisms which might later contaminate cultures. After boiling, the leaves were ground in a Virtis blender for 5 min. The ground mixture was then vacuum filtered and the resulting aqueous extract was brought to volume with distilled water (100 ml distilled water for each 5 g of ground leaves). Extracts were stored at 4.4° C until they were used to prepare cultures, usually within 3-5 days. Leaf leachates for each species were prepared by placing fronds (2 thick) on 1.5 mm mesh screen which was placed on top of a rectangular plastic box. Fronds having an area of ca. 630 cm? were then misted with 300 ml of distilled water. The droplets of water falling from the leaves were collected in the large plastic box. Leachates were stored at 4.4° C until used. Litter infusions were prepared by placing 10 g of chopped dry frond litter in a 500 ml beaker. Weekly additions of 100 ml of distilled water were made for a period of 3 weeks. Beakers were placed in direct light in the greenhouse. Filter paper was placed on top of each beaker to prevent external contamination. After 3 weeks, the aqueous portion of the infusion was decanted and vacuum filtered. The filtered liquid was used to prepare the experimental cultures. The methods used to prepare cultures were based on those reported by Munther (1975). All cultures using extracts, leachates, and infusions were prepared as follows: plastic seed germinators (clear plastic boxes measuring I1 x ikx 3 cm) were filled with 0.138 1 (dry) of autoclaved horticultural grade (medium) vermicu- lite, and various treatment solutions were added to the vermiculite. The treatment solutions contained two parts of extract, leachate, or infusion and one part of 2X Hoagland’s no. 1 solution plus trace elements (Hoshizaki, 1975). Fern spores generally require only moisture to germinate (Miller, 1968), but a nutrient solution was added because the cultures were examined over an extended time, and nutri- ents are necessary after the first few cell divisions of germinating spores. A total of 75 ml of solution was added to each experimental box. In controls, distilled water was mixed with Hoagland’s solution instead of ‘the extracts, leachates, or infu- sions. Fern spores were sown on the upper surface of pieces of unglazed clay pots placed on top of the vermiculite, 3 per germinator. New pots were used, and all were from the same manufacturer’s lot. The pots, broken into pieces of approxi- mately equal size, were boiled vigorously in water for 15 min three separate times, using fresh water for each boiling to remove the impurities. 128 AMERICAN FERN JOURNAL: VOLUME 70 1980) Cultures were placed under 285-315 ft-c of illumination provided by alternately placed 40-watt plant grow (G.E.) and cool white fluorescent tubes with a photo- period of 14 hrs. The light intensity used was based on field observations obtained with a light meter and on a recommendation reported by Miller and Miller (1961). Temperature was maintained at 22-24° C. Each culture was checked daily after sowing until the first spores began to germinate. At that time, a count was made to determine percent germination and then repeated every other day for thirteen days. Since experimental data indicated that the last two counts were not appreciably different, the time was later reduced to 11 days. The criterion used to determine when spore germination had occurred was the appearance of the rhizoid following the first division of the spore, not just the uptake of water as shown by swelling. All counts were made under low power (100) of a light microscope, and a mechanical counter was used to record the number of germinated and non-germinated spores within a randomly selected microscopic field. One count was taken from each chip in the box. All experi- ments were conducted in triplicate, using three germinator boxes for each repli- cate. To test for the production of volatile compounds by the fronds of each of the three species, 30 g of fresh fronds were placed in a 16 x 31 cm plastic box. Also within this plastic box were 9-cm petri dish bottoms containing culture medium similar to that used in the other experiments. Each plate contained 0.091 | (dry) of vermiculite plus 50 ml of control solution on top of which were the pieces of clay pot with spores. The box was sealed with clear plastic tape. For controls, 30 g of cotton moistened with water was used in place of the fresh fronds. This method was similar in principle to that described by Muller (1966), and allowed any vol- atile compounds produced by the fronds to concentrate in the closed atmosphere of the box. The only contact between the fronds and the spores was through the air. The sealed boxes were placed under 300 ft-c of light for 14 days, after which counts were taken to calculate percent germination. All experiments were con- ducted in tripicate. The replicated means of percent germination resulting from treatment with leaf leachates and extracts were compared statistically using Duncan’s multiple range test (Duncan, 1955; Steele & Torrie, 1960), with replicated control means, for every possible autotoxic and allelopathic interaction (Munther, 1978). Separate comparisons to controls were made using the litter infusion treatment means. Only the means from the first (day 1) and last (day 11 or 13) counts were used for statistical analysis. These statistical comparisons were therefore based on a minimum (the presence of any germinated spores in any replicate of a species and treatment) and a maximum (all spores capable of germination from a variable population) level of germination for every possible interaction. Replicated volatile treatment means were compared to the controls in a similar manner; however, only one series of means representing the maximum level of germination was used in these comparisons, since it was not possible to obtain a minimum level due to the design of the experiments. MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 129 RESULTS AND DISCUSSION In the following discussion, compounds produced in the leaves of a species which inhibit the germination of spores of the same species will be termed ‘‘au- totoxic.’’ The inhibitory effects on spores of another species will be termed “‘al- lelopathic.”’ Spores of both Osmunda species began to germinate in five or six days in Vermont and New Jersey populations. Dennstaedtia spores generally took longer to begin germination. Those from Vermont required at least 11 days, and those from New Jersey required at least 8 days. TABLE 1. EFFECTS OF AQUEOUS LEAF EXTRACTS AND LEACHATES ON SPORE GER- MINATION AND EARLY GAMETOPHYTE GROWTH IN DENNSTAEDTIA AND OSMUNDA. Species Vermont New Jersey Spores Leaves Leachate Extract Leachate Extract FC EC FC | FC Gy FC i & Autotoxic Interactions O. cinn. O. cinn a - ~ - - O. clay O. clay. _ “ a + ~ - - ~ D. punc D. punc. 2 = * aa = + 4 Allelopathic Interactions O. cinn O. clay. S - + os = as + O. cinn D. punc - ~ + + * - + sl O. clay. O. cinn. & = bid - ~~ re _ - O. clay D. punc. ~_ - - — ces D. pune O. cinn. _ = + ~ - - - - D. punc O. clay. - — + + = = = = FC = first count; LC = last count a ies + = Statistically significant inhibition (0.05 level); — = no significant inhibition In the Vermont populations, all species exhibited some degree of autotoxicity (Table 1). In each species, the leaf extract was found to inhibit germination signifi- cantly, particularly in the last count. InO. cinnamomea, leaf leachates also signif- icantly inhibited spore germination, and the inhibition was nearly as great as that caused by the extracts (Table 3). This differed considerably from the results obtained from the New Jersey populations, where only one species, D punctilobula, was found to be significantly autotoxic (Table 1). In this case, the inhibition was caused by the extract. Significant allelopathic interactions also were found in the Vermont popula- tions. Leaf extracts of bothO. claytoniana and D. punctilobula inhibited spores of O. cinnamomea, significantly lowering percent germination (Table /). Inhibition of germination was quite severe, particularly as revealed in the final counts (Ta- bles 4 and 5). Spores of D. punctilobula were inhibited by leaf extracts of O. cinnamomea and O. claytoniana (Table 1). Inhibition in this case also was quite marked, especially by extracts of O. claytoniana (Table 4). Aqueous leaf extracts of O. claytoniana often severely inhibited spores of the other two species, but €xtracts of both O. cinnamomea and D. punctilobula had little effect on O. claytoniana spore germination (Table 1). Leaf leachates of the three species tested Produced no significant allelopathic inhibition of spore germination. 130 AMERICAN FERN JOURNAL: VOLUME 70 (1980) The number of significant allelopathic interactions was lower in the New Jersey populations than in the Vermont populations (Table 1). Spores of O. cinnamomea were inhibited by leaf leachates of D. punctilobula, which was the only significant allelopathic inhibition by leaf leachates in either the New Jersey or Vermont populations (Tables ] and 5). The inhibition produced by the leachates was not so severe as that produced by the extracts (Table 5). Also in New Jersey, germina- tion of O. cinnamomea spores was inhibited by leaf extracts of both O. claytoniana and D. punctilobula able 1). Osmunda cinnamomea and O. claytoniana leaf extracts had little effect on germination of D. punctilobula spores (Table 1). This result is quite different from that obtained for the Vermont popula- tions (Table 1). However, as in the Vermont populations, spores of O. claytoniana were unaffected by leaf extracts (or leachates) of both O. cinnamomea and D. punctilobula (Table 1). Nevertheless, in the majority of cases, the Vermont and New Jersey populations reacted quite differently in terms of the number and type of allelopathic and autotoxic interactions. TABLE 2. EFFECT OF LEAF LITTER INFUSIONS ON SPORE GERMINATION AND EARLY GAMETOPHYTE GROWTH IN NEW JERSEY POPULATIONS OF DENNSTAEDTIA AND OSMUNDA. Species litter infusion Spores Leaves FC LC Autotoxic Interactions cinn. O. cinn. - = O. clay. O. clay. - = D. punc D. punc. = = Allelopathic Interactions O. cinn O. clay. = = O. cinn D. punc Sa x O. clay O. cinn = O. clay. D. punc = oa D. punc O. cinn. “ — D. punc O. clay. _ - FC = first count; LC = last count + = statistically significant inhibition (0.05 level); — = no significant effect Leaf litter infusions of New Jersey material did not inhibit germination greatly (Table 2). Dennstaedtia punctilobulaspores were inhibited by O. cinnamomea leaf litter infusions, the only allelopathic inhibition caused by treatment with litter infusions. The only species exhibiting autotoxicity was O. cinnamomea (Table 2), in which the litter infusion merely delayed germination, since there was no inhibi- tion in the last count. Two types of inhibition were observed in spores treated with extracts and leachates. Spores of all three species imbibed water readily when treated with leachates or extracts, but spores treated with extracts often did not divide. This indicates that the inhibitor was able to enter the spore through the wall and to prevent the first division. This type of inhibition causes a low percentage of spores to germinate, is evident in the first count when compared to the controls, and MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 131 continues through the last count (Table 1). This was the most common type of inhibition. In the second type, several divisions occurred and percent germination was not affected in the first count (see O. claytoniana, autotoxicity, Table 1). The inhibitor in this instance did not affect germination, but acted on the several-celled stage (young gametophyte). In the later counts, if a spore had germinated but had died (chlorophyll lost) at the several-celled stage, it was counted as not germi- nated. Therefore, the last count measures the inhibitor’s effect on early gametophyte growth and development. The presence of this type of inhibition can be recognized only at the last count. When leaf extracts caused significant inhibition of germination, gametophyte development from spores which germinated usually was affected, with the gametophyte usually arrested in the filamentous stage and little or no further growth occurring. Leaf leachates usually did not cause this type of response. Gametophytes resulting from spores which did germinate underwent normal de- velopment, but often exhibited somewhat slower growth than did the controls. TABLE 3. AUTOTOXIC EFFECTS OF LEAF LEACHATES AND EXTRACTS EXPRESSED AS A PERCENTAGE OF SPORE GERMINATION AND EARLY GAMETOPHYTE GROWTH IN VERMONT POPULATIONS OF O. CINNAMOMEA. STATISTICAL ANALYSIS USING DUN- CAN’S MULTIPLE RANGE TEST WITH A 0.05 SIGNIFICANCE LEVEL. (ALL EXTRACT AND LEACHATE MEANS SIGNIFICANTLY DIFFERENT FROM ALL CONTROL MEANS.) First count (sx = 5.421) is E2 Means El LI E3 ee eT eee ee s= ‘a Oe Be SE 8 X= ny 6 UMA OOS ee ee ee Last count (sx = 5.034) Means B3 El E2 L3 a S66 6S s= 7.1 ‘i a. ae 3 OU CU x= ot OW Wh - Us. MA Oe ee ee LM. SX = standard error (of the mean); s = standard deviation; X = mean percent germination C1-3 = control means; L1-3 = leachate treatment means; E1-3 = extract treatment means Phenolic compounds are known to inhibit ion uptake in flowering plants through reversible alterations in membrane permeability (Glass, 1973, 1974) and may af- fect IAA metabolism. The spore wall may prevent the phenolic inhibitor from entering the spore, with the first divisions of the spore occurring on stored re- serves. By the time the gametophyte reaches the several-celled stage, the phenolic inhibitor may prevent ion uptake sufficiently to inhibit growth or further cell division. The effects of the extracts on spore germination and on early gametophyte Stages were much greater than the effects of the leachates in almost every experi- ment. While experimental effects of the leachates may be easily extrapolated to field conditions, the extracts present an enigma. While nothing directly paralleling an extraction procedure exists in nature, experimental results obtained using ex- 132 AMERICAN FERN JOURNAL: VOLUME 70 (1980) tracts do possess some validity in determining the phytotoxic potential of plants under laboratory conditions. Extracts may concentrate compounds which are leachable and concentrated in the soil under field conditions. However, some compounds present in an extract may not be leached from healthy, growing tissue, and many substances can be altered by the extraction procedure itself. On the other hand, substances that are not normally leached from such tissue can be released by senescent tissue quite readily (Gliessman & Muller, 1972; Stewart, 1975). Various forms of stress, such as temperature extremes, drought, attack by pathogens, or mechanical injury also increase the leachability of metabolites from foliage (Tukey, 1966, 1969; Rice, 1974). Soils also vary in their composition, structure, moisture content, pH, and the kinds of microorganisms present. Thus, they would have different affinities for various inhibitors, or may even render them inactive or increase their activity due to microbial decomposition (del Moral & Cates, 1971; Rice, 1967, 1969; Wang et al., 1967). TABLE 4. ALLELOPATHIC EFFECTS OF O. CLAYTONIANA LEAF EXTRACTS AND LEACHATES EXPRESSED AS A PERCENTAGE OF SPORE GERMINATION AND EARLY GAMETOPHYTE GROWTH IN VERMONT POPULATIONS OF O. CINNAMOMEA AND D. PUNCTILOBULA. STATISTICAL ANALYSIS USING DUNCAN’S MULTIPLE RANGE TEST WITH A 0.05 SIGNIFICANCE LEVEL. (ALL EXTRACT (NOT LEACHATE) MEANS SIGNIFI- CANTLY DIFFERENT FROM ALL CONTROL MEANS.) . cinnamomea Last count (sx = 8.450) Means E2 E3 El L2 Li Cl C3 C2 s= 14.1 18.0 27.2 19.0 8.7 3.5 8.1 10.2 7.0 <= 12.2 24 39.4 43.6 54.8 58.4 76.8 80.5 81.3 D. punctilobula First count (sx = 4.618) Means El E3 E2 1 C3 L3 C2 S= 3.8 5.8 3.3 13.7 3.8 9.1 14.5 4.5 12 X= 15.1 18.5 28.0 36.0 41.9 47.5 49.7 S27 54.5 Last count (sx = 4.965) Means El E3 E2 LI L2 L3 C3 Cl C2 s= =a 12.5 15.6 2.3 10.9 3 6.4 2.9 6.2 x= 11.0 26.0 32.8 64.0 65.1 67.3 74.9 78:7 76.2 sx = standard error (of the mean); s = standard deviation; X inati cas ' 3 eviation; X = mean percent germination C1-3 = control means; L1-3 = leachate treatment means; E1-3 = extract treatment means In all samples except one of D. punctilobula from New Jersey, the pH of the extracts was lower than that of the leachates (Table 5 ). The pH of the extracts and leachates from New Jersey populations did not differ greatly from that obtained from the Vermont material (Table 6). Although in all samples the pH of the leachates from Vermont populations was higher than in leachates from New Jer- sey, the difference probably is not enough to be significant. It can be deduced from Table 6 that allelopathy is not strictly a low pH phenomenon. For example, MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 133 the leaf leachates and extracts from O. cinnamomea (New Jersey) exhibited very little phytotoxicity, yet they possessed the lowest pH. As already mentioned, litter infusions of O. cinnamomea were found to be autotoxic (Table 2) only in the first count, indicating delayed germination. Such a delay could be important under field conditions, however, increasing the chances of attack by microbial pathogens which could cause damping off of the young prothalli, as is often the case with seeds. Within the species studied in New Jersey, the pH values obtained for leaf-litter infusions were higher than expected when compared with those obtained for leaf leachates and extracts (Table 6). TABLE 5. ALLELOPATHIC EFFECTS OF D. PUNCTILOBULA LEAF EXTRACTS AND LEACHATES EXPRESSED AS A PERCENTAGE OF SPORE GERMINATION AND EARLY GAMETOPHYTE GROWTH IN VERMONT AND NEW JERSEY POPULATIONS OF 0. CIN- NAMOMEA. STATISTICAL ANALYSIS USING DUNCAN’S MULTIPLE RANGE TEST WITH A 0.05 SIGNIFICANCE LEVEL. (ALL EXTRACT (NOT LEACHATE) MEANS SIGNIFI- CANTLY DIFFERENT FROM ALL CONTROL MEANS IN VERMONT; ALL MEANS SIGNIF- ICANTLY DIFFERENT FROM ALL CONTROL MEANS IN NEW JERSEY.) Vermont Last count (sx = 5.488) LI 2 Means E2 E3 El LS Cl C3 C2 S = 27 3.5 9.5 15.8 14.7 74 8.1 10.2 7.0 X= 27.8 33.2 38.8 oe RE OOS ES. ES New Jersey First count (sx = 4.569) Means El E3 E2 L2 L3 Ll Cl C2 C3 a= 10.4 8.1 10.6 7.4 4.2 8.8 7.9 75 3.0 x= 29.9 36.7 37.1 ay. oe OF er Last count (sx = 4.608) Means El E2 E3 L3 Ll L2 C2 C3 Cl = 12.2 12.3 5.3 3.9 6.1 4.4 8.3 6.1 8.2 X = MS Ob eo aS se ee ee mean per cent germination SX = standard err -s= dard deviation; X = rt Tere 8 oe 3 = extract treatment means. C1-3 = control means; L1-3 = leachate treatment means; El- llowing the first he time of the parent Frond litter was collected while the fronds were still standing, fo killing frost in 1976. It is unknown whether or not it rained between t first frost and the time of collection. If it did, this may account for the ap lack of phytotoxicity in the infusions, since any water soluble phytotoxins could have been removed by the rainfall. The importance of the first rainfall after frond senescence already has been cited by Gliessman and Muller (1972), who studied the effects of P. aquilinum leaf litter extracts on Avena fatua and Bromus rigida radicle growth, and by Stewart (1975), who examined the effects of P. aquilinum litter extracts on seeds of Rubus sp. and Douglas-fir. 134 AMERICAN FERN JOURNAL: VOLUME 70 (1980) No inhibitory volatile compounds were detected in biologically significant levels from leaves removed from New Jersey or Vermont populations. Denn- staedtia punctilobula leaves are quite fragrant, particularly when crushed, but the substance producing this fragrance, presumably coumarin, had no apparent effect on spore germination. Water-soluble inhibitors predominate in more humid or wet environments, ac- cording to the hypothesis developed by Whittaker (1970) relating toxin production to climate. Since the Vermont populations exist in a wetter environment than do the New Jersey ones (Climatological Data, U. S. Dept. of Commerce, NOAA, 1976-1977), this may partially explain the increased number of allelopathic and autotoxic interactions caused by water-soluble leachates and extracts. Volatile inhibitors, according to Whittaker (1970) and Muller (1970), would be most com- mon in a hot, arid environment. The fact that none were found in either Vermont or New Jersey populations would also lend support to their theory relating toxin production to climate, as both areas are relatively moist. Summer temperatures in the study areas in Vermont and New Jersey were nearly identical (Climatological Data, NOAA, 1976-1977). TABLE 6. pH VALUES OBTAINED FROM AQUEOUS LEAF EXTRACTS AND LEACHATES FOR NEW JERSEY AND VERMONT POPULATIONS, AND LITTER INFUSIONS FOR NEW JERSEY POPULATIONS. pH Species NJ VT Osmunda cinnamomea leaf extract 5.40 ia leaf leachate 6.00 6.60 litter infusion 6.50 — soil (5/77) aoa. & os Osmunda claytoniana leaf extract 5.60 5.80 leaf leachate 6.20 6.65 litter infusion 6.75 —_ Soi 6.40 — Dennstaedtia punctilobula leaf extract 5.70 5.80 leaf leachate 5.50 6.00 litter infusion 5.95 — soil (5/77) 5.90 — We express our appreciation to R. W. Willemsen for his help during a portion of this research, and to B. F. Palser and J. A. Quinn for their helpful suggestions and comments. LITERATURE CITED BOHM, B. A, and R. M. Tryon. 1967. Phenolic compounds in ferns. I. A survey of some ferns for cinnamic acid and benzoic acid derivatives. Can. J. Bot. 45:585—593. : BORMANN, F. H., and G. E. LIKENS. 1977. The fresh air-clean water exchange. Nat. Hist. 86:63-71. COBB, B. 1963. A Field Guide to the Ferns. Houghton Mifflin, Bost CONWAY, E. 1953. Spore and sporeling survival in bracken (Peeriaium aquilinum). J. Ecol. 4: 289- 293. MUNTHER & FAIRBROTHERS: ALLELOPATHY AND AUTOTOXICITY 135 _ e H. 1976. The occurrence of thelypterin in ferns. Amer. Fern J. 66:107-108. —_—__——, . RUDDAT. 1973. Allelopathic compounds, thelypterin a and b, in the fern Thely- aa aa Planta 111:23-32. , and M. RUDDAT. 1974. Growth inhibition in gametophytes and oat coleoptiles by thely- erin a and b released from roots of the fern Thely pteris normalis. Amer. J. Bot. 61:925-930. del MORAL, R., and R. G. CATES. ith pinged srr potential of the dominant vegetation of western Washington. Ecology 52:1030— DUNCAN, D. B. 1955. Multiple range he mttins f tests. Biometrics 11:1 GLASS, A. D. M. 1973. Influence 8 phenolic acids on ion uptake. I. ak of phosphate uptake. Plant Physiol. 51:1037-1 . 1974. Influence of ats acids on ion uptake. III. Inhibition of potassium absorption. J. nae Bot. 25:1104-1113. = 1976. The seca potential of saga acids associated with the rhizosphere of Pteridium aquilinum. Can. J. Bot. 54:2440- , and B. M. 1969. A further alia “ ferns for cinnamic and benzoic acids. Phistichemistry 8:629-632. GLIESSMAN, S. R., and C. H. MULLER. 1972. The phytotoxic potential of bracken, Pteridium aquilinum (L.) Kuhn. Madrono 21:299-304. HILL, R. H. 1971. Comparative habitat —— for spore alg and prothallial growth of three ferns in southeastern Michigan. Amer. Fern J. 61:171-182. HORSLEY, S. B. 1977. Allelopathic inhibition of black ese II. Inhibition by woodland grass, erns, and club moss. Can. J. For. Res. 7:415-419. HOSHIZAKL B. J. 1975. Fern Growers Manual. Knopf, New York. LIKENS, G. E., F. H. BORMANN, N. M. JOHNSON, D. W. FISHER, and R. S. PIERCE. 1970. saree of forest cutting and herbicide treatment on nutrient budgets in the Hubbard Brook atershed-ecosystem. Ecol. Monogr. 40:23-4 MILLER. J. H. 1968. Fern gametophytes as esoceieetel material. Bot. Rev. 34:361-4 ———, and P. M. MILLER. 1961. The effects of different light conditions and sucrose on the growth and development of the gametophyte of the fern Onoclea sensibilis. Amer. J. Bot. 8:154-159. MULLER, C. H. 1966. The role of chemical inhibition (allelopathy) in vegetational composition. Bull. Torrey Bot. Club 93:332-351. — Phytotoxins as plant habitat variables. Jn C. Steelink and V. C. Runeckles, eds. Reo nt Advances in Phytochemistry, vol. 3. Appleton-Century-Crofts, New York. MUNTHER. Wy, E. 1975. Investigations into the effects of the photoperiod on spore germination and gametangial development in ferns. Honors thesis, Drew University, Madison, NJ. | . 1978. Allelopathy and autotoxicity in three species of ferns. M.S. thesis, Rutgers University, New Brunswick, NJ. ; PETERSEN, R. L. 1976. Chemical research in the genus Dryopteris Adanson; oh Oa mor- phogenesis, and allelopathy. Ph.D. thesis, Rutgers University, New Brunswick, N RICE, E. L. 1967. Chemical warfare between plants. _ 38:67-74. 1974. Allelopathy. Academic Press, New SHAVER, J. M. 1954. Ferns of the Eastern Central i Dover, New Yor STEELE, R. G. D. , and J. H. TORRIE. 1960. Principles and Procedures of Statiatics. McGraw-Hill, ew York. STEWART, R. E. 1975. Allelopathic potential of western bracken. J. Chem. Ecol. 1: 161-169. STOCKEY, A. G. 1951. Duration of viability of spores of the Osmundaceae. Amer. Fern J. 41:111- 115. Nae TUKEY, H. B., Jr. 1966. Leaching of metabolites from above-ground plant parts and its implications. Bull. Torrey Bot. Club 93:385—401. 16. oe Implications of allelopathy in agricultural plant science. Bot. Rev. 35:1— WANG, T. . C., T-K. YANG, and T-T. CHUANG. 1967. Soil phenolic acids as plant growth aie Soil Sci. 103:239-246. nd J. B. WHITTAKER, R. H. 1970. The biochemical ecology of higher or a © Simeone, eds. Chemical Ecology. Academic Press, New 136 AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) SHORTER NOTES SANDSTONE ROCK CREVICES, AN EXCEPTIONAL NEW HABITAT FOR THELYPTERIS SIMULATA.—The Massachusetts or Bog Fern, Thelypteris simulata (Davenp.) Niewl., is a common wetland fern in the New England states. Most manuals give its typical habitat as very acid, shaded bogs and swamps, frequently in association with sphagnum moss. In some New England cranberry bogs, the fern is so abundant as to be a weed. Hartley (Rhodora 67:399— 404. 1965) reported for the first time 7. simulata from Wisconsin, a disjunction of approximate- ly 600 miles from the nearest station in West Virginia. Since the original discovery, twelve other stations have been found for 7. simulata in west-central Wisconsin. The Wisconsin habitats are all flat, low-lying woods with a moist layer of peat about one foot thick overlying sand. The dominant trees usually are Pinus strobus and Acer rubrum, and the most common understory species Alnus rugosa and Ilex verticillata. The most abundant herbaceous plant at most sites is Osmunda cinnamomea. Other characteristic ground layer associates include Carex brunnescens, C. folliculata, C. trisperma, Dryopteris intermedia, D. spinulosa, D. X triploidea, Maianthemum canadense, Mitchella repens, Osmunda regalis, Rubus hispidus, Sphagnum spp., and Viola incognita. It seems likely that in presettlement times, before the natural character of central Wisconsin was vastly changed by drainage ditches, peat fires, farming, etc., 7. simulata was a common member of the low, wet-acid woods flora that once covered much of the sandy plain of extinct glacial Lake Wisconsin. In mid-August 1979, I found a single plant of 7. simulata growing one foot above the ground in the crevice of a sandstone cliff at Castle Mound, Castle Mound State Park, Jackson County, Wisconsin (T2IN, R4W, SWYsSWYs sec. 23). To my knowledge, there are no reports in the literature of 7. simulata occurring on rock cliffs. The plant consisted of only eight sterile fronds, one of which was taken as a voucher (Moran 995, MIL). Two more fronds were taken on 13 Aug 1980 and checked by Dr. W. H. Wagner, Jr. for correct identity (Moran 1267, MICH). One of the fronds was soriferous, although all the sporangia were still white and immature; maturation was late compared with several other Wisconsin populations that were already actively sporulating. Castle Mound is composed of Cambrian sandstone that rises 180 feet above the surrounding lake plain of extinct glacial Lake Wisconsin. It is typical of many weathered, castellated mounds found in the central Wisconsin sand plains. The T. simulata cliff habitat on Castle Mound faces north and is shaded by Pinus resinosa, Fr strobus, and Quercus rubra. The soil reaction in the crevice where 7. simulata was growing was pH 5.0. No other vascular plants were growing in the crevice with T. simulata. The cliff face immediately surrounding the plant was barren, except for one plant of Dryopteris spinulosa growing in a crevice about one foot above Z simulata. About ten feet above the plant was a ledge with numerous individuals of Polypodium virginianum, along with Aquilegia canadensis, Athyrium filix-femina, and Betula papyrifera. : AMERICAN FERN JOURNAL: VOLUME 70 NUMBER 4 (1980) 137 Although it seems odd that 7. simulata should occupy a dry sandstone cliff when its typical habitat is wet, acid bogs, usually in association with sphagnum moss, the two habitats are similar in certain respects. The wet, cold acid conditions of a bog make it physiologically difficult for roots to absorb water and mineral nutrients. Such a state of “physiological drought” simulates the dry, nutrient-poor crevices of a sandstone cliff. This type of habitat switch is known to experienced field botanists from other examples of swamp or bog plants growing on rocks, and vice-versa. A few such examples are: Cystopteris bulbifera, Dryopteris marginalis, D. spinulosa, Ledum groenlandicum, Lorinseria areolata, Matteuccia struthiopteris, Osmunda cinnamomea, Phegopteris connectilis, Sphagnum spp., and Thelypteris palustris. Although T. simulata may have been common in the low wooded acid swamps that surrounded Castle Mound in presettlement times, the nearest presently known locality is two miles away. The Castle Mound individual certainly is the result of relatively wide-range spore dispersal. It is important to point out that, in view of Klewkowski’s (Science 153:305—307. 1966) ideas on the adaptive value of polyploidy in homosporous pteridophytes, the Castle Mound individual is most probably the result of single spore establishment and intragametophytic selfing.—Robbin C. Moran, Wisconsin Scientific Areas Preservation Council, Department of Natural Resources, P.O. Box 7921, Madison, WI 53707. A SECOND ALABAMA LOCALITY FOR THE HART’S TONGUE.—The discovery of the Hart’s-tongue, Phyllitis scolopendrium (L.) Newm., in a sinkhole in Jackson County, Alabama (Amer. Fern J. 69:47-48. 1979) generated interest in further searches for this fern among members of the Huntsville Grotto of the National Speleological Society who had participated in the find. According to Eric Bachel- der, my guide to the Jackson County locality, these spelunkers found a second, larger population in a sinkhole in Morgan County soon after the original discovery (Huntsville Grotto Newsletter 20:49-50. 1979). On 31 May 1980, I visited the new locality with Mr. Bachelder again as my guide. The population is in a deep sinkhole in the area known as Newsome Sinks, a large sink-valley in northeastern Morgan County about 25 miles southwest of the Jackson County locality and 65 miles southwest of the one in Marion County, Tennessee. The sinkhole is about 70 feet deep and has sheer walls. A small stream falls into the sink, making the air very misty and humid, unlike the dry Jackson County sink. Also unlike the Jackson County sinkhole, it is necessary to rappel down to a wide ledge about half-way own, where most of the Phyllitis plants are. Fifty-three plants occur on the ledge, along with luxuriant C 'ystopteris bulbifera and Wood Nettle (Laportea canadensis), which may have obscured more Phyllitis plants. At least 20 Hart’s-tongues were mature adults; the juveniles ranged from sporelings to almost adults. The ledge is Partially overhung by the cliffs above, and the Phyllitis plants grow in a narrow strip beneath the overhang. The left end of the strip contains mostly adults and large Juveniles; the plants toward the right are gradually reduced in size, age, and density. Apparently the population is spreading towards the right. Four fairly large Hart’s- tongues also were seen at the bottom of the sinkhole. A number of fronds were collected as a voucher (Short 1/95, AUA and duplicates to be distributed).—John W. Short, 905 McKinley Ave., Auburn, AL 36830. 138 AMERICAN FERN JOURNAL: VOLUME 70 (1980) AMERICAN FERN JOURNAL Manuscripts submitted to the JOURNAL are reviewed for scientific content by one or more of the editors, and, ften, b During the past year we have received the kind assistance of J. Beitel, L. G. Hickok, I. Cousens, A. M. Evans, y one or more outside reviewers as well. B. M. Boom, A. C. Jermy, J. H. Miller, J. D. Montgomery, D. H. Nicolson, H. E. Robinson, J. Skog, A. R. Smith, A. Star, R. G. Stolze, W. C. Taylor, R. A. White, D. P. Whittier, and J. J. Wurdack, to whom we are deeply indebted. We welcome suggestions of other reviewers.—D. INDEX TO VOLUME 70 Acrostichum, 51 Se seat tad 97: agen 65: ep apni 102. u istii Pe pone: um cladotrichum, 97; conforme. 55: crinitum, 63 denaciioltarn, 99, 108, 109: decoratum, 60; diversifolium. 97; ellipsoideum 97: engleri, 97: eximium ei, 60: fulvum, 97; gossyp 97. guamanianum, 97; hieronymi. 97: hickenii, 97; hirtum, 66 huacsaro. 62; hybridum, 63: latifolium. 55: litanum, 97 longissimum. 97: molle, 97; muriculatum, 97; muscosum. 62: eandro 7. ovat: paleaceum, 60, 61: pangoanum, : pellucidum. 97; petiolosum. 62: sg ge 97: pilosel- loides, 65: pilosum, 61: pruinos 97; pte cio 97. i Hloen m: ii, 61: maxonii, 55; melanopus, 63: meride ubsect. Microlepidea 62: moorei, 59: moritzianum ity 62: obl | 64: ocoense x . 61: hace ie 48, 53. 59: ssa, 49, igo Gots Be S —— Polylepidia, 50, 53. 60-62: rent . 63: seti . 66: elect: Setosa, 49-53. bie teil. an siliquoides, 51, 64: sim- 63: spatulatum, 65: spor- D. Sh. 53¢ 3B, 57, 59, OO Way caneri. 61: wawrae, 55: wrightii. ~~ vatesih. 9. 41-43; arvense. 33. 42-44. 81: bogoten oe Sapiens 3 33-35, 81 Equisetum, 3. 36. 3 2 oe lustre, 42, 43: pratense, 42. 43: ecnorvie ri , 43, 81; 42. gang x ictale in Illinois, ial. Minnesota, and Wiscon- E — ELF range extension for aaa filix-mas, 113 = aan psa of Se I ‘airbrot worothers, D, E. Mur aig R. L. Petersen) : columbianu 59; orderoanum, 62: cordifolium, eae costari- 9; sect. Craspedoglossa. 52. 55; jo apinip ii 62: deci oratu . 60: etophytes, 93 Flora del Avila (rev.). 79 Flora de la provincia de Jujuy Republica Argentina. parte II. Pteridophyta (rev. ), Francis, P. C. (see R. L. Petersen) Futyma, R. P. bie banrenee and ecology of Phyllitis scolo- pendrium in Mic Gametophytes of sem diffusum, 39 Gastony, G. J. The deletion of Vietria graminifolia from the flora of Florida, on Gleichenia anions 26: — 27 Pb: Omez P. oths and ferns, 111 Gomez P.. L a & K. S. Walter. A double spore wall in Macroglossum, 45 ; Grammitis rigens, 26; taxifolia, 90: taenifolia, 90; xipho- teroides, 26 renee tg dryopteris. ei. Ga mnie Equisetum diffusum, 39 Henin margina’ 9 Hill, R.. A new pi record for Pilularia americana in exas, 28 How to know the ferns and ferns allies (rev.), 68 Huber, O. (see J. A. Steyermark) Hunter, D. M. (see E. E. Karrfalt) Hymenodium, 50, 63 Hymenophyllum caespitosum, 88: fucoides, 88: subg. Hymeno- phyllum, 88: — rminieri, 88: subg. Mecodium, 88: myriocar- um ndulatum, 88 o an Intersectional ap in owe 1 soé 2: drummondii. 7h: Sonera i‘ . 3) Var. iana, 3: . 1,2: mac wipers; 1-3: melanospora, 3: siesieadind, a eae 3 alt, ELE. . M. Hunter. Notes on the natural history of 69 Lellinger, D. B. Date of publication of Sodiro’s “Sertula Florae Ecuadorensis.” 96: New names for Polypodium chnoodes and P. nantes 30 loyd, R. M. Reproductive biology and tea apt ogy New World populations of Acrostichum au riopsidaceae, 48, Lonchitis, 88. 89; hirsuta, 88 rine areolata, 137 Lyco m, 42: caracasicum, 79: clavatum, 73: lucidulum. 113: cana astern japonicum, 111, 112; smithianum, 115 Maatsch, R.. Das Buch der Fendi (rev.), 92 Macroglossum, 45, 46; alidae Marattiales, sf 46 arsileac hodedcia aes. FOy8 15 a SF ay W. The economic uses = saci folklore of ferns and fern allies (rev.) Mertensia farinosa, 2 Mickel How to know the ferns and fern allies (rev.). 68 Mickel. J. T. & - spheate G. Subdivision of the Elaphoglossum, 4 Microstaphyla eed Miller, J. H. peed rences in the apparent permeability of spore walls my Serer cell walls in Onoclea sensibilis, 119 Sandstone rock crevices, an exceptional new habitat . Te simulata. 136 Moths wet fern Munt "ke pa EB; sap nici arsnhiggnre and auotoxicity in ser eastern Northern A n ferns, 124 Nephrodium cinereum. 98, var. in haicaiaeds: 98: kunthii, 89; 140 ry: 98 Jifolia ke — 89 A new count y reco n Te New ache Pol shar hnoodes and P. dsm. = A new record for Pellaea Detach in Mar . 30 New taxa and combinations of cndiatione: frou See Mexico, aierst on some ——- and Polypodium species of the Notes on the natural history of wa gemmifera, 59 Oleandra, 49: articulata a. 116, rap 20. 12 ay sersibiis. 5-81, rps 119 podum s Ree fa i : awe eee 60: cinnamomea, 73-77. 112. 116. 17, 125, 126. 129-134. 136. 137; claytoniana, 75, 116. 117. 126, 129-132: peltata, 59: E25 : . J, Petrik-Ott) Peck. J. H. Equisteum litorale in Illinois. lowa. Minnesota, and Wisconsin, 33 Pellaea oo 30: cardiomorpha, 26: cordifolia, 26: ovata. 26 79; sa sie var. cordata, Peart -_ 9.39 Petersen. R. L. & D. E. Fairbrothers. onan synthesis and ntheridium initiation in Dryopteris gametophytes, 93; Recipro- cal all ween the gametophytes of back da cin- namomea and Dryopteris intermedia. 73 Petersen. R. L & P. C. — Differential oe of fern and moss spores in response reuric chloride. 113 oe ee 2 & D. Or. ” Pilularia americana new to Pancopbi ee 79 P Phlebod . 125; aureum. en erythrolepis. 9 Psi "i. 137: scolopendrium, 81-87, 137, var. americana, si hace 28: americana. 28-30 Pil caine americana new to Tennessee. 29 de rogramma = m calomelanos. ae P y pis. 5. 9: mac . 10. 26. var. —— 26, var. macrocarpa. 26. var. chop 10, 26: polylepis. 5, pis, 9. 10 u = hispidum 24: hygro ‘ometricum, 23: um. 26: lycopodioi su ae scutulatum, “98: sororium, 30, is. 5-8: var. + 5-8. var. thysanolepis 5-9: virginianum. 81. 136: vulgare, 122: x roides. seen [ R. Supplemental notes on Lesser Antillean pterido- Polyps. 5. os psn Take 49 Polysti : bra . 113; fournieri, 27: lonchitis. 81. 82. 92 asta A 25: muelleri. 27: m sie 25; plat tea 25 Psilotum complanatum, 88: nudum. 88 AMERICAN FERN JOURNAL: VOLUME 70 (1980) Soy 88; acts 75. 124, 125, 133. var. arachonideum, 88, var. pubescens, Pteris, med 125: sent ee 98: biternata. 98: esmeraldensis, 98: falcata, 98: japonica, 75: lucida, 89; longifolia, 75; — 111, 125: procera, 98: rigida, 98: rimbachii, 98: robusta, 98: vittata, 125 ete is filix-m Reciprocal ‘tpt pent the adelante of Osmunda and Dryo as intermedia, eae D. ‘E. A new record for Pellaea etropusputes in Mary- hol of New World land, 30 scone biology and g populat “nee paris 2 urew h der ee = VIEWS: 2: The economic uses and sea Siti of ferns a bio de Jujuy Republica Argentina, parte II. Pteridophyta, 38: How o know the ferns and vee - es. ae Taxon nomy of Thelypteris subg. rei 2 (Pteridophyta), 98 hipidopteris. Sandstone rock crevices, an exceptional new habitat for Thely- Fciren simulata, 136 ond Alabama locality for the Hart’s-ton 137 sant 3; chiapensis, 25; See 25: panel: 25 Sho W. Di lazium japonic to Alabama, I11: A ‘on Alabama locality for ee! Ha rt’s-tongue, 13 ew taxa and combinations ee perido fie from ‘Gas Mexico, 15: Taxonomy of Thelypter pra pteris, including Glaphyropteris ( cpa ta) ( Smith, «1B. -E; “Wolford: & J: Li. Co sg Athyrium fives eastern Tennessee, ate Sota, E. R. de la Flora de la provincia de Jujuy Republica Argentina, parte il. Pteridophyta (rev.). 38 Sphuseropteris brunei, 111 Steyermark, J. A. & 0. Huber. Flora del Avila (rev.). 79 ge gece 72: stoma oe 69-72 Supplemental notes on Lesser An weshist oe lophytes er of eae sibs Steiropteris. inchadlsae cee 98 r aaa f W. C. (see C. R. Werth) rool heracleifolia, 27: incisa, 27. subsp. transiens. 27: transiens, 27 Thelypteris. = dentata, 75: amir nba he hispidula. 89. var. hispidula, 89. var. inconstans. ; kunthii, 89: normalis. 7S. 112, tS noveboracnesis, 75. st pina 89: palustis. tet simulata, 136, 137: el Steiropteris. 98: torresiana, 80. Pettis tela —— to the pteriodphyte flora of Escambia County. Flori Vittaria, 12: filifolia, 12, 13; graminifolia. 12-14: lineata, 12. 13. 90 Walter, K. S. ( Wendt, uae the . D. Gom pen - some Pleopelis and Polypodium species of rth, C. . WiC. Taylor: Asplenium x gravesii discovered in shisha ford. B. E(s see D. K. Smith) wardia areolata, 112 ERRATA Page 52, line 22: For “beaurepairii” read “beaurepairei.” Page 55, line 39: For “tuckerheimii” read “tuerckheimii.” Page 61, line 1: For “Petioloa” read “Petiolosa.” Page 66, line 10: “For beaurepairii” read “beaurepairei.” TRIARCH Over 5© Wears of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 AMERICAN FERN JOURNAL Volume 71 1981 PUBLISHED BY THE AMERICAN FERN SOCIETY EDITORS David W. Bierhorst Gerald J. Gastony David B. Lellinger John T. Mickel MERCURY PRESS, ROCKVILLE, MARYLAND 20852 CONTENTS Volume 71, Number 1, Pages 1—32, Issued March 30, 1981 Equisetum variegatum and E. X trachyodon in New Jersey JAMES D. MONTGOMERY Comparative Ecology of Woodsia scopulina Sporophytes and Gametophytes PAUL J. WATSON AND MARGARITA VAZQUEZ Scale Insects Feeding on Farinose Species of Pityrogramma ECKHARD WOLLENWEBER AND VOLKER H. DIETZ Spore Germination and Young Gametophyte Development of Botrychium and Ophioglossum in Axenic Culture DEAN P. WHITTIER New Species of Moonworts, Botrychium subg. Botrychium (Ophioglossaceae), from North America W. H. WAGNER, JR. and FLORENCE S. WAGNER Shorter Note: Range Extensions for Two Lycopods on Baranof Island, Southeastern Alaska Pteridophytes for the Flora Mesoamericana Reviews g; Volume 71, Number 2, Pages 33—64, Issued June 29, 1981 Azolla filiculoides New to the Southeastern United States VERNON M. BATES, JR. and EDWARD T. BROWNE, JR. The Genus Nephrolepis in Florida CLIFTON E. NAUMAN The Branching Pattern of Hypolepis repens THERESA M. GRUBER Diplazium japonicum and Selaginella uncinata Newly Discovered in Georgia WAYNE R. FAIRCLOTH Notes on Selaginella, with a New Variety of S. pallescens ROBERT G. STOLZE Lepisorus kashyapii in the Western Himalayas S. S. BIR and CHANDER K. SATIJA Taxonomic Notes on Jamaican Ferns-IlII GEORGE R. PROCTOR ~~ Notes: Notes on North American Lower ascular Plants—II; Equisetum arvense Alabama 10 13 48 Volume 71, Number 3, Pages 65—96, Issued September 30, 1981 Arachniodes simplicior New to South Carolina and the United States JUDITH E. GORDON A New Isoétes from Jamaica R. JAMES HICKEY Leaf Turnover Rates and Natural History of the Central American Tree Fern Alsophila salvinii RALPH L. SEILER Nomenclatural Notes on Micronesian Ferns F. R. FOSBERG and M.-H. SACHET x Asplenosorus shawneensis, a New Natural Fern Hybrid Between Asplenium trichomanes and Camptosorus rhizophyllus ROBBIN C. MORAN Notes on North American Ferns DAVID B. LELLINGE Shorter Notes: Salvinia minima New to Louisiana; An Unusual Record of Asplenium trichomanes from Northeastern Florida Reviews Suggestions to Contributors Volume 71, Number 4, Pages 97—124, Issued December 30, 1981 Bog Clubmosses (Lycopodiella) in Kentucky Chain Ferns of Florida TERRY W. LUCANSKY Spore Germination Patterns in Anogramma, Bommeria, Gymnopteris, Hemionitis and Pityrogramma CLARK S. HUCKABY, R. NAGMANI, and V. RAGHAVAN Shorter Notes: The Chemoidentity of the Holotype of Pityrogramma triangularis; A Major Range Extension for Thelypteris simulata in the Southern Appalachians; A New Indiana Station for Epiphytic Resurrection Fern Review American Fern Journal Index to Volume 71 Erratum R R. CRANFILL 69 101 109 MISSOURI BOTANICAL APR 44 son] AMERICAN ows FERN | aed J O U R N A @ January—March, 1981 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Equisetum variegatum and E. x trachyodon in New Jersey JAMES D. MONTGOMERY Comparative Ecology of Woodsia scopulina Sporophytes and Gametophytes PAUL J. WATSON AND MARGARITA VAZQUEZ = 3 Scale Insects Feeding on Farinose Species of Pityrogramma ECKHARD WOLLENWEBER AND VOLKER H. DIETZ 10 Spore Germination and Young Gametophyte Development of Botrychium and Ophioglossum in Axenic Culture DEAN P. WHITTIER = 13 New Species of Moonworts, Botrychium subg. Botrychium (Ophioglossaceae), from North America a eres : W. H. WAGNER, JR. and FLORENCE S. WAGNER 20 Shorter Note: Range Extensions for Two Lycopods on Baranof Island, Southeastern Alaska Pteridophytes for the Flora Mesoamericana Reviews The American Fern Society Council for 1981 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, OH 45701. President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN 37235. Vice President MICHAEL I. COUSENS, Faculty of Biology, University of West Florida, Pensacola, FL 32504 Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, VA 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, DC 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Berkeley, CA 94720. Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, NY 10458. Newsletter Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, DC 20560. ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of ae Indiana Padon Se Bloomington, IN 47401. JOHN T. MICKEL w York Botanical Garden, Bronx, NY 10458. The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and os at the Smithsonian Institution, Washington, DC 20560. Second-class postage paid at Washin Claims for missing issues, made 6 months (domestic) to 2 month (foreign) after the date of issue, and the matters for publication should be addressed to the Changes of address, dues, and applications for membership ae be sent to Dr. J. E. Skog. Dept. of Biology, George Mason University, Fairfax, VA 22030. Orders for back issues should be addressed to the Treasur General inquiries concerning ferns should be addressed to ae Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50): sent free to members of the American Fern Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or less, $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is editor of the newsletter “Fiddlehead Forum.” The editor welcomes contributions from members and non-members, including miscellaneous notes. offers to shoe or purchase materials, personalia, horticultural notes, and reviews of non-technical books on fern Spore Exchange Mr. Neill D. Hall, 1230 rire 88th Street, Seattle. WA 98115. is Director. Spores exchanged and collection lists sent on reque Gifts and Bequests Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Botanical books, back issues of the Journal, and cash or other gifts are always welcomed. and are tax-deductible. Inquiries should be addressed to the Secretary. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 1 (1981) l Equisetum variegatum and E. xX trachyodon in New Jersey JAMES D. MONTGOMERY* Equisetum variegatum Schleich., the Variegated Scouring-rush, occurs northward from the northern tier of states of the United States into Canada, on damp sand or shores, often on calcareous substrate (Fernald, 1950; Wherry, 1961). This paper reports the first verified record of E. variegatum from New Jersey and discusses a dubious record from that state. The hybrid E. X trachyodon A. Br. is also reported for New Jersey. Equisetum variegatum was listed for New Jersey by Taylor (1915) and Fernald (1950), but not by Chrysler and Edwards (1947), Wherry (1961), or Hauke (1963). The record for New Jersey was apparently based on an undated specimen collected by C. F. Austin, labeled “Closter, Bergen Co.” Chrysler and Edwards (1947) disallowed this record, indicating that E. variegatum and E. pratense were mounted on one sheet and concluding that E. variegatum was mounted by mistake with E. pratense. Further investigation by the author showed that there are specimens of many pteridophytes from New Jersey collected by C. F. Austin, usually labeled only “Closter, Bergen Co.” and without dates or specific locality. A total of 36 species of the more common northeastern ferns and fern allies are represented at CHRB or NY by such collections. It is interesting that specimens of plants known to be difficult to grow or transplant (e.g., Lycopodium spp., Botrychium spp. other than B. dissectum) are lacking. I suggest, therefore, that at least some of these C. F. Austin collections represent garden plants which were originally collected at various places in western New Jersey and probably elsewhere. This conclusion is supported by the fact that, in addition to the mixed collection of E. variegatum and E. pratense, there is a sheet of E. fluviatile at NY with labels from both “Closter, Bergen Co.” and “St. Lawrence Co., NY,” and another undated sheet of FE. pratense labeled “Closter and Sparta.” I agree with Chrysler and Edwards that the record for E. variegatum from Closter, Bergen Co. is highly suspect. Other Closter records by this collector without specific location or date are likewise dubious. A hybrid involving E. variegatum has been known from New Jersey since at least 1950: E. x trachyodon is the hybrid between E. variegatum and E. hyemale L. As far as is known from herbarium records, this hybrid was first collected by J. L. Edwards along the Delaware River, near Flatbrookville, Sussex Co., 28 October 1950 (CHRB, NY). On 19 August 1977, Vincent Abraitys collected plants of E. variegatum neat Marksboro, Warren Co. I came across this material in connection with a project to revise the Chrysler and Edwards book. Material was sent to Dr. Richard Hauke, who verified the identification. *Ichthyological Associates, Inc., R. D. 1, Berwick, PA 18603. Volume 70, number 4, of the JOURNAL was issued December 30, 1980. 9 AMERICAN FERN JOURNAL: VOLUME 71 (1981) The discovery of E. variegatum in Warren County, New Jersey, represents a southward range extension for this species. The plant is known in New York from a collection made in 1960 along the Hudson River in southern Ulster Co. see 7 from central Columbia and Greene Cos. (NYS). The plant is more commo central New York, and there are many records from the edges of the Adieaeer (New York State Botanist’s Office, pers. comm.). The distance from the Ulster Co. site is about 100 km (60 mi); other locations in eastern and central New York are more than 160 km (100 mi) distant. The only location for EF. variegatum in Pennsylvania is from Presque Isle, Erie Co., in the extreme northwestern corner of the state (Wherry, Fogg, & Wahl, 1979). Equisetum variegatum occurs on the damp shore of a lake underlain by limestone, a habitat similar to that recorded in herbarium records from New York and Pennsyl- vania Mr. Abraitys reports that it is unlikely that plants were present about 1960, so this is represents a recently established colony. The source of the plants and means of introduction are unknown, although cultivation can be ruled out. It is cet aa that both E. variegatum and E. X trachyodon should appear recently in New Jersey. The localities are separated by approximately 15 km (9 mi). The appearance of E. x trachyodon is more easily explained since the colony is on the shore of the Delaware River which extends northward into the range of E. variegatum in New York. Herbarium records (and observation of the E. variegatum site) indicate that the hybrid became established separately from and earlier than the species in New Jersey. LITERATURE CITED senior lee A., and J. L. EDWARDS. 1947. The Ferns of New Jersey. Rutgers Univ. , New Brunswick, NJ. FERNALD, M. L. 1950. Gray’s Manual of Botany, 8th ed. American Book, New York, NY. HAUKE, R. L. 1963. A taxonomic monograph of the genus Equisetum subgenus Hippochaete. Nova —123. TAYLOR, N. 1915. Flora of the vicinity of New York. Mem. N. Y. Bot. Gard. 5:1—683. WHERRY, E. T. 1961. The Fern Guide. Doubleday, Garden Cit Xs OGG, Jr., and H.A. WAHL. 1979. Atlas of the Flora of Pennsylvania. Morris Arbore- tum, Philadelphia, PA. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 1 (1981) 3 Comparative Ecology of Woodsia scopulina Sporophytes and Gametophytes PAUL J. WATSON and MARGARITA VAZQUEZ* Woodsia scopulina D. C. Eaton, an obligate rock fern, is widespread in the Rocky Mountains and has a few disjunct populations in eastern North America. This report concerns the ecology of W. scopulina gametophytes and sporophytes. An understanding of the life history of any fern is incomplete as long as the functioning of its gametophytes remains obscure. Yet, the study of gametophyte ecology is in its infancy. Adaptations of gametophytes to cold and dessication have been explored by Pickett (1914). Hill (1971) compared the habitat requirements for spore germination and gametophyte development for three ferns in Michigan. Gametophyte population divergence and general ecology have been studied by Cousens (1979). Wagner and Sharp (1963) found that free-living Vittaria prothalli occurred in areas far north of their sporophytes. Since this discovery, several other genera of tropical ferns have been found to possess gametophytes with geographic ranges much more extensive than the sporophyte (Wagner & Evers, 1963; Farrar, 1967; McAlpin & Farrar, 1978). Page (1979) sums up much of the research on fern gametophyte ecology. Gametophyte ecology is, of course, microecology. We find this a fascinating and unusual level at which to study plant ecology, a science in which investigation of systems at the macro level is the norm. PROCEDURES Our study areas consisted of xeromesic to xeric talus slopes and rock outcrops in the immediate vicinity of Bigfork, Montana. These rocky sites represent typical Woodsia scopulina habitat, and many hundreds of sporophytes as well as thousands of gametophytes are found there. Cystopteris fragilis (L.) Bernh. also occurs at the sites; however, it is only a minor contaminant of the nearly pure W. scopulina fern communities. Other herbaceous plants are only sparsely distributed on the sites. Shrubs such as Mountain Spray (Holodiscus discolor (Pursh) Maxim.), Serviceberry (Amelanchier alnifolia Nutt.), and Rocky Mountain Maple (Acer glabrum Torr. ) occur, especially on the less disturbed sections of the talus slopes. Douglas-fir (Pseudotsuga menziesii (Mirbel) Franco.) also occurs sparingly. A rich moss flora is present on the talus and the rock outcrops. We explored the microhabitats of numerous W. scopulina gametophytes, gameto- phytes harboring juvenile sporophytes (in which the sporophytes’ leaves were still dichotomous, not yet resembling those of the mature sporophyte), young, sterile Sporophytes, and fully-developed, fertile sporophytes. For each of these life cycle phases, we noted such environmental factors as substrate composition and pH, sunlight exposure, and nearby plant associates. Consistently occurring differences in the habitat of each of the four phases were recorded. *532 University Ave., Missoula, MT 59801. AMERICAN FERN JOURNAL: VOLUME 71 (1981) FIG. 1. Talus hollows where Woodsia scopulina gametophytes are found. FIG. 2. Two clumps of soil illustrating juvenile sporophytes (SP) in association with mosses and gametophytes (GP) growing on soil, WATSON & VAZQUEZ: COMPARATIVE ECOLOGY OF WOODSIA 5 Quantitative data on the substrate preferences of gametophytes were gathered by examining all the cave-like talus hollows on a chosen talus slope. Sixty-one hollows were inspected. For each, the presence or absence of gametophytes and nature of the substrate (bare soil, litter on soil, moss and litter on soil, etc.) were recorded. (We define litter as decomposing but still recognizable plant material upon soil or rock surfaces.) In order to compare the habitat preferences of W. scopulina gametophytes and sporophytes with those of a typical mesophytic forest fern, we familiarized ourselves with the habitats of the life cycle phases of Athyrium filix-femina (L.) Roth. The study sites for this fern were located in the moister forested areas of the University of Montana Biological Station, approximately 12 miles south of Bigfork. RESULTS The puberulent, slightly sticky gametophytes of W. scopulina were found only in cave-like hollows on the talus slopes (Fig. /) and in crevices on rock cliffs and tables. These talus hollows and rock crevices contain various amounts of soil, bare rock, mosses, and litter. Gametophytes existed only on bare soil over rock (Fig. 2). Gametophytes usually were found within talus hollows containing obvious patches of bare soil (23 out of 29 hollows), but seldom were found in hollows with soil mostly covered by litter or mosses (5 out of 32 hollows). Gametophytes were never found in hollows lacking soil. A mere dusting of soil about a millimeter thick was enough to support gametophytes. Soil pH readings from gametophyte substrates ranged from 6.4 to 7.2. Gametophytes in the various populations were sparse to dense, and were not closely associated with other plants, even with the young, leafy moss shoots. Favorable gameotophyte habitats typically were sheltered from direct sunlight. However, gametophytes did not grow so far back in hollows and crevices that illumination was too heavily diminished, even if other environmental conditions were favorable. Gametophytes were oriented with their apical notches farthest from the opening of the crevice or hollow. Most gametophytes did not lie flat on the substrate, but had the apical two-thirds of their thalli slanted steeply upwards (Fig. 3) Populations of gametophytes giving rise to juvenile sporophytes also grew in the crevices and hollows, but the soil substrate often was not bare (Fig. 2). This phase of W. scopulina was usually associated with the moss Brachythecium velutinum (Hedw.) B.S.G. This small, pleurocarpous moss grew sparingly around gametophytes with new juvenile sporophytes but more robustly around those with more advanced juvenile sporophytes having two or three well-developed leaves (in the latter case the gametophytic tissue was still visable but totally chlorotic). Where B - velutinum grew densely, a second moss, Encalypta vulgaris Hedw.., often occurred intermixed with it. Occasionally, a thin layer of litter also covered the soil. Clusters of tiny juve- nile sporophytes grew from gametophytes positioned in the more illuminated portions of the microhabitat as, for example, near the front of talus hollows, but still out of direct sunlight. Those gametophytes furthest back in such a hollow often had few or no sporophytes. F AMERICAN FERN JOURNAL: VOLUME 71 (1981) Young, sterile sporophytes were seen growing only out of talus hollows, rock crevices, and upon rock tables. When seen growing upon a rock table, examination of the plants’ bases showed them to be anchored to at least a small crack or other irregularity in the rock. These sporophytes usually grew in close association with several mosses such as Dicranum scoparium (L.) Hedw., Rhytidiadelphus triquetrus (Hedw.) Warnst., Tortula muralis Hedw.. Homalothecium sp., and Brachythecium sp. Openings that harbored young, sterile sporophytes supported only one AY G wiv, gh. "eich TED ONY f oS . ogo biel 4 : Aprreen : AA\SO1L .SBORELING GAMETOPHYTES Ay @ MMH FIG. 3. Position of gametophytes and juvenile sporophytes within a typical talus hollow. Fully developed, fertile sporophytes of W. scopulina were found growing out of talus hollows, rock crevices, and from irregularities on rock tables and cliffs. Their fronds always extended completely out of the rock opening of their origin, but grew in filtered to direct sunlight, depending on the presence of overstory shrubs or trees. Mature sporophytes had huge systems of fine roots. These roots were much greater in extent and density than we expected for a fern so small as W. scopulina. The roots never grew on bare rock. Instead, they grew within a layer of soil over rock and spread down into even the thinnest layers of soil between rocks. Thick mats of the same mosses associated with young, sterile sporophytes commonly covered the soil substrate. Woodsia sporophytes were not associated with thick growths of grasses and did not occur on sites with a great abundance of deciduous litter. The pH of the soil substrates we tested ranged from 6.2 to 6.8. The prothalli of Athyrium filix-femina were found only on bare soil or moist, heavily rotted wood in shady microhabitats. As was true for W. scopulina gameto- phytes, A. filix-femina gametophytes did not grow among mosses or upon forest litter. Sporophytes of A. filix-femina occurred in mesic forest conditions, often beneath gaps in the tree canopy. Mature sporophytes grew in association with many herbs, shrubs, and thick layers of mosses. WATSON & VAZQUEZ: COMPARATIVE ECOLOGY OF WOODSIA 7 DISCUSSION We conclude that there are at least three salient differences in the ecologies of Woodsia scopulina gametophytes and sporophytes: (1) Gametophytes become estab- lished only in secluded microhabitats where sunlight is diffuse for most or all of the day. Sporophytes, on the contrary, can tolerate exposure to intense direct sunlight for many hours each day. (2) Competitive abilities of the two generations differ. Gametophytes cannot coexist with other plant growth, including the mere leafy shoots of small bryophytes. On the other hand, even the youngest sporophytes are often surrounded by mosses with no apparent ill effects. Association with mosses may even benefit sporophytes by reducing desiccation of their roots. Sporophytes also grow well in close proximity to widely spaced herbaceous and woody angiosperms. (3) Gametophytes cannot grow upon or under litter, but sporophytes are not disadvantaged by moderate litter accumulations. The above conclusions conflict with certain general statements in the literature, such as that by Nayar and Kaur (1971) who claim that “sporophytes and gameto- phytes have nearly the same ecological requirements.” The first difference in ecological requirements mentioned above is a function of place and, of course, results from sporophytes growing towards sunlight. Gameto- phytes typically orient towards sunlight, but they do not grow towards it to any extent. The latter two ecological differences mentioned above are a function of time rather than place. The microhabitats that provide suitable conditions for gametophytic growth contain patches of fresh, bare soil in talus hollows and rock crevices. Such microhabitats are ephemeral and exist due to very recent accumulation of dust and soil or to soil-churning rock movements. Sporophytes grow in the same places, but after other plant life has invaded and litter has accumulated. Although the sporophytes of W. scopulina and A. filix-femina are adapted to different habitats, their gametophytes exist in strikingly similar habitats. Gametophytes of both genera grow on bare substrates in relatively moist and shady microhabitats. Difference in the sporophytes’ yet similarity in the gametophytes’ habitat require- ments of these two ferns suggests that specialization of the haploid generation of W. scopulina has lagged behind that of the diploid phase. This observation reinforces the concept that the evolution of ferns is primarily a diploid affair (Wagner, pers. comm.). It also is in harmony with the suggestion of Cousens (1979) that a potential exists for some degree of independent evolution by the two generations. Cousens suggestion is based upon the observation of Pray (1968) that differences among populations of Pellaea andromediifolia gametophytes were not correlated with differences in the sporophyte generation. Woodsia scopulina gametophytes are evolutionary conservatives, developing only in the most mesic, forest-like microhabitat available on the otherwise xeric, rocky macrohabitat to which the sporophyte has become adapted. 8 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Nonetheless, it may be that Woodsia gametophytes have made modest advances towards becoming xerophytes. Unicellular glandular hairs and a slightly sticky surface coating may be desiccation-inhibiting adaptations. (We note however that Stokey (1951) states that hairs of this type are widely distributed and usually not of generic significance.) Also, the nonoccurrence of Woodsia gametophytes on rotting wood in the immediate vicinity of the study sites indicates that spore germination on this substrate has been selected against. Perhaps the ability to grow on wood has been traded for some xerophytic adaptation. Our studies suggest that the frequency and abundance of W. scopulina gameto- phytes in suitable microsites is greater than that of A. filix-femina gametophytes. This has interesting implications. Perhaps ferns that do not reproduce extensively by means of their rhizome system, such as W. scopulina, produce gametophytes of greater vigor than ferns capable of such asexual reproduction and dispersal, such as A. filix-femina. In a study of this type, the tremendous individual mortality due to random phenomena and intraspecific competition that may take place as plants of a species struggle to advance from one developmental stage to another becomes highly evident. For example, only those spores of W. scopulina that happen to land on properly illuminated bare soil in specially protected areas among rocks will germinate and develop into gametophytes. Spore wastage must be incredibly high. Of the gametophytes that do develop, some are not in the right portion of the microhabitat to give rise to new sporophytes, such as gametophytes found towards the inner recesses of talus hollows. The threshold light intensity allowing gameto- phyte development may be too low for the growth of new sporophytes. Alternatively, it may be that not enough free water reaches the deeper parts of talus hollows to allow much sperm transfer from antheridia to archegonia. Thus, innumerable gameto- phytes are also wasted. Further, out of the mats of juvenile sporophytes that arise from part of a gametophyte population, only one ultimately survives to occupy the crevice or hollow as a young, sterile and latter as a fertile plant. In W. scopulina, almost all intraspecific competition takes place between juvenile sporophytes. New sporophytes promptly kill their parent gametophytes, possibly by secreting a toxic chemical or by parasitism, thereby avoiding competition from other sporophytes that otherwise might arise on the same gametophyte. Competition between juvenile sporophytes of different gametophytes probably is mostly for available light. The reason most intraspecific competition takes place between juvenile sporophytes is because of the small size and spotty distribution of microhabitats suitable to nurture the birth, growth, and continued survival of W. scopulina sporophytes. We extend deep graditude to Dr. W. H. Wagner, Jr., whose counsel made this study possible. Special thanks are also due Dr. C. N. Miller for valuable support throughout the study. We would like to extend our appreciation to the other readers of our final report, Dr. D. E. Bilderback, Dr. F. Wagner, and R. Moran. We also thank R. Hoham for aid in moss identification, Dr. C. A. Speer for providing darkroom facilities, D. Watson for help in preparing photographs, and the students and faculty of Yellow Bay Biological Station for their friendly support. WATSON & VAZQUEZ: COMPARATIVE ECOLOGY OF WOODSIA 9 LITERATURE CITED COUSENS, M. I. 1979. Gametophyte ontogeny, sex expression, and genetic load as measures of population divergence in Blechnum spicant. Amer. J. Bot. 66:116—132. FARRAR, D. R. 1967. Gametophytes of four tropical fern genera reproducing independently of their sporophytes in the southern Appalachians. Science 155:1266-1267. HILL, R. H. 1971. Comparative habitat requirements for spore germination and prothallial growth in three ferns from southern Michigan. Amer. Fern J. 61:171-182. McALPIN, B. & D. R. FARRAR. 1978. Trichomanes gametophytes in Massachusetts. Amer. Fern J. 68:97-98. NAYAR, B. K. and S. KAUR. 1971. Gametophytes of homosporus ferns. Bot. Rev. 37:295-396. PAGE, C. N. 1979. Experimental aspects of fern ecology. In A.F. Dyer (ed.). The Experimental Biology of Ferns. Academic Press, New Yor : PICKETT, F. L. 1914. Some ecological adaptations of certain fern prothalli—Camptosorus rhizophyllus Link., Asplenium platyneuron Oakes. Amer. J. Bot. 1:477-498 PRAY, T. R. 1968. Interpopulational variation in the gametophytes of Pellaea androemedaefolia. Amer. J. Bot. 51-960. STOKEY, A. G. 1951. The contribution of the gametophyte to classification of the homosporous ferns. Phytomorphology 1:39-58. WAGNER, W. H., Jr, and R. A. EVERS. 1963. Sterile prothallial clones (Trichomanes”) locally abundant on Illinois sandstones. Amer. J. Bot. 50:623. _and A. J. SHARP. 1963. A remarkably reduced vascular plant in the United States. Science 142:1483-1484. 7 REVIEW “FERNS, FERN ALLIES AND CONIFERS OF AUSTRALIA,” by H. T. Clifford and J. Constantine. xviii + 150 pp. illustr. University of Queensland Press. 1980. ISBN 0-7022-1447-7. $24.25.— Although this book is subtitled “A Laboratory Manual,” and does contain illustrations of morphological details, it is a good introduction to the pteridophyta of Australia, and, by extension, to the Old World tropics. Two-thirds of the book concerns pteridophyta. For those interested in identifying Australian ferns and fern allies, there are keys down to species, interesting and useful generic descriptions with notes on habitats, and tables of distribution by species within Australia and Tasmania. Species descriptions are missing, and are not really compensated for in the brief keys to species. Fortunately, the genera of Australian ferns are diverse and mostly with only a few species, and so identification is likely to be easy in most cases. References and literature cited, a table of vegetative characteristics of major vascular plant groups, a list of synonyms of Australian pteridophyta with the names accepted by the authors, a short glossary, and an index conclude this useful work —D. B. L. 10 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Scale Insects Feeding on Farinose Species of Pityrogramma ECKHARD WOLLENWEBER and VOLKER H. DIETZ* Several well known species of the fern genus Pityrogramma are called Silverback Ferns or Goldback Ferns because of the conspicuous white or yellow farinose deposit on the abaxial surface of their fronds. This material is composed of flavonoid aglycones, mostly of dihydrochalcones and chalcones (Wollenweber, 1978; Wollenweber & Dietz, 1980). The correct structural formulae of these compounds are shown in Fig. /; the formulae in both cited papers are incorrect. These lipophilic phenolic compounds are secreted by glandular trichomes and form quasi-crystalline rods or filaments on the surface of the enlarged terminal cell of the capitate glands (Wollenweber, 1978, pp. 13-15). R1 ea | On OU R! OH R2 Chalcones { * 1. R'=OCH3,R2=H 2. R'=R*=0CH, OH O Dihydrochalcones 7 8 OCH, Roe 4. R'=R*=0CH, FIG. 1. Correct structural formulae of chalcones and dihydrochalcones, which form the major consti- tuents of the farina on Pityrogramma fronds. No doubt, glandular trichomes are anatomical features developed for some definite function, and the natural compounds secreted ought not to be regarded merely as waste products (cf. Swain, 1977). The farinose indument of Pityrogramma has been the subject of speculation as to its physiological and ecological function for about a century. Among the postulated functions are protection of the lower frond surface and the young spores against wetting on the one hand, and against water loss by transpiration on the other hand. Antibacterial and antifungal effects of such *Institut fiir Botanik, Technische Hochschule Darmst: ittspahnstrz - é tadt, Federal Republic of Germany. armstadt, Schnittspahnstrasse 3, D-6100 Darms WOLLENWEBER & DIETZ: SCALE INSECTS ON PITYROGRAMMA 1] phenolic substances have some likelihood. Insect deterrence also has been discussed (Wollenweber, 1978). Field observations indicate that these ferns are rarely attacked by feeding insects (L. D. Gomez, pers. comm.). Hoéhlke (1902) favored the latter function of the farinose deposits because he had observed that “plants in the greenhouse are free of destructive insects even in the summer.” There is as yet no evidence for any of the foregoing functions; allelopathy from frond exudates has been demonstrated (Star, 1980). FIG. 2. Saissetia on the pinna costa and pinnules of Pityrogramma austroamericana, showing older and younger glandular trichomes covered with exudate (Leitz Reprovit II). FIG. 3. Close-up of Saissetia (Cambridge Stereoscan 600), x 30. For several years, we have been growing plants of Pityrogramma austromaericana, P. calomelanos, P. chrysophylla, and P. trifoliata in a greenhouse at Darmstadt. We Saissetia (Coccidae) thrives on our Silverback and Goldback ferns. The insects are found on the abaxial costae surface of old fronds, and also on the pinnule surface. The insects seem to be not harmed or irritated by the rich flavonoid deposits, and they grow and reproduce quite well while sitting in and between the exudate (Figs. 2 and 3). We are able to rid the plants of the insects for several months by removing infested fronds. It is not our intent to refute the possibility that flavonoid exudates may function to deter some insects, even though they do not deter Saissetia. Observations of other insects, especially in the field, would be welcome. We wish to thank Mrs. R. Heger 12 AMERICAN FERN JOURNAL: VOLUME 71 (1981) (Darmstadt) for the photograph and Dr. W. Bartholott ‘aie for the SEM picture. We are grateful to Dr. D. R. Miller of the Dept. of Agriculture (Beltsville) for determining the genus of the scale insect. LITERATURE CITED HOHLKE, F. 1902. Uber die Harzebehilter und die Harzbildung bei den Polypodiaceen und einigen Phanerogamen. Beih. Bot. Centralbl. 11:8-45 STAR, A.E. 1980. Frond exudate flavonoids as allelopathic agents in Pityrogramma. Bull. Torrey Bot. Club 107:146-153. SWAIN, T. 1977. Secondary compounds as protective agents. Ann. Rev. Pl. Physiol. 28:479-S01. WOLLENWEBER, E. 1978. The distribution and chemical constituents of the farinose exudate in iaiaiglanig ferns. Amer. Fern J. 68:13-28. ——, and V. H. DIETZ. 1980. Flavonoid patterns in the farina of goldenback and silverback ferns. Biochem: Syst. Ecol. 8:21-33. REVIEW “FERNS AND FERN ALLIES OF KENTUCKY,” by R. Cranfill, Kentucky, Nat. Pres. Comm. Sci. Techn. Ser. 1. 284 pp. 1980. $4.50.—This new pteridophyte Flora is a model for what state floristic treatments should be. General topics, covered in about 35 pages, include collecting history, phytogeography and ecology, pteridophyte life history, identifying pteridophytes, and a statistical summary of Kentucky’s pteridophytes (69 species and 10 hybrids in 29 genera are known). The technical treatment includes a synoptical key to families, a key to genera using fertile and another using sterile material (the latter a very worthwhile novelty for ecolo- gists), and treatments of the genera and species, including keys, synonymies, notes, and illustrations. Although there are no descriptions, the keys are fairly extensive and very thorough notes help to distinguish critical taxa. Distribution maps grouped at the back of the book show the location of each species and the herbarium where vouchers documenting the locations may be found. Seventeen herbaria were consult- ed, as well as the literature, in compiling the maps. An illustrated glossary, extensive literature cited, and an index conclude the volume. This book is a must for anyone studying ferns in the east-central portion of the United States. It is available rom the Kentucky Nature Preserves Commission, 407 Broadway, Frankfort, KY, 40601.—D.B.L. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 1 (1981) 13 Spore Germination and Young Gametophyte Development of Botrychium and Ophioglossum in Axenic Culture DEAN P. WHITTIER* Spores of the Ophioglossaceae have been sown by numerous investigators, but few have succeeded in germinating them (Boullard, 1963). Du Buysson (1889) reported germinating the spores of Ophioglossum vulgatum and Botrychium ternatum. The spores of B. virginianum and three tropical species of Ophioglossum (O. moluccanum, O. pendulum, and O. intermedium) were germinated by Campbell (1895, 1907). Recent studies have shown that spores of B. dissectum f. obliquum (Whittier, 1972) and B. multifidum (Gifford & Brandon, 1978) can germinate in axenic culture. Spores sown in axenic culture required darkness, took weeks to germinate, and, at least for B. dissectum, produced a three-dimensional gametophyte immediately after germination (Whittier, 1973; Gifford & Brandon, 1978). None of the earlier reports by Campbell and du Buysson gave any indication that darkness was required for germination. In soil cultures, O. moluccanum, O. vulgatum, and B. ternatum germinated in a few days, which is a period comparable to that for spores of the Polypodiaceae (Campbell, 1907; du Buysson, 1889). Young gametophytes of O. vulgatum and B. ternatum were illustrated by du Buysson (1889) with filamentous and two-dimensional growth habits which are similar to those for polypodiaceous gametophytes. Consequently, the present study of additional species was carried out in axenic culture to determine how the requirements for germination and the type of early gametophyte development for these species compared with the results from previous studies. MATERIALS AND METHODS Spores of eight Ophioglossaceae species, Botrychium biternatum ( Sav.) Underw., B. dissectum Spreng. var. dissectum, B. lunarioides (Michx.) Swartz, B. matri- cariifolium A. Braun, B. virginianum (L.) Swartz, Ophioglossum engelmannii Prantl, and O. vulgatum var. pycnostichum Fern. and var. pseudopodum (Blake) Farw., were used in this study. With the exception of the spores of B. matricartifolium and O. vulgatum var. pseudopodum, which were from Ontario, the spores were collected in the southeastern United States. Voucher specimens are on deposit in the Vanderbilt University Herbarium (VDB). The techniques of Whittier (1973) were employed. The spores were sown on ? ml of nutrient medium in culture tubes with a diameter of 20 mm. The tubes had screw caps which were tightened to reduce moisture loss. The nutrient medium was composed of Knudson’s solution of mineral salts, minor elements, FeEDTA, and 0.6% agar (Whittier, 1973). The medium was supplemented with 0.5% sucrose and had a pH of 6.3. The spores were cultured at 24+1°C in the light at an intensity of 1400 lux from cool white fluorescent lamps or in darkness. For some treatments, the cultures were maintained at 3+1°C in a cold room. *Dept. of Biology, Vanderbilt University, Nashville, TN 37235. 14 AMERICAN FERN JOURNAL: VOLUME 71 (1981) For the first eight weeks after sowing, the spores were removed weekly from at least six cultures of each species and examined for germination. After eight weeks, the cultures were sampled at irregular intervals. The percentage of germination was determined after examining 1000 or more spores. Early gametophyte stages of development were cleared and stained with acetocarmine-chloral hydrate according to Edwards and Miller (1972). The cleared and stained gametophytes were drawn with a camera lucida for study. No effort was made to indicate either the walls of cells lying behind illustrated cells or the spore coat if one was present. However, the nuclei of cells positioned behind other cells are denoted by dotted lines. OBSERVATIONS A few of the spores of most species studied germinated in the dark on the nutrient medium at 24°C (Table 1). Lengthening the dark periods for those spores which germinated under these conditions increased the amount of their germination. Even after extended periods of time, the spores did not germinate in illuminated cultures. TABLE |. GERMINATION OF SPORES OF THE OPHIOGLOSSACAE “ AXENIC CULTU S pecies Weeks in darkness Percent germination Botrychium biternatum ~ 0.5 B. dissectum var. dissectum 8 0.1 B. lunarioides 3 0.7 B. matricariifolium 8 0.5 B. virginianum be 0 Ophioglossum engelmannii! 52 0 O. vulgatum var. pseudopodum 16 0.7 0 O. vulgatum var. pycnostichum 52 'See text as conditions which promote germina’ fy th of time in darkness bette: Ber ipation varies considerably. The fastest pais was three weeks for B. lunarioides. The spores of O. vulgatum var. pseudopodum took the longest time to germinate at 24°C; although the cultures were not sampled on a weekly basis after eight weeks, initial gametophyte stages were found in the cultures of this species after 16 weeks in the dark, which indicated that the spores had germinated recently. Thus, under these conditions, spores of the Ophioglossaceae do not have identical dark requirements for germination. Sowing the spores in darkness at 24° did not promote germination in B. virginianum, O. vulgatum var. pycnostichum, or O. engelmannii. However, 0.1% of the spores of O. engelmannii germinated in the dark when maintained at 24°C for three months, followed by 3°C for three months, and returned to 24°C for six months. Germination did not occur during the cold treatment; consequently the final warm period was necessary for germination. When the first warm period was omitted and the spores were placed directly into the cold (3°C) for three months, no germination occurred. Other modifications of the warm and cold temperature regime are now under investigation in an attempt to increase the amount of germination. Although successful for O. engelmannii, the regime of warm and cold temperatures in the dark failed to bring about germination in B. virginianum and O. vulgatum vat. pycnostichum. D. P. WHITTIER: BOTRYCHIUM AND OPHIOGLOSSUM IN AXENIC CULTURE 15 The early developmental stages of gametophytes were studied in the five species which germinated in the dark at 24°C (Table 1). Spores of these species exhibited similar early development. The spore coat cracked open at the triradiate ridge and the spore divided transversely to its polar axis (Figs. /, 14, 20 and 27), producing a distal cell (away from the triradiate ridge) and a somewhat smaller proximal cell (near the triradiate ridge). The proximal cell enlarged, forcing the three lobes of the FS" (\B> fo oS BO oe wy Is =) G FIGS. 1-35. Stages in the early development of Botrychium and Ophioglossum gametophytes, x 275. The bottom cell in Figs. 1-33 is the proximal cell. FIGS. 1-6. B. matricariifolium. FIGS. 7-13. B. dissectum var. dissectum. FIGS. 14-19. B. lunarioides. FIGS. 20-26. B. biternatum. FIGS. 27-35. O. vulgatum var. pseudopodum. cracked spore coat apart, and bulged out of the spore coat. The distal cell, sch remained inside the spore coat for a longer period of time, continued to divide. The second division was longitudinal and divided the distal cell into two nearly equal cells (Figs. 2, 7, 15, 21 and 28). The third division, which was either longitudinal or transverse, occurred in one of the two distal cells (Figs. 3, 8, 16, and 29). The fourth division occurred in the other cell at the distal end. Usually the plane of division was perpendicular to the plane of the third division (Figs. 10, 17, 22, and 30) but occasionally its plane was almost the same as the plane of the third division (Figs. 4 and 9). The development to the 5-celled gametophyte usually occured inside the spore coat. Usually the gametophytes lost the spore coat by the 10-celled stage, although it was seen occasionally on gametophytes with more — 16 AMERICAN FERN JOURNAL: VOLUME 71 (1981) The exact pattern of the later divisions was not followed for two reasons: there was more variation in the sequence of the later divisions (Figs. 5, 6, 11, 12, 18, 23, 24, and 25) and the three dimensional form of the young prothalli made it difficult to follow the divisions. However, a small, more or less spherical or globular gameto- phyte formed as a result of these later divisions (Figs. 13, 19, 33, and 36). The mature gametophytes with morphology typical of the species developed from the globular stage. The early sequence of divisions in the gametophytes of O. vulgatum var. pseudopodum was similar to that for the species of Botrychium. However, at about the 7-celled stage (Fig. 32), the Ophioglossum gametophytes appeared to have an apical cell. Side and apical views of 10-celled gametophytes showed definite apical cells (Figs. 33 and 34). Older gametophytes had a typical apical cell with three cutting faces (Fig. 35). Botrychium gametophytes of similar sizes or cell numbers had no recognizable apical cells (Figs. 1/3 and 26). FIG. 36. Young globular gametophyte of O. engelmannii, x 275. The young gametophytes of O. vulgatum var. pseudopodum were larger than those of Botrychium because the individual cells of the Ophioglossum gametophytes were larger. More than likely, the larger cell size in Ophioglossum gametophytes is controlled by the same factor which caused the spores of O. vulgatum vat. pseudopodum to be larger than those of the Botrychium species. The average long dimension for O. vulgatum var. pseudopodum spores was 43.8 ym; Botrychium spore averages ranged from 31.0 to 34.9 um. The spores of O. vulgatum vat. pseudopodum were not exceptional; those of the other Ophioglossum species studied also were larger, with their longest dimensions averaging greater than 40 pm. DISCUSSION AND CONCLUSIONS The dormancy of those Botrychium and Ophioglossum spores which germinated is broken by culturing them in the dark. Thus, the findings of Whittier (1973) and Gifford and Brandon (1978) are corroborated by the results of the present study. However, not all the spores germinate after a dark period at 24°C. In O. engelmannii, germination occurs after a sequence of warm and cold periods in the dark. Although the germination in O. engelmannii is limited and does not increase with extensions of the final warm/dark period, the sequence of warm/dark, cold/dark, and warm/dark periods is the only treatment which promoted germination. It remains to be ¢ D. P. WHITTIER: BOTRYCHIUM AND OPHIOGLOSSUM IN AXENIC CULTURE 17 determined if increasing the duration of the first warm/dark period and/or the cold/dark period would increase germination in O. engelmannii. In B. virginianum and O. vulgatum var. pycnostichum spores, there is no germination after a dar period of 52 weeks at 24°C or following a sequence of warm and cold periods, as for O. engelmannii. Although the dormancy of most species’ spores is broken by a dark period, spores of O. engelmannii, O. vulgatum var. pycnostichum, and B. vir- ginianum demonstrate that either a longer period in the dark at 24°C or a more elaborate treatment is necessary for germination. Campbell (1895, 1907) germinated spores of four species of Ophioglossaceae, including those of B. virginianum, on soil or humus. These spores must have germinated in the light because Campbell noted whether or not there was a trace of chlorophyll in the young gametophytes. In three of the species, germination was slow (taking a month or more), but the spores of O. moluccanum (now O. petiolatum) from Java germinated in three days. Sussman (1965) suggested that Campbell possibly used old spores which had lost their dormancy. At least for the Ophioglossum species, this seems unlikely because Campbell ( 1907) discussed collecting and sowing the spores in a relatively short time. The germination of these Ophigglossum spores in the light is contrary to the findings for axenic culture. In Lycopodium, on the other hand, the speed of germination varies considerably (Barrows, 1935). Some of the more tropical species of Lycopodium tend to germinate rapidly and on the surface of the soil, compared to the slower underground germination of the more temperate species. Possibly a similar situation exists with Ophioglossum, with the spores of the tropical species germinating more rapidly and in the presence of light. The spores of B. virginianum which germinated for Campbell (1895) but which failed to germinate in axenic culture present a different problem. These spores are from a temperate zone species in which dormancy is not broken in axenic culture. Although the age of the spores used by Campbell is unknown, their age may not be that significant because spores of B. virginianum up to two years old failed to germinate in axenic culture. More time may be necessary for germination; Campbell (1895), for instance, found early stages of germination after 18 months. Jeffrey (1897), after failing to get germination after 18 months, suggested that the warmer climate in which Campbell worked may have been a factor. Increasing the duration of the experiments and employing higher temperatures along with other variations are now being tested on the spores of B. virginianum. ae The earliest stages of gametophyte development for the five taxa studied in detail are consistent with the reports of Campbell (1895, 1907). The gametophytes in axenic culture never produced a filament or plate of cells. Within a few divisions and usually before the young gametophyte had completely broken out of the spore Coat, a 3-dimensional growth pattern was established. Although Campbell (1907) reported that the proximal cell usually divided in the early stages of the species he Observed, the proximal cell remained undivided in the young gametophytes of the present study. However, there are indications that the proximal cell may divide at later stages in the development of the gametophytes in axenic culture. The undivided proximal cell in the young gametophytes and the total lack of chlorophyll (because 18 AMERICAN FERN JOURNAL: VOLUME 71 (1981) the gametophytes in axenic culture were grown in total darkness) are essentially the only differences found between the gametophytes from axenic culture and those observed by Campbell (1895, 1907). All the species that have been studied have a similar pattern of early development which produces a small, globular gametophyte. Campbell (1907) suggested that an apical cell was established very early in Ophioglossum gametophytes. In the present study, O. vulgatum, which has a prominent apical cell in the mature gametophyte (Bruchmann, 1904), set off an apical cell at possibly the sixth division. In slightly older gametophytes (10 or 11 cells), a characteristic apical cell with three cutting-faces was present. Campbell (1907) observed the earliest stages of apical cell formation in side view, but apparently he did not obtain apical views to confirm the presence of a typical apical cell. No apical cell is found in the young gametophytes of Botrychium, which probably is not surprising because the mature gametophytes of Botrychium are generally without a prominent apical cell or sometimes even without a recognizable apical cell (Bierhorst, 1958). The report of du Buysson (1889) on spore germination and early gametophyte growth of B. ternatum and O. vulgatum must be considered separately because of his unusual observations. He reported germination times comparable to that for spores of the Polypodiaceae, gametophytes with filamentous and two-dimensional growth patterns, and young gametophytes with large amounts of chlorophyll. The characters described by du Buysson (1889) are typical for the Polypodiaceae, but not for the Ophioglossaceae. The possibility that spores of the Ophioglossaceae he sowed never germinated and that stray spores of the Polypodiaceae (he was also studying the development of other fern gametophytes) germinated instead would explain his results. It is probably best to question or ignore the observations of du Buysson until they have been repeated. Some differences exist between the observations of Campbell and those from axenic culture. Whether these differences are related to the source of the spores (tropical as opposed to temperate) or to environmental variations (soil as opposed to axenic culture) remains to be investigated. Nevertheless, the studies in axenic culture basically substantiate Campbell’s descriptions of the early development of gametophytes of the Ophioglossaceae. I want to thank Mr. Alan Anderson (University of Guelph), Dr. Robert Kral (Vanderbilt University), and Dr. R. Dale Thomas (Northeastern Louisiana Universi- ty) for supplying spores of O. vulgatum var. pseudopodum, B. biternatum, and B. lunarioides, respectively; Dr. W. H. Wagner, Jr. (University of Michigan) for confirming the identification of B. biternatum; and the Vanderbilt University Research Council for support of this research. LITERATURE CITED BARROWS, F. L. 1935 Propagation of Lycopodium. I. Spores, cuttings, bulbils. Contrib. Boyce Thompson Inst. 7:267-294. BIERHORST, D. W. 1958. Observations on the gametophytes of Botrychium virginianum and B. dissectum. Amer. J. Bot. 45:1—9. BOULLARD, B. 1963. Le gametophyte des Ophioglossacées. Considérations biologiques. Bull. Soc. * Linn. Normandie 4:8 1-97. D. P. WHITTIER: BOTRYCHIUM AND OPHIOGLOSSUM IN AXENIC CULTURE 19 ogee H. 1904. ae das Prothallium und die Keimpflanze von Ophioglossum vulgatum L. , Zeit. 62:227-2 PAYSON. i du 1889. Monographs des cryptogames d’Europe. II. Filicinées. Rev. Sci. Bourbonnais Cent. France 2:153—16 CAMPBELL, D. H 1895. The ee and Development of Mosses and Ferns, Ist ed. Macmillan, dyin ork. ——— . Studies on the Ophioglossaceae. Ann. Jardin Bot. Buitenzorg, Il, 6:1 Trey yy E. and J. H. MILLER. 1972. Growth regulation by ethylene in fern 2p one Ill. Inhibition a spore germination. Amer. J. Bot. 59:458—465. GIFFORD, E. M., Jr. and D. D. BRANDON. 1978. Gametophytes of Botrychium multifidum as grown n axenic clue Amer. Fern J. 68:71-75. JEFFREY. E. C. 1897. The gametophyte of Botrychium virginianum. Trans. Roy. Canadian Inst. —2 i SUSSMAN, e S. 1965. Physiology of dormancy and germination in the propagules of i i plants. Pp. 931- eo in W. Ruhland (ed.). Encyclopedia of Plant Physiology, vol. 15, pt. Spri sateen 8 WHITTIER, D. P. 1972. igen! i of Botrychium dissectum as grown in sterile culture. Bot. Gaz. 133;336—339. . 1973. The effect of light and other factors on spore germination in Botrychium dissectum. Canad. J. Bot. 51:1791-1794. 20 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 1 (1981) New Species of Moonworts, Botrychium subg. Botrychium (Ophioglossaceae), from North America W. H. WAGNER, JR. and FLORENCE S. WAGNER* Since the genus Botrychium was last revised (Clausen, 1938), much research on it has been done, and many of our taxonomic ideas concerning the use of infraspecific categories (subspecies, variety, form) have changed (Wagner, 1960). Many Botrychium species can be distinguished only by subtle differences, and there is often a high level of variability, even in single populations. In the littke Moonworts, Botrychium subg. Botrychium, taxonomic difficulties are exacerbated by the small size of the plants and the tendency for the segments to become folded when pressed. To gain an idea of the distinctness of the species, careful field studies of sizeable populations are needed, and the specimens must be pressed so as to spread out the segments. Finding these usually rare and local plants requires “a labor of love.” Together with colleagues and students, we have spent an extaordinary amount of time exploring for Moonworts over the past several decades. Some of the fruits of this quest are described here. In making our taxonomic judgments, we have adhered to a strict requirement of consistency. Plants from one locality to another must display the same characteris- tics. Likewise, the differences from other species must be consistent; there should be no connecting forms. We have emphasized the ability of two critical taxa to occur together in the same habitats without intergrading. Mixed populations have thus been of value in making decisions. We also have used the method of mutual associations—the ability of two taxa to maintain their differences in their separate habitats, even though their mutually associated relatives stay alike. Thus if taxon A grows with taxon C in one habitat, and taxon B grows with C in another, and if taxon C remains uniform in both habitats, then the differences between A and B are probably genetically fixed. The four new Moonworts described here have been studied in wild populations, and in all cases the evidence upholds their validity as discrete species. We treat each only briefly, but we hope that the description and comments will give other workers a preliminary idea of their distinctions and also will lead to finding new populations. More detailed reports on all of these plants will be made in the future. The measurements are based upon the best-developed and most distinctive specimens from each of several collections. Small specimens may be hard to identify, and all of these species include tiny fertile individuals less than 3 cm tall. Spores were mounted in Hoyer’s solution (Beeks, 1955) and measured along the largest diameter. The ranges represent the averages of several different collections. *Herbarium and Department of Botany, University of Michigan, Ann Arbor, MI 48109. WAGNER & WAGNER: NEW SPECIES OF MOONWORTS 2] B D FIG. 1. Silhouette drawings of four new species of moonworts. A. Botrychium crenulatum. B. B. paradoxum. C. B. mormo. D. B. montanum. Botrychium crenulatum W. H. Wagner, sp. nov. Figs. 1A, 2. Folium 10 (6—16) cm altum, herbaceum, flavo-virens; segmentum sterile dispositum altum in axe, stipite 5 (1-17) mm longo, lamina ovato-oblonga vel lineari-oblonga, 2 (1.5-6.5) cm longa, 1.2 (8-18) cm lata, divisionibus lateralibus 3 (2-5) paribus, cuneatis, remotis, 6 (3-12) mm longis, 5 (3-12) mm latis; margines crenulati, et interdum paucincisi. Leaf 10 (6-16) cm tall, texture herbaceous, color (living and freshly pressed) yellowish green; common stalk 0.60 (0.50—0.80) of the total leaf length; sterile segment stalk 5 (1-17) mm long, the blade (pressed) ovate-oblong to linear-oblong, 2 (0.5-6.5) cm long, 1.2 (0.8—1.8) cm wide; segments 3 (2-5) pairs, wedge-shaped, the lower and upper margins of the medial pinnae an angle of 70° (20°—110°), their outer edges separated by 80 (30—120)% of the width of the pinnae, often narrowly cuneate toward the base, the largest pinnae 6 (3-12) mm long, 5 (3-12) mm wide, with 15 (10-25) veins reaching the distal margin; margins crenulate (entire to crenate or dentate), and in very large specimens sometimes with I—5 incisions 1/5 (up to 1/2) of the pinna length; fertile segment 4.5 (2.5-9.5) cm tall, the sporangial branches confined mainly to the upper 1/2 (1/3-2/3), without major proximal sporangial branches; spores 45 (40—48) pm in diam.; chromosomes 2n= 90. TYPE: Mt. Baden-Powell Trail, Hamell Springs, San Gabriel Mountains, Los Angeles Co., California, 7745 ft., L. L. Kiefer 1488 (MICH; isotype UCLA). 9 AMERICAN FERN JOURNAL: VOLUME 71 (1981) OTHER COLLECTIONS: CALIFORNIA: San Bernardino Co.: San Bernardino Mts.: Santa Ana River: Big Meadows, Hilend 373 (LA), Bluff Lake Meadow, | specimen with B. simplex, Hilend 238 (LA), S Fork of Santa Ana River, with B. simplex, Munz 6164 (DS). San Gabriel Mts.: N of Baldy Notch, headwaters of Coldwater Fork of Lytle Creek, Kiefer 1489 (LA). San Gorgonio Wild Area: S Fork of Santa Ana River, ca. halfway between Poopout Hill and Slushy Meadows, Kiefer 815 (LA). South Mountain: Fern Canyon, branch of Mill Creek near Mill Creek Falls, opposite Vivian Creek Falls, Robertson in 1907 (DS). Tulare Co.: Southern Sierra Nevada, ridge between forks of Monache Creek. SW of Olancha Fedak, Munz 15333 (DS). 3 » * / | > FIG. 2. Botrychium crenulatum from the type locality in the San Gabriel Mts., CA. A. View of living plant, showing outline of fertile segment. B. Same, showing spacing and shape of sterile divisions. Leaf 8 cm tal One might be tempted to treat this as a geographical subspecies of the worldwide and variable Common Moonwort, B. /unaria L. However, there are several reasons why we cannot do this. There are no intermediate populations, nor any evidence of geographical transition. The new species is approximately equally distinct morphologically from B. /unaria and the endemic North American tetraploid species B. minganense Vict. (cf. Wagner & Lord, 1956). Botrychium crenulatum (Figs. 1A and 2) may be distinguished from B. lunaria by its herbaceous (not fleshy) texture, yellow (not dark-green) color, higher placement of the sterile segment on the leaf, stalked rather than sessile sterile portion, an average of three pinna pairs (rather than five), much more remote, smaller, and narrower segments, with fewer veinlets WAGNER & WAGNER: NEW SPECIES OF MOONWORTS 23 terminating along the distal margin’, and the prevalence of crenulate, as opposed to nearly entire, outer margins (except in B. lunaria f. subincisum (Roeper) Milde, with the distal margins shallowly toothed, and f. incisum (Milde) Holmberg, with pinnae more or less deeply cut; both usually are highly sporadic). The spores of B. crenulatum average 6 wm larger in diameter than those of B. lunaria. Superficially, B. crenulatum resembles B. minganense more than B. lunaria because of the over-all proportions of the leaf and its parts. Several of the collections have been identified as B. lunaria var. minganense (Vict.) Clausen. However, B. crenulatum differs from B. minganense in its more delicate texture, the more abrupt reduction of the apex, with fewer, coarser, and more angular, rather than more rous, smaller, and rounder ultimate and penultimate segments, in the mostl crenulate rather than mostly entire distal pinna margins, in the apex of the sterile segment in the leaf primordium for the following year overtopping the fertile and clasping it (in contrast to equally and paralleling it), and in the chromosome number of 2n=90 rather than 2n = 180. Whether the plant described here is the same as the one named B. lunaria subsp. occidentalis by Léve et al. (1971) is questionable. They distinguish it from typical B. lunaria in being smaller and having a yellower color. However, the type locality, above Graymont in Clear Creek Co., Colorado, is remote from the known range of B. crenulatum. We know of no localities in Colorado or its surrounding states where B. crenulatum has been found. The specimens were probably depauperate sun-forms of B. lunaria, judging from the characters given. Crenulations are not mentioned. This Southwestern Moonwort is very rare, judging from the few collections that exist. It may occur locally, however, in large numbers. Most of the specimens come from southern California, especially San Bernardino County, at altitudes averaging around 2500 (2100-3400) m. The fern extends in the mountains of California as far north as Butte County, where we discovered it in 1949 in company with Dr. E. B. Copeland. Possibly B. crenulatum has a wider range than given here. Plants we previously identified as the South American B. dusenii (Christ) Alston from various western states may prove to be variations of B. crenulatum. Also, in Oregon and Montana, moonworts with extremely narrow segment divisions may be conspecific or closely related. Botrychium crenulatum grows in the drier places of damp meadows, boggy areas, and marshes, either on hillsides or flat lands where there are wet banks or springy spots. The plants are rooted in tussocks or “rises” around isolated trees or shrubs, or in depressions that dry out during the summer, or at the edges of marshes. They may occur either in sun or shade, but evidently prefer partial shade. The associated genera that have been recorded include woody plants Pinus and Salix, and herbs Dodecatheon, Hypericum, Liparis, Mimulus, Smilacina, and Veratrum. Various sedges and grasses under which these moonworts grow can make detecting the plants very difficult. Associated species of grapeferns include com- monly B. simplex, and rarely B. multifidum. ‘Some collections (e.g., Wagner 4609, Johnston 2080) have relatively wider segments, with angles approaching those of B. lunaria. 24 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Botrychium paradoxum W. H. Wagner, sp. nov. Figs. 1B, 3. lateral branches are sessile or nearly so and short, usually only 1-3 mm long; sporangia mostly 2 or 3 (1-12) per lateral branch. Spores 40 (36-43) wm in dia- meter. Chromosome number unknown. ad FIG. 3. Botrychium paradoxum. A. Small form of deep shade. Marias Pass, MT. Leaf 3 cm tall. Large form of open meadows from the type locality at Storm Lake, MT. Leaf 15 cm tall. TYPE: Storm Lake, ca. 6 mi S of Georgetown Lake, Deerlodge National Forest, Flint Ridge Mountains, Deerlodge Co., Montana, Wagner 80/28 (MICH; 9 plants observed). OTHER COLLECTIONS: CANADA: Alberta: Waterton Lakes National Park, 1 mi NW of Red Rock, Blair & Nagy 1280 (LEA: 6 plants observed). U.S.A.: Montana: Glacier Co.: 2 mi W of N end of Swiftcurrent Lake, Many Glaciers Trail to Fishermancap Lake above Wilbur Creek, Wagner 78547 (MICH: 2 plants). Ca. | mi W of Marias Pass, Pondera and Flathead county line, Wagner 78528, 80/17. (MICH; 45 plants). WAGNER & WAGNER: NEW SPECIES OF MOONWORTS 25 This extraordinary Moonwort is surely one of the most peculiar ferns in North America. It shows modifications of the leaf as profound as, for example, those in Schizaea pusilla (Schizaeaceae). Two species of Ophioglossum lacking sterile segments have been described: O. kawamurae Tagawa from Osumi Province, Japan, and O. lineare Schlechter & Brause from New Guinea. These plants are morphologically different from Botrychium paradoxum, however, in that the sterile segment has been lost rather than transformed. The Sumatran O. simplex Ridley has vestiges of the sterile segment (Tagawa, 1939; see also Sahashi, 1980). The fact that bladeless species of the sister genus Ophioglossum have been reported points up the fact that these are strongly mycorrhizal plants that may lose much or all of the photosynthetic lamina. In Botrychium we can find a progression from forms like B. lanceolatum (Gmel.) Angstr. to such extremes as B. montanum and B. mormo with the ample blade replaced by a narrow, much reduced blade. In B. paradoxum, the sterile lamina is not really lost, but has become converted to a second fertile segment. The two fertile segments are both apparently positively phototropic and negatively geotropic, so that they are held erect and parallel. In other words, what might be an expected loss of the sterile segment, as illustrated partially in B. montanum, B. mormo, and Ophioglossum simplex, and completely in O. kawamurae and O. lineare, does not take place. In regard to the western flowering plant, the Leafless Wintergreen, Pyrola aphylla Smith, Camp (1940) wrote that “...the Pyrolaceae are on the physiological borderline between autophytism and parasitism,” and he showed that several species of wintergreens were capable of developing aphyllous forms. Similar physiological conditions no doubt prevail also in the Ophioglossaceae. All of the fronds we have studied are without laminae. It is interesting to note, however, that there is a distinctive Moonwort, the sterile segments of which have remote and narrow divisions, that has been found at two of the four Botrychium paradoxum localities. This may represent the atavistic form of B. paradoxum, but our evidence for this 1s tenuous; it may be merely a coincidence. eee The four localities for B. paradoxum are all near the Continental Divide, extending 200 miles (325 km), and the altitudes range from ca. 1700 m to 2500 m. Plants of this species are very scarce. In spite of intensive searches by many people, the numbers of specimens observed are low, as noted in the citations of collections above. The plants occur in diverse habitats with two very different extremes—open, sunny meadows and closed, shaded fireweed clones. The Waterton Lakes specimens were taken in a moist drainage area on a grassy slope and those from Storm Lake in meadows only a few feet from the shore. The Marias Pass Population grows in black muck and rotting plant materials under a dense cover. Prominent woody plants growing near Botrychium paradoxum include Pinus contorta, Picea englemannii, Abies lyallit, Potentilla fruticosa, and shrubby species of Salix. The herbaceous vegetation is very different between the open and closed closed sites are dominated by Epilobium angustifolium (Fireweed), Geum macrophyllum, Heracleum lanatum, Osmorhiza occidentalis, and Senecio foetidus, % AMERICAN FERN JOURNAL: VOLUME 71 (1981) as in the habitat near Marias Pass. Weeds are found in small numbers (Taraxacum officinale and Bromus inermis). Our largest specimens of Botrychium paradoxum were found in open sites, where they reached 15 cm in height with stalks 2.0 mm in diameter (Fig. 3A). The heavily shaded plants of B. paradoxum are very delicate, reaching only 7 cm in height with stalks 0.5 mm or less in diameter (Fig. 3B). Some individuals reached only 3 cm in height and had fertile branches only 1-2 mm tall with only | or 2 sporangia each. The latter appeared to be very young juveniles. In the Fireweed habitats, plants are easily found by parting the large herbs and searching the soil below. But in the open meadows, the plants are challenging to find because of close interlacing with other vegetation and the presence of “look-alikes.” The inflorescences of various small grasses, sedges, and plantains may superficially resemble the fertile segments of B. paradoxum. The inflorescences of the tiny knotweed, Polygonum (Bistorta) viviparum are especially troublesome. Botrychium mormo W. H. Wagner. sp. nov. Figs. 1C, 4, 5. Folium 8.6 (7—12.5) cm altum, succulentissimum, flavo-virens, nitidum; segmentum sterile lineare, 2 (1.3-4.1) cm longum, 5 (3-7) mm latum, stipite | (0.51.6) cm longo, lobis 2 (1-3) paribus, acutis vel truncatis, marginibus distalibus integris vel parum crenatis, nec acute dentatis nec irregulariter laceratis; segmentum fertile 4.5 (2.5-7.5) cm altum, plerumque in parte tertia proximali ramosum Gametophyte commonly persisting at the bases of even the largest plants; leaf 8.6 (7-12.5) cm tall, very succulent, yellow-green, shiny; the common stalk making up 50 (20-70)% of the total length; sterile segment linear, 2 (1.3-4.1) cm long, 5 (3-7) mm wide, the stalk 1 (0.5—1.6) cm long; lobes 2 (1-3) pairs, round-pointed to truncate, the distal margins entire or shallowly crenate, not sharply dentate or irregularly lacerate, and with no tendency for exaggerated basal lobes; the tip usually with 2—4 angular triangular or squarish lobes; fertile segment 4.5 (2.4—7.5) cm tall, commonly branched in the lower third, the branches 1/3 to 2/3 as long as the main axis; sporangia large, sunken, not opening until late September and October, the aperture narrow, only 15—30°; spores 49 (45-53) wm in diameter; chromosomes 2n= 90. TYPE: Rich woods on W side of route 9, 8.15 mi N of Bagley City Park, Clearwater Co., Minnesota, Wagner 79314 (MICH). OTHER COLLECTIONS: MICHIGAN: Alger Co.: Grand Sable Lake, Hagenah 2550 (BLH). Gorge at Sable Falls, near Grand Marais, Hall & Hagenah 83] (BLH). Cheboygan Co.: Witkout definite locality, Hagenah 25/1] (BLH). 0.5 mi WSW of Riggsville Corners, Wagner 8062 (BLH). Chippewa Co.: E of Trout Lake, Hagenah 3480 (BLH). Dickinson Co.: 3 mi N of Norway, Hagenah 2920 (BLH). Luce Co.: N of Hendriks Quarry, Hagenah 3205 (BLH). Otsego Co.: Without definite locality, Hagenah 2717 (BLH). MINNE- SOTA: Becker Co.: N side of Route 143, NE of Twin Island Lake, Wagner 79327 (MICH). Beltrami Co.: S and N of Route 1, ca. 1.5 mi E of Clearwater Co. line. Wagner 73196 (MICH). Cass CoN side of Leech Lake, Ottertail Peninsula, Trana 756/5 (MIN). Clearwater Co.; U. of Minn. Biol. Station, Lake Itasca, Bearpaw Point, Wagner 73104 (MICH: | specimen mixed with B. minganense), Rosendahl! 5929 (MIN). Lake Itasca, Garrison Point, Wagner 7329] (MICH). Route 200, ca. 1/3 mi of Wild Rice River, Wagner 73127 (MICH). Mahnomen Co.: E side of county road 4, | mi N of Route 200, Waterway 122 (WIS). WISCONSIN: Ashland Co.: Chippewa Twp.. Sec. 10, Peck 79-590 (UWL), Moran 917 (MIL). Forest Co.: N of Route 8, 2 mi W of Crandon, Wagner in 1951 (MICH). Iron Co.: 3.5 mi NW of Hurley, R. M. Tryon 4023 (WIS). Wood Co.: Arpin, Goess! 3001 (MIL). WAGNER & WAGNER: NEW SPECIES OF MOONWORTS FIG. 4. Botrychium mormo, Mature plants. A. Leaf 8 cm tall. Note low branching of fertile segment. B. Leaf 7 cm tall. Note form of segment divisions. Lake Itasca, MN. FIG. 5. Botrychium mormo. Young plants 1-1.5 cm tall. Note highly reduced sterile and ; N segments at tip of petiole, with only 2-4 sporangia. Lake Itasca, MN. 28 AMERICAN FERN JOURNAL: VOLUME 71 (1981) We have studied this Moonwort for three decades and have determined that it is totally consistent. Among the many curious features of B. mormo, we include the tendency for persistence of the gametophyte even on full-sized sporophytes, the narrow range, the habitat in mesic forest, and the disappearance of the plants during drought periods. Figure 4 shows well developed, mature specimens and Fig. 5 shows the rudimentary, immature specimens. The extremely “chubby” form is nicely illustrated by Peck (1980). This, together with the following species, may constitute a distinct section of subg. Botrychium. Although some small leaves may resemble juvenile or shade specimens of B. simplex E. Hitchc., the fully developed leaves show important differences. Mature individuals have the sterile segment higher on the axis, unlike mature B. simplex (“Blade inserted almost basally or toward the middle of the plant,” Clausen, 1938). The sterile segment has lower divisions approximately equal to those above, but in B. simplex these are normally enlarged in mature individuals, in the fullest development producing a ternate blade. The strongly truncate and adnate lateral lobes of the two new species differ from the rounded lobes and spatulate to lunate, more or less stalked pinnae of B. simplex. (Indeed, many medium and large specimens of B. simplex have middle and upper lateral divisions so flabellate as to resemble B. /unaria). The tip of the sterile segment is usually toothed to deeply cleft, unlike the mainly undivided and entire tip of B. simplex. The usual habitats of B. mormo and B. montanum are shaded forest floors under mature trees, rather than those of B. simplex, which are open, marshy places, meadows, and edges of wet woodland pond shores. On the campus of the Lake Itasca Biological Station in Minnesota, sizeable populations of both B. mormo and B. simplex are found within a quarter of a mile of each other. Class studies have revealed over twenty differences between these species. Botrychium mormo is very difficult to find and apparently very rare. It if does occur in areas other than Michigan, Wisconsin, and Minnesota, it has been overlooked. Our students have made extensive searches for the plants in many localities. We estimate that only one in fifty seemingly suitable sites have yielded specimens. The plant grows in rich leaf mold in Basswood (Tilia americana)—Sugar Maple (Acer saccharum) forests. East of Marquette, Michigan, these dominants are joined by Beech (Fagus grandifolia). The little plants push their way through the leaf litter or simply lie under the litter, failing to appear at all. The goblinish appearance and behavior of this odd plant has inspired the English name of Little Goblin and the scientific epithet mormo. Often very small plants less than 1.5 cm tall (Fig. 5) will dominate a population. In drought years, even large plants may fail to send up a leaf, or one has to scrape off the litter to find any plants, these whitish and apparently lacking chlorophyll. It is not clear why as many as half the individuals, including specimens of all sizes, retain gametophytes at their bases. These are readily detected as swollen, brown masses 1-6 mm long, protruding among the roots. The grapeferns that most commonly grow with or near B. mormo are B. virginianum and B. minganense, and more rarely B. lunaria, B. lanceolatum, B. WAGNER & WAGNER: NEW SPECIES OF MOONWORTS 29 matricariifolium, B. dissectum, and B. multifidum. None of these other species shows the peculiarities of B. mormo such as the persistent gametophytes, extremely succulent texture, and peculiar shiny yellow color. We conclude, therefore, that the habitat is not responsible for its distinctive characteristics. They are evidently genetically fixed. Botrychium montanum W.H. Wagner, sp. nov. Figs. 1D, 6. Folium 8.7 (4-12.5) cm altum, herbaceum, glaucum, hebes; segmentum sterile FIG. 6. Botrychium montanum from the type locality at Swan Valley, MT. Largest plants 8 cm tall. Note Shape of divisions of sterile segments. Leaf 8.7 (412.5) cm tall, herbaceous, glaucous, dull; the common stalk making up 60 (40-90) per cent of the total length; sterile segment oblong to linear, 2.1 (0.7-4.0) cm long, 5.5 (2-9) mm wide, the stalk 0.7 (0.3-1.5) cm long: lobes 3 (1-6) pairs, irregular, narrow and pointed, oblong, square, often grouped or confluent, the distal margins irregularly toothed to lacerate, frequently with ew teeth, basal lobes lacking tendency for enlargement, the blade tip usually with 2- 30 AMERICAN FERN JOURNAL: VOLUME 71 (1981) TYPE: Crane Mountain Rd. (Route 498), 3.6 mi S of junction with Ferndale Road, Swan Valley, Lake Co, Montana, cedar swamp forest, Wagner 80/10 (MICH). OTHER COLLECTIONS: MONTANA: Flathead Co.: Glacier National Park, Johns Lake, Wagner 78510 (MICH). S end of Lake McDonald, N of Apgar, Wagner 78522 (MICH). Avalanche Creek, Wagner 78532 (MICH). Lake Co.: Station Creek, near Flathead Lake, Wagner 78534 (MICH). W side of Cedar Bay Road, Wagner 80114 (MICH). Soup Creek, | mi E of Route 83, Wagner 80121 (MICH). Botrychium montanum differs from B. mormo in characters that are especially evident in living plants (Fig. 6) The leaf is herbaceous (not succulent), glaucous (not eee green), and dull (not shiny). The position of the sterile segment along the vertical axis tends to be higher, and the fertile segment only rarely branches near the base. The cutting of the sterile segment is most distinctive: the lobes are more deeply cut and irregular in pattern, with some narrow and some broad (due to grouping or confluence), and most have sinuses of various sizes. Sporangial dehiscence takes place a couple of months earlier than in B. mormo, and the sporangial valves open wider. We were unable to 88 persistent gametophytes in.B. montanum as we did in B. mormo rychium montanum is relatively easy to find, and we are surprised at the paucity of pone collections. A single locality may have hundreds of plants in a small area (Fig. 6). It is most abundant in moist, springy Western Red Cedar (Thuja plicata) forests. The species may grow in habitats which are quite different, however, as along grassy trail edges (Logan Pass and Many Glaciers, Glacier National Park). Although all of the known localities for B. mormo are at less than 1500 ft (450 m) altitude, B. montanum grows at altitudes from 3200 ft (970 m) at Swan Lake to 6000 ft (1800 m) at Logan Pass. e fern may occur in pure stands, but generally it is associated with B. virginianum and more rarely with B. lunaria, B. lanceolatum, B. boreale, and B. minganense. We acknowledge the support of NSF Grant DEB 800 55 36, “Evolution and Systematics of the Grapefern genus Botrychium.” William R. Anderson kindly translated the Latin diagnoses. A large part of the information we have about the species described here was obtained by our students in Pteridology courses at the Lake Itasca Biological Station, Minnesota, and the Flathead Lake Biological Station, Montana. LITERATURE CITED BEEKS, R. M. 1955. Improvements in the squash technique for plant eigen Aliso 3:131-133. CAMP, W. H. 1940. Aphyllous forms in Pyrola. Bull. Torrey Bot. Club 67:453—465. CLAUSEN, R. T. 1938. A Monograph of the Ophioglossaceae. Mem ater a Club 19(2):1-177. LOVE, A., D. LOVE, and B. M. KAPOOR. 1971. Cytotaxonomy of a century of Rocky Mountain orophytes. Arctic and Alpine Res. 3:139-165. PECK, J. H. 1980. Discovery of the Goblin Fern in Wisconsin. Bull. Bot. Club Wisconsin 12:2-4 + cover illustration. SAHASHI, N. 1980. Morphological and taxonomical studies on Ophioglossales in Japan and the se regions (4). Comparative morphology of spores of some species in Ophioglossales. J. Jap. Bot. 55:73-80. TAGAWA, M. ne Ophioglossum kawamurae Tagawa, a new species from Japan. Acta Phytotax. Geobot. 8:134—136. WAGNER, W. H., Jr. 1960. Evergreen grapeferns and the ie of infraspecific categories as used in North American pteridophytes. Amer. Fern J. 50:32—4 , and L. P. LORD. 1956. The morphological and cytological distinctness of Botrychium minganense and B. lunaria in Michigan. Bull. Torrey Bot. Club 83:261—280. _~ AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 1 (1981) 31 SHORTER NOTE RANGE EXTENSIONS FOR TWO LYCOPODS ON BARANOF ISLAND, SOUTHEASTERN ALASKA.—Early in 1976, I discovered a colony of Tree Clubmoss (Lycopodium dendroideum Michx.) at Thimbleberry Bay, about 5 km southeast of Sitka, Alaska (57°02’N, 135°16’W). Sitka is located on the outer coast of Baranof Island, one of the larger islands of southeastern Alaska. The small, vigorous colony of about five plants grows on a moss-covered rock bench about | m above the extreme high tide line. This collection re-extends the range of the species in southeastern Alaska about 200 km to the northwest. It is the first authenticated report of L. dendroideum from the large outer islands. The voucher collected in 1979 (Muller 2854) is in the herbarium at the University of Alaska Museum, Fairbanks, Alaska (ALA). During the early 1840’s Eduard Blashke, a physician and botanist stationed in Sitka, found L. dendroideum. Franz Joseph Ruprecht, curator of the herbarium at St. Petersburg, Russia, published Blashke’s sighting. However, in 1941, Eric Hulten expressed doubts about the century-old report of L. dendroideum at Sitka (Hulten, E. 1941. Flora of Alaska and Yukon, |. Lunds Univ. Arsskr. N. F., avd. 2, vol. 37. pp. 69-70). The species occurs at widely scattered localities in the main part of Alaska and is known to grow in the Alaska panhandle (southeastern Alaska) south and east from Wrangell. A previously unreported collection of L. dendroideum was made in June of 1973 at Thomas Bay east of Petersburg, a northward range extension of 65 km within the Alaska panhandle. The specimen (Robuck 1403) is in the herbarium of the U.S.D.A. Forest Service, Forestry Sciences Laboratory, Juneau, Alaska. In October 1980, several colonies of the Bog Clubmoss (Lycopodium inundatum L.) were found about 25 km south of Sitka, 1/3 km northwest of Big Bay at the southeastern end of an unnamed lake system (56°49°N, 135°21’W). These plants were growing on a thin muskeg mat overlying granitic rock at the edge of a shallow lake. Vouchers (Muller 4181) are at ALA and the University of Washington herbarium (WTU). In Alaska, the plant is known to occur along lake shores and in muskeg areas in the vicinity of Ketchikan and Wrangell. The Big Bay collection is a range extension of about 200 km to the northwest. This is the first time L. inundatum has been reported from the large outer islands of the Alaska panhandle. Since southeastern Alaska is sparsely settled and travel is difficult and expensive, a lack of botanical information is not surprising. More range extensions can be expected as botanists investigate southeastern Alaska more thoroughly.—Mary Clay Muller, Chatham Area, Tongass National Forest, P.O. Box 1980, Sitka, AK 99835. AMERICAN FERN JOURNAL: VOLUME 71 (1981) PTERIDOPHYTES FOR THE FLORA MESOAMERICANA The Flora Mesoamericana Project, run by a consortium of museums, aims at producing a comprehensive vascular plant flora covering southern Mexico (Chiapas, Tabasco, Quintana Roo, and Yucatan) south to the Panama/Colombia_ border. Although pteridophytes initially were to be excluded, a decision was made in November 1980 to include them. The volume on Pteridophyta, scheduled to appear around 1986, will Se edited by Ramén Riba and Luis Diego Gémez. Pteridologists interested in contributing generic or family treatments are urged to contact the editors for general information. Please write to Flora Day eee Museo Nacional de Costa Rica, Apartado 749, San José, Costa Rica, STATEMENT OF OWNERSHIP, MANAGEMENT AND CIRCULATION equired by 39 U.S.C. ane B 2. DATE OF FILING aaarices pata pany Oct 1980 la. ANNUAL SUBSCRIPTION ANNUALLY quarterly | 4 pate. $9. subs U. S. Nat'l. Herbarium, 10th & Constitution Ave. N.W., Washington, DC 20560 5. LOCATION OF THE HEADQUARTERS OR GENERAL BUSINESS OFFICES OF THE sapere (Not printers) ma S. Nat'l. Herbarium, W., Washi NO CO OITOR, AND seh & Constitution Ave. N. ington, DC 20560 ISHER, E EDIT! PUBLISHER (Ni AMERICAN yen society, INC., (see 5.) Dr. David B. Lellinger, (see 5.) 7, OWNER (If owned by a corporation, its name and address must be stated and also immediately thereunder the names and addresses of stock- holders owning or holding I percent or more of total amount of stock. If not owned by a corporation, the names and addresses of the individual owners must be given. If owned by a partnership or other unincorporated firm, its name and — as well as that of each individual must be giving. If the publication is published by a nonprofit organization, its name ai and address must be ‘d.) NAME | ADDRESS | AMERICAN FERN SOCIETY, INC (see 5.) i j & KNOWN gsc kiss MORTGAGEES, AND OTHER SECURITY HOLDERS OWNING OR HOLDING 1 oa OR MORE OF L AMOUNT OF BONDS, MORTGAGES OR OTHER SECURITIES (If there are none, so stat. ee SEG RET ID SE + none i | AT SPECIAL HAVE CHANGED OURING Uf changed, publiahe f ch PRECEDING 12 MONT with this statement.) $$ $$ —$$—$$_$_$____—_ AVERAGE NO. COPIES EACH | ACTUAL NO. COPIES OF SINGLE ISSUE DURING PRECEDING ISSUE PUBLISHED NEAREST TO 12 MONTHS FILING DATE a. 1525 1600 B. PAID CIRCULATION Le ee RM yeeotngt ND COUNTER SALES 0 i) 2. MAIL SUBSCRIPTIONS 1166 1179 <. 1166 1179 D. FREE DISTRIBUTION BY MAIL, CARRIER OR OTHER MEANS SAMPLES, COMPLIMENTARY, AND OTHER FREE COPIES 1 1 bee Ces ion wowire so 1167 1180 , cones NOT DISTRIBUTED cores se) ER PR! 358 420 2 a eee. 2 0 Y) So iasps cs S E badd 1525 1600 1 i dca a a . BPSLISHER, BUSINESS y MANAGER. OR OWNER MS fee by 4 1 id David B. Lellinger, Editor ————————— | 2.121, Postal Service Manual) BRITISH PTERIDOLOGICAL SOCIETY Open to all who are interested in growing and studying ferns and fern-allies. Full members receive THE FERN GAZETTE and the BULLETIN OF THE BRITISH PTERIDOLOGICAL SOCIETY. Membership subscriptions are £ 7 for full members, £ 5 for ordinary members (not receiving the GAZETTE), and £5 for student members (under 25 years of age). For particulars, U. S. residents should apply to Dr. J. Skog, Biology Department, George Mason University, Fairfax, VA 22030. Non-U. S. residents should apply to Lt. Col. P. G. Coke, Robin Hill, Stinchcombe, Dursley, Gloucestershire, England. TRIARCH Over 5© Wears of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 ee aoe reorient AMERICAN = FERN = JOURNAL es QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Azolla filiculoides New to the Southeastern United States VERNON M. BATES, JR. and EDWARD T. BROWNE, JR. The Genus Nephrolepis in Florida CLIFTON E. NAUMAN The Branching Pattern of Hypolepis repens THERESA M. GRUBER Diplazium japonicum and Selaginella uncinata Newly Discovered in Georgia WAYNE R. FAIRCLOTH Notes on Selaginella, with a New Variety of S. pallescens ROBERT G. STOLZE Lepisorus kashyapii in the Western Himalayas S. S. BIR and CHANDER K. SATIJA Taxonomic Notes on Jamaican Ferns-IIl GEORGE R. PROCTOR Shorter Notes: Notes on North American Lower Vascular Plants—II; Equisetum arvense in ama mesoun BoTANCAT ms Wil 13 rR GARDEN LIBRARY. wa _ The American Fern Society Council for 1981 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, OH 45701. President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN 37235. Vice President MICHAEL I. COUSENS, Faculty of Biology, University of West Florida, Pensacola, FL 32504. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, VA 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, DC 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Berkeley, CA 94720. Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, NY 10458. Newsletter Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, DC 20560. ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington, IN 47401. JOHN T. MICKEL New York Botanical Garden, Bronx, NY 10458. The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and published at the Smithsonian Institution, Washington, DC 20560. Second-class postage paid at Washington. Claims for missing issues, made 6 months (domestic) to 12 months (foreign) after the date of issue, and the matters for publication should be addressed to the Editor. Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog, Dept. of Biology, George Mason University, Fairfax, VA 22030. Orders for back issues should be addressed to the Treasurer. General inquiries concerning ferns should be addressed to the Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Fern Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or less, $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is editor of the newsletter “Fiddlehead Forum.” The editor welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th Street, Seattle, WA 98115, is Director. Spores exchanged and collection lists sent on request. Gifts and Bequests Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Botanical books, back issues of the Journal, and cash or other gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secretary. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 33 Azolla filiculoides New to the Southeastern United States! VERNON M. BATES, JR.* and EDWARD T. BROWNE, JR.** When studying specimens of Azolla, one is often confronted with a large number of sterile collections that cannot be identified with confidence. Although Svenson (1944), in his monograph of the New World species of Azolla, based his classifica- tion on characteristics of the infrequently occurring sporocarps and paid little attention to vegetative characters, we believe that megaspore ornamentation pro- vides the only suitable means of identification. Through SEM examination of its megaspores, we have confirmed the identifica- tion of a collection of A. filiculoides Lam. from Georgia. The collection was made 15 April 1956 from plants floating in a large fresh-water lake at the north end of Sapelo Isaland, McIntosh Co., Georgia (Wilbur H. Duncan 1 9969, with William P. Adams and Clyde Connell, TEX). Apparently this is a new record for the southeastern United States. We found that the megaspore ornamentation agreed very closely with Svenson’s (1944) line drawings and description; the “raised, hexagonal markings” reported by Svenson actually are localized masses of fused excrescences which form a flattened to slightly convex reticulum (Fig. /). Although there has been no previous ultrastructural study of Azolla megaspores from North America, recent descriptions of European material support our identification (Pieterse, de Lange & van Vliet, 1977: Martin, 1978). Traditionally, Azolla in the southeastern United States has been thought to be A. caroliniana Willd. Megaspores apparently never have been described for this species in North America (Svenson, 1944; Correll & Correll, 1975). Di Fulvio (1961) illustrated a cross-section of the perispore of an A. caroliniana megaspore, but the source of her material and the voucher documenting it are not stated. Unfortunately, so little is known about the cytology of Azolla that it cannot be said whether or not A. caroliniana is a sterile hybrid; it has been reported as 2n=48 in Europe (Love, Love & Pichi Sermolli, 1977, p. 373). Finding consistently sterile material over the entire range of a species does, however, raise the possibility that the material is hybrid or otherwise incapable of forming spores, perhaps as a result of aneuploidy. Svenson reported the distribution of A. filiculoides as widely scattered along the Pacific Coast from Alaska southward into Washington, Oregon, California, and Mexico. Although he believed that this species was naturalized in the New England states, he never attempted to explain two New York collections of A. filiculoides which may not have been the result of artificial introductions: Riverhead, Long Island, Suffolk Co., 20 Aug 1938, Muenscher & Curtiss 6647 (US) and Oak Orchard, Orleans Co., Aug 1876, Herb. N. L. Britton 5. n. (NY). If, as some have suggested (Smith, 1955), Azolla is disseminated by adhering to the feet of migratory waterfowl, it seems strange that A. caroliniana and A. * Data Communications Corp., 3000 Directors Row, Memphis, TN 38131. ** Dept. of Biology, Memphis State University, Memphis, TN 38152. E Portion of a Master of Science thesis submitted to the Department of sity, 1980. Volume 71, number 1, of the JOURNAL was issued March 30, 1981. Biology, Memphis State Univer- 34 AMERICAN FERN JOURNAL: VOLUME 71 (1981) filiculoides would be limited to the Atlantic and Pacific coasts, respectively. In light of the large number of sterile collections of Azolla from the eastern United States which cannot be identified with certainty, it seems quite possible that A. filiculoides is more widely distributed in the eastern United States than has been thought. Our discovery of A. filiculoides in Georgia, considered with those specimens already known from New York, supports the hypothesis that this species may be widely distributed along the Atlantic coast. Unfortunately, this cannot be confirmed until more fertile collections are observed or until vegetative differences can be found to distinguish the various species of Azolla. . 1. Scanning electron micrograph showing characteristic features of the megaspore of Azolla filiculoides. Proximal to the girdle (g), elevated masses of excrescences form a distinct reticulum (r). We wish to express our appreciation to Dr. Lewis B. Coons and Mrs. Naomi Roberts for their kind assistance in obtaining the micrograph. We are also especially indebted to Dr. Billie L. Turner, curator of the Herbarium, University of Texas at Austin, who loaned us collections of Azolla to examine. In addition, we are very grateful to Dr. W. Carl Taylor, curator of the Vascular Herbarium, Milwaukee Public Museum, for reading the manuscript and making valuable suggestions. LITERATURE CITED CORRELL, D. S. and H. B. CORRELL. 1972. Aquatic and Wetland Plants of the Southwestern United States. Environmental Protection Agency, Washington, DC. : DI FULVIO, T. M. 1961. Sobre el episporio de las especies Americanas de Azolla con especial referencia a A. mexicana Presl. Kurtziana 1:299-302. OVE, A, D. LOVE, and R. E. G. PICHI SERMOLLI. 1977. Cytotaxonomical Atlas of the Pteridophyta. J. Cramer, Vaduz. MARTIN, A. R. H. 1976. Some structures in Azolla megaspores and an anomalous form. Rev. Paleobot. Palynol. 21:141-169. PIETERSE, A. H., L. de LANGE, and J. P. van VLIET. 1977. A comparative study of Azolla in the Netherlands. Acta Bot. Neerl. 26:433—449. SMITH, G. M. 1955. Cryptogamic Botany, vol. II. Bryophytes and Pteridophytes, ed. 2. McGraw-Hill, New York. SVENSON, H. K. 1944. The new world species of Azolla. Amer. Fern J. 34:69-84. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 35 The Genus Nephrolepis in Florida CLIFTON E. NAUMAN* The recent addition of N. multiflora (Roxb.) Jarrett ex Morton to the Florida flora (Gillis & Proctor, 1975) and the description of a new hybrid species (Nauman, 1979a) has created a need for a revised treatment of the genus in Florida. This treatment is based on a study of more than 120 morphological characters involving more than 1000 specimens. A transplant experiment and ecological observations (Nauman, 1979b) also aided in interpreting morphological variability. his study was based in part on a Masters thesis, and I would like to express my appreciation to D. F. Austin (Florida Atlantic University) for his guidance during the course of study and for advice and assistance in the gathering of data. Appreciation is also given to the curators and staff of the following herbaria for the loan o materials and types: B, BM, BR, F, FAU, FLAS, FSU, FTG, GH, K, MA, NY, P, PENN, PH, S, TENN, UC, US, USF, UWFP. KEY TO THE FLORIDA SPECIES OF NEPHROLEPIS 1. Adaxial costa surface of poe pinnae glabrous (sometimes with a few scales); indusium reniform, hippocrepiform, or lunate, .2 mm or more wide. 2. Pinnae falcate with acute to attenuate tips; plants never bearing tubers; rachis ere appearing concolorous or obscurely bicolorous . exaltata 2. Pinnae not falcate or slightly so with blunt tips; plants sometimes bearing oe "chs scales distinctly bicolorous (pale with a dark point of attachment) 2. N. cordifolia 1. Adaxial costa surface of medial pinnae sparsely to ad covered with short erect trichomes (often also with scales); indusium orbicular, ca. 1.0 mm Basal portions of mature stipes covered with jeg brown, appressed scales with a margins. 4. N. multiflora 3. Basal portions of mature stipes not covered with dark brown, appressed scales, but often with a few loose, reddish to light brown scales or the scales absent. 4. Adaxial costa surface sparsely = the trichomes 0.4 mm long; pinnae ered falcate; blade length/width ratio 6.8-17.9, mean 9.5 5.N. X averyi 4. Adaxial costa surface densely sinocen to tomentose (rarely glabrous), trichomes 0.3 mm long; pinnae not falcate or slightly so; blade length/width ratio 3.8-7.8, mean 5.1. 1. N. biserrata 1. Nephrolepis biserrata (Swartz) Schott, Gen. Fil. text to pl. 3. 1834. Aspidium biserratum Swartz, J. Bot. Schrad. 1800(2):32. 1802. TYPE: Mauritius. Groendal, (S-Hb. wartz !). Nephrodium biserratum (Swartz) Presl, Rel. Haenk. 1:31. 1825. Hypopeltis biserrata (Swartz) Bory in Bél. Voy. Indes. a Bot. 2:65. 1833. Lepidoneuron biserratum (Swartz) Fée, Gen. Fil. 301. ; Nephrolepis exaltata var. biserrata (Swartz) Baker in sii Fl. Brasil. a 493. 1870. Nephrolepis hirsutula 8 biserrata (Swartz) Kuntze, Rev. Gen. Pl. 2:816. _ Aspidium acutum Schkuhr, Krypt. Gew. 32, pl. 31. 1806. TYPE: =ge wal collector not state (Hb. Breyne presumably destroyed). Nephrodium acutum (Schkuhr) Presl, Rel. Haenk. 1:31. 1825. Nephrolepis acuta (Schkuhr) Presl, Tent. Pterid. 79. 1836. _Nephrolepis hirsutula « acuta (Schkuhr) Kuntze, Rev. Gen. Pl. 2:816. 1891. “Department of Biological Sciences, Florida Atlantic University, Boca Raton, FL 33431. AMERICAN FERN JOURNAL: VOLUME 71 (1981) N. biserrata N. cordifolia N. exaltata N. multiflora 200 miles N.x% averyi ~ FIG. 1. County distributions of the species of Nephrolepis in Florida. C. £. NAUMAN: NEPHROLEPIS IN FLORIDA 37 DISTRIBUTION AND HABITAT: This is a species of swamps and wet ham- mocks in which it is usually terrestrial, but may be epiphytic or epipetric. The general distribution is pantropical; in Florida it is largely restricted to Florida’s tropical fringe (Fig. /). Distribution in Florida appears to be limited by the plant’s ability to tolerate cold temperatures as revealed by transplant experiments. Nephrolepis biserrata is best distinguished from the other Florida Nephrolepis species by its size, hyaline margined scales on the rhizome and croziers, pubescent pinnae, and orbicular indusia. The species is closest to N. multiflora in indumentum and sori, to N. X averyi in size, and to N. cordifolia in spores. REPRESENTATIVE SPECIMENS: Broward Co.: Durand 58, 64, 72 (FAU); Leeds 342 (NY, PH); Moldenke 483 (NY, PH, US); Nauman & Nauman 380 (FAU); R. P. St. John 1751 (FLAS). Collier Co.: Correll 6081 (FLAS), Diddle 698 (FLAS); Evans (TENN); Lakela et al. 27996 (USF); Nauman & Austin 555 (FAU). Dade Co.: Avery & Loope 1918 (FAU); Britton 428 (F, NY); Buswell (FAU); Carter 204 (PH); Correll 5897 (GH, US); Eaton (F); Long et al. 1941 (NY, USF); Munroe (GH, NY, UC); Safford & Mosier 105 (US); Small & Mosier 5890 (NY, US). Highlands Co.: Garrett 60 (FLAS); McFarlin 10714 (GH). Monroe Co.: Delchamps & Wherry (PH); Lakela & Almeda 30526 (USF). Palm Beach Co.: Durand 25, 80 (FAV); Hill 152 (NY); McJunkin (FAV). 2. Nephrolepis cordifolia (L.) Presl, Tent. Pterid. 79. 1836. Polypodium cordifolium L., Sp. Pl. 2:1089. 1753. TYPE: Petiver, Pterigr. Amer. f. /, f. //. 1712 ('). The Petiver plate is supposedly a copy of Plumier, Tract. Fil. Amer. 1. 7/. 1705 (!). The Plumier plate is a poor drawing of what might be a species of Nephrolepis. Until plants can be examined from the area where Plumier obtained his material, application of the epithet N. cordifolia will be problematical. Aspidium cordifolium (L.) Swartz, J. Bot. Schrad. 1800(2):32. 1802. Aspidium tuberosum Bory ex Willd., Sp. Pl. ed. 4, 5:234. 1810. TYPE: “Bourbon sur les arbres, No. 111, Bory de St. Vincent” (B-Hb. Willd. 19759 photo FAU !, photo GH !; isotypes P photo FAU !, FI, not seen). Nephrodium tuberosum (Bory ex Willd.) Desv., Mém. Soc. Linn., Paris 6:252. 1827. Nephrolepis tuberosa (Bory ex Willd.) Presl, Tent. Pterid. 79. 1836. Nephrolepis cordifolia var. tuberosa (Bory ex Willd.) Baker in Mart. Fl. Brasil. (1)2:491. 1870. Nephrolepis exaltata B tuberosa (Bory ex Willd.) Kuntze, Rev. Gen. Pl. 2:816. 1891 . DISTRIBUTION AND HABITAT: Whether or not this species is native to Florida is uncertain. Wherry (1964) considered it possibly native to the southern- most portions of the State. Though widespread, the plants are almost entirely persistent from cultivation in dumps and at abandoned homesites. I know of only one site where the plants may have colonized without the help of man. This is on the Sturrock estate in West Palm Beach. The plants were reported to have been blown in by a hurricane in the late 1940’s (Sturrock, pers. comm.). This species’ distribution is scattered in Florida and doesn’t conform to any apparent trends in temperature extremes (Fig. ]). The general distribution is possibly worldwide in the tropics and subtropics, also in Japan and New Zealand. Tubers are the most distinctive feature of this species, although tuberless colonies are frequent throughout Florida. Whether tuber production, or the lack of it, is controlled by environmental or genetic factors is at present uncertain. There appears to be a correlation between the substrate in which the plant is growing and tuber production. In the Florida populations, tuber production seems restricted to plants growing in humus and has not been seen in epiphytic plants or plants growing in 38 AMERICAN FERN JOURNAL: VOLUME 71 (1981) drier sites. These observations suggest some specific soil and moisture requirements for tuber production. The distinctly bicolorous rachis scales are as diagnostic as tuber- production and may be used to distinguish this species from any of the other Florida species, even in the absence of other key features.. REPRESENTATIVE SPECIMENS: Brevard Co.: Hollister (US); Shuey M1084 (USF). Broward Co.: McCart & Snyder 9523 (FAV). Citrus Co.: R. P. St. John 197 (FLAS). Dade Co.: Evans (TENN); Hardy 5 (PH). Duval Co.: Darling (US). Escambia Co.: Burkhalter 5919 (UWFP). Hernando Co.: Cooley et al. 8307 (GH, USF); Mickel et al. 1782 (UC); Moldenke & Moldenke 29488 (US). Highlands Co.: McFarlin 8957 (FLAS). Hillsborough Co.: Evans 2303 (TENN); Mickel et al. 1789 (US); Scudder 441 (FAU). Leon Co.: Hume (FLAS). Marion Co.: Mickel et al. 1742 (UC). Martin Co.: Austin et al. 6469 (FAU); Orange Co.: Githens 2579 (PH); Medgser (UC). Palm Beach Co.: Austin (FAU); Baker (FAU); Cassen 89 (USF); Hill 149 (NY); Nauman et al. 218, 410 (FAU); Wilhelm 121 (FAU). Pasco Co.: Carpenter (GH). Pinellas Co.: Genelle & Fleming 2508 (USF); Scudder 436 (FAU). Polk Co.: Cooley 11783 (USF); 3. Nephrolepis exaltata (L.) Schott, Gen. Fil. text to pl. 3. 1834. Polypodium exaltatum L. Syst. Nat. ed. 10(2):1326. 1759. TYPE: Sloane, Jam Voy. 1. 31, 1107, which is based on a specimen communicated to Sloane by Dr. Sherard from Jamaica (BM not seen, photo FAU). Aspidium exaltatum (L.) Swartz, J. Bot. Schrad. 1800(2):32. 1802. Nephrodium exaltatum (L.) R. Brown, Prodr. Fl. Nov. Holl. 1:148. 1810. DISTRIBUTION AND HABITAT: This is the most common of the Florida species. It is found in a wide variety of habitats, such as tropical hammocks, . low hammocks, and swamps. Frequently N. exaltata is found as an epiphyte on Sabal palmetto or species of Quercus, but it also may occur epipetrically or terrestrially. In Florida, N. exaltata is found from Dade and Monroe to Duval Counties (Fig. /). Occasionally it occurs farther north, but only in cultivation. The plants are most common south of Lake Okeechobee, becoming occasional to rare northward. The general distribution has been traditionally construed as pantropical, but studies by Proctor (1977) and Stolze (pers. comm.) have implied that N. exaltata may have a more restricted range than previously thought. Examination of herbarium specimens shows that a large proportion of plants identified as N. exaltata is actually N. multiflora, N. biserrata, N. cordifolia, or N. rivularis (Vahl) Mettenius. The misapplication of the epithet exaltata is so widespread that a thorough monographic study will be necessary to determine the actual range of this species. Authentic x phar are known from Florida, Jamaica (the type locality), the Antilles, 10802 (NY). Broward Co.: Durand 15, 73 (FAU); Hopkins 31 (FAU): Janda 29 (FAU); Nauman 826 (FAU). Charlotte Co.: Ward A-69 (FLAS). Citrus Co.: E. P. St. John (FLAS). Collier Co.: Austin & Austin 6669 (FAU); Austin el al. 6763 (FAU); Clewell 286 (FSU); Cooley 786 (USF); Evans (TENN); Eyles & Eyles 8235 (GH); Hitchcock (F); Lakela & Almeda 2994] (USF), Long et al. 2393 (USF); Nauman et al. 328 (FAU); Sturtevant 44 (FLAS, US). Dade Co.: Garber (F, NY); Hill 3079 (FTG); Moldenke 436 (NY); Small 7384 (NY, TENN); Tracy 9135 (F, NY, PENN, US). Duval Co.: Calkine (F). Glades Co.: Ward 1-10 (FLAS). Hardee Co.: Kirk (FLAS). Hendry Co.: Jennings (USF); Ward C. E. NAUMAN: NEPHROLEPIS IN FLORIDA 39 et al. 2396 (FLAS, FSU). Hernando Co.: E. P. St. John (FLAS). Highlands Co.: Evans 2292 (TENN); Lakela 26793 (USF); Mickel et al. 1815 (FSU, UC); Nauman 90 (FAU); Porter & Porter 10618 (NY, UC); Wilbur & Webster 2610 (GH, NY, US). Hillsborough Co.: Mickel et al. 1789. (UC). Indian River Co.: Small 8849 (NY). Lake Co.: Nash 1288 (F, GH, NY, PH, UC, US); Underwood (F, GH). Co.: Brumbach 5348 (FAU, FLAS); Correll 5919 (GH); Hitchcock 542 (F, GH. US): Standley 124 (F, GH, NY). Leon Co.: Jackson (FSU). Martin Co.: Nauman 527 (FAU); Nauman & Tatje 520 (FAU); Popenoe & Popenoe 688 (FTG). Monroe Co.: Long et al. 1703 (USF). Okeechobee Co.: McCart 10287 (FAU). Osceola Co.: Mearns 3] (US). Palm Beach Co.: Austin 6636 (FAU): Cooley et al. 4877 (GH, USF); Durand 77, 81 (FAU); Hitchcock 2402 (F); Kral 5680 (FSU); Meagher 895 (FTG); Nauman et al. 702. (FAU); Randolph 140 (GH); Underwood 2219 (NY). Pasco Co.: Underwood 1931 (NY). Pinellas Co.: Curtiss 3764 (F, NY, UC); Genelle & Fleming 1640 (USF); Ralphs 750 (F); Scudder 104 (FAU); Thorne 10315 (UC). Polk Co.: Jennings & Jennings (USF); Smith (US); Wherry (PH). Saint Lucie Co.: Austin et al. 6454 (FAU); Leeds 344 (PH); Small & Matthaus 9640 (NY). Sarasota Co.: Smith (PH). Seminole Co.: Lambert 18 (PH); Nauman et al. 293 (FAV). Sumter Co.: E. P. St. John (FLAS, NY); Smith (US). Suwannee Co.: Leonard 6667 (FSU). 4. Nephrolepis multiflora (Roxb.) Jarrett ex Morton, Contrib. U. S. Natl. . 1974. Davallia multiflora Roxb., Calcutta J. Nat. Hist. 4:515, t. xxxi left hand. 1844. LECTOTYPE: India, Roxburgh (BR not found, fide Lawalrée in litt., fragment US!), chosen by Morton. DISTRIBUTION AND HABITAT: Occasional in disturbed sites, usually near canals and other bodies of water in loose, well drained soil, frequently in full sun. Scattered in southern Florida, but reaching as far north as Pinellas and Hillsborough Counties (Fig. 1). Like N. biserrata, the distribution appears limited by tolerance to cold temperatures. A native of the Old World tropics, N. mulitflora is widely naturalized in the New World. The species was first reported for Dade County, Florida by Gillis and Proctor (1975). Though the data are inconclusive, this species seems to have arrived in Florida in the late 1940’s or 1950’s somewhere in Lee County and is actively spreading. This conclusion is based on the dates of collection of this species throughout the State, but may be biased by the uneven collecting of certain areas. Nephrolepis multiflora was reported by Proctor (1977) to occur in the Bahamas and the Antilles. I have seen specimens from these areas, as well as Central America, Brazil, and Venezuela. REPRESENTATIVE SPECIMENS: : Broward Co.: Avery & McPherson 1327 (USF); Lakela & Long 1598 (USF). Collier Co.: Austin et al. 6767 (FAU); Correll & Popenoe 47290 (FTG); Lakela & Almeda 29988 (GH, USF); Lassiter-et al. 9 (USF). Dade Co.: Avery 1329 (FLAS, USF); Evans (TENN); Gillis 10856 (FTG). Hillsborough Co.: Long et al. 2942 (USF): Scudder 439 (FAU); Shuey (USF). Lee Co.: Austin 6625 (FAU); Brumbach 8743 (NY, US); Cooley 2548 (FLAS, GH, NY, US). Martin Co.: McCart 10406 (FAU, FLAS); Nauman & Tatje 258 (FAU); Nauman et al. 200 (FAU). Monroe Co.: Avery (FLAS). Palm Beach Co.: Brawner (FAU); Durand 5, 85 (FAU); Nauman 312 (FAU); Nauman & Austin 181 (FAU); Nauman et al. ). 5. Nephrolepis x averyi Nauman, Amer. Fern J. 69:69. 1979. TYPE: Fakahatchee Strand off West Grade, 50 ft E of Indian Mound Slough Bridge, Collier Co., Florida, 29 Jan 1979, Nauman et al. 635 (US; isotypes FAU, AS, GH, MSC, NY). oe DISTRIBUTION AND HABITAT: Terrestrial, epiphytic, or epipetric in ham- mocks and swamps. Known to occur only with its putative parents, N. biserrata and 40 AMERICAN FERN JOURNAL: VOLUME 71 (1981) N. exaltata, in mixed colonies, in Florida south of Lake Okeechobee to as far north as Pinellas County (Fig. /). Distribution outside Florida is uncertain; one specimen seen from Jamaica. Nephrolepis X averyi is best distinguished from N. biserrata by its falcate pinnae and narrower fronds, and from N. exaltata by its larger size and lightly pubescent adaxial costa surface. PRESENTATIVE SPECIMENS: Co.: Durand 4, 49, 56, 67, 70 (FAU); Nauman 318, 434 (FAU); Nauman et al. 647 (FAV). Collier Co.: Beck 2027 (FSU); Field & Lazon (US); Fischer 8 (US); Howell 855 (US); Nauman et al. 631. (FAU). Dade Co.: Curtiss 5460 (FLAS, UC, US); Eaton 99] (F); Garber (F, FLAS, PH); Hardy 10 (PH); Lakela 31449 (USF), Small 7384 (NY); Tatnall 805 (PH). Manatee Co.: Cuthbert (FLAS). Palm Beach Co.: Stevens (FAU). Pinellas Co.: Bebb (F). Polk Co.: White (FLAS). SPECIES AND FORMS EXCLUDED Several cultivated forms of Nephrolepis are occasionally found in Florida. These forms are usually in cultivation or are persistent from cultivation. The distribution is scattered from Dade to Duval Counties. Forms represented in the herbaria are: N. exaltata cv. ‘Bostoniensis,’ cv. ‘Elegantissima,’ Cv. ‘Florida Ruffles,’ and cv. ‘M. P. ills’; N. falcata f. furcans (Moore in Nicholson) Proctor [=N. biserrata cv. ‘Furcans’]; and N. hirsutula cv. ‘Superba.’ Nephrolepis pectinata (Willd.) Schott was reported by Wherry (1964) for southern Florida. I have seen no specimens of this species in the herbaria or field. It is doubtful that this species exists in Florida. LITERATURE CITED GILLIS, W. T. and G. R. PROCTOR. 1975. Additions and corrections to the Bahama flora—II. Sida 6:52-62 NAUMAN, C. E. 1979a. A new Nephrolepis hybrid from Florida. Amer. Fern J. 69:65—70. _______. 1979b. The genus Nephrolepis in Florida. Unpublished Master's Thesis, Florida Atlantic University. 126 pp. PROCTOR, G. R. 1977. Pteridophyta. In R. A. Howard, A Flora of the Lesser Antilles. Vol. II]. Arnold Arboretum, Harvard University, Jamaica Plain, MA. WHERRY, E. T. 1964. The Southern Fern Guide. Doubleday, Garden City, NY. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 4] The Branching Pattern of Hypolepis repens THERESA M. GRUBER* Hypolepis, a pantropical genus of the Dennstaedtiaceae, is a terrestrial fern with usually large leaves and rather slender, long-creeping stems bearing trichomes and long, fibrous roots. Hypolepis repens (L.) Presl may form large colonies, and a single plant was measured in Puerto Rico with 42 meters of stem (R. and A. Tryon, pers. comm.). The leaves may be up to three meters or more long and often are partially supported by surrounding vegetation. Roots occur along the entire length of the stem, but are especially numerous at the leaf bases. The morphology of Hypolepis repens and other species has been studied by Gwynne-Vaughan (1903), Bower (1923), Troop and Mickel (1968), and Imaichi and Nishida (1973). These treatments are limited, however, to describing the vas- cular system and analyzing single branch units. This paper considers the branching pattern of the stem and its components of entire plants of H. repens and the development and adaptive significance of the pattern. MATERIALS AND METHODS Stems of Hypolepis repens were studied in a Liquidambar cloud forest located about 12 km south of Misantla, Veracruz, Mexico at an altitude of 1400 meters. A large plant (Fig. 7) was excavated and exposed in situ; it was measured, mapped, and portions were taken for dissection. The plant was growing on a 35—45° slope in loose humus in association with herbaceous angiosperms. The stem length totaled 30 m and occupied an area of 13 m?. In clear weather, most of the colony received sunlight between 10:00 AM and 1:00 PM, despite the surrounding tree canopy. A smaller plant (Fig. 2) was collected intact from a nearby wooded area. It was growing in loose humus on a slope of about 30° around a tree base nearly a meter in diameter. The vascular system of the stems was studied by dissection and the anatomy of the branch unit by means of a cinematographic record of the surface between serial slices (Tomlinson, 1971). OBSERVATIONS The pattern of stem growth, with its branches, buds, and leaves is shown in Figs. I and 2. An analysis of these diagrams revealed a surprising regularity of pattern, which is illustrated schematically in Fig. 3. The following regular components of the whole stem and leaf system were noted. (1) Each bifurcation of the stem (Fig. 3, axis 1) produces a leaf (Fig. 3, axis 2) and a continuing stem (Fig. 3, axis 1). Mes The bifurcation may be described as dichotomous since the division of the parent Stele into stem and leaf is equal (Fig. 4). However, this designation may be inappropriate because dichotomous branching usually refers to the production of two *6206 Braidwood Drive, San Antonio, TX 78249. AMERICAN FERN JOURNAL: VOLUME 71 (1981) TR yo Je /! > T/ 7 TR TR uphill downhill i O y @: Dead o: Matur mo ©: Infant l cnaesaine t tuned Oo: Crozier ~~ ud or stem apex ] 1 METER ; TR : Truncat pan Dead leaf roe panel nfant lea cagatains ges we Crozier eoo® Bud or stem apex Truncate Aborted Leaf Decayed but continuous ; at 2 FIG. 1. Diagrammatic sketch of Hypolepis repens, mapped in situ. FIG. 2. Diagrammatic sketch of Hypolepis repens, collected intact and later mapped. 1 METER T. M. GRUBER: BRANCHING PATTERN OF HYPOLEPIS REPENS 43 initially equivalent organs rather than a stem and a leaf and because it is not known whether the bifurcation arises from a division of the apical cell. Sections of the vasculature illustrated in Fig. 4 agree with the observations of Gwynne-Vaughan (1903) and Troop and Mickel (1968), except that the leaf petioles arise on the lateral rather than the upper side of the stem. (2) Leaves are produced successively on alternate sides of the stem. The simplest assumption for descriptive purposes is that the system is monopodial, that is, the leaf is an appendage of the stem axis. (3) The petiole (Fig. 3, axis 2) bears 1-4 buds (Fig. 3, axes 3-6) which may develop into stems (Fig. 3, axes 3, 4) that function like the stem from which their parent leaf originated. An analysis of the number of buds on the petioles showed that of a total of 41 petioles on the plants in Figs. / and 2 in condition to study, two had no buds, five had one bud, ten had two, 22 had three, and two had four buds. The third and fourth buds, when present, were borne in the region of the petiole where it became erect and they grew down toward the ground, perhaps providing support for the leaf. (4) The proximal bud that develops into a branch (Fig. 3, axis 3) is always on the side of the petiole (Fig. 3, axis 2) opposite to the continuing main axis (Fig. 3, axis 1); successive buds that may develop into branches (Fig. 3, axes 4-6) are alternate beyond the proximal bud. It follows that the second bud is always on the side of the petiole nearest to the continuing main axis. (5) When the proximal bud develops into a stem (Fig.-3, axis 3), the symmetry of its resulting next higher-order branch unit (Fig. 3, leaf 7 and petiolar buds 8, 9) is identical to that of the unit (Fig. 3, leaf 2 and petiolar buds 3-6) from which the branch originated. ; (6) When the second bud develops into a stem (Fig. 3, axis 4), the symmetry of its resulting next higher-order branch unit (Fig. 3, leaf 10 and petiolar buds 11, 12) is the mirror image of that unit (Fig. 3, leaf 2 and petiolar buds 3-6) from which the branch originated. ; (7) All stems developing beyond the first branch unit repeat the patterns laid down in (1) through (6). The distance between leaves and the angle between bifurcations of axes were both rather variable. The spacing of the leaves varied widely in each of the plants investigated. The small plant (Fig. 2), which was growing in a heavily shaded locality, had the leaves an average of 19.5 cm apart, with two 28 cm apart. The large plant (Fig. /), which was growing in a somewhat exposed site, had the leaves an average of 33.9 cm apart and two were as much as 68 cm apart. Apparently the distance between leaves is influenced by the immediate environment and is not a fixed characteristic of the branching pattern. : . The angle between bifurcations of the stem and leaf approximates 60 - In a tota of 17 branch units of the small plant (Fig. 2), the angle was 40 in two units, 50° in four, 55° in one, 60° in eight and 90° in two (an average of 59°). The angle between the petiole and its first bud averaged 63°, for the nine that were measurable. AMERICAN FERN JOURNAL: VOLUME 71 (1981) A 666 0000000000000 0000000000000 Bud nh Leaf = Fn Bud \ : : ae mxAy G. 3. The Branch Complex. Branch order is represented by the different patterned axes: main axis, solid; first-order, solid dots; second-order, stripes; third-order, open circles, fourth-order, stippled. The individual axes are numbered according to their ontogenetic appearance. An axis terminated by a circular tip represents a leaf; a triangle represents a stem apex. _ 4. Serial sections of the vasculature of the branching unit of Hypolepis repens, from a cinematographic record. T. M. GRUBER: BRANCHING PATTERN OF HYPOLEPIS REPENS 45 DISCUSSION Ecological Implications. — Bell and Tomlinson (1980) noted the infrequency of adequately detailed accounts of basic branching patterns in rhizomatous systems, especially among the ferns. Ferns are ideal for this kind of study because they have no secondary cambial growth. A fern with a creeping, branching stem, as exemplified by Hypolepis repens, has important advantages over its relatives which have short-creeping or erect stems. An environment such as the Liquidambar cloud forest presents a substantial limitation to gametophytes in the form of competition for open space. An elongate, branching stem provides a means of vegetative propagation which alleviates the necessity for local propagation through sexual reproduction. Interestingly enough, not a single fertile frond was found during several days of mapping and excavating. In addition, a regular method of branching enables the plant to survive destruction of one or more stem apices. When growth along one axis is unsuccessful, that shoot may abort without the loss of the entire plant. The closely related H. punctata Mett. ex Kuhn responds to the destruction of the leaf apical cell by the development of the most proximal bud into a leaf or shoot (traumatic reiteration, Imaichi and Nishida, 1973). Hypolepis repens probably responds in a similar manner, as indicated by the development of the proximal bud in two out of four aborted leaves (Fig. 2). A branched network with bifurcation angles of 60° tends to form hexagons, a pattern which covers a maximum amount of surface area with a minimum path length (Stevens, 1974). This patterned branching enables the plant to make a systematic exploration of the local environment with a minimal energy output. This same pattern enables the plant to reenter territory, over a span of time, which previously proved favorable to growth. A shoot might be expected to be sufficiently plastic to respond to an especially favorable environment. This theory of “adaptive reiteration under supra-optimal conditions” is difficult to prove, as was pointed out by Bell and Tomlinson (1980). The difficulty lies in there being no established basis for quantification. A favorable environment could exist in time, such as especially favorable weather during a growing season, or in space, such as nutrient-rich soil or good light availability. Adaptive reiteration could refer to the production of a maximum amount of photosynthetic material within a limited amount of time or space, exceeding that observed to be average. Hypolepis repens appears to support this theory, as indicated by the inordinate number of croziers within the area delineated by broken lines in Fig. 1. Furthermore, within this same region, there are three examples of advanced development of the second of the second-order branch (as in Fig. 3, axis 4), a relatively rare occurrence in the two plants. These ecological advantages and the predictability of the morphology of H gd repens suggest that this particular pteridophyte has genetically fixed growth an branching patterns. Development. — It is interesting to spe ment for a branching system as well defined a examination of the rhizome raises the question 0 culate on the possible mode of develop- s that of H. repens. A superficial f whether branching in this species 46 AMERICAN FERN JOURNAL: VOLUME 71 (1981) is sympodial, monopodial, or dichotomous. If the first were the case, the question of development would be simplified from a descriptive point of view; all the leaves would be terminal and all branches could be produced in the same way. However, the anatomy of H. repens and the developmental studies of the closely related H. punctata by Imaichi and Nishida (1973) indicate otherwise. Instead, there are two types of branches in a branch complex: (1) the stem (Fig. 3, axis 1) bifurcates to produce a leaf (Fig. 3, axis 2) and a continuing stem; and (2) and one to four buds (Fig. 1, axes 3-6) branch off from the base of the leaf (Fig. 3, axis 2). The first type of branching may be dichotomous in the strictest sense, that is, by a division of the apical cell of the parent stem axis (Fig. 3, axis 1). The equal distribution of the vasculature between the daughter stem and leaf favors this explanation (see Fig. 4). On the other hand, the two daughter products are not the same. This might result from a lateral mode of branching, that is, one where meristematic tissue proximal to the apical cone of the continuing stem differentiates into and terminates as a leaf. — The methods by which the buds branch from the leaf are somewhat less clearly defined. Bower (1923) theorized that the formation of extra-axillary buds in H. repens represented a modification of dichotomous branching of the rhizome system, that, in effect, the basal bud (Fig. 3, axis 3) bore the leaf (Fig. 3, axis 2). The regular pattern of buds on the leaf base strengthened his argument that these buds were part of a regular branching system and therefore did not develop. adventitiously. Imaichi and Nishida’s (1973) decapitation experiments and ontogenetical observa- tions of the related H. punctata demonstrated that the bud meristem is formed after the leaf is established, making Bower’s theory untenable. In his developmental studies of Onoclea sensibilis, Dryopteris aristata, and D. filix-mas, Wardlaw (1943) demonstrated that adventitious buds occur only in specific positions corresponding to those occupied by detached meristems. These superficial bud meristems develop from the region of the shoot apical meristem not involved in the development of leaves. In the case of D. filix-mas, every bud initially occupies an axillary position. It may then become separated from the leaf with which it originally was in an axillary relationship by displacement onto the enlarging base of another leaf which was lateral to it in the earlier developmental phase. Since the bud develops more slowly than the leaf, its vasculature tends to become joined with that of the leaf it had been carried up on, rather than with that of the stem from which it originated. In summary, Wardlaw (1943) demonstrated that each bud of D. filix-mas, despite its position on the petiole of an adult leaf, occupies an approximately axillary position on the shoot at the time of its formation. Wardlaw concluded that extra-axillary buds in other fern species might originate in a similar fashion. It is possible that some interaction of hormones produced by the petiolar roots and/or developing frond may trigger the differentiation of meristematic cells pro- duced by the apical cone of the leaf. Only a thorough investigation of the developmental process will provide the insight needed to explain the origin of the interesting branching pattern of Hypolepis repens. Field work in Veracruz, Mexico was supported by the Atkins Garden Fund, Harvard University, and assistance in Mexico was obtained from the Instituto de T. M. GRUBER: BRANCHING PATTERN OF HYPOLEPIS REPENS 47 Biologia, Universidad Nacional Autonoma de México. I am grateful for the help and encouragement of ching 2 Rolla Tryon and Dr. Alice Tryon and especially for the guidance of Professor P. B. Tomlinson in preparation of the film. Appreciation is also extended to Scott Clempson, Kathleen Buckley, David Karachuk and the staff of the Gray Herbarium and Arnold Arboretum Library. LITERATURE CITED BELL, A. D. and P. B. TOMLINSON, 1980. Adaptive architecture in rhizomatous plants. Bot. J. Linn. Soc. 80:125—160. BOWER, F. O. 1923. The Ferns, Vol. I. University Press, Cambridge. GWYNNE-VAUGHAN, D. T. 1903. Observations on the anatomy of solenostelic ferns. Ann. Bot. :689-742. IMAICHI, R. and M. NISHIDA. 1973. Studies on the extra-axillary buds of Hypolepis punctata. J. Jap. Bot. 48:268-279. STEVENS, P. S. 1974. Patterns in Nature. Little, Brown, Boston. TOMLINSON, P. B. 1971. Monocotyledons—Towards an understanding of their morphology and peri In R. D. Preston (ed.) Advances in Botanical Research, vol. 3. Academic Press, TROOP, J. ra aa Yes 8 ae 1968. Petiolar shoots in the Dennstaedtioid and related ferns. Amer. Fern J. 58:64—7 WARDLAW, C. W., 1943. aA and analytical studies of pteridophytes, II. Ann. Bot. n. s., 7:349-377. 48 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) Diplazium japonicum and Selaginella uncinata Newly Discovered in Georgia WAYNE R. FAIRCLOTH* The Japanese Twin-sorus Fern, Diplazium japonicum (Thunb.) Bedd., has been reported as escaped from cultivation in Gadsden County, Florida since 1957 (Wherry, 1964), and recently Short (1980) published an account of its discovery in Lee County, Alabama. This species subsequently has been found in three widely disjunct locations in Georgia (Fig. /). In late summer of 1980, my wife and I led a fern field trip in central south Georgia for the Georgia Botanical Society. During the field trip, Marge White and Frieda Polsfuss (fern hobbyists from middle Georgia) showed several of us some pressed fronds of a fern which Mrs. White had collected in Houston County, Georgia. Upon superficial examination, most of us identified the plant as a form of Athyrium, although the possibility of its being a Diplazium was suggested. A frond was given to Lloyd H. Snyder, Jr. (a fern hobbyist from Atlanta), who discovered the same fern beneath a highway bridge in Berrien County, Georgia the following day. Specimens from Houston and Berrien Counties were shown to Dr. Murray Evans of the University of Tennessee, who confirmed the identity of both as Diplazium japonicum. The two sites were rechecked in October to determine the extent and condition of the colonies. The team of White, Polsfuss, and Snyder visited the Houston County site, which is located in dense, deciduous woods between Hatcher Road and Fagin Mill Road southwest of Warner Robins. Two groups of approximately a dozen plants each about 1000 m apart were found along the edge of a small stream in association with Asplenium platyneuron, Athyrium asplenioides, Onoclea sensibilis, Osmunda cinnamomea, O. regalis, Polystichum acrostichoides, and Thelypteris torresiana. My wife and I found the Diplazium in Berrien County growing beneath a slough bridge of the Withlacoochee River, west of Nashville on Georgia Highway 125. Thirteen plants bearing fertile fronds were clustered together in an area of three square meters. Eighty-two sporelings were widely scattered beneath the bridge, indicating that the colony was successfully reproducing and expanding. Other ferns found beneath the bridge were: Asplenium platyneuron, Lorinseria areolata, Lygodium japonicum, Onoclea sensibilis, Ophioglossum petiolatum, Osmunda regalis, Pteridium aquilinum, Thelypteris dentata, T. kunthii, T. torresiana, and Woodwardia virginica. On 2 January 1981, I discovered an immense colony of the Japanese Twin-sorus Fern on the Farmer’s Branch prong of Sofkee Creek in Grady County, Georgia. Hundreds of plants were growing thickly on both banks of this spring-fed stream for a distance of 28 m; in addition, scattered plants were found downstream for a distance of approximately 120 m. Most of the fertile plants were robust, with fronds *Department of Biology, Valdosta State College, Valdosta, GA 31601. W. R. FAIRCLOTH: DIPLAZIUM AND SELAGINELLA IN GEORGIA 49 commonly as long as 65 cm. The only ferns growing in association with D. japonicum at this site were Dryopteris ludoviciana, Lorinseria areolata, and Osmun- da cinnamomea. Houston County is located in the Upper Coastal Plain Province, almost in the center of the state and very near the Fall Line (the junction with the Piedmont Province). Geographically, it is similar to the Fall Line location in Lee County, Alabama. Grady County also is in the Upper Coastal Plain Province, but is about FIG. 1. Distribution of Diplazium japonicum (circles) and Selaginella uncinata (triangles) in Alabama, Florida, and Georgia. 140 miles to the southwest and borders Gadsden County, Florida. Berrien County is located in the Lower Coastal Plain Province, a region which differs from the Upper both in elevation and in having predominantly sandy soils. Unlike the rolling, well drained Upper Coastal Plain, the Lower is very flat, with numerous ponds, vast Swamps, and broad river systems. ae Not only do these new locations extend the range of Diplazium japonicum Significantly, the species is a new addition to the vascular flora of Georgia. Recent papers by Bruce, Jones, and Coile (1980) and by Duncan and Kartesz (1981) do not include D. japonicum as a component of Georgia’s flora. 50 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Specimens from the three locations (Berrien County, Snyder 697, Faircloth 8518; Houston County, Snyder 757; Grady County, Faircloth 8521) are on deposit in the herbaria at the University of Georgia (GA) and Valdosta State College (VSC). Another addition to the vascular flora of Georgia is Selaginella uncinata (Desv.) Spring. The location of an unfamiliar Spike-moss in Decatur County was first mentioned to me by Angus Gholson, U. S. Army Corps of Engineers, Resource Manager for Lake Seminole, Chattahoochee, Florida. My wife and I visited the site in late March, 1980. We found an extensive colony of the Blue Spike-moss growing luxuriantly on the banks of a small stream in a ravine immediately north of Greenshade Cemetery between Faceville and Fowltown (Fig. /). Shoot development at that season of the year was strictly sterile because a late freeze on 1 March 1980 had killed the aerial shoots back to ground level. There was evidence of abundant strobili development the preceeding year, although I could find no microspores or macrospores that had been retained. My identification, based upon sterile material, was confirmed by Dr. John T. Mickel (pers. comm.). Fertile material was collected later in the summer; specimens (Faircloth 8531) are on deposit in the Valdosta State College Herbarium. Naturalized Selaginella uncinata, a native of China, has been known from Florida and some of the Gulf Coast states for more than 20 years (Brown & Correll, 1942; Lakela & Long, 1976). With the return of horticultural interest in hanging baskets, it has lately been cultivated and widely sold as a hanging basket plant under the names of Rainbow-fern and Parlor-moss. Its iridescent, blue-green foliage and arching- trailing growth habit make it a choice plant for this purpose. The origin of the colony in Decatur County is puzzling. There are no homesites nearby, and its location in close proximity to a cemetery appears coincidental because the plant is not likely to be used either as a potted plant or in other types of floral arrangements for cemetery ornamentation. Mr. Gholson’s familiarity with the site indicates that the colony is at least four years old but judging from its size, it is perhaps as much as 8 to 10 years old. I wish to thank the following people for their help in field work and in providing information for this report: Juanita N. Faircloth, Marge White, Frieda Polsfuss, Leslie Garland, Lloyd H. Snyder, Jr., and Angus Gholson. LITERATURE CITED BROWN, J. W. and D. S. CORRELL. 1942. Ferns and Fern Allies of Louisiana. Louisiana State University Press, Baton Rouge, G. JO BRUCE, J. oe NES, Jr, and N. C. COILE. 1980. The Pteridophytes of Geo; sae Castanea 45:185-193. DUNCAN, W. and J. T. KARTESZ. 1981. The Vascular Flora of Georgia; An Annotated oe University of Georgia Press, Athens, GA. LAKELA, O. and R. W. LONG. 1976. Ferns of Florida; An Illustrated Manual and Identification Gade Banyan Books, Miami, FL. SHORT, J. W. 1980. Diplazium japonicum New to Alabama. Amer. Fern. J. 70-111. WHERRY, E. T. 1964. The Southern Fern Guide. Doubleday, Garden City, NY. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 51 Notes on Selaginella, with a New Variety of S. pallescens ROBERT G. STOLZE* One of the most variable neotropical species in the genus Selaginella is S. pallescens (Presl) Spring. It may grow on soil or rocks, in sun or in deep shade, from sea level to over 3000 m elevation, and it occurs in Mexico, Central America, Cuba, Jamaica, and parts of South America. The species belongs to the heterophyl- lous subg. Stachygynandrum, which is characterized by the stems (at least distally) and the vegetative leaves of branches having 2 rows of smaller, usually appressed, median leaves and 2 rows of larger, spreading lateral ones. Contrasting with this is subg. Selaginella, with plants homophyllous throughout, i.e., leaves are borne on all sides of the stem and branches and all are appressed for most of their length. Subgenus Stachygynandrum is sharply divided into two groups, one having stems articulate, or at least constricted at or near the nodes and here usually discolored, and with rhizophores produced dorsally. The other group, containing S. pallescens, has stems neither articulate nor with discolored or constricted nodes, and with thizophores produced ventrally, Selaginella pallescens and its nearest relatives commonly have the stems densely caespitose, often forming rosettes. They tend to curl inward when dry, then uncurl again when moisture is introduced, thus giving rise to the common name “Resurrection Plant.” Taxonomy of the entire S. pallescens complex in the neotropics needs careful re-examination. There are several species which I feel are not truly distinct, and yet there are some hitherto unrecognized variants, which perhaps should be formally treated as varieties or forms. Some very closely related species are: S. cuspidata (Link) Link and var. elongata Spring, S. harrisii Underw. & Hieron., S. microdendron Bak., S. millspaughii Hieron., and S. pulcherrima Liebm. During a study of the genus for the “Ferns and Fern Allies of Guatemala,” it appeared to me that most of these species might better be included under S. pallescens, for whatever differences were noted by previous authors appear to be thoroughly inconsistent. On the other hand, some new features have come to light which seem Significant and consistent enough to indicate recognition of some specimens at the varietal level. In his study of the spores of heterophyllous Selaginellae, Hellwig (Ann. Mo. Bot. Gard. 56:444—464. 1969) annotated a number of specimens in various herbaria as S. pallescens, “red-stemmed variant.” Some other minor features have been discovered to be consistent with stem color on these specimens, and the combination of characters support a decision to name the following new variety. Selaginella pallescens (Presl) Spring var. acutifolia Stolze, var. nov. arietas haec a varietate typica /S. pallescens (Presl) Spring var. pallescens] differt caulibus rubellis, foliis lateralibus acutis (mec acuminatis nec aristatis), et interdum pallide roseo-tinctis, et foliis medianis acutis (nec acuminatis nec aristatis). TYPE: Rocky hills near Santa Rosalia, 2 mi south of Zacapa, alt. 200 m; Depto. Zacapa, Guatemala, 1939, Steyermark 29293 (F). *Department of Botany, Field Museum of Natural History, Chicago, IL 60605. 52 AMERICAN FERN JOURNAL: VOLUME 71 (1981) In forests or wooded ravines; commonly on rocks, cliffs, or rock outcrops, from sea level to 1100 m; Guatemala, Honduras, El Salvador, Nicaragua, and Costa Rica. Plants epipetric or terrestrial; stems reddish, at least at base; lateral leaves acute, rarely subacute, sometimes a few of the older ones becoming streaked or tinged with pale red; median leaves commonly acute at apex. SELECTED SPECIMENS EXAMINED: GUATEMALA: Chiquimula: Damp thicket along road between Chiquimula and Zacapa, 400-600 m, Standley 74517 (F, US). Escuintla: Medio Monte, Palén, Mario Dary Rivera 755 (F). Jutiapa: Quebrada near Mangoy, 550 m, L. O. Williams 1420] (F). HONDURAS: Choluteca: Moist bank above San Antonio de Flores, 30 m, Williams & Molina 16714 (F). Morazan: Sabana Grande, 1100 m, J. desided ip 3260 (F, UC, US). El Paraiso: Drainage of Rio Yeguare, 600 m, Molina 4014 (F, US). EL SALVADOR: Chalatenango: Hills outside San José Cancasque, 400 m, Seiler 365 (F). San sees Riverside, Canton San Antonio Chavez, 300 m, Seiler 54] (F). La Unién: Humus, woods N of La Union, Morrison & Beetle 8761 (F, US). NICARAGUA: Leon: Canyon of Rio Sinecapa, near Santa Rosa, 200 m, Williams & Molina 42442 (F). Nueva Segovia: Ravine W of Ocotal, Seymour 840 (F). COSTA RICA: Guanacaste: Margen rocosa del Rio Recreo, 80 m, Jiménez 1172 (F). e€ most conspicuous difference between this and the typical variety is the reddish coloration present in the plants (a phenomenon not uncommon in the. subgenus). The stems are always reddish at base, and the color often extends halfway to the apex. Also, as the lateral leaves begin to age, some become tinged with red. Typical S. pallescens has stems pale greenish to stramineous throughout; and if lateral leaves turn color with age, it is to a dull whitish or tawny hue. Most median and lateral leaves in var. acutifolia are never more than acute, whereas in var. pallescens the leaves are acuminate or even aristate. The new variety prefers low altitudes from sea level to 800(1000) m, and is most common in rocky habitats. The typical variety is found occasionally near sea level, but most frequently occurs between 800 and 2500 m. It, too, is found in rocky situations, but seems equally at home on the forest floor. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 53 Lepisorus kashyapii in the Western Himalayas S. S. BIR and CHANDER K. SATIJA* Lepisorus kashyapii (Mehra) Mehra is a polymorphic species closely related to L. excavatus (Bory) Ching. The two species are known to hybridize in the Himalayas (Bir & Trikha, 1969, p. 271). The latter species was studied in detail by Bir and Trikha (1974). These two species of Lepisorus are more easily separated in the field than in the herbarium. The rhizomes of L. kashyapii are loosely attached to trees or rocks by long, straight roots, whereas L. excavatus has rather tightly adherent rhizomes. Lepisorus kashyapii laminae are thicker and have more obscure veins, and their color is paler yellow-green. According to a note by R. R. Stewart in the U.S. National Herbarium, L. kashyapii has quickly deciduous, dark brown, contorted hairs on the abaxial surface. Unfortunately, these characteristics are mostly difficult to observe in herbarium material. The rhizome scales of the two species are very similar. Although there is some variation, the sori of L. kashyapii are more round and only slightly immersed in the laminae, whereas those of L. excavatus are oval and more deeply immersed. Lepisorus kashyapii tends to dry brown, whereas L. excavatus usually dries green. These characteristics will aid in separating the two species in the herbarium. During a study of various collections of Lepisorus from the Himalayas, L. kashyapii was found to be morphologically very interesting. Some specimens from various localities around Nainital, a popular health resort located in the northwestern Himalayas, exhibit variations from typical L. kashyapii that proved to be two new varieties. These, along with var. kashyapii, are distinguished in the following key. KEY TO THE VARIETIES OF LEPISORUS KASHYAPII . Fronds linear to linear-lanceolate, 7.5—11.5 cm long, 0.5—1 cm wide; rhizome scales long-acuminate at 3. L. kashyapii var. minor ea the apex . 2 Finsae narrowly to sometimes broadly elliptic-lanceolate, (12)15—35(42) cm long, 1.5-3.5(5.5) cm wide; rhizome scales acute to acuminate at the apex. 2. Laminae bright brown on drying; rhizome scales peltate-lanceolate, dark brown, strongly clathrate in the center, yellowish and not clathrate at the margins, especially around the peltate base, the margins contorted and finely toothed when young, usually worn away and merely erose in age, sporangial paraphyses all peltate, clathrate .........-.-.+-ssessseeseeeeees Lak, kashyapii var. kashyapii . Laminae dull brown on drying; rhizome scales narrowly ovate-lanceolate, yellowish-brown through- out, clathrate in the center, less so at the margins, the margins erose and hyaline; sporangial paraphyses hair-like as well as peltate, clathrate ............-:sssesseesseeeeenersseeseen es 2. L. kashyapii var. major 1. Lepisorus kashyapii (Mehra) Mehra in Bir, Res. Bull. Panjab Univ. n.s., 13:23. 1962, var. kashyapii. Polypodium satslinagh arte tee Univ. Publ. 24, f. 5. 1939. TYPE: Not stated; a lectotype should be chosen from Mehra’s material at Panjab University, Lahore (LAH). Pleopeltis kashyapii (Mehra) Alston & Bonner, Candollea 15:208. 1956. Rhizomes long-creeping, 3-5 mm in diam.; rhizome scales ig abel pe central portion clathrate, brown to dark brown in mass, the marginal porti — nN *Department of Botany, Punjabi University, Patiala 147002, India. AMERICAN FERN JOURNAL: VOLUME 71 (1981) 005mm ; figs.4,5 Ff etagd ds Gege > sO. 24 e,, Cer boat ar y 49,¢8. “MN j i] a%, 4 MERE »_Smm lig Pr) figs.2,7 BIR & SATUA: LEPISORUS IN THE WESTERN HIMALAYAS 55 scarcely or not clathrate, yellowish, broader around the peltate base than toward the apex, the margins contorted, finely toothed, usually worn away and merely erose in age. Stipes 2-5 cm long, yellowish, bearing a few, usually somewhat contorted scales. Laminae narrowly to sometimes broadly aes -lanceolate, (12)15—35(40) cm long, 1.5—3.5(5.5) cm wide, bright brown on drying, acute above an acuminate bas acuminate at the apex, entire or slightly wavy along the margin; sporangia protected by subpersistent, clathrate, peltate paraphyses; spores reniform or oval, plane to concavo-convex, (idtie minutely verrucose, ca. 42-60 wm long, 30-50 pm wide; mber n=36. OTHER CI TATIONS: Mehra & Bir, Res. Bull. Panjab Univ. n.s., 15:168. 1964; Bir & Trikha, Bull. Bot. Surv. India 11:271-273. 1969 [1971]. ILLUSTRATIONS: Mehra (1939, pp. 24-25, fig. 5a—g); Bir & Trikha (1969, pp. 271-273, figs. 41-45). SPECIMENS EXAMINED: INDIA: Himachal Pradesh: Simla: Near Shali Peak, 2400 m, Sept 1969, Bir 1046 (PAN); Sanjouli, 2100 m, Aug 1969, Bir (PAN), Sept 1958, Bir (PAN); Chharbara, 2400 m, Aug 1960, Bir (PAN). Dalhousie: Satdhara, 2000 m, Sept 1968, Trikha 106 (PUN). Uttar Pradesh: Mussoorie: 1800 m, Aug 1959, Bir (PAN); Nag Tiba, 2400 m, Sept 1947, Fleming 82 (US); Lal Tiba, 2400 m, Sept 1968, Bir 44 (PUN), 2250 m, Sept 1968, Trikha 1070 (PUN). Nainital: Laria Kanta, 2400 m, Sept 1967, Bir (PUN); Tiffon Top, 2100 m, July 1971, Trikha 1901 (PUN); Khurpatal, Sariyatal, 1600 m, July 1971, Trikha 1904 (PUN). West Bengal: Darjeeling: Senchal forest, 1500 m, July 1957 Malhotra 744 (PAN); Manibhangang—-Tonglu Road, 2400 m, July 1957, Bir 774 (PAN), 2000 m, July 1957, Bir (PAN); Near Dingle Kothi, 1800 m, July 1969, Tritha 1025 (PUN); Senchal Lake, 2400 m, Aug 1969, Trikha 1030 (PUN); without definite locality, Thomson (US). NEPAL: Shotibas, 3000 m, Oct 1958, Polunin, Sykes & Williams 5568 (US); Garpung Kholen, 3000 m, Sept 1962, Polunin, Sykes & Williams 5410 (US). 2. Lepisorus kashyapii var. major, Bir & Trikha, var. nov. Figs. 1-5. Squamae rhizomatis ovatae, acuminatae, flavescenti-brunneae, margine eroso, kg lamina lanceolata, 29-42 cm longa, 2.7-3.5 cm lata, statu sicco obscure brunnea; sporangia paraphysibus biformibus, unis, umbelliformibus, parietibus cellularum comparate tenuibus, alteris bicellulatis, uniseriatis, piliformibus; sporae verrucosae vel tuberculatae 55-67 um longae, 33-55 pm latae; chromosomatum numerus n=36. TYPE: Cheena Peak, Nainital, Uttar Pradesh, India, epiphyte, 2400 m, July 1971, Trikha 1906 (PUN 1240; isotypes PUN 1241, 1242). PARATYPE: Tiffon ae Nainital, Uttar Pradesh, India, lithophyte, 2100 m, July 1971, Trikha 1907 (PUN 1357). 3. sami kashyapii var. minor Bir & Trikha, var. nov. Figs. 6-8. quamae rhizomatis ovatae, longe acuminatae, brunnescenti-atrae, margine eroso; lamina Hipearié vel lanceolato-linearis, 7.5-11.5 cm longa 0.5—1 cm lata, statu a laete fusca; sporangia paraphysibus solum umbelliformibus, clathratis, parietibus cellularum fortibus, Seas: sporae verrucosae, 50-67 jm longae, 38-50 jum latae; chromosomatum numerus FIGS. 1-5. Holotype of Lepisorus kashyapii var. major (Trikha 1906, PUN). FIG. 1. es FIG. 2. Rhizome scale. FIG. 3. Peltate sporangial paraphysis. FIG. 4. Hair-like paraphysis. Fl ae . FIGS. 6-9. Holotype of Lepisorus kashyapii var. minor (Trikha 1096, PUN). FIG. 6. Habit Scale from rhizome apex. FIG. 8. Peltate iolawaiat paraphysis. FIG. 9. Spore. 56 AMERICAN FERN JOURNAL: VOLUME 71 (1981) TYPE: Tiffon Top, Nainital, Uttar Pradesh, India, epiphyte, 2100 m, July 1971, Trikha 1096 (PUN 1175; isotype PUN 1243). Our grateful thanks are due to Prof. K. U. Kramer (Zurich) for the Latin diagnoses, to Dr. D. B. Lellinger for advice on taxonomic matters, and to the keepers of the cited herbaria for the loan of material. LITERATURE CITED ALSTON, A. H. G. & C. E. B. BONNER. 1956. Résultats des expéditions scientifiques genevoises au Népal en 1952 et 1954 (Partie botanique), 5.—Pteridophyta. Candollea 15:193-220 BIR, S. a and C. K. TRIKHA. 1969. Taxonomic revision of the polypodiaceous genera of India—IV. Polypodium lineare complex and allied species. Bull. Bot. Surv. India 11:260—276. [Pub- lished in 1971]. , and C. K. TRIKHA. 1974. Taxonomic revision of the polypodiaceous genera of India—VI. Legis orus excavatus group. Amer. Fern J. 64:49-63. MEHRA, P. N. 1939. Ferns of Mussoorie. Panjab Univ. (Lahore) Publ. 29 pp. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 37 Taxonomic Notes on Jamaican Ferns—III GEORGE R. PROCTOR* This paper continues my series concerning the taxonomy of the ferns of Jamaica (see Proctor, 1965, 1968). The genus Thelypteris is the largest and one of the most complex genera of ferns in Jamaica, when construed in the broad sense, with a total of 59 known species. In recent years, this world-wide taxonomic group has received intensive scrutiny, especially by R. E. Holttum (Old World) and A. R. Smith (New World). The former has presided over the disintegration of the genus into numerous small “splinter” genera, a process already initiated by Ching and others; Smith has developed an integrated classification exercising the concept of subgenera and sections. The present writer prefers the latter approach. The first of Smith’s subgenera is Amauropelta. This is the largest and most difficult group of Thelypteris species occurring in Jamaica, with a total of 24 species, nine of them believed to be endemic. Amauropelta was classified by Smith (1974) in nine sections, all but one of which are represented in Jamaica. The position of the Jamaican species in these sections may be summarized as follows: 2. Phacelothrix: thomsonii. 3. Uncinella: negligens, oligocarpa, germaniana, linkiana, gracilis, heteroclita. 4. Amauropelta: firma, basiattenuata, ba bisii, trelawniensis, randallii, sancta, nockiana, underwoodiana, harrisii, gracilenta, resinifera vats. resinifera and caribaea. 5. Blennocaulon: cheilanthoides. 6. Pachyrhachis: pachyrhachis, malangae vat. sitiorum. 7. Lepidoneuron: rudis. 8. Blepharitheca: concinna. 9. Apelta: (not represented in Jamaica). Three of the above species are new to science and are described herein, along with a new minor form of 7. rudis; four other names represent new combinations requiring validation. The localities of these taxa are shown in Fig. /. The writer is grateful to Dr. John Mickel for helping to locate, at the New York Botanical Garden, types of species described by Jenman. Thelypteris decrescens Proctor, sp. nov. he, Subg. Amauropelta, sect. Adenophyllum. Ex affinitate 7. pilosulae a qua stipitibus multo brevioribus, densissime minutissimeque stipitato-glandulosis, glabratis vel parce pilosulis; laminis minoribus, utrinque, sed plerumque subtus abundanter glandulosis, glandulis stipitatis, flavis resinaceis differt. hizome stout, erect, its scales yellow-brown, lance-attenuate, the margins subentire with a few minute colorless stipitate glands. Stipe very short, 3-4 cm long, scaly at base, puberulous, and, to ether with the rhachis, mec): and minutely stipitate-glandular, the glands colorless; also bearing few to many lon soft, pluricellular hairs. Blades narrowly elliptic to oblanceolate, 35-50 cm long, *Harvard University Herbaria, 22 Divinity Ave., Cambridge, MA 02138. 58 AMERICAN FERN JOURNAL: VOLUME 71 (1981) 10-15 cm broad at or above the middle, markedly decrescent downward with 7 pairs or more of reduced pinnae; longest pinnae oblong-linear, sessile, acuminate, up to 7 cm long, 1-1.7 cm wide at the base, with up to 20 pairs of oblong or narrowly deltate- oblong segments, these blunt at the apex and | .5— 2.5 mm wide, the margins flat or very narrowly reflexed; veins 5-8 pairs, simple, scarcely prominulous but with slightly enlarged tips (seen from the upper side). Sori supramedial, round; indusium delicately reniform, densely glandular (the glands minute, globular, and ales to stipitate), soon withering. Rhachis and other vascular parts densely pubescent on the upper (adaxial) side with whitish pluricellular hairs; similar hairs less dense beneath; all parts on both sides, but more abundantly beneath, beset with stipitate yellow-resinous glands TYPE: Upper west slope of Blue Mt. Peak, Parish of St. Thomas, Jamaica, 6500-7325 ft (1981-2233 m), L. M. Underwood 1513, 11-12 Feb 1903 (NY). Paratype from lower western ridge of Blue Mt. Peak, 5500-6318 ft (1700-1950 m), 4-9 July 1926, W. R. Maxon 10025, (NY, US). This is the first species of sect. Adenophyllum to be reported from the West Indies. Most of the other species of this section are South American, but the closely related Thelypteris pilosula (Mett.) Tryon has an extensive range from southern Mexico to Peru, and has been reported from Jamaica and Hispaniola. Thelypteris negligens (Jenm.) Proctor, comb. nov. Nephrodium negligens Jenm. Bull. Bot. Dept. Jamaica, n.s. 3:21. 1896. TYPE: Jamaica, without exact locality, Jenman s.n., (NY). bag ns ote abe naga Proctor, sp. nov. uropelta, sect. Amauropelta. Ex affintate T. balbisii a quo laminis ne Selon thachidi et costis laminarum supra sulcatis earum marginibus pilis incurvatis multicellularibus ca. 0.2 mm longis; venis 9-11 paribus differt. Rhizome decumbent-ascending or suberect, clothed at the apex with dark brown, lustrous, pabraie lance-attenuate scales 3-4 mm long. Fronds few, erect-arching, up to 65 cm long; stipes 4-8 cm long, deciduously scaly toward the base, minutely stipitate-glandular throughout and lightly clothed with small, curved, pluricellular hairs. Blades lanceolate, 45-60 cm long and up to 18 cm broad below the middle, rather abruptly narrowed at the base, acuminate at the apex. Rhachis yellowish- rown, together with the costae densely clothed just inside the adaxial groove with short (ca. 0.2 mm long), stiffly incurved, pluricellular hairs; underside of the pegs minutely stipitate- ore and sparingly clothed with a few long, transpar- t, septate hairs up to 1.5 mm long. Pinnae mostly at right-angles to the rhachis, Bae the lower reduced ones somewhat reflexed, the largest linear- to narro owly deltate-oblong, 1.5—-2 cm bi at the base, sessile, acuminate, deeply pinnatifid, with up to 23 pairs of segments; very small, brown aerophores present at abaxial base of costae. Segments Sbtong. subfalcate, 2.5—-3.5 mm wide and not over 6 mm long, subacute at the apex, the margins strigillose-ciliate, and with 9-11 pairs of simple veins. Veins lightly strigillose on upper (adaxial) side; veins and tissue beneath with small, erect, unicellular, straight hairs and numerous sessile, reddish- G. R. PROCTOR: NOTES ON JAMAICAN FERNS-III 59 resinous glands. Sori medial to supramedial; indusium erect, glabrous, densely resinous-glandular; sporangia glabrous. TYPE: 1 mile N of Spring Garden, Parish of Trelawny, Jamaica, 1500-1700 ft (457-518 m), 2 Mar 1978, G. R. Proctor 37704 (IJ). This species is known only from the type specimen. It appears to be related to Thelypteris balbisii (Spreng.) Ching, but differs in the nature of its indument, in the deflexed lower pinnae, and in the more oblique segments with fewer veins. From the related T. randallii Maxon & Morton ex Morton it differs in the much thicker thachis clothed on the sides of the adaxial groove with stiff, incurved hairs, and in having sessile, reddish-resinous glands only beneath. These three species and the next (7. harrisii) differ from the rest of sect. Amauropelta in having pluricellular hairs; in T. balbisii such hairs, however, may be present or absent. JAMAICA FIG 1. Known localities of cited species and varieties of Thelypteris. In order from west to east ( left to right): T. trelawniensis, T. resinifera var. caribaea, T. harrisii and T. malangae vat. sitiorum, i. negligens, and T. decrescens. Thelypteris harrisii Proctor, sp. nov. tdi kaulfussii sensu Jenm. Ferns Brit. W. Ind. & Guiana 210. 1908, non Hook., 1862. Jenman believed that this plant had been treated by Grisebach (Fl. Brit. West Ind. 691. 1864) as a variety of what is now called Thelypteris oligocarpa, and rightly rejected this assignment. However, taking up the name kaulfussii, he resurrected an epithet whose antecedents had been hopelessly confused by Hooker (Sp. Fil. 4:97. 1862), and which cannot be applied to the present pie = ubg. Amauropelta, sect. Amauropelta. Ex affinitate T. underwoo coger q y rachide dense pilosa, pilis ca. 1 mm longis vel longioribus; indusiis pilis se f nnae, the lowest often being mere auricles; rhe throughout and also soft-pilose with long, spreading, whit mm long or longer. Pinnae mostly linear-ob 1.5-1.8 cm broad above the sessile base; costae and other va 60 AMERICAN FERN JOURNAL: VOLUME 71 (1981) throughout with soft hairs; tissue freely resinous-glandular beneath and nearly glabrous. Segments close, strongly oblique, oblong, acute, up to 3 mm broad, the margins narrowly reflexed; veins 7-9 pairs, the lower ones often forked, all strongly prominulous on the upper (adaxial) side. Sori supramedial; indusium relatively large, round-reniform, densely long-pilose and ciliate, but without glands, decidu- ous; sporangia glabrous. TYPE: Moody’s Gap, border of Parishes of St. Andrew and Portland, Jamaica, ca. 3000 ft (914 m), 22 Oct 1898, W. Harris 7430, (IJ; isotypes BM nae Paratype from same locality, 13 Feb 1900, W. N. Clute 173 (NY). Thelypteris harrisii seems to be related to T. underwoodiana (Maxon) Ching, but clearly differs in its much longer stipes, denser, longer, and softer pilosity, and in having the indusium pilose but without glands, instead of ciliate and resinous- glandular. Thelypteris gracilenta (Jenm.) Proctor, comb. nov. Polypodium gracilentum Jenm. Bull. Bot. Dept. Jamaica, n.s. 4:129. 1897. TYPE: Jamaica, without exact locality, Jenman s. n. (N.Y.) is species somewhat resembles a large. T. gracilis in general appearance, but markedly differs from that species in details of the indument, sorus, and indusium. It is not clear why Jenman placed it in Polypodium, in view of its evident indusium, unless through faulty observation. Also, he stated that it is “common from 3,500 to 5,000 ft. altitude, in grass by the sides of open shallow streams and in similar wet exposed places.” The total lack of any subsequent collections indicates, however, that it must in fact be very rare, and its continued existence needs confirmation. Thelypteris resinifera var. caribaea (Jenm.) Proctor, comb. & stat. nov. Nephrodium caribaeum Jenm., J. Bot. Brit. For. 24:270. 1886. TYPE: North slopes of Mt. Diablo, Parish of St. Ann, Jamaica. Sherring s. n., (K photo US; isotypes IJ. US). Dryopteris caribaea (Jenm.) C. Chr., Ind. Fil. 257. 1905. Thelypteris caribaea (Jenm.) Morton, Amer. Fern J. 53:65. 1963. In recognizing this plant as a distinct species, Morton stressed characters which do not provide clear and sharp differentiation from 7. resinifera, but which in each case are more a question of degree (e.g., relative cell width vs. cell length in the clathrate rhizome scales, relative hairiness of the indusium). Although the ensemble of differences suggests a recognizable local variant of T. resinifera, more can hardly be said until living plants are rediscovered. — Thelypteris malangae var. sitorium (Jenm.). Proctor, comb. nov. Nephrodium jenmanii var. sitiorum Jenm. J. Bot. Brit. For. 17:261. 1879. TYPE: Jamaica, without exact locality, Jenman 38, in 1878 (K; isotype US). Nephrodium conterminum sensu Jenm. Bull. Bot. Dept. Jamaica n.s. 3:45. 1896, non Aspidium conterminum Willd. in L., 1810. Dryopteris consanguinea var. aequalis C. Chr. Smiths. Misc. Coll. 52:380. 1909. TYPE: Second Breakfast Spring, Parish of St. Andrew, Jamaica, W. R. Maxon 997 (US) (= Underwood 2131, NY). Differs from typical 7. malangae of Hispaniola in having narrower, more tapering pinnae (mostly 1.5—-1.8 cm wide vs. usually over 2 cm wide), the segments usually distinctly crenulate and relatively shorter and broader, and in the presence on G. R. PROCTOR: NOTES ON JAMAICAN FERNS-—III 61 Jamaican plants of widely scattered, very minute, colorless, stipitate glands, especially in the adaxial grooves of the rhachis and costae. In addition, the sori of var. malangae are approximately medial, whereas those of var. sitiorium are submarginal. Thelypteris rudis f. cristata Proctor, f. nov. A forma typica marginibus pinnarum apicem versus integerrimis, apice ipso cristato-laciniato differt. Differs from the typical form in having the distal part of the pinnae entire, at the end expanding into a cristate-laciniate apex. PE: Jamaica, without definite locality or collector, J. P. 1232-a (K; isotype IJ). This plant was gathered in 1885, probably by J. H. Hart or one of his colleagues in the then Botanical Department of Jamaica. LITERATURE CITED PROCTOR, G. R. 1965. Taxonomic notes on Jamaican ferns. Brit. Fern Gaz. 9: Sie —. 1968. Taxonomic notes on Jamaican ferns—II. Brit. Fern Gaz. 10:21-25, SMITH, A. R. 1974. A revised classification of Thelypteris subg. eamiaabe "Amer. Fern J. 64:83-95. 62 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) SHORTER NOTES NOTES ON NORTH AMERICAN LOWER VASCULAR PLANTS-II.—Field work in Arizona and examination of herbarium specimens at ASU, LL, and NY have revealed several new state records for Arizona and various states in northern Mexico. Also presented here are range extensions for two species within states where previously reported, one for Arkansas and one for Arizona. I am especially grateful to have been able to examine the numerous fine collections of Marshall C. Johnston and his former students, Thomas L. Wendt and Fernando Chiang C. A second locality in Arkansas is now known for Cheilanthes eatonii Baker in Hook. & Baker. The collection data are: Benton Co., Arkansas, E. N. Plank s. n. in 1899 (NY). The species was first reported in Arkansas in Baxter Co., by W. C. Taylor & D. Demaree (Rhodora 81:514. 1979). They reported it as C. castanea Maxon, which I do not recognize as distinct from C. eatonii [A monograph of the fern genus Cheilanthes section Physapteris (Adiantaceae), Ph.D. Dissertation Ari- zona State University, 1979]. Cheilanthes < parishii Davenp. (pro sp.) has been found new to Arizona. The collection data are: Dushey Canyon, Harquahala Mts., Maricopa Co., Arizona, growing near C. covillei Maxon and C. parryi (D. C. Eaton) Domin, desert scrub vegetation with Saguaro, Ocotillo and Jojoba, igneous substrate, 3000 ft elevation, Reeves 7127 (ASU). Reported previously from three localities in southern California by A. R. Smith (Madrofio 22:377. 1974). I agree with Smith that this is a sterile hybrid between the species listed above. A single plant was found at this site where the presumed parents are abundant. The first collection from Chihuahua, Mexico for Notholaena bryopoda Maxon has been made. The collection data are: Sierra del Roque, N of Julimes and N and NW of Rancho el Sauz, 28°39’—28°41'N, 105°20’18”—105°20'30"W, 1450-2150 m ele- vation, matorral desertico con espinos laterales, steep slopes of limestone mountains, limestone gravel, with Acacia neovernicosa, Dasylirion, Agave lecheguilla, Fouqueria splendens, and Parthenium incanum, M. D. Johnston et al. 12314 (LL). Previously known from Coahuila and Nuevo Leén, where it occurs on gypsum, according to R. M. Tryon (Contr. Gray Herb. 179:77. 1956). Notholaena greggii (Kuhn) Maxon is now known from Nuevo Leén. The collec- tion data are: Sierra Madre Oriental, Nuevo Leén, Mexico, calcite and limestone hills beyond Pablillo toward Santa Clara, 15 mi SW of Galeana, scattered on bank of calcite, C. H. & M. T. Muller 1080 (LL). Previously known from Texas, Chihua- hua, Coahuila, and Durango (Tryon, 1956, p. 76). The first collection of Notholaena neglecta Maxon in Nuevo Leon has been made. The collection data are: Minas “Manto Blanco” y “Sabana Blanca” just N of the Canon de Potrerillos, Nuevo Leén, Mexico, 26°04’N, 100°45’W, 950-1000 m elevation, crasi-rosulifolios espinos, limestone ridge, gypsiferous clay loam, with Agave lecheguilla, Hechtia, Fouqueria, Larrea, and Opuntia rufida, M. C. Johnston et al. 10248C sae Previously known in Texas, Arizona, Chihuahua, and Coahuila (Tryon, 1956, p. 76). AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) 63 Notholaena parvifolia Tryon has been collected in San Luis Potosi. The collection data are: Estaci6n Microondas “Pastoriza” about 22 km S of Matehuala, San Luis Potosi, Mexico, 23°25'05"—23°25'25"N, 100°38’50’—100°39'00"W, 1550-1650 m elevation, crasi-rosulifolio espinoso, a few patches of matorral, limestone hills, calcareous gravelly soil, with Orthosphenia mexicana, Cnidoscolus sp., Eysenhardtia sp., and Agave lecheguilla, M.C. Johnston et al. 11111B (LL). Previously known from New Mexico, Texas, Chihuahua, Coahuila, Nuevo Leén, Tamaulipas and Zacatecas (Tryon, 1956, p. 99). The first record for Pellaea intermedia Mett. ex Kuhn in San Luis Potosi has the following collection data: 1 km by winding road below and W of Real de Catorce, on road to Estacién Catorce, above Socavon La Purisima, San Luis Potosi, Mexico, 23°41'40"N, 100°53’50”W, 2400-2450 m elevation, crasi-rosulifolio espinoso, badly disturbed agriculturally, very steep canyon slopes of metamorphic rock, with Agave spp. and Opuntia spp., M. C. Johnston et al. 11070A (LL). Known previously from Arizona, New Mexico, Texas, Chihuahua, Sonora, Coahuila, Nuevo Leon and Zacatecas, according to A. F. Tryon. (Ann. Missouri Bot. Gard. 44:179. 1957). Pitryogramma triangularis (Kaulf.) Maxon var. triangularis is now known to occur in Arizona. The collection data are: Frehner Canyon, Virgin Mountains, Mohave Co., Arizona, very scarce, steep rocky N slope, granite, with Pinon, Quercus turbinella, Ephedra, Galium, Stipa, Sitanion, and Thamnosma, 5000 ft elevation, R. Gierisch 4598 (ASU). This variety was previously known from southern British Columbia, Washington, Oregon, California, Baja California, south- ern Nevada, and southwestern Utah, according to K. S. Alt and V. Grant (Brittonia 12:155. 1960). Variety maxonii Weath. occurs in central and southern Arizona. The locality reported here for variety triangularis is in the extreme northwestern corner of Arizona. The first collection of Polypodium glycyrrhiza D. C. Eaton in Arizona has been made. The collection data are: Devil’s Chasm, Sierra Ancha, Gila Co., Arizona, narrow, deep gorge, one large patch ca. 6 X 6 ft on cliff, ca. 5000 ft elevation, B. Warner s.n., 10 Jan 1979 (ASU). Known previously from Kamtchatka, the Aleutian Islands, Alaska and coastal British Columbia, Washington, Oregon, and California (south to central part of state), according to R. M. Lloyd and F. A. Lang (Brit. Fern Gaz. 9:171. 1964). The specimen examined has the “sweet” rhizome and free venation of P. glycyrrhiza, in contrast to the acrid rhizome and usually anastomosed venation of P. californicum Kaulf. The material does not resemble P. hesperium Maxon, which is rather widely distributed in Arizona. The presence of P. glycyrrhiza in Arizona brings to three the number of predominantly Pacific Coast ferns found disjunctly in central Arizona. The other two are Dryopteris arguta (Kaulf.) Watt. and Woodwardia fimbriata J. E. Smith in Rees. na ae Selaginella eremophila Maxon is now known from several additional localities in southwestern Arizona. The collection data are: Sierra Estrella Regional Park, Maricopa Co., Arizona, W-facing wash on west face of Squaw Tit, moist desert under rocks with Penstemon antirrhinoides, Salvia mohavensis, Notholaena standleyi, and Castilleja lanata, 2400 ft elevation, E. & M. Sundell 177 (ASU); West side of White Tank Mountains, Maricopa Co., Arizona, slope leading up to light 1, lower 64 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 2 (1981) sonoran desert slope, at base of cliff, D. Keil 4088 (ASU); 7 mi S of Buckeye on U.S. 80, 2.4 mi along Buckeye Recreation Area Park road, on S-facing slope, under rocky ledge, 1189 ft elevation, A. Pierce 2/1 p. p. (a small fragment is apparently S. arizonica Maxon); Cabeza Prieta Game Range, Yuma Co., Arizona, S4, T14S, RI5W, ca. 6 mi NW of Tule Well, E-facing slope of low mountain, E. Lehto et al. 23548 (ASU, NY, US). This species previously was known in Arizona from a single collection near Tinajas Altas, Yuma Co., collected by Jaeger in 1934, according to R. Tryon (Ann. Missouri Bot. Gard. 42:80. 1955). All of the cited collections previously had been identified as S. arizonica, which is more common in Arizona than is S. eremophila. So far as I know at present, the two species occur together only at the Buckeye locality. I have seen material of S. arizonica from the White Tank Mountains, but from a different locality than the one cited above for S. eremophila. Selaginella eremophila is otherwise known from southern California and Baja California (Tryon, 1955). Field work in Arizona was supported by National Science Foundation Dissertation Improvement Grant 77-00182 to Dr. D. J. Pinkava and the author. I thank the curators of the cited herbaria for permission to examine their collections.—Timothy Reeves, Biological Science Center, Boston University, Boston, MA 02215. EQUISETUM ARVENSE IN ALABAMA.—The Common Horsetail, Equisetum arvense L., has been reported from Alabama by various authors from Small (Ferns of the Southeastern States, 1938) to Evans in Radford, Ahles, and Bell (Manual of the Vascular Flora of the Carolinas, 1968), none of whom gave any indication of locality. According to Dean (Ferns of Alabama, 1969, p. 153), “A large colony was discovered in Marengo County by Dr. R. M. Harper.” This was puzzling because Marengo County is in the coastal plain of southwestern Alabama, but E. arvense is a northern plant reaching its southern limit in Alabama. One would expect the plant to be present only in the northern part of the state; indeed, Wheatstone and Atkinson (Castanea 44:1—8. 1979) found it in Morgan County in central northern Alabama in 1974. It also has been found recently in Calhoun County (R. R. Haynes, UNA). While boating on the Black Warrior River in Greene County in October 1978, I was quite surprised to find a thriving colony of E. arvense on the west bank of the river. This locality is in the coastal plain about 20 miles upstream from Demopolis, which is in Marengo County. Therefore, this find lent credence to the report from that county. The Greene County colony was much larger and denser than the Morgan County population. It was growing in damp, partially shaded sand at the base of a wet chalk bluff, just above the normal level of the river (Short 1183, AUA, and duplicates to be distributed). On a visit to the U.S. National Herbarium in April, 1980, I examined a specimen of E. arvense (R. M. Harper 121, US) collected on October 11, 1908 in Marengo County on the bank of the Tombigbee River about 10 miles downstream from Demopolis. The plants were growing in sand at the base of a wet chalk bluff. This evidently is the collection referred to by earlier authors.—John W. Short, 905 McKinley Ave., Auburn, AL 36830. BRITISH PTERIDOLOGICAL SOCIETY Open to all who are interested in growing and studying ferns and fern-allies. Full members receive THE FERN GAZETTE and the BULLETIN OF THE BRITISH PTERIDOLOGICAL SOCIETY. Membership subscriptions are £7 for full members, £ 5 for ordinary members (not receiving the GAZETTE), and £5 for student members (under 25 years of age). For particulars, U. S. residents should apply to Dr. J. Skog, Biology Department, George Mason University, Fairfax, VA 22030. Non-U. S. residents should apply to Lt. Col. P. G. Coke, Robin Hill, Stinchcombe, Dursley, Gloucestershire, England. TRIARCH Over 50 Years of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 AMERICAN Boe FERN Aas J O U R NA L July-September, 1981 QUARTERLY JOURNAL OF THE AMERICAN FERN SOCIETY Arachniodes simplicior New to South Carolina and the United States JUDITH E. GORDON 65 A New Isoétes from Jamaica R. JAMES HICKEY 69 Leaf Turnover Rates and Natural History of the Central American Tree Fern Alsophila salvinii RALPH L. SEILER 75 Nomenclatural Notes on Micronesian Ferns F. R. FOSBERG and M.-H. SACHET 82 x Asplenosorus shawneensis, a New Natural Fern Hybrid Between Asplenium trichomanes and : Camptosorus rhizophyllus ROBBIN C. MORAN 85 Notes on North American Ferns DAVID B,. LELLINGER 90 Shorter Notes: Salvinia minima New to Louisiana; An Unusual Record of Asplenium trichomanes from Northeastern Florida 95 Reviews 68, 8&4 S i : : ss aici 06 uggestions to Contributors + MISSOURI ome ocT 17 s196f" GARE WT The American Fern Society Council for 1981 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, OH 4570 President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville. Ane 47235. -Vice Pah i! MICHAEL I. COUSENS, Faculty of Biology, University of West Florida, Pensacola. FL 325 eee JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, VA 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, DC 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, ih CA 94720. Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, NY 104 Newsletter Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, DC 20560. ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington. IN 47401. JOHN T. MICKEL New York Botanical Garden. Bronx, NY 10458. The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and or: at the Smithsonian Institution, Washington, DC 20560. Second-class postage paid at Washin Claims for missing issues, made 6 months (domestic) to ie suis (foreign) -after the date of issue, and the matters for a should be addressed to the Editor. Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog. Dept. of Biology, George se University, Fairfax, VA 22030. Orders for back issues should be addressed to the Treasure General inquiries oe ferns should be addressed to a Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0. os sent free to members of the Aenchiah Fern Society (annual dues, $8.00; life membership. $160. Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or 68 $1.25: 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458. is Librarian. Members may borrow books at any time. the borrower paying all shipping costs. Newsletter John T. Mickel, New York Botanical Garden, Bronx, NY 10458. is editor of the newsletter “Fiddlehead Forum.” The editor welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia. horticultural notes, and reviews of non-technical books on ferns. Spore Exchange Mr. Neill D. Hall, 1230 Northeast 88th Street, Seattle, WA 98115, is Director. Spores exchanged and collection lists sent on request. Gifts and Bequests Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Botanical books, back issues of the Journal, and cash or other gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secretary. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 3 (1981) 65 Arachniodes simplicior New to South Carolina and the United States JUDITH E. GORDON* During the past year, a population of about 100 plants of Arachniodes simplicior (Mak.) Ohwi was found along nearly 100 m of the bank of a small, unnamed tributary of the Savannah River in North Augusta, South Carolina. The species was identified using Ching’s (1934) treatment of Asiatic species of Rumohra Raddi, a genus now correctly designated Arachniodes Blume, as explained by Tindale (1961). The identification was confirmed by the United States National Herbarium. Speci- mens from the American population are described as ee (see also Figs. 1, 2): Rhizome moderately creeping, commonly S—8 mm in diameter, densely covered with tan to cinnamon-brown, papery scales, these long triangular with auricular- clasping bases, 0.5-2.5 mm wide, 3-8 mm long, the tips one cell wide and 2-4 cells long, the margins entire except for | "3 trichomes, 3—5 cells lo geod near scale bases and a few basal teeth A ose from outward distensions of two cell end junctures. Stipes commonly 0.5—4 cm apart, 15—46 cm long, mm diane: at the base, often equal to or longer than aie laminae, sub-terete, latened on the upper surface and grooved toward the apex, tan at the base, pale green above, aging to a straw-yellow color, somewhat scabrous from small preiberences on ica scattered scales are borne, the scales more oe sehr ae tibet the stipe ,2-I12m long, and similar to those of the rhizome ex cept more elongate and hocolale: brown, ~ Len consisting of 2 large and 3 small fanilles at the base, fe he to 3 at the Rachis unwinged, pale green, the upper surface grooved, the scales like those of the stipe except smaller. Costae yellow-green, with grooves continuous with Spa o me rachis, the scales smaller and more abruptly ogee Laminae ts 26-34 cm wide, deltoid- Sonal 2—3-pinnate, coriaceous, Tare ial, the mata surfaces glossy dark green with yellow- as Be along the costes and adjacent bases of pinnules, the abaxial surfaces uniformly pale green and not glossy, the apex gradually acuminate. Pinnae generally |-pinnate, 3—4 Bette pairs below i i m long cm the terminal pinna, 1-6 cm distant, at 30—45° to the rachis, 8-1 , wide at the base except the basal pair, 2-pinnate with the base 8-10 cm wide, the petiolules 1-6 mm long, the apices gradually acuminate, the pinnules and venation anadromically arated on all pinnae. Pinnules, except the lower pairs on the basal pinna pair, 16—22 alternate pairs, asymmetrically ovate-auriculate with an acroscopic basal lobe, 13-22 mm long, 5-10 mm wide, the petiolules narrowly whe gies sessile or up to | mm long, the apices sie, the margins with spines up to 1 mm long, with a few trichomes along the larger veins; basal pinnule pair on the nee pinna pair basiscopically produced with the upper about /- ys the length of the lower, the latter 6-10 cm long, the ultim ee segments resembling the pinnules of the 2-pinnate pinnae. Pintiale venation free, semi-dichotomous, not extending to the margins, the first branch arising anadromically. Sori dorsal, globular, arranged in a single row on each side of the pinnule main vein, terminal on smaller lateral veins, somewhat closer to the main vein than the margin. Indusia 0.5-1.0 mm in diameter, slightly darker brown in the center, glabrous, the margins entire, persisting for about two months before being shed. Sporangia with stalks composed of three rows of *Department of Biology, Augusta College, Augusta, GA 30910. Volume 71, number 2, of the JOURNAL, was issued June 29, 1981. 66 AMERICAN FERN JOURNAL: VOLUME 71 (1981) FIGS. 1 and 2. Photographs of Arachniodes simplicior. FIG. 1. Frond as seen in natural habitat, FIG. 2. Lowermost basal pinna, dorsal view. J. E. GORDON: ARACHNIODES NEW TO THE UNITED STATES 67 cells, 0.3-0.4 mm long, englandular, the annulus of 13-16 thickened cells. Spores bilateral, dark brown, rugose-reticulate, 4144ym X 28—-324m. Chromosome number 2n = 164 (Léve et al., 1977), undetermined for this population. The unusual nature of the habitat deserves further mention. The area, although within the city limits, is accessible only by foot and is characterized by steep, southwest-facing cliffs and heavy undergrowth within southern hardwood forest that probably has not been logged. The area surrounding the creek is dominated by Fagus grandifolia with a scattering of Celtis occidentalis and various species of Quercus. The understory is dominated by a species of Aesculus, apparently a hybrid between A. pavia and A. sylvatica. Polystichum acrostichoides (Michx.) Schott plants grow intermixed with the population of A. simplicior, but are less numerous. A small population (ca. eight plants) of Pteris multifida Poir., a naturalized species, 1s also present along the creek bank. The site has the following coordinates: 33°30’ 11” N Lat., 81°58'54” W Long., 42-53 m elev. This places the site along the western edge of the Hammond Hills subdivision. The area is seldom visited by hikers, but there have been two major disturbances within the last 50-55 years according to city officials. About 1927-29, the Georgia and Florida Railroad extended a line northward through the area. The rail bed was built with fill (source unknown) rising about seven meters above the creek bed. the creek water being piped through at the base. Plants of A. simplicior grow within a meter of the pipe. Use of the rail line was discontinued about 1954, at which time the city laid a sanitary trunk sewer line through the area. The sewer pipe, which is elevated about seven meters, parallels the old rail bed and is downstream about 60 m. If the population were present in 1954, it would have been disturbed by the construction of the sewer line. Observations of greenhouse and field plants show a growth pattern based on two new fronds produced yearly, usually in April in the North Augusta area. The persistence of the old frond bases permits a rough estimate of individual rhizome age, obtained by counting old bases and dividing by two. Ten rhizomes were examined; ages ranged from 2-15 years, using this method of calculation. Consider- ing the likely decay of older rhizome portions, the population size, and possible establishment from a spore source, it appears that the population is at least 20-25 years old. There are several possible spore or rhizome sources in the vicinity, including several nurseries, although none is within a mile radius of the site. One of the nurseries reported selling A. simplicior about three years ago, but not prior to that time nor within the last year. Another source of spores or rhizomes would include plants purchased by residents of the Hammond Hills subdivision, which was established about 1955. The probability of someone having actually planted the fern at the site is highly unlikely since access to the area is difficult. Based on the available information, I believe the population was probably established from a spore source shortly after the construction of the sewer line in 1954. Further studies centering on chromosome counts and gametophyte developmental stages are antici- pated. 68 AMERICAN FERN JOURNAL: VOLUME 71 (1981) LITERATURE CITED CHING, R. C. 1934. A revision of the compound leaved Polysticha and other related species in the continental Asia including Japan and Formosa. Sinensia 5:23-91. LOVE. pt D. LOVE, and R. E. G. PICHI SERMOLLI. 1977. Cytotaxonomical Atlas of the erido’ saa rene Vaduz. TINDALE, M. D. 1961. Aspidiaceae in nals of South Eastern Australia. Contr. N. S. W. Natl. erb. Ben Leong 208-21 1:1 REVIEW VASCULAR PLANTS OF CONTINENTAL NORTHWEST TERRITORIES, CANADA, by A. Erling Porsild and William J. Cody. 1980. 667 pp. National Museums of Canada Publ. Div., Ottawa, Canada K1A OM8. ISBN 0-660-001 19-5. Can$80.00.—This admirable volume provides descriptions, keys, habitat data, and maps for 1113 species of vascular plants which occur in the region immediately north of British Columbia, Alberta, and Saskatchewan. Information also is included on a number of species expected by the authors to occur in the region. There is an excellent survey of the history of collection in the region and a worthwhile bibliogra- h y. The pteridophytes are treated as 40 species, one subspecies, and one variety present, with four more species expected. Nearly half are fern allies. There are no great surprises in this subarctic flora; no hybrids or taxonomic novelties were noted. The limestone plants are of interest. I found the distribution maps to be curious and unexpected. They include Alaska and the Aleutians to the west, Greenland and Iceland to the east, and Chicago to the south. Thus, they are too small for detailed distributions within the region covered by the book. They do nothing to support the six phytogeographic zones designated by the authors, and certainly do not give a clear picture of the tree line, areas well collected, or other extrapolations one usually extracts from dot maps. Although we are warned (p. 3) that the maps are “in no way complete” for the Canadian distributions and are meant to show a “broad picture,” they have serious shortcomings. A person interested in Newfoundland might conclude that Woodsia alpina or Phegopteris connectilis does not occur there. Also, no fiddleheads in New Brunswick where they are canned in quantity? More serious are the maps which record more than the taxon being treated, e.g., Botrychium lanceolatum versus subsp. lanceolatum, Gymnocarpium dryopteris versus subsp. disjunctum, and Wood- sia oregana versus W. cathcartiana. I think it is a great pity that the authors fall back on names such as Dryopteris disjuncta for Gymnocarpium dryopteris and D. phegopteris for Phegopteris connectilis. 1 do not like Polypodium vulgare subsp. virginianum for P. virginianum or Dryopteris dilatata (the common tetraploid of Europe) for D. expansa, but these are newer and more problematical changes All in all, this book is a most useful compendium of the pteridophytes of a broad area of boreal, sub-arctic and arctic Canada and will be a constant source of reference for anyone interested in this part of the world.—D. M. Britton, Dept. of Botany and Genetics, University of Guelph, Guelph, Ont. NIG 2W1, Canada. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 3 (1981) 69 A New Isoétes from Jamaica R. JAMES HICKEY* The genus /soétes L. is poorly represented in the Caribbean, with only two species reported to date. /soétes cubana Baker collections made by Wright in the western, lowlands of Cuba. To my knowledge, there have been no subsequent collections of this species from Cuba, although certain collections from Belize strongly resemble the Cuban material and probably are conspecific with it. /soétes tuerckheimii Brause is known from several collections made in the Cordillera Central in the Dominican Republic at altitudes of 2200-2900 meters. Thanks to the generosity of Dr. George Proctor, who kindly supplied spirit and herbarium material, I am able to report the discovery of a third Caribbean species which is endemic to Jamaica. Isoétes jamaicensis Hickey, sp. nov. Figs. 1-4, 6, and 8 Radices dichotome ramosae. Cormus trilobatus, 1-2 cm diametro. Folia 20-55 (X =38), 8-20 cm longa, 0.5—0.8 mm lata in medio, attenuatissima, recurvata. Stomata et fasciculi fibrosi peripherales praesentes. Velum parvum. Margines ligulae hyalini, ephemeri; area centralis triangularis, attenuata versus apicem. Megasporae albidae, 320-440 (X=398) pm, tuberculatae. Microsporae cinereae, ellipticae, 30-40 (x = 36) ym longae, 22.5-32.5 (X= 27.3) um latae, cum papillis cavis. TYPE: In mud of drying seasonal rain pools, ca. 350 ft, Harris Savanna, Clarendon Parish, Jamaica, 26 Nov 1974, G. R. Proctor 34357 (1). PARATYPE: Harris Savanna, Clarendon Parish, Jamaica, 26 Nov 1974, G. R. Proctor 34358 (IJ). In habit and spore morphology, /. jamaicensis most closely resembles /. montezumae Eaton of Mexico (see Table | for comparison with other Mexican and Caribbean species). Although the distinctness of the latter species from /. mexicana Underw. has been questioned (e.g., Reed, 1953; Pfeiffer, 1922), preliminary examination of numerous Mexican collections, including type specimens, suggests that hybridization, possibly involving the formation of partially fertile hybrids, has caused the confusion between these species. In light of this possibility, /. montezumae certainly merits provisional specific status until my more detailed study of these species and /. pringlei Underw. is completed. Isoétes jamaicensis and I. montezumae differ in several respects, of which the more salient are discussed here. The corm of /. jamaicensis is distinctly three-lobed, whereas in the latter it is deeply bilobed. The membranaceous margin at the base of the leaves is less pronounced in /. jamaicensis than in /. montezumae. In the former, it is 1-2 mm wide and merges into the leaf proper 5-10 mm above the sporangium; in the latter, the margin is 1.5-2 mm wide and extends 10-20 mm above the sporangium. Both species have similar velum coverage of the sporangium. Eaton (1897) reported a very narrow velum in megasporophylIs and virtually none in microsporophylls of /. montezumae. In I. jamaicensis, there 1s little dimorphism in *Biological Sciences Group, U-43, University of Connecticut, Storrs, CT 06268. TABLE |. COMPARISON OF ec sca the AND CARIBBEAN ISOETES SPECIES. ! Character Habitat Pana igi Corm Leaf a Leaf length (cm) Leaf diameter (mm) Stomates Fibrous strands Megaspore ornamentation Perispore strands Megaspore size (wm) Microspore color Microspore ornamentation Microspore length £m Microspore width £m + > white tuberculate nsis I, mexicana amphibious 1830-2150 grey-white h wn smooth-echinate 25-38 25-33 I. montezumae amphib-terr 2 8-20 8-14 0.6-1.2 + + > white tuberculate + united 350-510 ash-grey papillate 23-28 I. pringlei amphibious A 10-20 16-25 0.5-1 + >lA white cristate-echinate 'Data obtained from original descriptions, Pfeiffer’s monograph (1922), and personal observations. I. cubana amphibious >50 rudimentary grey-white tuberculate all united 290-400 fawn echinate-papillate 25-33 20-25 1. tuerckheimii amphib-aquat 2200-2900 (1861) TZ SWATOA “IWNUNO! NY34 NYOMI R. J. HICKEY: NEW ISOETES FROM JAMAICA 71 velum coverage between the megasporohylls and the microsporophylls: most of the leaves show some velum coverage (Fig. / and 2), although in a single microsporo- phyll of the type collection there was no appreciable velum development. Ligule characteristics show considerable variability due to the ephemeral nature of the hyaline margin and due to the delicate nature of the central region, which deterio- rates rapidly with age (Figs. / and 2). The sporangium wall is unspotted and consists of thin-walled cells in /. jamaicensis (Fig. 3), just as it does in many specimens of J. montezumae. The tuberculate megaspores of both species are quite similar (Figs. 4 and 5). However, /. montezumae has consistently larger equatorial ridges and is more variable in the extent of ornamentation; the spores range from distinctly to indistinctly tuberculate. /soétes jamaicensis is more consistent in possessing well developed tubercles. Specimens of /. montezumae also show variation in tubercle distribution. On the distal surface, the tubercles diminish in size and increase in frequency close to the equatorial ridge. The tubercles of the proximal surface are always less well developed. In /. jamaicensis, on the other hand, the tubercles are uniformly developed and distributed. The tubercles of both species consist of solid masses of perispore material extending outward from the spore surface (Figs. 6 and 7). In both species, cords of perispore material radiate out from the tubercles, giving the spores a cobwebby appearance. While superficially similar, a closer examination of the spores and tubercles shows some distinct morphological differences. In /. montezumae, the perispore layer between the tubercles consists of very slender, interwoven strands which form a mat-like surface (Fig. 7, lower right). These small strands merge and two groups of them twist together to form the cords which can be seen radiating out from the tubercles. The cords in /. jamaicensis are also composed of smaller strands, but they are larger than those in /. montezumae, parallel, and are never interwoven. Between the tubercles, these small strands are isolated from one another and do not form any sort of solid structure. The microspores of /. jamaicensis are ornamented with numerous, conical, hollow papillae (Fig. 8). In some spores of the type collection, the papillae are quite pointed, approaching echinate projections in form, and are occasionally branched apically. The bases of the papillae are confluent. In /. montezumae (Fig. 9), the papillae are sparse, rounded, and give no indication of being hollow (as evidenced by their invariably unbroken appearance). The papillae are quite distant, and between them the microspores are covered with a finely granular perispore. Unlike I. jamaicensis, the equatorial ridges are visible in /. montezumae when the micro- spores are viewed from the side. Both species have a prominent proximal suture. The holotype of /. montezumae (Pringle 3459, MO) shows no indication of spore abortion, nor do either of the collections of /. jamaicensis. Spore abortion is not uncommon in many of the Mexican collections of /soétes. Isoétes jamaicensis inhabits seasonal rain pools in open xerophytic scrub of the Harris Savanna. This region receives 35-40 inches of rain annually during seasonal rainfalls. However, according to Dr. Proctor, the pools containing I. jJamaicensis form only once every four or five years, when the rains are particularly heavy. AMERICAN FERN JOURNAL: VOLUME 71 (1981) FIGS. |I-3. Leaf morphology of the type of /svétes jamaicensis (Proctor 34357, \J). FIG. 1. Megasporophyll showing incomplete velum and degraded ligule, x 9.6. FIG. 2. Microsporophyll from a more central region of the corm showing velum and ligule which still retains a portion of the apex of the central region, x 9.6. FIG. 3. Cellular detail from outer sporangial wall of a megasporangium, x 396. All drawings made with a camera lucida drawing tube. R. J. HICKEY: NEW ISOETES FROM JAMAICA 73 FI 4-9. peels of the types of /soétes jamaicensis (Proctor 34357, IJ) and J. montezumae (Pringle IGS. 3459, MO). FIG. 4. Megaspore of /. jamaicensis, equatorial view, F Megaspore of /. jamaicensis, close up of a partially FIG. 7. gen of J. montezumae, close up of a _ 8. Microspore of /. jamaicensis, equatorial 95 x 125. FIG. 5. Megaspore of /. montezumae, distal view, * 125 developed tubercle, distal surface. 2500. partially splape er tubercle, distal surface, < 2500. view, X 1665. FIG. 9. Microspore of /. montezumae, near proximal view, 74 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Whether the corms break their dormancy in drier years is not known, but observa- tions on other species of similar habitats suggest that they may not (Hall, 1971). Considering the unusual habitat and apparent sporadic growth of this plant, it is not surprising that it has eluded detection this long. It will not be surprising if this species is collected in other areas of Jamaica or perhaps on some of the other islands of the Greater Antilles. LITERATURE CITED EATON. A. A. 1897. A new quillwort from Mexico (1. seaegn Fern Bull. 5:25. HALL, J. B. 1971. Observations on Isoetes in Ghana. Bot. J. Linn. Soc. 64:117-139. PFEIFFER. N. E. 1922. Monograph of the Isoétaceae. a ar Bec Gard. 9:79-232, pl. 12-19. REED, C. F. 1953. Index Isoétales. Bol. Soc. Brot. 27:5 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 3 (1981) 75 Leaf Turnover Rates and Natural History of the Central American Tree Fern Alsophila salvinii RALPH CL. SEILER* There are few studies of growth rates of tropical trees. Moreover. the growth of tree ferns in their native habitats has seldom, if ever, been studied because of their slow growth and the need to make observations over a long period of time. As a Peace Corps volunteer in El Salvador I had the chance to study the growth of the cloud forest tree fern Alsophila salvinii Hooker (Cyatheaceae). Bosque Montecristo is a cloud forest that covers about 15 km* and is located at the common borders of El Salvador, Guatemala, and Honduras. Approximately 60 percent of the forest is in El Salvador and has been set aside as a national park and wildlife preserve. The highest point in the forest, the summit of Cerro Montecristo, has an altitude of 2414 meters. Bosque Montecristo receives an average of 2250 mm of rain annually. Fog drip during the night deposits about the same amount of water on the forest (Reyna, 1979). March and April are the driest months. The rainy season begins in mid-May or early June and continues until December. Bosque Montecristo is a subtropical, lower montane, very humid forest according to the Holdridge life zone classification (Holdridge. 1975). The cloud forest proper begins at ca. 2100 m and extends to ca. 2350 m, where it is replaced by an ericaceous shrub association on the mountain summits. The ecotone between the cloud forest and the pine-cypress association below occurs between 1900 and 2100 meters. Secondary forest is found in the whole zone wherever there has been logging (Reyna, 1979). The majority of the trees in the cloud forest are evergreen. The emergent trees in the forest canopy are principally oaks, which are thickly covered with epiphytes (Reyna, 1979). Tree trunks near the forest floor support lush growths of mosses, filmy ferns, and many other ferns such as Asplenium, Grammitis, and Elaphoglos- sum, Reyna (1979) reports 175 species of trees in the Montecristo area and 71 species from the cloud forest proper. In the area where the present study was undertaken three tree ferns are common: Dicksonia gigantea Karst.. Trichopteris schiedeana (Presl) Tryon, and Alsophila salvinii Hooker. Alsophila salvinii is a cloud forest tree fern known from southern Mexico, Guatemala, Honduras, and El Salvador (Stolze, 1976). It is one of the most conspicuous ferns in Bosque Montecristo and is the most common tree fern. In very wet parts of the forest, A. salvinii forms large, dense thickets, often in areas where a large tree or branch has fallen. The thicket shown in Fig. / is brightly lit during the morning from holes in the forest canopy. The thicket appears to be perpetuating itself, as there are hundreds of very young tree ferns present with trunks as yet undeveloped. Fronds of these are only about 30 cm long. whereas fronds of adults are about 2.5 m long. ; Alsophila salvinii is also found in relatively dry, open secondary forest at ca. 2050 meters in Bosque Montecristo. The ferns in the secondary forest don’t form dense *3977 S. 775 W., Bountiful, UT 84010. FIG. 1. The thicket of tree ferns in ae Sk ee which the study was undertaken. FIG. 2. Alsophila salvinii showing persistent stipes a nd =! 3 rachises of dead fronds. (1861) TZ SWMIOA “IWNYNOF NY34 NVOINIWY R. L. SEILER: LEAF TURNOVER OF ALSOPHILA SALVINII 77 thickets, are less common, and in general appear less robust than the ferns of the cloud forest proper. Young tree ferns are not seen in the secondary forest. The absence of young tree ferns in the secondary forest might be partially due to grazing by cattle; cattle from a nearby pasture have been observed in the forest eating fiddleheads from the crowns of adult tree ferns. The fronds of A. salvinii are tripinnate, and when the tronds die only the secondary pinnules fall; the stipe and main rachis remain attached to the trunk for several years (Fig. 2). The base of the stipe is closely appressed to the trunk for about 15 cm and then arches out to support the frond. Eventually the dead rachis and most of-the stipe will break off, but the appressed stipe bases remain attached to the trunk. The bases decay slowly and for many years will hold water either from the roots or captured from rainfall and condensation. The outer trunk of A. salvinii is not thickly covered with adventitious roots as it is in some tree ferns such as Dicksonia gigantea. The goals of this study were to investigate frond production rates, death rates, and lifespan; frond phenology; the relation between trunk length and tree age: the effects of water stored in the appressed stipe bases on frond production and death rates: and the factors affecting tree mortality. METHODS In late June 1978, twenty-eight mature, healthy trees of Alsophila salvinii: were chosen from a dense thicket (Fig. /) in very wet forest at 2300 m altitude on Cerro Montecristo and were tagged with forester’s flagging tape. All fiddleheads on these trees were separately tagged and their subsequent development followed. All dead fronds were cut from the tree ferns, but the appressed stipe bases were left intact. The mean trunk length of the 28 selected ferns was 1.18 m (s=0.47). The terns were randomly divided into equal control and experimental groups. Holes were cut in the bottoms of the dead stipe bases of the trees in the experimental group with a pocket knife so that they could no longer hold water. At irregular intervals over the next two years, eleven visits were made to Bosque Montecristo to observe the tree ferns. All mature, immature, and dead fronds present on each fern were counted. Fiddleheads and young fronds in which the pinnules had not fully expanded were considered immature. All living fronds with fully expanded pinnules were considered mature. Because dead fronds were removed at the beginning of the study, the ferns had no dead fronds at the first observation. Obtaining an accurate count of the number of fronds that died over the period of the study would have been difficult if dead fronds fell without leaving a trace; this was not a problem, however, because dead fronds remained attached to the trunk. RESULTS AND DISCUSSION Means of mature, immature, and dead fronds per tree in the control group during the two year period from late June 1978 to May 1980 are shown in Fig. 3. There is a rapid flush of fronds produced starting at the end of the dry season in early May. The immature fronds of A. salvinii do not develop synchronously. The first fiddlehead of the season may be nearing maturity as the last fiddlehead just barely CVG) | (14) (12) (13) (10) (13) C13}. C08} (13) (14) HME MATURE FRONDS JE IMMATURE FRONDS 2K DEAD FRONDS MEAN WKUMBER OF MATURE, IMMATURE AMD DEAD FROMDS PER TREE : ; i 24 / / / / : / / / tbl as 0 4 T T T T t TT 1 Sah cus T qT T 1 & i T T T T 7 af Fa oa T z J A $ 0 N D J F M A M J J A 5 0 N 1) J F M A M J MONTH 1978 ‘Beet 1979 et 1980. ————-——») _The mean number of mature, immature, and dead fronds per tree of Alsophilia salvinii in the control group during the period June 1978 to May G. 1980. eo bars indicate 95% contidence limits for the means; number of individuals is shown in parentheses. (1861) TZ SWNIOA “IWNYNOF NY34 NVOIMSWY R. L. SEILER: LEAF TURNOVER OF ALSOPHILA SALVINII 19 protrudes from the crown. The asynchronous production of fronds by A. salvinii is not typical of all tree ferns; in some tree ferns such as Dicksonia gigantea, all the fiddleheads produced during one growing season develop and are mature nearly synchronously (personal observation). Throughout the year, even during the dry season, fiddleheads are present on a few ferns, but these normally just barely protrude above the trunk apex. In February and April 1979, several very young fiddleheads were tagged and the time required for their maturation was observed. Fiddleheads initiated in February, just before the driest part of the year, required 3-4 months to mature, and one of the tagged * fiddleheads died during this period. Many of the pinnae of the surviving fronds were undeveloped, dead, and dried out when observed in April. Fiddleheads tagged in April 1979, at the end of the dry season, required 2-3 months to mature and all survived. Frond death rate fluctuates seasonally (Fig. 3). Few fronds died during the dry part of the year (January to June). Nearly dead fronds with only a few pinnules remaining survived this entire period. In June, at about the time the pinnules of the fronds initiated in April began to expand, the frond death rate sharply accelerated and was quite constant from June to January. However, the end of the period of rapid frond death does not coincide with the time the fiddleheads were finally mature. On | April 1979, the ferns averaged 1.7 dead fronds per tree; one year later they averaged 4.7 dead fronds per tree (data interpolated from Fig. 3). This indicates an average frond death rate of three fronds per year per tree. During this same period, there was no net gain or loss of mature fronds, which together with the calculated death rate, indicates an average frond production of three fronds per tree per year. Since the number of mature fronds per fern averages about six, the death rate of three fronds annually implies a frond lifespan of about two years. The period trom April to April was used to measure net changes in the number of mature and dead fronds because April is in the middle of a 3—4 month period when there are no rapid changes in the number of mature and dead fronds. The lives of the 27 fiddleheads tagged in late June 1978 were followed for two years. By October 1978 all the tagged fiddleheads were mature. In February 1979, about half the fronds had begun to drop some pinnules. In July 1979, one year after the fronds were initiated, all those not killed by catastrophe were alive, although nearly all were dropping pinnules. By December 1979, one of the fronds was dead, but at the last observation in May 1980 (22 months after the fronds were initiated) no more had died. It is probable that the fiddleheads tagged in June 1978 died during the 1980 wet season. If the number of fronds produced by a given length of trunk is constant, the age of a tree fern can be estimated. To determine the number of fronds produced in 50 cm of trunk, ten ferns not previously included in the study were selected from the same thicket. The number of stipe bases between 20 and 70 cm below the crown was counted on the ten tree ferns; the mean was 18.1 fronds/ 50 cm (s= 1.59). Since about three fronds are produced annually, about 12 years are needed to grow a meter of trunk under the wet conditions of the primary forest. The growth rate may be different in the drier secondary forest. The exact age of a tree fern cannot be 80 AMERICAN FERN JOURNAL: VOLUME 71 (1981) calculated from its trunk length because neither the time needed for establishment nor the rate of frond production in very young individuals is known. The lifespan of some of the tree ferns may be in excess of 50 years. In the primary forest, one vigorous fern with a trunk 4.6 meters long was observed; assuming a growth of | meter/I2 years, the plant was about 55 years old. Such longevity apparently is not common. Of 92 healthy adult tree ferns measured in both the primary and secondary forests, only one other trunk over 4 meters long was noted and the third largest was only 3.3 m long. The median length of the 92 tree trunks was |.3 meters. One of the most common natural catastrophes in the forest is the falling of large epiphyte-covered trees or their branches. Winds in the forest can reach 80 kph (Reyna, 1979). In February 1979 a large oak branch fell at the edge of the thicket and completely crushed and broke one of the tagged tree ferns: although it did no harm to the trunk of another fern, all the fronds except one immature frond were broken off. The ferns seem to be able to withstand extraordinary damage without being killed. Few of the tree ferns in the thicket have erect trunks, most are inclined away from the vertical, and some nearly lie on the ground. Probably this is caused by falling branches or trees knocking the ferns over. The ferns recover from being knocked over by turning the crown and growing upright (Fig. /). One fern was seen which had had its direction of growth radically changed at least twice, and possibly three times, with no apparent ill effects. Water is retained in the persistent petiole bases of the ferns during the entire wet season and part of the dry season. It was originally thought that this might be advantageous to the ferns during the dry season, perhaps providing additional water trapped from rain or condensation. Observations indicate otherwise; the tree ferns which could not store water in their punctured petiole bases did not differ significantly from the control in number of fronds produced annually or in the frond death rate. The results might have been different in the drier secondary forest, where the ferns are less vigorous. Possible reasons for tree fern mortality were investigated. Falling branches kill some trees; nevertheless, being knocked over or the loss of fronds may not kill the fern. The tree that lost all except one immature frond subsequently developed normal fronds. No sign of insect damage to fronds was ever observed. Two fronds on one tree fern were stricken by a blight; normal fiddleheads developed into twisted, stunted fronds with only a few functional pinnules. This did not kill the fern during the period of the study, and no other fronds seen were affected by a similar blight. Relatively few dead tree ferns were seen, as the soft wood of the trunk probably decays rapidly in the wet conditions of the forest. Nevertheless, several dead individuals were observed in which the trunk diameter was sharply constricted at the crown. Living trees with this syndrome were easily recognized by their extremely small fronds, which usually were only 40-50 cm long. The crowns of several of these living trees were examined by splitting them open; there was no evidence of rot or decay in the crown or of damage to the base or roots of the trees. It may be that the constricted trunk and the very small fronds are symptoms of senescence in R. L. SEILER: LEAF TURNOVER OF ALSOPHILA SALVINII 81 the trees. Further research on these ferns could provide evidence supporting the hypothesis that death from old age is fairly common in the population. Data supporting the hypothesis would include a relatively high frequency of ferns with this syndrome occuring in the population and evidence that these trees are old compared with the population median. ACKNOWLEDGMENTS Agradezco sinceramente a la Unidad de Parques Nacionales y Vida Silvestre del Ministerio de Agricultura y Ganaderia y a la Fundacion Freund por el apoyo recibido para la realizaci6n de este estudio. Expreso mi aprecio a la Lic. Maria Luisa Reyna V. por su valiosa ayuda con la descripcion del bosque y ademas a la Lic. Kathy DeRiemer, Lic. Dennis Witsberger y Senor Gabriel Calderon Hernandez quienes hicieron las observaciones en la parte final del estudio por encontrarme afuera de El Salvador. Asimismo a los Sefores Amadeo, Julio y Ricardo Martinez. encargados de los jardines de Los Planes de Montecristo; al Senor Juan Escobar, motorista de Parques Nacionales; y al Senior Pedro Hernandez y los demas vigilantes por su plena cooperacion en el trabajo del campo. LITERATURE CITED HOLDRIDGE, L. R. (Consultor). 1975. Zonas de Vida Ecoldgica de El Salvador: informe preparado r la Organizacién de la Naciones Unidas para la agricultura y la alimentacion. PNUD/FAO/ELS/73/004, Documento de trabajo No. 6. San Salvador, EI Salvador. REYNA V., M. L. 1979. Vegetacién Arborea del Bosque Nebuloso de Montecristo. Thesis (Licenciatura en Biologia), Facultad de Ciencias y Humanidades, Universidad Nacional de El Salvador. San Salvador, El Salvador. 177 pp. STOLZE. R. 1976. Ferns and Fern Allies of Guatemala. Part 1. Ophioglossaceae through Cyatheaceae. Fieldiana Bot. 39:1—130 82 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 3 (1981) Nomenclatural Notes on Micronesian Ferns F. R. FOSBERG and M.-H. SACHET* Three problems concerning the names of ferns found in Micronesia have arisen during the preparation of a geographical checklist of Micronesian pteridophytes. Our solutions to these problems are as follows. In addition, a new synonymy is given for the western Pacific species Thelypteris immersa. List gris guamensis CHont: ) Fosb. & Sachet, comb. O77 ay PE23 mi-S o hte at. on Agat-Umatac — m, “Marin ana Is Sands by ssnull stream. Grether 4354 (BISH not seen; isotype US). PARATYP Ercthce 4385 (US Since we ie the genus Thelypteris Schmidel in a broad sense, including such segregates as Christella Léveillé, it is necessary to provide the appropriate combina- tion for C. guamensis in Thelypteris. Wagner and Grether (Occ. Pap. Bishop Mus. 19:56. 1948) list as Cyclosorus dentatus (Forsk.) Ching the two collections cited by Holttum (Grether 4384, 4385). We could not match these exactly in the consider- able series of specimens from Guam labelled 7. dentata or T. parasitica. The thickened, reddish, glandular hairs scattered over the underside of the lamina, the pubescence of two ranges of length, the broad fronds with the lowest pinnae only slightly or not reduced, and the short-creeping rhizome would seem to place T. guamensis between T. dentata and T. parasitica. We follow Holttum in recognizing T. guamensis as a separate species, but with some misgivings. The Guam plants of this affinity do not sort well into three, and even less well into two, populations. BE 9 = zt. Ds aa z AS = S as) = S = 3 is = = a oe co = = oO & ms = = =5 3 ho tO S) Thelypteris immersa (Blume) Ching, Bull. Fan Mem. Inst. Biol. 6:306. 1936. Aspidium immersum gon Enum. Pl. Javae 156. 1828. TYPE: Gaenaeng Parang, Java. Blume (L not seen, Morton photo 116 Drvopteris immersa oe ca Kuntze, Rev. Gen. PI. 2:313. 1891; Hosokawa. Trans. Nat. Hist. Soc. Formosa 28:147. | eatin en see Hosokawa, Trans. Nat. Hist. Soc. Formosa 32:285. 1942. TYPE: Peliliu Island, Palau, Hosokawa 922/ (TAI not seen), Amphineuron immersum (Blume) Holtt. in Nayar & Kaur, Comp. Bedd. Handb. 203. 1974. Thelypteris peliliuensis Fosb. Smiths. Contr. Bot. 45:4. 1980, nom. nov. for G. palauensis Hosokawa, non T. palauensis (Hosokawa) Reed. Professor R. E. Holttum (in /itt. 1 Mar 1981) informed us that he studied the type of G. palauensis and found it to be his Amphineuron immersum, a widely distributed East Asian and Western Pacific species. Hosokawa himself had reported this earlier as Dryopteris immersa. Since we regard Amphineuron as a part o Thelypteris, T. peliliuensis must go into synonymy under 7. immersa. Trichomanes falsinervulosum (Nishida) Fosb., Smiths. Contr. Bot. 45:4. 1980. Mic Secs falsinervulosum Nishida, J. Jap. Bot. 32:156. 1957. TYPE: Aimiriik. Babeldaob. Palau. 8 Sept . Tsuyama (TI not seen). This. . was made after checking recent indices and publications on co of Botany, National Museum of Natural History, Smithsonian Institution, Washington, FOSBERG & SACHET: NOMENCLATURAL NOTES ON MICRONESIAN FERNS 83 Malesian and Micronesian ferns. After its publication, Dr. Lellinger called our attention to C. V. Morton’s combination 7. “falsivenulosum” (Nishida) Morton (Contr. U. S. Natl. Herb. 38:192. 1974), based on Microgonium “falsivenulosum™ Nishida (J. Jap. Bot. 32:156. 1957). We find no such name published by Nishida: because the references are identical, it seems likely that Morton misread the epithet or unintentionally changed the epithet to one that meant exactly the same thing. If this were an obvious typographical error or a simple orthographic error that involved only a connecting vowel or similar change, it could be regarded under ICBN Art. 73 as a correctible error. However, when the change produces a different Latin root which could properly be used in this position in an epithet, it seems most advisable to treat the two as different epithets and to index the resulting combina- tions as different names. This will save future workers from rechecking to account for the difference. The present wording of Art. 73 does not make it clear where the line should be drawn between correctible and non-correctible errors. Therefore. one’s best judgment must be followed. In this case we choose to use the 1980 combination. Trichomanes motleyi (v. d. Bosch) v. d. Bosch, Nederl. Kruid. Arch. 5(2):145. 1861. Microgonium motleyi v. d. Bosch, Hym. Jay. 5. t. 1. f. /. 2. 1861. SYNTYPE: “Hab. Insula Borneo (pr. Laboan), Motley No. 203 (comm. ill. W. J. Hooker)” (L not seen). SYNTYPE CITED WITH T. MOTLEYI: “Hab. Ins. Borneo (pr. Laboan) Motley (H. Hook.)” (L not seen). In verifying the original publication of this species, a Curious circumstance was noticed. Van den Bosch published M. motleyi and T. motleyi independently in different publications in the same year, each with a slightly different description and statement of type material, and each with no reference to the other. No firm evidence is available to us as to which publication is earlier, other than that the “Hymenophyllaceae Javanicae” has an introduction dated “Mart. 1860” and a prefatory note in the other publication is dated “Aug. 1860.” Presumably Motley’s two sheets are different sendings of the same material, although they could be different collections from the same place. Since van den Bosch published the two species only a few months apart, it is likely that he had both sheets at hand while both publications were being prepared. Therefore, both collections should be considered as syntypes. In response to an inquiry directed to the Leiden Herbarium concerning the Motley specimens, G. J. de Joncheere informed us (in litt. 17 Oct 1980) that one sheet is labelled in van den Bosch’s hand “Trichomanes motleyi v. d. B. from Laboan (Borneo) no. 203 Mr. Motley ad truncos arborum.” This sheet bears two small specimens which are, according to de Joncheere, apparently the plants from which the excellent illustrations in the “Hymenophyllaceae Javanicae” were made. Attached to the other sheet 1s a descrip- tion written by van den Bosch which is similar to but not identical with the two published descriptions, which are themselves similar but not identical. Interestingly, none of the annotations on either of these sheets mentions Microgonium; all refer to Trichomanes motleyi. We here designate the sheet with the description as the lectotype of M. motleyi. 84 AMERICAN FERN JOURNAL: VOLUME 71 (1981) As to the exact status of the two names, we can do no better than to quote from de Joncheere’s letter: From the above you will see that indeed the names Microgonium motleyi and Trichomanes motleyi are independently published, strictly speaking. However, as the dual names originated at about the same time from the same man from the same type and were apparently just the result of a wavering mind, one cannot regard these names as truthfully unconnected: on the contrary these publications are very much connected and should in my opinion be treated as the publication of a new species and a new combination. depending on what exact date of publication of the V. D. B. articles in question can eventually be ascertained. We are prepared to accept this well considered opinion. Since the scanty evidence from the introductory statements indicates that M. motleyi may have appeared earlier, we treat the 7richomanes as a transfer. Copeland (Phil. J. Sci. 51:201—202. 1933; 67:62. 1938), in his major works on Hymenophyllaceae, used both names without raising any question of the propriety of their publication. his species was reported as Microgonium motleyi by Ito (Bot. Mag. Tokyo 67:219. 1954) from Aimiriik, Babeldaob, Palau, Caroline Islands, Okabe /6 (TI not seen). REVIEW “SINOPSIS DE LAS ESPECIES DE LYCOPODIUM L. (LYCOPODIACEAE PTERIDOPHYTA) DE LA SECCION CRASSISTACHYS HERTER,” by Cristina H. Rolleri. Revista del Museo de La Plata, n.s. 13:61-114. 1981.—1In the first half of this century, Herter and Nessel published several papers and one book (“Die Barlappgewachse”) which thoroughly confounded the taxonomy and nomenclature of Lycopodium. So complete was the chaos introduced by these authors, especially in the tropical members of the genus, that subsequent botanists often have avoided studying them. The synopsis now published by Dr. Rolleri goes a long way toward resolving the problems that were introduced in sect. Crassistachys, a largely terrestrial group of mostly Andean species. Dr. Rolleri has studied the anatomy and morphology of the plants in great detail, which is evident from her key to the 56 species of the section. A synonymy, statement of habitat and range, and brief description are given for each species. A rather long list of excluded names, many of which are of uncertain application because the author was not able to obtain type material, indicates that there is yet more work which must be done before the taxonomy and nomenclature of this section of Lycopodium can be completely understood, but the progress in the present synopsis is very great and will be most helpful to all who must deal with these confusing plants.—D. B. L. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 3 (1981) 85 x Asplenosorus shawneensis, a New Natural Fern Hybrid Between Asplenium trichomanes and Camptosorus rhizophyllus ROBBIN C. MORAN* The Appalachian spleenworts constitute one of the most interesting and diverse groups of ferns in North America. They are popular with amateur fern growers due to their beauty and ease of culture. To professional botanists, the study of these small ferns has contributed to advances in cytology, chemical systematics, reticulate evolution, hybridization, and ecology. Knowledge of the Appalachian Spleenwort complex is still far from complete: indeed, new hybrid ferns and cytogenetic races have only recently been described. The purpose of this paper is to describe a new, naturally occurring spleenwort hybrid. In October, 1979, while botanizing in the Shawnee Hills of southern Illinois. I discovered one plant of a hybrid between Asplenium trichomanes and Camptosorus rhizophyllus. The hybrid was found on a north-facing crevice in a sandstone canyon where a large number of the parent species were present. The elongate-triangular shape of the blade at once suggested C. rhizophyllus; however, the base of the blade was cut into opposite, pinnate lobes that looked very much like enlarged basal pinnae of A. trichomanes. The stipe resembled that of A. trichomanes due to its narrow diameter and dark brown color that extended a short distance up the rachis (Figs. / and 3). Furthermore, the frond venation was both dichotomously free branched. as in A. trichomanes, and anastomosing, as in C. rhizophyllus (Fig. 3). The suspected hybrid was removed and brought to Southern I[Ilinois University. where it was grown for further stud To gain further evidence of the plant’s hybrid nature, alpha and beta esterase isoenzymes were studied using starch gel electrophoresis in the suspected hybrid and the putative parents. It was hypothesized that the suspected hybrid would show isoenzyme banding patterns present in both parents and perhaps intermediate bands as well. The electrophoretic system followed was that of Steiner and Johnson (1973). Their instructions for specific esterase stains and gel and electrode buffer were used exclusively. Only mature blade tissue was ground. This eliminated variation in the results which could be caused by using fronds of different ages. Plants of A. trichomanes and C. rhizophyllus used for the analysis were collected at the same site where the hybrid was found. Electrophoresis was run until a standard marker dye had migrated about 3.5 cm. After staining the starch gel. it was possible to differentiate the beta esterase enzymes that stained pink—purple from the alpha esterases that stained black (Brewer, 1970, p. 88). The resulting zymograms were wrapped with clear plastic wrap, stored in a refrigerator at 1° C, and later photographed with Ektachrome film. Several replicates were made to confirm the validity of the results. *Department of Botany, Southern Illinois University. Carbondale. IL 62901. Present ron Illinois Natural History Survey, Natural Resources Bldg., 60 )7 E. Peabody Dr.. Champaign. IL 6182¢ AMERICAN FERN JOURNAL: VOLUME 71 (1981) An Te 1 2 | ic . i — — “5 “ Ce a aye ad “te, om, 7 AEN : a A “hall ea +») we ed Ag aed oa * ee . Wi tk ee ; ‘ he ® a et, r a % ae SS wa pa” eo ge ut t A Se - ¥ ft eile t¢ i Beet | ’ el pe —_ $*o, oo A + fn ee eh 8 e Pe Karn, py bos " pes 2 . ‘ : F a “ ey : a, tl e f * : a * ¥ a, 4 Le i “as ” ’ ; — * bk > a a! AC gees % i % be a, we PT to : a ‘ye * ‘ " * i . 4 ight) e. unreduced spore mother cells (ri < oO 4 & FIG. |. Mature fronds cultivated from the type of x Asplenosorus shawneensis (Moran 1269. MICH). S. Wagner. nosorus Shawneensis showing 16 lar and 64 aborted spores (left). Photo by F. FIG. 2. Sporangia of x Asple R. C. MORAN: x ASPLENOSORUS SHAWNEENSIS, A NATURAL HYBRID 87 The hybrid combines the esterase isoenzyme banding patterns of A. trichomanes and C. rhizophyllus (Fig. 4). Isoenzyme band numbers seven and nine. present in C. rhizophyllus, were the only parental bands undetected in the hybrid: they may have been present in the hybrid but simply not seen due to their extreme faintness. Isoenzyme band number eight, present in A. trichomanes and the hybrid. is distinct from all other bands because it is a purple-staining beta esterase band. All other bands are black-staining alpha esterases. No intermediate bands were observed. The hybrid esterase isoenzyme banding patterns provide strong evidence that trichomanes and C. rhizophyllus are indeed the parents. F rhizophyllus trichomanes shawneensis ~ FIG. 3. Frond outlines and venation patterns drawn from cleared fronds of * Asplenosorus shawneensis and its parents. Esterase zymograms of two other common spleenworts in southern Illinois, x Asplenosorus pinnatifidus and Asplenium platyneuron, were made in order to ascertain if they were involved in the hybrid’s formation (Fig. 4). Both species produced banding patterns suggesting that they were not involved. x Asplenosorus shawneensis R. C. Moran, hybr. nov. Figs. 1 and 3 (L.) Link. Rhizoma breve, reptans; squamae nigricantes, clathratae. Frondes u humilo-patentes, caespitosae, sempervirentes, usque ad IS cm longae. Stipites filiformes, atrobrunnei, nitidi, usque ad 4 cm longi. Lamina leviter coriacea, usque ad I cm latae, marginibus non profunde crenatis. Rachis atrobrunnea usque ad 7 cm ad basim, viridis ad apicem. Sori usque ad 4 mm longi. 88 AMERICAN FERN JOURNAL: VOLUME 71 (1981) TYPE: East of Devil’s Kitchen Lake, Williamson County, Illinois, T10S, RIE, sec. 15. in crevice of a shaded, north-facing sandstone canyon, with Asplenium trichomanes and Camptosorus rhizophyllus growing abundantly nearby, 21 O 979, Robbin C. Moran 1269 (MICH Observations of the hybrid’s chromosomes showed 72 univalents at early meiotic metaphase. Thus, x Asplenosorus shawneensis is an allodiploid with the genomic constitution R'T2, with no pairing between members of the two different genomes (R and T are the respective genomes and superscripts the pairing controls). The spores of XA. shawneensis vary between 64 aborted spores and 16 unreduced spores per sporangium (Fig. 2). = 4r SLPS oT Se rs sitetetats oO 2 eaoet Ti 1 — merece : — trichomanes hybrid rhizophyllum = pinnatifidum = platyneuron FIG. 4. Esterase isoenzyme banding patterns obtained with starch gel electrophoresis in five spleen- worts, Relative darkness of the bands is indicated by cross-hatching. Band 8 is a beta esterase: all other bands are alpha esterases. The following key has been provided to distinguish x A. shawneensis from the similar Index to Volume 71 an? eg PAN 3 122 Erratum The American Fern Society Council for 1981 ROBERT M. LLOYD, Dept. of Botany, Ohio University, Athens, OH 45701. President DEAN P. WHITTIER, Dept. of Biology, Vanderbilt University, Nashville, TN 37235. Vice President MICHAEL I. COUSENS, Faculty of Biology, University of West Florida, Pensacola, FL 32504. Secretary JAMES D. CAPONETTI, Dept. of Botany, University of Tennessee, Knoxville, TN 37916. Treasurer JUDITH E. SKOG, Dept. of Biology, George Mason University, Fairfax, VA 22030. Records Treasurer DAVID B. LELLINGER, Smithsonian Institution, Washington, DC 20560. Journal Editor ALAN R. SMITH, Dept. of Botany, University of California, Berkeley, CA 94720. Memoir Editor JOHN T. MICKEL, New York Botanical Garden, Bronx, NY 10458. Newsletter Editor American Fern Journal EDITOR DAVID B. LELLINGER Smithsonian Institution, Washington, DC 20560. ASSOCIATE EDITORS DAVID W. BIERHORST Rt. 3, Box 188, Picayune, MS 39466. GERALD J. GASTONY Dept. of Biology, Indiana University, Bloomington, IN 47401. JOHN T. MICKEL New York Botanical Garden, Bronx, NY 10458. The “American Fern Journal” (ISSN 0002-8444) is an illustrated quarterly devoted to the general study of ferns. It is owned by the American Fern Society, and care: at the Smithsonian Institution, Washington, DC 20560. Second-class postage paid at Washin Claims for missing issues, made 6 months (domestic) to if sacs (foreign) after the date of issue, and the matters for publication should be addressed to the Editor Changes of address, dues, and applications for membership should be sent to Dr. J. E. Skog, Dept. of es George Mason University, Fairfax, VA 22030. rs for back issues should be midressed to the Treasure Sea inquiries concerning ferns should be addressed to ihe Secretary. Subscriptions $9.00 gross, $8.50 net if paid through an agency (agency fee $0.50); sent free to members of the American Fern Society (annual dues, $8.00; life membership, $160.00). Back volumes 1910-1978 $5.00 to $6.25 each; single back numbers of 64 pages or less, $1.25; 65-80 pages, $2.00 each; over 80 pages, $2.50 each, plus shipping. Back volumes 1979 et seq. $8.00 each; single back numbers $2.00 each, plus shipping. Ten percent discount on orders of six volumes or more. Library Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is Librarian. Members may borrow books at any time, the borrower paying all shipping costs. Newsletter Dr. John T. Mickel, New York Botanical Garden, Bronx, NY 10458, is editor of the newsletter “Fiddlehead Forum.” The editor welcomes contributions from members and non-members, including miscellaneous notes, offers to exchange or purchase materials, personalia, horticultural notes, and reviews of non-technical books on ferns. Spore Exchange Mr. Neill D. Hall, 1230 a 88th Street, Seattle, WA 98115, is Director. Spores exchanged and collection lists sent on reques Gifts and Bequests Gifts and bequests to the Society enable it to expand its services to members and to others interested in ferns. Botanical books, back issues of the Journal, and cash or other gifts are always welcomed, and are tax-deductible. Inquiries should be addressed to the Secretary AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) 97 Bog Clubmosses (Lycopodiella) in Kentucky R. CRANFILL* Distributed largely on the Atlantic and Gulf Coastal Plains, the southern members of Lycopodiella Holub (= Lycopodium subg. Lepidotis (Pal. Beauv.) Baker) are rare and local in the interior of the continent. Inland occurrences are scattered as far north as the middle Mississippi Valley, and often are markedly disjunct from populations on the coastal plain. The purposes of this paper are to record recent observations on this genus in Kentucky, to give some explanation of its recent appearance and spread, to describe a new hybrid, and to make a series of new combinations. t is now apparent that the genus Lycopodium in the broad sense may be divided into several smaller, more natural genera (Beitel & Bruce, 1980). To facilitate the use of the segregate genus Lycopodiella, the following new hybrid and combinations are necessary. Lycopodiella x brucei Cranfill, hybr. nov. Planta hybrida inter L. appressam et L. prostratam intermedia, a L. appresso microphyllis ciliatis et sporophyllis patentibus, a L. prostrato rhizomatibus crassiore et strobilis angustiore, minus quam 10 mm latis differt; sporae non abortivae. TYPE: Borrow pit ca. 200 m ENE of road KY-280, 0.6 mi from its junction with road KY-121, Calloway County, Kentucky, 15 May 1975, J.G. Bruce 76006 (MICH). Named in honor of James G. Bruce, III. Lycopodiella alopecuroides (L.) Cranfill, comb. nov. Lycopodium alopecuroides L. Sp. P|. 2:1102. 1753. Lycopodiella appressa (Chapman) Cranfill, comb. & stat. nov. Lycopodium inundatum var. appressum Chapman, Bot. Gaz. 3:21. 1878. Lycopodiella < copelandii (Eiger) Cranfill, comb. nov. Lycopodium X copelandii Eiger, Biol. Rev. City Coll. New York 18:21. 1956. Lycopodium inundatum var. elongatum Chapman, FI. So. States, ed. 2:671. 1883. Lycopodiella prostrata (Harper) Cranfill, comb. nov. Lycopodium prostratum Harper, Bull. Torrey Bot. Club 33:229. 1906. os _ Although much has been written about the Lycopodiella population in Kentucky (Johnson & McCoy, 1975; Bruce 1975; Cranfill, 1980), confusion still exists over the number and kind of taxa present. Johnson and McCoy reported L. appressa and L. prostrata from a large gravel pit in eastern Calloway County. Bruce’s (1975) extensive survey of the North American members of the genus includes comments on the Kentucky material, which he concluded is comprised of L. appressa, ae copelandii (L. alopecuroides appressa), and L. X brucei (as L. appressa X *Division of Biological Sciences, University of Michigan, Ann Arbor, MI 48109. Present address: Department of Botany, University of California, Berkeley, CA 94720. Volume 71, number 3, of the JOURNAL was issued September 30, 1981. 98 AMERICAN FERN JOURNAL: VOLUME 71 (1981) prostrata). For want of authentic L. x brucei, | admitted only the first two species to the flora (Cranfill, 1980). In an attempt to resolve these differences, I undertook fieldwork in Calloway County during the summers of 1979 and 1980. I recorded the topographic and ecological features of each site where Lycopodiella was found and made representative collections of each population for further analysis. Results of these studies are reported below. Reconnaissance on 10 Aug 1979 in the vicinity of the gravel pit where Johnson and McCoy had collected (Site 1) revealed three additional populations (see Fig. /) of L. appressa (Site 2, Cranfill 4730; Site 3, Cranfill 4731; Site 4, Cranfill 4738a). All sites are scattered along the bases of Cretaceous ridges composed of sand and partially consolidated gravel and are characterized by low-lying flats of sands with virtually no contamination from silt or organic matter. These plants seem especially sensitive to siltation, for no Bog Clubmoss sites have been found where siltation had occurred. Ground water in the area is soft and acidic and lies at or just below the surface of the flats for much of the growing season. The original pit is the largest and best developed of the four sites, has been abandoned for some time, and contains the most extensive populations of the clubmosses. Lycopodiella appressa is abundant, occupying nearly all the moist, open areas, whereas L. X copelandii is confined to the wettest spots, often around pools. The moss Aulacomnium palustre is an associate of both species, and Cladonia and Leucobryum albidum are scattered about on dryer hummocks. Vascular plant associates include species with decidedly southern affinities, e.g., Boltonia asteroides, Gymnopogon ambiguus, Scirpus koilolepis, Spiranthes odorata, Vaccinium atrococcum, Woodwardia areolata, and FIG. 1. Portion of south-central Calloway County, Kentucky, showing extant sites of Lycopodiella. R. CRANFILL: BOG CLUBMOSSES IN KENTUCKY 99 Xyris torta. Pinus taeda is naturalized along the banks above the pit. Sites 2-4 have been disturbed recently and contain only small populations of L. appressa. Few species other than the clubmosses have become established along the roadside ditches and sandy flats, although Boltonia was seen at Site 2. Winter and early spring visits reveal almost complete dieback of the clubmosses during the winter. Bruce (1975) reports the formation of “tubers” in L. alopecuroides and L. prostrata which serve as food storage organs for the following spring. I observed a similar situation in L. appressa at Site 1. Growth is initiated from the tips of these organs in early April. From these observations, it is evident that any management programs designed to protect these stations must (1) maintain the open nature of the site by controlling the invasion of woody vegetation, (2) attempt to maintain a high water table, which is important not only for reproduction but also perhaps in moderating the harsh winters, and (3) minimize siltation, which appears to have adverse effects on these ants. Although I have been unable to confirm the presence of L. X brucei at any of the sites out of ca. 100 individuals that I collected from all populations, recently | received herbarium material of this species (Bruce 76006) from Dr. J. G. Bruce. Since L. prostrata is the most southern of the species in the complex, it is plausible that it and its hybrids are less winter hardy than other taxa in the pits. Therefore, a combination of harsh winters in the late 1970’s and competition with other clubmosses may be responsible for the hybrid’s disappearance. On the other hand, hybrids which occur without one or both parents often have been explained by “long distance hybridization,” which may apply in this case. In Lycopodiella, hybrids between parents with the same chromosome number exhibit normal pairing of chromosomes at meiosis, and produce what appear to be functional spores (Bruce, 1975). If the spores be viable, the occurrence of isolated hybrids may represent the introduction of single spores from hybrid plants. Such a mechanism for hybrid reproduction and dispersal would be novel among the pteridophytes, which generally produce infertile hybrids that reproduce largely vegetatively. Although the bog clubmosses may have been present in the area for some time, the evidence points to a relatively recent introduction. In Calloway County, these plants occur in situations which have been created by man’s activities in the recent past. This also appears to be the situation with populations in west Tennessee. Natural habitat suitable to Lycopodiella is very limited in the northern portion of the Mississippi Embayment and is occupied quickly by the adjacent floodplain forests. It seems likely that these clubmosses have migrated from their metropolis to the south by hopping from one gravel pit to another along the Cretaceous hills in western Kentucky and west Tennessee. If migration is occurring still, close inspection of suitable sites to the north may reveal the presence of these taxa on Cretaceous outcrops in southern Illinois. I am indebted to M. E. Medley, Laurina Lyle, and D. M. Johnson for assistance in the field, to D. M. Johnson and J. M. Beitel for reading and commenting on the manuscript, and especially to J. G. Bruce, III who helped me in various ways. 100 AMERICAN FERN JOURNAL: VOLUME 71 (1981) LITERATURE CITED BEITEL. J. and J. G. BRUCE, III. 1980. Generic concepts in the Lycopodiaceae. Abstract. Bot. Soc. r. Misc. Ser. Publ. TEL. BRUCE, J. G., Ill. 1975. Systematics and morphology of subgenus Lepidotis of the genus Lycopodium. Unpublished Ph.D. dissertation, University of Michigan, Ann Arbor, MI. 15 CRANFILL, R. 1980. The Ferns and Fern Allies of Kentucky. Ky. Nat. Pres. Comm. Monogr. Tech. Ser. 1:1-284. JOHNSON. R. G. and T. N. McCOY. 1975. Some Lycopodiums new to western Kentucky. Amer. Fern J. 65:29 REVIEW FERNS AND FERN ALLIES in the “FLORA OF BAJA CALIFORNIA” by Ira L. Wiggins, viii + 1025 pp. Stanford University Press, 1980. ISBN 0-8047-1016-3. $65.00—This is the first publication to treat all the known ferns and fern allies of Baja California, Mexico. Sixty-five species and an additional five varieties are included in the treatment, which appears on pp. 51-71 of the book. Descriptions are provided for families and genera. The species and varieties are delimited through rather detailed keys which include comments on the species’ habitat, distribution in Baja California, and overall oe Line drawings illustrate one member of each genus treated. These are the work of the author in many instances; others are taken from L. R. Abrams’ “lllustrated Flora of the Pacific Coast States” (1923). The illustrations are nae of very high quality and are detailed enough to give an accurate representation of the species illustrated. Two exceptions are the illustrations of Woodsia plummerae and Equisetum laevigatum, which are not very useful below the genus level. The taxonomy used is conservative and traditional in treatment of families (e.g., Polypodiaceae s. lat. is used) and species and varieties. Three examples of outdated nomenclature to be corrected are: Pteridium entry corrected to Pteridium aquilinum (L.) Kuhn var. pubescens Underw.; Anemia anthriscifolia corrected to A. tomentosa (Savigny) Swartz var. mexicana (Presl) Mickel; and Pellaea longimucronata cor- rected to P. truncata Goodd. | have noted several cases of misspelling or incorrect citation of authors. The overall distributions given for the species vary considerably in completeness and accuracy. Several species should be added to the list of ferns known in Baja California (based upon recent literature reports): Bommeria pedata (Swartz) Fourn., Pellaea skinneri — and Pellaea seemannii Hook. (both are Cheilanthes, pers. obs. and R. M. Tryon, pers. comm., but there are not available names in Cheilanthes), Cheilanthes den Hook., and C. wootonii Maxon. In addition, I have seen specimens of C. eatonii Baker in Hook. & Baker from Baja California. This treatment of the ferns and fern allies of the exciting botanical region of Baja California is most welcome, as is the treatment of the entire vascular flora. Dr. Wiggins is to be commended for producing such a monumental work.—Timothy Reeves, Biological Science Center, Boston University, Boston, MA 0221 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) 101 Chain Ferns of Florida TERRY W. LUCANSKY* Woodwardia is a genus of rather large, terrestrial ferns with mostly ascending to erect rhizomes. The species native in the United States, however, typically have creeping rhizomes. The genus includes 11 or 12 species, is worldwide in its distribution, and. is one of a comparatively few genera characterized by boreal distribution (Copeland, 1947). Three species of Woodwardia, commonly known as chain ferns, occur in Florida. Woodwardia virginica and W. areolata are native species which occur primarily on the Atlantic Coastal Plain and extend from Florida to Texas and northward to Nova Scotia. Scattered inland localities also exist for both species. Woodwardia radicans is native to the Old World, but is cultivated and has reportedly escaped in peninsular Florida (Small, 1938). All three species possess distinctive leaf venation, with elongate areoles arranged in chain-like fashion along the costae and/or costules. In W. areolata, an extensive network of areoles extends to the margins of the leaf, whereas in W. virginica and W. radicans the veins are simple and forked beyond the areoles and are free at the margins. Elongate sori also are arranged in chain-like fashion along the costae and/or costules in each species. Woodwardia virginica, the Virginia or Giant Chain Fern, and Woodwardia radicans, the European Chain Fern, were originally placed in the genus Blechnum (Linnaeus, 1771), although Smith (1793) later transferred them to his genus Woodwardia, with W. radicans the type species. Subsequently, Pres! (1851) founded the genus Anchistea for W. virginica, although only the glandular indusia and the presence of a single row of areoles distinguish this species from the other species of Woodwardia (Morton & Neidorf, 1956). Today both species are typically included in the genus Woodwardia (Copeland, 1947; Wherry, 1964; Fernald, 1970; Lakela & Long, 1976). Some workers, however, still recognize the genus Anchistea (Radford et al., 1964; Small, 1938; McVaugh & Pyron, 1951). Woodwardia areolata, the Net-vein or Dwarf Chain Fern, was first named Acrostichum areolatum (Linnaeus, 1753). Smith (1793) included the species in Woodwardia (as W. angustifolia), although it was not until much later that Moore (1857) called the species W. areolata. Today some botanists (Fernald, 1970; Lakela and Long, 1976) use this name. Pres! (1851), however, established the genus Lorinseria, and many taxonomists (Copeland, 1947; Small, 1938; McVaugh and Pyron, 1951) consider L. areolata to be the correct name for the species. Wherry (1964) felt that the completely areolate venation and marked dimorphism of this species justified its segregation into a separate genus. Much morphological data are available for these species (Shaver, 1954; Wherry, 1964: Small, 1938; Fernald, 1970), but comparative anatomical data are almost totally lacking. In this study, the anatomy of W. virginica and W. radicans is compared with that of W. areolata, and these data are correlated with the two taxonomic systems currently in use for these species. Plant materials of W. virginica and W. areolata were collected in Alachua County, Florida. Plants of W. radicans were obtained from the Strybling Arboretum in San *Department of Botany, University of Florida, Gainesville, FL 32611. 102 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Francisco, California. The plant parts were killed and fixed in formalin-acetic acid-alcohol (FAA), dehydrated in a tertiary-butyl alcohol series and infiltrated with Tissuemat (Johansen, 1940). Sections (8 4m) were stained with a safranin-fast green series and photographed with a 35 mm Zeiss C35 camera. Habit photographs were taken with a 35 mm single-lens reflex camera. FIG. 1. Prostrate underground rhizomes of W. areolata (above), x 0.4 and W. virginica (below), 0.5 Note adventitious roots. FIG. 2. Erect stem of W. radicans, X0.9. Note large mass of adventitious roots. RESULTS AND DISCUSSION Both Woodwardia virginica and W. areolata occur in wet pinelands, bogs, marshes, and alongside streams and roadside ditches, whereas W. radicans is cultivated in fern gardens and has reportedly escaped to swamps and hammocks in the state. Woodwardia areolata and W. radicans thrive in partial shade around the bases of trees, but W. virginica does best in sunny locations and can tolerate a lower degree of acidity in the soil (Cody, 1963; Wherry, 1964) Both W. virginica and W. areolata have a branched, creeping rhizome (Fig. /), whereas W. radicans has an ascending to erect stem (Fig. 2). The stems of W. virginica (x = 10 mm) and W. radicans (x = 18 mm) are stouter than the rhizome of W. areolata (x = 5 mm) and more deeply buried in the soil. Golden brown, elongate scales with either blunt, rounded, or acuminate apices occur on the rhizome and petiole bases of W. virginica and W. radicans. Pale brown, cordate scales with acuminate apices are found on the rhizome and petiole bases of W. areolata. Shaver (1954) reported brown-black, oblong scales with acute apices in W. virginica and brown, ovate scales on the rhizome of W. areolata. I found small, light-brown, subcordate to cordate scales with acuminate apices abaxially on or proximal to the T. W. LUCANSKY: CHAIN FERNS OF FLORIDA 103 rachis and midveins of all three species. Woodwardia virginica has rather large, subcoriaceous, pinnate leaves, whereas W. radicans possesses large, coriaceous, pinnate-pinnatifid leaves with scaly buds produced on the rachis at the bases of the upper pinnae. Woodwardia areolata, however, has marked dimorphism, with the taller fertile leaves typically found in the summer and fall, distinct from the pinnatifid sterile leaves. Stem transections of the three species reveal similar anatomical features, although differences are noted (Fig. 3-5). A single-layered epidermis composed of elongate, bulbous, or irregularly shaped, thick-walled cells is partially sloughed off in mature sporophytes of both W. virginica and W. areolata, whereas the epidermal layer is typically intact and composed of small, thick-walled cells filled with tanniferous substances in W. radicans. A hypodermis composed of two zones occurs in the three species, although the outer zone may be sloughed off in W. virginica and W. radicans. In W. areolata, the outer zone consists of irregularly shaped, thick-walled parenchyma cells (Fig. 6), whereas in W. virginica and W. radicans the parenchyma cells of this zone are thicker-walled and more heavily lignified. The inner zone in all species is composed of sclerified, thick-walled parenchyma cells that are filled with tannins or contain starch grains (Fig. 6), although the zone is less extensive in W. areolata than in the other two species. In all three species, the bulk of the stem is composed of ground tissue (Figs. 3-5). The cortical region is distinguishable from the pith region primarily by the position of the meristeles. The cortex is composed primarily of irregularly shaped, thin-walled parenchyma cells filled with tanniferous substances, starch grains, or both. The pith region is composed of large, thin-walled parenchyma cells filled primarily with tannins, although starch grains are occasionally found in these cells. In W. virginica, thick-walled parenchyma cells occasionally comprise this region. In both W. virginica and W. radicans, certain cells in the cortical and pith regions enlarge singly or in groups, and tanniferous substances partially or totally fill the lumens of these cells (Fig. 7). These cell contents are abundant in W. radicans, especially around the meristeles, but are lacking in W. areolata. Large intercellular spaces occur in the cortical and pith regions of W. virginica and W. areolata and possibly serve as an adaptation to a marsh habitat. In W. radicans, very small intercellular spaces are occasionally noted in the ground tissues. The stelar pattern in all three species is a dictyostele with overlapping leaf gaps and variously-sized meristeles (Figs. 3-5). Each meristele is an amphicribral bundle delimited by an endodermis with distinct Casparian strips (Fig. 8). A pericycle of thin-walled parenchyma cells (1- 4 layers) filled with tannins completely encircles the primary phloem. The latter tissue is composed of sieve cells and phloem parenchyma, although protophloem and metaphloem are typically indistinguishable in mature stems. The primary xylem of all species is composed primarily of scalariform-pitted tracheids, although spiral and annular-thickened protoxylem ele- ments do occur. Interspersed among the xylary elements are thin-walled parenchyma cells filled with tannins. Xylem maturation is mesarch. In all three species, the sterile leaves exhibit similar anatomical features (Figs. 9-11), whereas the fertile leaf of W. areolata usually is reduced, although it may be 104 AMERICAN FERN JOURNAL: VOLUME 71 (1981) FIG. 3. Stem transection of W. nee x10. Note meristeles. FIG. 4. Stem transection of W. radicans, X10. Note dictyostele. FIG. 5. Stem transection of W. areolata, x 10. Note meristeles a dictyostelic pattern. FIG Two-zon hs hypodermis of W. areolata, x2 FAG: Teste seninestils substances in pith region of W. radicans, x91 . 8. Individual meristele of W. areolata, Pid. 9. Transection of leaf of W. areolata, x 245. extensive ae pesge layer. 10. Transection of leaf of W. ote tannin-filled palisade layer.. FIG. . Transection of leaf of W. radicans, x 295. Note extensive tannin-filled wee layer. The pomhens are: abe = abaxial epidermis, ade = adaxial epidermis, c = cortex, e = endodermis, g = gap, h= hypodermis, iz = inner zone, m= meristele, oz = outer zone, p=pith, pe= ewe ph= coldiens pl =palisade layer, s = stomate, sm=spongy mesophyll, t=tannins, and x= xylem. ote amphicribral arrangement of vascular tissue Note reduced palisade layer an virginica, X 279 T. W. LUCANSKY: CHAIN FERNS OF FLORIDA 105 partially sterile (Halsted, 1899) or even may resemble a fertile leaf (Waters, 1903). In each of the three species, the single-layered adaxial epidermis is composed of variously sized cells, is covered with a thin cuticle, and lacks stomates. The mesophyll is differentiated into palisade and spongy mesophyll layers, although the palisade layer is much reduced and less extensive in W. areolata than in the other two species (Figs. 9-11). The palisade layer is composed of loosely arranged, irregularly shaped chlorenchyma cells in W. areolata; a more compact and tannin- filled palisade layer characterizes W. virginica and W. radicans. Although Payne and Peterson (1973) noted an abaxial hypodermis in the leaves of W. virginica, none was found in the present study (Fig. 10). The spongy mesophyll in all three species consists of loosely-arranged chlorenchyma cells with numerous, large intercellular spaces. This layer is much more extensive and constitutes a greater proportion of the leaf in W. areolata than in the other two species (Figs. 9-//). In W. radicans, the spongy mesophyll layer is greatly reduced and contains small intercellular spaces. Depending upon the species, the spongy mesophyll cells are either larger (W. areolata) or smaller (W. virginica and W. radicans) than the cells of the palisade layer. The chlorenchyma cells of the mesophyll in the latter two species are typically filled with tannins, whereas this ergastic substance is less evident or lacking in the mesophyll cells of W. areolata (Figs. 9-11). In all species, the abaxial epidermis is composed of variously shaped cells and possesses anomocytic stomates (Fig. 12). Typically two or three epidermal cells abut the guard cells, with one cell nearly surrounding the stomatal apparatus. Chloroplasts are infrequently noted in the adaxial epidermis of W. areolata, but do not occur in the epidermal layers of the other two species. The presence of chloroplasts in the epidermal cells of land plants is correlated with a deep-shade habitat (Sculthorpe, 1967), and their occurrence in W. areolata may be an adaptation to the shaded habitat of this species. The mid-vein (costa) in all three species is surrounded by a bundle sheath of thick-walled parenchyma cells and is delimited by a single-layered endodermis with distinct Casparian strips (Figs. 13 and 14). The innermost layer of the sheath may have tanniferous substances in W. virginica and W. radicans, whereas these cell contents are lacking in W. areolata. A pericycle composed of thin-walled parenchyma cells encircles the vascular tissues. The primary phloem is composed of sieve cells and phloem parenchyma. The primary xylem is U- or V-shaped and consists primarily of scalariform-pitted metaxylem elements. Protoxylem is generally restricted to a median position in the concavity of the xylary elements. The fertile leaves of W. areolata show a much reduced structure; no differentiation of the mesophyll occurs, and only a few small intercellular spaces are noted. Stomates are lacking in both epidermal layers of a fertile leaf. Ground tissue comprises the bulk of the petiole base in the three species (Figs. 15-17). This tissue is aerenchymatous and consists of either thick-walled (W. virginica) or thin-walled (W. areolata) parenchyma cells with numerous intercellular spaces. In W. radicans, this tissue consists of thin-walled parenchyma cells or is composed of two zones in the larger petioles (Fig. 17). The outer zone consists of thick-walled parenchyma cells, and the inner zone is composed of thin-walled parenchyma cells. Numerous, small intercellular spaces occur in these zones, and 106 AMERICAN FERN JOURNAL: VOLUME 71 (1981) the cells are filled with tannins. Occasionally the parenchyma cells of the ground tissues in W. radicans enlarge and possess tanniferous substances that partially or totally fill the cell lumen. A single-layered epidermis characterizes the petiole and a hypodermis is found in all three species (Figs. /5—/7). The latter tissue is much more extensive in W. radicans than in the other two species. The number of vascular strands at the base of the petiole differs in each species In W. virginica, typically five small spherical strands and two large U- or V-shaped strands occur (Fig. /6), although a total of five or six strands has been noted. 12. Anomocytic stomates of W. virginica, x 225. FIG. 13. Transection of midvein (costa) of W. ‘ . Transection of midvein (costa) of W. virginica, x 100. substances in bundle s heath cells: FIG? 1S, small, spherical abaxial b ede ie Note Saas ak Transection of petiole base of W. areolata, x21. 2 two FIG. 16. Transection of petiole base of W. virginica, x2 yao a ee Transection of petiole base of W. eri x11. FIG. 18. Transection of adventitious root of W. virginica, X97. Note diarch pattern. FIG. 19. Shoot meristem of W. areolata, x2 IG. 20. Shoot of meristem of W. virginica, x 295. Note pamia apical cell. The abbreviations are: aoe apical cell, bs=bundle sheath, gt=ground tissu ner cortex, mx =metaxylem oc =outer cortex, ph = phloem, ps = petiole strand, px = hein s=stomate, x = xylem. T. W. LUCANSKY: CHAIN FERNS OF FLORIDA 107 Waters (1903) reported seven oval bundles in the petioles of W. virginica. In W. radicans, four or five petiole strands are noted in the petiole base (Fig. /7). Two or three small, spherical to oval bundles occur abaxially and two large crescent-shaped strands are found adaxially in the petiole (Fig. 17). In W. areolata, typically two large petiole strands are found in a median position in the petiole base, although occasionally four strands occur at a comparable level in the petiole (Fig. /5). Waters (1903), however, reported only two oval petiole bundles in W. angustifolia (W. areolata). When four strands are present, they include two large, median, crescent- shaped strands and two very small, spherical strands located proximal to the hypodermis on the abaxial surface of the petiole. The small abaxial strands originate from the division of the larger strands in the petiole base. In both W. virginica and W. radicans, small abaxial strands also arise from the division of the two large strands in the base of the petiole. In all species, a bundle sheath composed of large parenchyma cells with tanniferous substances surrounds each petiole strand. Cellular composition and arrangement of the stelar tissues is similar to the midrib (rachis) of a leaf. Transections of the adventitious roots of all three species show similar anatomical features (Fig. 18). The epidermis is typically sloughed off in mature roots, and the outer cortex, which is composed of irregularly shaped, thick-walled parenchyma cells, forms the outer boundary of the organ. The inner cortex consists of isodiamet- ric, sclerified, thicker-walled parenchyma cells that may be filled with tanniferous substances. A single-layered endodermis with Casparian strips delimits the stele. A pericycle composed of 1-3 layers of thin-walled parenchyma cells surrounds the vascular tissue. The primary phloem consists of sieve cells and phloem parenchyma, and the primary xylem is composed primarily of scalariform-pitted metaxylem and some spiral and transitional (reticulate-scalariform) protoxylem. In all three species, the primary xylem is diarch with exarch maturation (Fig. 18), although a triarch pattern is infrequently noted in W. radicans. In the species studied, the shoot apical meristem consists of a highly-vacuolated pyramidal apical cell averaging 50 x 80 ym (Figs. 19-20). Recent derivatives of the apical cell constitute the promeristem and comprise two groups of cells, the surface and subsurface zones. The promeristem in W. areolata is slightly dome- shaped (Fig. 19), whereas this tissue has a pronounced dome shape in W. virginica and W. radicans (Fig. 20). The surface zone in these three species consists of large, vacuolated, rectilinear cells and a few isodiametric cells on the periphery. McAlpin and White (1974) found a similar arrangement of the superficial cells in the genera Dryopteris and Quercifilix. The subsurface zone in W. areolata is distinct and consists of small, isodiametric cells (Fig. 19), whereas this zone is indistinct in the other two species. ; Comparative data indicate a close relationship between W. virginica, W. radicans, and W. areolata. Although the marked dimorphism and venation of the leaves justify the segregation of Lorinseria areolata from the genus Woodwardia, chromosome numbers (Wherry, 1964) and spore morphology (McVaugh, 1935; Hires, 1965) favor the retention of these species in the same genus. Comparative anatomical data also support the placement of all three species in the genus Woodwardia. 108 AMERICAN FERN JOURNAL: VOLUME 71 (1981) Grateful acknowledgment is made to Messrs. Alan Jones and James Neumann for their technical assistance in this study. LITERATURE CITED CODY, W. J. 1963. Woodwardia in Canada. Amer. Fern J. : COPELAND, E. B. 1947. Genera Filicum. Chronica wats imal A: pone M. L. 1970. Gray’s Manual of Botany. 8th corrected ed. D. Van Nostrand, Co., New York, pepe B. D. 1899. Partial sterility of fertile Woodwardia fronds. Plant World 2:55— HIRES, C. S. 1965. Spores. Ferns. Microscopic Illusions Analyzed, vol. 1. Mistaire kdaesee Millburn, NJ. JOHANSEN, D. A. 1940. Plant Microtechnique. McGraw-Hill, New York, AM LAKELA, O. and R. LONG. 1976. Ferns of Florida. Banyan Books, Miam ; LINNAEUS, C. 1753. Species Plantarum, vol. 2. Reprint ed. 1959. Ray ee, London. ———. 177]. Mantissa Plantarum Altera. Reprint ed. 1961. Hafner, New York, NY McALPIN, s and R. WHITE. 1974. Shoot organization in the Filicales: the promeristeti. Amer. J. Bot. 61:562-579. McVAUGH, R. 1935. Studies on the spores of some northeastern ferns. Amer. Fern. J. 25:73-85. d J. PYRON. 1951. Ferns of Georgia. Reprint ed. 1968. Univ. Georgia Press, Athens, GA. MOORE, T. 1857-62. Index Lome William Pamplin, London MORTON, C. V. and C. NEIDORF. 1956. The Virginia Chain- fern. Amer. Fern. J. PAYNE, W. W. and K. M. PereRseny 1973. Observations of the hypodermises of pitas Asher. Fern. 34-42. PRESL, K. B. 1851. = forage — Abhandl. K. Boehm, Gesell. Wiss. V, 6:361-624. RADFORD, m E., H. E. AHLES, and C. R. BELL. 1964. Guide to the Vascular Flora of the Carolinas. The hee paces Univ. of North Carolina, Chapel Hill, NC. SCULTHORPE. C. D. 1967. The Biology of Aquatic Vascular Plants. Edward Arnold, London SHAVER, J. M. 1954. Ferns of Tennessee. Bureau of Publications, Geo. Peabody College for Teacher’: Nashville. SMALL, J. K. 1938. Fert of the Southeastern States. Reprint ed. 1964. Hafner, New York, NY. SMITH, J. E. 1793. —— Botanicum. Mém. Acad. Turin. 5:401—413. WATERS, C. E. 1903. Ferns. Henry Holt, New York, NY. WHERRY, E. T. 1964. The powals Fern Guide. Doubleday, Garden City, NY. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) 109 Spore Germination Patterns in Anogramma, Bommeria, Gymnopteris, Hemionitis and Pityrogramma CLARK S. HUCKABY*, R. NAGMANI** and V. RAGHAVAN** Based on the morphological characters of the sporophyte and gametophyte, palynological evidence, and spore germination patterns seen in whole mount preparations, the genera Anogramma, Bommeria, Gymnopteris, Hemionitis, and Pityrogramma have been assigned to different families by fern taxonomists. Bower (1928) included these genera in the group designated as gymnogrammoid ferns, whereas Christensen (1938) placed all except Gymnopteris in the polypodiaceous group. Copeland (1947), however, regarded these genera as part of the complex group of pteroid ferns. Holttum (1949) followed Bower in his conception of gymnogrammoid ferns and combined the gymnogrammoid and vittarioid ferns of Bower into the family Adiantaceae. Nayar (1970) placed Anogramma in the Adiantaceae, Bommeria, Hemionitis, and Gymnopteris in the Cheilanthaceae and Pityrogramma in the Pteridaceae, whereas Crabbe et al. (1975) treated all five genera as members of the Adiantaceae. According to Haufler and Gastony (1978), gametophytes of Bommeria and Hemionitis responded differently from those of Anogramma and Pityrogramma to antheridiogen A, the male sex hormone isolated from the gametophytes of Pteridium aquilinum, to merit their further subgrouping. Previous light microscopic studies (Endress, 1974; Raghavan & Huckaby, 1980; Huckaby & Raghavan, 1981a; Rutter & Raghavan, 1978) on the orientation of the initial cell divisions during germination of spores of certain ferns using techniques generally employed in electron microscopy (fixation in an aldehyde fixative and embedding in glycol methacrylate plastic) have contradicted earlier work on spores of the same ferns based on whole mount preparations. Comparative studies of spore germination in Anemia, Lygodium, and Mohria (Schizaeaceae) (Raghavan & Huckaby, 1980) and in Cyathea and Dicksonia (Huckaby & Raghavan, 1981a) have also demonstrated the value of the early division sequence during spore germination as a stable taxonomic criterion in defining affinities of problematic genera. In view of the disagreement among pteridologists on the relationships of Anogramma, Bommeria, Gymnopteris, Hemionitis, and Pityrogramma, the present work is aimed at evaluating the importance of spore germination patterns in these genera as studied by modern histological techniques and in the scanning electron microscope (SEM) as a new source of taxonomic evidence. a Relatively little information is available in the literature on the initial cell division patterns during germination of spores of Anogramma, Bommeria, Gymnopteris, *Department of Biology, Syracuse University, Syracuse, NY 13210. **Department of Botany, The Ohio State University, Columbus, OH 43210. AMERICAN FERN JOURNAL: VOLUME 71 (1981) cell, HUCKABY ET AL.: SPORE GERMINATION PATTERNS 111 Hemionitis and Pityrogramma. Rao (1949) and Nayar (1956, 1962), who followed the early development of H. arifolia gametophytes, make no reference to the spore germination. Nayar’s (1964) study of P. calomelanos and P. chrysophylla gameto- phytes also lacks details of spore germination. In a study of G. vestita gametophytes, Kaur (1972) states, without giving any details, that spore germination is of the Vittaria type described earlier by Nayar and Kaur (1968). In this type of germination, a rhizoid is cut off at the proximal pole of the spore by a wall perpendicular to the polar axis. The protonemal cell is formed by a subsequent division of the large distal cell perpendicular to the first division wall. Elongation of the rhizoid parallel to the polar axis of the spore and that of the protonemal cell along the equatorial plane are also characteristic of this type of germination. In their review of the gametophytes of homosporous ferns, Nayar and Kaur (1971) make statements which imply that the germination pattern of Anogramma, Bommeria, Gymnopteris, Hemionitis, and Pityrogramma spores follows the Vittaria type, Based on whole mount observations, Haufler (1979) recently showed that spores of Bommeria hispida germinated by a division wall perpendicular to the polar axis, the small proximal cell differentiating into the rhizoid; origin of the protonemal cell was not traced in this work. However, in a proportion of spores of B. subpaleacea, the small (proximal?) cell formed from the first division is believed to yield the protonemal cell, the rhizoid being derived by a division of the same cell as the protonemal. According to Baroutsis (unpublished work cited by Haufler, 1979 and pers. comm.), who followed in whole mounts the pattern of cell division during germination of spores of several species of Anogramma, the first division of the spore to form the rhizoid is oblique or nearly parallel to the polar axis. According to this investigator, in A. osteniana, the protonemal cell sometimes appeared to form before the rhizoid. MATERIAL AND METHODS Spores of Anogramma chaerophylla, Bommeria pedata, B. ehrenbergiana, B. hispida, B. subpaleacea, Gymnopteris rufa, Hemionitis arifolia, H. palmata, H. pedata, Pityrogramma calomelanos, and P. chrysophylla used in this work were obtained from various sources! and stored at 5°C until used. To follow cell division pattern during germination, spores were sown on the surface of 10 ml modified Knop’s liquid medium (Raghavan, 1965) contained in 5 cm diameter Petri dishes and allowed to imbibe in the dark for 48 hr. They were then irradiated continuously with red light or exposed to weak fluorescent light during a photoperiod of ca. 12 hr as described earlier (Huckaby & Raghavan, 1981a, b). Samples were collected at intervals of 12-24 hr during an experimental period of 6 days, fixed in 10% acrolein, 'Spores were obtained from the following sources: A. chaerophylla, B. pedata and H. pedata from Dr. J. T. Mickel, New York Botanical Garden, Bronx, New York; B. ehrenbergiana, B. hispida, B. subpaleacea, and H. palmata from Dr. C. H. Haufler, University of Kansas, Lawrence, Kansas; H. arifolia and P. calomelanos from Dr. C. N. Page, Royal Botanic Gardens, Edinburgh, Scotland; G. rufa and P. chrysophylla from Dr. T. Walker, Royal Botanic Gardens, Kew, England. 112 AMERICAN FERN JOURNAL: VOLUME 71 (1981) FIGS. 5-8. Germination of Bommeria spores. FIG. 5. B. pedata section owe the rupture of the the exine; the spore nucleus is in mitosis. FIG. 6. B. ehrenbergiana section showing the formation of rhizoid initial (arrow). oe a B. pie section showing the rhizoid initial (r). Arrow points to the nucleus of the distal cell. . 8. B. pedata section a the protonemal cell (p); (r) is part of the rhizoid. The nucleus of e pee cell (arrow) is in divisio HUCKABY ET AL.: SPORE GERMINATION PATTERNS 113 and embedded in glycol methacrylate according to routine procedures previously used (Raghavan, 1976, 1977; Raghavan & Huckaby, 1980; Huckaby & Raghavan, 1981a, b; Rutter & Raghavan, 1978). Sections cut at 7 pm thickness on a rotary microtome equipped with a steel knife were stained in 0.05% toluidine blue, mounted in Euparal, and examined in the light microscope. For examination in SEM, spores were fixed in 70% ethanol-acetic acid (3:1) at 4°C, dehydrated in ascending series of ethanol and subjected to critical point drying. They were subsequently mounted on SEM specimen mounts using double-stick cellophane tape, vaccuum-coated with gold, and examined in a Hitachi S-500 instrument. SPORE GERMINATION RESULTS Anogramma chaerophylla.—Spore germination was initiated by the splitting of the exine at the proximal pole along the lines of the trilete scar after about 36—48 hr in red light (Fig. 1). Following this, the spore protoplast divided by a wall perpendicular to the polar axis, resulting in a small proximal cell and a large distal cell (Fig. 2). The former elongated into the rhizoid, with a corresponding enlarge- ment of the latter. The distal cell also acquired chloroplasts and appeared outside through the opening in the exine as a green protrusion, displacing the rhizoid laterally, so that the wall delimiting the rhizoid from the distal cell appeared parallel to the polar axis (Fig. 3). The protonemal cell was formed «about 72 hr after exposure of spores to red light by the division of the distal cell by a wall perpendicular to the first wall (Fig. 4), although due to the elongation of the distal cell, the second division wall barely intercepted the first. The division sequences observed in spores exposed to white light were similar to those seen in red light. In spores germinated in either light regimen, the protonemal cell grew parallel to the polar axis of the spore and the rhizoid grew at a right angle to the protonemal cell. Initial division of the spore protoplast oblique or nearly parallel to the polar axis as described in this genus by Baroutsis (cited by Haufler, 1979 and pers. comm.) was not observed in sections made from our sample. Bommeria ehrenbergiana, B. hispida and B. pedata.—Spores of B. ehrenbergiana, B. hispida, and B. pedata germinated by cracking of the spore coats at the proximal pole after exposure to red light for about 48 hr. As in the case of A. chaerophylla, the first division of the spore protoplast perpendicular to the polar axis yielded a small proximal rhizoid and a large distal cell (Figs. 5-7). When spores initially exposed to red light were transferred to white light for 24 hr, protonemal initiation occurred by the division of the distal cell by a wall perpendicu- lar to the first. Occasionally after cutting off the protonemal cell, the nucleus of the distal cell was found to divide again, probably giving rise to a secondary rhizoid (Fig. 8). Exposure of spores of B. ehrenbergiana and B. pedata to white light or red light alone resulted in the elongation of the distal cell through the opening in the exine accompanied by chloroplast accumulation, but its division to form the protonemal cell was not observed during the experimen hispida exposed to white light gave rise to the rhizoid and protonemal cell according 114 AMERICAN FERN JOURNAL: VOLUME 71 (1981) oo TN it FIGS. 9 and 10. Germination of Bommeria hispida spores. FIG. 9. Section showing the rhizoid (r). Arrow points to the distal cell enlarging preparatory to division to form the protonemal cell. FIG. 10. Section showing the distal cell (arrow), partially loose from the wall, after cutting off the protonemal cell (p); (r) is the rhizoid. FIGS. 11 and 12. Germination of Gymnopteris rufa spores. FIG. 11. Section showing the first division giving rise to a small proximal rhizoid initial (r) and a large distal cell (dc). Arrow points to the division wall. FIG. 12. Section showing the elongation of the distal cell (dc) and the migration of its nucleus (arrow) to the tip. HUCKABY ET AL.: SPORE GERMINATION PATTERNS 115 to the same division sequence observed in the other two species. In spores of all three species examined, due to lateral expansion of the distal cell, the orientation of the wall delimiting the rhizoid was changed so that the wall appeared parallel to the polar axis of the spore (Figs. 9 and 10). In spores of Bommeria examined by us, the orientation of the initial division wall yielding the rhizoid is similar to that described by Haufler (1979) in B. hispida. However, a sufficient quantity of spores of B. subpaleacea was not available to us to confirm or refute Haufler’s (1979) observation that in some spores of this species, the proximal cell arising out of the first division gave rise to the protonemal cell. ymnopteris rufa.—The first division, which occurred in spores exposed to red or white light regimens for 72-96 hr, was perpendicular to the polar axis and yielded a small proximal cell and a large distal cell (Fig. //). This was the only division observed in the germinating spore during the experimental period. As the proximal cell differentiated into a rhizoid, the distal cell elongated, became chlorophyllous and appeared outside as a green cell. Later the nucleus moved from the basal part of the cell enclosed within the exine to its exposed tip (Fig. /2). A transverse division occurred at the tip of the distal cell during its further growth in red or white light. A similar pattern of cell disposition has been described in germinating spores of Preris vittata (Raghavan, 1977). Hemionitis arifolia, H. palmata and H. pedata.—Spores of all three species followed a uniform pattern of germination in which the first division of the spore protoplast by a wall perpendicular to the polar axis gave rise to a small proximal cell and a large distal cell. As seen in other genera, the small cell differentiated into the rhizoid (Figs. 13 and 14). Rhizoid initiation occurred after exposure of fully imbibed spores to a red or a white light regime for 48 hr and was preceded by the opening of the exine at the trilete mark. Formation of the rhizoid was followed by enlargement of the distal cell through the opening in the exine and its division by a wall perpendicular to the first to give rise to the protonemal cell (Fig. 15). As seen in SEM preparations (Fig. 16), expansion of the distal cell resulted in the displacement of the wall separating the rhizoid from the distal cell to a plane parallel to the polar axis of the spore. In spores of H. palmata grown in red or white light regimes, the division of the distal cell to form the protonemal cell was not observed during the experimental period. : Pityrogramma calomelanos and P. chrysophylla.—Spores of both species are characterized by the presence of equatorial flanges of sporoderm material which can be used as a marker for the equatorial plane. Spores responded to white or red light regimes in about 24 hr by the cracking of the spore walls at the trilete mark and by the emergence of the rhizoid (Figs. /7 and /8). The latter was traced to the proximal cell formed by the division of the spore protoplast by a wall perpendicular to the polar axis (Fig. 19). Following rhizoid formation, the distal cell expanded laterally and elongated through the opening in the exine parallel to the polar axis, displacing the rhizoid in a plane parallel to the equatorial axis (Fig. 20). The division of the distal cell to form the protonemal cell was delayed until the nucleus migrated from the base of the cell to its tip, and occurred by a wall perpendicular to the first division wall. AMERICAN FERN JOURNAL: VOLUME 71 (1981) 116 13 50um hy Pr 16 FIGS. 13-16. Germination of Hemionitis spores. FIG. 13. Section of H. pedata showing the first division to form the rhizoid initial (r) and a distal cell. Small arrow points to the nucleus of the distal cell; part of the distal cell has become loose from the spore wall (large arrow). FIG. 14. Section of H. arifolia showing the rhizoid (r) and the undivided distal cell; arrow points to the nucleus of the distal cell. FIG. 15. Section of H. pedata showing the protonemal cell (p) and rhizoid (r). White arrow points to the wall delimiting the rhizoid; black arrow points to the wall delimiting the protonemal cell. FIG. 16. Scanning electron micrograph of H. pedata showing the displacement of the rhizoid wall (arrow): (r) is the rhizoi HUCKABY ET AL.: SPORE GERMINATION PATTERNS 117 DISCUSSION Planes of cell division during germination of Anogramma, Bommeria, Gymno- pteris, Hemionitis, and Pityrogramma spores described here follow the general pattern of the Vittaria type of Nayar and Kaur (1968). The spore is divided by two walls, the first one perpendicular to its polar axis yielding a rhizoid and the second, perpendicular to the first, giving rise to the protonemal cell. However, due to the displacement of the rhizoid initial by the expansion of the distal cell, the rhizoid and protonemal cell have been found to elongate in planes opposite to those described by Nayar and Kaur (1968). The only difference noted between the initial division patterns of spores of different genera is in the timing of the second division under the experimental conditions employed. As seen in spores of Gymnopteris rufa, the distal cell apparently functions as the protonemal cell due to a delay in the second division, giving the impression that a single division of the spore protoplast is sufficient to give rise to a functional gametophyte. However, the fact that under extended periods in both red and white light regimes, division of the spore protoplast follows the Vittaria type, reinforces the stability of this character in the early gametophyte development of ferns. On the basis of our observations, we believe that the report of Baroutsis (cited by Haufler, 1979 and pers. comm.) on the formation of the rhizoid in spores of several species of Anogramma by a wall oriented obliquely or nearly parallel to the polar axis is due to failure to identify the first division wall as soon as it is formed. As seen in Figs. 5, 6, //, 13, and 19, this wall appears before or immediately after the exine is ruptured and by the time the wall is visible in whole mounts, the distal cell would have expanded, displacing the original wall and giving the false impression of its occurrence parallel to the polar axis (Figs. 3, 9, and /5). If the cell division pattern during spore germination can be considered a stable character for taxonomic purposes, along with other features of the gametophyte and of the sporophyte, the uniformity in the pattern observed in the five genera investigated here tends to support their assignment to a single family, Adiantaceae, as done by Crabbe et al. (1975). Since no distinctive variants of the germination pattern were seen between groups of genera, a further subgrouping separating Bommeria and Hemionitis from Anogramma and Pityrogramma as suggested by Haufler and A comparison between the cell division patterns observed in sectioned spores of certain genera of Schizaeaceae investigated earlier (Raghavan, 1976; Raghavan & According to Nayar (1970), Vittaria type germination is evolutionarily more ad- vanced than the pattern observed in the Schizaeaceae, designated as the Anemia type ore germination based on whole mount prepara- patterns of cell division during sp m here is an increasing body of opinion (Holttum, tions may be overstated. However, t 118 AMERICAN FERN JOURNAL: VOLUME 71 (1981) 19 50um FIGS. 17-20. Germination of Pityrogramma calomelanos spores. FIGS. 17 and 18. Scanning electron micrographs showing the trilete mark and e flange. j 2g of rhizoid (r). Arrows point to the ee 19. Section showing the first division. Arrow points to the rhizoid initial. FIG. 20. Scanning electron micrograph showing elongation of the distal cell (de) and displacement of the rhizoid (r). HUCKABY ET AL.: SPORE GERMINATION PATTERNS 119 1949; Crabbe et al., 1975) on the possibie origin of adiantoid ferns as a distinct group from the schizaeaceous stock. Raghavan and Huckaby (1980) have shown that spores of M. caffrorum follow the same route as the five genera studied here to form the rhizoid and protonemal cell, exhibiting typical Vittaria type germination. The existence of similar patterns of division during spore germination in the Adiantaceae and in a member of the Schizaeaceae might suggest a close relationship between the two families, but examination of the germination patterns of spores of other species of Mohria and of Schizaea and Actinostachys is necessary before this evidence can be used to support the possible origin of adiantoid ferns from a schizaeaceous ancestry. This work was ud aos by grant (DEB 78-01297) from the National Science Foundation to V. Raghava LITERATURE CITED BOWER, F. O. 1928. The Ferns (Filicales), vol. III. University Press, Cambridge. CHRISTENSEN, C. 1938. Filicinae. Jn F. Verdoorn (ed.), Manual of Pteridology. Martinus Nijhoff, The Hague. COPELAND, E. B. 1947. Genera Filicum. Chronica Botanica, Waltham, MA. CRABBE, J. A., A. C. JERMY, and J. T. oar 1975. A new generic sequence for the a herbarium. Fern. Gaz. 11:141- ENDRESS, A. G. 1974. Spore germination of cet thalictroides (L.) Brongn. Ann. Bot. 38: a 7-381. HAUFLER, C. H. 1979. A biosystematic revision of Bommeria. J. Arnold Arbor. 60:445—-4 _ and G. J. GASTONY. 1978. Antheridiogen and the breeding system in the fern genus Bon meria. Canad. J. Bot. 56:1594—1601. HOLLTUM, R. E. 1949. The classification of ferns. Biol. Rev. 24:267-296. HUCKABY, C. S. and V. RAGHAVAN. 1981a. Spore germination patterns in the ferns Cyathea and Dicksonia. Ann. Bot. 47:397—403. . and V. RAGHAVAN. 1981b. Germination of the spores of the thelypteroid ferns. Amer. J. Bot. 68:517-523. KAUR, S. 1972. Morphology of the prothallus of Gymnopteris vestita. sociopath 22:46-49. NAYAR, B. K. 1956. Studies in Pteridaceae Il. Hemionitis Linn. J. Indian Bot. ip stage ________. 1962. Ferns of India—V. Hemionitis. Bull. Lucknow Natl. Bot. Gard. 6 . 1964. Some aspects of the morphology of Pityrogramma calomelanos = es aie ‘. Indian Bot. Soc. 43:203-213. . 1970. A phylogenetic classification of the homosporous ferns. Taxon 19:229-236. _ and S. KAUR. 1968. Spore germination in homosporous ferns. J. Palynol. 1:10—26. a - KAUR. 1971. Gametophytes of homosporous ferns. Bot. Rev. 37:295-396. RAGHAVAN, V. 1965. Action of purine and | analogs on the growth and differentiation of the gametophytes of the fern Asplenium nidus. Amer. J. Bot. 52:900-910. 1976. Gibberellic acid-induced germination ie? spores oH Anemia phyllitidis: Nucleic acid and nade synthesis during germination. Amer. J. Bot. 972. Cell morphogenesis and macromolecule synthesis during phytochrome-controlled germination of spores of the fern, Pteris vittata. J. Exp. Bot. 28: 439-456. _ HUCKABY. 1980. A comparative study of the cell division patterns paring , and C ria (Schizaeaceae). Amer. germination ‘of spores of Anemia, Lygodium and Mohri 67:653-663. RAO, A. R. 1949. The prothallus of Hemionitis arifolia Sm. Curr. Sci. 18:349- RUTTER, M. R., and V. RAGHAVAN. 1978. DNA synthesis and cell division during spore germination in Lygodium japonicum. Ann. Bot. 42:957-965. 120 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) SHORTER NOTES THE CHEMOIDENTITY OF THE HOLOTYPE OF PITYROGRAMMA TRIANGULARIS.—In recent years, the Pityrogramma triangularis complex has been the subject of extensive phytochemical studies, for example, by Smith (Bull. Torrey Bot. Club 107:134—145. 1980 and references therein). The nominal variety tacitly recognized by Weatherby (Rhodora 22:113-120. 1920), var. triangularis, has proved to be rather complex in itself with respect to the composition of frond-surface flavonoids and chromosome numbers. The plants subsumed under var. triangularis are morphologically very similar, but their extreme chemical differences and distinct regional distribution suggest that further taxonomic revision within this group may prove warranted. It is critical, therefore, to establish the chemical identity of the holotype of var. triangularis, since ultimately the one biological entity identifiable with the holotype must bear the name var. triangularis. The holotype, which is in the herbarium of the Berlin Botanical Garden, was collected by the German poet, botanist, and globetrotter Adelbert von Chamisso. It bears a note in his handwriting “Gymnogramma triangularis Kaulf. Enum. p. 73, legit deditque A. v. Chamisso, California.” According to Eaton, quoted by Alt and Grant (Brittonia 12:153—170. 1960), the specimen was collected in 1816 near San Francisco. It is filed at Berlin under “Polypodiaceae, Gattung No. 62a. Ceropteris, Art No. 21 triangularis.” A minute fragment of the holotype was made available for analysis of its farina; 22 mg of material was rinsed with acetone to dissolve the exuded flavonoids. These were identified by direct comparison with authentic markers on polyamide-TLC; for experimental details see Wollenweber, Dietz, Schillo and Schilling (Z. Naturforsch. 35c:685-690. 1980). The major constituent is ceroptin; minor components are triangularin and another compound which is not yet fully elucidated, “tvt-11,” according to Dietz (unpubl. dissertation, Darmstadt). This flavonoid pattern is characteristic of those plants representing the ceroptin chemotype of var. triangularis. The holotype specimen itself chemically resembles those plants of the diploid ceroptin type collected by D. M. Smith from Refugio Pass, Santa Barbara County, California according to Star, Seigler, Mabry and Smith (Biochem. Syst. Ecol. 2:109-112. 1975). We think it is a most remarkable result that the holotype of P. triangularis can be equated unambiguously with the typical and well defined ceroptin chemotype of var. triangularis. This illustrates the powerful role that chemotaxonomy can play in determining the application of names by the type method. Thanks are due to Dr. D. Meyer, Berlin, for kindly supplying the fern fragment used in this study.—Eckhard Wollenweber, Institut fiir Botanik, Technische Hochschule Darmstadt, Schnittspahn- strasse 3, D-6100 Darmstadt, Federal Republic of Germany and Dale M. Smith, Department of Biological Sciences, University of California, Santa Barbara, CA 93106. AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) 121 A MAJOR RANGE EXTENSION FOR THELYPTERIS SIMULATA IN THE SOUTHERN APPALACHIANS.—Recently, while examining fern speci- mens in the Illinois Natural History Survey Herbarium (ILLS), I found a collection of Thelypteris simulata (Davenp.) Nieuwl. collected in August 1931 at New Found Gap, Sevier Co., in eastern Tennessee at the North Carolina border (Pepoon 935). Herman S. Pepoon was a botanist from Chicago, Illinois and the author of “Flora of the Chicago Region.” Pepoon himself had identified the specimen as Aspidium simulatum Davenp., and wrote on the label that the plant occurred at “summit elevations, damp woods near New Found Gap.” Pepoon deposited several hundred specimens from eastern Tennessee in ILLS, and it is unlikely that a person of his competence would have correctly identified this unusual fern and then would have made a gross error in describing its location. Furthermore, Pepoon’s collecting notebooks, which are preserved at ILLS, show that he did not collect in west-central Wisconsin or in the New England states from which this fern presently is known. Thus, it seems unlikely that the specimen represents a false record due to a label mix-up. The specimen is significant because it represents a disjunction of about 400 miles from the nearest station recorded for this species, which is in northern West Virginia (Tryon & Tryon, Amer. Fern J. 63:66. 1973). In addition, this find may bear upon the species distribution during the late Pleistocene; it seems likely that 7. simulata was present and probably more widespread in the southern Appalachians during Wisconsinan glacial times, and then gradually became restricted to higher elevations and latitudes as the climate changed and the glaciers receded during the Holocene. Although long-distance spore dispersal can never be completely ruled out, it seems unlikely. Many other pteridophytes show a similar widespread distribution in the north and become gradually restricted or disjunct at high elevations southward in the Appalachians, including Botrychium matricariifolium B. multifidum, Dryopteris campyloptera, Gymnocarpium dryopteris, Lycopodium annotinum, L. selago, and Phegopteris connectilis. | encourage pteridologists to look for populations of this interesting species in acid, boggy areas at high elevations in the Great Smoky Mountains region.—Robbin C. Moran, Herbarium, Illinois Natural History Survey, Natural Resources Bldg., 607 E. Peabody Dr., Champaign, IL 68201. A NEW INDIANA STATION FOR EPIPHYTIC RESURRECTION FERN.— In 1975, the author discovered a large, healthy population of Polypodium polypodioides (L.) Watt growing approximately 30 feet up on a branch of a dead tree at Carnes Mill, Crawford County, Indiana. Hemlock (Tsuga canadensis), Yellow Buckeye (Aesculus octandra), and Goatsbeard (Aruncus dioicus) flourish on the sandstone cliffs and the tight valley slopes encasing the Little Blue River at this point. Although commonly found at similar heights in the south, literature research indicates that this is a relatively rare habitat in the interior northern reaches of this fern’s range. It is reported by Deam (Flora of Indiana, 1940) from nine counties in 122 AMERICAN FERN JOURNAL: VOLUME 71 NUMBER 4 (1981) southern Indiana. He stated that he once found it growing in the crotch of a Bur Oak tree in Wabash County, but that “this is the only specimen I have ever seen growing on a tree in Indiana, although it is common in this habitat in the south.” Welch in Lindsey (The Natural Features of Indiana, 1966) stated that she had not seen the Resurrection Fern growing on a tree in Indiana.—Ronald R. Van Stockum, Jr., 810 Kentucky Ave., Frankfort, KY 40601. AMERICAN FERN JOURNAL Manuscripts submitted to the JOURNAL are reviewed for scientific content by one or more of the editors, and, often, by one or more outside reviewers as well. During the past year we have received the kind assistance of J. D. Caponetti, A. M. E , R. L. Petersen, H. E. Robinson, R. G. § tolze, W. C. Taylor, R. A. White, D. P. Whittier, and J. J. Wurdack, to whom we are deeply indebted. We welcome suggestions of other reviewers.—D.B.L INDEX TO VOLUME 71 Acrostichum areolatum, 101 Actinostachys. 119 Adiantaceae, 109, 119 Adiantum capillus-veneris, 91 per ore salvinii, 75-79 Amphine , 82: immersum, 82 "101 117; anthriscifolia Anemia, 109, . 100; tomentosa var. mexicana, 00 erica 109, 111, 117: chaerophylla. 110, 111, 113; osteniana, Manes 65: simplicior, 65-67 Arachniodes simplicior new to South Carolina and the United States, 65 — acutum, 35; alpestre, 91; biserratum, 35; conterminum, 60: co sdibolisien, be exaltatum, 38: immersum, 82; simulatum. 121; whereas gases fe aon -nigrum, 91: x ere 89; ebeneum, 90: riophyllum, 91; . 90; ruta-muraria, tts p. trichomanes, 95: ‘ ; trichomanes-ramosum, 91: rene! Ac 88; x herb-wagneri, 89; pinnatifidus. 87. ; Shawneensis, 85-8 * Asplenosorus nsevionmsnanee . new — fern hybrid between hizophy Illus, Athyrium, 48: alpestr var. ame no subsp. american- » 91; Reinnasper pe descsithction: ot: filix-femina, 5-8: Me cum, 91 Azolla, 33; filiculoides, 33; carolinian Azolla filiculoides new to the vmiionis United States, 33 Bates, V. M., E. T. Browne, Jr. Azolla filiculoides new to the Da cae Unie States, 33 By, -S.°3: . Satija. Lepisorus kashyapii in the western Himalayas, i. aie clubmosses (Lycopodiella) in Kentucky, 9 Bom 1695-499, (E925; 899; pi ee 111-113: hsp, VW, 113-115; pedata, 100, 111-113; hialnicet: 11, 115 Botrychium, 1, 15, 16, 8. 20, 25, 28; biternatum, 13-15, 18; bo 5,30: besperium, sie lavaciacdaiicha: 25, 28, 30, 68, nari 8, 30, f. incisum, var. * cbiaiinan. pe 2 30; —30; multifidum, ; m, 21, 24- 26: ternatum, 13; simplex, 2, 23, 28, 92; virgini . 13, 14, 17, 28, 30 = rach pattern of Hypolepis ye 41 . T., Jr. (see V. M. Bates, Jr. raved si sorus rhizophyllus, 85, 87-89 Chain ferns of Florida, | Cheil S ag castanea, 62; ni 62, 100; fendleri, 100: microphylla, 91; hii, 62: , 62: wootonii, 100 The peaei of vi py of | Plate triangularis. 120 Christella, Ph reser s, 82 Clifford, t cole Ferns, fern allies and conifers of pean ioe” Cody, W. J. ( . (see A. E. Porsild) INDEX TO VOLUME 71 Comparative ecology of Woodsia scopulina sporophytes and ga- metophytes, Constantine, Ay ee yk i = Cranfill, diella) in Kentucky, 97; Ferns = fem alee of Kentucky pe 12 , 109 c ve Cy aoa 4 Cystopteris dickinson fs alee 31, 92, var. mackayii, 92. subsp. tenuifo rusa, 92: reevesiana, 92; tenuis, 92 Diplazium ghee and ‘Sclaginella uncinata newly discovered in eorgia, 4: Dryopteris, 107; arist consanguinea var. aequalis, ea 68; filix-mas, 46; phegopteris, 68 Hah 75 Equisetum arvense, be fluviatile, 1; beara ats 1; rachyodon, |, 2; variegatum, ma, ata, 46; campyloptera, 121; baea, 60: 60; dilatata, 68: ones neta, 68: immersa, 82; ludoviciana, 49: ; eat 100: 2 trachyodon in New Jer Fairc wees ths R. Diplazium japonicum and Sclvginatla: ‘ancinata Fete oat _ eae of Kentucky (rev.), 12 Ferns, fi f Australia (rev.), 9 Flora of Baja Calife ), 100 Fosberg, F R. & M--H. Sachet. Nomenclatural notes on Micro- ee sapen seis hrolepis in Florida, 35 Clapper soy Gordon, Judith iaahaaiiins simplicior new to South Carolina and the ne ae 6 Grammitis, 75 Gruber, T. M b f Hypolepis repens, 41 68, 121, subsp. disjunctum, 68 Gy <, +, Gymn nogrecnea ang, 120 Gymnopteris, 109, , 117; rufa, 11 , 114, 115, 117; vestita, 111 ae 117; ee 115135,-1 16; palmata, 69 an. Spore germination ; Vi ing mma, Bommeria, Gymnopteris, Hemionitis ont 70; ne 69-73; mexicana, 69, 70: .¢ oss gus 70: tuerckheimii, 70 alvinia minima new to Louisiana, 95 Leaf turnover rates and vat iisioey of the Cetera American tree tm Alsophila orcbees oo Lepidoneuron biserratu Lepisorus, 53; e avatus, S38 oe 53, var. kashyapii, 53. var. . Var. S35 eae kasapi in ak western Himalayas. 53 Lellinger, . Notes on North American Pen 90 Lorinseria me olata, 48, 49, 101 Lucansky, T. W. Chain ferns of Florida pe Lycopodiella, 97 des, 97. 99: appressa, 97-99: x brucei, 97-99: x copelandii, 97: prostrata, 97-99 opodium 17, 97; alopecuroides, 97: pees ik x Lyco} bopelendii, 97: dendroideum, 31; inundatum, 31, var. appressum. 123 97, var. elongatum, 97; eke Lepidotis, 97: selago, 121 Lygodium, 109, 117; japoni A major atl extension Pai Thetypeeris simulata in the southern Appalac hep Marsilea pone a a3, — 93 Mi rog 29: falsivenulosum. 83: motleyi Mohria, 109, 117, 119: cafforum, 119 Montgomery, J. D. Equisetum variegatum and E. x trachyodon in New Jers sis, a new natural fern and Cam) tosorus rhizo- , R. C. XAsplenosorus shawneen hybrid between nar trichomane ee = =] a g ae SRO eo @ +* nn 3 3 e ai Nau . E. The genus Nephrolepis in Florida, Nem acutum, 35; biserratum, 35; caribaeum, oo conterm m, 60; exaltatum, 38; ayaa var. sitioram, 60; kaulfussii. $9. er pentane 58; tu Nephrolepis, 35-37, os acuta, 35; x averyi, 35- 39, 40: biserrata, 35-40, Furcans, 40 sewers 163 var. tuberosa, 37; exaltata, 35, 36, . bise ] Bostoniensis, 40, cv. ntissima, 40. cv. tephra s. 4 M. P. s, 40, var. tuberosa, 37: falcata f. asi 40: utula var. acuta, 35, serrata, 35, cv. Superba, 40: multiflora, 35-39; pectinata, 40; rivularis, 38; tuberosa, 37 A new BAe station for a tic resurrection fern, 121 A new Isoétes from sa ca, e — S ie moonwo = Brn subg. Botrychium (Ophio- olen ssaceae) from sip ae apace notes on Micronesia ferns, 82 Notes on North American sa on North American serie vascular plants— ge 2 Notes on Selaginella, with a new variety of S. pal oe Notholaena bryopoda, 62: greggii, 62: neglecta, = gi 63 O nsibilis, 46, 48 10, 25 be hs singe 13, 14, 16, 17: intermedi- are, paagtaorantt 13: pendu- 35 vul 13. va Ma ; 25; ti lum, 13; petiolatum, cS pe : pron m, 13-16, 18, var. igen 13, eve 7 mea ; ‘es regalis, 4 ppv ee 7 interm ile 63; longimucronata, 100: , 100; skinneri, 100 tilis fa _ Programa 10, a austroamericana, | 1: calomelan- we me 5 by: cy il, 111, 115, 118: PSOE 63, 12 0, var. maxonii. 63: trifoliata. 11 BF <= Polypodiaceae, 13, 100 Polypodium, 60, 93; californicum, 63: coetitotint, so exaltatum. i : ilentum, 60: 38; glycyrthiza, 63, subg. G le , hesperium, 63; kashyapii, 53: subg. arginaria, 93: pectinatum, 93. subg. Pectin: 93: polypodioides, 121. subg Polypodium, m, = scouleri, 93; virginianum, 68: vulgare. 93, subsp. virginian- m, 68 Pol inh tichoides , 67 aay. Vascular plants of continental Pa Porsild, E. & W. cites territories, da (rev.). Proct ‘axonomic notes on Jamaican — 57 Pteridium, “100; aq uilinum, 48, 109, var. pubescen’ Pteris multifida, ‘Gr: vittata, 115 Quercifilix, 107 Raghavan, V. (see S. C. Huckaby) 124 Range ~~ for two lycopods on Baranof Island, southeastern Alaska Reeves, T. ‘oie on North American lower vascular gt —Il, 62 Reviews: Ferns and fern allies of Kentucky, 12; Ferns, fern allies and conifers of Australia, 9; Flora of Baja California, 100; L. inopsis de las especies Lycopodiu (Lycopodiaceae Pteridophyta) de la seccién Crassistachys Herter, 84: Vascular plants of continental a territ Spey 68 ra, a t, M.-H. (see F. R. Fosberg) 95 ima new to Louisiana, 95 i K. eS. S. Bir Scale ge Paar on fariose species of Pityrogramma, 10 Schizaea, 119; pusilla, 25 Schizaeaceae ci oh 117, i Seiler, R. B bs es and natural history of the Central merican tree fern A sein salvinii, 75 — sh arizonica, 64; cuspidata, 51, var. elongata, 51: a. 63 i gynandrum, 51: ae 49, 50 Short, J. W. Equisetum arvense in Alabama, 64 Sinopsis de las especies de Lycopodium L. ridophy la ( Soho nashud seccion Crassistachys Herter (rev. le Smith, D. M. ( Spore germination and young gametophyte i At of Botry- chium and Hoglossum it in axenic culture S rminat! Anogramma, oN Gymno- rogramma, 109 on Selaginella, with a new variety of S. yea ret ferns—III, 57 Thelypteris. 57, 59, Adenophyllum, 58: subg. Adeno- phyllum. 57, 58: sect. "nea 58, 59; whe Amauropelta, AMERICAN FERN JOURNAL: VOLUME 71 (1981) 57-59; subg. Apelta, 57; balbisii, 585.59; ~~ —— , 60 a3 57; subg. Blepharitheca, 57; baea, 60: dec 60; dentata, 48, 82: gracilenta, 60: gracilis, 60: ns sis, “s harrisii, 60; hispidula var. versicolor, 94; immersa, 82 if . uen at | 57, 58: ran ea, . 57, f. crista zt is mulata, 121; torresiana, a8 tela, x 59; subg. ttncinelia. 37: underwoodian versicolor, ea 8A: a 82: falsivenulosum, 83; motleyi, Trichopteris schiedeana, 75 An unusual record of Asplenium trichomanes from northeastern Van ‘Stock R. rection fern 542 satus plants of ponienand northwest territories, Canada (rev.), 68 . Jr. A new Indiana station for epiphytic erage (see P. J. Watson) ee —, eae) sr = Rae Ba: speci moonworts, ae pas iain ee from North America, 20 Watson, P. J. & Margarita Vazquez. Comparative ecology of Wood- sia scopulina ee wR auaesgaaaile Whittier, D d you ng ipiarieiea-tail develop- ment of Botrychi 1 Ophi 13 Wiggins, I. L. Flora of Baja Cali 100 ele VOB, ‘Dietz. Scale insects feeding on farinose a of Ptyrogramma, 10 Wollenwe .&D ~~ Smith. The chemoidentity of the holo- type o e paid rogramma triangularis, Woodsia, 8; alpina, 68; jr ace 68; oregana, 68; plummerae, = bi lin = 3-8 wardia, 101, 107, angustifolia, 101; areolata, 98, rae ns, im om 105-107; virginica, 48, 101-107 7 - 101-107: ERRATUM FOR 1980 Page 67, line 39: For “39: in press.” read “ns. 6:210-238" BRITISH PTERIDOLOGICAL SOCIETY Open to all who are interested in growing and studying ferns and fern-allies. Full members receive THE FERN GAZETTE and the BULLETIN OF THE BRITISH PTERIDOLOGICAL SOCIETY. Membership subscriptions are £ 7 for full members, £ 5 for ordinary members (not receiving the GAZETTE), and £5 for student members (under 25 years of age). For particulars, U. S. residents should apply to Dr. J. Skog, Biology Department, George Mason University, Fairfax, VA 22030. Non-U. S. residents should apply to Lt. Col. P. G. Coke, Robin Hill, Stinchcombe, Dursley, Gloucestershire, England. TRIARCH Over 5@ Wears of slide manufacture and service to botanists. We welcome samples of your preserved research material for slide-making purposes, and we invite your suggestions for new slides that would be use- ful in your teaching. Your purchases have made our 50 years of existence possible. To satisfy your con- tinued need for quality prepared slides, address your requests for catalogs or custom preparations to: TRIARCH INCORPORATED P.O. Box 98 Ripon, Wisconsin 54971 eee